mathx”17
Quasi-invariance of the Gaussian measure for the two-dimensional stochastic cubic nonlinear wave equation
Abstract.
We consider the stochastic damped nonlinear wave equation on the two-dimensional torus , where denotes a space-time white noise and . We show that the measure corresponding to the unique invariant measure for the flow of the associated linear equation is quasi-invariant under the nonlinear stochastic flow.
Key words and phrases:
stochastic nonlinear wave equation; invariant measure; white noise, quasi-invariance, gaussian measure, absolute continuity2020 Mathematics Subject Classification:
35L15, 60H15Contents
1. Introduction
We consider here the following stochastic damped nonlinear wave equation (SDNLW):
(1.1) |
where , is a space-time white noise, and denotes the smoothing operator
(1.2) |
In the following, we consider (1.1) as a first-order system in the variable ,
(1.3) |
The interest for the equation stemmed from recent results pertaining to the PDE construction of the Euclidean quantum field theory [11, 1, 2, 23, 24, 33]. Namely, if one considers the stochastic quantisation equation
(1.4) |
as time goes to infinity, we expect a generic solution to (1.4) to converge to the (massive) -measure.
In the above context, the stochastic wave equation
(1.5) |
corresponds to the so-called canonical stochastic quantisation equation, or equivalently, to the kinetic Langevin equation for the -measure with momentum . Therefore, one can sample the -measure by sampling the law of the first component of the solution of (1.5). This procedure has the name of Hamiltonian Montecarlo (HCM). It has numerically been observed that in the finite dimensional setting, HCM converges faster than MCMC (Markov Chain Montecarlo) in many situations, see [14] (and references within), for a rigorous justification of these faster convergence rates.
As a consequence, equation (1.5) has received a substantial amount of attention in recent years, as an alternative approach to stochastic quantisation of the measure. However, due to the worse analytical properties of the wave propagator compared to those of the heat propagator , there are far fewer results available, namely, uniqueness and convergence to the invariant measure have been proven by the second author only in [51] for dimension and in [52] for dimension . Local and global well posedness results for related models, but without uniqueness of the measure, can be found in [25, 26, 50, 35, 36, 42].
In light of the above discussion, we see the equation (1.1) as a version of (1.5) in fractional dimension . A similar approach to fractional dimension has already been considered for the parabolic case of (1.4), see for instance [9]. In the work by the authors [18], it has been shown that equation (1.1) also has a unique invariant measure, to which we have convergence for every initial data belonging to a certain class. However, strictly speaking, the equation (1.1) is not a Langevin equation, because of the presence of the smoothing operator . This prevents us from writing an explicit formula for the invariant measure, which in turn leaves many open questions. A fairly natural question to ask is the following. Denoting by the invariant measure for the linear equation
or more precisely, to the system
is it true that the invariant measure of (1.1) satisfies ? As opposed to the invariant measure for the nonlinear equation, the measure has an explicit formula, namely, it formally holds that
(1.6) |
Therefore, this absolute continuity property would allow us to bootstrap relevant properties of the invariant measure from the analogous properties of . However, while the analogous of this statement is relatively simple to show for the parabolic equation (1.4), for the wave case, it turns out to be a very difficult problem. Indeed, even in the case of the stochastic Navier-Stokes equation on ,
(1.7) |
this question is still open since the original proof of the existence of the invariant measure in [16, 15]. Indeed, the current state-of-the art is that one can show that the invariant measure is absolutely continuous with respect to the stationary law of the linear solution when is replaced with for some , see [32].
A closely related question is the following: if is it true that ? If this happens, we say that the measure is quasi-invariant under the (stochastic) flow of (1.3). In the parabolic case of (1.4) and (1.7), or more in general, whenever one has the strong Feller property, the two questions are actually equivalent. Unfortunately, the strong Feller property fails for equation (1.3), and one cannot deduce absolute continuity of the invariant measure from quasi-invariance.
Quasi-invariance for the deterministic flow of Hamiltonian and dispersive PDEs in situations where the initial data is not distributed according to an invariant measure has been studied extensively in the last decade. This pursuit was initiated by Tzvetkov in [53] for the BBM equation, with many results following for a number of other Hamiltonian PDEs, see [53, 40, 41, 38, 39, 43, 20, 28, 45, 13, 21, 22, 44, 7, 17, 37, 19, 48, 10, 31] and references therein for a list of results in this direction. We point out the results in [41, 28, 45] prove quasi-invariance results in the case of the deterministic and undamped version of (1.1).
Nevertheless, the question of quasi-invariance for stochastic dispersive models is yet to be answered. In this paper, we prove the first of such results; in particular, we show the following.
Theorem 1.1.
For each , let denote the solution to SDNLW 1.1 with . Then for each , the probability measure is absolutely continuous with respect to .
1.1. Outline of the proof
The proof of Theorem 1.1 will follow closely the strategy of the deterministic results in [41, 28, 45], hence we first describe how one obtains quasi-invariance of the measure for the deterministic equation
(1.8) |
One can formally see this equation as an infinite-dimensional system of ODEs in the form
with on some infinite-dimensional vector space . Therefore, denoting by the flow of this system of ODEs, for a measure and for any Borel set , one has that
This allows to perform a Gronwall-type argument for the quantity , from which one obtains that quasi-invariance holds whenever for every ball , there exists some such that
(1.9) |
While not codified in this way, one can find a proof of this result in [43]. As the condition (1.9) does not hold for the measures , what is shown in [41, 28, 45] is that, for an appropriate range of , if one defines the energy functional
with , then
and the property (1.9) holds.111The expression above must be considered formally, as . Nevertheless, one can define the measure via a renormalisation procedure.
In order to prove Theorem 1.1, we follow a similar approach in the stochastic setting. We define the measure
and we try to show that under a condition analogous to (1.9), we can deduce quasi-invariance of the measure in the sense of Theorem 1.1. Indeed, in Section 5, we first prove that the measure satisfies the alternative condition
(1.10) |
and then we prove that this condition can indeed replace (1.9) in the stochastic setting, thus completing the proof. While the argument is specialised to equation (1.1), we believe that a similar condition can be proven more in general.
1.2. Further remarks
Remark 1.2.
In the proof of the condition (1.10), we follow a slightly different approach from the cited works [41, 45]. This allows us to prove that the estimate holds in the full non-singular regime . When applied to the deterministic wave equation (1.8), this allows us to show the estimate (1.9), and thus quasi-invariance of , in the regime , thus improving the result in [45], which instead requires .
Remark 1.3.
In the deterministic case (1.8), it is in principle possible to write an explicit formula for the Radon-Nikodym derivative . While this formula has not been used in the context of wave equations, for other deterministic models, it allowed to improve quasi-invariance results. See [13, 17, 19, 21, 22, 10, 31] for some examples. However, in the stochastic case, we do not have such an explicit formula for the density, as we cannot solve the Fokker-Plank equation (5.5) explicitly.
Remark 1.4.
A very recent result by Hairer-Kusuoka-Nagoji [29] shows the reverse of Theorem 1.1 for a wide class of parabolic equations, i.e. that under certain assumptions, the invariant measure for singular parabolic SPDEs is singular with respect to the law of the stationary linear solution. Interestingly, the main assumption of their result, i.e. the fact that does not belong to the Cameron-Martin space for , also applies to equation (1.1). Nevertheless, the result of this paper shows that the measure is quasi-invariant. The discrepancy between these results is due to the following.
-
-
The result in [29] heavily relies on the fact that the nonlinearity of the equation is not an function when computed in the linear solution, and
-
-
The dispersive/Hamiltonian structure of the wave equation introduces further cancellations that are not present in the parabolic case.
2. Preliminaries
Notations. For a function , its Fourier series is given by , where
Moreover, for a function , the expression will denote the distribution with Fourier coefficients given by
For , we denote . Given , we define as the sharp projection onto frequencies in the square , that is,
where for , , and similarly, when is vector valued,
We also define and for , extend these projections to and . As is the Fourier restriction to a cube with sides of length , we can rewrite as the composition of the Hilbert transform on each of the variables. Thus, from the boundedness of the Hilbert transform on for , we obtain
uniformly in the parameter and for . When , we set .
Given a vector , we use and to denote the projection onto the first and second components of , respectively. Namely, and .
2.1. Function spaces
We define the Sobolev spaces and via the norms
with the usual modification when . Moreover, when , we write . With this definition and (2.38), it holds that for every ,
(2.1) |
Following [18], for , we define the space
with the norm
(2.2) |
We define to be the closure of trigonometric polynomials in under the norm , which makes into a Polish space. Note that for any ,
(2.3) |
Moreover the following embeddings hold (see [18]):
When , we simply define , and, when , the embedding implies that is an algebra.
Similar to the spaces and , given , we let
and define to be the closure of trigonometric polynomials in under the norm
(2.4) |
Now, is a Polish space (see [51, Lemma 1.2]) and .
We define Littlewood-Payley projections as follows. Let be a smooth bump function supported on and on . For , we set and for . For , we define the Littlewood-Payley projector as the Fourier multiplier operator with symbol
where . This gives a decomposition of a distribution
We recall the definition of the Besov spaces which are defined by norm:
(2.5) |
We have for any , and we define the Besov-Hölder spaces . We also define vector versions of these spaces as with the norm
The Besov spaces have the following properties. See for instance [3].
Lemma 2.1.
The following estimates hold.
(i) (interpolation) For and , it holds that
(2.6) |
(ii) (embeddings) Let and . Then,
(2.7) |
(iii) (algebra property) For any ,
(2.8) |
(iv) (duality) Let and such that . Then,
(2.9) |
(iv) (fractional Leibniz rule) Let such that . Then, for all , we have
(2.10) |
(iv) (Sobolev embedding) Let , , and . Then,
(2.11) |
Given two functions and on , we can expand the product as
(2.12) |
We will use the shorthand for and for .
Lemma 2.2.
Let and such that . Then, we have
(2.13) |
When , we have
(2.14) |
When , we have
(2.15) |
In particular, if and , the mapping extends to a continuous bilinear map from to .
See [3] for proofs in the non-periodic case. We then have some useful consequences.
Lemma 2.3.
Let and . Then,
(2.16) | |||
(2.17) |
2.2. A commutator estimate
Our main goal in this subsection is to give a proof of the following commutator estimate.
Proposition 2.4.
For any and , it holds that
(2.19) |
Whilst the estimate (2.19) is far from optimal, it will be sufficient for our purposes as we only require the regularities on the right-hand side of (2.19) to be strictly less than . We need two preliminary results before we can give a proof of (2.19).
Lemma 2.5.
For any and , it holds that
(2.20) |
Proof.
We have
so that
(2.21) | ||||
(2.22) | ||||
(2.23) |
We provide details only for estimating (2.21) as the (2.22) and (2.23) follow from similar ideas exploiting the existence of at least two similar and large frequencies. We write
(2.24) |
Let . For the first piece on the right-hand side of (2.24), we have
(2.25) |
Using the characterisation
we see that this contribution is controlled provided that , enforcing . For the second piece on the right hand side of (2.24), we further write it as
Again, we have at least that in each of these terms, and we can proceed similar to (2.25) provided that . This completes the proof. ∎
Lemma 2.6.
Given and , there exists a multilinear operator such that
(2.26) |
and satisfies
Lemma 2.6 is well known in the case of ; see the more general result in [3, Theorem 2.92]. For an argument in the periodic setting, we refer to [45, Lemma 13].
Proof of Proposition 2.4.
We begin by writing
(2.27) | ||||
(2.28) |
By the mean-value theorem,
(2.29) |
If , then (2.29) implies
(2.30) |
Otherwise, if , (2.10) implies
(2.31) |
for any . We also have by (2.29) that
(2.32) |
Now (2.30), (2.31), and (2.32) give
For (2.27), we write
(2.33) | ||||
(2.34) | ||||
(2.35) |
First, we have
Using (2.15) and (2.8), we have
We obtain the same bound for by using (2.14). Thus,
For (2.33), we use (2.20). It remains to control (2.34) for which we use Lemma 2.6. For the remainder term, we obtain the bound , after using (2.8). For the first term, we fix , and use (2.13) and (2.8),
Now, if , taking we may sum over the dyadics. If , we need . ∎
2.3. The white noise and stochastic convolution
The space-time white noise is the unique (in law) random distribution such that is a family of centred Gaussian random variables on a probability space with the property that
(2.36) |
for any . For , we set
and we denote by the usual augmentation of the filtration (see [47, p. 45]).
We define the stochastic convolution in a pathwise manner as the space-time distribution satisfying
(2.37) |
for all test functions . Here, is the linear propagator for the (damped) wave equation, and it is given by the formula
(2.38) |
where . For smooth , the definition (2.38) corresponds exactly to
(2.39) |
It enjoys the following regularity property.
Lemma 2.7.
Let and . Then, almost surely. Moreover, for every ,
(2.40) |
where .
Before we give the proof, we first recall a quantitative version of the Kolmogorov continuity criterion, which is a slight modification of [30, Lemma 2.4].
Lemma 2.8.
Given , , , , a Banach space , and a process that is almost surely continuous,
(2.41) |
Proof of Lemma 2.7.
The almost sure properties and the moment bound (2.40) when were proved in [18, Proposition 2.7 and 2.8]. Note that this does not automatically imply (2.40), since the operator is not bounded on . However, the estimate immediately follows form the arguments in [18, Proposition 2.8], so here we highlight the main modifications to the argument. For (2.40), a telescoping series argument and Minkowski’s inequality shows that it suffices to prove that
(2.42) | ||||
(2.43) |
for some , and every .
For shorthand, we define . To control the supremum over in (2.43), we decompose into unit intervals and then apply Lemma 2.8 on each interval. We only discuss the first term on the right-hand side of (2.41). The estimates required for the time-differences term in (2.41) follow similarly using the strategy below and the estimates in the proof of [18, Proposition 2.7], up to the modification we give below. Fix . Recalling the definition of the -norm in (2.4), for finite , Minkowski’s inequality implies
To control the supremum over , we again need to use Lemma 2.8. We again only discuss the contribution from the first term from (2.41). By Sobolev embedding and the Gaussianity of , we have
where . Using (2.37) and the frequency localisations from the difference in , which forces , we bound this by
which gives (2.43), as long as . Finally, (2.42) follows similarly. ∎
3. The flow for (1.1)
3.1. Review of well-posedness
In this section, we recall elements from [18] of the notion of solution and associated well-posedness for the equation (1.1). We view (1.1) in the following integral formulation:
(3.1) |
where denotes the projection to the first coordinate. Motivated by (3.1) and these considerations about the stochastic convolution (2.39), we define a solution of in terms of the first-order expansion [34, 6, 11]:
(3.2) |
where solves the equation
(3.3) |
We note that the equation (3.3) is chosen to be the mild formulation of the differential equation
which we obtain by inserting the ansatz (3.2) into the equation (1.1). We recall the well-posedness statement for (3.3) below in Proposition 3.1.
In order to make the arguments in this paper rigorous, we require approximating finite dimensional versions of (1.1), which we describe now. Given , we define the truncated system of (1.3) by
(3.4) |
Solutions to (3.4) naturally decompose as:
where the low-frequency part solves
(3.5) |
and the high-frequency part solves the linear damped stochastic wave equation:
(3.6) |
We write solutions to (3.5) as
(3.7) |
where
(3.8) |
is the solution map for the linear stochastic damped NLW, and solves
(3.9) |
We will occasionally use the fact that without explicit reference. We have the following global well-posedness results for (1.1) and (3.9).
Proposition 3.1.
[18, Section 3] Let , , , , , and . Then, the equations
(3.10) | ||||
(3.11) |
are almost surely globally-well posed in . Moreover, the following hold true:
(i) With denoting the solution to (3.11) with , for every and for , it holds that
(3.12) |
(ii) For every , we have
The existence and uniqueness of and is guaranteed by an application of Banach fixed point theorem, which uses only the fractional Leibniz rule and Sobolev embedding. The global bound (3.12) follows from a Gronwall argument on an energy quantity resembling that of the Hamiltonian for the undamped, non-stochastic NLW; see [18, (3.12)]. The a priori bound (3.12) is only of interest when . When , we still have global well-posedness for by a Gronwall argument with the energy
(3.13) |
We omit the details. See for example [8].
We remark that by regularity counting arguments, we expect to be more regular than and to belong to . This additional regularity can be justified a posteriori from Proposition 3.1, and we will make use of it later; see Lemma 4.5.
Recall that for, and , we have
(3.14) |
where has the decomposition (3.7) and solves the linear equation (3.6). When , we write
(3.15) |
where solves (3.10). We have the following convergence statement for the nonlinear flows.
Lemma 3.2.
Let , , , and . Then, converges to almost surely in as .
3.2. Markov semigroups
In this section, we recall how the flows (3.14) and (3.15) generate Markov semigroups on . We denote by the space of measurable, bounded functions from to , and by the space of bounded, continuous functions from to , both endowed with the norm
Given and , we define the family of bounded linear operators , by
(3.16) |
for every , and . We recall from [18, Section 4], that defines a Markov semigroup. The same is true for the operators , , defined using the truncated flow. We state the main properties of the semigroups . For the proof, see [18, Proposition 4.1, Lemma 4.2, and Proposition 4.3].
Lemma 3.3.
Let . Then, the following hold true:
(i) Let . Then the map defined by is adapted with respect to the filtration . In particular, for every ,
(3.17) |
(ii)For a.e. realisation of , and for every , we have that
(3.18) |
where is the (right) time translation by , which is defined on space-time distributions as an extension of the map for .
(iii) is a Markov process and is a Markov semigroup.
4. Modified measures and sample path properties
4.1. Modified measures
The Gaussian measure in (1.6) can be formally written as
where
(4.1) |
Rigorously, we define , where denote the following pair of random Fourier series:
where and are families of standard complex-valued Gaussian random variables conditioned so that and . More precisely, we define the index set
and let be standard complex valued Gaussian random variables, with real-valued, and then define , for .
Given , we define the marginals
(4.2) | ||||
(4.3) |
By independence, we then have the decomposition
(4.4) |
We let be the real vector space
where , and put the usual scalar product on . We endow with a Lebesgue measure as follows: with
we put , , and so that
(4.5) |
We then define as the Lebesgue measure on with respect to the orthogonal basis
For every , the Gaussian measure is invariant under the dynamics of the high-frequency equation (3.6). Indeed, a tedious but straightforward computation shows that
for all , and since evolves linearly, Gaussianity is preserved.
As discussed in [41, 28], it is not clear how to study the quasi-invariance of the measure in (1.6) under the nonlinear flow of (1.1) directly. Indeed, a renormalisation is needed to make sense of the time evolution of in (4.1). Thus, following [28], we define
(4.6) |
Then, we set
(4.7) |
We now define a modified version of (4.1) by
(4.8) |
where
(4.9) |
It follows from Parseval’s theorem that
(4.10) |
where
(4.11) |
where , and is the energy in (3.13).
The following lemma justifies the construction of the modified measures.
Lemma 4.1.
Let and . Then, there exists such that in as . Moreover, there exists such that
(4.12) |
Furthermore,
(4.13) |
and thus, for any , the restriction to of the measures
(4.14) |
converge to a measure
(4.15) |
in total variation, where are defined so that and are probability measures. In particular, and are equivalent; that is, and vice versa.
The convergence of was essentially proved in [41, Lemma 3.4]. The uniform bound (4.12) is possible in view of the positivity of the term in (4.9). Our proof of Lemma 4.1 is by now a standard application of the Boué-Dupuis variational formula [5, 54, 1]. However, to our knowledge a proof for the whole range does not appear in the literature, so for the reader’s convenience, we will present some of the details in Section 4.3.
4.2. Pathwise properties
The goal of this section is to establish some properties that sample paths of and satisfy, building up to the uniform bounds on the truncated flows in Lemma 4.4 and Lemma 4.5. We begin with samples of , which enjoy the following regularity property.
Lemma 4.2.
Let . For , we have almost surely. Moreover,
(4.18) |
Proof.
Clearly, the membership in is stronger than that in , so we focus on the former. Using Sobolev embedding for some large enough so that , with , Lemma 2.8, we have
for some to chosen later. It will suffice to control the norms of just the first components of the vectors. For each fixed , is a mean zero Gaussian random variable with a variance bounded by
(4.19) |
By Minkowski’s inequality, the Gaussianity of and (4.19),
The factor ensures the summability of the first term in . For the second term we argue similarly using the mean value theorem to see that for each fixed , is a mean zero Gaussian random variable with variance bounded by
provided that . Thus,
provided that which is possible by taking sufficiently large and sufficiently small at the beginning. We then obtain
As for (4.18), this follows in a similar way as for (2.40). In particular, by (3.8) and the triangle inequality, we only need to prove the bound for , which further reduces to showing
(4.20) | ||||
(4.21) |
for some , and every . Here, (4.20) just follows from Sobolev embedding and (2.1). Now, (4.21) follows similar arguments as above in showing , and we gain the factor from localisation to frequencies around in the sum (4.19). ∎
As our arguments in Section 5.3 require us to also control norms of the time evolution of the function , we need to be more precise here about this quantity. For , we define
(4.22) |
in , whenever this limit exists.
Lemma 4.3.
Let but sufficiently close to . Suppose that is such that exists in the sense of (4.22), and let . Then, exists in the sense of (4.22) and satisfies
(4.23) |
In particular, if is distributed according to , then almost surely exists, and moreover, for any , and in . Finally, for distributed according to , exists almost surely and satisfies
(4.24) |
for any , , and where is the linear flow defined in (3.7).
Proof.
For each fixed , it follows by the definition (4.7) that
(4.25) |
Then, the existence of is ensured by verifying that the right hand side of (4.25) converges as . The term converges by our assumption on . For the mixed term , we use (2.17) which is controlled provided that , which holds as long as is sufficiently close to . Similarly, converges to in , as long as .
The second claim regarding functions distributed according to is essentially shown in [28, (4.6), Proposition 4.3] and the required details are already encapsulated in showing (4.24), which we move onto now.
By a telescoping series argument, it is enough to show that
(4.26) | ||||
(4.27) |
for and some . In particular, (4.27) also establishes that the sequence is Cauchy in this sense and thus exists in . We only consider (4.27) as simpler arguments suffice for (4.26).
By Lemma 2.8, we have
(4.28) |
where , for some , and
For each fixed , the law of is , so then
which is acceptable as . The Wiener Chaos estimate (see [46, Theorem I.22] and [49, Proposition 2.4]) then implies a similar result for any finite . Moreover, by the Wiener Chaos estimate, for the second term in (4.28), it is enough to show that
(4.29) |
where and for some . By expanding the definition of the norm, (4.29) follows from
(4.30) |
for any . We compute
(4.31) |
where
Then, taking the difference of the first term in (4.31) at times , with the absolute value squared, and using the mean value theorem, we find
(4.32) |
provided that . When , we need to use that fact that for each fixed and , is a standard complex Gaussian random variable so
Then, (4.30) follows in this case since
and thus
for any . This completes the proof of (4.30) and thus also (4.29). ∎
We now show that (4.18) and (4.24) also hold for the nonlinear flow of (1.1). To this end, we need to exploit that belongs to the higher regularity space for any , namely, one full degree better than that of the initial data and the stochastic convolution . It is convenient to verify this a posteriori relative to the results from Proposition 3.1. In particular, the following results hold pathwise.
Lemma 4.4.
Let , , and . Then, for every , , and whenever
(4.33) |
for some , there exists such that
(4.34) |
Proof.
We now obtain a uniform control on the transport of the renormalised quantities .
Lemma 4.5.
Let such that . Given and , suppose that
Then, there exists a constant such that
(4.36) |
4.3. Proof of Lemma 4.1
First, we note that (4.13) follows from (4.12) and the convergence in measure from of (4.9), which in turn follows from immediately from Lemma 4.3. We thus focus on establishing (4.12). We set as the argument is identical for any finite . We now state the Boué-Dupuis variational formula [5, 54]; the current version can be found in [19, Lemma 2.6].
Lemma 4.6.
Let be distributed according to and . Suppose that is measurable such that for some , where denotes the negative part of . Then, we have
(4.37) |
and the expectation is an expectation with respect to the underlying probability measure .
In the following, we simply write for and for .
Lemma 4.7.
Let . Then, given any finite , we have
for any , and uniformly in
Proof of Lemma 4.1 (4.12).
In view of (4.16), we apply Lemma 4.6 with . Note that the assumptions in Lemma 4.6 are trivially satisfied as the bounds may depend on . From (4.37), we have
(4.38) |
By Hölder and Young’s inequalities, we have
(4.39) |
Thus, (4.39) implies
(4.40) |
Next, by (2.9), (2.10), (2.7), for every , there exists such that
(4.41) |
By Hölder and Young’s inequalities, we also have
(4.42) |
for any . Thus, by (4.41) and (4.42), we have
(4.43) |
where is polynomial in its arguments. Therefore Lemma 4.7 implies
Combining (4.40) and (4.43) in (4.38) and choosing sufficiently small, we have shown
uniformly in . This completes the proof of (4.12). ∎
5. On the transported densities
5.1. Functional derivatives
For vector-valued functions and , we define an inner product on by
and we denote the Fréchet derivatives as follows: if , then
and, if is also , then we define the second functional derivative
and the Hessian can be further expanded as
In the finite dimensional setting, we can relate the functional derivatives to the derivatives with respect to the Fourier coefficients of as in (4.5). In particular, by the chain rule, we have
so that
5.2. Transported densities
In the following, we fix and consider the truncated system (3.5). Using , (3.5) is then a finite dimensional system of SDEs for the Fourier coefficients . We use (4.5) as a bijection between and the coefficients . We write (3.5) in the Ito formulation:
(5.1) |
where
and is the gradient with respect to the variables . With respect to , we associate to (5.1) the (adjoint of the) infinitesimal generator:
(5.2) |
As , we rewrite (5.2) as
(5.3) |
where
(5.4) |
and
Note that
Let denote the almost sure global-in-time solution to (5.1) with initial data , as guaranteed by Proposition 3.1, and recall the decomposition
These solutions generate a Markov semigroup
and the push-forward measure is a weak solution to the Fokker-Planck equation:
(5.5) |
Lemma 5.1.
For each , there exists such that , and satisfies
(5.6) |
and
(5.7) |
Proof.
As the coefficients of the operator are smooth and the initial datum is smooth, Hörmander’s condition ensures that and . The first equality in (5.6) follows from (5.5) and the second equality follows from (5.3), (5.4), (4.10), and the fact that . In particular, we have
Finally, for (5.7), a direct computation from (4.11) shows that
and from (3.13), we have
We define
(5.8) |
which is the transported density of the truncated probability measure
(5.9) |
We extend to all of by defining . In the following key result, we show that is the transported density of the measure with respect to as well as obtain good long-time -bounds for when restricted to compact sets.
Lemma 5.2.
Let and . Then, it holds that
(5.10) |
Moreover, for such that and a compact set, there exists such that
(5.11) |
for all and . Furthermore, given , let
(5.12) |
and . Then, is pre-compact in and
(5.13) |
for all , and where the constants do not depend on .
Proof.
We first show (5.10). Recalling (4.14) and (4.4), we have
(5.14) |
Let and be measurable sets. Then, as and are independent and, for fixed , and depend only on and , respectively, the random variables and are independent. Thus, using (5.14) and the invariance of under the linear high-frequency flow , we obtain
As this holds for all such sets and , we conclude (5.14).
We now move onto the latter claims in the lemma. By Lemma 5.1 and (5.8), and satisfies and . As is uniformly continuous on , there exists such that for all , and ,
(5.15) |
For , and , Taylor’s theorem implies
We move onto the claims regarding the set . Since the norm controls the Fourier coefficients of , is bounded, and hence pre-compact. We now show the bound (5.13). Let . From (5.11), the inequality for any , (4.12), (4.4) and Cauchy-Schwarz, we have
(5.16) |
Recall that for , the random variable is a centered complex-valued Gaussian with variance . Moreover from (5.7), can be written as , where
Thus, from (5.7) and the independence of the random variables and , we have
It follows from (2.16), (2.19), and the definition of , that there exists such that
for . Inserting this bound back into (5.16) and using (4.17) yields (5.13). ∎
5.3. Lifted Markov semigroups
For and , we define the space
with the norm
It is clear from (2.37) that is a Polish space. We view as a probability measure on . More precisely, for any , we have
In the following proposition, we show that for each , the Markov semigroups lift to Markov semigroups on the product space .
Proposition 5.3.
Given , there exists a Markov semigroup on such that
(5.17) |
Moreover, the Radon-Nikodym derivative
(5.18) |
exists and is equal to for almost every with respect to , where was defined in (5.8).
Proof.
In order to establish the absolute continuity result, we need to argue locally in the function spaces, which necessitates keeping track of the flow. For a fixed measurable set and , we define
(5.21) |
These sets behave well with respect to the Markov semigroup in the following sense.
Lemma 5.4.
For every , and measurable, it holds that
(5.22) |
Proof.
Suppose that . Using (5.19), (5.21), and (3.18), we have
Moreover, by the inclusion
(5.23) |
for any , we have that .
Next, assume that . Then, there is a such that . If , then . If instead , then . ∎
We now obtain an analogue of Proposition 5.3 for the localised measures .
Lemma 5.5.
Let , be measurable, , and be as in (5.21). Then, the measure has a density with respect to the probability measure and we set
(5.24) |
In particular, the following properties hold:
(i) for almost every ,
(ii) for almost every ,
(iii) for almost every .
Proof.
The existence of (5.24) and consequently (i) is immediate from the existence of which we showed in Proposition 5.3. We move onto showing the properties (ii) and (iii). First, by the definition (5.24) we have
(5.25) |
for any . By choosing in (5.25) and noting that is disjoint from , we obtain
(5.26) |
Property (ii) now follows from property (i) and that .
5.4. bounds on densities
The key ingredient for the proof of Theorem 1.1 is the following long-time bounds on the truncated densities .
Proposition 5.6 (Uniform long-time bounds on densities).
Proof.
To simplify the notation, we write in place of . First, we will prove that
(5.29) |
for any such that and .
Let . Then, by (5.22), the semigroup property of , (5.23), and Hölder’s inequality,
(5.30) |
Now using the definition (5.19) and (5.22), we have
(5.31) |
for any . Using (5.31) in (5.30), we then have
Thus, by duality we establish (5.29).
We now iterate (5.29). We choose the time-step from Lemma 5.2 and break the time interval into many sub-intervals of width . We let for , and note that . For to be chosen sufficiently small, we define and the sequence recursively by
Note that for all and . Using that the sequence is decreasing, we have
Thus, by choosing
we ensure that and . In particular, . For , (5.29) implies
Therefore,
for each . As
Hölder’s inequality then implies
(5.32) |
Now recalling that we chose such that and using Lemma 5.5 (ii)-(iii), Proposition 5.3, and (5.13), we have
(5.33) |
uniformly in . Returning this bound to (5.32), establishes that
(5.34) |
uniformly in .
Now we obtain bounds for every . Let and assume that is not an integer multiple of . If , then Lemma 5.5 (ii)-(iii), Proposition 5.3, (5.13) imply
(5.35) |
uniformly in . Now we assume that . There is a (depending on ) such that . By repeating the arguments leading to (5.29), we find that
for any . We choose so that . Moreover, since , we have . It now follows from (5.34), (5.35), and Hölder’s inequality that
(5.36) |
As was arbitrary, (5.35) and (5.36) complete the proof of (5.28). ∎
5.5. Proof of Theorem 1.1
Let , be as in Lemma 5.2. Let . For , define the set
and let be sufficiently close to so that also satisfy the conditions in Lemmas 4.4, 4.5, and 5.2. Then, for any such that , Lemma 3.2, and Lemma 4.1 imply
(5.37) |
It follows by Markov’s inequality with (3.7), (2.40), (4.18), (4.24) that
(5.38) |
From Lemma 4.4 and Lemma 4.5, we see that if , then , where is a set of the same form as in (5.12) but with radius and replaced by . Consequently,
(5.39) |
It follows from (5.28) that, up to a subsequence,
converges weakly in to some element . It then follows from (5.37), (5.38), and (5.39) that
(5.40) |
By density of continuous, bounded functions in
we can extend (5.40) to hold for all . Let be a Borel set such that . Therefore, by (5.40), we obtain
for every . By taking , we obtain that . Since this holds for every with , we deduce that . This completes the proof of Theorem 1.1.
Acknowledgements.
J. F. acknowledges support from Tadahiro Oh’s ERC consolidator grant (grant no. 864138 “SingStochDispDyn”).
References
- [1] N. Barashkov, M. Gubinelli, A variational method for , Duke Math. J. 169 (2020), no. 17, 3339–3415.
- [2] N. Barashkov, M. Gubinelli, The measure via Girsanov’s theorem, Electron. J. Probab. 26(2021), Paper No. 81, 29 pp.
- [3] H. Bahouri, J.Y. Chemin, R. Danchin, Fourier analysis and nonlinear partial differential equations, Vol. 343. Springer Science & Business Media, 2011.
- [4] V. I. Bogachev, N. V. Krylov, M. Röckner, S. V. Shaposhnikov, Fokker-Planck-Kolmogorov equations, Math. Surveys Monogr., 207 American Mathematical Society, Providence, RI, 2015. xii+479 pp.
- [5] M. Boué, P. Dupuis, A variational representation for certain functionals of Brownian motion, Ann. Probab. 26 (1998), no. 4, 1641–1659.
- [6] J. Bourgain, Invariant measures for the 2D-defocusing nonlinear Schrödinger equation, Comm. Math. Phys. 176(2): 421-445 (1996).
- [7] N. Burq, L. Thomann, Almost sure scattering for the one dimensional nonlinear Schrödinger equation, Mem. Amer. Math. Soc. 296(2024), no.1480, vii+87 pp.
- [8] N. Burq, N. Tzvetkov, Random data Cauchy theory for supercritical wave equations. II. A global existence result, Invent. Math. 173 (2008), no. 3, 477–496.
- [9] A. Chandra, A. Moinat, H. Weber, A priori bounds for the equation in the full sub-critical regime, Arch. Ration. Mech. Anal.247(2023), no.3, Paper No. 48, 76 pp.
- [10] J. Coe, L. Tolomeo, Sharp quasi-invariance threshold for the cubic Szegö equation, arXiv:2404.14950 [math.AP].
- [11] G. Da Prato, A. Debussche, Two-dimensional Navier-Stokes equations driven by a space-time white noise, J. Funct. Anal. 196 (2002), no. 1, 180–210.
- [12] G. Da Prato, J. Zabczyk, Stochastic equations in infinite dimensions, Second edition. Encyclopedia of Mathematics and its Applications, 152. Cambridge University Press, Cambridge, 2014. xviii+493 pp. ISBN: 978-1-107-05584-1.
- [13] A. Debussche, Y. Tsutsumi, Quasi-invariance of Gaussian measures transported by the cubic NLS with third-order dispersion on , J. Funct. Anal. 281 (2021), no. 3, 109032, 23 pp.
- [14] A. Eberle, A. Guillin, R. Zimmer, Couplings and quantitative contraction rates for Langevin dynamics, Ann. Probab. 47 (2019), no. 4, 1982–2010.
- [15] B. Ferrario, Ergodic results for stochastic Navier-Stokes equation, Stochastics and Stochastics Reports 60 (1997), 271–288.
- [16] F. Flandoli, B. Maslowski, Ergodicity of the 2-D Navier-Stokes equation under random perturbations, Comm. Math. Phys. 172 (1995), no. 1, 119–141.
- [17] J. Forlano, K. Seong, Transport of Gaussian measures under the flow of one-dimensional fractional nonlinear Schrödinger equations, Comm. PDE, 47 (2022), no. 6, 1296–1337.
- [18] J. Forlano, L. Tolomeo, On the unique ergodicity for a class of 2 dimensional stochastic wave equations, Trans. Amer. Math. Soc., 377 (2024), 345–394.
- [19] J. Forlano, L. Tolomeo, Quasi-invariance of Gaussian measures of negative regularity for fractional nonlinear Schrödinger equations, arXiv:2205.11453 [math.AP].
- [20] J. Forlano, W. J. Trenberth, On the transport of Gaussian measures under the one-dimensional fractional nonlinear Schrödinger equations, Ann. Inst. H. Poincaré Anal. Non Linéaire 36 (2019), no. 7, 1987–2025.
- [21] G. Genovese, R. Lucá, N. Tzvetkov, Transport of Gaussian measures with exponential cut-off for Hamiltonian PDEs, J. Anal. Math.150(2023), no.2, 737–787.
- [22] G. Genovese, R. Lucá, N. Tzvetkov, Quasi-invariance of Gaussian measures for the periodic Benjamin-Ono-BBM equation, Stoch. Partial Differ. Equ. Anal. Comput.11(2023), no.2, 651–684.
- [23] M. Gubinelli, M. Hofmanov’a, Global solutions to elliptic and parabolic models in Euclidean space, Comm. Math. Phys. 368 (2019), no. 3, 1201–1266.
- [24] M. Gubinelli, M. Hofmanov’a, A PDE construction of the Euclidean quantum field theory, Comm. Math. Phys. 384 (2021), 1–75.
- [25] M. Gubinelli, H. Koch, T. Oh, Renormalization of the two-dimensional stochastic nonlinear wave equations, Trans. Amer. Math. Soc. 370 (2018), no 10, 7335–7359.
- [26] M. Gubinelli, H. Koch, T. Oh, Paracontrolled approach to the three-dimensional stochastic nonlinear wave equation with quadratic nonlinearity, J. Eur. Math. Soc. 26 (2024), no. 3, 817–874.
- [27] M. Gubinelli, H. Koch, T. Oh, L. Tolomeo, Global dynamics for the two-dimensional stochastic nonlinear wave equations, Int. Math. Res. Not. (2021), rnab084, https://doi.org/10.1093/imrn/rnab084.
- [28] T. Gunaratnam, T. Oh, N. Tzvetkov, H. Weber, Quasi-invariant Gaussian measures for the nonlinear wave equation in three dimensions, Probab. Math. Phys. 3 (2022), no. 2, 343–379.
- [29] M. Hairer, S. Kusuoka, H. Nagoji, Singularity of solutions to singular SPDEs, arXiv:2409.10037 [math.PR]
- [30] R. Killip, J. Murphy, M. Vişan, Invariance of white noise for KdV on the line, Invent. Math. 222 (2020), no. 1, 203–282.
- [31] A. Knezevitch, Transport of low regularity Gaussian measures for the 1d quintic nonlinear Schrödinger equation arXiv:2406.07116 [math.AP].
- [32] J. C. Mattingly, T. M. Suidan, The small scales of the stochastic Navier-Stokes equations under rough forcing, J. Stat. Phys. 118 (2005), no. 1–2, 343–364.
- [33] J.-C. Mourrat, H. Weber, The dynamic model comes down from infinity, Comm. Math. Phys. 356 (2017), no. 3, 673–753.
- [34] H. P. McKean, Statistical mechanics of nonlinear wave equations. IV. Cubic Schrödinger, Comm. Math. Phys. 168 (1995), no. 3, 479–491. Erratum: Statistical mechanics of nonlinear wave equations. IV. Cubic Schrödinger, Comm. Math. Phys. 173 (1995), no. 3, 675.
- [35] T. Oh, M. Okamoto, L. Tolomeo, Focusing model with a Hartree nonlinearity, to appear in Mem. Amer. Math. Soc.
- [36] T. Oh, M. Okamoto, L. Tolomeo, Stochastic quantisation of the -model, to appear in Mem. Eur. Math. Soc.
- [37] T. Oh, K. Seong, Quasi-invariant Gaussian measures for the cubic fourth order nonlinear Schrödinger equation in negative Sobolev spaces, J. Funct. Anal. 281 (2021), no. 9, 109150, 49 pp.
- [38] T. Oh, P. Sosoe, N. Tzvetkov, An optimal regularity result on the quasi-invariant Gaussian measures for the cubic fourth order nonlinear Schrödinger equation, J. Éc. polytech. Math. 5 (2018), 793–841.
- [39] T. Oh, Y. Tsutsumi, N. Tzvetkov, Quasi-invariant Gaussian measures for the cubic nonlinear Schrödinger equation with third order dispersion, C. R. Math. Acad. Sci. Paris 357 (2019), no. 4, 366–381.
- [40] T. Oh, N. Tzvetkov, Quasi-invariant Gaussian measures for the cubic fourth order nonlinear Schrödinger equation, Probab. Theory Related Fields 169 (2017), 1121–1168.
- [41] T. Oh, N. Tzvetkov, Quasi-invariant Gaussian measures for the two-dimensional defocusing cubic nonlinear wave equation, J. Eur. Math. Soc. 22 (2020), no. 6, 1785–1826.
- [42] T. Oh, Y. Wang, Y. Zine, Three-dimensional stochastic cubic nonlinear wave equation with almost space-time white noise, Stoch. Partial Differ. Equ. Anal. Comput.10(2022), no.3, 898–963.
- [43] F. Planchon, N. Tzvetkov, N. Visciglia, Transport of Gaussian measures by the flow of the nonlinear Schrödinger equation, Math. Ann. 378 (2020), no. 1-2, 389–423.
- [44] F. Planchon, N. Tzvetkov, N. Visciglia, Modified energies for the periodic generalized KdV equation and applications, Ann. Inst. H. Poincaré C Anal. Non Linéaire 40 (2023), no.4, 863–917.
- [45] P. Sosoe, W. Trenberth, T. Xian, Quasi-invariance of fractional Gaussian fields by the nonlinear wave equation with polynomial nonlinearity, Differential Integral Equations 33 (2020), no. 7-8, 393–430.
- [46] B. Simon, The Euclidean (quantum) field theory, Princeton Series in Physics. Princeton University Press, Princeton, N.J., 1974. xx+392 pp
- [47] D. Revuz, M. Yor, Continuous martingales and Brownian motion. Third edition, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], 293. Springer-Verlag, Berlin, 1999. xiv+602 pp. ISBN: 3-540-64325-7.
- [48] C. Sun, N. Tzvetkov, Quasi-invariance of Gaussian measures for the 3d energy critical nonlinear Schrödinger equation, arXiv:2308.12758 [math.AP].
- [49] L. Thomann, N. Tzvetkov, Gibbs measure for the periodic derivative nonlinear Schrödinger equation, Nonlinearity 23 (2010), no. 11, 2771–2791.
- [50] L. Tolomeo, Global well-posedness of the two-dimensional stochastic nonlinear wave equation on an unbounded domain, Ann. Probab. 49(3): 1402-1426 (May 2021).
- [51] L. Tolomeo, Unique ergodicity for a class of stochastic hyperbolic equations with additive space-time white noise, Comm. Math. Phys. 377 (2020), no. 2, 1311–1347.
- [52] L. Tolomeo, Ergodicity for the hyperbolic -model, arXiv:2310.02190 [math.PR].
- [53] N. Tzvetkov, Quasi-invariant Gaussian measures for one dimensional Hamiltonian PDE’s, Forum Math. Sigma 3 (2015), e28, 35 pp.
- [54] A. Üstünel, Variational calculation of Laplace transforms via entropy on Wiener space and applications, J. Funct. Anal. 267 (2014), no. 8, 3058–3083.