Energy solutions of KPZ are unique

By Massimiliano Gubinelli and Nicolas Perkowski

Abstract

The Kardar–Parisi–Zhang (KPZ) equation is conjectured to universally describe the fluctuations of weakly asymmetric interface growth. Here we provide the first intrinsic well-posedness result for the stationary KPZ equation on the real line by showing that its energy solutions, as introduced by Gonçalves and Jara in 2010 and refined by Gubinelli and Jara, are unique. This is the first time that a singular stochastic PDE can be tackled using probabilistic methods, and the combination of the convergence results of the first work and many follow-up papers with our uniqueness proof establishes the weak KPZ universality conjecture for a wide class of models. Our proof builds on an observation of Funaki and Quastel from 2015, and a remarkable consequence is that the energy solution to the KPZ equation is not equal to the Cole–Hopf solution, but it involves an additional drift .

1. Introduction

The aim of this paper is to establish the well-posedness of the martingale problem for the stationary conservative stochastic Burgers equation (SBE) on ,

where is a continuous process in taking values in the space of (Schwartz) distributions over , , , and is a cylindrical Wiener process such that is a space-time white noise. A direct consequence will be the well-posedness of the martingale problem for the quasi-stationary Kardar–Parisi–Zhang (KPZ) equation

where is a continuous process, and for any the law of is a two-sided Brownian motion on . The SBE describes the evolution of the weak derivative of the solution to the KPZ equation . Our uniqueness proof also establishes that is related to the solution of the linear multiplicative stochastic heat equation (SHE),

by the Cole–Hopf transformation

The SHE allows a formulation via standard Itô calculus and martingale, weak, or mild solutions in suitable weighted spaces of continuous adapted processes. The SBE and the KPZ equation, on the other hand, cannot be studied in standard spaces due to the fact that the nonlinearity is ill-defined, essentially because the trajectories of the solutions do not possess enough spatial regularity. Indeed, solutions of the KPZ equation are of Hölder regularity less than in space, so a priori the pointwise square of their derivatives cannot be defined.

Despite this mathematical difficulty, the KPZ equation is expected to be a faithful description of the large scale properties of one-dimensional growth phenomena. This was the original motivation which led Kardar, Parisi, and Zhang Reference KPZ86 to study the equation, and both experimental and theoretical physics arguments have, since then, confirmed their analysis. The rigorous study of the KPZ equation and its relation with the SHE started with the work of Bertini and Giacomin Reference BG97 on the scaling limit of the weakly asymmetric exclusion process (WASEP). Starting from this discrete Markov process on and performing a suitable space-time rescaling and recentering, they were able to prove that its density fluctuation field converges to a random field which is linked to the solution of the SHE by the Cole–Hopf transformation Equation 4. Incidentally they had to add exactly the strange drift in order to establish their result. Their work clarifies that any physically relevant notion of solution to the (still conjectural) equations Equation 1 and Equation 2 needs to be transformed to the SHE by the Cole–Hopf transformation and also that the SBE should allow the law of the space white noise as invariant measure. A priori these insights are of little help in formulating the SBE/KPZ equation, since given a solution to the SHE it is not possible to apply Itô’s formula to , and in particular the inverse Cole–Hopf transformation is ill-defined. It should be noted that the main difficulty of equations Equation 1 and Equation 2 lies in the spatial irregularity and that no useful martingales in the space variable are known, a fact which prevents an analysis via Itô’s stochastic integration theory. Moreover, the convergence result of Reference BG97 relies strongly on the particular structure of the WASEP and does not have many generalizations because most models behave quite badly under exponentiation (Cole–Hopf transformation); see Reference DT16Reference CT17Reference CST16Reference Lab17 for examples of models that do admit a useful Cole–Hopf transformation.

After the work of Bertini and Giacomin, there have been various attempts to study the SBE via Gaussian analysis tools taking into account the necessary invariance of the space white noise. A possible definition based on the Wick renormalized product associated to the driving space-time white noise has been ruled out because it lacks the properties expected from the physical solution Reference Cha00. Assing Reference Ass02 has been the first, to our knowledge, to attempt a martingale problem formulation of the SBE. He defines a formal infinite-dimensional generator for the process essentially as a quadratic form with dense domain, but he has not been able to prove its closability. The singular drift, which is ill-defined pointwise, make sense as a distribution on the Gaussian Hilbert space associated to the space white noise, however this distributional nature prevents the identification of a suitable domain for the formal generator.

The martingale problem approach has been subsequently developed by Gonçalves and Jara Reference GJ10Reference GJ14.⁠Footnote1 Their key insight is that while the drift in Equation 1 is difficult to handle in a Markovian picture (that is, as a function on the state space of the process), it makes perfect sense in a pathwise picture. They proved in particular that a large class of particle systems (which generalize the WASEP studied by Bertini and Giacomin) has fluctuations that subsequentially converge to random fields which are solutions of a generalized martingale problem for Equation 1 where the singular nonlinear drift is a well-defined space-time distributional random field. Avoiding a description of a Markovian generator for the process, they manage to introduce an auxiliary process which plays the same role in the formulation of the martingale problem. Subsequent work of Jara and Gubinelli Reference GJ13 gave a different definition of the martingale problem via a forward-backward description. The solution of the martingale problem is a Dirichlet process, that is the sum of a martingale and a zero quadratic variation process. This property and the forward-backward decomposition of the drift are reminiscent of Lyons–Zheng processes and in general of the theory of Markov processes described by Dirichlet forms; however, a complete understanding of the matter is at the moment not well developed, and the martingale problem formulation avoids the subtleties of the Markovian setting. Gonçalves and Jara called the solutions of this generalized martingale problem energy solutions for the SBE/KPZ equation.

1

The paper Reference GJ14 is the revised published version of Reference GJ10.

Following Reference GJ14, it has been shown for a variety of models that their fluctuations subsequentially converge to energy solutions of the KPZ equation or the SBE, for example for zero range processes and kinetically constrained exclusion processes in Reference GJS15, various exclusion processes in Reference GJS17Reference FGS16Reference BGS16Reference GJ16, interacting Brownian motions in Reference DGP17, and Hairer–Quastel type SPDEs in Reference GP16. This is coherent with the conjecture that the SBE/KPZ equation describes the universal behavior of a wide class of conservative dynamics or interface growth models in the particular limit where the asymmetry is “small” (depending on the spatial scale), the so-called weak KPZ universality conjecture; see Reference Cor12Reference Qua14Reference QS15Reference Spo16. In order to fully establish the conjecture for the models above, the missing step was a proof of uniqueness of energy solutions. This question remained open for some time during which it was not clear if the notion is strong enough to guarantee uniqueness or if it is too weak to expect well-posedness. Here we present a proof of uniqueness for the refined energy solutions of Reference GJ13, on the full line and on the torus, thereby finally establishing the well-posedness of the martingale problem and its expected relation with the SHE via the Cole–Hopf transform. In fact we even prove that the equation formulated in Reference GJ13 leads to strongly unique solutions, which directly gives uniqueness in law. The reason we emphasize the (weaker) uniqueness in law is that energy solutions frequently arise as scaling limits for particle systems.

Our proof follows the strategy developed by Funaki and Quastel in Reference FQ15. Namely, we map a mollified energy solution to the SHE via the Cole–Hopf transform, and we use a version of the Boltzmann–Gibbs principle to control the various error terms arising from the transformation and to derive the relation Equation 4 in the limit as we take the mollification away. A direct corollary of our results is the proof of the weak KPZ universality conjecture for all the models in the literature which have been shown to converge to energy solutions.

Shortly after the introduction of energy solutions, the fundamental work Reference Hai13 of Hairer on the KPZ equation appeared, where he established a pathwise notion of solution using Lyons’s theory of rough paths to provide a definition of the nonlinear term as a continuous bilinear functional on a suitable Banach space of functions. Existence and uniqueness were then readily established by fixed point methods. This breakthrough developed into a general theory of singular SPDEs, Hairer’s theory of regularity structures Reference Hai14, which provides the right analytic setting to control the singular terms appearing in stochastic PDEs, such as the SBE/KPZ equations and their generalizations, but also in other important SPDEs, such as the stochastic Allen–Cahn equation in dimensions and the (generalized) parabolic Anderson model in . The work of the authors of this paper together with P. Imkeller on the use of paradifferential calculus Reference GIP15 and the work of Kupiainen based on renormalization group (RG) techniques Reference Kup16Reference KM17 opened alternative ways to tackle singular SPDEs. All these approaches have in common that they control the a priori ill-defined nonlinearities in the equation using pathwise (deterministic) arguments, and it was not clear if a probabilistic understanding of such singular SPDEs is possible at all. Our uniqueness proof for the stationary martingale solution to the KPZ equation is a first indication that it is possible and is a problem worth investigating closer—even if our method of proof does not extend to other equations because we extensively use the specific structure of the KPZ equation, including the facts that its invariant measure is Gaussian, that its nonlinearity is antisymmetric, and most restrictively that it can be mapped to the SHE through the Cole–Hopf transform.

From the point of view of the weak KPZ universality conjecture, the pathwise approach is difficult to use and, for now, there are only a few convergence results using either regularity structures, paracontrolled distributions, or RG techniques; see Reference HQ15Reference HS15Reference GP17Reference Hos16. The martingale approach has the advantage that it is easy to implement, especially starting from discrete particle systems which often do not have the semilinear structure that is at the base of the pathwise theories.

The main limitation of the martingale approach to the SBE/KPZ equation is that currently it works only at stationarity. Using tools from the theory of hydrodynamic limits, it seems possible to extend the results to initial conditions with small relative entropy with respect to the stationary measure. However, this has not been done yet and dealing with even more singular initial conditions is a completely open problem. On the other hand, with energy solutions it is relatively easy to work on the real line, while in the pathwise approach this requires dealing with weighted function spaces, and the question of uniqueness seems still not clear.

To summarize, the main contribution of the present paper is a proof of uniqueness of energy solutions (in the refined formulation of Jara and Gubinelli Reference GJ13) on the real line and on the torus. We start in section 2 by introducing the notion of solution and the space of trajectories where solutions live. Subsequently, we discuss in section 3 several key estimates available in this space, estimates which allow us to control a large class of additive functionals. After these preliminaries we show in section 4 how to implement the Cole–Hopf transformation at the level of energy solutions and, by a careful control of some error terms, how to establish the Itô formula which proves the mapping from the SBE to the SHE. Using the uniqueness for the SHE, we conclude the uniqueness of energy solutions. In Appendix A we add some detail on how to modify the proof to deal with the case of periodic boundary conditions. Appendix B contains an auxiliary moment bound for the Cole-Hopf transformation of the stochastic Burgers equation.

Notation.

We use the notation if there exists a constant , independent of the variables under consideration, such that , and we write if and . The Schwartz space on is denoted with and its dual is the space of tempered distributions. The notation refers to the distributions on , the dual space of . The Fourier transform of is denoted with , and we use the normalization . For , we use the following slightly unusual (but of course equivalent to the usual) norm for the space

Throughout this paper we work with the quadratic variation in the sense of Russo and Vallois Reference RV07: A real-valued stochastic process has quadratic variation if

where the convergence is uniform on compacts in probability. If is a continuous semimartingale, then is its semimartingale quadratic variation. Despite the fact that we deal with continuous processes, we use the notation for the quadratic variation because will be reserved for the inner products in various Hilbert spaces.

2. Controlled processes and energy solutions

2.1. Burgers equation

In this section we follow Gonçalves and Jara Reference GJ14 and Gubinelli and Jara Reference GJ13 in defining stationary energy solutions to the SBE ,

where , , and is a space-time white noise. Recall that from a probabilistic point of view the key difficulty in making sense of Equation 5 is that we expect the law of (a multiple of) the white noise on to be invariant under the dynamics, but the square of the white noise can only be defined as a Hida distribution and not as a random variable. To overcome this problem, we first introduce a class of processes which at fixed times are distributed as the white noise but for which the nonlinear term is defined as a space-time distribution. In this class of processes it then makes sense to look for solutions of the Burgers equation Equation 5.

If is a filtered probability space, then for an adapted process with trajectories in we say that is a space-time white noise on that space if for all the process is a Brownian motion in the filtration with variance for all . A (space) white noise with variance is a random variable with values in such that is a centered Gaussian process with covariance . If , we simply call a white noise. Throughout, we write for the law of the white noise on .

Definition 2.1 (Controlled process).

Let , let be a space-time white noise on the filtered probability space , and let be an -measurable space white noise. Denote with the space of pairs of adapted stochastic processes with trajectories in such that

i)

and the law of is that of a white noise with variance for all ;

ii)

for any test function , the process is almost surely of zero quadratic variation, satisfies , and the pair solves the equation

iii)

for any the time-reversed processes , satisfy

where is a space-time white noise in the filtration generated by .

If there exist a space-time white noise and a white noise such that , then we simply write . For and we omit the parameters in the notation and write respectively .

We will see that contains the probabilistically strong solution to Equation 5, while is the space in which to look for probabilistically weak solutions.

Controlled processes were first introduced in Reference GJ13 on the circle, and the definition on the real line is essentially the same. For the process is the stationary Ornstein–Uhlenbeck process. It is the unique-in-law solution to the SPDE

with initial condition . Allowing has the intuitive meaning of considering perturbations of the Ornstein–Uhlenbeck process with antisymmetric drifts of zero quadratic variation. In this sense we say that a couple is a process controlled by the Ornstein–Uhlenbeck process.

As we will see below, for controlled processes we are able to construct some interesting additive functionals. In particular, for any controlled process, the Burgers drift makes sense as a space-time distribution:

Proposition 2.2.

Let , let with , and write for . Then for all the process

converges uniformly on compacts in probability to a limiting process that we denote with

As the notation suggests, this limit does not depend on the function .

The proof will be given in section 3.3. Note that the convolution is well-defined for and not only for because at fixed times is a white noise.

Now we can define what it means for a controlled process to solve the SBE.

Definition 2.3.

Let be a space-time white noise on , let be an -measurable space white noise, and let . Then is called a strong stationary solution to the SBE

if for all . If and for all , then is called an energy solution to Equation 7.

The notion of energy solutions was introduced in a slightly weaker formulation by Gonçalves and Jara in Reference GJ14, and the formulation here is due to Reference GJ13. Note that strong stationary solutions correspond to probabilistically strong solutions, while energy solutions are probabilistically weak in the sense that we do not fix the probability space and the noise driving the equation. Our main result is that strong and weak uniqueness hold for our solutions.

Before we state this precisely, let us introduce some notation. Given with and we write for the unique solution to the linear multiplicative SHE

where . If is another function of this form, then , where is an -measurable random variable that is independent of , and in particular for all and , which implies for the distributional derivative .

Theorem 2.4.

Let be a space-time white noise on the filtered probability space and let be an -measurable space white noise. Then the strong stationary solution to

is unique up to indistinguishability. Moreover, for we have , where the derivative is taken in the distributional sense, and is the unique solution to the linear multiplicative SHE Equation 8 for an arbitrary with and . Consequently, any two energy solutions of Equation 9 have the same law.

The proof will be given in section 4.1.

Remark 2.5.

As far as we are aware, this is the first time that existence and uniqueness for an intrinsic notion of solution to the Burgers equation on the real line is obtained. The results of Reference Hai13Reference GP17Reference KM17 are restricted to the circle . In Reference HL15 the linear multiplicative heat equation is solved on the real line using regularity structures, and by the strong maximum principle of Reference CGF17 the solution is strictly positive, so in particular its logarithm is well-defined. Then it should be possible to show that the derivative of the logarithm is a modeled distribution and solves the SBE in the sense of regularity structures. However, it is not at all obvious if and in which sense this solution is unique.

Remark 2.6 (Reduction to standard parameters).

To simplify notation, we will assume from now on that and which can be achieved by a simple transformation. Indeed, it is easy to see that if and only if , where

and is a space-time white noise. Moreover, is a strong stationary solution to Equation 9 if and only if is a strong stationary solution to

By also rescaling the space variable, it would be possible to set in Equation 5. However, in the periodic setting discussed below, the rescaling of the space variable would change the size of the torus on which the solution lives, so we prefer to work with general instead.

Remark 2.7 (Martingale problem).

Energy solutions can be understood as solutions to a martingale problem. Indeed, given a pair of stochastic processes with trajectories in , we need to check the following criteria to verify that is the unique-in-law energy solution to Equation 9:

i)

The law of is that of the white noise with variance for all .

ii)

For any test function the process is almost surely of zero quadratic variation, satisfies , and the pair solves the equation

where is a continuous martingale in the filtration generated by , such that and has quadratic variation .

iii)

For any the time-reversed processes , satisfy

where is a continuous martingale in the filtration generated by , such that and has quadratic variation .

iv)

There exists such that and with we have

for all and .

Remark 2.8 (Different notions of energy solutions).

In Reference GJ14 a slightly weaker notion of energy solution was introduced. Roughly speaking, Gonçalves and Jara only make assumptions i), ii), and iv) of Remark 2.7 but do not consider the time reversal condition iii). They then show that a wide class of weakly asymmetric exclusion type processes have subsequential scaling limits that satisfy conditions i), ii), and iv). We do not know whether this weaker formulation still identifies the subsequential limits uniquely. However, the proof in Reference GJ14 actually shows that any subsequential scaling limit of their particle system also satisfies iii), even if Gonçalves and Jara do not mention this explicitly. Indeed, the time-reversal condition iii) is satisfied for the particle system, and it trivially carries over to the limit. Therefore, the combination of Reference GJ14 with our uniqueness result proves the weak KPZ universality conjecture for a wide class of particle systems.

Remark 2.9 (Relaxations).

It suffices to assume and to verify the conditions of Remark 2.7 for instead of . For our proof of uniqueness we do need to handle test functions in , but since all terms in the decomposition

come with good continuity estimates, it follows from the above conditions that and have trajectories in and satisfy the same equation also for test functions in . It is even possible to only assume that is a family of continuous adapted stochastic processes indexed by such that the conditions i)–iv) in Remark 2.7 hold up to a null set that may depend on . In that case we can find a version with values in such that for all . These relaxations may come in handy when proving the convergence of fluctuations of microscopic systems to the Burgers equation.

2.2. KPZ equation

Once we understand how to deal with the SBE, it is not difficult to also handle the KPZ equation. Let be a space-time white noise on the filtered probability space and let be an -measurable random variable with values in such that is a space white noise. Then we denote with the space of pairs of adapted stochastic processes with trajectories in which solve for all the equation

and are such that , and for which , satisfy . Similarly, we write if there exist as above such that . In Proposition 3.15 we show in fact the following generalization of Proposition 2.2: If and is an approximate identity (i.e., , for all , for all ), then there exists a process , defined by

where the convergence takes place in uniformly on compacts, and the limit does not depend on the approximate identity .

So we call a strong almost-stationary solution to the KPZ equation

if . Similarly, we call an energy solution to Equation 11 if and . The terminology “almost-stationary” comes from Reference GJ14, and it indicates that for fixed the process is always a two-sided Brownian motion; however, the distribution of may depend on time. The analogous result to Theorem 2.4 is then the following.

Theorem 2.10.

Let be a space-time white noise on and let be an -measurable random variable with values in such that is a space white noise. Then the strong almost-stationary solution to

is unique up to indistinguishability. Moreover, for we have

where is the unique solution to the linear multiplicative SHE

Consequently, any two energy solutions of Equation 12 have the same law.

The proof will be given in section 4.1.

Remark 2.11.

It is maybe somewhat surprising that the energy solution to the KPZ equation is not equal to the Cole–Hopf solution but instead we have to add the drift to to obtain . Remarkably, this drift often appears in results about the Cole–Hopf solution of the KPZ equation. For example in Reference BG97, Theorem 2.3 it has to be added to obtain the Cole–Hopf solution as the scaling limit for the fluctuations of the height profile of the weakly asymmetric exclusion process (there the drift is because Bertini and Giacomin consider different parameters for the equation). The same drift also appears in Reference ACQ11, Theorem 1.1, in the key formula Reference SS10, (4.17), and in Reference FQ15, Theorem 1.1.

In Reference GJ14, Theorem 3 it is claimed that the Cole–Hopf solution is an energy solution to the KPZ equation, and as we have seen this is not quite correct. The reason is that the proof in Reference GJ14 is based on the convergence result of Reference BG97, but they did not take the drift into account which Bertini and Giacomin had to add to obtain the Cole–Hopf solution in the limit.

Remark 2.12 (Martingale problem).

Given a pair of stochastic processes with trajectories in we need to check the following criteria to verify that is the unique-in-law energy solution to Equation 12:

i)

For all we have and

for a continuous martingale starting in and with quadratic variation .

ii)

The pair , satisfies conditions i), ii), iii) in Remark 2.7.

iii)

There exists such that and with we have

for all and .

2.3. The periodic case

It is also useful to have a theory for the periodic model , where and

for a periodic space-time white noise and a periodic space white noise . For a process with trajectories in , where are the (Schwartz) distributions on the circle, we say that is a periodic space-time white noise if for all the process is a Brownian motion with variance . A periodic space white noise is a centered Gaussian process with trajectories in , such that for all we have , where is the projection of onto the mean-zero functions. The reason for setting the zero Fourier mode of equal to zero is that the SBE is a conservation law and any solution to Equation 15 satisfies for all , and therefore shifting simply results in a shift of by the same value, for all . So for simplicity we assume . Controlled processes are defined as before, except that now we test against and all noises are replaced by their periodic counterparts. Then it is easy to adapt the proof of Proposition 3.15 to show that also in the periodic setting the Burgers drift is well-defined; alternatively, see Reference GJ13, Lemma 1. Thus, we define strong stationary solutions, respectively energy solutions, to the periodic Burgers equation exactly as in the nonperiodic setting. We then have the analogous uniqueness result to Theorem 2.4:

Theorem 2.13.

Let be a periodic space-time white noise on and let be an -measurable periodic space white noise. Then the strong stationary solution to

is unique up to indistinguishability. Moreover, for we have , where the derivative is taken in the distributional sense and is the unique solution to the linear multiplicative SHE

for , , , and where denotes the Fourier transform on . Consequently, any two energy solutions of Equation 16 have the same law.

We explain in Appendix A how to modify the arguments for the nonperiodic case in order to prove Theorem 2.13.

3. Additive functionals of controlled processes

3.1. Itô trick and Kipnis–Varadhan inequality

Our main method for controlling additive functionals of controlled processes is to write them as a sum of a forward and a backward martingale which enables us to apply martingale inequalities. For that purpose we first introduce some notation. Throughout this section we fix .

Definition 3.1.

The space of cylinder functions consists of all of the form for some , and with polynomial growth of its partial derivatives up to order .

For we define the action of the Ornstein–Uhlenbeck generator as

With the help of Itô’s formula it is easy to verify that if is the stationary Ornstein–Uhlenbeck process (see the discussion below Definition 2.1) and , then , , is a martingale and in particular is indeed the action of the generator of on . We will see in Corollary 3.8 that can be uniquely extended from to a closed unbounded operator on , also denoted by , so is a core for . We also define the Malliavin derivative

for all , and since is the law of the white noise we are in a standard Gaussian setting and is closable as an unbounded operator from to for all ; see for example Reference Nua06. Similarly also is closable from to for all , and we denote the domain of the resulting operator by . Then is the completion of with respect to the norm . So writing

we have for all . Finally, we denote

The following martingale or Itô trick is well known for Markov processes; see for example the monograph Reference KLO12, and in the case of controlled processes on it is due to Reference GJ13. The proof is in all cases essentially the same.

Proposition 3.2 (Itô trick).

Let , and . Then we have for all

For we get in particular

Proof.

We first assume that for all and that for all . Since is a Dirichlet process for all (the sum of a local martingale and a zero quadratic variation process), we can then apply the Itô formula for Dirichlet processes (see Reference RV07) to and obtain for

for a continuous martingale with and quadratic variation . Similarly, we get for

for a continuous backward martingale with and quadratic variation . Adding these two formulas, we get

and thus the Burkholder–Davis–Gundy inequality yields

where the last step follows from Minkowski’s inequality.

For a general we first approximate in by a step function that is piecewise constant in time, then we approximate each of the finitely many values that the step function takes by a cylinder function, and finally we mollify the jumps of the new step function. In that way our bound extends to all of .

Remark 3.3.

The right-hand side of Equation 18 does not involve the -norm of and indeed it is possible to extend the result to the following space. Identify all with and write for the completion of the resulting equivalence classes with respect to the norm . Then Equation 18 holds for all provided that the integral on the left-hand side in Equation 18 makes sense. But we will not need this.

Remark 3.4.

If in the setting of Proposition 3.2 has a finite chaos expansion of length for all (see section 3.2 for the definition), then also has a chaos expansion of length and therefore Gaussian hypercontractivity shows that for all

The bound in Proposition 3.2 allows us to control provided that we are able to solve the Poisson equation

for all . Note that this is an infinite-dimensional PDE which a priori is difficult to solve, but that we only need to consider it in which has a lot of structure as a Gaussian Hilbert space. We will discuss this further in section 3.2. Nonetheless, we will encounter situations where we are unable to solve the Poisson equation explicitly, and in that case we rely on the method of Kipnis and Varadhan allowing us to bound in terms of a certain variational norm of . We define for

and we write if the right-hand side is finite. For details on and , see Reference KLO12, Chapter 2.2. Here we just remark that for and we have

and on the other hand we have for all with and all , and therefore

which proves that

We will need a slightly refined version of the Kipnis–Varadhan inequality which also controls the -variation. Recall that for the -variation of is defined as

Corollary 3.5 (Kipnis–Varadhan inequality).

Let and , and let be a controlled process. Then for all

Proof.

The nonreversible version of the Kipnis–Varadhan inequality is due to Reference Wu99, and our proof is essentially the same as in Reference KLO12Reference FQ15. But we are not aware of any reference for the statement about the -variation. Note that since the integral vanishes in zero, its supremum norm can be controlled by its -variation. Let and decompose

For the first term on the right-hand side, we apply the same martingale decomposition as in the proof of the Itô trick to get . By Reference Lep76, Proposition 2 we can therefore control the -variation by

where the second inequality follows from Proposition 3.2. For the second term on the right-hand side of Equation 20, we get

and therefore overall

Now take as the solution to the resolvent equation . Note that unlike the Poisson equation, the resolvent equation is always solvable and , where is the semigroup generated by . Then and by Lemma 3.9 we have , which yields

from which we get , and then . Therefore,

and now it suffices to send .

3.2. Gaussian analysis

To turn the Itô trick or the Kipnis–Varadhan inequality into a useful bound, we must be able either to solve the Poisson equation for a given or to control the variational norm appearing in the Kipnis–Varadhan inequality. Here we discuss how to exploit the Gaussian structure of in order to do so. For details on Gaussian Hilbert spaces we refer to Reference Jan97Reference Nua06. Since is a Gaussian Hilbert space, we have the orthogonal decomposition

where is the closure in of the span of all random variables of the form , with being the th Hermite polynomial and where with . The space is called the th homogeneous chaos, and is the th inhomogeneous chaos. Also, is the law of the white noise on and therefore we can identify

where is the multiple Wiener–Itô integral of , that is

Here are the equivalence classes of that are induced by the seminorm

where denotes the set of permutations of . Of course, is a norm on and we usually identify an equivalence class in with its symmetric representative. The link between the multiple stochastic integrals and the Malliavin derivative is explained in the following partial integration-by-parts rule, which will be used for some explicit computations below.

Lemma 3.6.

Let and let be Malliavin differentiable in . Then

Proof.

The proof is virtually the same as for Reference Nua06, Lemma 1.2.1. Since the span of functions of the form is dense in , it suffices to argue for such . By polarization it suffices to consider with , for which for the th Hermite polynomial . By another approximation argument we may suppose that for orthonormal that are also orthogonal to and for . So if denotes the -dimensional standard normal distribution, then

which concludes the proof.

Recall that so far we defined the operator acting on cylinder functions. If we consider a cylinder function for some given , then the action of is particularly simple.

Lemma 3.7.

Let and be such that in we have for , the twice weakly differentiable symmetric functions from to that satisfy

Then

Proof.

Consider first a functional of the form , where with . In that case

where in the second step we used that for and . Now we use that to rewrite (see Reference Nua06, Proposition 1.1.4) and note the additional factor in our definition of compared to the one in Reference Nua06. Thus, we can apply Reference Nua06, Proposition 1.1.2 to compute the first term on the right-hand side:

where in the first term on the right-hand side denotes the Laplacian on . Plugging this back into Equation 21, we obtain . By polarization this extends to , and then to general by taking the closure of the span of functions of the form with .

Corollary 3.8.

The operator is closable in and the domain of its closure, still denoted with , is

For we have

Proof.

Let be a cylinder function with chaos expansion . By a standard approximation argument it follows that . But then Reference Nua06, formula (iii) on p. 9 yields

from which our claim readily follows because the Laplace operator on is a closed operator with domain .

Before we continue, let us link the -norm defined in section 3.1 with the operator .

Lemma 3.9.

For we have

Proof.

See Reference GP15, section 2.4 for a proof in the periodic case which works also in our setting.

Next, we define two auxiliary Hilbert spaces that will be useful in controlling additive functionals of controlled processes.

Definition 3.10.

We identify all with , and we write for the completion of the equivalence classes of with respect to .

Similarly, we identify with if and the space is defined as the completion of the equivalence classes with respect to .

Definition 3.11.

Recall that for and , the homogeneous Sobolev space is constructed in the following way: We set for

and we consider only those with , where we identify and if . The space is then the completion of the equivalence classes with respect to .

We write for the space that is obtained if we perform the same construction replacing by

Remark 3.12.

By construction, is always a Hilbert space. For there is an explicit description (see Reference BCD11, Propositions 1.34 and 1.35) for the nonsymmetric case:

Here we write for those tempered distributions with for all , where is the symmetrization of .

Lemma 3.13.

For and we have

Proof.

For the -norm it suffices to note that

where the last equality follows from Plancherel’s formula. For the -norm let us consider a test function . Then

and Plancherel’s formula and then the Cauchy–Schwarz inequality give

from which Equation 19 shows that . To see the converse inequality, let . Then we have for

and . Of course may not be in , but we can approximate it by functions in and this concludes the proof.

Corollary 3.14.

We have an explicit representation of and via

Moreover, there is a unique extension of from to for which is an isometry from to .

Proof.

We only have to prove the statement about the extension of . Since for all we have and for all . Thus,

For we have

and

which proves that and thus that is an isometry from to which can be uniquely extended to all of because is dense in .

3.3. The Burgers and KPZ nonlinearity

With the tools we have at hand, it is now straightforward to construct the KPZ nonlinearity (and in particular the Burgers nonlinearity) for all controlled processes.

Proposition 3.15.

Let , , let , and let . Then

Therefore, the integral is well-defined also for . Moreover, there exists a unique process such that for every approximate identity (i.e., , for all , for all we have

where the convergence is in , uniformly on compact subsets of . If is such that , and , then

Proof.

Let us set

where in the last step we applied the stochastic Fubini theorem. Due to infrared problems, it seems impossible to directly solve the Poisson equation , so instead we consider with which means that for

or in Fourier variables

Then we have

For the first term on the right-hand side, we further apply Gaussian hypercontractivity to estimate

where we used the completion of the square to compute

By Proposition 3.2 together with Remark 3.4 and Lemma 3.13, the second term on the right-hand side of Equation 25 is bounded by

Since , the claimed estimate Equation 23 follows.

If is an approximate identity, then converges to for all , and since and thus is uniformly bounded the convergence of follows from the above arguments together with dominated convergence theorem. If is another approximate identity, then converges pointwise to 0 and is uniformly bounded, so the independence of the limit from the approximate identity follows once more from the dominated convergence theorem.

Finally, for we can solve the Poisson equation directly (strictly speaking, we would have to first approximate by , but for simplicity we argue already in the limit ). We get for with

Let us concentrate on the first term, the second one being essentially of the same form (start by bounding in that case):

Now and , and therefore our claim follows from Proposition 3.2 together with Remark 3.4 and Lemma 3.13.

Proposition 2.2 about the Burgers drift follows by setting

Remark 3.16.

For the Ornstein–Uhlenbeck process one can check that the process has regularity . But given the bounds of Proposition 3.15, we cannot even evaluate in a point, because the Dirac delta just fails to be in . The reason is that the martingale argument on which our proof is based gives us at least regularity in time, and this prevents us from getting better space regularity. On the other hand we are able to increase the time regularity by an interpolation argument as shown in the following corollary which will be useful for controlling certain Young integrals below.

Corollary 3.17.

For all and all , the process is almost surely in and we have for all with and for all

where the convergence is in for all and we write again .

Proof.

Proposition 3.15 yields for all , and

and a direct estimate gives

The expectation on the right-hand side is

Plugging this into Equation 27, we get for all and

Now if we apply Equation 26 and Equation 28 with and we get , from which Kolmogorov’s continuity criterion yields the local Hölder-continuity of order . Moreover, for equation Equation 26 gives for all

while for we get

so that choosing small and applying once more Kolmogorov’s continuity criterion, we get the convergence of to in for all and all .

4. Proof of the main results

4.1. Mapping to the stochastic heat equation

Let be a space-time white noise on the filtered probability space and let be an -measurable space white noise. Let be a strong stationary solution to the SBE

Our aim is to show that is unique up to indistinguishability, that is to prove the first part of Theorem 2.4. The case is well understood, so from now on let . The basic strategy is to integrate in the space variable and then to exponentiate the integral and to show that the resulting process solves (a variant of) the linear SHE. However, it is not immediately obvious how to perform the integration in such a way that we obtain a useful integral process. Note that any integral of is determined uniquely by its derivative and the value for a test function with . So the idea, inspired by Reference FQ15, is to fix one such test function and to consider the integral with for all .

More concretely, we take with and and consider the function

which satisfies for any

and

so is the unique integral of which vanishes when tested against . Moreover,

and in particular

where the notation means that the -norm is taken in the variable . Now let be an even function with and such that on a neighborhood of . We write , and by our assumptions on there exists for every an such that for all . This will turn out to be convenient later. We also write and and we define

Using that is even, we get . So since is a strong stationary solution to Equation 29, we have

and Equation 31 yields

Now set . Then the Itô formula for Dirichlet processes of Reference RV07 gives

and from Equation 32 we get

Therefore, . Since moreover

we obtain

Expanding the -inner product and noting that by Equation 33, we deduce that

where we introduced the processes

for the deterministic function

and

Integrating Equation 34 against we get

In section 4.2 we will prove the following three lemmas.

Lemma 4.1.

We have for all , and all

Lemma 4.2.

The deterministic function converges to as and is uniformly bounded in the sense that .

Lemma 4.3.

For all the process converges in probability in to the zero quadratic variation process

With the help of these results it is easy to prove our main theorem.

Proof of Theorem 2.4.

Consider the expansion Equation 38 of . Denoting , the stochastic integrals converge to

by the stochastic dominated convergence theorem; see Reference RY99, Proposition IV.2.13 for a formulation in the finite-dimensional setting whose proof carries over without problems to our situation. Lemma 4.1 shows that the -variation of converges to zero in whenever . Combining this with Lemmas 4.2 and 4.3, the -variation of stays uniformly bounded in and therefore we can use once more that Lemma 4.3 gives us local convergence in for to obtain that converges as a Young integral to . In conclusion, we get

So let us define

for which and

This shows that is a weak solution to the SHE. By Lemma B.1 in Appendix B there exist such that . This allows us to extend Equation 39 on to which together with their derivatives up to order decay superexponentially. We can also extend Equation 39 to time-dependent which are such that

for all , for which we get for

see, e.g., Reference Wal86, Exercise 3.1. We apply this with for and fixed and where is the heat kernel generated by . Then , and

Since and does not depend on the space variable , it is easy to see that and have modifications which are continuous in . Therefore, as , the term on the left-hand side converges to , and the first term on the right-hand side converges to . For the stochastic integral we have

By the dominated convergence theorem, the right-hand side converges to zero, and therefore is the unique mild solution to the multiplicative SHE as defined in Walsh Reference Wal86, Theorem 3.2. Walsh considers the equation on with Neumann boundary conditions, but as Reference Qua14, Theorem 2.4 shows, the arguments carry over to as long as we have at most exponential growth of in .

So far we have only showed uniqueness on a small time interval, to extend this to, say, we need a uniform -bound of the type for all . Lemma B.1 gives us such a bound if we know that for and some . But now we know that is the unique mild solution to the SHE, and for and we have for some

Therefore, the Burkholder–Davis–Gundy inequality together with the arguments from the proof of Reference Qua14, Lemma 2.3 show that for and some . Now the uniqueness extends to , and of course we can iterate the argument to see that is the unique mild solution to the SHE on .

But we know that

where the derivative is taken in the distributional sense. Since for fixed we have and does not depend on the space variable, we get

and therefore the strong stationary solution is unique up to indistinguishability.

The uniqueness in law of energy solutions follows in the same way from the weak uniqueness of .

Proof of Theorem 2.10.

Let be a strong almost-stationary solution to the KPZ equation

Since by definition of the pair we have and , we get

Again, by definition, is a strong stationary solution to the SBE. So in the proof of Theorem 2.4 we showed that

where solves the linear multiplicative heat equation with initial condition and was defined by and

This shows that , and therefore we get with Equation 40 and Equation 41

Finally, , where solves Equation 14, the linear multiplicative heat equation with initial condition , and this concludes the proof of the strong uniqueness for strong almost-stationary solutions. The weak uniqueness of energy solutions follows from the weak uniqueness of .

4.2. Convergence of the remainder terms

We now proceed to prove Lemmas 4.14.3 on the convergence of , and , respectively.

4.2.1. Proof of Lemmas 4.1 and 4.2

To treat we introduce the auxiliary process

for

We will show in Lemma 4.5 that converges to for all and is uniformly bounded in and . Since by assumption and is a nice test function, we get from Corollary 3.17 that converges to in .

We also define

so that . Using Corollary 3.5, we can estimate for

where we recall that

where are the cylinder functions and in terms of the Malliavin derivative associated to the measure . We prove below that we can choose so that for all . This is necessary for to be finite for all . At this point everything boils down to controlling and to showing that it goes to zero as first and then .

Observe that the random variable is an element of the second homogeneous chaos of . Let us compute its kernel. From Equation 31 we get

and furthermore

Therefore, let

so that

We also let . Using the partial integration by parts derived in Lemma 3.6, we are able to bound by a constant:

Lemma 4.4.

Setting we have for all

and in particular , where

and

Here the notation means that the norm is taken in the -variable and

is a partial Wick contraction in the sense that .

Proof.

Consider

Partially integrating by parts , we have

The second term on the right-hand side can be integrated by parts again to obtain

while the first term can be written as

In conclusion, we have the decomposition

with

and

So it suffices to bound the three terms , , independently. In order to proceed, observe that

so that by definition of

and

Using the duality of and and the Cauchy–Schwarz inequality, we bound further

where is the constant defined in Equation 46. Similarly, we obtain

where is the constant in Equation 47. This proves Equation 45.

So to control it remains to show that the two constants and vanish in the limit . Before doing so, let us prove Lemma 4.2. More precisely, we show the following refined version:

Lemma 4.5.

We have and as well as for all .

Proof.

Recall that , so

By Equation 33 we know that for the first term on the right-hand side converges to

where in the last step we used that . Moreover,

and by similar arguments also the second term on the right-hand side of Equation 50 stays bounded in . Recalling that , we get by symmetry of

as well as in for any . In particular,

For the last term on the right-hand side we further get

and the three-dimensional integral takes the value

by symmetry of the variables . To compute the two-dimensional integral, observe first that

and integrating this against and using the symmetry in , we get

Plugging all this back into Equation 51, we have

which concludes the proof.

The following computation will be useful for controlling both and , which is why we outsource it in a separate lemma.

Lemma 4.6.

Define the kernel

Then there exists such that for all there is with

Proof.

We argue by duality. For we have

where the random variables are independent and , , and (note that are all probability densities). The observation that we can simplify the notation in this way is taken from Reference FQ15. Note that by assumption for sufficiently large , and therefore

Similarly by the independence of and , hence we can regroup

Let us estimate for example the most complicated term

The other terms can be controlled using the same arguments, and therefore we get

which yields by the density of in .

Lemma 4.7.

We have .

Proof.

We expand the squared -norm as

Integrating by parts the terms and taking into account the cancellations due to the partial Wick contractions, we get

The second term can be written as

so letting , we have

Let us consider first , which according to Lemma 4.6 can be bounded by

for all large . We continue by estimating the term in Equation 56 which is bounded by

To treat the -norms, we argue again by duality, as in the proof of Lemma 4.6. Therefore, let and consider

where are independent random variables as above. Now observe that by our assumptions on

if is large enough, and similarly . So for large we can decompose the expectations in Equation 58 as

Bounding each term individually as in the proof of Lemma 4.6, we get

which yields

By the same computation we get a similar bound for , and plugging these back into Equation 57, we generate a number of products between different expectations. Let us treat three prototypical cases: Writing , we have

where we introduced a new independent copy of ; this is a trick that we will apply several times in the following. Another situation occurs if only one of the two expectations depends on (respectively ), for example,

Finally, we have to handle the case where none of the expectations depend on or , for example,

In conclusion also vanishes as first and then , and this concludes the proof.

Lemma 4.8.

We have .

Proof.

Recall that

So by Lemma 4.6 we get directly for all large , from where the convergence immediately follows.

Lemma 4.1 now follows by combining Lemma 4.4, Lemma 4.7, and Lemma 4.8.

4.2.2. Proof of Lemma 4.3

Recall that

By Corollary 3.17 the first term on the right-hand side converges in to whenever and . The convergence of the remaining terms is obvious, and overall we get

where the convergence takes place in .

Appendix A. The periodic case

For the periodic equation described in section 2.3 most of the analysis works in the same way. The Itô trick and the Kipnis–Varadhan inequality are shown using exactly the same arguments, and also the Gaussian analysis of section 3.2 works completely analogously. We only have to replace all function spaces over by the corresponding spaces over , say by . The construction of the Burgers nonlinearity and the proof of its time-regularity also carry over to the periodic setting, although we have to replace the integrals over in Fourier space by sums over . But since those sums can be estimated by the corresponding integrals, we get the same bounds.

The first significant difference is in the construction of the integral. As discussed in section 4.1, any integral of is determined uniquely by its derivative and the value for some with . The same is true on the circle, and here there is a canonical candidate for the function , namely the constant function . So let be a pair of controlled processes, where is a periodic space-time white noise and is a periodic space white noise. Let be an even function with and such that on a neighborhood of and define

where respectively denotes the Fourier transform (respectively inverse Fourier transform) on the torus, is the convolution on the torus, and is the periodization of . For the last identity in Equation 59 we applied Poisson summation; see for example Reference GP15, Lemma 6. We then integrate by setting

where , which corresponds to

or equivalently for . From the representation as a Fourier multiplier, it is obvious that , and since we assumed that for all we get . Writing for , we get from the fact that is a strong stationary solution of the periodic Burgers equation that

and . From the expression for we see that , where denotes the Dirac delta, and therefore for . So setting , we have

and since we get

where we expanded the -norm and defined

for the constant (which is independent of ) with defined in Equation 61, and

From here on the proof is completely analogous to the nonperiodic setting provided that we establish the following three lemmas.

Lemma A.1.

We have for all , and all

Lemma A.2.

The constant converges to as .

Lemma A.3.

For all the process converges in probability in to the zero quadratic variation process

To prove these lemmas, we follow the argumentation in section 4.2. Here the kernel takes the form

and as in section 4.2 we see that we should choose

The proof of Lemma A.2 is not a trivial modification of the one of Lemma 4.2, so we provide the required arguments.

Proof of Lemma A.2.

Sending and using that and we get

and the first term on the right-hand side is . Moreover, also

and therefore we remain with

which concludes the proof.

The rest of the proof is completely analogous to the nonperiodic case. Let us just point out that if , then

where is a random variable with density , and that

and therefore the same line of argumentation as in section 4.2 yields Lemmas A.1 and A.3. From here we follow the same steps as in the proof of Theorem 2.4 to establish Theorem 2.13.

Appendix B. Exponential integrability

Lemma B.1.

Let for and , and

Then there exist such that

Moreover, if and there exists with , then for some , independent of ,

Proof.

Note that

for some , so the second statement is a generalization of the first one. We have

and therefore with Hölder’s and Jensen’s inequalities

It remains to bound the expectation on the right-hand side. To simplify notation, we will write the following computations directly for , but to make them rigorous we should mollify before taking the square and derive bounds that are uniform in the mollification. Let be the functional from the proof of Proposition 3.15 with . By Reference GJ13, Lemma 3 we have

where the third step follows from Jensen’s inequality. Now and are random variables in the second inhomogeneous Gaussian chaos generated by , and therefore they have small exponential moments. Indeed, if is a random variable living in a second-order homogeneous Gaussian chaos, then for all (see Reference Nua06, p. 62), and therefore by Stirling’s formula

which is finite as long as . Since variables in the first- and zero- order Gaussian chaos have all exponential moments and we can make small by choosing sufficiently small, the proof is complete once we know that . But this we already showed in the proof of Proposition 3.15.

Mathematical Fragments

Equation (1)
Equation (2)
Equation (4)
Equation (5)
Definition 2.1 (Controlled process).

Let , let be a space-time white noise on the filtered probability space , and let be an -measurable space white noise. Denote with the space of pairs of adapted stochastic processes with trajectories in such that

i)

and the law of is that of a white noise with variance for all ;

ii)

for any test function , the process is almost surely of zero quadratic variation, satisfies , and the pair solves the equation

iii)

for any the time-reversed processes , satisfy

where is a space-time white noise in the filtration generated by .

If there exist a space-time white noise and a white noise such that , then we simply write . For and we omit the parameters in the notation and write respectively .

Proposition 2.2.

Let , let with , and write for . Then for all the process

converges uniformly on compacts in probability to a limiting process that we denote with

As the notation suggests, this limit does not depend on the function .

Definition 2.3.

Let be a space-time white noise on , let be an -measurable space white noise, and let . Then is called a strong stationary solution to the SBE

if for all . If and for all , then is called an energy solution to 7.

Equation (8)
Theorem 2.4.

Let be a space-time white noise on the filtered probability space and let be an -measurable space white noise. Then the strong stationary solution to

is unique up to indistinguishability. Moreover, for we have , where the derivative is taken in the distributional sense, and is the unique solution to the linear multiplicative SHE Equation 8 for an arbitrary with and . Consequently, any two energy solutions of 9 have the same law.

Remark 2.7 (Martingale problem).

Energy solutions can be understood as solutions to a martingale problem. Indeed, given a pair of stochastic processes with trajectories in , we need to check the following criteria to verify that is the unique-in-law energy solution to Equation 9:

i)

The law of is that of the white noise with variance for all .

ii)

For any test function the process is almost surely of zero quadratic variation, satisfies , and the pair solves the equation

where is a continuous martingale in the filtration generated by , such that and has quadratic variation .

iii)

For any the time-reversed processes , satisfy

where is a continuous martingale in the filtration generated by , such that and has quadratic variation .

iv)

There exists such that and with we have

for all and .

Equation (11)
Theorem 2.10.

Let be a space-time white noise on and let be an -measurable random variable with values in such that is a space white noise. Then the strong almost-stationary solution to

is unique up to indistinguishability. Moreover, for we have

where is the unique solution to the linear multiplicative SHE

Consequently, any two energy solutions of 12 have the same law.

Equation (15)
Theorem 2.13.

Let be a periodic space-time white noise on and let be an -measurable periodic space white noise. Then the strong stationary solution to

is unique up to indistinguishability. Moreover, for we have , where the derivative is taken in the distributional sense and is the unique solution to the linear multiplicative SHE

for , , , and where denotes the Fourier transform on . Consequently, any two energy solutions of 16 have the same law.

Proposition 3.2 (Itô trick).

Let , and . Then we have for all

For we get in particular

Remark 3.4.

If in the setting of Proposition 3.2 has a finite chaos expansion of length for all (see section 3.2 for the definition), then also has a chaos expansion of length and therefore Gaussian hypercontractivity shows that for all

Equation (19)
Corollary 3.5 (Kipnis–Varadhan inequality).

Let and , and let be a controlled process. Then for all

Equation (20)
Lemma 3.6.

Let and let be Malliavin differentiable in . Then

Equation (21)
Corollary 3.8.

The operator is closable in and the domain of its closure, still denoted with , is

For we have

Lemma 3.9.

For we have

Lemma 3.13.

For and we have

Proposition 3.15.

Let , , let , and let . Then

Therefore, the integral is well-defined also for . Moreover, there exists a unique process such that for every approximate identity (i.e., , for all , for all we have

where the convergence is in , uniformly on compact subsets of . If is such that , and , then

Equation (25)
Corollary 3.17.

For all and all , the process is almost surely in and we have for all with and for all

where the convergence is in for all and we write again .

Equation (26)
Equation (27)
Equation (28)
Equation (29)
Equation (31)
Equation (32)
Equation (33)
Equation (34)
Equation (38)
Lemma 4.1.

We have for all , and all

Lemma 4.2.

The deterministic function converges to as and is uniformly bounded in the sense that .

Lemma 4.3.

For all the process converges in probability in to the zero quadratic variation process

Equation (39)
Equation (40)
Equation (41)
Lemma 4.4.

Setting we have for all

and in particular , where

and

Here the notation means that the norm is taken in the -variable and

is a partial Wick contraction in the sense that .

Lemma 4.5.

We have and as well as for all .

Equation (50)
Equation (51)
Lemma 4.6.

Define the kernel

Then there exists such that for all there is with

Lemma 4.7.

We have .

Equation (56)
Equation (57)
Equation (58)
Lemma 4.8.

We have .

Equation (59)
Lemma A.1.

We have for all , and all

Lemma A.2.

The constant converges to as .

Lemma A.3.

For all the process converges in probability in to the zero quadratic variation process

Equation (61)
Lemma B.1.

Let for and , and

Then there exist such that

Moreover, if and there exists with , then for some , independent of ,

References

Reference [ACQ11]
G. Amir, I. Corwin, and J. Quastel, Probability distribution of the free energy of the continuum directed random polymer in dimensions, Comm. Pure Appl. Math. 64 (2011), no. 4, 466–537, DOI 10.1002/cpa.20347. MR2796514,
Show rawAMSref \bib{bib:amirCorwinQuastel}{article}{ label={ACQ11}, author={Amir, Gideon}, author={Corwin, Ivan}, author={Quastel, Jeremy}, title={Probability distribution of the free energy of the continuum directed random polymer in $1+1$ dimensions}, journal={Comm. Pure Appl. Math.}, volume={64}, date={2011}, number={4}, pages={466--537}, issn={0010-3640}, review={\MR {2796514}}, doi={10.1002/cpa.20347}, }
Reference [Ass02]
S. Assing, A pregenerator for Burgers equation forced by conservative noise, Comm. Math. Phys. 225 (2002), no. 3, 611–632, DOI 10.1007/s002200100606. MR1888875,
Show rawAMSref \bib{bib:assing}{article}{ label={Ass02}, author={Assing, Sigurd}, title={A pregenerator for Burgers equation forced by conservative noise}, journal={Comm. Math. Phys.}, volume={225}, date={2002}, number={3}, pages={611--632}, issn={0010-3616}, review={\MR {1888875}}, doi={10.1007/s002200100606}, }
Reference [BCD11]
H. Bahouri, J.-Y. Chemin, and R. Danchin, Fourier analysis and nonlinear partial differential equations, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 343, Springer, Heidelberg, 2011, DOI 10.1007/978-3-642-16830-7. MR2768550,
Show rawAMSref \bib{bib:bahouri}{book}{ label={BCD11}, author={Bahouri, Hajer}, author={Chemin, Jean-Yves}, author={Danchin, Rapha\"el}, title={Fourier analysis and nonlinear partial differential equations}, series={Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={343}, publisher={Springer, Heidelberg}, date={2011}, pages={xvi+523}, isbn={978-3-642-16829-1}, review={\MR {2768550}}, doi={10.1007/978-3-642-16830-7}, }
Reference [BG97]
L. Bertini and G. Giacomin, Stochastic Burgers and KPZ equations from particle systems, Comm. Math. Phys. 183 (1997), no. 3, 571–607, DOI 10.1007/s002200050044. MR1462228,
Show rawAMSref \bib{bib:bertiniGiacomin}{article}{ label={BG97}, author={Bertini, Lorenzo}, author={Giacomin, Giambattista}, title={Stochastic Burgers and KPZ equations from particle systems}, journal={Comm. Math. Phys.}, volume={183}, date={1997}, number={3}, pages={571--607}, issn={0010-3616}, review={\MR {1462228}}, doi={10.1007/s002200050044}, }
Reference [BGS16]
O. Blondel, P. Gonçalves, and M. Simon, Convergence to the stochastic Burgers equation from a degenerate microscopic dynamics, Electron. J. Probab. 21 (2016), Paper No. 69, 25, DOI 10.1214/16-EJP15. MR3580035,
Show rawAMSref \bib{bib:blondelGoncalvesSimon}{article}{ label={BGS16}, author={Blondel, Oriane}, author={Gon\c {c}alves, Patr\'\i cia}, author={Simon, Marielle}, title={Convergence to the stochastic Burgers equation from a degenerate microscopic dynamics}, journal={Electron. J. Probab.}, volume={21}, date={2016}, pages={Paper No. 69, 25}, issn={1083-6489}, review={\MR {3580035}}, doi={10.1214/16-EJP15}, }
Reference [CGF17]
G. Cannizzaro, P. K. Friz, and P. Gassiat, Malliavin calculus for regularity structures: the case of gPAM, J. Funct. Anal. 272 (2017), no. 1, 363–419, DOI 10.1016/j.jfa.2016.09.024. MR3567508,
Show rawAMSref \bib{bib:cannizzaroGassiatFriz}{article}{ label={CGF17}, author={Cannizzaro, G.}, author={Friz, P. K.}, author={Gassiat, P.}, title={Malliavin calculus for regularity structures: the case of gPAM}, journal={J. Funct. Anal.}, volume={272}, date={2017}, number={1}, pages={363--419}, issn={0022-1236}, review={\MR {3567508}}, doi={10.1016/j.jfa.2016.09.024}, }
Reference [Cha00]
T. Chan, Scaling limits of Wick ordered KPZ equation, Comm. Math. Phys. 209 (2000), no. 3, 671–690, DOI 10.1007/PL00020963. MR1743612,
Show rawAMSref \bib{bib:chan}{article}{ label={Cha00}, author={Chan, Terence}, title={Scaling limits of Wick ordered KPZ equation}, journal={Comm. Math. Phys.}, volume={209}, date={2000}, number={3}, pages={671--690}, issn={0010-3616}, review={\MR {1743612}}, doi={10.1007/PL00020963}, }
Reference [Cor12]
I. Corwin, The Kardar-Parisi-Zhang equation and universality class, Random Matrices Theory Appl. 1 (2012), no. 1, 1130001, 76, DOI 10.1142/S2010326311300014. MR2930377,
Show rawAMSref \bib{bib:corwin}{article}{ label={Cor12}, author={Corwin, Ivan}, title={The Kardar-Parisi-Zhang equation and universality class}, journal={Random Matrices Theory Appl.}, volume={1}, date={2012}, number={1}, pages={1130001, 76}, issn={2010-3263}, review={\MR {2930377}}, doi={10.1142/S2010326311300014}, }
Reference [CST16]
I. Corwin, H. Shen, and L.-C. Tsai, ASEP converges to the KPZ equation. arXiv preprint arXiv:1602.01908 (2016).
Reference [CT17]
I. Corwin and L.-C. Tsai, KPZ equation limit of higher-spin exclusion processes, Ann. Probab. 45 (2017), no. 3, 1771–1798, DOI 10.1214/16-AOP1101. MR3650415,
Show rawAMSref \bib{bib:corwinTsai}{article}{ label={CT17}, author={Corwin, Ivan}, author={Tsai, Li-Cheng}, title={KPZ equation limit of higher-spin exclusion processes}, journal={Ann. Probab.}, volume={45}, date={2017}, number={3}, pages={1771--1798}, issn={0091-1798}, review={\MR {3650415}}, doi={10.1214/16-AOP1101}, }
Reference [DGP17]
J. Diehl, M. Gubinelli, and N. Perkowski, The Kardar-Parisi-Zhang equation as scaling limit of weakly asymmetric interacting Brownian motions, Comm. Math. Phys. 354 (2017), no. 2, 549–589, DOI 10.1007/s00220-017-2918-6. MR3663617,
Show rawAMSref \bib{bib:diehlGubinelliPerkowski}{article}{ label={DGP17}, author={Diehl, Joscha}, author={Gubinelli, Massimiliano}, author={Perkowski, Nicolas}, title={The Kardar-Parisi-Zhang equation as scaling limit of weakly asymmetric interacting Brownian motions}, journal={Comm. Math. Phys.}, volume={354}, date={2017}, number={2}, pages={549--589}, issn={0010-3616}, review={\MR {3663617}}, doi={10.1007/s00220-017-2918-6}, }
Reference [DT16]
A. Dembo and L.-C. Tsai, Weakly asymmetric non-simple exclusion process and the Kardar-Parisi-Zhang equation, Comm. Math. Phys. 341 (2016), no. 1, 219–261, DOI 10.1007/s00220-015-2527-1. MR3439226,
Show rawAMSref \bib{bib:demboTsai}{article}{ label={DT16}, author={Dembo, Amir}, author={Tsai, Li-Cheng}, title={Weakly asymmetric non-simple exclusion process and the Kardar-Parisi-Zhang equation}, journal={Comm. Math. Phys.}, volume={341}, date={2016}, number={1}, pages={219--261}, issn={0010-3616}, review={\MR {3439226}}, doi={10.1007/s00220-015-2527-1}, }
Reference [FGS16]
T. Franco, P. Gonçalves, and M. Simon, Crossover to the stochastic Burgers equation for the WASEP with a slow bond, Comm. Math. Phys. 346 (2016), no. 3, 801–838, DOI 10.1007/s00220-016-2607-x. MR3537337,
Show rawAMSref \bib{bib:francoGoncalvesSimon}{article}{ label={FGS16}, author={Franco, Tertuliano}, author={Gon\c {c}alves, Patr\'\i cia}, author={Simon, Marielle}, title={Crossover to the stochastic Burgers equation for the WASEP with a slow bond}, journal={Comm. Math. Phys.}, volume={346}, date={2016}, number={3}, pages={801--838}, issn={0010-3616}, review={\MR {3537337}}, doi={10.1007/s00220-016-2607-x}, }
Reference [FQ15]
T. Funaki and J. Quastel, KPZ equation, its renormalization and invariant measures, Stoch. Partial Differ. Equ. Anal. Comput. 3 (2015), no. 2, 159–220, DOI 10.1007/s40072-015-0046-x. MR3350451,
Show rawAMSref \bib{bib:funakiQuastel}{article}{ label={FQ15}, author={Funaki, Tadahisa}, author={Quastel, Jeremy}, title={KPZ equation, its renormalization and invariant measures}, journal={Stoch. Partial Differ. Equ. Anal. Comput.}, volume={3}, date={2015}, number={2}, pages={159--220}, issn={2194-0401}, review={\MR {3350451}}, doi={10.1007/s40072-015-0046-x}, }
Reference [GJ10]
P. Gonçalves and M. Jara, Universality of KPZ equation. arXiv preprint arXiv:1003.4478 (2010).
Reference [GJ14]
P. Gonçalves and M. Jara, Nonlinear fluctuations of weakly asymmetric interacting particle systems, Arch. Ration. Mech. Anal. 212 (2014), no. 2, 597–644, DOI 10.1007/s00205-013-0693-x. MR3176353,
Show rawAMSref \bib{bib:goncalvesJara}{article}{ label={GJ14}, author={Gon\c {c}alves, Patr\'\i cia}, author={Jara, Milton}, title={Nonlinear fluctuations of weakly asymmetric interacting particle systems}, journal={Arch. Ration. Mech. Anal.}, volume={212}, date={2014}, number={2}, pages={597--644}, issn={0003-9527}, review={\MR {3176353}}, doi={10.1007/s00205-013-0693-x}, }
Reference [GJ16]
P. Gonçalves and M. Jara, Stochastic Burgers equation from long range exclusion interactions. arXiv preprint arXiv:1606.06655 (2016).
Reference [GJS15]
P. Gonçalves, M. Jara, and S. Sethuraman, A stochastic Burgers equation from a class of microscopic interactions, Ann. Probab. 43 (2015), no. 1, 286–338, DOI 10.1214/13-AOP878. MR3298474,
Show rawAMSref \bib{bib:goncalvesJaraSethuraman}{article}{ label={GJS15}, author={Gon\c {c}alves, Patr\'\i cia}, author={Jara, Milton}, author={Sethuraman, Sunder}, title={A stochastic Burgers equation from a class of microscopic interactions}, journal={Ann. Probab.}, volume={43}, date={2015}, number={1}, pages={286--338}, issn={0091-1798}, review={\MR {3298474}}, doi={10.1214/13-AOP878}, }
Reference [GJS17]
P. Gonçalves, M. Jara, and M. Simon, Second order Boltzmann-Gibbs principle for polynomial functions and applications, J. Stat. Phys. 166 (2017), no. 1, 90–113, DOI 10.1007/s10955-016-1686-6. MR3592852,
Show rawAMSref \bib{bib:goncalvesSimon}{article}{ label={GJS17}, author={Gon\c {c}alves, Patr\'\i cia}, author={Jara, Milton}, author={Simon, Marielle}, title={Second order Boltzmann-Gibbs principle for polynomial functions and applications}, journal={J. Stat. Phys.}, volume={166}, date={2017}, number={1}, pages={90--113}, issn={0022-4715}, review={\MR {3592852}}, doi={10.1007/s10955-016-1686-6}, }
Reference [GIP15]
M. Gubinelli, P. Imkeller, and N. Perkowski, Paracontrolled distributions and singular PDEs, Forum Math. Pi 3 (2015), e6, 75, DOI 10.1017/fmp.2015.2. MR3406823,
Show rawAMSref \bib{bib:paracontrolled}{article}{ label={GIP15}, author={Gubinelli, Massimiliano}, author={Imkeller, Peter}, author={Perkowski, Nicolas}, title={Paracontrolled distributions and singular PDEs}, journal={Forum Math. Pi}, volume={3}, date={2015}, pages={e6, 75}, issn={2050-5086}, review={\MR {3406823}}, doi={10.1017/fmp.2015.2}, }
Reference [GJ13]
M. Gubinelli and M. Jara, Regularization by noise and stochastic Burgers equations, Stoch. Partial Differ. Equ. Anal. Comput. 1 (2013), no. 2, 325–350, DOI 10.1007/s40072-013-0011-5. MR3327509,
Show rawAMSref \bib{bib:gubinelliJara}{article}{ label={GJ13}, author={Gubinelli, M.}, author={Jara, M.}, title={Regularization by noise and stochastic Burgers equations}, journal={Stoch. Partial Differ. Equ. Anal. Comput.}, volume={1}, date={2013}, number={2}, pages={325--350}, issn={2194-0401}, review={\MR {3327509}}, doi={10.1007/s40072-013-0011-5}, }
Reference [GP17]
M. Gubinelli and N. Perkowski, KPZ reloaded, Comm. Math. Phys. 349 (2017), no. 1, 165–269, DOI 10.1007/s00220-016-2788-3. MR3592748,
Show rawAMSref \bib{bib:kpzReloaded}{article}{ label={GP17}, author={Gubinelli, Massimiliano}, author={Perkowski, Nicolas}, title={KPZ reloaded}, journal={Comm. Math. Phys.}, volume={349}, date={2017}, number={1}, pages={165--269}, issn={0010-3616}, review={\MR {3592748}}, doi={10.1007/s00220-016-2788-3}, }
Reference [GP15]
M. Gubinelli and N. Perkowski, Lectures on singular stochastic PDEs, Ensaios Matemáticos [Mathematical Surveys], vol. 29, Sociedade Brasileira de Matemática, Rio de Janeiro, 2015. MR3445609,
Show rawAMSref \bib{bib:ebp}{book}{ label={GP15}, author={Gubinelli, Massimiliano}, author={Perkowski, Nicolas}, title={Lectures on singular stochastic PDEs}, series={Ensaios Matem\'aticos [Mathematical Surveys]}, volume={29}, publisher={Sociedade Brasileira de Matem\'atica, Rio de Janeiro}, date={2015}, pages={89}, isbn={978-85-8337-079-6}, review={\MR {3445609}}, }
Reference [GP16]
M. Gubinelli and N. Perkowski, The Hairer–Quastel universality result at stationarity, Stochastic analysis on large scale interacting systems, RIMS Kôkyûroku Bessatsu, B59, Res. Inst. Math. Sci. (RIMS), Kyoto, 2016, pp. 101–115. MR3675927,
Show rawAMSref \bib{bib:gubinelliHQ}{article}{ label={GP16}, author={Gubinelli, Massimiliano}, author={Perkowski, Nicolas}, title={The Hairer--Quastel universality result at stationarity}, conference={ title={Stochastic analysis on large scale interacting systems}, }, book={ series={RIMS K\^oky\^uroku Bessatsu, B59}, publisher={Res. Inst. Math. Sci. (RIMS), Kyoto}, }, date={2016}, pages={101--115}, review={\MR {3675927}}, }
Reference [Hai13]
M. Hairer, Solving the KPZ equation, Ann. of Math. (2) 178 (2013), no. 2, 559–664, DOI 10.4007/annals.2013.178.2.4. MR3071506,
Show rawAMSref \bib{bib:hairerKPZ}{article}{ label={Hai13}, author={Hairer, Martin}, title={Solving the KPZ equation}, journal={Ann. of Math. (2)}, volume={178}, date={2013}, number={2}, pages={559--664}, issn={0003-486X}, review={\MR {3071506}}, doi={10.4007/annals.2013.178.2.4}, }
Reference [Hai14]
M. Hairer, A theory of regularity structures, Invent. Math. 198 (2014), no. 2, 269–504, DOI 10.1007/s00222-014-0505-4. MR3274562,
Show rawAMSref \bib{bib:hairerRegularity}{article}{ label={Hai14}, author={Hairer, M.}, title={A theory of regularity structures}, journal={Invent. Math.}, volume={198}, date={2014}, number={2}, pages={269--504}, issn={0020-9910}, review={\MR {3274562}}, doi={10.1007/s00222-014-0505-4}, }
Reference [HL15]
M. Hairer and C. Labbé. Multiplicative stochastic heat equations on the whole space. arXiv preprint arXiv:1504.07162 (2015).
Reference [Hos16]
M. Hoshino, Paracontrolled calculus and Funaki–Quastel approximation for the KPZ equation. arXiv preprint arXiv:1605.02624 (2016).
Reference [HQ15]
M. Hairer and J. Quastel, A class of growth models rescaling to KPZ. arXiv preprint arXiv:1512.07845 (2015).
Reference [HS15]
M. Hairer and H. Shen, A central limit theorem for the KPZ equation. arXiv preprint arXiv:1507.01237 (2015).
Reference [Jan97]
S. Janson, Gaussian Hilbert spaces, Cambridge Tracts in Mathematics, vol. 129, Cambridge University Press, Cambridge, 1997, DOI 10.1017/CBO9780511526169. MR1474726,
Show rawAMSref \bib{bib:janson}{book}{ label={Jan97}, author={Janson, Svante}, title={Gaussian Hilbert spaces}, series={Cambridge Tracts in Mathematics}, volume={129}, publisher={Cambridge University Press, Cambridge}, date={1997}, pages={x+340}, isbn={0-521-56128-0}, review={\MR {1474726}}, doi={10.1017/CBO9780511526169}, }
Reference [KLO12]
T. Komorowski, C. Landim, and S. Olla, Fluctuations in Markov processes: Time symmetry and martingale approximation, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 345, Springer, Heidelberg, 2012, DOI 10.1007/978-3-642-29880-6. MR2952852,
Show rawAMSref \bib{bib:komorowskiLandimOlla}{book}{ label={KLO12}, author={Komorowski, Tomasz}, author={Landim, Claudio}, author={Olla, Stefano}, title={Fluctuations in Markov processes: Time symmetry and martingale approximation}, series={Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={345}, publisher={Springer, Heidelberg}, date={2012}, pages={xviii+491}, isbn={978-3-642-29879-0}, review={\MR {2952852}}, doi={10.1007/978-3-642-29880-6}, }
Reference [KM17]
A. Kupiainen and M. Marcozzi, Renormalization of generalized KPZ equation, J. Stat. Phys. 166 (2017), no. 3-4, 876–902, DOI 10.1007/s10955-016-1636-3. MR3607594,
Show rawAMSref \bib{bib:kupiainenMarcozzi}{article}{ label={KM17}, author={Kupiainen, Antti}, author={Marcozzi, Matteo}, title={Renormalization of generalized KPZ equation}, journal={J. Stat. Phys.}, volume={166}, date={2017}, number={3-4}, pages={876--902}, issn={0022-4715}, review={\MR {3607594}}, doi={10.1007/s10955-016-1636-3}, }
Reference [KPZ86]
M. Kardar, G. Parisi, and Y.-C. Zhang, Dynamic scaling of growing interfaces. Phys. Rev. Lett. 56 (1986), no. 9, 889.
Reference [Kup16]
A. Kupiainen, Renormalization group and stochastic PDEs, Ann. Henri Poincaré 17 (2016), no. 3, 497–535, DOI 10.1007/s00023-015-0408-y. MR3459120,
Show rawAMSref \bib{bib:kupiainen}{article}{ label={Kup16}, author={Kupiainen, Antti}, title={Renormalization group and stochastic PDEs}, journal={Ann. Henri Poincar\'e}, volume={17}, date={2016}, number={3}, pages={497--535}, issn={1424-0637}, review={\MR {3459120}}, doi={10.1007/s00023-015-0408-y}, }
Reference [Lab17]
C. Labbé, Weakly asymmetric bridges and the KPZ equation, Comm. Math. Phys. 353 (2017), no. 3, 1261–1298, DOI 10.1007/s00220-017-2875-0. MR3652491,
Show rawAMSref \bib{bib:labbe}{article}{ label={Lab17}, author={Labb\'e, Cyril}, title={Weakly asymmetric bridges and the KPZ equation}, journal={Comm. Math. Phys.}, volume={353}, date={2017}, number={3}, pages={1261--1298}, issn={0010-3616}, review={\MR {3652491}}, doi={10.1007/s00220-017-2875-0}, }
Reference [Lep76]
D. Lépingle, La variation d’ordre des semi-martingales (French), Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 36 (1976), no. 4, 295–316, DOI 10.1007/BF00532696. MR0420837,
Show rawAMSref \bib{bib:lepingle}{article}{ label={Lep76}, author={L\'epingle, D.}, title={La variation d'ordre $p$ des semi-martingales}, language={French}, journal={Z. Wahrscheinlichkeitstheorie und Verw. Gebiete}, volume={36}, date={1976}, number={4}, pages={295--316}, review={\MR {0420837}}, doi={10.1007/BF00532696}, }
Reference [Nua06]
D. Nualart, The Malliavin calculus and related topics, 2nd ed., Probability and its Applications (New York), Springer-Verlag, Berlin, 2006. MR2200233,
Show rawAMSref \bib{bib:nualart}{book}{ label={Nua06}, author={Nualart, David}, title={The Malliavin calculus and related topics}, series={Probability and its Applications (New York)}, edition={2}, publisher={Springer-Verlag, Berlin}, date={2006}, pages={xiv+382}, isbn={978-3-540-28328-7}, isbn={3-540-28328-5}, review={\MR {2200233}}, }
Reference [Qua14]
J. D. Quastel, The Kardar-Parisi-Zhang equation and universality class, XVIIth International Congress on Mathematical Physics, World Sci. Publ., Hackensack, NJ, 2014, pp. 113–133. MR3204148,
Show rawAMSref \bib{bib:quastel}{article}{ label={Qua14}, author={Quastel, J. D.}, title={The Kardar-Parisi-Zhang equation and universality class}, conference={ title={XVIIth International Congress on Mathematical Physics}, }, book={ publisher={World Sci. Publ., Hackensack, NJ}, }, date={2014}, pages={113--133}, review={\MR {3204148}}, }
Reference [QS15]
J. Quastel and H. Spohn, The one-dimensional KPZ equation and its universality class, J. Stat. Phys. 160 (2015), no. 4, 965–984, DOI 10.1007/s10955-015-1250-9. MR3373647,
Show rawAMSref \bib{bib:quastelSpohn}{article}{ label={QS15}, author={Quastel, Jeremy}, author={Spohn, Herbert}, title={The one-dimensional KPZ equation and its universality class}, journal={J. Stat. Phys.}, volume={160}, date={2015}, number={4}, pages={965--984}, issn={0022-4715}, review={\MR {3373647}}, doi={10.1007/s10955-015-1250-9}, }
Reference [RV07]
F. Russo and P. Vallois, Elements of stochastic calculus via regularization, Séminaire de Probabilités XL, Lecture Notes in Math., vol. 1899, Springer, Berlin, 2007, pp. 147–185, DOI 10.1007/978-3-540-71189-6_7. MR2409004,
Show rawAMSref \bib{bib:russoVallois}{article}{ label={RV07}, author={Russo, Francesco}, author={Vallois, Pierre}, title={Elements of stochastic calculus via regularization}, conference={ title={S\'eminaire de Probabilit\'es XL}, }, book={ series={Lecture Notes in Math.}, volume={1899}, publisher={Springer, Berlin}, }, date={2007}, pages={147--185}, review={\MR {2409004}}, doi={10.1007/978-3-540-71189-6\_7}, }
Reference [RY99]
D. Revuz and M. Yor, Continuous martingales and Brownian motion, 3rd ed., Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 293, Springer-Verlag, Berlin, 1999, DOI 10.1007/978-3-662-06400-9. MR1725357,
Show rawAMSref \bib{bib:revuzYor}{book}{ label={RY99}, author={Revuz, Daniel}, author={Yor, Marc}, title={Continuous martingales and Brownian motion}, series={Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={293}, edition={3}, publisher={Springer-Verlag, Berlin}, date={1999}, pages={xiv+602}, isbn={3-540-64325-7}, review={\MR {1725357}}, doi={10.1007/978-3-662-06400-9}, }
Reference [Spo16]
H. Spohn, The Kardar–Parisi–Zhang equation—a statistical physics perspective. arXiv preprint arXiv:1601.00499 (2016).
Reference [SS10]
T. Sasamoto and H. Spohn, Exact height distributions for the KPZ equation with narrow wedge initial condition, Nuclear Phys. B 834 (2010), no. 3, 523–542, DOI 10.1016/j.nuclphysb.2010.03.026. MR2628936,
Show rawAMSref \bib{bib:sasamotoSpohnExact}{article}{ label={SS10}, author={Sasamoto, Tomohiro}, author={Spohn, Herbert}, title={Exact height distributions for the KPZ equation with narrow wedge initial condition}, journal={Nuclear Phys. B}, volume={834}, date={2010}, number={3}, pages={523--542}, issn={0550-3213}, review={\MR {2628936}}, doi={10.1016/j.nuclphysb.2010.03.026}, }
Reference [Wal86]
J. B. Walsh, An introduction to stochastic partial differential equations, École d’été de probabilités de Saint-Flour, XIV—1984, Lecture Notes in Math., vol. 1180, Springer, Berlin, 1986, pp. 265–439, DOI 10.1007/BFb0074920. MR876085,
Show rawAMSref \bib{bib:walsh}{article}{ label={Wal86}, author={Walsh, John B.}, title={An introduction to stochastic partial differential equations}, conference={ title={\'Ecole d'\'et\'e de probabilit\'es de Saint-Flour, XIV---1984}, }, book={ series={Lecture Notes in Math.}, volume={1180}, publisher={Springer, Berlin}, }, date={1986}, pages={265--439}, review={\MR {876085}}, doi={10.1007/BFb0074920}, }
Reference [Wu99]
L. Wu, Forward-backward martingale decomposition and compactness results for additive functionals of stationary ergodic Markov processes (English, with English and French summaries), Ann. Inst. H. Poincaré Probab. Statist. 35 (1999), no. 2, 121–141, DOI 10.1016/S0246-0203(99)80008-9. MR1678517,
Show rawAMSref \bib{bib:wu}{article}{ label={Wu99}, author={Wu, Liming}, title={Forward-backward martingale decomposition and compactness results for additive functionals of stationary ergodic Markov processes}, language={English, with English and French summaries}, journal={Ann. Inst. H. Poincar\'e Probab. Statist.}, volume={35}, date={1999}, number={2}, pages={121--141}, issn={0246-0203}, review={\MR {1678517}}, doi={10.1016/S0246-0203(99)80008-9}, }

Article Information

MSC 2010
Primary: 60H15 (Stochastic partial differential equations)
Author Information
Massimiliano Gubinelli
Hausdorff Center for Mathematics & Institute for Applied Mathematics, Universität Bonn, Bonn, Germany
gubinelli@iam.uni-bonn.de
MathSciNet
Nicolas Perkowski
Institut für Mathematik, Humboldt–Universität zu Berlin, Berlin, Germany
perkowsk@math.hu-berlin.de
MathSciNet
Additional Notes

The first author gratefully acknowledges financial support by the DFG via CRC 1060.

The second author gratefully acknowledges financial support by the DFG via Research Unit FOR 2402.

Journal Information
Journal of the American Mathematical Society, Volume 31, Issue 2, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2017 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/jams/889
  • MathSciNet Review: 3758149
  • Show rawAMSref \bib{3758149}{article}{ author={Gubinelli, Massimiliano}, author={Perkowski, Nicolas}, title={Energy solutions of KPZ are unique}, journal={J. Amer. Math. Soc.}, volume={31}, number={2}, date={2018-04}, pages={427-471}, issn={0894-0347}, review={3758149}, doi={10.1090/jams/889}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.