Superlinear stochastic heat equation on
Abstract
In this paper, we study the stochastic heat equation (SHE) on subject to a centered Gaussian noise that is white in time and colored in space. We establish the existence and uniqueness of the random field solution in the presence of locally Lipschitz drift and diffusion coefficients, which can have certain superlinear growth. This is a nontrivial extension of the recent work by Dalang, Khoshnevisan and Zhang [DKZ19], where the one-dimensional SHE on subject to space-time white noise has been studied.
Keywords. Global solution; Stochastic heat equation; Reaction-diffusion; Dalang’s
condition; superlinear growth.
1 Introduction
In this paper, we study the following stochastic heat equation (SHE) on with a drift term:
(1.1) |
with both and being locally Lipschitz continuous. The noise is a centered Gaussian noise which is white in time and colored in space with the following covariance structure
(1.2) |
We assume that the correlation function in (1.2) satisfies the improved Dalang’s condition:
(1.3) |
where is the Fourier transform of , namely, . Recall that condition (1.3) with refers to Dalang’s condition [Dal99]:
(1.4) |
The case when refers to the space-time white noise. The solution to (1.1) is understood in the mild formulation:
(1.5) | ||||
where the stochastic integral is the Walsh integral [Wal86, Dal+09], is the heat kernel, and “” denotes the convolution in the spatial variable.
Motivated by the work of Fernandez Bonder and Groisman [FG09], Dalang, Khoshnevisan and Zhang [DKZ19] established the global solution with superlinear and locally Lipschitz coefficients for the one-dimensional SHE on subject to the space-time white noise. In particular, they assumed that
(1.6) |
Foondun and Nualart [FN21] studied SHE with an additive noise, i.e., , and showed that the solution to (1.1) blows up in finite time if and only if satisfies the Osgood condition:
(1.7) |
Salins [Sal21] studied this problem for SHE on a compact domain in under the following Osgood-type conditions, which are weaker than (1.6): There exists a positive and increasing function that satisfies
such that for some (which depends on the noise),
(1.8) |
Extending the above results to the SHE on the whole space with both superlinear drift and diffusion coefficients is a challenging problem due to the non-compactness of the spatial domain. Indeed, for the wave equation on (), the compact support of the corresponding fundamental solution can help circumvent this difficulty; see Millet and Sanz-Solé [MS21]. The aim of this present paper is to carry out such extension by proving the following theorem:
Theorem 1.1.
Assume the improved Dalang’s condition (1.3) is satisfied for some . Let be the solution to (1.1) starting from for some . Suppose that and are locally Lipschitz functions such that .
- (a)
-
(b)
(Local solution) If
(1.10) then for some deterministic time , there exists a unique solution solution to (1.1) for all .
-
(c)
In either case (a) or (b), the solution is Hölder continuous: a.s., where denotes the Hölder continuous function on the space-time domain with exponents and in time and space, respectively, for any small .
Remark 1.2 (Critical vs sub-critical cases).
We call the case under conditions in (1.10) the critical case and the one in (1.9) the sub-critical case. Dalang et al [DKZ19] established the global solution for the critical case using the semigroup property of the heat equation. In this paper, we cannot restart our SHE (1.1) to pass the local solution to global solution due to the fact that it is not clear whether at time , as the initial condition for the next step is again an element in a.s. This issue does not present for a continuous random field on a compact spatial domain, which is the case in [DKZ19] and [Sal21].
Remark 1.3 (Regularity of the initial conditions).
In [DKZ19], the initial condition is assumed to be a Hölder continuous function on . In contrast, in this paper, we only assume that for some large . This is one example of the smoothing effect of the heat kernel in the stochastic partial differential equation context. This improvement from a Hölder continuous function to a measurable function is due to the factorization representation of the solution (see (2.10) below). Similar arguments using this factorization have also been carried out by Salins [Sal21].
Example 1.4 (Examples of and in Theorem 1.1).
(1) The function for is locally Lipschitz, but not globally Lipschitz, continuous with linear growth and . Hence Theorem 1.1 holds when either or takes the form of . (2) For the function for , it is easy to see that the conditions and imply that and is locally Lipschitz continuous, respectively. The growth condition of either (1.9) or (1.10) makes the further restriction on the suitable choices of .
The extension given in Theorem 1.1 from a compact spatial domain to the entire space critically relies on the sharp moment formulas obtained in Theorem 1.5 below. These moment formulas, as extensions of those in [CD15, CK19, CH19] to allow a Lipschitz drift term, constitute the second and independent contribution of the paper. Indeed, when the drift term is linear, i.e., , then one can work with the following heat kernel . However, when is a Lipschitz nonlinear function, the situation is much more trickier, especially if one wants to allow rough initial conditions [CD15, CK19, CH19], namely, being a signed Borel measure such that
(1.11) |
where and is the Jordan decomposition of the signed measure . The existence and uniqueness of the solution is proved in [Hua17] (the proof still works for signed Borel measure). We will prove the following theorem:
Theorem 1.5 (Moment formulas with a Lipschitz drift term).
Let be the solution to (1.1) and suppose that and are globally Lipschitz continuous functions and the correlation function satisfies the improved Dalang’s condition (1.3) for some . Then we have the following:
-
(a)
If , then for all , and , it holds that
(1.12) where and denote the -norm and -norm, respectively,
(1.13) , and
(1.14) -
(b)
If is a rough initial condition (see (1.11)), then for all , and ,
(1.15) where and the constant does not depend on .
-
(c)
If for some and if , then for all ,
(1.16) where denotes the -norm and the constant does not depend on .
Remark 1.6.
Part (a) of Theorem 1.5 can be derived from part (b) by noticing that . However, we still keep part (a) due to the simplicity of its proof.
Using the moment bounds in (1.15), one can extend the Hölder regularity from the SHE without drift (see [SS02] for the bounded initial condition case and [CH19] for the rough initial condition case) to the one with a Lipschitz drift.
Corollary 1.7 (Hölder regularity).
Parts (a), (b), and (c) of Theorem 1.5 are proved in Sections 2.1, 2.3, and 2.4, respectively. Corollary 1.7 is proved in Section 2.5.
Finally, we list a few open questions for future exploration: (1) Theorem 1.1 cannot handle either the constant one initial condition or the Dirac delta initial condition. It is interesting to investigate if either global or local solution exists for these two special initial conditions. (2) Can one improve Theorem 1.1 by relaxing the growth conditions in (1.9) and (1.10) to the Osgood-type conditions in (1.8) as in [Sal21]?
2 Moment bounds with a Lipschitz drift term
2.1 The bounded initial data case – Proof of part (a) of Theorem 1.5
Proof of Theorem 1.5 (a):.
By Minkowski’s inequality,
where is the constant coming from the Burkholder-Davis-Gundy inequality and as ; see [CK12, Theorem 1.4] and references therein. For , consider the following norm
Then we see that
Hence,
By the improved Dalang’s condition (1.3) and by assuming that , we see that
Therefore,
Now by choosing large enough, namely,
we form a contraction map, which can be easily solved:
Notice that , we have that
Therefore, for all , when , we can take
to have that
This proves part (a) of Theorem 1.5. ∎
2.2 A Gronwall-type lemma
Let us introduce some functions. For , denote
(2.1) |
By the Fourier transform, this function can be written in the following form
(2.2) |
Define and for ,
(2.3) |
Let
(2.4) |
When we have and , we will use , and to denote , and , respectively. Note that this convention makes our notation in case of and consistent with those in [CH19], [CK19] or [BC18]. The following lemma generalizes Lemma 2.5 in [CK19] or Lemma 3.8 in [BC18] from the case and to the case with general parameters and .
Lemma 2.1.
Proof.
Corollary 2.2.
Suppose that the correlation function satisfies the improved Dalang’s condition (1.3) for some . Then for all and , when is large enough, it holds that
(2.6) |
where the constant can be chosen to be
(2.7) |
2.3 Moment bounds for rough initial data – Proof of part (b) of Theorem 1.5
In this part, we extend the moment bounds obtained in [CH19] to allow a Lipschitz drift term.
Proof of Theorem 1.5 (b).
Taking the -th norm on both sides of the mild form (1.5) with and applying the Minkowski inequality, we see that
(2.8) |
By the Burkholder-Davis-Gundy inequality (see also a similar argument in the step 1 of the proof of Theorem 1.7 of [CH19] on p. 1000), we see that
Then by the sub-additivity of square root,
(2.9) |
By the Cauchy-Schwartz inequality applied to the integral, the square of second term on the right-hand side of (2.8) is bounded by
Hence, by raising both sides of (2.8) by a power two and recalling that the constant is defined in (1.13), we obtain that
2.4 Uniform moment bounds – Proof of part (c) of Theorem 1.5
Proof of Theorem 1.5 (c).
Fix arbitrary and recall that as in (1.3). The proof relies on the factorization lemma (see, e.g., Section 5.3.1 of [DZ14]), which says that
(2.10) |
where
and
It is clear that
Step 1. In this step, we will show that
(2.11) |
Let and be a conjugate pair on positive numbers, i.e., , whose values will be determined below. By Hölder’s inequality, we see that
where we have used the fact that in the second inequality. Hence, since
we have
Notice that
Since , by (1.15),
Combining the above three bounds shows that
with
By the same arguments as the proof of Theorem 1.8 of [CH19] (see, in particular, the bound for on p. 1006 ibid.), we see that
By Hölder’s inequality, we see that
Therefore,
Combining the last two inequalities proves (2.11).
2.5 Hölder regularity – Proof of Corollary 1.7
Proof of Corollary 1.7.
Denote the last two parts of right-hand side of (1.5) by and . One can use the same arguments as those in the proof of Theorem 1.8 of [CH19], but with the slightly different moment formula (1.15), to show that . It remains to show that . Now choose and fix arbitrary and . For any , with , an application of the Minkowski inequality shows that
By the moment formula (1.15) and by setting , we see that
It is clear that is a rough initial condition, i.e., condition (1.11) is satisfied for . Denote . It is straightforward to see that . As for and , for any , by Lemma 3.1 of [CH19], we have that
and similarly,
Combining the above bounds proves Corollary 1.7. ∎
3 Proof of Theorem 1.1
Proof of Theorem 1.1.
For , let us consider the truncated stochastic heat equation:
(3.1) | |||
where
(3.2) |
Recall that and denote the growth rate; see (1.14). According to Theorem 1.1 of [Hua17], there exists a unique solution to (3.1). In the following, we will use to denote a generic constant that may change its value at each appearance, does not depend on , but may depend on .
Step 1. In this step, we will prove (a). For any fixed, consider the following stopping time
Noticing that for all , we have that and
we can construct the solution via
(3.3) |
From the definition, it is clear that on ,
By the Chebyshev inequality and the moment formula (1.16),
(3.4) |
The sub-critical conditions in (1.9) implies that
which ensure that above probability in (3.4) goes to zero as . Therefore, by sending , we see that is well defined on . The uniqueness is inherited from the uniqueness of in (3.1).
Step 2. Now we prove part (b), the proof of which is similar to that of part (a). Fix an arbitrary . Denote
We claim that
(3.5) |
Indeed, for all , by replacing by in (3.4), we see that
(3.6) |
By the critical conditions in (1.10), for some ,
Hence, when is small enough, by plugging the above constants into (3.6), we see that the probability in (3.6) goes to zero as . Therefore, by choosing any positive constant , we prove the claim (3.5). The uniqueness is proved in the same way as the proof of part (a).
Acknowledgement
J. Huang thanks Mohammud Foondun for pointing out the reference [Sal21] when J. H. presented this paper at a conference.
References
- [BC18] Raluca M. Balan and Le Chen “Parabolic Anderson model with space-time homogeneous Gaussian noise and rough initial condition” In J. Theoret. Probab. 31.4, 2018, pp. 2216–2265 DOI: 10.1007/s10959-017-0772-2
- [CD15] Le Chen and Robert C. Dalang “Moments and growth indices for the nonlinear stochastic heat equation with rough initial conditions” In Ann. Probab. 43.6, 2015, pp. 3006–3051 DOI: 10.1214/14-AOP954
- [CH19] Le Chen and Jingyu Huang “Comparison principle for stochastic heat equation on ” In Ann. Probab. 47.2, 2019, pp. 989–1035 DOI: 10.1214/18-AOP1277
- [CK19] Le Chen and Kunwoo Kim “Nonlinear stochastic heat equation driven by spatially colored noise: moments and intermittency” In Acta Math. Sci. Ser. B (Engl. Ed.) 39.3, 2019, pp. 645–668 DOI: 10.1007/s10473-019-0303-6
- [CK12] Daniel Conus and Davar Khoshnevisan “On the existence and position of the farthest peaks of a family of stochastic heat and wave equations” In Probab. Theory Related Fields 152.3-4, 2012, pp. 681–701 DOI: 10.1007/s00440-010-0333-4
- [DZ14] Giuseppe Da Prato and Jerzy Zabczyk “Stochastic equations in infinite dimensions” 152, Encyclopedia of Mathematics and its Applications Cambridge University Press, Cambridge, 2014, pp. xviii+493 DOI: 10.1017/CBO9781107295513
- [Dal99] Robert C. Dalang “Extending the martingale measure stochastic integral with applications to spatially homogeneous s.p.d.e.’s” In Electron. J. Probab. 4, 1999, pp. no. 6\bibrangessep29 DOI: 10.1214/EJP.v4-43
- [DKZ19] Robert C. Dalang, Davar Khoshnevisan and Tusheng Zhang “Global solutions to stochastic reaction-diffusion equations with super-linear drift and multiplicative noise” In Ann. Probab. 47.1, 2019, pp. 519–559 DOI: 10.1214/18-AOP1270
- [Dal+09] Robert Dalang, Davar Khoshnevisan, Carl Mueller, David Nualart and Yimin Xiao “A minicourse on stochastic partial differential equations” Held at the University of Utah, Salt Lake City, UT, May 8–19, 2006, Edited by Khoshnevisan and Firas Rassoul-Agha 1962, Lecture Notes in Mathematics Springer-Verlag, Berlin, 2009, pp. xii+216
- [FG09] Julian Fernández Bonder and Pablo Groisman “Time-space white noise eliminates global solutions in reaction-diffusion equations” In Phys. D 238.2, 2009, pp. 209–215 DOI: 10.1016/j.physd.2008.09.005
- [FN21] Mohammud Foondun and Eulalia Nualart “The Osgood condition for stochastic partial differential equations” In Bernoulli 27.1, 2021, pp. 295–311 DOI: 10.3150/20-BEJ1240
- [Hua17] Jingyu Huang “On stochastic heat equation with measure initial data” In Electron. Commun. Probab. 22, 2017, pp. Paper No. 40\bibrangessep6 DOI: 10.1214/17-ECP71
- [MS21] Annie Millet and Marta Sanz-Solé “Global solutions to stochastic wave equations with superlinear coefficients” In Stochastic Process. Appl. 139, 2021, pp. 175–211 DOI: 10.1016/j.spa.2021.05.002
- [Sal21] Michael Salins “Global solutions to the stochastic heat equation with superlinear accretive reaction term and superlinear multiplicative noise term on a bounded spatial domain” In preprint arXiv:2110.10130, 2021
- [SS02] M. Sanz-Solé and M. Sarrà “Hölder continuity for the stochastic heat equation with spatially correlated noise” In Seminar on Stochastic Analysis, Random Fields and Applications, III (Ascona, 1999) 52, Progr. Probab. Birkhäuser, Basel, 2002, pp. 259–268
- [Wal86] John B. Walsh “An introduction to stochastic partial differential equations” In École d’été de probabilités de Saint-Flour, XIV—1984 1180, Lecture Notes in Math. Springer, Berlin, 1986, pp. 265–439 DOI: 10.1007/BFb0074920