In this paper, we considered the parabolic Anderson model with a class of time-independent generalized Gaussian fields on Rd, which included fractional white noise, Bessel field, massive free field, and other nonstationary Gaussian fields. Under the rough initial conditions, we constructed the Feynman-Kac formula as a solution in the Stratonovich integral by Brownian bridge, and then proved the Hölder continuity of the solution with respect to the time variable. As a comparison, we also studied the Hölder continuity under the regular initial conditions that u0≡C and u0∈Cκ(Rd) with κ∈(0,1].
Citation: Hui Sun, Yangyang Lyu. Temporal Hölder continuity of the parabolic Anderson model driven by a class of time-independent Gaussian fields with rough initial conditions[J]. AIMS Mathematics, 2024, 9(12): 34838-34862. doi: 10.3934/math.20241659
[1] | Carey Caginalp . A minimization approach to conservation laws with random initialconditions and non-smooth, non-strictly convex flux. AIMS Mathematics, 2018, 3(1): 148-182. doi: 10.3934/Math.2018.1.148 |
[2] | Xiaodong Zhang, Junfeng Liu . Solving a class of high-order fractional stochastic heat equations with fractional noise. AIMS Mathematics, 2022, 7(6): 10625-10650. doi: 10.3934/math.2022593 |
[3] | Ugur G. Abdulla . Generalized Newton-Leibniz formula and the embedding of the Sobolev functions with dominating mixed smoothness into Hölder spaces. AIMS Mathematics, 2023, 8(9): 20700-20717. doi: 10.3934/math.20231055 |
[4] | Jia Li, Zhipeng Tong . Local Hölder continuity of inverse variation-inequality problem constructed by non-Newtonian polytropic operators in finance. AIMS Mathematics, 2023, 8(12): 28753-28765. doi: 10.3934/math.20231472 |
[5] | Yudong Sun, Tao Wu . Hölder and Schauder estimates for weak solutions of a certain class of non-divergent variation inequality problems in finance. AIMS Mathematics, 2023, 8(8): 18995-19003. doi: 10.3934/math.2023968 |
[6] | Ruimin Xu, Kaiyue Dong, Jingyu Zhang, Ying Zhou . Linear-quadratic-Gaussian mean-field games driven by Poisson jumps with major and minor agents. AIMS Mathematics, 2025, 10(5): 11086-11110. doi: 10.3934/math.2025503 |
[7] | Quincy Stévène Nkombo, Fengquan Li, Christian Tathy . Stability properties of Radon measure-valued solutions for a class of nonlinear parabolic equations under Neumann boundary conditions. AIMS Mathematics, 2021, 6(11): 12182-12224. doi: 10.3934/math.2021707 |
[8] | Pakeeza Ashraf, Ghulam Mustafa, Husna A. Khan, Dumitru Baleanu, Abdul Ghaffar, Kottakkaran Sooppy Nisar . A shape-preserving variant of Lane-Riesenfeld algorithm. AIMS Mathematics, 2021, 6(3): 2152-2170. doi: 10.3934/math.2021131 |
[9] | Kelei Zhang . Orlicz estimates for parabolic Schrödinger operators with non-negative potentials on nilpotent Lie groups. AIMS Mathematics, 2023, 8(8): 18631-18648. doi: 10.3934/math.2023949 |
[10] | Chao Wei . Parameter estimation for partially observed stochastic differential equations driven by fractional Brownian motion. AIMS Mathematics, 2022, 7(7): 12952-12961. doi: 10.3934/math.2022717 |
In this paper, we considered the parabolic Anderson model with a class of time-independent generalized Gaussian fields on Rd, which included fractional white noise, Bessel field, massive free field, and other nonstationary Gaussian fields. Under the rough initial conditions, we constructed the Feynman-Kac formula as a solution in the Stratonovich integral by Brownian bridge, and then proved the Hölder continuity of the solution with respect to the time variable. As a comparison, we also studied the Hölder continuity under the regular initial conditions that u0≡C and u0∈Cκ(Rd) with κ∈(0,1].
In this paper, we study the following stochastic heat equation
∂∂tu(t,x)=12△u(t,x)+θV(x)u(t,x),(t,x)∈R+×Rd, | (1.1) |
which is also called parabolic Anderson model. Here, parameter θ>0 and V is a centered generalized Gaussian field which is defined by the Gaussian family {⟨V,φ⟩;φ∈S(Rd)} with mean zero and covariance
E[⟨V,ϕ⟩⟨V,ψ⟩]=∫Rd∫Rdϕ(x)ψ(y)k(x,y)dxdy,∀ϕ,ψ∈S(Rd), | (1.2) |
where S(Rd) is the Schwartz space, and k(x,y) is a symmetric positive definite kernel function. We assume that there exists a constant C>0 such that for almost everywhere (x,y)∈R2d,
|k(x,y)|≤C(γh(x−y)+1). | (1.3) |
Here, γh is a nonnegative and nonnegative definite function which satisfies that γh(x)∈L1loc(Rd), and there exists a α∈(0,2∧d) such that γh(rx)=r−αγh(x) for all r>0.
There exist many Gaussian fields satisfying (1.3). For example, the stationary case includes Bessel field [1], Gaussian field with Riesz potential covariance [2], and fractional white noise [3] (Hurst parameters Hi∈(1/2,1) for 1≤i≤d), while the nonstationary case partly includes 2-d massive free field [4] and log-correlated Gaussian field [5]. In these fields, the covariances of Gaussian field with Riesz potential covariance and fractional white noise are homogeneous themselves, and γh in (1.3) can be taken as them. The covariance of Bessel field is represented as the Bessel function Gb(x), which is not homogeneous but satisfies the asymptotic behaviours for when x→0,
Gb(x)∼{Γ(d−b2)2bπb/2|x|b−d,if0<b<d,12d−1πd/2ln1|x|,ifb=d,Γ(b−d2)2bπb/2,ifb>d, | (1.4) |
when |x|→∞, Gb(x)∼(2d+b−12πd−12Γ(b2))−1|x|b−1−d2e−|x|. The covariances of 2-d massive free field and log-correlated Gaussian field satisfy that when x→y, k(x,y)∼ln1|x−y|, which are bounded away from the diagonal region {x=y}. It can be observed that the covariance of Bessel field (0<b<d) is asymptotically homogeneous, where it requires that b>d−2 such that (1.3) is satisfied when α=d−b; the covariances of Bessel field (b≥d), 2-d massive free field, and log-correlated Gaussian field are bounded or asymptotically logarithmic, satisfying (1.3) for all α sufficiently closed to 0. In addition, we can construct a series of nonstationary fields satisfying (1.3) by setting V(x)=g(x)˜V(x) for the nontrivial, bounded, and measurable function g and stationary Gaussian field ˜V satisfying (1.3).
At present, the rough initial conditions are getting more and more attention in the field of stochastic partial differential equations. Bertini and Giacomin [6] focused on the initial conditions with growing tails in stochastic Burgers and Kardar-Parisi-Zhang (abbr. KPZ) equations. Amir, Corwin, and Quastel [7] utilized the Dirac δ initial condition (or narrow wedge initial conditions) to study the distribution of stochastic heat (or KPZ) equations. Until the publishing of [8], Chen and Dalang first introduced and studied the rough initial conditions for the nonlinear stochastic heat equation, which are quite extensive, including Dirac δ measure, non-tempered measure with exponentially growing tails, etc.
For (1.1), we consider the rough initial condition: the initial value u0 is a Borel measure on Rd owing a Jordan decomposition u0=u+0−u−0. Let |u0|:=u+0+u−0 be the variation measure of u0. We assume that for t>0 and x∈Rd,
pt∗|u0|(x):=∫Rdpt(x−y)|u0|(dy)<∞, | (1.5) |
where "∗" represents the convolution and pt(x):=(2πt)−d/2exp{−|x|2/(2t)} is the usual heat kernel function. It is worth noting that due to the temporal continuity of pt(x) on (0,∞), condition (1.5) implies that for 0<δ<T and x∈Rd,
supt∈[δ,T]pt∗|u0|(x)<∞. | (1.6) |
There have been many results for the Hölder continuity of the stochastic heat equation in the Itô-Skorokhod integral and rough initial conditions, such as [9,10,11,12,13]. In the earlier literatures [8,14], Chen and Dalang studied the continuity for the nonlinear stochastic heat and fractional heat equations with rough initial conditions in the Itô-Skorokhod integral, including the parabolic Anderson model. In Chen and Huang [15], the time-space Hölder continuity was established for nonlinear stochastic heat equations driven by time-white and space-colored Gaussian fields, with rough initial conditions concerning Itô-Skorokhod integral. However, the published papers about Hölder continuity in the Stratonovich sense are not as rich as in the Itô-Skorokhod sense due to the technical complexity. When initial value u0≡1, Hu, Huang, Nualart and Tindel [16] proved the time-space Hölder continuity for the stochastic heat equation driven by time-space stationary Gaussian fields in the Stratonovich integral. For the similar model, under the rough initial condition, Lyu [17] obtained the spatial Hölder continuity in the case of time-space stationary Gaussian fields, which are homogeneous on space. Later, Lyu and Li [18] proved the time-space Hölder continuity for time-independent log-correlated Gaussian field and initial value u0≡1. As far as we know, there are very few results for temporal Hölder continuity in the case of nonstationary Gaussian field and rough initial condition.
In this paper, under the conditions (1.3) and (1.5), we tend to prove the temporal Hölder continuity for the Feynman-Kac formula of (1.1) in the Stratonovich integral. According to [[17], Lemma 3.1], the Feynman-Kac formula is a mild solution to (1.1) in the Stratonovich integral. As mentioned in [16], the path-wise solution in the Young integral can be viewed as a version of the Feynman-Kac formula in the Stratonovich integral. Thus, to obtain the Hölder continuity in the Stratonovich sense, we only need to prove the Hölder continuity in the Young sense. However, the strategy is usually unsuccessful for the rough initial condition.
According to (5.13) in [16], when the initial value u0 belongs to the weighted Besov-Hölder space Bκ,eλ∞,∞(Rd) (κ∈(0,1)), it was obtained as the temporal Hölder continuity of solution in the sense of the norm of Bκu,wt∞,∞(Rd) (κu∈(κ,1)). Because the weighted Besov space Bκu,wt∞,∞ coincides with the weighted Hölder space Cκu(Rd;wt), we can directly obtain the temporal Hölder continuity in the point-wise sense. Unluckily, if u0 is a measure, it usually does not belong to Bκ,eλ∞,∞(Rd) (κ∈(0,1)), such as Dirac δ0∈B−d(1−1/q),eλq,∞ (q∈[1,∞]) but ∉Bκ,eλ∞,∞(Rd) (κ∈(0,1)). When u0 belongs to the Besov space on torus B−κq,∞ (κ∈[0,1/2)), by reference to [19,20], the temporal Hölder continuity of solution was obtained in the sense of the norm of Bκuq,∞(Td) (κu∈(κ,1)), but q cannot arrive at infinity in solution space Bκuq,∞(Td). This leads to that we still have no way to prove the temporal Hölder continuity in the point-wise sense.
Instead of the above method, we directly prove the Hölder continuity for the Feynman-Kac formula by the Kolmogorov continuity theorem. It has been known that under the rough initial condition, the previous Feynman-Kac formula based on Brownian motion is not well-defined any more. Hence, we will use the Feynman-Kac formula based on Brownian bridge. In the earlier work [21], Chen, Hu, and Nualart proved the Feynman-Kac formula for the nonlinear stochastic heat equation on R in the Itô-Skorokhod integral with time-space white noise and rough initial conditions. Hu, Nualart, and Song [3] (also see [16]) obtained the Feynman-Kac formula for the stochastic heat equation driven by time-space Gaussian fields with function-valued initial data in the Itô-Skorokhod and Stratonovich integral. After it, Huang, Lê, and Nualart [22] obtained the Feynman-Kac moment representation based on Brownian bridge for the stochastic heat equation in the Itô-Skorokhod integral, driven by time-white Gaussian fields with rough initial conditions. Inspired by it, Lyu [17] proved the Feynman-Kac formula for the stochastic heat equation in the Stratonovich integral, with time-space Gaussian fields and rough initial condition. Similarly, this paper also obtained the Feynman-Kac formula based on Brownian bridge uθ(t,x) defined in (2.1) in the case of nonstationary Gaussian field and rough initial condition, but the Feynman-Kac moment representation of uθ(t,x) that we get in (2.6) is different from the representation in [17].
Different from Brownian motion and stationary Gaussian field, the computations of Hölder continuity are complex in the case of Brownian bridge and nonstationary Gaussian field. To overcome the difficulty, on the one hand, we construct a novel decomposition of Brownian bridge in Lemma 2.5; on the other hand, because the technique of Fourier transform cannot be directly applied to estimate positive definite kernel k(x,y), we will use the estimates of the heat kernel in Lemma 2.1.
We state the temporal Hölder continuity of the Feynman-Kac formula uθ(t,x) in (2.1) as follows.
Theorem 1.1. Assume that conditions (1.3) and (1.5) hold. Set 0<δ<1≤T and β∈(0,1−α/2), where α is taken from (1.3). Then, there exists some constant C>0 such that for all θ>0, t,s∈[δ,T], x∈Rd, and integer n≥1,
E|uθ(t,x)−uθ(s,x)|n≤CneCθ2n2T2exp{Cθ42−αT4−α2−αn4−α2−α}((2n−1)!!)1/2T(βd/2+1)n⋅δ−(d/2+1)βn(supr∈[δ,T/(1−β)]pr∗|u0|(x))n|t−s|βn/2. | (1.7) |
Moreover, there exists a temporal β2-Hölder continuous modification of uθ(t,x) on (0,∞).
As an extension of temporal Hölder continuity in [[16], Theorem 4.12], where the Gaussian fields are stationary and initial value u0≡1, Theorem 1.1 contains the case of nonstationary Gaussian fields and initial value of measure. However, patient readers may observe from Theorem 1.1 that when initial value u0≡1, on the one hand, the order of Hölder continuity is not optimal, where β/2<1/2; on the other hand, the Hölder continuity of the solution is limited on open interval (0,∞) excluding the zero point. For this reason, we intend to make some technical explanations as follows:
(1) Because the measure-valued initial data u0 is considered, we choose to use the Feynman-Kac formula based on Brownian bridge (2.1). In the estimates of the Hölder continuity, (2.1) leads to the need to utilize the continuity of bridge B0,t with respect to t; see Proposition 5.3. Here, remark that the continuity of ∫s0V(Bx,z0,t(r))dr at the t=s point is necessary for our estimates. If we consider the Feynman-Kac formula based on Brownian motion with function-valued initial data, then the continuity of the term can be bypassed. So, when u0≡1, the order of Hölder continuity is low in Theorem 1.1.
(2) Under condition (1.5), the proof of Hölder continuity can only depend on the regularity of heat kernel pt(x) rather than of u0. However, in the step of estimates of the heat kernel, the terms t and s with negative power are produced; see Lemma 2.1. For the Feynman-Kac formula (2.1), in the computations of (5.37)–(5.39), we have no way to get rid of the term (t−d/2−1+s−d/2−1)βn produced in estimates of the heat kernel. Moreover, we obtain an additional term δ−(d/2+1)βn in (1.7) relative to the estimates of moment in Proposition 4.1, which implies that δ cannot tend to 0. Thus, when u0≡1, the coefficient in the right side of (1.7) is not exact, such that the Hölder continuity cannot be proved at zero point.
In order to compensate the defect of Theorem 1.1 in the case of function-valued initial data, we specifically show the following result in which the Hölder continuity is extended to the zero point.
Theorem 1.2. Under condition (1.3), the following results hold:
(i) When initial value u0 is a κ-Hölder continuous function in Cκ(Rd) with κ∈(0,1], for ρ∈(0,κ), θ>0, and x∈Rd, there exists a modification of uθ(t,x), which is ρ2-Hölder continuous on [0,∞).
(ii) When initial value u0 is a constant, that is, u0≡C, for ν∈(0,1−α/4), θ>0, and x∈Rd, there exists a modification of uθ(t,x), which is ν-Hölder continuous on [0,∞).
The order of Hölder continuity in Theorem 1.2 (ⅰ) coincides with it in [8,15], though their settings are different from ours, where they considered the Itô-Skorokhod integral and time-white Gaussian fields which are colored in space. The order in Theorem 1.2 (ⅱ) is the same as it is in [[16], Theorem 4.12].
Next, we make some comments on the results in Theorems 1.1 and 1.2 as follows:
(a) With respect to the special fields, including Bessel field (b≥d), 2-d massive free field, and log-correlated Gaussian field, the orders of Hölder continuity are sufficiently closed to 1/2 and 1 in Theorem 1.1 and Theorem 1.2 (ⅱ), respectively, because these fields always satisfy (1.3) for all small α. For the special fields, we can prove a more precise modulus of continuity in Theorem 1.1 than (1.7), which is similar to [[18], Proposition 2.4]. However, the more precise modulus does not impact on the order of Hölder continuity. For the homogeneous or asymptotically homogeneous Gaussian fields, like Gaussian field with Riesz potential covariance, fractional white noise, and Bessel field (0<b<d), the orders of Hölder continuity in Theorems 1.1 and 1.2 are optimal within our framework if we take α equal to the (asymptotically) homogeneous degree of these fields in condition (1.3).
(b) For Theorem 1.2, the Hölder continuity in (ⅰ) is limited by the Hölder continuity of u0. Though the initial condition in (ⅱ) is a special case in (ⅰ), the order in (ⅱ) is obviously higher than it is in (ⅰ), i.e., ν>1/2>ρ/2. It is found that, different from (ⅱ), the Hölder continuity in (ⅰ) is only determined by the regularity of u0 by comparing (5.41) and (5.47) in the proof of Theorem 1.2.
(c) Notice that the order in Theorem 1.2 (ⅰ) is not necessarily higher than Theorem 1.1, because the Hölder continuity at zero point is considered in Theorem 1.2 (ⅰ). To sum up Theorem 1.1 and Theorem 1.2 (ⅰ), the order of Hölder continuity on (0,∞) is (β∨ρ)/2 when u0∈Cκ(Rd). On the other hand, it is found that the order in Theorem 1.1 is lower than it is in Theorem 1.2 (ⅱ), i.e., β/2<1/2<ν. Obviously, the initial condition is very special in Theorem 1.2 (ⅱ).
Methodology: In the sense of the Stratonovich integral, our method heavily depends on the Feynman-Kac formula based on the Brownian bridge (2.1) and Feynman-Kac formula based on Brownian motion (2.4), which produce the different Hölder continuities and modulus of continuity in Theorems 1.1 and 1.2. Meanwhile, our method can only be applied to the linear model. However, in the sense of the Itô-Skorokhod integral, the method in [8,13,14,15] can cover the case of the nonlinear model, the advantages of which are that the estimates of Hölder continuity are stable for rough and regular initial conditions. In fact, the above settings of the integral are different, and our method mainly compensates the lack of result in the Stratonovich integral (or Young integral) rather than the Itô-Skorokhod integral.
Organisation: Section 2 is the preliminaries about Fourier transform, estimates of heat kernel, and Brownian bridge. In Section 3, we give the definitions of the Feynman-Kac formula, Feynman-Kac functional, and Feynman-Kac moment representation. In Section 4, we show the well-definiteness and moment estimates of the Feynman-Kac functional and Feynman-Kac formula. Section 5 is the proof of temporal Hölder continuity in Theorem 1.1.
Notations: Write R+:=[0,∞) and N+:={1,2,3,⋯}. Let (Ω,F,P) be the probability space with expectation E. Set p∈[1,∞], and denote the Lebesgue space on (Ω,F,P) by Lp(Ω). For region D⊆Rd, let Lp(D) be Lebesgue space on D. Denote by L1loc(Rd) the space composed of locally integrable functions on Rd. For κ∈(0,1], Cκ(Rd) is the space composed of κ-Hölder continuous functions. S(Rd) is Schwartz space on Rd, and its dual space S′(Rd) is the space of tempered distributions. Let C be a universal nonnegative constant. f≲g represents that there is a constant C>0 not dependent on variables such that f≤Cg.
Fourier transform: The Fourier transform of a function f∈S(Rd) is defined as
Ff(ξ):=∫Rdeiξ⋅xf(x)dx, |
and the inverse Fourier transform is given by F−1f(ξ)=(2π)−dFf(−ξ). The generalized Fourier transform of f∈S′(Rd) is defined by the dual
⟨Ff,g⟩=⟨f,Fg⟩,∀g∈S(Rd). | (2.1) |
For nonnegative definite function γh in (1.3), according to the Bochner theorem (e.g., p.158, [23]), there exists a nonnegative and symmetric tempered measure μh such that γh=Fμh. Noticing that γh(x) is a function, it is found that
γh(x)=∫Rdeiξ⋅xμh(dξ),a.e., | (2.2) |
by (2.1) and the Fubini theorem. Because γh satisfies that γh(rx)=r−αγh(x) for all r>0, μh is homogeneous, that is, μh(d(rξ))=rαμh(dξ) for all r>0.
Estimates of heat kernel: We give the estimates of heat kernel used to prove the Hölder continuity. The results similar to (ⅰ) and (ⅲ) in Lemma 2.1 have been proved in [[15], Lemma 3.1], but our proof is slightly different from [15] in details.
Lemma 2.1. For the heat kernel pt(x)=(2πt)−d/2exp{−|x|2/(2t)}, the following results hold.
(i) For all x,y∈Rd and t>0, it holds that
|pt(x)−pt(y)|≲t−(d+1)/2|x−y|. | (2.3) |
(ii) For all z1,z2,x,y∈Rd, and t>0, it holds that
|pt(z1+x)−pt(z1+y)−pt(z2+x)+pt(z2+y)|≲t−d/2−1|z1−z2||x−y|. | (2.4) |
(iii) For all x∈Rd and t,s>0, it holds that
|pt(x)−ps(x)|≲(t−d/2−1+s−d/2−1)|t−s|. | (2.5) |
Proof. (ⅰ) By pt=F−1e−t2|⋅|2, the inequality |eiξ⋅y−eiξ⋅x|≤|ξ||x−y|, and the integral substitution, we have
|pt(x)−pt(y)|=(2π)−d|∫Rd(e−iξ⋅x−e−iξ⋅y)e−t|ξ|2/2dξ|≤(2π)−d∫Rd|ξ|e−t|ξ|2/2dξ|x−y|≲t−(d+1)/2|x−y|. | (2.6) |
(ⅱ) According to the arguments similar to (2.6), it holds that
|pt(z1+x)−pt(z1+y)−pt(z2+x)+pt(z2+y)|=(2π)−d|∫Rd(e−iξ⋅z1−e−iξ⋅z2)(e−iξ⋅x−e−iξ⋅y)e−t|ξ|2/2dξ|≤(2π)−d∫Rd|ξ|2e−t|ξ|2/2dξ|z1−z2||x−y|≲t−d/2−1|z1−z2||x−y|. |
(ⅲ) From pt=F−1e−t2|⋅|2, |eiξ⋅x|=1, and the inequality |ea−eb|≤|a−b|(ea+eb), it implies that
|pt(x)−ps(x)|=(2π)−d|∫Rde−iξ⋅x(exp{−t2|ξ|2}−exp{−s2|ξ|2})dξ|≤2−d−1π−d∫Rd|ξ|2(exp{−t2|ξ|2}+exp{−s2|ξ|2})dξ|t−s|=2−d−1π−d(t−d/2−1+s−d/2−1)∫Rd|ξ|2exp{−12|ξ|2}dξ|t−s|≲(t−d/2−1+s−d/2−1)|t−s|, |
where the second to last step is due to the integral substitution.
So, we complete the proof.
Lemma 2.2. Under condition (1.3), for β>0, there exist some C>0 dependent on α and β such that for all t>0,
∫Rd(γh(y+x)+1)pβt(y)dy≤Ct(1−β)d/2(t−α/2+1). | (2.7) |
Proof. By the spherical substitution, γh∈L1loc(Rd), and γh(tx)=t−αγh(x) (α∈(0,2∧d)), it gives that
∫Rdγh(y)pβ1(y)dy=(2π)−βd/2∫∞0r−α+d−1e−βr2/2dr∫{|y|=1}γh(y)dS<∞. | (2.8) |
By the facts that γh=Fμh and Fpt(ξ)=e−t|ξ|22, pβt(x)=(2π)(1−β)d/2β−d/2t(1−β)d/2pt/β(x), and |eia|=1,
∫Rd(γh(y+x)+1)pβt(y)dy=(2π)(1−β)d/2β−d/2t(1−β)d/2(∫Rdγh(y+x)pt/β(y)dy+1)=(2π)(1−β)d/2β−d/2t(1−β)d/2(∫Rdeiξ⋅xexp{−t2β|ξ|2}μh(dξ)+1)≤(2π)(1−β)d/2β−d/2t(1−β)d/2(∫Rdexp{−t2β|ξ|2}μh(dξ)+1)=(2π)(1−β)d/2β−d/2t(1−β)d/2(∫Rdγh(y)pt/β(y)dy+1). | (2.9) |
Moreover, using the integral substitution, γh(tx)=t−αγh(x), and (2.8), it gives that
∫Rd(γh(y+x)+1)pβt(y)dy≤(2π)(1−β)d/2β−d/2t(1−β)d/2((t/β)−α/2∫Rdγh(y)p1(y)dy+1)≤Ct(1−β)d/2(t−α/2+1). | (2.10) |
Thus, (2.7) is proved.
Brownian bridge: Let B(s) or Bs be a d-dimensional standard Brownian motion on R+, which is independent of V. Set Bxs:=Bs+x as a Brownian motion starting from point x∈Rd. Moreover, for t>0, the d-dimensional standard Brownian bridge is defined as
B0,t(s):=Bs−stBt,∀s∈[0,t]. | (2.11) |
For 0≤s≤t and x,y∈Rd, write ax,ys,t:=t−stx+sty. Based on the notations B0,t and ax,ys,t, the Brownian bridge from x to y is defined as
Bx,y0,t(s):=B0,t(s)+ax,ys,t,∀s∈[0,t]. | (2.12) |
Write B0=B and B0,00,t=B0,t without ambiguity.
By the relation that ax,ys,t=ay,xt−s,t and the computations of covariances, it can be directly checked that the following two elementary lemmas hold.
Lemma 2.3. {Bx,y0,t(s)}s∈[0,t] is identically distributed as {By,x0,t(t−s)}s∈[0,t].
Lemma 2.4. {Bx,y0,t(s)}s∈[0,t] is independent of {Bx(s)}s≥t.
Based on Lemma 2.4, we obtain a decomposition of the Brownian bridge.
Lemma 2.5. For 0<t2<t1 and 0≤r≤t2, let
Gt2,t1:=B(t2)t2−B(t1)t1. | (2.13) |
Then,
B0,t1(r)=B0,t2(r)+rGt2,t1, | (2.14) |
where Gt2,t1 is independent of {B0,t2(r)}r∈[0,t2] and Gt2,t1∼N(0,t1−t2t2t1).
Lemma 2.6. Let F be a nonnegative measurable functional on C([0,λt]), where C([0,λt]) is the space composed of continuous functions on [0,λt] for t>0 and λ∈(0,1). Then,
EF({B0,t(s)}0≤s≤λt)≤(1−λ)−d/2EF({B(s)}0≤s≤λt). | (2.15) |
Proof. Using [[22], (2.8)] in the case of x=y=0 and the nonnegativity of F, we obtain
EF({B0,t(s)}0≤s≤λt)=(1−λ)−d/2E[F({B(s)}0≤s≤λt)exp{−|B(λt)|22(1−λ)t}]≤(1−λ)−d/2EF({B(s)}0≤s≤λt). | (2.16) |
Thus, the proof is completed.
When u0 is a measure satisfying (1.5), we consider the following Feynman-Kac formula:
uθ(t,x):=∫RdEBexp{θ∫t0V(Bx,y0,t(s))ds}pt(y−x)u0(dy), | (2.1) |
for (t,x)∈R+×Rd. Here, Bx,y0,t is the d-dimensional Brownian bridge from x to y, and the integral ∫t0V(Bx,y0,t(s))ds is defined as a L2(Ω)-limit, that is,
∫t0V(Bx,y0,t(s))ds:=limε→0∫t0Vε(Bx,y0,t(s))ds,∀(t,x,y)∈R+×R2d, | (2.2) |
where we set Vε(x):=⟨V(⋅),p2ε(x−⋅)⟩ with p2ε(x)=(4πε)−d/2e−|x|2/(4ε). To simplify it, we also use the notation
ˆVx,y(t)=∫t0V(Bx,y0,t(s))ds. | (2.3) |
We will prove the well-definiteness of ˆVx,y(t) in Lemma 3.1. Based on it, if the Feynman-Kac formula uθ(t,x) is a L1(Ω)-integrable stochastic process, we call uθ(t,x) well-defined, which will be proved in Corollary 4.2.
When u0 is a measurable function, (2.1) is rewritten as
uθ(t,x):=EB[exp{θ∫t0V(Bxs)ds}u0(Bxt)], | (2.4) |
where Bxt is a d-dimensional Brownian motion at starting point x∈Rd, and the integral ∫t0V(Bxs)ds is similarly defined, like (2.2).
Let EV be the expectation with respect to V, and EB be the expectation with respect to B. Then, by the independence between V and B, E can be represented as EB⊗EV. Conditioning on the Brownian motion, ˆVx,y(t) is a centered Gaussian process with conditional covariance
EV[ˆVx,y1(t)ˆVx,y2(t)]=∫t0∫t0k(Bx,y10,t(s),Bx,y20,t(r))dsdr,∀y1,y2∈Rd. | (2.5) |
Let {Bj;j=1,⋯,n} be a family of d-dimensional independent standard Brownian motions for positive integer n. Set {Bx,yj,0,t(s):=Bj(s)−stBj(t)+ax,ys,t,∀s∈[0,t];j=1,⋯,n} as a family of independent Brownian bridges from x to y. Then, based on (2.1) and (2.5), the n-order Feynman-Kac moment representation satisfies that
Eunθ(t,x)=∫RdnEexp{θ22n∑j,k=1∫t0∫t0k(Bx,yjj,0,t(s),Bx,ykk,0,t(r))drds}n∏j=1pt(yj−x)u0(dy1)⋯u0(dyn). | (2.6) |
Similar to (2.1), for (t,s,x)∈(0,∞)2×Rd, we define the Feynman-Kac functional ˉuθ(t,s,x) as
ˉuθ(t,s,x):=∫RdEBexp{θˆVx,y(t)}ps(y−x)|u0|(dy). | (2.7) |
When s=t, we write ˉuθ(t,x):=ˉuθ(t,t,x). Through (2.5) and (2.7), we can obtain the n-order moment representation
Eˉunθ(t,s,x)=∫RdnEexp{θ22n∑j,k=1∫t0∫t0k(Bx,yjj,0,t(s),Bx,ykk,0,t(r))drds}n∏j=1ps(yj−x)|u0|(dy1)⋯|u0|(dyn). | (2.8) |
Lemma 3.1. If condition (1.3) holds, then ˆVx,y(t) in (2.2) is well-defined.
Proof. By the similar method to [[16], Proposition 4.2.] and [[17], Proposition 3.1.], we only need to show that for T>0,
supε>0sup(t,x,y)∈[0,T]×R2dE|∫t0Vε(Bx,y0,t(s))ds|2<∞. | (2.9) |
In fact, using the inequality |a+b|2≤2|a|2+2|b|2 and the integral substitution, we obtain
E|∫t0Vε(Bx,y0,t(s))ds|2≤2E|∫t/20Vε1(Bx,y0,t(s))ds|2+2E|∫tt/2Vε1(Bx,y0,t(s))ds|2≤2E|∫t/20Vε1(Bx,y0,t(s))ds|2+2E|∫t/20Vε1(Bx,y0,t(t−s))ds|2≤2E|∫t/20Vε(Bx,y0,t(s))ds|2+2E|∫t/20Vε1(By,x0,t(s))ds|2, | (2.10) |
where the last step is due to {By,x0,t(s)}s∈[0,t]d={Bx,y0,t(t−s)}s∈[0,t]. Notice that the above two terms are similar, and we only need to show the estimates of the first term. Recall that ax,ys,t=t−stx+sty. Then, by Lemma 2.6 for {B0,t(s)}s∈[0,t/2] and the integral substitution, we have
E|∫t/20Vε(Bx,y0,t(s))ds|2≤2d/2EB[EV|∫t/20Vε1(B(s)+ax,ys,t)ds|2]≤2d/2E∫t/20∫t/20∫Rd∫Rdk(x1+B(s)+ax,ys,t,y1+B(r)+ax,yr,t)pε(x1)pε(y1)dx1dy1dsdr≲2d/2E∫t/20∫t/20∫Rd∫Rd(γh(x1+B(s)+ax,ys,t−y1−B(r)−ax,yr,t)+1)pε(x1)pε(y1)dx1dy1dsdr≲∫t/20∫t/20∫RdEeiξ⋅(B(s)−B(r))eiξ⋅(ax,ys,t−ax,yr,t)e−ε|ξ|2μh(dξ)dsdr+t2≲∫t/20∫t/20∫Rde−12|s−r||ξ|2μh(dξ)dsdr+t2≲t2−α2+t2, | (2.11) |
where the second to last step is due to (2.2), Fp2ε(ξ)=e−ε|ξ|2, and |eia|=1, and the last step is due to (2.8), γh=Fμh, and μh(d(rξ))=rαμh(dξ) for all r>0.
Finally, substituting (2.11) into (2.10), and by α<2, we can obtain that
supε>0sup(t,x,y)∈[0,T]×R2dE|∫t0Vε(Bx,y0,t(s))ds|2≲T2−α2+T2, | (2.12) |
which shows that (2.9) holds.
Lemma 4.1. Under condition (1.3), there exist some C>0 dependent on k(x,y) such that for all θ,t>0 and n∈N+,
Eexp{θ2n∑j,k=1∫t0∫t0γh(Bj(s)−Bk(r))drds}≤Cnexp{Cθ42−αt4−α2−αn4−α2−α}. | (2.1) |
Proof. By (2.2), the Jensen inequality, and the independence of {Bj}1≤j≤n, we have
Eexp{θ2n∑j,k=1∫t0∫t0γh(Bj(s)−Bk(r))dsdr}=Eexp{θ2n2∫Rd|1nn∑j=1∫t0eiξ⋅Bj(s)ds|2μh(dξ)}≤Eexp{θ2nn∑j=1∫Rd|∫t0eiξ⋅Bj(s)ds|2μh(dξ)}≤(Eexp{θ2n∫Rd|∫t0eiξ⋅B(s)ds|2μh(dξ)})n. | (2.2) |
By Brownian scaling {B(rs)}s∈R+d={r1/2B(s)}s∈R+ and μh(d(rξ))=rαμh(dξ) for any r>0 and the integral substitution, we find that for r>0,
∫Rd|∫t0eiξ⋅B(s)ds|2μh(dξ)d=rα2−2∫Rd|∫rt0eiξ⋅B(s)ds|2μh(dξ). | (2.3) |
Set process At:=∫Rd|∫t0eiξ⋅B(s)ds|2μh(dξ). Then, taking r=(θ2nt)22−α and using (2.3), we obtain
Eexp{θ2nAt}=Eexp{(rt)−1Art}. | (2.4) |
Using the similar methods to [[17], (3.20)], we find that there exist some C>0 such that for all t>0,
Eexp{(rt)−1Art}≤CeCrt, | (2.5) |
by Lemma 2.2 in [17]. At last, summing up (2.2), (2.4), and (2.5), and using (θ2nt)22−α instead of r, the proof of (2.5) can be completed.
Proposition 4.1. Under conditions (1.3) and (1.5), there exist some C>0 such that for all t,s,θ>0, x∈Rd, and n∈N+,
Eˉunθ(t,s,x)≤CneCθ2n2t2exp{Cθ42−αt4−α2−αn4−α2−α}(ps∗|u0|(x))n. | (2.6) |
Proof. By (2.8) and (1.3), we obtain
Eˉunθ(t,s,x)≤eCθ2n2t2∫RdnEexp{Cθ2n∑j,k=1∫t0∫t0γh(Bx,yjj,0,t(s)−Bx,ykk,0,t(r))drds}⋅n∏j=1ps(yj−x)|u0|(dy1)⋯|u0|(dyn). | (2.7) |
By (2.2) and the inequality |a+b|2≤2|a|2+2|b|2, we obtain
n∑j,k=1∫t0∫t0γh(Bx,yjj,0,t(s)−Bx,ykk,0,t(r))drds≤2∫Rd|n∑j=1∫t/20eiξ⋅Bx,yjj,0,t(s)ds|2μh(dξ)+2∫Rd|n∑j=1∫tt/2eiξ⋅Bx,yjj,0,t(s)ds|2μh(dξ). | (2.8) |
In addition, by the integral substitution and {Bx,yj,0,t(s)}s∈[0,t]d={By,xj,0,t(t−s)}s∈[0,t], we have
∫Rd|n∑j=1∫tt/2eiξ⋅Bx,yjj,0,t(s)ds|2μh(dξ)d=∫Rd|n∑j=1∫t/20eiξ⋅Byj,xj,0,t(s)ds|2μh(dξ). | (2.9) |
Recall that ax,ys,t=t−stx+sty. To substitute (2.8) and (2.9) into (2.7), and by using (2.2) and the Cauchy-Schwartz inequality, we obtain
Eˉunθ(t,s,x)≤eCθ2n2t2(∫RdnEexp{Cθ2n∑j,k=1∫t/20∫t/20γh(Bj,0,t(s)−Bk,0,t(r)+ax,yjs,t−ax,ykr,t)drds}⋅n∏j=1ps(yj−x)|u0|(yj)dy1⋯dyn)1/2⋅(∫RdnEexp{Cθ2n∑j,k=1∫t/20∫t/20γh(Bj,0,t(s)−Bk,0,t(r)+ayj,xs,t−ayk,xr,t)drds}⋅n∏j=1ps(yj−x)|u0|(yj)dy1⋯dyn)1/2. | (2.10) |
Let a(s,r,t,x,y,z) be a measurable fucntion from R3+×R3d to Rd. We claim that for all t,θ>0 and x,y1,⋯,yn∈Rd, it holds that
Eexp{θ2n∑j,k=1∫t0∫t0γh(Bj,0,t(s)−Bk,0,t(r)+a(s,r,t,x,yj,yk))drds}≤Eexp{θ2n∑j,k=1∫t0∫t0γh(Bj,0,t(s)−Bk,0,t(r))dsdr}. | (2.11) |
In fact, through the Taylor expansion, we only need to compare their m-order moments. Precisely, using (2.2), we find that for any positive integer m,
E[n∑j,k=1∫t0∫t0γh(Bj,0,t(s)−Bk,0,t(r)+a(s,r,t,x,yj,yk))dsdr]m=∫Rdm∫[0,t]m∫[0,t]mn∑jl,⋯,jm=1n∑kl,⋯,km=1Em∏l=1eiξl⋅(Bjl,0,t(sl)−Bkl,0,t(rl))⋅m∏l=1eiξl⋅a(s,r,t,x,yj,yk)ds1⋯dsmdr1⋯drmμh(dξ1)⋯μh(dξm)≤E[n∑j,k=1∫t0∫t0γh(B0,t(s)−B0,t(r))dsdr]m. |
Here in the last inequality, we have used |eia|=1, the nonnegativity of μh, and the fact that
Em∏j=1eiξj⋅(Bjl,0,t(sl)−Bkl,0,t(rl))=exp{−12Var(m∑j=1ξj⋅(Bjl,0,t(sl)−Bkl,0,t(rl)))}≥0. |
Then, by (2.10), (2.11), and Lemma 2.6, we obtain
Eˉunθ(t,s,x)≤eCθ2n2t2Eexp{Cθ2n∑j,k=1∫t/20∫t/20γh(Bj,0,t(s)−Bk,0,t(r))drds}(ps∗|u0|(x))n≤2d/2eCθ2n2t2Eexp{Cθ2n∑j,k=1∫t/20∫t/20γh(Bj(s)−Bk(r))drds}(ps∗|u0|(x))n≤CneCθ2n2t2exp{Cθ42−αt4−α2−αn4−α2−α}(ps∗|u0|(x))n, | (2.12) |
where the last step is due to Lemma 4.1. Hence, we complete the proof of (2.6).
Corollary 4.1. Under conditions (1.3) and (1.5), there exist some C>0 such that for all t,θ>0, x∈Rd, and n∈N+,
E|uθ(t,x)|n≤CneCθ2n2t2exp{Cθ42−αt4−α2−αn4−α2−α}(pt∗|u0|(x))n. | (2.13) |
Proof. By the Cauchy-Schwartz inequality and (2.6), it is readily checked that
E|uθ(t,x)|n≤(Eu2nθ(t,x))1/2≤(Eˉu2nθ(t,x))1/2. | (2.14) |
Recalling ˉuθ(t,x):=ˉuθ(t,t,x), and by (2.14) and Proposition 4.1, we complete the proof of (2.13).
By Proposition 4.1 and Corollary 4.1, we directly obtain the following result.
Corollary 4.2. Under conditions (1.3) and (1.5), for t,s>0, x∈Rd, and n∈N+, ˉuθ(t,s,x) and uθ(t,x) are well-defined as the Ln(Ω)-integrable stochastic processes.
In this section, we will prove Theorem 1.1. Before it, the following results are required.
Proposition 5.1. Under conditions (1.3) and (1.5), for all t≥s>0, n∈N+, and x∈Rd,
E|∫RdEB[exp{ˆVx,z(t)}−exp{ˆVx,z(s)}]pt(z−x)u0(dz)|n≤2n−1θn((2n−1)!!)1/2{(Eˉun2θ(t,x))1/2+(Eˉun2θ(s,t,x))1/2}⋅(∫RdE|ˆVx,z(t)−ˆVx,z(s)|2pt(z−x)|u0|(dz))n/2. | (2.1) |
Proof. Using the inequalities |ea−eb|≤|a−b|(ea+eb), (|a|+|b|)n≤2n−1(|a|n+|b|n) and the Cauchy-Schwartz inequality, we obtain
E|∫RdEB[exp{ˆVx,z(t)}−exp{ˆVx,z(s)}]pt(z−x)u0(dz)|n≤θnE[∫RdEB[(exp{θˆVx,z(t)}+exp{θˆVx,z(s)})|ˆVx,z(t)−ˆVx,z(s)|]pt(z−x)|u0|(dz)]n≤2n−1θnE[∫RdEB[exp{θˆVx,z(t)}|ˆVx,z(t)−ˆVx,z(s)|]pt(z−x)|u0|(dz)]n+2n−1θnE[∫RdEB[exp{θˆVx,z(s)}|ˆVx,z(t)−ˆVx,z(s)|]pt(z−x)|u0|(dz)]n≤2n−1θnE[(∫RdEBexp{2θˆVx,z(t)}pt(z−x)|u0|(dz))1/2⋅(∫RdEB|ˆVx,z(t)−ˆVx,z(s)|2pt(z−x)|u0|(dz))1/2]n+2n−1θnE[(∫RdEBexp{2θˆVx,z(s)}pt(z−x)|u0|(dz))1/2⋅(∫RdEB|ˆVx,z(t)−ˆVx,z(s)|2pt(z−x)|u0|(dz))1/2]n≤2n−1θn{(Eˉun2θ(t,x))1/2+(Eˉun2θ(s,t,x))1/2}⋅{EV[∫RdEB|ˆVx,z(t)−ˆVx,z(s)|2pt(z−x)|u0|(dz)]n}1/2. | (2.2) |
Using the Minkowsky integral inequality and (conditional) Gaussian variance property, we get
{EV[∫RdEB|ˆVx,z(t)−ˆVx,z(s)|2pt(z−x)|u0|(dz)]n}1/2≤(∫RdEB[EV|ˆVx,z(t)−ˆVx,z(s)|2n]1npt(z−x)|u0|(dz))n/2≤((2n−1)!!)1/2(∫RdE|ˆVx,z(t)−ˆVx,z(s)|2pt(z−x)|u0|(dz))n/2. | (2.3) |
Substituting (2.3) into (2.2), we can complete the proof of (2.1).
Proposition 5.2. Under condition (1.3), there exists a C>0 dependent on α such that for all x,z∈Rd, T>1, and 0≤s≤t≤T,
E|∫tsV(Bx,z0,t(r))dr|2≤CTα/2|t−s|2−α/2. | (2.4) |
Proof. Case I: t/2≤s≤t. Recall that az,xr,t:=t−rtz+rtx and Bz,x0,t(r)=B0,t(r)+az,xr,t. Then, by the integral substitution, {Bx,z0,t(s)}s∈[0,t]d={Bz,x0,t(t−s)}s∈[0,t], and Lemma 2.6, we get
E|∫tsV(Bx,z0,t(r))dr|2=E|∫t−s0V(B0,t(r)+az,xr,t)dr|2≤(ts)d/2E|∫t−s0V(B(r)+az,xr,t)dr|2≤2d/2∫t−s0∫t−s0Ek(B(r1)+az,xr1,t,B(r2)+az,xr2,t)dr1dr2≲∫t−s0∫t−s0E[γh(B(r1)+az,xr1,t−B(r2)−az,xr2,t)+1]dr1dr2≲∫t−s0∫t−s0∫Rd(γh(y+az,xr1,t−az,xr2,t)+1)p|r1−r2|(y)dydr1dr2, | (2.5) |
where the second to last step is due to (1.3). By Lemma 2.2, we have
∫Rd(γh(y+az,xr1,t−az,xr2,t)+1)p|r1−r2|(y)dy≲(|r1−r2|−α/2+1). | (2.6) |
Substituting (2.6) into (2.5), it is obtained that
E|∫tsV(Bx,z0,t(r))dr|2≲∫t−s0∫t−s0(|r1−r2|−α/2+1)dr1dr2≲((1−α/2)−1|t−s|2−α/2+|t−s|2)≤CTα/2|t−s|2−α/2, | (2.7) |
by the relations that s≤t≤T, T>1, and α∈(0,2∧d).
Case II: 0≤s<t/2. From the inequality |a+b|2≤2(|a|2+|b|2), it gives that
E|∫tsV(Bx,z0,t(r))dr|2≤2E|∫tt/2V(Bx,z0,t(r))dr|2+2E|∫t/2sV(Bx,z0,t(r))dr|2. | (2.8) |
Using Lemma 2.6, (1.3), and the integral substitution, we have
E|∫t/2sV(Bx,z0,t(r))dr|2≤2d/2E|∫t/2sV(B(r)+az,xr,t)dr|2≲∫t/2s∫t/2sE[γh(B(r1)+az,xr1,t−B(r2)−az,xr2,t)+1]dr1dr2≲∫t/2−s0∫t/2−s0∫Rd(γh(y+az,xr1+s,t−az,xr2+s,t)+1)p|r1−r2|(y)dydr1dr2≲∫t/2−s0∫t/2−s0∫Rd(γh(y+az,xr1+s,t−az,xr2+s,t)+1)p|r1−r2|(y)dydr1dr2≤CTα/2|t/2−s|2−α/2, | (2.9) |
where we have used the computations similar to (2.7) in the last step.
To combine (2.8) with (2.7) and (2.5), it is found that
E|∫tsV(Bx,z0,t(r))dr|2≤CTα/2(|t/2|2−α/2+|t/2−s|2−α/2)≤CTα/2|t−s|2−α/2. | (2.10) |
So, to sum up (2.7) and (2.10) in the above two cases, we can complete the proof.
Proposition 5.3. Under condition (1.3), set β∈(0,1−α/2), and there exists C>0 dependent on α and β such that for all x,z∈Rd, T>1, and 0<s≤t≤T,
E|∫s0V(Bx,z0,t(r))dr−∫s0V(Bx,z0,s(r))dr|2≤CTα/2+βs2−β−α/2t−β|t−s|β(|x−z|2β+1). | (2.11) |
Proof. By Bx,z0,t(r)=B0,t(r)+ax,zr,t, Lemma 2.5, and the inequality |a+b|n≤2n−1(|a|n+|b|n), we have
I:=E|∫s0V(Bx,z0,t(r))dr−∫s0V(Bx,z0,s(r))dr|2=E|∫s0V(B0,s(r)+rGs,t+ax,zr,t)dr−∫s0V(B0,s(r)+ax,zr,s)dr|2≤∫RdE|∫s/20V(B0,s(r)+ry+ax,zr,t)dr−∫s/20V(B0,s(r)+ax,zr,s)dr|2pt−sst(y)dy+∫RdE|∫ss/2V(B0,s(r)+ry+ax,zr,t)dr−∫ss/2V(B0,s(r)+ax,zr,s)dr|2pt−sst(y)dy≤∫RdE|∫s/20V(B0,s(r)+ry+ax,zr,t)dr−∫s/20V(B0,s(r)+ax,zr,s)dr|2pt−sst(y)dy+∫RdE|∫s/20V(B0,s(r)+(s−r)y+ax,zs−r,t)dr−∫s/20V(B0,s(r)+ax,zs−r,s)dr|2pt−sst(y)dy≤2d/2∫RdE|∫s/20V(B(r)+ry+ax,zr,t)dr−∫s/20V(B(r)+ax,zr,s)dr|2pt−sst(y)dy+2d/2∫RdE|∫s/20V(B(r)+(s−r)y+ax,zs−r,t)dr−∫s/20V(B(r)+ax,zs−r,s)dr|2pt−sst(y)dy:=I1+I2, | (2.12) |
where the second to last inequality is due to the integral substitution and {B0,t(s)}s∈[0,t]d={B0,t(t−s)}s∈[0,t], and the last inequality is due to Lemma 2.6.
For I1, using the symmetry of k(x,y) and the integral substitution, it is obtained that
I1=2d/2∫Rd∫s/20∫s/20E[k(Br1+r1y+ax,zr1,t,Br2+r2y+ax,zr2,t)−k(Br1+r1y+ax,zr1,t,Br2+ax,zr2,s)−k(Br1+ax,zr1,s,Br2+r2y+ax,zr2,t)+k(Br1+ax,zr1,s,Br2+ax,zr2,s)]dr1dr2pt−sst(y)dy=2d/2+1∫Rd∫s/20∫r10ηr1,r2,ys,t,x,zdr1dr2pt−sst(y)dy, | (2.13) |
where we set
ηr1,r2,ys,t,x,z:=E[k(Br1+r1y+ax,zr1,t,Br2+r2y+ax,zr2,t)−k(Br1+r1y+ax,zr1,t,Br2+ax,zr2,s)−k(Br1+ax,zr1,s,Br2+r2y+ax,zr2,t)+k(Br1+ax,zr1,s,Br2+ax,zr2,s)]. | (2.14) |
By r2≤r1 and the independence of Brownian increments and the integral substitutions, we obtain
ηr1,r2,ys,t,x,z=E[k(Br1−Br2+Br2+r1y+ax,zr1,t,Br2+r2y+ax,zr2,t)−k(Br1−Br2+Br2+r1y+ax,zr1,t,Br2+ax,zr2,s)−k(Br1−Br2+Br2+ax,zr1,s,Br2+r2y+ax,zr2,t)+k(Br1−Br2+Br2+ax,zr1,s,Br2+ax,zr2,s)]=∫∫R2dk(¯x+¯y,¯y)[pr1−r2(¯x+(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+r2y+ax,zr2,t−ax,zr1,s)]⋅[pr2(¯y−r2y−ax,zr2,t)−pr2(¯y−ax,zr2,s)]d¯xd¯y+∫∫R2dk(¯x+¯y,¯y)[pr1−r2(¯x+(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+r2y+ax,zr2,t−ax,zr1,s)−pr1−r2(¯x−r1y+ax,zr2,s−ax,zr1,t)+pr1−r2(¯x+ax,zr2,s−ax,zr1,s)]pr2(¯y−ax,zr2,s)d¯xd¯y. | (2.15) |
We write bs,t:=(t−sst)1/2. To substitute (2.15) into (2.13), and by the absolute-value inequality and the integral substitutions about y, we get
I1≤2d/2+1∫Rd∫s/20∫r10∫∫R2d|k(¯x+¯y,¯y)||pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)||pr2(¯y−bs,tr2y−ax,zr2,t)−pr2(¯y−ax,zr2,s)|d¯xd¯ydr1dr2p1(y)dy+2d/2+1∫Rd∫s/20∫r10∫∫R2d|k(¯x+¯y,¯y)||pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)−pr1−r2(¯x−bs,tr1y+ax,zr2,s−ax,zr1,t)+pr1−r2(¯x+ax,zr2,s−ax,zr1,s)|⋅pr2(¯y−ax,zr2,s)d¯xd¯ydr1dr2p1(y)dy=:I11+I12. | (2.16) |
Notice that β∈(0,1). Thanks to (2.3) and (1.3), it holds that
˜J11:=∫∫R2d|k(¯x+¯y,¯y)||pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)||pr2(¯y−bs,tr2y−ax,zr2,t)−pr2(¯y−ax,zr2,s)|d¯xd¯y≲∫∫R2d(γh(¯x)+1)|pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)|1−β⋅|pr2(¯y−bs,tr2y−ax,zr2,t)−pr2(¯y−ax,zr2,s)|1−βd¯xd¯y(r1−r2)−β(d+1)/2r−β(d+1)/22⋅|−bs,tr1y−ax,zr1,t+ax,zr1,s|β|−bs,tr2y−ax,zr2,t+ax,zr2,s|β. | (2.17) |
On the one hand, by bs,t=(t−sst)1/2 and ax,zr,t−ax,zr,s=(t−s)st(x−z)r, it is found that
|−bs,tr1y−ax,zr1,t+ax,zr1,s|β|−bs,tr2y−ax,zr2,t+ax,zr2,s|β=(t−sst)βrβ1rβ2|y+bs,t(x−z)|2β. | (2.18) |
On the other hand, notice the fact that pβt(x)=(2π)(1−β)d/2β−d/2t(1−β)d/2pt/β(x). Then, by the inequality |a+b|β≤|a|β+|b|β(β∈[0,1]) and Lemma 2.2, we have
∫∫R2d(γh(¯x)+1)|pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)|1−β⋅|pr2(¯y−bs,tr2y−ax,zr2,t)−pr2(¯y−ax,zr2,s)|1−βd¯xd¯y≤∫Rd(γh(¯x)+1)(p1−βr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)+p1−βr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s))d¯x⋅∫Rd(p1−βr2(¯y−bs,tr2y−ax,zr2,t)+p1−βr2(¯y−ax,zr2,s))d¯y≤C(r1−r2)βd/2((r1−r2)−α/2+1)rβd/22∫Rd(pr2/(1−β)(¯y−bs,tr2y−ax,zr2,t)+pr2/(1−β)(¯y−ax,zr2,s))d¯y≤C(r1−r2)βd/2((r1−r2)−α/2+1)rβd/22, | (2.19) |
where the last step is due to the integral substitutions about ¯y and ‖pt‖L1(Rd)=1.
To substitute (2.18) and (2.19) into (2.17), we get
˜J11≤C(t−sst)β(r1−r2)−β/2((r1−r2)−α/2+1)rβ1rβ/22|y+bs,t(x−z)|2β. | (2.20) |
In addition, by the inequality |a+b|2β≤22β−1∨1(|a|2β+|b|2β) (β∈(0,1)),
∫Rd|y+bs,t(x−z)|2βp1(y)dy≤C∫Rd(|y|2β+(t−sst)β|x−z|2β)p1(y)dy≤C((t−sst)β|x−z|2β+1). | (2.21) |
Noticing that −α/2−β/2>−1 (i.e., β<1−α/2<2−α), and by (2.20), (2.21), and the Fubini theorem,
I11≤C(t−sst)β∫s/20∫r10(r1−r2)−β/2((r1−r2)−α/2+1)rβ1rβ/22dr1dr2∫Rd|y+bs,t(x−z)|2βp1(y)dy≤Cs2−α/2(sα/2+1)t−β(t−s)β∫Rd|y+bs,t(x−z)|2βp1(y)dy≤Cs2−α/2(sα/2+1)t−β(t−s)β((t−sst)β|x−z|2β+1). | (2.22) |
For β∈(0,1), by (1.3) and (2.4),
˜J12:=∫∫R2d|k(¯x+¯y,¯y)||pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)−pr1−r2(¯x−bs,tr1y+ax,zr2,s−ax,zr1,t)+pr1−r2(¯x+ax,zr2,s−ax,zr1,s)|pr2(¯y−ax,zr2,s)d¯xd¯y≲∫∫R2d(|γh(¯x)|+1)|pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)−pr1−r2(¯x−bs,tr1y+ax,zr2,s−ax,zr1,t)+pr1−r2(¯x+ax,zr2,s−ax,zr1,s)|1−βpr2(¯y−ax,zr2,s)d¯xd¯y(r1−r2)−βd/2−β|bs,tr1y+ax,zr1,t−ax,zr1,s|β|bs,tr2y+ax,zr2,t−ax,zr2,s|β. | (2.23) |
Using the inequality |a+b|β≤|a|β+|b|β(β∈[0,1]) and Lemma 2.2,
∫Rd(|γh(¯x)|+1)|pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)−pr1−r2(¯x−bs,tr1y+ax,zr2,s−ax,zr1,t)+pr1−r2(¯x+ax,zr2,s−ax,zr1,s)|1−βd¯x≤∫Rd(|γh(¯x)|+1)(p1−βr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)+p1−βr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)+p1−βr1−r2(¯x−bs,tr1y+ax,zr2,s−ax,zr1,t)+p1−βr1−r2(¯x+ax,zr2,s−ax,zr1,s))d¯x≤C(r1−r2)βd/2((r1−r2)−α/2+1). | (2.24) |
Using the Fubini theorem for (2.23), and substituting (2.24) and (2.18) into (2.23),
˜J12≲∫Rd(|γh(¯x)|+1)|pr1−r2(¯x+bs,t(r2−r1)y+ax,zr2,t−ax,zr1,t)−pr1−r2(¯x+bs,tr2y+ax,zr2,t−ax,zr1,s)−pr1−r2(¯x−bs,tr1y+ax,zr2,s−ax,zr1,t)+pr1−r2(¯x+ax,zr2,s−ax,zr1,s)|1−βd¯x∫Rdpr2(¯y−ax,zr2,s)d¯y(r1−r2)−βd/2−β|bs,tr1y+ax,zr1,t−ax,zr1,s|β|bs,tr2y+ax,zr2,t−ax,zr2,s|β≤C(t−sst)β(r1−r2)−β((r1−r2)−α/2+1)rβ1rβ2|y+bs,t(x−z)|2β. | (2.25) |
Recalling that −α/2−β>−1, and by (2.25), (2.21), and the similar computations to (2.22),
I12≤C(t−sst)β∫s/20∫r10(r1−r2)−β((r1−r2)−α/2+1)rβ1rβ2dr1dr2∫Rd|y+bs,t(x−z)|2βp1(y)dy≤Cs2−α/2(sα/2+1)t−β(t−s)β((t−sst)β|x−z|2β+1). | (2.26) |
To substitute (2.22) and (2.26) into (2.16),
I1≤Cs2−α/2(sα/2+1)t−β(t−s)β((t−sst)β|x−z|2β+1). | (2.27) |
Notice that I2 is similar to I1. By ax,zs−r,t−ax,zs−r,s=t−sst(x−z)(r−s) and the similar computations to (2.27), we obtain
I2≤Cs2−α/2(sα/2+1)t−β(t−s)β((t−sst)β|x−z|2β+1). | (2.28) |
At last, substituting (2.27) and (2.28) into (2.12), and by the relations that 2−β−α/2>0 (because of β∈(0,1−α/2) and α∈(0,2∧d)), T>1, and s≤t≤T,
I≤Cs2−α/2(sα/2+1)t−β(t−s)β((t−sst)β|x−z|2β+1)≤CTα/2+βs2−β−α/2t−β(t−s)β(|x−z|2β+1). | (2.29) |
So, we complete the proof.
The proof of Theorem 1.1. Without loss of generality, we assume that t≥s. Firstly, by (2.1), we have
uθ(t,x)−uθ(s,x)=∫RdEB[exp{ˆVx,z(t)}−exp{ˆVx,z(s)}]pt(z−x)u0(dz)+∫RdEBexp{ˆVx,z(s)}[pt(z−x)−ps(z−x)]u0(dz). | (2.30) |
Then, by the inequality |a+b|n≤2n−1(|a|n+|b|n), we obtain
E|uθ(t,x)−uθ(s,x)|n≤2n−1E|∫RdEB[exp{ˆVx,z(t)}−exp{ˆVx,z(s)}]pt(z−x)u0(dz)|n+2n−1E|∫RdEBexp{ˆVx,z(s)}[pt(z−x)−ps(z−x)]u0(dz)|n:=I1+I2. | (2.31) |
In I1, by the elementary inequality (a+b)2≤2(a2+b2), we find that for x,z∈Rd,
E|ˆVx,z(t)−ˆVx,z(s)|2=E|∫t0V(Bx,z0,t(r))dr−∫s0V(Bx,z0,s(r))dr|2≤2E|∫tsV(Bx,z0,t(r))dr|2+2E|∫s0V(Bx,z0,t(r))dr−∫s0V(Bx,z0,s(r))dr|2. | (2.32) |
Thanks to β<1−α/2 and α>0, it holds that 2−α/2−2β>α/2>0. To combine (2.32) with Propositions 5.2 and 5.3, and by the relations that T>1 and s≤t≤T,
E|ˆVx,z(t)−ˆVx,z(s)|2≤CTα/2|t−s|2−α/2+CTα/2+βs2−β−α/2t−β|t−s|β(|x−z|2β+1)≤CT2−β|t−s|β(|x−z|2β+1). | (2.33) |
In addition, by the inequality |a|2β≤e|a|2(β∈(0,1)), we find that
∫Rd(|x−z|2β+1)pt(z−x)|u0|(dz)≲tβpt/(1−β)∗|u0|(x)+pt∗|u0|(x). | (2.34) |
Hence, by (2.33), (2.34), and T>1, we find that
∫RdE|ˆVx,z(t)−ˆVx,z(s)|2pt(z−x)|u0|(dz)≤CT2−β|t−s|β∫Rd(|x−z|2β+1)pt(z−x)|u0|(dz)≤CT2supr∈[δ,T/(1−β)]pr∗|u0|(x)|t−s|β, | (2.35) |
where the last step is due to δ≤s≤t, too.
Using Proposition 5.1 and (2.35), we obtain
I1≤Cnθn((2n−1)!!)1/2Tn{(Eˉun2θ(t,x))1/2+(Eˉun2θ(s,t,x))1/2}⋅(supr∈[δ,T/(1−β)]pr∗|u0|(x))n/2|t−s|βn/2. | (2.36) |
Second, from (2.5), we find that for β∈(0,1),
I2≤CnE[∫RdEBexp{ˆVx,z(s)}|pt(z−x)−ps(z−x)||u0|(dz)]n≤Cn(t−d/2−1+s−d/2−1)βn|t−s|βn⋅E[∫RdEBexp{ˆVx,z(s)}|pt(z−x)−ps(z−x)|1−β|u0|(dz)]n. | (2.37) |
Then, by the inequalities that |a+b|β≤|a|β+|b|β(β∈[0,1]) and |a+b|n≤2n−1(|a|n+|b|n), and p1−βt(x)=(2π)βd/2(1−β)−d/2tβd/2pt/(1−β)(x),
E[∫RdEBexp{ˆVx,z(s)}|pt(z−x)−ps(z−x)|1−β|u0|(dz)]n≤E[∫RdEBexp{ˆVx,z(s)}(p1−βt(z−x)+p1−βs(z−x))|u0|(dz)]n≤Cn(tβd/2+sβd/2)nE[∫RdEBexp{ˆVx,z(s)}(pt1−β(z−x)+ps1−β(z−x))|u0|(dz)]n≤Cn(tβd/2+sβd/2)n[Eˉunθ(s,t/(1−β),x)+Eˉunθ(s,s/(1−β),x)], | (2.38) |
where we recall that ˉuθ(t,s,x) is defined in (2.7).
To substitute (2.38) into (2.37), and by the relation δ≤s≤t≤T,
I2≤Cn(tβd/2+sβd/2)n(t−d/2−1+s−d/2−1)βn[Eˉunθ(s,t/(1−β),x)+Eˉunθ(s,s/(1−β),x)]|t−s|βn≤CnTβdn/2δ−(d/2+1)βn[Eˉunθ(s,t/(1−β),x)+Eˉunθ(s,s/(1−β),x)]|t−s|βn. | (2.39) |
To combine (2.31) with (2.36) and (2.39),
E|uθ(t,x)−uθ(s,x)|n≤Cnθn((2n−1)!!)1/2Tn{(Eˉun2θ(t,x))1/2+(Eˉun2θ(s,t,x))1/2}⋅(supr∈[δ,T/(1−β)]pr∗|u0|(x))n/2|t−s|βn/2+CnTβdn/2δ−(d/2+1)βn[Eˉunθ(s,t/(1−β),x)+Eˉunθ(s,s/(1−β),x)]|t−s|βn. |
Moreover, by Proposition 4.1 and the relations that δ<1≤T, β<1 and θ≤eθ2, we can obtain that for all n∈N+,
E|uθ(t,x)−uθ(s,x)|n≤CnθneCθ2n2t2((2n−1)!!)1/2Tnexp{Cθ42−αt4−α2−αn4−α2−α}⋅(supr∈[δ,T/(1−β)]pr∗|u0|(x))n|t−s|βn/2+CneCθ2n2t2Tβdn/2δ−(d/2+1)βnexp{Cθ42−αt4−α2−αn4−α2−α}⋅(supr∈[δ,T/(1−β)]pr∗|u0|(x))n|t−s|βn≤CneCθ2n2T2exp{Cθ42−αT4−α2−αn4−α2−α}((2n−1)!!)1/2T(βd/2+1)n⋅δ−(d/2+1)βn(supr∈[δ,T/(1−β)]pr∗|u0|(x))n|t−s|βn/2. | (2.40) |
At last, by (1.6), (2.40), and the classic Kolmogorov continuity theorem, we find that for all β∈(0,1−α/2), there exists a temporal β2-Hölder continuous modification of uθ(t,x) on [δ,T]. Because δ and T are any, the proof can be completed.
The proof of Theorem 1.2. Assume that T>1 and 0≤s≤t≤T. Let n be a positive integer.
(ⅰ) Through (2.4) and Lemma 2.4, it can be proved that
uθ(t,x)−uθ(s,x)=EB[(exp{θ∫t0V(Bxr)dr}−exp{θ∫s0V(Bxr)dr})u0(Bxt)]+EB[exp{θ∫s0V(Bxr)dr}u0(Bxt)]−EB[exp{θ∫s0V(Bxr)dr}u0(Bxs)]=∫RdEB[exp{θ∫t0V(Bx,z0,t(r))dr}−exp{θ∫s0V(Bx,z0,t(r))dr}]pt(z−x)u0(dz)+EB[exp{θ∫s0V(Bxr)dr}u0(Bxt)]−EB[exp{θ∫s0V(Bxr)dr}u0(Bxs)]. | (2.41) |
Next, by the similar computations to (2.31), we obtain
E|uθ(t,x)−uθ(s,x)|n≤2n−1E|∫RdEB[exp{θ∫t0V(Bx,z0,t(r))dr}−exp{θ∫s0V(Bx,z0,t(r))dr}]pt(z−x)u0(dz)|n+2n−1E|EB[exp{θ∫s0V(Bxr)dr}u0(Bxt)]−EB[exp{θ∫s0V(Bxr)dr}u0(Bxs)]|n:=D1+D2. | (2.42) |
For D1, using the method of proof similar to Proposition 5.1, it not difficult to check that
D1≤2n−1θn((2n−1)!!)1/2{(Eˉun2θ(t,x))1/2+(Eˉun2θ(s,t,x))1/2}⋅(∫RdE|∫tsV(Bx,z0,t(r))dr|2pt(z−x)|u0|(dz))n/2. | (2.43) |
To associate the above (2.43) with Propositions 5.2 and 4.1,
D1≤Cnθn((2n−1)!!)1/2Tαn/4{(Eˉun2θ(t,x))1/2+(Eˉun2θ(s,t,x))1/2}|t−s|(1−α/4)n≤CneCθ2n2t2((2n−1)!!)1/2Tαn/4exp{Cθ42−αt4−α2−αn4−α2−α}(pt∗|u0|(x))n|t−s|(1−α/4)n. | (2.44) |
For D2, from the independence of Brownian increments and κ-Hölder continuity of u0, it is found that
D2=2n−1E|EB[exp{θ∫s0V(Bxr)dr}u0(Bt−Bs+Bxs)]−EB[exp{θ∫s0V(Bxr)dr}u0(Bxs)]|n≤2n−1E|∫RdEB[exp{θ∫s0V(Bxr)dr}|u0((t−s)1/2y+Bxs)−u0(Bxs)|]p1(y)dy|n≤CnE[EBexp{θ∫s0V(Bxr)dr}]n(∫Rd|y|κp1(y)dy)n(t−s)κn/2≤CneCθ2n2t2exp{Cθ42−αt4−α2−αn4−α2−α}(t−s)κn/2, | (2.45) |
where the last step is due to Proposition 4.1.
Notice that 0≤s≤t≤T. To combine (2.42) with (2.44) and (2.45), it is found that for all x∈Rd and integer n≥1,
E|uθ(t,x)−uθ(s,x)|n≤CneCθ2n2T2exp{Cθ42−αT4−α2−αn4−α2−α}((2n−1)!!)1/2T(1−κ/2)n⋅(supr∈[0,T]pr∗|u0|(x))n|t−s|κn/2, | (2.46) |
where we have used the fact that κ/2<1−α/4 for κ∈(0,1] and α∈(0,2∧d).
So, by (2.46) and the Kolmogorov continuity theorem, we can prove the result.
(ⅱ) By u0≡C and the method similar to (2.41), it is obtained that
uθ(t,x)−uθ(s,x)=Cn∫RdEB[exp{θ∫t0V(Bx,z0,t(r))dr}−exp{θ∫s0V(Bx,z0,t(r))dr}]pt(z−x)dz. | (2.47) |
Moreover, using the computations similar to (2.44) and 0≤s≤t≤T, we find that for all x∈Rd and integer n≥1,
E|uθ(t,x)−uθ(s,x)|n≤CneCθ2n2T2((2n−1)!!)1/2Tαn/4exp{Cθ42−αT4−α2−αn4−α2−α}|t−s|(1−α/4)n. | (2.48) |
Lastly, through (2.48) and the Kolmogorov continuity theorem, we can complete the proof.
This work mainly studies the temporal Hölder continuity for the Feynman-Kac formula of the parabolic Anderson model under the rough initial condition pt∗|u0|(x)<∞. As a comparison, we also consider the function-valued initial conditions u0≡C and u0∈Cκ(Rd) with κ∈(0,1]. However, many function-valued initial data have not been considered in this paper, which will be a future work. Besides, our future work is also going to investigate the case of time-space generalized Gaussian field and rough initial condition.
Hui Sun: Dealt with conceptualization, supervision, formal analysis, writing-original draft, review, edition; Yangyang Lyu: Investigation, methodology, writing-original draft, edition. All authors have read and approved the final version of the manuscript for publication.
The author(s) declare(s) that they have not used Artificial Intelligence (AI) tools in the creation of this article.
The author(s) would like to thank the anonymous reviewer(s) and referee(s) for their patient reviews and earnest suggestions. H. Sun was supported by the Education Department of Fujian Province (No. JAT210257) and High-level cultivation program in Minnan Normal University (No. MSGJB2022009). Y. Lyu was supported by NSFC (No. 12201282) and Natural Science Foundation of Fujian Province, China (No. 2023J05176).
The authors declare that they have no competing interests.
[1] |
L. D. Pitt, R. Robeva, On the sharp Markov property for Gaussian random fields and spectral synthesis in spaces of Bessel potentials, Ann. Probab., 31 (2003), 1338–1376. http://dx.doi.org/10.1214/aop/1055425783 doi: 10.1214/aop/1055425783
![]() |
[2] |
X. Chen, Spatial asymptotics for the parabolic Anderson models with generalized time-space Gaussian noise, Ann. Probab., 44 (2016), 1535–1598. http://dx.doi.org/10.1214/15-AOP1006 doi: 10.1214/15-AOP1006
![]() |
[3] |
Y. Hu, D. Nualart, J. Song, Feynman-Kac formula for heat equation driven by fractional white noise, Ann. Probab., 39 (2011), 291–326. http://dx.doi.org/10.1214/10-AOP547 doi: 10.1214/10-AOP547
![]() |
[4] |
B. Duplantier, R. Rhodes, S. Sheffield, V. Vargas, Renormalization of critical Gaussian multiplicative chaos and KPZ relation, Commun. Math. Phys., 330 (2014), 283–330. http://dx.doi.org/10.1007/s00220-014-2000-6 doi: 10.1007/s00220-014-2000-6
![]() |
[5] |
T. Madaule, Maximum of a log-correlated Gaussian field, Ann. Inst. H. Poincaré Probab. Statist., 51 (2015), 1369–1431. http://dx.doi.org/10.1214/14-AIHP633 doi: 10.1214/14-AIHP633
![]() |
[6] |
L. Bertini, G. Giacomin, Stochastic Burgers and KPZ equations from particle systems, Commun. Math. Phys., 183 (1997), 571–607. https://doi.org/10.1007/s002200050044 doi: 10.1007/s002200050044
![]() |
[7] |
G. Amir, I. Corwin, J. Quastel, Probability distribution of the free energy of the continuum directed random polymer in 1+1 dimensions, Comm. Pure Appl. Math., 64 (2011), 466–537. https://doi.org/10.1002/cpa.20347 doi: 10.1002/cpa.20347
![]() |
[8] |
L. Chen, R. C. Dalang, Moments and growth indices for the nonlinear stochastic heat equation with rough initial conditions, Ann. Probab., 43 (2015), 3006–3051. https://doi.org/10.1214/14-AOP954 doi: 10.1214/14-AOP954
![]() |
[9] |
R. M. Balan, L. Quer-Sardanyons, J. Song, Hölder continuity for the parabolic Anderson model with space-time homogeneous Gaussian noise, Acta Math. Sci., 39 (2019), 717–730. https://doi.org/10.1007/s10473-019-0306-3 doi: 10.1007/s10473-019-0306-3
![]() |
[10] |
R. Balan, L. Chen, Parabolic Anderson model with space-time homogeneous Gaussian noise and rough initial condition, J. Theor. Probab., 31 (2018), 2216–2265. http://dx.doi.org/10.1007/s10959-017-0772-2 doi: 10.1007/s10959-017-0772-2
![]() |
[11] |
R. Balan, L. Chen, Y. Ma, Parabolic Anderson model with rough noise in space and rough initial conditions, Electron. Commun. Probab., 27 (2022), 1–12. http://dx.doi.org/10.1214/22-ECP506 doi: 10.1214/22-ECP506
![]() |
[12] |
L. Chen, R. Dalang, Hölder-continuity for the nonlinear stochastic heat equation with rough initial conditions, Stoch. PDE: Anal. Comp., 2 (2014), 316–352. http://dx.doi.org/10.1007/s40072-014-0034-6 doi: 10.1007/s40072-014-0034-6
![]() |
[13] |
L. Chen, K. Kim, Nonlinear stochastic heat equation driven by spatially colored noise: moments and intermittency, Acta Math. Sci., 39 (2019), 645–668. https://doi.org/10.1007/s10473-019-0303-6 doi: 10.1007/s10473-019-0303-6
![]() |
[14] |
L. Chen, R. C. Dalang, Moments, intermittency and growth indices for the nonlinear fractional stochastic heat equation, Stoch. PDE: Anal. Comp., 3 (2015), 360–397. https://doi.org/10.1007/s40072-015-0054-x doi: 10.1007/s40072-015-0054-x
![]() |
[15] |
L. Chen, J. Huang, Comparison principle for stochastic heat equation on Rd, Ann. Probab., 47 (2019), 989–1035. https://doi.org/10.1214/18-AOP1277 doi: 10.1214/18-AOP1277
![]() |
[16] |
Y. Hu, J. Huang, D. Nualart, S. Tindel, Stochastic heat equations with general multiplicative Gaussian noises: Hölder continuity and intermittency, Electron. J. Probab., 20 (2015), 1–50. http://dx.doi.org/10.1214/EJP.v20-3316 doi: 10.1214/EJP.v20-3316
![]() |
[17] |
Y. Lyu, Spatial asymptotics for the Feynman-Kac formulas driven by time-dependent and space-fractional rough Gaussian fields with the measure-valued initial data, Stochastic Process. Appl., 143 (2022), 106–159. http://dx.doi.org/10.1016/j.spa.2021.10.003 doi: 10.1016/j.spa.2021.10.003
![]() |
[18] |
Y. Lyu, H. Li, Almost surely time-space intermittency for the parabolic Anderson model with a log-correlated Gaussian field, Acta Math. Sci., 43 (2023), 608–639. http://dx.doi.org/10.1007/s10473-023-0209-1 doi: 10.1007/s10473-023-0209-1
![]() |
[19] |
M. Gubinelli, N. Perkowski, KPZ reloaded, Commun. Math. Phys., 349 (2017), 165–269. http://dx.doi.org/10.1007/s00220-016-2788-3 doi: 10.1007/s00220-016-2788-3
![]() |
[20] | N. Perkowski, SPDEs, classical and new, Freie Universität Berlin, 2020. |
[21] |
L. Chen, Y. Hu, D. Nualart, Two-point correlation function and Feynman-Kac formula for the stochastic heat equation, Potential Anal., 46 (2017), 779–797. https://doi.org/10.1007/s11118-016-9601-y doi: 10.1007/s11118-016-9601-y
![]() |
[22] |
J. Huang, K. Lê, D. Nualart, Large time asymptotics for the parabolic Anderson model driven by spatially correlated noise, Ann. Inst. Henri. Poincar. Probab. Stat., 53 (2017), 1305–1340. http://dx.doi.org/10.1214/16-AIHP756 doi: 10.1214/16-AIHP756
![]() |
[23] | M. I. Gelfand, N. Ya. Vilenkin, Applications of harmonic analysis, Academic Press, 1964. https://doi.org/10.1016/C2013-0-12221-0 |