Processing math: 100%
Research article

Fractional stochastic heat equation with mixed operator and driven by fractional-type noise

  • Received: 27 May 2024 Revised: 04 October 2024 Accepted: 08 October 2024 Published: 14 October 2024
  • MSC : 35R11, 60G22, 60H15

  • We investigated a novel stochastic fractional partial differential equation (FPDE) characterized by a mixed operator that integrated the standard Laplacian, the fractional Laplacian, and the gradient operator. The equation was driven by a random noise, which admitted a covariance measure structure with respect to the time variable and behaved as a Wiener process in space. Our analysis included establishing the existence of a solution in the general case and deriving an explicit form for its covariance function. Additionally, we delved into a specific case where the noise was modeled as a generalized fractional Brownian motion (gfBm) in time, with a particular emphasis on examining the regularity of the solution's sample paths.

    Citation: Mounir Zili, Eya Zougar, Mohamed Rhaima. Fractional stochastic heat equation with mixed operator and driven by fractional-type noise[J]. AIMS Mathematics, 2024, 9(10): 28970-29000. doi: 10.3934/math.20241406

    Related Papers:

    [1] Xueqi Wen, Zhi Li . $ p $th moment exponential stability and convergence analysis of semilinear stochastic evolution equations driven by Riemann-Liouville fractional Brownian motion. AIMS Mathematics, 2022, 7(8): 14652-14671. doi: 10.3934/math.2022806
    [2] Kinda Abuasbeh, Ramsha Shafqat, Azmat Ullah Khan Niazi, Muath Awadalla . Nonlocal fuzzy fractional stochastic evolution equations with fractional Brownian motion of order (1,2). AIMS Mathematics, 2022, 7(10): 19344-19358. doi: 10.3934/math.20221062
    [3] Weiguo Liu, Yan Jiang, Zhi Li . Rate of convergence of Euler approximation of time-dependent mixed SDEs driven by Brownian motions and fractional Brownian motions. AIMS Mathematics, 2020, 5(3): 2163-2195. doi: 10.3934/math.2020144
    [4] 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
    [5] Fawaz K. Alalhareth, Seham M. Al-Mekhlafi, Ahmed Boudaoui, Noura Laksaci, Mohammed H. Alharbi . Numerical treatment for a novel crossover mathematical model of the COVID-19 epidemic. AIMS Mathematics, 2024, 9(3): 5376-5393. doi: 10.3934/math.2024259
    [6] Rajesh Dhayal, Muslim Malik, Syed Abbas . Solvability and optimal controls of non-instantaneous impulsive stochastic neutral integro-differential equation driven by fractional Brownian motion. AIMS Mathematics, 2019, 4(3): 663-683. doi: 10.3934/math.2019.3.663
    [7] Noorah Mshary, Hamdy M. Ahmed . Discussion on exact null boundary controllability of nonlinear fractional stochastic evolution equations in Hilbert spaces. AIMS Mathematics, 2025, 10(3): 5552-5567. doi: 10.3934/math.2025256
    [8] 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
    [9] Obaid Algahtani, M. A. Abdelkawy, António M. Lopes . A pseudo-spectral scheme for variable order fractional stochastic Volterra integro-differential equations. AIMS Mathematics, 2022, 7(8): 15453-15470. doi: 10.3934/math.2022846
    [10] Xinyi Wang, Jingshen Wang, Zhidong Guo . Pricing equity warrants under the sub-mixed fractional Brownian motion regime with stochastic interest rate. AIMS Mathematics, 2022, 7(9): 16612-16631. doi: 10.3934/math.2022910
  • We investigated a novel stochastic fractional partial differential equation (FPDE) characterized by a mixed operator that integrated the standard Laplacian, the fractional Laplacian, and the gradient operator. The equation was driven by a random noise, which admitted a covariance measure structure with respect to the time variable and behaved as a Wiener process in space. Our analysis included establishing the existence of a solution in the general case and deriving an explicit form for its covariance function. Additionally, we delved into a specific case where the noise was modeled as a generalized fractional Brownian motion (gfBm) in time, with a particular emphasis on examining the regularity of the solution's sample paths.



    In recent decades, fractional partial differential equations (FPDEs) have garnered significant attention across various fields such as mathematics [5,34], physics [14], engineering [20], chemistry [26], fluid mechanics [1], nuclear reactor dynamics [15], chaotic dynamical systems [19], mechanics of materials [8], biology [18], hydrology [7], finance [21], and social sciences [9]. In this paper, we focus on the following specific FPDE for a fixed d1:

    u(t,x)t=La,bu(t,x), (1.1)

    with (t,x)[0,T]×Rd,T>0. Here, La,b is the mixed fractional operator given as:

    La,b=Δ+aαΔα/2+b., (1.2)

    where α(1,2], a(0,M], b=(b1bd)Rd, Δ=di=12x2i is the Laplacian on Rd, =(x1xd) is the gradient on Rd, b.=di=1bixi, and Δα/2 is the operator defined by:

    Δα/2u(x)=A(d,α)limδ0{zRd;zx>δ}u(z)u(x)zxd+αdz,uC2c(Rd), (1.3)

    with A(d,α)=α2α1πd/2Γ(d+α2)Γ(1α2). Here, Γ is denoting the Gamma function and C2c(Rd) is the space of twice continuously differentiable functions on Rd with compact support.

    Operator (1.2) was introduced in [5] in a more general case, where b is a function in a certain Kato class, and it can be seen as the infinitesimal generator of some diffusion processes related to anomalous diffusion (see [5,34] and references therein).

    In the present paper, we introduce a stochastic counterpart of Eq (1.1), defined by

    {u(t,x)t=La,bu(t,x)+˙W(t,x),t>0,xRd,u(0,x)=0, (1.4)

    where ˙W denotes the formal derivative of a centered Gaussian field W, which behaves as a Wiener process with respect to the space variable, and as a process that admits a covariance measure structure, in the sense of [13], with respect to the time variable. In particular, for fixed xRd, W(.,x) extends many interesting Gaussian fractional processes such as: mixed fractional Brownian motion (mfBm) (see e.g., [28]), mixed subfractional Brownian motion (msfBm) (see, e.g., [6,16]), generalized fractional Brownian motion (gfBm) [29,31,32], and so on. Investigation of Eq (1.4) represents a novel mathematical problem that has not been explored before. In [34], Zili and Zougar have investigated equation of the form (1.4) with a different type of random force, specifically space-time white noise {Wt}t0, that is, a centered Gaussian process with covariance function EWtWs=ts. In this paper, we deal with a distinct and more general stochastic problem that involves a different framework and additional complexity.

    Equation (1.4) illustrates heat propagation in inhomogeneous media, influenced by anomalous diffusion and subject to stochastic perturbations. Recent studies have uncovered atypical behaviors in diffusion processes within nonhomogeneous media. These anomalous diffusion phenomena are best described by fractional-order models, as classical integer-order models fail to capture their unique characteristics (see, for example, [12]). This can be considered as a main motivation for this work.

    Stochastic FPDEs of type (1.4) have been widely studied in the literature with specific operators that are special cases of the general operator considered in this work. For instance, the case where the operator is limited to the fractional Laplacian Δα/2 has been thoroughly investigated in works like [2,11], considering various types of additive Gaussian noises. In [25,27], the authors examined equations resembling (1.4), particularly when a=1 and b=0, resulting in the operator L1,0=Δ+Δα/2. They explored these equations under the influence of diverse additive drifts and various types of fractional noises.

    Therefore, in addition to the generalization of the stochastic FPDE introduced in [34], Eq (1.4) represents a further extension of the fractional models investigated in [10,11,25,27,34], and this can be regarded as another important motivation for the investigation of such an equation's solution. It's worth noting that in numerous other papers (e.g., [17,23,24,30,33], and the references therein), researchers have explored different types of fractional stochastic PDEs. In these cases, the term "fractional" typically pertains to the additive noise rather than to the fractional Laplacian operator, as is the case in our study.

    This research represents a pioneering study of the solution for the novel, unexplored FPDE (1.4). Our primary contribution lies in laying the foundational groundwork, providing a solid basis for future investigations in this area. We first provide a sufficient condition for the existence of the solution, and we give an explicit expression of its covariance function. Then, we focus our attention on the interesting specific case, where W(.,x) behaves as a gfBm, in the sense introduced by M. Zili in [29,31,32]. The gfBm is an extension of both fBm and sfBm, defined as a linear combination of two independent fBm and sfBm. So, it is about a process which depends on three parameters: the Hurst index and the coefficients of the linear combination. This should allow researchers to construct more adequate models, permitting, for example, to control the level of correlation between the increments of the studied phenomena, and, consequently, to overcome the deficiency of fBm and sfBm models due to their dependence on one single constant, which is the Hurst parameter. More information about the gfBm and the motivations of its introduction can be found in [31].

    The results of this paper are obtained by introducing the canonical Hilbert space associated to the Gaussian noise W by applying many integration techniques, calculation, and analysis tools, and especially by suitably exploiting the two-sided estimates of the fundamental solution Ga,b of the operator La,b, already established in the case where d1, by Zili and Zougar in [34], and moreover by the explicit expression of its Fourier transform.

    The paper is organized as follows: In next section, we give some examples of applications of the fractional PDE. In Section 3, we first introduce the random noise that drives our stochastic FPDE and, in particular, the processes admitting a covariance measure structure. We give some interesting examples, and we explain the mode of Wiener integration with respect to our noise that we will use. Then, we specify our meaning of solution and give some characteristics of the Green fundamental function Ga,b of Eq (1.1) that will play a main role in the whole of this paper. After that, we give a sufficient condition for the existence of the mild solution of the stochastic FPDE (1.4), and we present an explicit expression of its covariance function. In Section 4, we focus on the interesting particular case when the process is a white-space Gaussian field, behaving as a gfBm in time. We especially analyze the regularity of the sample paths of the solution with respect to the time variable. Finally, Section 5 offers a discussion of the results, while Section 6 presents the conclusion of the paper.

    As mentioned in the introduction, Eq (1.1) serves as a good model across various domains. It has become increasingly important in modeling complex systems where classical integer-order models fail to capture the full dynamics. By incorporating fractional derivatives, these equations provide a more comprehensive framework for handling anomalous diffusion, nonlocal processes, and long-range interactions. Below, we explore two examples of applications of FPDEs: one in fluid dynamics and the other in financial mathematics.

    FPDEs are vital in fluid dynamics, especially for systems involving turbulence, anomalous diffusion, and non-Newtonian fluids. While traditional PDEs using the classical Laplacian operator Δ effectively model normal diffusion, many real-world systems, such as turbulent or porous media, involve more complex transport processes. In turbulent flows, like those in atmospheric dynamics or ocean currents, particles exhibit super-diffusive behavior, where their mean squared displacement grows faster than under normal diffusion. This can be modeled using the fractional Laplacian, which captures nonlocal transport mechanisms like Lévy flights, where particles make large, unpredictable jumps. The drift term b. models advection, representing directional transport driven by external forces or velocity fields, such as wind or ocean currents. In oceanography, for example, this describes how water is carried by currents, while diffusion terms capture smaller-scale turbulence and mixing. FPDEs are also effective for modeling non-Newtonian fluids, like polymers or biological fluids, where the stress-strain relationship is nonlinear, and memory effects are important.

    Example 2.1. In the study of pollutant dispersion in turbulent ocean currents, the drift term accounts for the overall flow of water carrying the pollutant, while the standard Laplacian Δ represents local diffusion. The fractional Laplacian Δα/2 captures the nonlocal effects, such as sudden shifts in concentration caused by turbulence. Together, these terms provide a full description of how pollutants spread in complex fluid environments. For further details, readers are advised to consult [4].

    FPDEs are also extensively used in financial mathematics, particularly for modeling asset prices and option pricing in markets with jumps and volatility clustering. In these models, FPDEs capture the stochastic processes governing financial instruments. The classical Laplacian Δ corresponds to Brownian motion, which assumes continuous price changes, as seen in models like the Black-Scholes equation. However, real-world markets often exhibit large price jumps and heavy-tailed distributions, which the standard diffusion operator cannot account for. So, to address this, the fractional Laplacian is used to model jumps and heavy-tailed returns. This operator captures the behavior of Lévy processes, which describe sudden, unpredictable price changes, such as those caused by market shocks or economic news. It is particularly useful for pricing exotic options or derivatives sensitive to these large, infrequent movements. The drift term represents the expected return or trend of an asset, influenced by predictable market factors like economic growth or interest rates. Combined with the fractional Laplacian, this term models both the overall direction and the random, jump-like fluctuations of asset prices.

    Example 2.2. In option pricing under jump-diffusion models, the drift term accounts for the average return on an asset, while the fractional Laplacian captures the sudden jumps in asset prices caused by market shocks. This approach is particularly relevant for pricing credit derivatives or insurance contracts against market crashes, as it allows for a more accurate assessment of risk and pricing, considering both gradual changes in asset prices and the possibility of rare but significant market events. For more information, readers are encouraged to see [21].

    Let us describe the random noise driving the stochastic FPDE (1.4), subject to this study.

    To start, we define the processes with covariance measure structure.

    Consider a zero mean square integrable process (Xt)t[0,T] with covariance function RX(t,r)=E[XtXr], for (t,r)[0,T]2. The covariance RX defines naturally a finite additive measure μRX:=μ on the algebra R of finite disjoint rectangles included in the set [0,T]2 by:

    μ(J)=ΔJRX

    where ΔJRX denotes the rectangular increment of RX over the rectangle J=[a1,b1)×[a2,b2) given by:

    ΔJRX=RX(b1,b2)RX(a1,b2)RX(a2,b1)+RX(a1,a2).

    The process X is said to have a covariance measure structure if μ can be extended to a signed sigma finite measure on B([0,T]2). Some important characteristics of such processes can be found in [13]. In particular, we have:

    Lemma 3.1. Any zero mean square integrable process (Xt)t[0,T] with covariance function R such that

    2RXrtis integrable on  [0,T]2,

    has a covariance measure structure. Furthermore, the measure μ generated by RX admits a density with respect to the Lebesgue measure on [0,T]2 given by 2RXrt.

    Let us give a few examples of processes with covariance measure structure.

    Example 3.1. Let us denote by MH={MHt(θ,ν);t0}={MHt;t0} the mixed-fBm of parameters θ,ν, and H such that H(0,1), (θ,ν)R2{(0,0)}; that is, the centered Gaussian process, starting from zero, with covariance

    RH,θ,νM(t,r):=Cov(MHt(θ,ν),MHr(θ,ν))=θ2(tr)+ν22(t2H+r2H|tr|2H), (3.1)

    where tr=12(t+r|tr|). Some specific examples of this process include: MH(0,1)=BH, which represents an fBm and MH(1,0)=B, which corresponds to standard Bm. So, the mfBm is clearly an extension of the fBm and of the Wiener process. We refer to [28] for further information on this process.

    If (θ,ν,H)R×R×(12,1) or (θ,ν)R×{0}, MH(θ,ν) admits a covariance measure structure on [0,T]2 which has a density given by:

    2RH,θ,νMtr(t,r)=θ2δ0(tr)+ν2σH|tr|2H2,

    where σH=H(2H1), δ0 is the Dirac measure and R is the set R{0}.

    Example 3.2. Consider θ and ν two real constants such that (θ,ν)(0,0) and H(0,1). A gfBm of parameters θ,ν, and H is a process ZH={ZHt(θ,ν);t0}={ZHt;t0}, defined on the probability space (Ω,F,P) by:

    tR+ZHt=ZHt(θ,ν)=θBHt+νBHt (3.2)

    where (BHt)tR is a two-sided fBm of parameter H. We offer several examples of this process: ZH(1,0) represents an fBm, while ZH(12,12) denotes the sfBm. So, the gfBm is in the same time, a generalization of the fBm, of the sfBm, and of course of the standard Brownian motion. For more information, the reader can read [29,31,32].

    The gfBm (ZHt(θ,ν))tR+ is a centered Gaussian process with covariance function

    RH,θ,νZ(t,r)=Cov(ZHt(θ,ν),ZHr(θ,ν))=12(θ+ν)2(t2H+r2H)νθ(t+r)2Hθ2+ν22|tr|2H (3.3)

    for every t,r(0,+). Then, when H(12,1), ZH(θ,ν) admits a covariance measure structure on [0,T]2 which has a density given by:

    2RH,θ,νZtr(t,r):=σH[(θ2+ν2)|tr|2H22νθ(t+r)2H2]. (3.4)

    In this section, we introduce the random noise that drives the parabolic Eq (1.4). On a complete probability space (Ω,F,P), we consider a zero-mean Gaussian field W={W(t,A);t[0,T],ABb(Rd)} with covariance:

    Cov(W(t,A)W(r,B))=λd(AB)RW(t,r) (3.5)

    where λd is the Lebesgue measure, and RW is the covariance of a stochastic process that generates a covariance measure μ.

    To the Gaussian field W, we can associate a Hilbert space that will be called the canonical Hilbert space of W and will be denoted by H. Consider E the set of linear combinations of elementary functions 1[0,t]×A, (t,A)[0,T]×Bb(Rd), and let H be the Hilbert space defined as the closure of E with respect to the inner product

    <1[0,t]×A,1[0,r]×B>H:=Cov(W(t,A)W(s,B)).

    We have the following expression of the scalar product in H:

    <φ,ψ>H=T0T0μ(dr,dw)Rdφ(r,z)ψ(w,z)dz (3.6)

    for any φ,ψH such that

    T0T0|μ|(dr,dw)Rd|φ(r,z)||ψ(w,z)|dz<, (3.7)

    where |μ| denotes the total variation measure associated to μ.

    Following [24], by a routine extension of the construction done in [13], it is possible to define Wiener integrals with respect to the process W whose covariance is given by (3.5). This Wiener integral will act as an isometry between the Hilbert space H and L2(Ω) in the sense that:

    E[T0Rdφ(r,z)W(dr,dz)T0Rdψ(r,z)W(dr,dz)]=T0T0μ(dr,dw)Rdφ(r,z)ψ(w,z)dz. (3.8)

    In this part, we will analyze the existence of the solution to Eq (1.4) driven by a random noise characterized by (3.5). The notion of the solution to Eq (1.4) is defined in the mild sense. We call a mild solution to (1.4) the stochastic process

    ua,b(t,x)=T0RdGa,b(tr,x,z)1(0,t)(r)W(dr,dz),t0,xRd, (3.9)

    where W is the Gaussian noise with covariance given by (3.5), Ga,b denotes the fundamental solution for the operator La,b, and the integral in (3.9) is a Wiener integral with respect to the Gaussian noise W.

    Let us first recall some useful properties of Ga,b.

    In [5], the authors established existence and uniqueness of a fundamental solution Ga,b of the operator La,b, and they provide some characterizations and estimates, some of which we quote in the following lemma.

    Lemma 3.2. Let d1. There exist two positive constants c1,c2 such that, for all t>0,x,z,bRd, and a[0,M], we have

    c11pac2(t,x,z)Ga,b(t,x,z)c1pa1/c2(t,x,z). (3.10)

    with pac(t,x,z)=td/2exp(cxz2t)+td/2aαtxzd+α,c>0.

    The following useful characteristic of Ga,b was proved in [34].

    Lemma 3.3. For every t>0 and xRd, the Fourier transform of Ga,b(t,x,.), denoted by F(Ga,b(t,x,.)), is given by

    F(Ga,b(t,x,.))(ξ)=exp(i(xtb).ξ)exp(tAaα(ξ)) (3.11)

    for every ξRd, with Aaα(ξ)=ξ2+aαξα.

    In the rest of the paper, we will denote the following function: at,r=tr,(t,r)[0,T]2.

    It is well-known that the mild solution to (1.4) exists when the Wiener integral in (3.9) is well-defined, and this happens when the integrand Ga,b belongs to H=L2([0,T]×Rd). Let us now give a sufficient condition for the existence of the mild solution defined in (3.9).

    Theorem 1. We assume that t0t0(2trw)d/α|μ|(dr,dw) is finite, for every t[0,T], then the mild solution given in (3.9) is well-defined. Moreover, if supt[0,T]t0t0(2trw)d/α|μ|(dr,dw)<, then

    sup(t,x)[0,T]×RdE[|ua,b(t,x)|2]<.

    Proof. Consider t[0,T]. By the Wiener isometry characteristic (3.8), we have

    E[|ua,b(t,x)|2]=t0t0μ(dr,dw)RdGa,b(at,r,x,z)Ga,b(at,w,x,z)dz. (3.12)

    By applying the Plancherel theorem, we obtain

    E[|ua,b(t,x)|2](2π)dt0t0|μ|(dr,dw)Rd|FGa,b(at,r,x,.)(ξ)|¯|FGa,b(at,w,x,.)(ξ)|dξ=(2π)dt0t0|μ|(dr,dw)×Rd|exp(i(xat,rb).ξ)|eat,rAaα(ξ)|exp(i(xat,wb).ξ)|eat,wAaα(ξ)dξ=(2π)dt0t0|μ|(dr,dw)Rd|eib.ξ(wr)|e(2trw)Aaα(ξ)dξ.

    Since α/2(1/2,1), the function xxα/2 is concave on [0,+). Therefore, using the Jensen inequality, we get:

    Aaα(ξ)aαξα=aα(dj=1|ξj|2)α/2aαdα21dj=1|ξj|α (3.13)

    for every ξRd. Hence,

    E[|ua,b(t,x)|2](2π)dt0t0|μ|(dr,dw)(ReDα(2trw)|ξ1|αdξ1)d

    with Dα=aαdα/21. By the change variable z1=(2trw)1/αξ1, we get

    ReDα(2trw)|ξ1|αdξ1=(2trw)1/αReDα|z1|αdz1=γα(2trw)1/α,

    with

    γdα=(ReDα|z1|αdz1)d. (3.14)

    Consequently,

    E[|ua,b(t,x)|2](2π)dγdαt0t0(2trw)d/α|μ|(dr,dw)<+, (3.15)

    which achieves the proof of Theorem 1.

    Remark 3.1. In the particular case where the noise W in Eq (1.4) is defined by a Wiener process, that is, RW(t,s)=ts, W defines a covariance measure μ given by μ(du,dv)=δ0(uv)dudv, where δ0 is the Dirac measure. This case was analyzed by Zili and Zougar in [34]. They established that a mild solution exists precisely when the integral t0(tr)d/αdr is finite, which is equivalent to the condition d=1.

    In the following theorem, we give an explicit expression of the covariance function of the mild solution for Eq (1.4).

    Theorem 2. For every t,s[0,T], x,bRd, and a(0,M], we have:

    E[ua,b(t,x)ua,b(s,x)]=(2π)dt0s0μ(dr,dw)Rdeib.ξ(at,ras,w)e(at,r+as,w)Aaα(ξ)dξ. (3.16)

    Proof. Again, by the Wiener isometry characteristic (3.8) and the Plancheral formula, we get

    E[ua,b(t,x)ua,b(s,x)]=t0s0μ(dr,dw)RdGa,b(at,r,x,z)Ga,b(as,w,x,z)dz=(2π)dt0s0μ(dr,dw)RdFGa,b(at,r,x,.)(ξ)¯FGa,b(as,w,x,.)(ξ)dξ=(2π)dt0s0μ(dr,dw)×Rdexp(i(xat,rb).ξ)eat,rAaα(ξ)exp(i(xas,wb).ξ)eas,wAaα(ξ)dξ=(2π)dt0s0μ(dr,dw)Rdeib.ξ(trs+w)e(t+srw)Aaα(ξ)dξ.

    Then, the proof is established.

    An immediate consequence of the previous theorem.

    Corollary 3.1. For every t[0,T], x,bRd, and a(0,M], we have:

    E[|ua,b(t,x)|2]=(2π)dt0t0μ(dr,dw)Rdeib.ξ(rw)e(2trw)Aaα(ξ)dξ. (3.17)

    Remark 3.2. (1) In the case where the noise W in Eq (1.4) is defined by a Wiener process, the covariance and variance expressions, respectively, given by Eqs (3.16) and (3.17) become

    E[ua,b(t,x)ua,b(s,x)]=(2π)1ts0Reib.ξ(ts)e(t+s2r)Aaα(ξ)dξdrE[|ua,b(t,x)|2]=(2π)1t0Re2(tr)Aaα(ξ)dξdr (3.18)

    for any x,yR and t,s[0,T], leading to exactly the same expressions obtained in [34].

    (2) In the particular case where a=0 and b=0, Eq (1.4) coincides with the standard stochastic heat equation, which was studied in many references (see, for example, [24]). In fact, the expressions for covariance and variance can be deduced from Theorem 2 and Corollary 3.1.

    In this section, we will focus on the particular case where the noise is the gfBm ZH(θ,ν) with respect to the time variable (see Example 3.2), in the particular case when H>12. Consider R=RH,θ,νZ the covariance function given in (3.3) and denote

    αH=2H(2H1),cH1(θ,ν)=αHθ2+ν22andcH2(θ,ν)=αHθν,

    for any (θ,ν)R2{(0,0)}. Throughout, Cte denotes a generic positive constant, and, in what follows, for any α(1,2], we denote

    λH,αd=2Hdα. (4.1)

    We first justify the existence of the solution defined by (3.9).

    Corollary 4.1. Suppose that the noise is the gfBm ZH with respect to the time variable. If λH,αd>0, then the mild solution defined in (3.9) exists and, for every T>0, we have

    sup(t,x)[0,T]×RdE[|ua,b(t,x)|2]<. (4.2)

    Moreover, t,s[0,T], x,bRd, and a(0,M],

    E[ua,b(t,x)ua,b(s,x)]=(2π)dt0s0drdwhH,θ,ν(r,w)Rdeib.ξ(at,ras,w)e(at,r+as,w)Aaα(ξ)dξ, (4.3)

    with hH,θ,ν(r,w)=cH1(θ,ν)|rw|2H2+cH2(θ,ν)(r+w)2H2.

    Proof. From (3.4), we have clearly that

    2RH,θ,νZrw(r,w)=cH1(θ,ν)|rw|2H2+cH2(θ,ν)(r+w)2H2:=hH,θ,ν(r,w),

    for any (r,w)[0,T]×Rd. We first note that hH,θ,ν(r,w)0. Indeed, the constant cH1(θ,ν) is clearly positive. Moreover, on the one hand, if θν0 then, since H>12, the constant cH2(θ,ν) is nonnegative, and as consequence, hH,θ,ν is positive. Also, if θν0, then by writing the covariance function of ZH(θ,ν) in the following form

    hH,θ,ν(r,w)=αH{(θν)22|rw|2H2+θν[|rw|2H2(r+w)2H2]},

    we clearly see that hH,θ,ν is positive too, because, for H>1/2, we have |rw|2H2(r+w)2H2. Hence, the covariance measure generated by RH,θ,ν is positive, and by applying Theorem 3.1, we deduce that, if t0t0(2trw)d/αμ(dr,dw) is finite for any t[0,T], then the mild solution defined in (3.9) is well-defined. Therefore, there exists a positive constant depending on H,θ,ν, such that

    t0t0(2trw)d/αhH,θ,ν(r,w)drdwc(H,θ,ν)t0t0(2trw)d/α|rw|2H2drdw.

    Then, by the change of variables ˜r=tr and ˜w=tw, we get

    t0t0(2trw)d/αhH,θ,ν(r,w)drdwc(H,θ,ν)t0t0(r+w)d/α(rw)2H2drdw=c(H,θ,ν)[t0r0(r+w)d/α(rw)2H2drdw+t0tr(r+w)d/α(wr)2H2drdw]=c(H,θ,ν)[t0r0(r+w)d/α(rw)2H2drdw+t0w0(r+w)d/α(wr)2H2drdw]=2c(H,θ,ν)t0r0(r+w)d/α(rw)2H2dvdu=2c(H,θ,ν)t0r2H2dαr0(1wr)2H2(1+wr)d/αdwdr=2c(H,θ,ν)C1t0r2H1d/αdr,

    where in the last line we used the change of variables ˜w=wr and the constant C1 is given by C1=10(1˜w)2H2(1+˜w)d/αd˜w, which is finite because 2H2>1.

    Since d<2αH, we have t0r2H1d/αdr<, for every 0tT. Therefore, using Inequality (3.15), we get

    sup(t,x)I×RdE[|ua,b(t,x)|2]Ctesupt[0,T]t0t0(2trw)d/αhH,θ,ν(r,w)drdwCtesupt[0,T]t0r2H1d/αdr<.

    All this with Theorems 1 and 2 allow us to finish the proof of Corollary 4.1.

    Remark 4.1. For any α(1,2] and H(12,1), the sufficient condition, λH,αd>0, for the existence of the mild solution defined in (3.9), is equivalent to

    d=1or(d=2andαH>1)or(d=3andαH>3/2).

    In this section, we will focus on the study of the regularity of the trajectories of the solution to the stochastic FPDE (1.4) with respect to the time variable. In what follows, we suppose that H(12,1) and α(1,2] such that λH,αd>0.

    Theorem 3. Let ua,b be the mild solution to Equation (1.4). There exists a positive constant Cte such that,

    E[|ua,b(t,x)ua,b(s,x)|2]Cte|ts|λH,αd (4.4)

    for any (t,s)[0,T]2 and xRd.

    In order to prove Theorem 3, we need the following technical lemmas.

    Lemma 4.1. For every λ0, β(1,2], H(12,1), and d{1,2,3} such that λH,βd>0, the improper double integral is

    +0+0|rw|2H2gd,β,λ(r,w)drdw<,

    with gd,β,λ(r,w):=(2(1+λ)+r+w)dβ2(1+λ+r+w)dβ+(r+w)dβ.

    Proof. Denoting J(β,d,λ) as the above double integral, we have

    J(β,d,λ)=+0r0|rw|2H2gd,β,λ(r,w)drdw++0+r|rw|2H2gd,β,λ(r,w)drdw=2+0r0|rw|2H2gd,β,λ(r,w)drdw.

    For any r,w(0,), we have 2(1+λ)+r+w1+λ+r+wr+w. So, by using the fact that the function xR+xd/β is decreasing, we get

    |rw|2H2gd,β,λ(r,w)=|rw|2H2[(2(1+λ)+r+w)dβ2(1+λ+r+w)dβ+(r+w)dβ]4|rw|2H2(r+w)dβ.

    Hence, by the change of variables ˜w=wr, we get

    10r0|rw|2H2gd,β,λ(r,w)drdw410r0(rw)2H2(r+w)dβdrdw=410r2H1dβdr10(1˜w)2H2(1+˜w)dβd˜w. (4.5)

    Since H>1/2, the integral 10(1˜w)2H2(1+˜w)dβd˜w is finite, and since 2βH>d, the integral 10r2H1dβdr is also finite. Therefore, we obtain 10r0|rw|2H2gd,β,λ(r,w)dwdr<.

    Now, when u+v is close to infinity,

    |rw|2H2[(2(1+λ)+r+w)dβ2(1+λ+r+w)dβ+(r+w)dβ]=|rw|2H2(r+w)dβ[(1+2(1+λ)r+w)dβ2(1+(1+λ)r+w)dβ+1]dβ(dβ+1)(1+λ)2|rw|2H2(r+w)dβ2.

    Therefore, for every (u,v)[0,+)2 such that r1 and 0wr, we have

    |rw|2H2gd,β,λ(r,w)Cte|rw|2H2(r+w)dβ2.

    Hence, by the change of variables ˜w=wr, we get

    +1r0|rw|2H2gd,β,λ(r,w)Cte+1r0|rw|2H2(r+w)2dβdrdwCte1r2H3dβdr10|1˜w|2H2(1+˜w)2dβd˜w.

    Both integrals appearing in the last line are finite because H>12 and 2Hdβ<2. As a consequence,

    +1r0|rw|2H2gd,β,λ(r,w)dwdr<,

    which implies that J(β,d,λ) is finite.

    The following useful lemma was obtained by Balan and Tudor in [3].

    Lemma 4.2. We have

    s0s0|rw|2H2exp((r+w)z2)drdwcH(s2H+1)(11+z)2H (4.6)

    for any s[0,T] and z0 where cH denotes a positive constant depending only on H.

    Consider t,s[0,T] and xRd. Without loss of generality, we assume that st. By the Wiener isometry characteristic (3.8), we have

    E[|ua,b(t,x)ua,b(s,x)|2]=T0duT0dvhH,θ,ν(u,v)Rdz[Ga,b(at,u,x,z)1(u)(0,t)Ga,b(as,u,x,z)1(u)(0,s)]×[Ga,b(at,v,x,z)1(v)(0,t)Ga,b(as,v,x,z)1(v)(0,s)], (4.7)

    where we recall that hH,θ,ν(u,v)=cH1(θ,ν)|uv|2H2+cH2(θ,ν)(u+v)2H2. Therefore,

    E[|ua,b(t,x)ua,b(s,x)|2]tstsdvduhH,θ,ν(u,v)RdGa,b(at,u,x,z)Ga,b(at,v,x,z)dz+2tss0dvduhH,θ,ν(u,v)RdGa,b(at,u,x,z)[Ga,b(at,v,x,z)Ga,b(as,v,x,z)]dz+s0s0dvduhH,θ,ν(u,v)Rd[Ga,b(at,u,x,z)Ga,b(as,u,x,z)][Ga,b(at,v,x,y)Ga,b(as,v,x,z)]dz=Aa,b1(t,s,x)+Aa,b2(t,s,x)+Aa,b3(t,s,x).

    Let us start by the first term. By applying the Plancherel theorem, we obtain

    |Aa,b1(t,s,x)|=(2π)d|tstsdvduhH,θ,ν(u,v)RdFGa,b(at,u,x,z)¯FGa,b(at,v,x,z)dz|=(2π)d|tstsdvduhH,θ,ν(u,v)Rdeib.ξ(vu)e(2tuv)Aaα(ξ)dξ|(2π)dtstsdvdu|hH,θ,ν(u,v)|Rd|eib.ξ(vu)|e(2tuv)Aaα(ξ)dξ.

    For every H(12,1), u(0,t), and v(0,s), we have |uv|2H2(u+v)2H2, which implies that

    |hH,θ,ν(u,v)|cH(θ,ν)|au,v|2H2, (4.8)

    with cH(θ,ν)=αH(|θ|+|ν|)22. This with Inequality (3.13) implies that

    |Aa,b1(t,s,x)|(2π)dcH(θ,ν)tstsdvdu|au,v|2H2Rdeaα(2tuv)ξαdξ(2π)dcH(θ,ν)tstsdvdu|au,v|2H2(ReDα(2tuv)|ξ1|αdξ1)d

    with Dα=aαdα21. By the change of variables z1=(2tuv)1/αξ1, we get

    ReDα(2tuv)|ξ1|αdξ1=(2tuv)1/αReDα|z1|αdz1=γα(2tuv)1/α,

    with γα is defined in (3.14), which is clearly finite. This, with the change of variables U=at,u and V=at,v, then, ˜u=uat,S and ˜v=vat,S, allowing us to get:

    |Aa,b1(t,s,x)|(2π)dγdαcH(θ,ν)tsts|au,v|2H2(2tuv)d/αdvdu=(2π)dγdαcH(θ,ν)ts0ts0|au,v|2H2(u+v)d/αdvdu=(2π)dγdαcH(θ,ν)(ts)2Hdα1010|au,v|2H2(u+v)d/αdvdu=Cdα(H)(ts)2Hdα,

    with Cdα(H)=(2π)dγdαcH(θ,ν)1010|au,v|2H2(u+v)d/αdvdu, which is finite because 2αH>d.

    Now, let us consider the second term. By applying the Plancherel theorem and from Lemma 3.11, we obtain

    |Aa,b2(t,s,x)|=2|tss0hH,θ,ν(u,v)RdGa,b(at,u,x,y)[Ga,b(at,v,x,z)Ga,b(as,v,x,z)]dzdvdu|=21dπd|tss0dvduhH,θ,ν(u,v)Rd¯FGa,b(at,u,x,z)[FGa,b(at,v,x,z)FGa,b(as,v,x,z)]dz|=21dπd|tss0dvduhH,θ,ν(u,v)Rddξexp(i(xat,ub).ξ)eat,uAaα(ξ)×[exp(i(xat,vb).ξ)eat,vAaα(ξ)exp(i(xas,vb).ξ)eas,vAaα(ξ)]| (4.9)

    for any s,t[0,T] and xRd. With a simple simplification, we get

    |Aa,b2(t,s,x)|=21dπd|tss0hH,θ,ν(u,v)Rdeix.ξei(tu)b.ξe(tu)Aaα(ξ)eix.ξeivb.ξ×[exp(itb.ξ)e(tv)Aaα(ξ)exp(isb.ξ)e(sv)Aaα(ξ)]dξdvdu|=21dπd|tss0dvduhH,θ,ν(u,v)Rddξe(tu)Aaα(ξ)ei(vu)b.ξ×[e(tv)Aaα(ξ)ei(ts)b.ξe(sv)Aaα(ξ)]|=21dπd|tss0dvduhH,θ,ν(u,v)Rddξe(t+suv)Aaα(ξ)ei(vu)b.ξ×[e(ts)Aaα(ξ)ei(ts)b.ξ]|21dπdtss0dvdu|hH,θ,ν(u,v)|Rddξe(t+suv)Aaα(ξ)|ei(vu)b.ξ|×|e(ts)Aaα(ξ)ei(ts)b.ξ|21dπdtss0dvdu|hH,θ,ν(u,v)|Rddξe(t+suv)Aaα(ξ)|e(ts)Aaα(ξ)ei(ts)b.ξ|. (4.10)

    We discuss here three cases:

    First case: if d=3: It is clear that for any α(1,2], 13α<0. From (4.10), we have

    |Aa,b2(t,s,x)|22dπdtss0|hH,θ,ν(u,v)|(R3e(t+suv)Aaα(ξ)dξ)dvdu.

    Following the same technique employed above, we get

    |Aa,b2(t,s,x)|22dπdtss0|hH,θ,ν(u,v)|(Re(t+suv)|ξ1|2dξ1)3dvdu.

    By (4.8) and by the change of variables z1=(t+suv)1/2ξ1, we get

    |Aa,b2(t,s,x)|22dπ3/2dcH(θ,ν)tss0(uv)2H2(t+suv)3/2dvdu.

    Now, since u(s,t) and v(0,s), we have (uv)2H2(us)2H2. Therefore, by the change of variables ˜u=us, and ˜u=uat,s, we get

    |Aa,b2(t,s,x)|Ctets(us)2H2[(t+su)1/2(tu)1/2]duCtets(us)2H2(tu)1/2du=Cte(ts)2H3210u2H2(1u)1/2du=Cte(ts)2H32,

    where the last line is due to the fact that the integral 10u2H2(1u)1/2du is finite since 22H<1.

    Second case: if d=2: It is clear that for any α(1,2], 12α<0. From (4.10), we have

    |Aa,b2(t,s,x)|22dπdtss0|hH,θ,ν(u,v)|R2e(t+suv)Aaα(ξ)dξdvdu.22dπdtss0|hH,θ,ν(u,v)|(ReDα(t+suv)|ξ1|αdξ1)2dvdu.

    By (4.8) and by the change of variables z1=(t+suv)1/αξ1, we get

    |Aa,b2(t,s,x)|Ctetss0(uv)2H2(t+suv)2/αdvdu.

    Now, since u(s,t) and v(0,s), we have (uv)2H2(us)2H2. Therefore, by the change of variables ˜u=us, and ˜u=uat,s, we get

    |Aa,b2(t,s,x)|Ctetss0(us)2H2(t+suv)2/αdvdu=Ctets(us)2H2[(tu)2α+1(t+su)2α+1]duCtets(us)2H2(tu)2α+1du=Cte(ts)2H2α10u2H2(1u)2α+1du=Cte(ts)2H2α,

    where in the last line we used the fact that the integral 10u2H2(1u)2/α+1du is finite because 2H1>0 and 22α>0.

    Third case: if d=1: Starting from (4.10) and using (4.8), we get

    |Aa,b2(t,s,x)|π1cH(θ,ν)tss0(uv)2H2Re(t+suv)Aaα(ξ)|e(ts)Aaα(ξ)ei(ts)b.ξ|dξdvduCte{Aa,b2,1(t,s,x)+Aa,b2,2(t,s,x)},with (4.11)
    Aa,b2,1(t,s,x)=tss0(uv)2H2(Re(t+suv)Aaα(ξ)|1e(ts)Aaα(ξ)|dξ)dvduAa,b2,2(t,s,x)=tss0(uv)2H2(Re(t+suv)Aaα(ξ)|1ei(ts)bξ|dξ)dvdu.

    Concerning the first term A2,1, from the mean theorem, we have for any ξR and (t,s)I2

    |1e(ts)Aaα(ξ)||ts|Aaα(ξ).

    Moreover, since u(s,t) and v(0,s), we have (uv)2H2(us)2H2. Therefore,

    Aa,b2,1(t,s,x)=|ts|tss0(uv)2H2Re(t+suv)Aaα(ξ)Aaα(ξ)dξdvdu|ts|tss0(us)2H2Re(t+suv)Aaα(ξ)Aaα(ξ)dξdvdu=|ts|RAaα(ξ)ts(us)2H2e(tu)Aaα(ξ)s0e(sv)Aaα(ξ)dvdξdu=|ts|RAaα(ξ)ts(us)2H2e(tu)Aaα(ξ)(1esAaα(ξ)Aaα(ξ))dξdu|ts|ts(us)2H2Re(tu)Aaα(ξ)dξdu|ts|ts(us)2H2Reaα(tu)|ξ|αdξdu.

    By the change of variables z=(tu)1/αξ, v=us, and w=vat,s, we get

    Aa,b2,1(t,s,x)|ts|Reaα|z|αdzts(us)2H2(tu)1/αdu=|ts|2H1/αReaα|z|αdz10w2H2(1w)1/αdw=Cte|ts|2H1α,

    where the last line is due to the fact that both integrals Reaα|z|αdz and 10w2H2(1w)1/αdw are finite.

    Now, concerning the second term A2,2, we have

    |1eiat,sbξ|=|2ieiat,sbξ2||eiat,sbξ2eiat,sbξ22i|2|sin(at,sbξ2)||at,s||b||ξ|

    for any ξR. Therefore,

    Aa,b2,2(t,s,x)|b||ts|tss0(us)2H2Re(t+suv)Aaα(ξ)|ξ|dξdvdu=|b||ts|R|ξ|ts(us)2H2e(tu)aα|ξ|α(s0e(sv)aα|ξ|αdv)dξdu=|b||ts|R|ξ|ts(us)2H2e(tu)aα|ξ|α(1eaαs|ξ|αaα|ξ|α)dξduaα|b||ts|ts(us)2H2Re(tu)aα|ξ|α1|ξ|α1dξdu.

    By the change of variables z=(tu)1/αξ, v=us, and w=vts, we get

    Aa,b2,1(t,s,x)aα|b||ts|Reaα|z|α|z|α1dzts(us)2H2(tu)12αdu=Cte|ts|2H2α+110w2H2(1w)12αdwCte|ts|2H1α

    where the second line is due to the fact that the integral Reaα|z|α|z|α1dz is finite, and in the last line we used that 10w2H2(1w)12αdw<, because 12α>1 and 2H2>1.

    Now, let us consider the third term Aa,b3. By applying the Plancherel theorem, using again (4.8), and with some simple computations, we get:

    |Aa,b3(t,s,x)|cH(θ,ν)s0s0dvdu|uv|2H2×Rd|Ga,b(at,u,x,z)Ga,b(as,u,x,z)||Ga,b(at,v,x,z)Ga,b(at,v,x,z)|dz=(2π)dcH(θ,ν)s0s0dvdu|uv|2H2×Rd¯|FGa,b(at,u,x,z)FGa,b(as,u,x,z)||FGa,b(at,v,x,z)FGa,b(as,v,x,z)|dz=Ctes0s0dvdu|uv|2H2Rd|exp(i(xat,ub).ξ)eat,uAaα(ξ)exp(i(xas,ub).ξ)eas,uAaα(ξ)||exp(i(xat,vb).ξ)eas,vAaα(ξ)exp(i(xas,vb).ξ)eas,vAaα(ξ)|dξ=Ctes0s0dvdu|au,v|2H2Rde(2suv)Aaα(ξ)|exp((ts)[Aaα(ξ)ib.ξ])1|2dξ.

    Note that for any ξRd,s,t[0,T], we have,

    |e(ts)(ξ2+aαξαib.ξ)1|2=|ex+iy1|2

    with x=(ts)(ξ2+aαξα) and y=(ts)b.ξ, and that

    |ex+iy1|2=|ex+iy2|2|ex+iy2exiy2|2=2ex(coshxcosy)

    for all real numbers x and y. Therefore,

    |Aa,b3(t,s,x)|CteRds0s0dvdu|uv|2H2e(2suv)(ξ2+aαξα)e(ts)(ξ2+aαξα)×[cosh((ts)(ξ2+aαξα))cos((ts)b.ξ)]dξ=Cte{Aa,b3,1(t,s,x)+Aa,b3,2(t,s,x)},
    with Aa,b3,1(t,s,x)=Rds0s0dvdu|uv|2H2e(2suv)(ξ2+aαξα)e(ts)(ξ2+aαξα)×[1cos((ts)b.ξ)]dξ
    and Aa,b3,2(t,s,x)=Rds0s0dvdu|uv|2H2e(2suv)(ξ2+aαξα)e(ts)(ξ2+aαξα)×[cosh((ts)(ξ2+aαξα))1]dξ.

    Concerning the term Aa,b3,1, we first bound it in the case where d=1. From the expression of Aa,b3,1, we easily get that

    Aa,b3,1(t,s,x)R|1cos((ts)bξ)|s0s0dvdu|uv|2H2e(2suv)|ξ|2dξ.

    We note that, for any ξR, we have

    |1cos((ts)bξ)||1cos((ts)bξ)|H×21H.

    Moreover, |1cos((ts)bξ)|H=|2sin2((ts)bξ2)|H2H2H|ts|2H|b|2H|ξ|2H. It follows that

    |1cos((ts)bξ)||ξ|2H×|ts|2H|b|2H212H,

    and, consequently,

    Aa,b3,1(t,s,x)Cte|ts|2HR|ξ|2Hs0s0dvdu|uv|2H2e(2suv)|ξ|2dξCteTdα|ts|2HdαR|ξ|2Hs0s0dvdu|uv|2H2e(2suv)|ξ|2dξ=Cte|ts|2Hdα[J1+J2],

    with

    J1=|ξ|1|ξ|2Hs0s0dvdu|uv|2H2e(2suv)|ξ|2dξJ2=1|ξ||ξ|2Hs0s0dvdu|uv|2H2e(2suv)|ξ|2dξ.

    On the one hand, by the fact that |au,v|2H2e(2suv)|ξ|2s2H2 for every u,v(0,s) and ξR, we can write that, for every s[0,T] and ξR,

    J1T2H|ξ|1|ξ|2Hdξ<.

    On the other hand, by Lemma 4.2 and the Assumption 2H>1, we get

    J2Cte1|ξ||ξ|2H(1+|ξ|2)2HdξCte1|ξ|1|ξ|2Hdξ<.

    All this implies that

    Aa,b3,1(t,s,x)Cte|ts|2Hdα.

    Now, let us bound Aa,b3,1 in the case where d{2,3}. By the change of variables ˜u=2as,u and ˜v=2as,v, we get

    Aa,b3,1(t,s,x)Rde(ts)Aaα(ξ)|1cos((ts)b.ξ)|s0s0|uv|2H2e(2suv)Aaα(ξ)dvdudξ=22HRde(ts)Aaα(ξ)|1cos((ts)b.ξ)|2s02s0|˜u˜v|2H2e(˜u+˜v)2Aaα(ξ)d˜ud˜vdξ. (4.12)

    Using again Lemma 4.2, we get

    2s02s0|uv|2H2exp((u+v)Aaα(ξ)2)dvduCte((2s)2H+1)(11+Aaα(ξ))2H (4.13)

    for any s[0,T] and ξRd.

    Moreover, for every fixed a,xR+, xeax(1+x)2H is decreasing. All this with the change of variables z=(ts)ξ, allows us to get

    Aa,b3,1(t,s,x)Cte(22HT2H+1)Rde(ts)Aaα(ξ)[1+Aaα(ξ)]2H[1cos((ts)b.ξ)]dξCteRdeaα(ts)ξα[1+aαξα]2H[1cos((ts)b.ξ)]dξCteRd1cos((ts)b.ξ)[1+aαξα]2Hdξ,=Cte|ts|2αHdRd(1|ts|α+aαzα)2H[1cos(b.z)]dz. (4.14)

    Let us discuss two cases: |ts|1 and |ts|<1.

    First case: When |ts|1,

    Aa,b3,1(t,s,x)Cte|ts|2αHdRd(11+aαzα)2H|1cos(b.z)|dzCteT2αH2H+dαd|ts|2HdαRd(11+aαzα)2Hdz,

    where in the last line we used that 2αH2H+dαd>0, due to the assumptions α>1 and 2H>dα.

    We note that the integral Rd(11+aαzα)2Hdz is finite, because 2Hα>d. All this implies the existence of a nonnegative constant Cte such that

    Aa,b3,1(t,s,x)Cte|ts|2Hdα. (4.15)

    Second case: Suppose that |ts|<1: We first note that, since α(1,2], H(12,1), 2αH>d, and since we are in the case where d{2,3}, we necessarily have αH(1,2).

    Denoting D1={zRd;z1}, and D2={zRd;z1} from (4.14), we get

    Aa,b3,1(t,s,x)Cte|ts|2αHdRd1cos(b.z)a2αHz2αHdzCte|ts|2αHd[D12zα2Hdz+D2|1cos(b.z)|zα2Hdz]Cte|ts|2Hdα[2IH1+IH2],

    with IH1=D11zα2Hdz and IH2=D2|1cos(b.z)|zα2Hdz.

    Since 2αH>d, the integral IH1 is clearly finite. As regards IH2, for any zRd, we have

    |1cos(b.z)|=|1cos(b.z)|αH2×|1cos(b.z)|1αH2

    and, as consequence, |1cos(b.z)||1cos(b.z)|αH2×21αH2. Moreover,

    |1cos(b.z)|αH2=|2sin2(b.z2)|αH22αH2bαHzαH.

    It follows that

    |1cos(b.z)|21αHbαHzαH

    and, consequently, since αH<2, we have

    IH2CteD21zαHdz<.

    Therefore, Inequality (4.15) is satisfied for every (t,s)[0,T]2.

    Let us consider the term Aa,b3,2. We discuss two cases:

    First case: if |ts|1: ξRd, we have e(ts)Aaα(ξ)[cosh((ts)Aaα(ξ))1]2. Then, by using again Lemma 4.2 and by applying the change of variables z=(ts)1/αξ, we get:

    Aa,b3,2(t,s,x)CteRd1[1+aαξα]2Hdξ=Cte(ts)2HdαRd1[(ts)+aαzα]2HdzCte(ts)2HdαRddz[1+aαzα]2H.

    Since the last integral is finite because 2αH>d, we deduce that

    Aa,b3,2(t,s,x)Cte(ts)2Hdα.

    Second case: if |ts|1: Denoting E1={ξRd;ξ(at,s)1/2} and E2={ξRd;ξ(at,s)1/2}, we can write

    Aa,b3,2(t,s,x)=Ca,b1(t,s,x)+Ca,b2(t,s,x),with (4.16)
    Ca,b1(t,s,x)=E1s0s0dvdu|uv|2H2e(2suv)(ξ2+aαξα)e(ts)(ξ2+aαξα)×[cosh((ts)(ξ2+aαξα))1]dξ
    Ca,b2(t,s,x)=E2s0s0dvdu|uv|2H2e(2suv)(ξ2+aαξα)e(ts)(ξ2+aαξα)×[cosh((ts)(ξ2+aαξα))1]dξ.

    We note that if ξE1, then, ξD1. Moreover, the function xex[cosh(x)1] is increasing on R+, and Aaα(ξ)(1+aα)ξ2 for every ξD1. Therefore,

    e(ts)Aaα(ξ)[cosh((ts)Aaα(ξ))1]e(ts)(1+aα)ξ2[cosh((ts)ξ2(1+aα))1],

    for any ξD1, and, consequently, using Lemma 4.2, then by the change variable z=(ts)1/2ξ, we get:

    Ca,b1(t,s,x)E1e(ts)(1+aα)ξ2[cosh((ts)(1+aα)ξ2)1]×s0s0|uv|2H2e(2suv)Aaα(ξ)dvdudξCteE1e(ts)(1+aα)ξ2[cosh((ts)ξ2(1+aα))1][1+Aaα(ξ)]2HdξCteξ(ts)1/2e(1+aα)(ts)ξ2[cosh((ts)ξ2(1+aα))1]ξ4Hdξ=Cte|ts|2Hd2z1e(1+aα)z2[cosh(z2(1+aα))1]z4HdzCte|ts|2Hdαz11z4Hdz=Cte|ts|2Hdα,

    where in the last line we used that the integral z1dzz4H is finite since 4H>2αH>d. Now, concerning Ca,b2, it can be written as: Ca,b2(t,s,x)=Ca,b2,1(t,s,x)+Ca,b2,2(t,s,x),with

    Ca,b2,1(t,s,x)=ξ1[cosh((ts)Aaα(ξ))1]s0s0|uv|2H2e(2suv)Aaα(ξ)e(ts)Aaα(ξ)dvdudξ,Ca,b2,2(t,s,x)=1ξ(ts)1/2[cosh((ts)Aaα(ξ))1]×s0s0|uv|2H2e(2suv)Aaα(ξ)e(ts)Aaα(ξ)dvdudξ.

    Using again the fact that the function xex[cosh(x)1] is increasing on R+; and that Aaα(ξ)(1+aα)ξα for any ξD2, we obtain

    Ca,b2,1(t,s,x)D2s0s0|uv|2H2e(2suv)Aaα(ξ)e(1+aα)(ts)ξα×[cosh((ts)(1+aα)ξα)1]dvdudξD2s0s0|uv|2H2eaα(2suv)ξαe(1+aα)(ts)ξα×[cosh((ts)(1+aα)ξα)1]dvdudξ

    We note that for any ξD2 and u,v(0,s)[0,T], we have

    e(2suv)aαξα=e(2suv)(1+aα)ξαe(2suv)ξαe(2suv)(1+aα)ξαe2T.

    Moreover, for any a>0,u,v(0,s) and (t,s)[0,T]2, we have:

    ea(ts)ea(2suv)[cosh(a(ts))1]=12[ea(2tuv)2ea(t+suv)+ea(2suv)]. (4.17)

    Therefore, making the change of variables U=as,uat,s,V=as,vat,s, and z=(ts)1/αξ, we get

    Ca,b2,1(t,s,x)CteD2s0s0|uv|2H2[e(2tuv)(1+aα)ξα2e(t+suv)(1+aα)ξα+e(2suv)(1+aα)ξα]dξdvduCteRds0s0|uv|2H2[e(2tuv)(1+aα)ξα2e(t+suv)(1+aα)ξα+e(2suv)(1+aα)ξα]dξdvduCteRds0s0dvdu|uv|2H2[e(ts)(2+suts+svts)(1+aα)ξα2e(ts)(1+suts+svts)(1+aα)ξα+e(ts)(suts+svts)(1+aα)ξα]dξ=Cte|ts|2HRdsat,s0sat,s0dVdU|UV|2H2×[eat,s(2+U+V)(1+aα)ξα2eat,s(1+U+V)(1+aα)ξα+eat,s(U+V)(1+aα)ξα]dξCte|ts|2HdαRd+0+0|uv|2H2×[e(2+u+v)(1+aα)zα2e(1+u+v)(1+aα)zα+e(u+v)(1+aα)zα]dzdvdu.

    Now, making the changes of variables ξ=(2+u+v)1/αz,ξ=(1+u+v)1/αz, and ξ=(u+v)1/αz, we obtain

    Ca,b2,1(t,s,x)CteI(α,d)Rde(1+aα)ξαdξ|ts|2Hdα,

    with

    I(α,d)=+0+0|uv|2H2[(2+u+v)dα2(1+u+v)dα+(u+v)dα]dvdu.

    The above Gaussian integral Rde(1+aα)ξαdξ is clearly finite. Applying Lemma 4.1 with λ=0 and β=α, we get that I(α,d) is also finite. As a consequence,

    Ca,b2,1(t,s,x)Cte|ts|2Hdα. (4.18)

    Now, let us investigate Ca,b2,2. The function xex[cosh(x)1] is increasing on R+ and for any ξ, such that ξ1, we have Aaα(ξ)(1+aα)ξ2. Proceeding as above, we get:

    Ca,b2,2(t,s,x)1ξ(ts)1/2e(1+aα)(ts)ξ2[cosh((ts)(1+aα)ξ2)1]×s0s0|uv|2H2e(2suv)Aaα(ξ)dvdudξ1ξ(ts)1/2e(1+aα)(ts)ξ2[cosh((ts)(1+aα)ξ2)1]×s0s0e(2suv)(1+aα)ξ2eaα(2suv)ξ2|uv|2H2dvdudξ.

    Now, using (4.17) and denoting E={ξRd;1ξ(ts)1/2}, we get;

    Ca,b2,2(t,s,x)Edξs0s0dvdueaα(2suv)ξ2|uv|2H2×[e(2tuv)(1+aα)ξ22e(t+suv)(1+aα)ξ2+e(2suv)(1+aα)ξ2]=Edξs0s0dvdu|uv|2H2×[eaα(2suv)ξ2e(2tuv)(1+aα)ξ22eaα(2suv)ξ2e(t+suv)(1+aα)ξ2+eaα(2suv)ξ2e(2suv)(1+aα)ξ2]=Es0s0|uv|2H2×[e2aα(ts)ξ2e(2tuv)ξ22eaα(ts)ξ2e(t+suv)ξ2+e(2suv)ξ2]dξdvdu.

    By the change variable U=suat,s,V=svat,s, and z=(ts)1/2ξ, we obtain

    Ca,b2,2(t,s,x)|ts|2HEsts0sts0dξdvdu|uv|2H2×[e(ts)(2+u+v)ξ2e2aα(ts)ξ22e(ts)(1+u+v)ξ2eaα(ts)ξ2+e(ts)(u+v)ξ2]|ts|2HE+0+0dξdvdu|uv|2H2×[e(ts)(2(1+aα)+u+v)ξ22e(ts)(1+aα+u+v)ξ2+e(ts)(u+v)ξ2]|ts|2Hd2+0+0|uv|2H2×Rd[e(2(1+aα)+u+v)z22e(1+aα+u+v)z2+e(u+v)z2]dξdvdu.

    Therefore, by the changes of variables ξ=(2(1+aα)+u+v)1/2z,ξ=(1+aα+u+v)1/2z, and ξ=(u+v)1/2z, we obtain

    Ca,b2,2(t,s,x)Cte|ts|2HdαRdeξ2dξ×+0+0|uv|2H2[(2(1+aα)+u+v)d22(1+aα+u+v)d2+(u+v)d2]dvdu.

    The Gaussian integral Rdeξ2dξ is finite. Applying Lemma 4.1 with λ=aα and β=2, we also get that the improper double integral above is finite. It follows that

    Ca,b2,2(t,s,x)Cte|ts|2Hdα. (4.19)

    Gathering (4.16), (4.18), and (4.19) we get

    Aa,b3,2(t,s,x)Cte|ts|2Hdα. (4.20)

    This, with (4.15) allows us to achieve the proof of Theorem 3.

    As an immediate consequence of Theorem 3, applying Kolmogorov's criterion of continuity, we get:

    Corollary 4.2. Let ua,b be the mild solution to Equation (1.4) and assume that λH,αd>0. Then, for every xRd, the process tua,b(t,x) is Hölder continuous of order δ(0,λH,αd2).

    The following interesting theorem will allow us to show the non-differentiability of the trajectories of the process ua,b(.,x).

    Theorem 4. Let ua,b be the mild solution to Equation (1.4) and assume that λH,αd>0. There exists a positive constant Cte such that

    Cte|ts|λH,2dE[|ua,b(t,x)ua,b(s,x)|2] (4.21)

    for any (t,s)[0,T]2 and xRd.

    Proof. Denoting: c1(θ,ν)=(θ+ν)221{θν0}+(θν)221{θν0}, c2(θ,ν)=θν1{θν0}, and c3(θ,ν)=θν1{θν0}, using (4.7) and (3.4), we get

    E[|ua,b(t,x)ua,b(s,x)|2]=3i=1Ta,bi(t,s,x),

    with

    Ta,bi(t,s,x)=ci(θ,ν)T0T02RH,iZuv(u,v)Rd[Ga,b(at,u,x,z)1(0,t)(u)Ga,b(as,u,x,z)1(0,s)(u)]×[Ga,b(at,v,x,z)1(0,t)(v)Ga,b(as,v,x,z)1(0,s)(v)]dzdudv,

    for every i{1,2,3}, with RH,1Z=RH,1,0Z, RH,2Z=RH,1,1Z, and RH,3Z=RH,1,1Z as the covariance functions of ZH(1,0), ZH(1,1), and ZH(1,1), respectively (see Example 3.1).

    The three terms Ta,bi,i{1,2,3}, are nonnegative. Indeed, for every i{1,2,3}, we have

    Ta,bi(t,s,x)=ci(θ,ν)E[(Va,bi(t,x)Va,bi(s,x))2];
    Va,bi(t,x)Va,bi(s,x)=T0Rd[Ga,b(at,u,x,z)1(0,t)(u)Ga,b(as,u,x,z)1(0,s)(u)]WHi(dz,du)

    where WHi={WHi(t,A);(t,A)[0,T]×Bb(Rd)} is a centered Gaussian field with covariance:

    E(WHi(t,A)WHi(s,B))=RH,iZ(t,s)λd(AB). (4.22)

    Therefore, E[(ua,b(t,x)ua,b(s,x))2]Ta,b1(t,s,x). Since ZH(1,0) is none other than the fBm, by the known transfer formula (see, e.g., Proposition 2.4 in [22]), we obtain:

    Va,b1(t,x)Va,b1(s,x)=RdR(T0[Ga,b(at,u,x,y)1(0,t)(u)Ga,b(as,u,x,y)1(0,s)(u)](uz)H32+du)W(dy,dz),

    where the process W={W(t,A);t[0,T],ABb(Rd)} is a space time white Noise with covariance

    E(W(t,A)W(s,B))=(ts)λd(AB). (4.23)

    Therefore, by Wiener isometry, we get

    E[(Va,b1(t,x)Va,b1(s,x))2]=RRd(R1(0,T)(u)[Ga,b(at,u,x,y)1(0,t)(u)Ga,b(as,u,x,y)1(0,s)(u)](au,z)H32+du)2dydztsRd(R1(0,T)(u)[Ga,b(at,u,x,y)1(0,t)(u)Ga,b(as,u,x,y)1(0,s)(u)](au,z)H32+du)2dydz=tsRd(tzGa,b(at,u,x,y)aH32u,zdu)2dydz=tsRd(tzGa,b(at,u,x,y)aH32u,zdu)(tzGa,b(at,v,x,y)aH32v,zdv)dydz=tsdutsdvRddyGa,b(at,u,x,y)Ga,b(at,v,x,y)uvsaH32v,zaH32u,zdz.

    By the change of variables Z=av,zau,z1v<u+au,zav,z1v>u, we see that

    uvsaH32v,zaH32u,zdz=|uv|2H2av,sau,sav,sau,s0(1Z)12HZH32dZ.

    This, with (3.10), allows us to obtain:

    E[|ua,b(t,x)ua,b(s,x)|2]Ctetstsdudv|uv|2H2av,sau,sav,sau,s0(1z)12HzH32dz×Rddyad/2t,uad/2t,vexp(c2xy2at,u)exp(c2xy2at,v)dy=Ctetstsdudv|uv|2H2(vs)(us)(vs)(us)0(1z)12HzH32dz×((tu)1/2(tv)1/2Rexp(c2|x1y1|2(tu))exp(c2|x1y1|2(tv))dy1)d.

    By the change variable Y=(x1y1)c2(2tuv)(tu)(tv), we get Rexp(c2(x1y1)2(2tuv)(tu)(tv))dy1=π(tu)(tv)c2(2tuv). This, with the changes of variables U=au,s and V=av,s, then ˜u=Uat,s and ˜v=Vat,s, we get:

    E[|ua,b(t,x)ua,b(s,x)|2]Ctetstsdudv|uv|2H2(2tuv)d/2(vs)(us)(vs)(us)0(1z)12HzH32dzCte|ts|2Hd21010dudv|uv|2H2(2uv)d/2vuvu0(1z)12HzH32dzCte|ts|2Hd2,

    where in the last line we used that the last double integral is finite because 4H>2αH>d.

    Corollary 4.3. Let ua,b be the mild solution to Equation (1.4) and assume that λH,αd>0. For every fixed xRd, we have

    limϵ0supt[t0ϵ,t0+ϵ]|ua,b(t,x)ua,b(t0,x)tt0|=+

    with probability one for every t0. Consequently, the trajectories of the process ua,b(.,x) are not differentiable.

    Proof. The corollary can be obtained by applying Theorem 4 and by proceeding as in the proof of Theorem 3.3, page 88, in [16].

    Remark 4.2. (1) The particular case where a=1 and b=0, La,b reduced to L1,0=Δα/2, which has been examined by various authors, such as in [10,11].

    (2) The case where the noise is fBm or sfBM can be directly derived from this paper, as it presents a specific instance of gfBm.

    In this paper, we have introduced and analyzed a novel stochastic FPDE that integrates a mixed operator, combining the standard Laplacian, fractional Laplacian, and gradient operator. This approach provides an effective framework for modeling complex phenomena where standard and fractional diffusion processes interact with spatially dependent randomness.

    Our investigation allows us to analyze the complex behaviors of the solution under different random noise structures. The explicit form of the covariance function derived from our analysis reveals how the stochastic component affects the solution's properties. This result is crucial for understanding the interplay between the deterministic and stochastic elements in such FPDEs.

    The specific case of noise which behaves as a Wiener process with respect to the space variable, and as a gfBm with respect to the time variable offers additional insights into the regularity of sample paths. By focusing on this case, we explore how the fractional nature of the noise influences the solution's smoothness and continuity. This examination is particularly relevant for applications where the underlying random processes exhibit long-range dependencies or memory effects.

    Our results provide a foundation for further exploration into more complex scenarios and applications. For instance, the mixed operator FPDEs could be applied to areas such as turbulence modeling, financial mathematics, and environmental sciences where both local and nonlocal effects are significant. Future work could extend our results by considering more general forms of noise or by developing numerical methods to simulate such FPDEs effectively.

    In summary, we have developed and analyzed a novel stochastic FPDE incorporating a mixed operator with standard, fractional, and gradient components. Our study successfully derived an explicit covariance function for solutions influenced by spatially-dependent random noise. Additionally, by considering noise which behaves as a Wiener process with respect to the space variable, and as a gfBm with respect to the time variable, we provided insights into how fractional noise affects the regularity of solutions. These contributions advance the field of stochastic FPDEs and offer a robust framework for future research in complex systems with both local and nonlocal dynamics.

    A visual summary of our research contributions is presented below.

    Feature Existing research Contribution of this research
    Operator Often restricted to General mixed fractional operator
    Δ+Δα/2, Δ+aαΔα/2+b.
    Δ or Δα/2
    Random Noise White-space Gaussian field with White-space Gaussian field with
    temporal-covariance measure structure temporal covariance measure structure
    restricted to some with focus on the more
    particular gaussian processes general gaussian process: gfBm,
    as e.g. fBm, sfBm, etc. extending both fBm and sfBm
    Results Characterization of the solution Groundbreaking investigation of the
    especially in fBm and sfBm cases generalized gfBm case

    M. Zili: Conceptualization, formal analysis, investigation, resources, writing-original draft and editing. E. Zougar: Conceptualization, formal analysis and methodology, investigation, resources, writing-original draft, review and editing. M. Rhaima: Methodology, funding acquisition, project administration, writing-review and editing. All authors have read and agreed to the published version of the manuscript.

    The authors extend their appreciation to King Saud University in Riyadh, Saudi Arabia for funding this research work through researchers Supporting Project Number (RSPD2024R683).

    The authors declare that they have no conflict of interest.



    [1] P. S. Addison, The illustrated wavelet transform handbook: Introductory theory and applications in science, engineering, medicine and finance, CRC Press, 2016. https://doi.org/10.1201/9781315372556
    [2] R. Balan, D. Conus, A note on intermittency for the fractional heat equation, Stat. Probab. Lett., 95 (2014), 6–14. https://doi.org/10.1016/j.spl.2014.08.001 doi: 10.1016/j.spl.2014.08.001
    [3] R. M. Balan, C. A. Tudor, The stochastic wave equation with fractional noise: A random field approach, Stoch. Proc. Appl., 120 (2010), 2468–2494. https://doi.org/10.1016/j.spa.2010.08.006 doi: 10.1016/j.spa.2010.08.006
    [4] G. Boffetta, R. E. Ecke, Two-dimensional turbulence, Annu. Rev. Fluid Mech., 44 (2012), 427–451. https://doi.org/10.1146/annurev-fluid-120710-101240
    [5] Z. Q. Chen, E. Hu, Heat kernel estimates for Δ+Δα/2 under gradient perturbation, Stoch. Proc. Appl., 125 (2015), 2603–2642. https://doi.org/10.1016/j.spa.2015.02.016 doi: 10.1016/j.spa.2015.02.016
    [6] C. Elnouty, M. Zili, On the sub-mixed fractional Brownian motion, Appl. Math. J. Chin. Univ., 30 (2015), 27–43. https://doi.org/10.1007/s11766-015-3198-6 doi: 10.1007/s11766-015-3198-6
    [7] A. W. Jayawardena, Environmental and hydrological systems modelling, CRC Press, 2013. https://doi.org/10.1201/9781315272443
    [8] Z. Jie, M. Ijaz Khan, K. Al-Khaled, E. El-Zahar, N. Acharya, A. Raza, et al., Thermal transport model for Brinkman type nanofluid containing carbon nanotubes with sinusoidal oscillations conditions: a fractional derivative concept, Wave. Random Complex, 2022 (2022), 1–20. https://doi.org/10.1080/17455030.2022.2049926 doi: 10.1080/17455030.2022.2049926
    [9] B. Guo, X. Pu, F. Huang, Fractional partial differential equations and their numerical solutions, World Scientific, 2015.
    [10] C. Tudor, Z. Khalil-Mahdi, On the distribution and q-variation of the solution to the heat equation with fractional Laplacian, Probab. Math. Stat. 39 (2019), 315–335. https://doi.org/10.19195/0208-4147.39.2.5
    [11] Z. Khalil-Mahdi, C. Tudor, Estimation of the drift parameter for the fractional stochastic heat equation via power variation, Mod. Stoch. Theory App., 6 (2019), 397–417. https://doi.org/10.15559/19-VMSTA141 doi: 10.15559/19-VMSTA141
    [12] A. A. Kilbas, H. M. Srivastava, J. J. Trujillo, Theory and applications of fractional differential equations, Elsevier, 2006.
    [13] I. Kruk, F. Russo, C. A. Tudor, Wiener integrals, Malliavin calculus and covariance measure structure, J. Funct. Anal., 249 (2007), 92–142. https://doi.org/10.1016/j.jfa.2007.03.031 doi: 10.1016/j.jfa.2007.03.031
    [14] A. Lejay, Monte Carlo methods for fissured porous media: a gridless approach, Monte Carlo Methods, 10 (2004), 385–392. https://doi.org/10.1515/mcma.2004.10.3-4.385 doi: 10.1515/mcma.2004.10.3-4.385
    [15] J. C. Long, R. C. Ewing, Yucca mountain: Earth-science issues at a geologic repository for high-level nuclear waste, Annu. Rev. Earth Pl. Sc., 32 (2004), 363–401. https://doi.org/10.1146/annurev.earth.32.092203.122444 doi: 10.1146/annurev.earth.32.092203.122444
    [16] Y. Mishura, M. Zili, Stochastic analysis of mixed fractional Gaussian processes, Elsevier, 2018.
    [17] Y. Mishura, K. Ralchenko, M. Zili, E. Zougar, Fractional stochastic heat equation with piecewise constant coefficients, Stoch. Dynam., 21 (2021), 2150002. https://doi.org/10.1142/S0219493721500027 doi: 10.1142/S0219493721500027
    [18] S. Nicaise, Some results on spectral theory over networks, applied to nerve impulse transmission, In: Polynomes orthogonaux et applications, Berlin: Springer, 1985. https://doi.org/10.1007/BFb0076584
    [19] A. M. Selvam, Self-organized criticality and predictability in atmospheric flows, Cham: Springer, 2017. https://doi.org/10.1007/978-3-319-54546-2
    [20] K. Sobczyk, Stochastic differential equations with applications to physics and engineering, Springer Science & Business Media, 1991. https://doi.org/10.1007/978-94-011-3712-6
    [21] P. Tankov, Financial modelling with jump processes, Chapman and Hall/CRC, 2003. https://doi.org/10.1201/9780203485217
    [22] C. Tudor, Analysis of variations for self-similar processes, Cham: Springer, 2013. https://doi.org/10.1007/978-3-319-00936-0
    [23] C. Tudor, M. Zili, Covariance measure and stochastic heat equation with fractional noise, Fract. Calc. App. Anal., 17 (2014), 807–826. https://doi.org/10.2478/s13540-014-0199-8 doi: 10.2478/s13540-014-0199-8
    [24] C. Tudor, M. Zili, SPDE with generalized drift and fractional-type noise, Nonlinear Differ. Equ. Appl., 23 (2016), 53. https://doi.org/10.1007/s00030-016-0407-9 doi: 10.1007/s00030-016-0407-9
    [25] D. Xia, L. Yan, W. Fei, Mixed fractional heat equation driven by fractional Brownian sheet and Levy process, Math. Probl. Eng., 2017 (2017), 8059796. https://doi.org/10.1155/2017/8059796 doi: 10.1155/2017/8059796
    [26] B. J. West, Nature's patterns and the fractional calculus, Boston: De Gruyter, 2017. https://doi.org/10.1515/9783110535136
    [27] D. Xia, L. Yan, On a semi-linear mixed fractional heat equation driven by fractional Brownian sheet, Bound. Value Probl., 2017 (2017), 7. https://doi.org/10.1186/s13661-016-0736-y doi: 10.1186/s13661-016-0736-y
    [28] M. Zili, On the mixed fractional Brownian motion, J. Math. Anal. Appl., 2006 (2006), 032435. https://doi.org/10.1155/JAMSA/2006/32435 doi: 10.1155/JAMSA/2006/32435
    [29] M. Zili, Mixed sub-fractional Brownian motion, Random Operators Sto., 22 (2014), 163–178. https://doi.org/10.1515/rose-2014-0017 doi: 10.1515/rose-2014-0017
    [30] M. Zili, Mixed sub-fractional-white heat equation, J. Numer. Math. Stoch., 8 (2016), 17–35.
    [31] M. Zili, Generalized fractional Brownian motion, Mod. Stoch. Theory App., 4 (2017), 15–24. https://doi.org/10.15559/16-VMSTA71 doi: 10.15559/16-VMSTA71
    [32] M. Zili, Stochastic calculus with a special generalized fractional Brownian motion, Int. J. Appl. Math. Simul., 1 (2024), 1.
    [33] M. Zili, E. Zougar, Stochastic heat equation with piecewise constant coefficients and generalized fractional type-noise, Theor. Probab. Math. St., 104 (2021), 123–144. https://doi.org/10.1090/tpms/1150 doi: 10.1090/tpms/1150
    [34] M. Zili, E. Zougar, Mixed stochastic heat equation with fractional Laplacian and gradient perturbation, Fract. Calc. Appl. Anal., 25 (2022), 783–802. https://doi.org/10.1007/s13540-022-00037-z doi: 10.1007/s13540-022-00037-z
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(564) PDF downloads(40) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog