This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the local well-posedness of fractionally dissipated primitive equations with transport noise

Ruimeng Hu Department of Mathematics
Department of Statistics and Applied Probability
University of California
Santa Barbara, CA 93106, USA.
[email protected]
Quyuan Lin School of Mathematical and Statistical Sciences
Clemson University
Clemson, SC 29634, USA.
[email protected]
 and  Rongchang Liu Department of Mathematics, University of Arizona, Tucson, AZ 85721, USA [email protected]
Abstract.

We investigate the three-dimensional fractionally dissipated primitive equations with transport noise, focusing on subcritical and critical dissipation regimes characterized by (Δ)s/2(-\Delta)^{s/2} with s(1,2)s\in(1,2) and s=1s=1, respectively. For σ>3\sigma>3, we establish the local existence of unique pathwise solutions in Sobolev space HσH^{\sigma}. This result applies to arbitrary initial data in the subcritical case (s(1,2)s\in(1,2)), and to small initial data in the critical case (s=1s=1). The analysis is particularly challenging due to the loss of horizontal derivatives in the nonlinear terms and the lack of full dissipation. To address these challenges, we develop novel commutator estimates involving the hydrostatic Leray projection.

MSC Subject Classifications: 35Q86, 60H15, 76M35, 35Q35, 86A10

Keywords: stochastic primitive equations; transport noise; well-posedness; fractional dissipation; hydrostatic Leray projection

1. Introduction

In this paper, we establish the local existence and uniqueness of pathwise solutions in Sobolev spaces for the following three-dimensional fractionally dissipated primitive equations (PE, also called the hydrostatic Navier-Stokes equations) with Stratonovich transport noise, defined on 𝕋3=3/3\mathbb{T}^{3}=\mathbb{R}^{3}/\mathbb{Z}^{3},

dV+(VhV+wzV+f0V+hp)dt=ΛsVdt+k=1[hp~k+bkV]dWk,zp=0,zp~k=0,hV+zw=0,\displaystyle\begin{split}&dV+(V\cdot\nabla_{h}V+w\partial_{z}V+f_{0}V^{\perp}+\nabla_{h}p)dt=-\Lambda^{s}Vdt+\sum_{k=1}^{\infty}\left[\nabla_{h}\widetilde{p}_{k}+b_{k}\cdot\nabla V\right]\circ dW^{k},\\ &\partial_{z}p=0,\quad\partial_{z}\widetilde{p}_{k}=0,\\ &\nabla_{h}\cdot V+\partial_{z}w=0,\end{split} (1.1)

where Λs=(Δ)s/2\Lambda^{s}=(-\Delta)^{s/2} denotes the fractional Laplacian with s[1,2)s\in[1,2), h=(x1,x2)\nabla_{h}=(\partial_{x_{1}},\partial_{x_{2}}) is the 2D horizontal gradient and =(x1,x2,z)\nabla=(\partial_{x_{1}},\partial_{x_{2}},\partial_{z}) is the full 3D gradient. Here VV and ww represent the horizontal and vertical velocity component, respectively, f0f_{0}\in\mathbb{R} is the Coriolis parameter, pp denotes the pressure, and (p~k)k1(\widetilde{p}_{k})_{k\geq 1} are the components of the turbulent pressure [51]. (Wk)k1(W^{k})_{k\geq 1} is a sequence of independent standard Brownian motions, and (bk)k1(b_{k})_{k\geq 1} are divergence-free vectors representing the coefficients of the transport noise. The system satisfies the following boundary conditions:

V,w are periodic in (x1,x2,z) with period 1,V is even in z and w is odd in z.V,w\text{ are periodic in }(x_{1},x_{2},z)\text{ with period }1,\qquad V\text{ is even in }z\text{ and }w\text{ is odd in }z. (1.2)

Note that when the noise bk=(bk1,bk2,bk3)b_{k}=(b_{k}^{1},b_{k}^{2},b_{k}^{3}) satisfies (bk1,bk2)(b_{k}^{1},b_{k}^{2}) being even in zz and bk3b_{k}^{3} being odd in zz, the domain consisting of periodic functions adhering to such symmetry conditions remains invariant under the dynamics of system (1.1).

The fully viscous PE (s=2s=2) is derived from the Navier-Stokes equations [9, 49] and is widely used in the study of large-scale oceanic and atmospheric dynamics. In the deterministic setting, the global existence and uniqueness of strong solutions have been demonstrated in [14, 41]. On the other hand, the inviscid PE (s=0s=0) is ill-posed in Sobolev spaces and the Gevrey class of order strictly greater than 1 [53, 33, 38]. While they are locally well-posed in the analytic class [44, 32], finite time blowup of solutions has been established [13, 58, 38, 17]. In the stochastic setting, the well-posedness of the fully viscous PE has been investigated under the influence of multiplicative noise [23, 24] and transport noise [10, 6, 7, 4]. The inviscid PE with multiplicative noise perturbation has been studied more recently in [35, 36, 37]. These studies collectively highlight the critical role of viscosity in determining the well-posedness/ill-posedness and global existence/finite time blowup of solutions to the PE system. Contrasting to the Navier-Stokes and Euler equations, such a property is unique to the PE system. This naturally motivates the study of PE systems with fractional dissipation, which interpolate between the fully viscous and inviscid cases, and the exploration of how the system’s behavior evolves as the dissipation index varies.

Fractional dissipation introduces “weaker” and “non-local” dissipation compared to the classical dissipation arising from the full Laplacian. This poses significant challenges in mathematical analysis while offering intriguing opportunities for modeling physical phenomena. Fractional dissipation effectively captures anomalous diffusion, memory effects, and non-local interactions, making it a powerful tool in turbulence modeling [11, 15]. Its applications have expanded across fluid dynamics, appearing in studies of the Euler equations [18, 20], Boussinesq equations [60, 61], quasi-geostrophic equations [40, 12, 1], and magnetohydrodynamics [59, 22]. However, very little is known about the PE with fractional dissipation. To the best of our knowledge, the only result in this direction is an ongoing work [2] by the second author of this paper and his collaborators, which explores 2D fractional dissipative PE in the deterministic setting. On the other hand, nothing has been investigated for the 3D fractional dissipative PE, and the study of this system in the stochastic setting is completely open.

Transport noise, initially introduced in [42, 43] to model small-scale turbulence effects, has also gained prominence in stochastic fluid mechanics. It has been extensively studied in the context of stochastic Navier-Stokes equations [51, 52, 31] and has more recently been applied to the fully viscous PE [6, 7, 4]. However, existing results do not extend to the fractionally dissipated PE due to the lack of full dissipation. In this work, we focus on the Stratonovich formulation of transport noise for several reasons. First, Stratonovich noise is more compatible with numerical simulations due to Wong-Zakai convergence results [27] and two-scale type arguments [31, 25]. Second, energy estimates in Sobolev spaces are unattainable in the Itô formulation because weak dissipation cannot sufficiently counteract energy input from the noise. Finally, recent studies have demonstrated the regularization effects of Stratonovich transport noise [30, 28, 50, 5], suggesting that such noise may play a pivotal role in understanding the fractionally dissipated PE. These insights motivate our exploration of the system’s behavior under combined fractional dissipation and transport noise.

Challenges and key innovations. We highlight some mathematical difficulties in the analysis of system (1.1) and discuss important innovations. In contrast to the 2D case, where global existence of solutions in the subcritical regime is established [2], the 3D setting lacks several favorable properties present in two dimensions. As a result, the techniques and methodologies employed in [2] cannot be directly applied, and global existence may not be expected in the 3D case. Instead, we focus on establishing local well-posedness.

A primary difficulty arises from the lack of full dissipation, requiring careful treatment of the second order derivatives from the Itô-Stratonovich corrector. A crucial step involves leveraging the cancellation between the Itô-Stratonovich corrector and the energy input from the noise after applying Itô’s formula. Specifically, we derive the key estimate:

Λσ𝒫(bk𝒫(bk)V),ΛσV+Λσ𝒫(bkV),Λσ𝒫(bkV)bkVσ+122,\displaystyle\langle\Lambda^{\sigma}\mathcal{P}(b_{k}\cdot\nabla\mathcal{P}(b_{k}\cdot\nabla)V),\Lambda^{\sigma}V\rangle+\langle\Lambda^{\sigma}\mathcal{P}(b_{k}\cdot\nabla V),\Lambda^{\sigma}\mathcal{P}(b_{k}\cdot\nabla V)\rangle\lesssim_{b_{k}}\|V\|_{\sigma+\frac{1}{2}}^{2}, (1.3)

where 𝒫\mathcal{P} is the hydrostatic Leray projection defined in (2.1), and Vσ\|V\|_{\sigma} denotes the Sobolev HσH^{\sigma} norm.

In the context of the Euler equations [21, 47] and Boussinesq equations [8] in vorticity form, similar estimates to (1.3), though without 𝒫\mathcal{P}, have been derived by exploiting commutator cancellations of pseudo-differential operators [34, 57], resulting in bounds of Vσ2\|V\|_{\sigma}^{2}. These results were further generalized in [56], encompassing the usual Leray projection and various fluid models. However, the hydrostatic Leray projection behaves less favorably than the usual Leray projection. Consequently, prior results cannot be directly applied. For the hydrostatic Leray projection, we establish the following commutator estimate (see Lemma B.4):

[𝒫,bk]fσbkσ+1fσ+zbkhσ1fσ+1,\displaystyle\|[\mathcal{P},b_{k}\cdot\nabla]f\|_{\sigma}\lesssim\|b_{k}\|_{\sigma+1}\|f\|_{\sigma}+\|\partial_{z}b_{k}^{h}\|_{\sigma-1}\|f\|_{\sigma+1},

for sufficiently smooth ff. Here the commutator of two operators AA and BB is denoted by [A,B]:=ABBA[A,B]:=AB-BA, and bkhb_{k}^{h} is the horizontal component of the vector field bkb_{k}. Notably, it is known that under the usual Leray projection a better bound fσ\|f\|_{\sigma} can be achieved instead of fσ+1\|f\|_{\sigma+1}. This is essentially due to the fact that the symbol of the hydrostatic Leray projection is singular along the entire k3k_{3}-axis in frequency space, whereas the usual Leray projection has only a single singularity at the origin. Therefore, proving (1.3) requires rewriting the left-hand side into a suitable combination of commutators and carefully balancing the derivatives using commutator estimates in negative Sobolev norms (see Lemma B.3).

The bound in (1.3), which matches the order of the nonlinear term, can be controlled via interpolation with fractional dissipation in the subcritical case (s(1,2)s\in(1,2)). For the critical case (s=1s=1), we require both small initial conditions and small noise (through zbkh\partial_{z}b_{k}^{h}). In the supercritical case (s<1s<1), the dissipation is insufficient to control the highest order term due to the noise and the nonlinear term, necessitating analytic initial data, as in [35]. This is consistent with the results in [2] where the ill-posedness in the supercritical case and in the critical case with large initial data is established. Furthermore, in the supercritical case, the cancellation observed in (1.3) fails in the analytic setting unless bkb_{k} is independent of spatial variables, see remarks in Section 4. Thus, establishing well-posedness for the supercritical case with general transport noise remains an open problem.

To prove pathwise uniqueness, we apply a double cutoff technique to address difficulties arising from the nonlinear term. This approach can be applied to improve existing results, such as those in [35], demonstrating the existence of pathwise solutions in the space of analytic functions.

Organization of the paper. The rest of this paper is organized as follows. In Section 2, we introduce the mathematical preliminaries, set up the problem, and state the main result. Section 3 is devoted to proving the main result. Specifically, we derive uniform estimates for the truncated cutoff system in Section 3.1, establish the existence of martingale solutions using standard compactness arguments in Section 3.2, and prove pathwise uniqueness, thereby completing the proof of the main result in Section 3.3. Some remarks on the supercritical case are provided in Section 4. Finally, auxiliary lemmas and technical commutator estimates are presented in Appendices A and B, respectively.

2. Preliminaries and the main Result

In this section, we introduce notations and assumptions and state our main result. The universal constant CC appearing in the paper may change from line to line. When necessary, we shall use subscripts to emphasize the dependence of CC on some parameters.

2.1. Functional settings

Let x:=(x,z)=(x1,x2,z)𝕋3x:=(x^{\prime},z)=(x_{1},x_{2},z)\in\mathbb{T}^{3}, where xx^{\prime} and zz represent the horizontal and vertical variables, respectively, and 𝕋3=3/3\mathbb{T}^{3}=\mathbb{R}^{3}/\mathbb{Z}^{3} denotes the three-dimensional torus with unit volume. Denote the L2L^{2} norm of a function ff as

f:=fL2(𝕋3)=(𝕋3|f(x)|2𝑑x)12,\|f\|:=\|f\|_{L^{2}(\mathbb{T}^{3})}=\Big{(}\int_{\mathbb{T}^{3}}|f(x)|^{2}dx\Big{)}^{\frac{1}{2}},

associated with the inner product f,g=𝕋3f(x)g(x)𝑑x\langle f,g\rangle=\int_{\mathbb{T}^{3}}f(x)g(x)dx for f,gL2(𝕋3)f,g\in L^{2}(\mathbb{T}^{3}). For a function fL2(𝕋3)f\in L^{2}(\mathbb{T}^{3}), let f^k\widehat{f}_{k} denote its Fourier coefficient such that

f(x)=k2π3f^keikx,f^k=𝕋3eikxf(x)𝑑x.f(x)=\sum\limits_{k\in 2\pi\mathbb{Z}^{3}}\widehat{f}_{k}e^{ik\cdot x},\qquad\widehat{f}_{k}=\int_{\mathbb{T}^{3}}e^{-ik\cdot x}f(x)dx.

For σ0\sigma\geq 0, the Sobolev HσH^{\sigma} norm and the H˙σ\dot{H}^{\sigma} semi-norm are defined as

fHσ:=(k2π3(1+|k|2σ)|f^k|2)12,fH˙σ:=(k2π3|k|2σ|f^k|2)12.\displaystyle\hskip-7.22743pt\|f\|_{H^{\sigma}}:=\Big{(}\sum\limits_{k\in 2\pi\mathbb{Z}^{3}}(1+|k|^{2\sigma})|\widehat{f}_{k}|^{2}\Big{)}^{\frac{1}{2}},\qquad\|f\|_{\dot{H}^{\sigma}}:=\Big{(}\sum\limits_{k\in 2\pi\mathbb{Z}^{3}}|k|^{2\sigma}|\widehat{f}_{k}|^{2}\Big{)}^{\frac{1}{2}}.

Note that these two norms are equivalent for functions with zero mean. For convenience, we denote H:=L2H:=L^{2}. When a function f=f(x)f=f(x^{\prime}) depends only on the horizontal variables, we denote by HhσH^{\sigma}_{h} the corresponding Sobolev spaces. For further details on Sobolev spaces, we refer the reader to [3].

The divergence-free condition from (1.1) and the oddness of ww in zz imply that

01hV(x,z)𝑑z=0.\int_{0}^{1}\nabla_{h}\cdot V(x^{\prime},z)dz=0.

Since bkb_{k} are divergence-free, integrating the momentum equation in system (1.1) over 𝕋3\mathbb{T}^{3} shows that VV has zero mean for any t>0t>0, provided the initial data V0V_{0} has zero mean. Furthermore, as we consider VV to be even in the zz variable, it follows that VV\in\mathbb{H} where

:={fL2(𝕋3):01hf(x,z)𝑑z=0,f has zero mean and is even in z}.\mathbb{H}:=\left\{f\in L^{2}(\mathbb{T}^{3}):\int_{0}^{1}\nabla_{h}\cdot f(x^{\prime},z)dz=0,\,f\text{ has zero mean and is even in }z\right\}.

We define σ=Hσ\mathbb{H}^{\sigma}=H^{\sigma}\cap\mathbb{H} for σ>0\sigma>0 and use the notation σ\|\cdot\|_{\sigma} to denote the corresponding Sobolev norm.

The hydrostatic Leray projection 𝒫\mathcal{P} is defined as

𝒫φ=φhΔh1hφ¯\displaystyle\mathcal{P}\varphi=\varphi-\nabla_{h}\Delta_{h}^{-1}\nabla_{h}\cdot\overline{\varphi} (2.1)

for any regular velocity field φ\varphi, where φ¯(x)=01φ(x,z)𝑑z\bar{\varphi}(x^{\prime})=\int_{0}^{1}\varphi(x^{\prime},z)dz represents the barotropic part of φ\varphi. We also define

φ=(I𝒫)φ=hΔh1hφ¯,\mathbb{Q}\varphi=(I-\mathcal{P})\varphi=\nabla_{h}\Delta_{h}^{-1}\nabla_{h}\cdot\overline{\varphi},

which is a function depending only on the horizontal variables xx^{\prime}.

2.2. Assumptions on the noise

Let 𝒮=(Ω,,𝔽,)\mathcal{S}=\left(\Omega,\mathcal{F},\mathbb{F},\mathbb{P}\right) be a stochastic basis with filtration 𝔽=(t)t0\mathbb{F}=\left(\mathcal{F}_{t}\right)_{t\geq 0} satisfying the usual conditions. Let 𝒰\mathscr{U} be a separable Hilbert space, and let 𝕎\mathbb{W} be an 𝔽\mathbb{F}-adapted cylindrical Wiener process with reproducing kernel Hilbert space 𝒰\mathscr{U} on 𝒮\mathcal{S}. Let {ek}k=1\{e_{k}\}_{k=1}^{\infty} be an orthonormal basis of 𝒰\mathscr{U}, then 𝕎\mathbb{W} may be formally written as 𝕎=k=1ekWk\mathbb{W}=\sum_{k=1}^{\infty}e_{k}W^{k}, where {Wk}k=1\{W^{k}\}_{k=1}^{\infty} are independent standard Brownian motions on 𝒮\mathcal{S}. If we define the linear operators P,:𝒰HP,\mathcal{B}:\mathscr{U}\to H by

Pek=p~k,ek=bk,k1,Pe_{k}=\widetilde{p}_{k},\quad\mathcal{B}e_{k}=b_{k},\quad k\geq 1,

then the noise term in (1.1) is obtained from 𝒰\mathscr{U} as

k=1[hp~k+bkV]dWk=[hP+V]d𝕎.\displaystyle\sum_{k=1}^{\infty}\left[\nabla_{h}\widetilde{p}_{k}+b_{k}\cdot\nabla V\right]\circ dW^{k}=[\nabla_{h}P+\mathcal{B}\cdot\nabla V]\circ d\mathbb{W}.

The pressure term (p~k)k1(\widetilde{p}_{k})_{k\geq 1} will be eliminated after applying the hydrostatic Leray projector (2.1). We assume that each bk=(bk1,bk2,bk3)b_{k}=(b_{k}^{1},b_{k}^{2},b_{k}^{3}) is divergence-free and has zero mean, with bk3b_{k}^{3} being odd in zz and (bk1,bk2)(b_{k}^{1},b_{k}^{2}) being even in zz. The regularity assumption on b:=(bk)k1b:=(b_{k})_{k\geq 1} is

(bk)k12(,Hσ+3).\displaystyle(b_{k})_{k\geq 1}\in\ell^{2}(\mathbb{N},H^{\sigma+3}). (2.2)

In the case of s=1s=1, we additionally assume that

kbkσ+3zbkhσ321C,\displaystyle\sum_{k}\|b_{k}\|_{\sigma+3}\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}\leq\frac{1}{C}, (2.3)

where CC is a universal constant from Sobolev inequalities.

2.3. Notion of solutions

In this section, we introduce the notions of pathwise solutions (strong solutions in the stochastic sense) and martingale solutions (weak solutions in the stochastic sense) for system (1.1). For notational simplicity, we define:

Q(g,f)=ghf(0z(hg)(z~)𝑑z~)zf,F(g):=f0g,Bkg=bkg.Q(g,f)=g\cdot\nabla_{h}f-\left(\int_{0}^{z}(\nabla_{h}\cdot g)(\widetilde{z})d\widetilde{z}\right)\partial_{z}f,\ \ F(g):=f_{0}g^{\perp},\ \ B_{k}g=b_{k}\cdot\nabla g.

After applying the hydrostatic Leray projection (2.1), the pressure terms are eliminated, and system (1.1) can be rewritten as:

dV+[𝒫Q(V,V)+𝒫F(V)+ΛsV]dt=k=1𝒫BkVdWk,V(0)=V0.\displaystyle\begin{split}&dV+\big{[}\mathcal{P}Q(V,V)+\mathcal{P}F(V)+\Lambda^{s}V\big{]}dt=\sum_{k=1}^{\infty}\mathcal{P}B_{k}V\circ dW^{k},\\ &V(0)=V_{0}.\end{split} (2.4)

The corresponding Itô form (see, for example, [6] the conversion from Stratonovich noise to Itô noise) of the equation is:

dV+[𝒫Q(V,V)+𝒫F(V)+ΛsV]dt=12k=1(𝒫Bk)2Vdt+k=1𝒫BkVdWk,V(0)=V0.\displaystyle\begin{split}&dV+\big{[}\mathcal{P}Q(V,V)+\mathcal{P}F(V)+\Lambda^{s}V\big{]}dt=\frac{1}{2}\sum_{k=1}^{\infty}(\mathcal{P}B_{k})^{2}Vdt+\sum_{k=1}^{\infty}\mathcal{P}B_{k}VdW^{k},\\ &V(0)=V_{0}.\end{split} (2.5)
Definition 2.1 (Pathwise solution).

Let the initial condition V0L2(Ω;σ)V_{0}\in L^{2}\left(\Omega;\mathbb{H}^{\sigma}\right) be 0\mathcal{F}_{0}-measurable.

  1. (i)

    A pair (V,η)(V,\eta) is called a local pathwise solution of system (2.4) if η\eta is a strictly positive 𝔽\mathbb{F}-stopping time and V(η)V(\cdot\wedge\eta) is a progressively measurable stochastic process satisfying \mathbb{P}-a.s.,

    V(η)L2(Ω;C([0,),σ)),V\left(\cdot\wedge\eta\right)\in L^{2}\left(\Omega;C\left([0,\infty),\mathbb{H}^{\sigma}\right)\right),

    and for every t0t\geq 0, the following identity holds in \mathbb{H}:

    V(tη)+0tη[𝒫Q(V,V)+𝒫F(V)+ΛsV]𝑑r=V(0)+12k=10tη(𝒫Bk)2V𝑑r+k=10tη𝒫BkV𝑑Wrk.\displaystyle\begin{split}&V\left(t\wedge\eta\right)+\int_{0}^{t\wedge\eta}\big{[}\mathcal{P}Q(V,V)+\mathcal{P}F(V)+\Lambda^{s}V\big{]}dr\\ =&V(0)+\frac{1}{2}\sum_{k=1}^{\infty}\int_{0}^{t\wedge\eta}(\mathcal{P}B_{k})^{2}Vdr+\sum_{k=1}^{\infty}\int_{0}^{t\wedge\eta}\mathcal{P}B_{k}VdW_{r}^{k}.\end{split}
  2. (ii)

    A triple (V,(ηn)n1,ξ)\left(V,(\eta_{n})_{n\geq 1},\xi\right) is called a maximal pathwise solution if each pair (V,ηn)(V,\eta_{n}) is a local pathwise solution, ηn\eta_{n} is an increasing sequence of stopping times with limnηn=ξ\lim\limits_{n\to\infty}\eta_{n}=\xi almost surely, and

    supt[0,ηn]Vσn, on the set {ξ<}.\displaystyle\sup\limits_{t\in[0,\eta_{n}]}\|V\|_{\sigma}\geq n,\quad\text{ on the set }\{\xi<\infty\}.
Definition 2.2 (Martingale solution).

Let T>0T>0 and V0L2(Ω;σ)V_{0}\in L^{2}\left(\Omega;\mathbb{H}^{\sigma}\right) be an 0\mathcal{F}_{0}-measurable random variable. A martingale solution to equation (2.4) on [0,T][0,T] is a quadruple (V~0,V~,𝕎~,𝒮~)(\widetilde{V}_{0},\widetilde{V},\widetilde{\mathbb{W}},\widetilde{\mathcal{S}}) such that

  1. (i)

    𝒮~=(Ω~,~,𝔽~,~)\widetilde{\mathcal{S}}=(\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{F}},\widetilde{\mathbb{P}}) is a stochastic basis, over which 𝕎~\widetilde{\mathbb{W}} is an 𝔽~\widetilde{\mathbb{F}}-adapted cylindrical Brownian motion with components (W~k)k0(\widetilde{W}^{k})_{k\geq 0};

  2. (ii)

    V~0L2(Ω~,σ)\widetilde{V}_{0}\in L^{2}(\widetilde{\Omega},\mathbb{H}^{\sigma}) has the same law as V0V_{0};

  3. (iii)

    V~\widetilde{V} is a progressively measurable process such that

    V~L2(Ω~;C([0,T];σ)),\widetilde{V}\in L^{2}\left(\widetilde{\Omega};C\left([0,T];\mathbb{H}^{\sigma}\right)\right),

    ~\widetilde{\mathbb{P}}-a.s., and for every t[0,T]t\in[0,T], the following identity holds in \mathbb{H}:

    V~(t)+0t[𝒫Q(V~,V~)+𝒫F(V~)+ΛsV~]𝑑r=V~(0)+12k=10t(𝒫Bk)2V~𝑑r+k=10t𝒫BkV~𝑑W~rk.\displaystyle\begin{split}&\widetilde{V}\left(t\right)+\int_{0}^{t}\big{[}\mathcal{P}Q(\widetilde{V},\widetilde{V})+\mathcal{P}F(\widetilde{V})+\Lambda^{s}\widetilde{V}\big{]}dr\\ =&\widetilde{V}(0)+\frac{1}{2}\sum_{k=1}^{\infty}\int_{0}^{t}(\mathcal{P}B_{k})^{2}\widetilde{V}dr+\sum_{k=1}^{\infty}\int_{0}^{t}\mathcal{P}B_{k}\widetilde{V}d\widetilde{W}_{r}^{k}.\end{split}

2.4. Main result

The main result of this paper is stated in the following theorem.

Theorem 2.3.

Let σ>3\sigma>3, s(1,2)s\in(1,2) and assume the noise coefficient satisfies (2.2). Then, for any V0L2(Ω,σ)V_{0}\in L^{2}(\Omega,\mathbb{H}^{\sigma}), there exists a maximal pathwise solution (V,(ηn)n1,ξ)(V,(\eta_{n})_{n\geq 1},\xi) to system (2.4).

The same result holds for the case s=1s=1 if we assume additionally that the noise satisfies (2.3) and that V0σ<1C0\|V_{0}\|_{\sigma}<\frac{1}{C_{0}}, where C0C_{0} is a universal constant from Sobolev embeddings.

3. Proof of the main result

This section is dedicated to proving Theorem 2.3. We begin by deriving uniform estimates for the truncated cutoff system in Section 3.1. Next, we establish the existence of a martingale solution to the cutoff system in Section 3.2 using standard compactness arguments. Finally, Theorem 2.3 is proved at the end of Section 3.3 after we demonstrate pathwise uniqueness.

As mentioned in the introduction, a major difficulty arises from the singularity of the hydrostatic Leray projector and the fractional dissipation. In particular, obtaining the uniform estimates in Section 3.1 requires carefully controlling commutator estimates involving the hydrostatic Leray projector. In addition, the usual cutoff scheme is inadequate for proving pathwise uniqueness in Section 3.3 due to the nonlinear terms. To address this, a double cutoff scheme is introduced to overcome the associated challenges. Throughout this section, we assume that the noise coefficients satisfy the conditions outlined in Section 2.2.

Let θρ(x)C()\theta_{\rho}(x)\in C^{\infty}(\mathbb{R}) be a non-increasing cutoff function defined as

θρ(x)={1, if |x|ρ2,0 if |x|>ρ.\theta_{\rho}(x)=\begin{cases}1,&\text{ if }|x|\leq\frac{\rho}{2},\\ 0&\text{ if }|x|>\rho.\end{cases} (3.1)

Consider the cutoff system associated with the equation (2.5),

dV+[θρ2𝒫Q(V,V)+𝒫F(V)+ΛsV]dt=12k=1(𝒫Bk)2V+k=1𝒫BkVdWk,V(0)=V0,\displaystyle\begin{split}&dV+\big{[}\theta_{\rho}^{2}\mathcal{P}Q(V,V)+\mathcal{P}F(V)+\Lambda^{s}V\big{]}dt=\frac{1}{2}\sum_{k=1}^{\infty}(\mathcal{P}B_{k})^{2}V+\sum_{k=1}^{\infty}\mathcal{P}B_{k}VdW^{k},\\ &V(0)=V_{0},\end{split} (3.2)

where we denote by θρ=θρ(VW1,)\theta_{\rho}=\theta_{\rho}(\|V\|_{W^{1,\infty}}) for convenience.

3.1. Analysis of the Galerkin system

In this subsection, we derive a uniform energy estimate for the Galerkin system associated with the cutoff system (3.2). For k2π3k\in 2\pi\mathbb{Z}^{3}, define

ϕk=ϕk1,k2,k3:={2ei(k1x1+k2x2)cos(k3z)ifk30,ei(k1x1+k2x2)ifk3=0,\phi_{k}=\phi_{k_{1},k_{2},k_{3}}:=\begin{cases}\sqrt{2}e^{i\left(k_{1}x_{1}+k_{2}x_{2}\right)}\cos(k_{3}z)&\text{if}\;k_{3}\neq 0,\\ e^{i\left(k_{1}x_{1}+k_{2}x_{2}\right)}&\text{if}\;k_{3}=0,\end{cases}

and

:={ϕL2(𝕋3)|ϕ=k2π3akϕk,ak1,k2,k3=ak1,k2,k3,k2π3|ak|2<},\displaystyle\hskip-20.2355pt\mathcal{E}:=\left\{\phi\in L^{2}(\mathbb{T}^{3})\;|\;\phi=\sum\limits_{k\in 2\pi\mathbb{Z}^{3}}a_{k}\phi_{k},\;a_{-k_{1},-k_{2},k_{3}}=a_{k_{1},k_{2},k_{3}}^{*},\;\sum\limits_{k\in 2\pi\mathbb{Z}^{3}}|a_{k}|^{2}<\infty\right\},

where aa^{*} denotes the complex conjugate of aa. Notice that \mathcal{E} is a closed subspace of L2(𝕋3)L^{2}(\mathbb{T}^{3}), consisting of real-valued functions that are even in the zz variable.

For any nn\in{\mathbb{N}}, define the finite-dimensional subspace

n:={ϕL2(𝕋3)|ϕ=|k|nakϕk,ak1,k2,k3=ak1,k2,k3},\displaystyle\hskip-20.2355pt\mathcal{E}_{n}:=\left\{\phi\in L^{2}(\mathbb{T}^{3})\;|\;\phi=\sum\limits_{|k|\leq n}a_{k}\phi_{k},\;a_{-k_{1},-k_{2},k_{3}}=a_{k_{1},k_{2},k_{3}}^{*}\right\},

For any function fL2(𝕋3)f\in L^{2}(\mathbb{T}^{3}), denote its Fourier coefficients by

fk:=𝕋3f(x)ϕk(x)𝑑x,f_{k}:=\int_{\mathbb{T}^{3}}f(x)\phi^{*}_{k}(x)dx,

and define the projection Pnf:=|k|nfkϕkP_{n}f:=\sum_{|k|\leq n}f_{k}\phi_{k} for nn\in{\mathbb{N}}. Then, PnP_{n} is an orthogonal projection from L2(𝕋3)L^{2}(\mathbb{T}^{3}) to n\mathcal{E}_{n}.

For nn\in\mathbb{N}, the Galerkin approximation of system (3.2) at order nn is given by

dVn+[θρ2Pn𝒫Q(Vn,Vn)+Pn𝒫F(Vn)+ΛsVn]dt=12k=1Pn(𝒫Bk)2Vndt+k=1Pn𝒫BkVndWk,Vn(0)=PnV0,\displaystyle\begin{split}&dV_{n}+\big{[}\theta_{\rho}^{2}P_{n}\mathcal{P}Q(V_{n},V_{n})+P_{n}\mathcal{P}F(V_{n})+\Lambda^{s}V_{n}\big{]}dt=\frac{1}{2}\sum_{k=1}^{\infty}P_{n}(\mathcal{P}B_{k})^{2}V_{n}dt+\sum_{k=1}^{\infty}P_{n}\mathcal{P}B_{k}V_{n}dW^{k},\\ &V_{n}(0)=P_{n}V_{0},\end{split} (3.3)

where θρ=θρ(VnW1,)\theta_{\rho}=\theta_{\rho}(\|V_{n}\|_{W^{1,\infty}}). Since the coefficients are locally Lipschitz, the Galerkin system has a unique local solution. As the cancellation of the nonlinear term holds for the Galerkin system, the solution is indeed global.

We now establish the following uniform energy estimate.

Proposition 3.1.

Let p2p\geq 2, T0T\geq 0, s[1,2)s\in[1,2), σ>3\sigma>3, and let V0Lp(Ω,σ)V_{0}\in L^{p}(\Omega,\mathbb{H}^{\sigma}) be 0\mathcal{F}_{0}-measurable. In the case s=1s=1, we additionally assume that the cutoff parameter satisfies ρ<12Cσ\rho<\frac{1}{2C_{\sigma}}, where CσC_{\sigma} is a constant arising from the Sobolev embedding. Let VnV_{n} be the solution to the Galerkin system (3.3). Then for some universal constant CC independent of nn, the following results hold:

  • (1)

    Uniform energy bound:

    𝔼[supr[0,T]Vn(r)σp+0TVn(r)σp2Vn(r)σ+s22𝑑r]C(1+𝔼V0σp).\mathbb{E}\Big{[}\sup\limits_{r\in[0,T]}\|V_{n}(r)\|_{\sigma}^{p}+\int_{0}^{T}\|V_{n}(r)\|_{\sigma}^{p-2}\|V_{n}(r)\|_{\sigma+\frac{s}{2}}^{2}dr\Big{]}\leq C(1+\mathbb{E}\|V_{0}\|_{\sigma}^{p}).
  • (2)

    For any α[0,12)\alpha\in[0,\frac{1}{2}),

    𝔼|0kPn𝒫BkVndWrk|Wα,p(0,T;σ1)pC(1+𝔼V0σp).\mathbb{E}\left|\int_{0}^{\cdot}\sum_{k}P_{n}\mathcal{P}B_{k}V_{n}dW_{r}^{k}\right|_{W^{\alpha,p}(0,T;\mathbb{H}^{\sigma-1})}^{p}\leq C(1+\mathbb{E}\|V_{0}\|_{\sigma}^{p}).
  • (3)

    The following bound holds:

    𝔼Vn0kPn𝒫BkVndWrkW1,2(0,T;σ2+s2)C(1+𝔼V0σ2).\mathbb{E}\Big{\|}V_{n}-\int_{0}^{\cdot}\sum_{k}P_{n}\mathcal{P}B_{k}V_{n}dW_{r}^{k}\Big{\|}_{W^{1,2}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}})}\leq C(1+\mathbb{E}\|V_{0}\|_{\sigma}^{2}).
Proof.

Applying Itô’s formula to the functional ΛσVnp\|\Lambda^{\sigma}V_{n}\|^{p}, by (3.3) we have

dΛσVnp=pθρ2Pn𝒫Q(Vn,Vn)+Pn𝒫F(Vn)+ΛsVn,Λ2σVnΛσVnp2dt+p2k(Pn(𝒫Bk)2Vn,Λ2σVn+ΛσPn𝒫BkVn2)ΛσVnp2dt+p(p2)2ΛσVnp4kΛ2σVn,Pn𝒫BkVn2dt+pΛσVnp2kPn𝒫BkVn,Λ2σVndWk=I0dt+I1dt+I2dt+I3d𝕎.\displaystyle\begin{split}d\|\Lambda^{\sigma}V_{n}\|^{p}&=-p\langle\theta_{\rho}^{2}P_{n}\mathcal{P}Q(V_{n},V_{n})+P_{n}\mathcal{P}F(V_{n})+\Lambda^{s}V_{n},\Lambda^{2\sigma}V_{n}\rangle\|\Lambda^{\sigma}V_{n}\|^{p-2}dt\\ &+\frac{p}{2}\sum_{k}\left(\langle P_{n}(\mathcal{P}B_{k})^{2}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\|\Lambda^{\sigma}P_{n}\mathcal{P}B_{k}V_{n}\|^{2}\right)\|\Lambda^{\sigma}V_{n}\|^{p-2}dt\\ &+\frac{p(p-2)}{2}\|\Lambda^{\sigma}V_{n}\|^{p-4}\sum_{k}\langle\Lambda^{2\sigma}V_{n},P_{n}\mathcal{P}B_{k}V_{n}\rangle^{2}dt+p\|\Lambda^{\sigma}V_{n}\|^{p-2}\sum_{k}\langle P_{n}\mathcal{P}B_{k}V_{n},\Lambda^{2\sigma}V_{n}\rangle dW^{k}\\ &=I_{0}dt+I_{1}dt+I_{2}dt+I_{3}d\mathbb{W}.\end{split}

We first consider

0tI1𝑑r=0tp2k(Pn(𝒫Bk)2Vn,Λ2σVn+ΛσPn𝒫BkVn2)ΛσVnp2dr.\displaystyle\int_{0}^{t}I_{1}dr=\int_{0}^{t}\frac{p}{2}\sum_{k}\left(\langle P_{n}(\mathcal{P}B_{k})^{2}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\|\Lambda^{\sigma}P_{n}\mathcal{P}B_{k}V_{n}\|^{2}\right)\|\Lambda^{\sigma}V_{n}\|^{p-2}dr.

Note that the self-adjoint operator 𝒫\mathcal{P} commutes with Λ\Lambda in the periodic case, 𝒫Vn=Vn\mathcal{P}V_{n}=V_{n}, and each bkb_{k} in Bk=bkB_{k}=b_{k}\cdot\nabla is divergence-free. Additionally, PnP_{n} commutes with Λ\Lambda and satisfies Pn\|P_{n}\cdot\|\leq\|\cdot\|. Therefore, one has

Pn(𝒫Bk)2Vn,Λ2σVn+ΛσPn𝒫BkVn2Bk𝒫BkVn,Λ2σVn+ΛσBkVn,Λσ𝒫BkVn=Bk[𝒫,Bk]Vn,Λ2σVn+BkBkVn,Λ2σVn+ΛσBkVn,Λσ[𝒫,Bk]Vn+ΛσBkVn,ΛσBkVn=[Λσ,Bk][𝒫,Bk]Vn,ΛσVn+Λσ[𝒫,Bk]Vn,[Λσ,Bk]Vn+BkBkVn,Λ2σVn+ΛσBkVn,ΛσBkVn=[Λσ,Bk][𝒫,Bk]Vn,ΛσVn+Λσ[𝒫,Bk]Vn,[Λσ,Bk]Vn+[Λσ,Bk]BkVn,ΛσVn+[Λσ,Bk]Vn,ΛσBkVn=[Λσ,Bk][𝒫,Bk]Vn,ΛσVn+Λσ[𝒫,Bk]Vn,[Λσ,Bk]Vn+[[Λσ,Bk],Bk]Vn,ΛσVn+[Λσ,Bk]Vn,[Λσ,Bk]Vn.\displaystyle\begin{split}&\hskip 14.22636pt\langle P_{n}(\mathcal{P}B_{k})^{2}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\|\Lambda^{\sigma}P_{n}\mathcal{P}B_{k}V_{n}\|^{2}\\ &\leq\langle B_{k}\mathcal{P}B_{k}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\langle\Lambda^{\sigma}B_{k}V_{n},\Lambda^{\sigma}\mathcal{P}B_{k}V_{n}\rangle\\ &=\langle B_{k}[\mathcal{P},B_{k}]V_{n},\Lambda^{2\sigma}V_{n}\rangle+\langle B_{k}B_{k}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\langle\Lambda^{\sigma}B_{k}V_{n},\Lambda^{\sigma}[\mathcal{P},B_{k}]V_{n}\rangle+\langle\Lambda^{\sigma}B_{k}V_{n},\Lambda^{\sigma}B_{k}V_{n}\rangle\\ &=\langle[\Lambda^{\sigma},B_{k}][\mathcal{P},B_{k}]V_{n},\Lambda^{\sigma}V_{n}\rangle+\langle\Lambda^{\sigma}[\mathcal{P},B_{k}]V_{n},[\Lambda^{\sigma},B_{k}]V_{n}\rangle\\ &\qquad+\langle B_{k}B_{k}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\langle\Lambda^{\sigma}B_{k}V_{n},\Lambda^{\sigma}B_{k}V_{n}\rangle\\ &=\langle[\Lambda^{\sigma},B_{k}][\mathcal{P},B_{k}]V_{n},\Lambda^{\sigma}V_{n}\rangle+\langle\Lambda^{\sigma}[\mathcal{P},B_{k}]V_{n},[\Lambda^{\sigma},B_{k}]V_{n}\rangle\\ &\qquad+\langle[\Lambda^{\sigma},B_{k}]B_{k}V_{n},\Lambda^{\sigma}V_{n}\rangle+\langle[\Lambda^{\sigma},B_{k}]V_{n},\Lambda^{\sigma}B_{k}V_{n}\rangle\\ &=\langle[\Lambda^{\sigma},B_{k}][\mathcal{P},B_{k}]V_{n},\Lambda^{\sigma}V_{n}\rangle+\langle\Lambda^{\sigma}[\mathcal{P},B_{k}]V_{n},[\Lambda^{\sigma},B_{k}]V_{n}\rangle\\ &\qquad+\langle[[\Lambda^{\sigma},B_{k}],B_{k}]V_{n},\Lambda^{\sigma}V_{n}\rangle+\langle[\Lambda^{\sigma},B_{k}]V_{n},[\Lambda^{\sigma},B_{k}]V_{n}\rangle.\\ \end{split} (3.4)

Since σ>52\sigma>\frac{5}{2}, it follows from Lemma A.1 and the Sobolev embedding that

[Λσ,Bk]Vnj[Λσ,bkj](jVn)bkLVnσ1+ΛσbkVnLbkσVnσ,\displaystyle\begin{split}\|[\Lambda^{\sigma},B_{k}]V_{n}\|\leq\sum_{j}\|[\Lambda^{\sigma},b_{k}^{j}](\partial_{j}V_{n})\|&\lesssim\|\nabla b_{k}\|_{L^{\infty}}\|\nabla V_{n}\|_{\sigma-1}+\|\Lambda^{\sigma}b_{k}\|\|\nabla V_{n}\|_{L^{\infty}}\\ \lesssim\|b_{k}\|_{\sigma}\|V_{n}\|_{\sigma},\end{split} (3.5)

which leads to

[Λσ,Bk]Vn,[Λσ,Bk]Vnbkσ2Vnσ2.\displaystyle\langle[\Lambda^{\sigma},B_{k}]V_{n},[\Lambda^{\sigma},B_{k}]V_{n}\rangle\lesssim\|b_{k}\|_{\sigma}^{2}\|V_{n}\|_{\sigma}^{2}.

Thanks to Lemma B.2, we also have

[[Λσ,Bk],Bk]Vn,ΛσVn[[Λσ,Bk],Bk]VnΛσVnbkσ+12Vnσ2.\displaystyle\langle[[\Lambda^{\sigma},B_{k}],B_{k}]V_{n},\Lambda^{\sigma}V_{n}\rangle\leq\|[[\Lambda^{\sigma},B_{k}],B_{k}]V_{n}\|\|\Lambda^{\sigma}V_{n}\|\lesssim\|b_{k}\|_{\sigma+1}^{2}\|V_{n}\|_{\sigma}^{2}.

Applying Lemmas B.1 and B.4 we obtain

Λσ[𝒫,Bk]Vn,[Λσ,Bk]VnΛσ12[𝒫,Bk]VnΛ12[Λσ,Bk]Vn(bkσ+12Vnσ12+zbkhσ32Vnσ+12)(bkσVnσ+12+bkσ+3Vnσ)zbkhσ32bkσVnσ+122+bkσ+32VnσVnσ+12+bkσ+32Vnσ2,\displaystyle\begin{split}\langle\Lambda^{\sigma}[\mathcal{P},B_{k}]V_{n},[\Lambda^{\sigma},B_{k}]V_{n}\rangle&\leq\|\Lambda^{\sigma-\frac{1}{2}}[\mathcal{P},B_{k}]V_{n}\|\|\Lambda^{\frac{1}{2}}[\Lambda^{\sigma},B_{k}]V_{n}\|\\ &\lesssim\left(\|b_{k}\|_{\sigma+\frac{1}{2}}\|V_{n}\|_{\sigma-\frac{1}{2}}+\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}\|V_{n}\|_{\sigma+\frac{1}{2}}\right)\left(\|b_{k}\|_{\sigma}\|V_{n}\|_{\sigma+\frac{1}{2}}+\|b_{k}\|_{\sigma+3}\|V_{n}\|_{\sigma}\right)\\ &\lesssim\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}\|b_{k}\|_{\sigma}\|V_{n}\|_{\sigma+\frac{1}{2}}^{2}+\|b_{k}\|_{\sigma+3}^{2}\|V_{n}\|_{\sigma}\|V_{n}\|_{\sigma+\frac{1}{2}}+\|b_{k}\|_{\sigma+3}^{2}\|V_{n}\|_{\sigma}^{2},\end{split}

where bkh=(bk1,bk2)b_{k}^{h}=(b_{k}^{1},b_{k}^{2}). By Lemma B.4, we further have

[Λσ,Bk][𝒫,Bk]Vn,ΛσVn\displaystyle\langle[\Lambda^{\sigma},B_{k}][\mathcal{P},B_{k}]V_{n},\Lambda^{\sigma}V_{n}\rangle Λ12[Λσ,Bk][𝒫,Bk]VnΛσ+12Vn\displaystyle\leq\|\Lambda^{-\frac{1}{2}}[\Lambda^{\sigma},B_{k}][\mathcal{P},B_{k}]V_{n}\|\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|
(bkσ+32Vnσ12+bkσ+3zbkhσ32Vnσ+12)Vnσ+12\displaystyle\lesssim\left(\|b_{k}\|_{\sigma+3}^{2}\|\|V_{n}\|_{\sigma-\frac{1}{2}}+\|b_{k}\|_{\sigma+3}\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}\|V_{n}\|_{\sigma+\frac{1}{2}}\right)\|V_{n}\|_{\sigma+\frac{1}{2}}
bkσ+32VnσVnσ+12+bkσ+3zbkhσ32Vnσ+122.\displaystyle\lesssim\|b_{k}\|_{\sigma+3}^{2}\|\|V_{n}\|_{\sigma}\|V_{n}\|_{\sigma+\frac{1}{2}}+\|b_{k}\|_{\sigma+3}\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}\|V_{n}\|_{\sigma+\frac{1}{2}}^{2}.

Combining the estimates above, it follows that

Pn(𝒫Bk)2Vn,Λ2σVn+ΛσPn𝒫BkVn2bkσ+3zbkhσ32Vnσ+122+bkσ+32VnσVnσ+12+bkσ+32Vnσ2.\displaystyle\begin{split}&\langle P_{n}(\mathcal{P}B_{k})^{2}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\|\Lambda^{\sigma}P_{n}\mathcal{P}B_{k}V_{n}\|^{2}\\ &\qquad\lesssim\|b_{k}\|_{\sigma+3}\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}\|V_{n}\|_{\sigma+\frac{1}{2}}^{2}+\|b_{k}\|_{\sigma+3}^{2}\|V_{n}\|_{\sigma}\|V_{n}\|_{\sigma+\frac{1}{2}}+\|b_{k}\|_{\sigma+3}^{2}\|V_{n}\|_{\sigma}^{2}.\end{split} (3.6)

For s>1s>1, by interpolation and Young’s inequalities, and noting that VnV_{n} has zero mean, one has

Vnσ+122Λσ+12Vn2εΛσ+s2Vn2+CεΛσVn2.\|V_{n}\|_{\sigma+\frac{1}{2}}^{2}\lesssim\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}\leq\varepsilon\|\Lambda^{\sigma+\frac{s}{2}}V_{n}\|^{2}+C_{\varepsilon}\|\Lambda^{\sigma}V_{n}\|^{2}.

Consequently,

k(Pn(𝒫Bk)2Vn,Λ2σVn+ΛσPn𝒫BkVn2)14Λσ+s2Vn2+Cb2(,Hσ+3)2ΛσVn2.\displaystyle\sum_{k}\left(\langle P_{n}(\mathcal{P}B_{k})^{2}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\|\Lambda^{\sigma}P_{n}\mathcal{P}B_{k}V_{n}\|^{2}\right)\leq\frac{1}{4}\|\Lambda^{\sigma+\frac{s}{2}}V_{n}\|^{2}+C\|b\|_{\ell^{2}(\mathbb{N},H^{\sigma+3})}^{2}\|\Lambda^{\sigma}V_{n}\|^{2}. (3.7)

For s=1s=1, we impose the assumption that

kbkσ+3zbkhσ3215C,\sum_{k}\|b_{k}\|_{\sigma+3}\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}\leq\frac{1}{5C},

where CC is a universal constant. Note that this assumption is automatically satisfied when zbkh=0\partial_{z}b_{k}^{h}=0, i.e., bkhb_{k}^{h} is independent of the zz variable. Under this assumption, from (3.6) we have

k(Pn(𝒫Bk)2Vn,Λ2σVn+ΛσPn𝒫BkVn2)\displaystyle\sum_{k}\left(\langle P_{n}(\mathcal{P}B_{k})^{2}V_{n},\Lambda^{2\sigma}V_{n}\rangle+\|\Lambda^{\sigma}P_{n}\mathcal{P}B_{k}V_{n}\|^{2}\right)
CΛσ+12Vn2kbkσ+3zbkhσ32+b2(,Hσ+3)2(VnσVnσ+12+Vnσ2)\displaystyle\qquad\leq C\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}\sum_{k}\|b_{k}\|_{\sigma+3}\|\partial_{z}b_{k}^{h}\|_{\sigma-\frac{3}{2}}+\|b\|_{\ell^{2}(\mathbb{N},H^{\sigma+3})}^{2}\left(\|V_{n}\|_{\sigma}\|V_{n}\|_{\sigma+\frac{1}{2}}+\|V_{n}\|_{\sigma}^{2}\right)
14Λσ+12Vn2+C(1+b2(,Hσ+3)4)ΛσVn2.\displaystyle\leq\frac{1}{4}\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}+C\left(1+\|b\|_{\ell^{2}(\mathbb{N},H^{\sigma+3})}^{4}\right)\|\Lambda^{\sigma}V_{n}\|^{2}. (3.8)

Thus, under the given condition, for 1s<21\leq s<2, we derive

0tI1𝑑r140tΛσ+s2Vn2ΛσVnp2𝑑r+Cb0tVnσp𝑑r.\displaystyle\int_{0}^{t}I_{1}dr\leq\frac{1}{4}\int_{0}^{t}\|\Lambda^{\sigma+\frac{s}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr+C_{b}\int_{0}^{t}\|V_{n}\|_{\sigma}^{p}dr. (3.9)

To estimate I2I_{2} and I3I_{3}, note that since each bkb_{k} is divergence-free, we have

Λ2σVn,Pn𝒫BkVn\displaystyle\langle\Lambda^{2\sigma}V_{n},P_{n}\mathcal{P}B_{k}V_{n}\rangle =ΛσVn,ΛσBkVn=ΛσVn,ΛσBkVnΛσVn,BkΛσVn\displaystyle=\langle\Lambda^{\sigma}V_{n},\Lambda^{\sigma}B_{k}V_{n}\rangle=\langle\Lambda^{\sigma}V_{n},\Lambda^{\sigma}B_{k}V_{n}\rangle-\langle\Lambda^{\sigma}V_{n},B_{k}\Lambda^{\sigma}V_{n}\rangle
=ΛσVn,[Λσ,Bk]Vnbkσ+2Vnσ2,\displaystyle=\langle\Lambda^{\sigma}V_{n},[\Lambda^{\sigma},B_{k}]V_{n}\rangle\leq\|b_{k}\|_{\sigma+2}\|V_{n}\|_{\sigma}^{2},

where the last inequality follows from estimate (3.5). Consequently, we obtain

0tI2𝑑r=p(p2)20tΛσVnp4kΛ2σVn,Pn𝒫BkVn2drCb0tVnσp𝑑r.\displaystyle\begin{split}\int_{0}^{t}I_{2}dr=\frac{p(p-2)}{2}\int_{0}^{t}\|\Lambda^{\sigma}V_{n}\|^{p-4}\sum_{k}\langle\Lambda^{2\sigma}V_{n},P_{n}\mathcal{P}B_{k}V_{n}\rangle^{2}dr\leq C_{b}\int_{0}^{t}\|V_{n}\|_{\sigma}^{p}dr.\end{split}

Using the Burkholder-Davis-Gundy inequality, we then estimate I3I_{3}:

𝔼supr[0,t]|0rI3𝑑𝕎|=p𝔼supr[0,t]|0rΛσVnp2kPn𝒫BkVn,Λ2σVndWk|Cp𝔼(0tVnσ2(p2)kBkVn,Λ2σVn2dr)12=Cp𝔼(0tVnσ2(p2)k[Λσ,Bk]Vn,ΛσVn2dr)12Cb𝔼(0tVnσ2p𝑑r)1214𝔼supr[0,t]Vnσp+4Cb2𝔼0tVn(r)σp𝑑r.\displaystyle\begin{split}\mathbb{E}\sup_{r\in[0,t]}\left|\int_{0}^{r}I_{3}d\mathbb{W}\right|&=p\mathbb{E}\sup_{r\in[0,t]}\left|\int_{0}^{r}\|\Lambda^{\sigma}V_{n}\|^{p-2}\sum_{k}\langle P_{n}\mathcal{P}B_{k}V_{n},\Lambda^{2\sigma}V_{n}\rangle dW^{k}\right|\\ &\leq C_{p}\mathbb{E}\left(\int_{0}^{t}\|V_{n}\|_{\sigma}^{2(p-2)}\sum_{k}\langle B_{k}V_{n},\Lambda^{2\sigma}V_{n}\rangle^{2}dr\right)^{\frac{1}{2}}\\ &=C_{p}\mathbb{E}\left(\int_{0}^{t}\|V_{n}\|_{\sigma}^{2(p-2)}\sum_{k}\langle[\Lambda^{\sigma},B_{k}]V_{n},\Lambda^{\sigma}V_{n}\rangle^{2}dr\right)^{\frac{1}{2}}\\ &\leq C_{b}\mathbb{E}\left(\int_{0}^{t}\|V_{n}\|_{\sigma}^{2p}dr\right)^{\frac{1}{2}}\\ &\leq\frac{1}{4}\mathbb{E}\sup_{r\in[0,t]}\|V_{n}\|_{\sigma}^{p}+4C_{b}^{2}\mathbb{E}\int_{0}^{t}\|V_{n}(r)\|_{\sigma}^{p}dr.\end{split}

Note that the constant CbC_{b} above depends on bb through b2(,Hσ+2)\|b\|_{\ell^{2}(\mathbb{N},H^{\sigma+2})}.

To estimate I0I_{0}, we first use Lemma A.1 and the cutoff function θ\theta to obtain

p0tθρ2Pn𝒫Q(Vn,Vn),Λ2σVnΛσVnp2𝑑r=p0tθρ2Pn𝒫Λσ12Q(Vn,Vn),Λσ+12VnΛσVnp2𝑑rpCσ0tθρ2VnW1,Λσ+12Vn2ΛσVnp2𝑑rpCσρ0tΛσ+12Vn2ΛσVnp2𝑑r.\displaystyle\begin{split}&-p\int_{0}^{t}\theta_{\rho}^{2}\langle P_{n}\mathcal{P}Q(V_{n},V_{n}),\Lambda^{2\sigma}V_{n}\rangle\|\Lambda^{\sigma}V_{n}\|^{p-2}dr\\ =&-p\int_{0}^{t}\theta_{\rho}^{2}\langle P_{n}\mathcal{P}\Lambda^{\sigma-\frac{1}{2}}Q(V_{n},V_{n}),\Lambda^{\sigma+\frac{1}{2}}V_{n}\rangle\|\Lambda^{\sigma}V_{n}\|^{p-2}dr\\ \leq&pC_{\sigma}\int_{0}^{t}\theta_{\rho}^{2}\|V_{n}\|_{W^{1,\infty}}\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr\\ \leq&pC_{\sigma}\rho\int_{0}^{t}\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr.\end{split} (3.10)

Besides, we have Pn𝒫F(Vn),Λ2σVn=0\langle P_{n}\mathcal{P}F(V_{n}),\Lambda^{2\sigma}V_{n}\rangle=0 since F(Vn)F(V_{n}) is orthogonal to VnV_{n}. Therefore, when s=1s=1 one has

0tI0𝑑rp0t(1Cσρ)Λσ+12Vn2ΛσVnp2𝑑r.\displaystyle\begin{split}\int_{0}^{t}I_{0}dr\leq-p\int_{0}^{t}(1-C_{\sigma}\rho)\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr.\end{split}

For 1<s<21<s<2, applying interpolation and Young’s inequalities, we deduce

Λσ+12Vn2εΛσVn2+CεΛσ+s2Vn2.\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}\leq\varepsilon\|\Lambda^{\sigma}V_{n}\|^{2}+C_{\varepsilon}\|\Lambda^{\sigma+\frac{s}{2}}V_{n}\|^{2}.

Then we can bound (3.10) to obtain

0tI0𝑑rpCσρ0tΛσ+12Vn2ΛσVnp2𝑑rp0tΛσ+s2Vn2ΛσVnp2𝑑rp20tΛσ+s2Vn2ΛσVnp2𝑑r+Cp,σ,ρ0tVnσp𝑑r.\displaystyle\begin{split}\int_{0}^{t}I_{0}dr&\leq pC_{\sigma}\rho\int_{0}^{t}\|\Lambda^{\sigma+\frac{1}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr-p\int_{0}^{t}\|\Lambda^{\sigma+\frac{s}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr\\ &\leq-\frac{p}{2}\int_{0}^{t}\|\Lambda^{\sigma+\frac{s}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr+C_{p,\sigma,\rho}\int_{0}^{t}\|V_{n}\|_{\sigma}^{p}dr.\end{split}

Summarizing the above estimates for I0I_{0} to I3I_{3}, we obtain for any t[0,T]t\in[0,T]:

𝔼supr[0,t]ΛσVn(r)p+𝔼0tΛσ+s2Vn2ΛσVnp2𝑑rC(𝔼ΛσVn(0)p+𝔼0tVnσp𝑑r).\displaystyle\mathbb{E}\sup_{r\in[0,t]}\|\Lambda^{\sigma}V_{n}(r)\|^{p}+\mathbb{E}\int_{0}^{t}\|\Lambda^{\sigma+\frac{s}{2}}V_{n}\|^{2}\|\Lambda^{\sigma}V_{n}\|^{p-2}dr\leq C\left(\mathbb{E}\|\Lambda^{\sigma}V_{n}(0)\|^{p}+\mathbb{E}\int_{0}^{t}\|V_{n}\|_{\sigma}^{p}dr\right).

Since VnV_{n} has zero mean, this leads to

𝔼supr[0,t]Vn(r)σp+𝔼0tVnσ+s22Vnσp2𝑑rC(𝔼Vn(0)σp+𝔼0tVnσp𝑑r).\displaystyle\mathbb{E}\sup_{r\in[0,t]}\|V_{n}(r)\|_{\sigma}^{p}+\mathbb{E}\int_{0}^{t}\|V_{n}\|_{\sigma+\frac{s}{2}}^{2}\|V_{n}\|_{\sigma}^{p-2}dr\leq C\left(\mathbb{E}\|V_{n}(0)\|_{\sigma}^{p}+\mathbb{E}\int_{0}^{t}\|V_{n}\|_{\sigma}^{p}dr\right).

Applying Grönwall’s inequality, we conclude

𝔼[supr[0,T]Vnσp+0TVnσ+s22Vnσp2𝑑r]C(1+𝔼V0σp).\displaystyle\begin{split}\mathbb{E}\left[\sup_{r\in[0,T]}\|V_{n}\|_{\sigma}^{p}+\int_{0}^{T}\|V_{n}\|_{\sigma+\frac{s}{2}}^{2}\|V_{n}\|_{\sigma}^{p-2}dr\right]\leq C(1+\mathbb{E}\|V_{0}\|_{\sigma}^{p}).\end{split}

This proves (1).

Next, by the Burkholder-Davis-Gundy inequality, for α[0,12)\alpha\in[0,\frac{1}{2}) and p2p\geq 2, one has

𝔼|0kPn𝒫BkVndWrk|Wα,p(0,T;σ1)pCp𝔼0T(kBkVnσ12)p2𝑑r𝔼Cp0TVnσp(kbkσ12)p2𝑑rCT,b,p𝔼supr[0,T]Vnσp.\displaystyle\begin{split}\mathbb{E}\left|\int_{0}^{\cdot}\sum_{k}P_{n}\mathcal{P}B_{k}V_{n}dW_{r}^{k}\right|_{W^{\alpha,p}(0,T;\mathbb{H}^{\sigma-1})}^{p}&\leq C_{p}\mathbb{E}\int_{0}^{T}\left(\sum_{k}\|B_{k}V_{n}\|_{\sigma-1}^{2}\right)^{\frac{p}{2}}dr\\ &\leq\mathbb{E}C_{p}\int_{0}^{T}\|V_{n}\|_{\sigma}^{p}\left(\sum_{k}\|b_{k}\|_{\sigma-1}^{2}\right)^{\frac{p}{2}}dr\\ &\leq C_{T,b,p}\mathbb{E}\sup_{r\in[0,T]}\|V_{n}\|_{\sigma}^{p}.\end{split}

Combining this result with (1), we conclude (2).

Finally, in view of (3.3), we have

Vn(t)0tkPn𝒫BkVndWrk=PnV00t[θρ2Pn𝒫Q(Vn,Vn)+Pn𝒫F(Vn)+ΛsVn]𝑑r+120tkPn(𝒫Bk)2Vndr.\displaystyle\begin{split}V_{n}(t)&-\int_{0}^{t}\sum_{k}P_{n}\mathcal{P}B_{k}V_{n}dW_{r}^{k}\\ &=P_{n}V_{0}-\int_{0}^{t}\big{[}\theta_{\rho}^{2}P_{n}\mathcal{P}Q(V_{n},V_{n})+P_{n}\mathcal{P}F(V_{n})+\Lambda^{s}V_{n}\big{]}dr+\frac{1}{2}\int_{0}^{t}\sum_{k}P_{n}(\mathcal{P}B_{k})^{2}V_{n}dr.\end{split}

Since σ>3\sigma>3 and s[1,2)s\in[1,2), one has σ2+s2>32\sigma-2+\frac{s}{2}>\frac{3}{2}. Hence

θρ2Pn𝒫Q(Vn,Vn)+Pn𝒫F(Vn)+ΛsVnσ2+s2C(1+Vnσ+s22),\displaystyle\|\theta_{\rho}^{2}P_{n}\mathcal{P}Q(V_{n},V_{n})+P_{n}\mathcal{P}F(V_{n})+\Lambda^{s}V_{n}\|_{\sigma-2+\frac{s}{2}}\leq C(1+\|V_{n}\|_{\sigma+\frac{s}{2}}^{2}),

and

Pn(𝒫Bk)2Vnσ2+s2Cbkσ2+s2𝒫BkVnσ1+s2Cbkσ1+s22Vnσ+s2.\displaystyle\|P_{n}(\mathcal{P}B_{k})^{2}V_{n}\|_{\sigma-2+\frac{s}{2}}\leq C\|b_{k}\|_{\sigma-2+\frac{s}{2}}\|\mathcal{P}B_{k}V_{n}\|_{\sigma-1+\frac{s}{2}}\leq C\|b_{k}\|_{\sigma-1+\frac{s}{2}}^{2}\|V_{n}\|_{\sigma+\frac{s}{2}}.

We thus obtain

𝔼Vn(t)0tkPn𝒫BkVndWrkW1,2(0,T;σ2+s2)C𝔼[1+V0σ2+supr[0,T]Vnσ2+0TVnσ+s22𝑑r],\displaystyle\begin{split}\mathbb{E}\Big{\|}V_{n}(t)&-\int_{0}^{t}\sum_{k}P_{n}\mathcal{P}B_{k}V_{n}dW_{r}^{k}\Big{\|}_{W^{1,2}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}})}\\ &\leq C\mathbb{E}\left[1+\|V_{0}\|_{\sigma}^{2}+\sup_{r\in[0,T]}\|V_{n}\|_{\sigma}^{2}+\int_{0}^{T}\|V_{n}\|_{\sigma+\frac{s}{2}}^{2}dr\right],\end{split}

which implies (3) by utilizing (1) with p=2p=2.

3.2. Compactness and martingale solutions

In this subsection, we establish the existence of martingale solutions to system (3.2), leveraging the energy estimates obtained in Proposition 3.1. Before proceeding with the proof, we outline the necessary settings.

Recall that we have fixed a stochastic basis 𝒮=(Ω,,𝔽,)\mathcal{S}=(\Omega,\mathcal{F},\mathbb{F},\mathbb{P}). Define the path space

𝒳=σ×L2(0,T;σ)C([0,T];σ2)×C([0,T];𝒰).\mathcal{X}=\mathbb{H}^{\sigma}\times L^{2}\left(0,T;\mathbb{H}^{\sigma}\right)\cap C\left(\left[0,T\right];\mathbb{H}^{\sigma-2}\right)\times C\left(\left[0,T\right];\mathscr{U}\right).

Given any random initial data V0L2(Ω,σ)V_{0}\in L^{2}(\Omega,\mathbb{H}^{\sigma}), we let μ0n\mu_{0}^{n} be the law of PnV0P_{n}V_{0}, and μVn\mu_{V}^{n} be the law of the corresponding solution VnV_{n} to the approximating system (3.3) with initial data Vn(0)=PnV0V_{n}(0)=P_{n}V_{0}, and also μ𝕎\mu_{\mathbb{W}} the law of the cylindrical Wiener process on C([0,T];𝒰)C\left(\left[0,T\right];\mathscr{U}\right). Define μn=μ0nμVnμ𝕎\mu^{n}=\mu_{0}^{n}\otimes\mu_{V}^{n}\otimes\mu_{\mathbb{W}} as their joint law in the path space 𝒳\mathcal{X}.

Proposition 3.2.

Assume the same conditions as in Proposition 3.1 and let p>2p>2. Then there exists a stochastic basis (Ω~,~,𝔽~,~)(\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{F}},\widetilde{\mathbb{P}}) and an 𝒳\mathcal{X}-valued random variable (V~0,V~,𝕎~)(\widetilde{V}_{0},\widetilde{V},\widetilde{\mathbb{W}}) over Ω~\widetilde{\Omega} such that V~\widetilde{V}, adapted to the filtration 𝔽~\widetilde{\mathbb{F}}, is a solution to (3.2) on [0,T][0,T], with driving noise 𝕎~\widetilde{\mathbb{W}} and initial data V~0\widetilde{V}_{0} having the same distribution as V0V_{0}. In short, (V~0,V~,𝕎~)(\widetilde{V}_{0},\widetilde{V},\widetilde{\mathbb{W}}) over the stochastic basis (Ω~,~,𝔽~,~)(\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{F}},\widetilde{\mathbb{P}}) is a martingale solution to (3.2) on [0,T][0,T]. Moreover, V~\widetilde{V} satisfies

V~L2(Ω~;C([0,T];σ)L2(0,T;σ+s2)).\widetilde{V}\in L^{2}\left(\widetilde{\Omega};C\left([0,T];\mathbb{H}^{\sigma}\right)\cap L^{2}\left(0,T;\mathbb{H}^{\sigma+\frac{s}{2}}\right)\right). (3.11)
Proof.

The proof is divided into two main steps. First, we establish the compactness of μn\mu^{n} to obtain a limit as a random variable with values in 𝒳\mathcal{X} using the Skorokhod theorem. Then, we show that the limit gives a martingale solution to (3.2).

Step 1: Compactness. By Lemma A.2, the embedding

L2(0,T;σ+s2)W14,2(0,T;σ1)L2(0,T;σ)L^{2}(0,T;\mathbb{H}^{\sigma+\frac{s}{2}})\cap W^{\frac{1}{4},2}(0,T;\mathbb{H}^{\sigma-1})\subset L^{2}(0,T;\mathbb{H}^{\sigma})

is compact. In addition, by choosing α(0,1/2)\alpha\in(0,1/2) such that αp>1\alpha p>1, both spaces W1,2(0,T;σ2+s2)W^{1,2}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}}) and Wα,p(0,T;σ2+s2)W^{\alpha,p}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}}) are compactly embedded in C([0,T];σ2)C([0,T];\mathbb{H}^{\sigma-2}). These compact embeddings imply the tightness of the sequence {μn}n1\{\mu^{n}\}_{n\geq 1} over 𝒳\mathcal{X}. Indeed, if we let

BR1={VL2(0,T;σ+s2)W14,2(0,T;σ1):VL2(0,T;σ+s2)2+VW14,2(0,T;σ1)2R2},B_{R}^{1}=\left\{V\in L^{2}(0,T;\mathbb{H}^{\sigma+\frac{s}{2}})\cap W^{\frac{1}{4},2}(0,T;\mathbb{H}^{\sigma-1}):\|V\|_{L^{2}(0,T;\mathbb{H}^{\sigma+\frac{s}{2}})}^{2}+\|V\|_{W^{\frac{1}{4},2}(0,T;\mathbb{H}^{\sigma-1})}^{2}\leq R^{2}\right\},

and BR2=BR2,1+BR2,2B_{R}^{2}=B_{R}^{2,1}+B_{R}^{2,2}, where BR2,1B_{R}^{2,1} and BR2,2B_{R}^{2,2} are the closed balls in W1,2(0,T;σ2+s2)W^{1,2}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}}) and Wα,p(0,T;σ2+s2)W^{\alpha,p}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}}) centered at 0 with radius RR, then BR1BR2B_{R}^{1}\cap B_{R}^{2} is precompact in L2(0,T;σ)C([0,T];σ2)L^{2}(0,T;\mathbb{H}^{\sigma})\cap C([0,T];\mathbb{H}^{\sigma-2}). By Markov’s inequality and Proposition 3.1, one has

μVn((BR1)c)\displaystyle\mu_{V}^{n}\left((B_{R}^{1})^{c}\right) (VnL2(0,T;σ+s2)R2)+(VnW14,2(0,T;σ1)R2)\displaystyle\leq\mathbb{P}\left(\|V_{n}\|_{L^{2}(0,T;\mathbb{H}^{\sigma+\frac{s}{2}})}\geq\frac{R}{\sqrt{2}}\right)+\mathbb{P}\left(\|V_{n}\|_{W^{\frac{1}{4},2}(0,T;\mathbb{H}^{\sigma-1})}\geq\frac{R}{\sqrt{2}}\right)
C(1+𝔼V0σ2)R,\displaystyle\leq\frac{C(1+\mathbb{E}\|V_{0}\|_{\sigma}^{2})}{R},

and

μVn((BR2)c)\displaystyle\mu_{V}^{n}\left((B_{R}^{2})^{c}\right) (Vn0kPn𝒫BkVndWrkW1,2(0,T;σ2+s2)R)\displaystyle\leq\mathbb{P}\left(\left\|V_{n}-\int_{0}^{\cdot}\sum_{k}P_{n}\mathcal{P}B_{k}V_{n}dW_{r}^{k}\right\|_{W^{1,2}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}})}\geq R\right)
+(0kPn𝒫BkVndWrkWα,p(0,T;σ2+s2)pRp)\displaystyle\qquad+\mathbb{P}\left(\left\|\int_{0}^{\cdot}\sum_{k}P_{n}\mathcal{P}B_{k}V_{n}dW_{r}^{k}\right\|_{W^{\alpha,p}(0,T;\mathbb{H}^{\sigma-2+\frac{s}{2}})}^{p}\geq R^{p}\right)
C(1+𝔼V0σp)R.\displaystyle\leq\frac{C(1+\mathbb{E}\|V_{0}\|_{\sigma}^{p})}{R}.

Thus the sequence {μn}n1\{\mu^{n}\}_{n\geq 1} is tight over 𝒳\mathcal{X}. By Prokhorov’s theorem, the sequence is precompact. Applying the Skorokhod theorem yields the existence of a probability space (Ω~,~,~)(\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{P}}), a subsequence njn_{j}\rightarrow\infty as jj\rightarrow\infty, and a sequence of 𝒳\mathcal{X}-valued random variables (V~nj0,V~nj,𝕎~nj)\left(\widetilde{V}^{0}_{n_{j}},\widetilde{V}_{n_{j}},\widetilde{\mathbb{W}}_{n_{j}}\right) such that:

  • The sequence converges almost surely under ~\widetilde{\mathbb{P}} to (V~0,V~,𝕎~)\left(\widetilde{V}_{0},\widetilde{V},\widetilde{\mathbb{W}}\right) in 𝒳\mathcal{X};

  • Each triple (V~nj0,V~nj,𝕎~nj)\left(\widetilde{V}^{0}_{n_{j}},\widetilde{V}_{n_{j}},\widetilde{\mathbb{W}}_{n_{j}}\right) is a martingale solution to (3.3) for n=njn=n_{j} with initial data V~nj0\widetilde{V}^{0}_{n_{j}}.

  • The law of V~0\widetilde{V}^{0} coincides with that of V0V_{0}.

Moreover, the sequence V~nj\widetilde{V}_{n_{j}} satisfies the same energy estimate under the new probability space as in Proposition 3.1.

Step 2: Identify the limit as a martingale solution. We now show that the limit (V~0,V~,𝕎~)\left(\widetilde{V}_{0},\widetilde{V},\widetilde{\mathbb{W}}\right) is the desired martingale solution over the stochastic basis (Ω~,~,𝔽~,~)(\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{F}},\widetilde{\mathbb{P}}) with 𝔽~\widetilde{\mathbb{F}} being the filtration generated by V~,𝕎~\widetilde{V},\widetilde{\mathbb{W}}. Firstly, by the convergence V~njV~\widetilde{V}_{n_{j}}\to\widetilde{V} in 𝒳\mathcal{X}, Proposition 3.1 and the Banach-Alaoglu theorem, we infer:

V~L2(Ω~;L2(0,T;σ+s2)L(0,T;σ)).\displaystyle\widetilde{V}\in L^{2}\left(\widetilde{\Omega};L^{2}\left(0,T;\mathbb{H}^{\sigma+\frac{s}{2}}\right)\cap L^{\infty}\left(0,T;\mathbb{H}^{\sigma}\right)\right). (3.12)

Moreover, the Vitali convergence theorem implies:

V~njV~ in L2(Ω~;L2(0,T;σ)).\displaystyle\widetilde{V}_{n_{j}}\rightarrow\widetilde{V}\text{ in }L^{2}\left(\widetilde{\Omega};L^{2}\left(0,T;\mathbb{H}^{\sigma}\right)\right). (3.13)

Thus there exists a subsequence, still denote by V~nj\widetilde{V}_{n_{j}}, such that:

V~njV~ in L2(0,T;σ) for a.a. ωΩ~.\displaystyle\widetilde{V}_{n_{j}}\rightarrow\widetilde{V}\text{ in }L^{2}\left(0,T;\mathbb{H}^{\sigma}\right)\text{ for a.a. }\omega\in\widetilde{\Omega}\text{. } (3.14)

Since each pair (V~nj,𝕎~nj)(\widetilde{V}_{n_{j}},\widetilde{\mathbb{W}}_{n_{j}}) is a martingale solution to system (3.3) with n=njn=n_{j}, we have

V~nj(t),ϕ+0tθρ2Pnj𝒫Q(V~nj,V~nj)+Pnj𝒫F(V~nj)+ΛsV~nj,ϕ𝑑r=V~nj(0),ϕ+12k0tPnj(𝒫Bk)2V~nj,ϕ𝑑r+k0tPnj𝒫BkV~nj,ϕ𝑑W~rk,nj,\displaystyle\begin{split}&\left\langle\widetilde{V}_{n_{j}}(t),\phi\right\rangle+\int_{0}^{t}\left\langle\theta_{\rho}^{2}P_{n_{j}}\mathcal{P}Q(\widetilde{V}_{n_{j}},\widetilde{V}_{n_{j}})+P_{n_{j}}\mathcal{P}F(\widetilde{V}_{n_{j}})+\Lambda^{s}\widetilde{V}_{n_{j}},\phi\right\rangle dr\\ &\qquad=\left\langle\widetilde{V}_{n_{j}}(0),\phi\right\rangle+\frac{1}{2}\sum_{k}\int_{0}^{t}\left\langle P_{n_{j}}(\mathcal{P}B_{k})^{2}\widetilde{V}_{n_{j}},\phi\right\rangle dr\\ &\qquad\qquad+\sum_{k}\int_{0}^{t}\left\langle P_{n_{j}}\mathcal{P}B_{k}\widetilde{V}_{n_{j}},\phi\right\rangle d\widetilde{W}_{r}^{k,n_{j}},\end{split} (3.15)

for any ϕ\phi\in\mathbb{H} and t[0,T]t\in[0,T]. Here and below, W~k,nj\widetilde{W}^{k,n_{j}} is the kk-th component of the Wiener process 𝕎~nj\widetilde{\mathbb{W}}_{n_{j}}. We now prove that (V~,𝕎~)(\widetilde{V},\widetilde{\mathbb{W}}) is a martingale solution to (3.2) by passing the limit jj\to\infty in (3.15).

Linear Terms. The convergence of the linear terms follows straightforwardly from (3.14). Here, we only address the linear term corresponding to the Itô-Stratonovich corrector. Observe that

|12k0tPnj(𝒫Bk)2V~nj,ϕ𝑑r12k0t(𝒫Bk)2V~,ϕ𝑑r|\displaystyle\left|\frac{1}{2}\sum_{k}\int_{0}^{t}\left\langle P_{n_{j}}(\mathcal{P}B_{k})^{2}\widetilde{V}_{n_{j}},\phi\right\rangle dr-\frac{1}{2}\sum_{k}\int_{0}^{t}\left\langle(\mathcal{P}B_{k})^{2}\widetilde{V},\phi\right\rangle dr\right|
12k0t|(𝒫Bk)2(V~njV~),Pnjϕ|𝑑r+12k0t|(𝒫Bk)2V~,Pnjϕϕ|𝑑r\displaystyle\leq\frac{1}{2}\sum_{k}\int_{0}^{t}\left|\left\langle(\mathcal{P}B_{k})^{2}(\widetilde{V}_{n_{j}}-\widetilde{V}),P_{n_{j}}\phi\right\rangle\right|dr+\frac{1}{2}\sum_{k}\int_{0}^{t}\left|\left\langle(\mathcal{P}B_{k})^{2}\widetilde{V},P_{n_{j}}\phi-\phi\right\rangle\right|dr
Cb2(,W1,)2(ϕ0TV~njV~2𝑑r+Pnjϕϕ0TV~2𝑑r)0,\displaystyle\leq C\|b\|_{\ell^{2}(\mathbb{N},W^{1,\infty})}^{2}\left(\|\phi\|\int_{0}^{T}\|\widetilde{V}_{n_{j}}-\widetilde{V}\|_{2}dr+\|P_{n_{j}}\phi-\phi\|\int_{0}^{T}\|\widetilde{V}\|_{2}dr\right)\to 0,

almost surely by (3.12) and (3.14).

Nonlinear Terms. Denote θρ2(V)=θρ2(VW1,)\theta_{\rho}^{2}(V)=\theta_{\rho}^{2}(\|V\|_{W^{1,\infty}}) for convenience. For the nonlinear term, we have

|0tθρ2(V~nj)Pnj𝒫Q(V~nj,V~nj),ϕ𝑑r0tθρ2(V~)𝒫Q(V~,V~),ϕ𝑑r|0t|θρ2(V~nj)𝒫Q(V~nj,V~nj),Pnjϕϕ|𝑑r+0t|(θρ2(V~nj)θρ2(V~))𝒫Q(V~nj,V~nj),ϕ|𝑑r+0t|θρ2(V~)(𝒫Q(V~nj,V~nj)𝒫Q(V~,V~)),ϕ|𝑑r=N1+N2+N3.\displaystyle\begin{split}&\left|\int_{0}^{t}\left\langle\theta_{\rho}^{2}(\widetilde{V}_{n_{j}})P_{n_{j}}\mathcal{P}Q(\widetilde{V}_{n_{j}},\widetilde{V}_{n_{j}}),\phi\right\rangle dr-\int_{0}^{t}\left\langle\theta_{\rho}^{2}(\widetilde{V})\mathcal{P}Q(\widetilde{V},\widetilde{V}),\phi\right\rangle dr\right|\\ &\leq\int_{0}^{t}\left|\left\langle\theta_{\rho}^{2}(\widetilde{V}_{n_{j}})\mathcal{P}Q(\widetilde{V}_{n_{j}},\widetilde{V}_{n_{j}}),P_{n_{j}}\phi-\phi\right\rangle\right|dr+\int_{0}^{t}\left|\left\langle\left(\theta_{\rho}^{2}(\widetilde{V}_{n_{j}})-\theta_{\rho}^{2}(\widetilde{V})\right)\mathcal{P}Q(\widetilde{V}_{n_{j}},\widetilde{V}_{n_{j}}),\phi\right\rangle\right|dr\\ &\qquad+\int_{0}^{t}\left|\left\langle\theta_{\rho}^{2}(\widetilde{V})\left(\mathcal{P}Q(\widetilde{V}_{n_{j}},\widetilde{V}_{n_{j}})-\mathcal{P}Q(\widetilde{V},\widetilde{V})\right),\phi\right\rangle\right|dr=N_{1}+N_{2}+N_{3}.\end{split}

By the Cauchy-Schwarz inequality, Lemma A.1 and Sobolev embedding we have

N1CPnjϕϕ0tQ(V~nj,V~nj)𝑑rCPnjϕϕ0tV~njσ2𝑑r0,\displaystyle N_{1}\leq C\|P_{n_{j}}\phi-\phi\|\int_{0}^{t}\|Q(\widetilde{V}_{n_{j}},\widetilde{V}_{n_{j}})\|dr\leq C\|P_{n_{j}}\phi-\phi\|\int_{0}^{t}\|\widetilde{V}_{n_{j}}\|_{\sigma}^{2}dr\to 0,

almost surely. Similarly, one has

N3\displaystyle N_{3} Cϕ0tV~njV~σ(V~njσ+V~σ)𝑑r\displaystyle\leq C\|\phi\|\int_{0}^{t}\|\widetilde{V}_{n_{j}}-\widetilde{V}\|_{\sigma}(\|\widetilde{V}_{n_{j}}\|_{\sigma}+\|\widetilde{V}\|_{\sigma})dr
Cϕ(0TV~njV~σ2𝑑r)1/2(0TV~njσ2+V~σ2dr)1/20,\displaystyle\leq C\|\phi\|\left(\int_{0}^{T}\|\widetilde{V}_{n_{j}}-\widetilde{V}\|_{\sigma}^{2}dr\right)^{1/2}\left(\int_{0}^{T}\|\widetilde{V}_{n_{j}}\|_{\sigma}^{2}+\|\widetilde{V}\|_{\sigma}^{2}dr\right)^{1/2}\to 0,

almost surely. By Hölder’s inequality, the Lipschitz continuity of the cutoff function, Lemma A.1, and Sobolev embeddings, we have

N2\displaystyle N_{2} C(0TV~njV~σ2𝑑r)12(0TQ(V~nj,V~nj)2ϕ2𝑑r)12\displaystyle\leq C\left(\int_{0}^{T}\|\widetilde{V}_{n_{j}}-\widetilde{V}\|_{\sigma}^{2}dr\right)^{\frac{1}{2}}\left(\int_{0}^{T}\|Q(\widetilde{V}_{n_{j}},\widetilde{V}_{n_{j}})\|^{2}\|\phi\|^{2}dr\right)^{\frac{1}{2}}
Cϕ(0TV~njV~σ2𝑑r)12(0TV~njW1,2V~nj12𝑑r)12\displaystyle\leq C\|\phi\|\left(\int_{0}^{T}\|\widetilde{V}_{n_{j}}-\widetilde{V}\|_{\sigma}^{2}dr\right)^{\frac{1}{2}}\left(\int_{0}^{T}\|\widetilde{V}_{n_{j}}\|_{W^{1,\infty}}^{2}\|\widetilde{V}_{n_{j}}\|_{1}^{2}dr\right)^{\frac{1}{2}}
Cϕ(0TV~njV~σ2𝑑r)12(0TV~njσ2𝑑r)12supr[0,T]V~njσ20,\displaystyle\leq C\|\phi\|\left(\int_{0}^{T}\|\widetilde{V}_{n_{j}}-\widetilde{V}\|_{\sigma}^{2}dr\right)^{\frac{1}{2}}\left(\int_{0}^{T}\|\widetilde{V}_{n_{j}}\|_{\sigma}^{2}dr\right)^{\frac{1}{2}}\sup_{r\in[0,T]}\|\widetilde{V}_{n_{j}}\|_{\sigma-2}\to 0,

almost surely by (3.13) and the convergence of V~njV~\widetilde{V}_{n_{j}}\to\widetilde{V} in 𝒳\mathcal{X} (ensuring that supr[0,T]V~njσ2\sup_{r\in[0,T]}\|\widetilde{V}_{n_{j}}\|_{\sigma-2} remains finite almost surely in the limit jj\to\infty).

Stochastic Terms. Lastly, we look at the stochastic integral term. Since σ>3\sigma>3 and V~njV~\widetilde{V}_{n_{j}}\to\widetilde{V} in C([0,T];σ2)C([0,T];\mathbb{H}^{\sigma-2}), we have

supr[0,T]|kPnj𝒫BkV~nj,ϕk𝒫BkV~,ϕ|\displaystyle\sup_{r\in[0,T]}\left|\sum_{k}\left\langle P_{n_{j}}\mathcal{P}B_{k}\widetilde{V}_{n_{j}},\phi\right\rangle-\sum_{k}\left\langle\mathcal{P}B_{k}\widetilde{V},\phi\right\rangle\right|
Csupr[0,T]k(|𝒫BkV~nj,Pnjϕϕ|+|𝒫Bk(V~njV~),ϕ|)\displaystyle\leq C\sup_{r\in[0,T]}\sum_{k}\left(\left|\left\langle\mathcal{P}B_{k}\widetilde{V}_{n_{j}},P_{n_{j}}\phi-\phi\right\rangle\right|+\left|\left\langle\mathcal{P}B_{k}\left(\widetilde{V}_{n_{j}}-\widetilde{V}\right),\phi\right\rangle\right|\right)
Cb2(,L)(Pnjϕϕsupr[0,T]V~nj(r)σ2+ϕsupr[0,T]V~njV~σ2)0,\displaystyle\leq C\|b\|_{\ell^{2}(\mathbb{N},L^{\infty})}\left(\|P_{n_{j}}\phi-\phi\|\sup_{r\in[0,T]}\|\widetilde{V}_{n_{j}}(r)\|_{\sigma-2}+\|\phi\|\sup_{r\in[0,T]}\|\widetilde{V}_{n_{j}}-\widetilde{V}\|_{\sigma-2}\right)\to 0,

almost surely. This fact, together with the almost sure convergence of 𝕎~nj𝕎~\widetilde{\mathbb{W}}_{n_{j}}\to\widetilde{\mathbb{W}} in C([0,T];𝒰)C\left(\left[0,T\right];\mathscr{U}\right) and Theorem 4.2 in [45] imply the following convergence

k0tPnj𝒫BkV~nj,ϕ𝑑W~rk,njk0t𝒫BkV~,ϕ𝑑W~rk\sum_{k}\int_{0}^{t}\left\langle P_{n_{j}}\mathcal{P}B_{k}\widetilde{V}_{n_{j}},\phi\right\rangle d\widetilde{W}_{r}^{k,n_{j}}\to\sum_{k}\int_{0}^{t}\left\langle\mathcal{P}B_{k}\widetilde{V},\phi\right\rangle d\widetilde{W}_{r}^{k}

in probability in C([0,T];)C([0,T];\mathbb{R}). Passing to a subsequence if necessary, we obtain the above convergence in C([0,T];)C([0,T];\mathbb{R}) almost surely.

We have just proved that for each ϕ\phi\in\mathbb{H} and t[0,T]t\in[0,T], the triple (V~0,V~,𝕎~)(\widetilde{V}_{0},\widetilde{V},\widetilde{\mathbb{W}}) satisfies

V~(t),ϕ+0tθρ2𝒫Q(V~,V~)+𝒫F(V~)+ΛsV~,ϕ𝑑r=V~(0),ϕ+12k0t(𝒫Bk)2V~,ϕ𝑑r+k0t𝒫BkV~,ϕ𝑑W~rk,\displaystyle\begin{split}\left\langle\widetilde{V}(t),\phi\right\rangle+\int_{0}^{t}\left\langle\theta_{\rho}^{2}\mathcal{P}Q(\widetilde{V},\widetilde{V})+\mathcal{P}F(\widetilde{V})+\Lambda^{s}\widetilde{V},\phi\right\rangle dr&=\left\langle\widetilde{V}(0),\phi\right\rangle+\frac{1}{2}\sum_{k}\int_{0}^{t}\left\langle(\mathcal{P}B_{k})^{2}\widetilde{V},\phi\right\rangle dr\\ &\qquad\qquad+\sum_{k}\int_{0}^{t}\left\langle\mathcal{P}B_{k}\widetilde{V},\phi\right\rangle d\widetilde{W}_{r}^{k},\end{split}

almost surely over (Ω~,~,~)(\widetilde{\Omega},\widetilde{\mathcal{F}},\widetilde{\mathbb{P}}). This result shows that tV~(t),ϕt\to\left\langle\widetilde{V}(t),\phi\right\rangle is continuous for each ϕ\phi\in\mathbb{H}. It remains to prove that V~L2(Ω~;C([0,T],σ))\widetilde{V}\in L^{2}(\widetilde{\Omega};C([0,T],\mathbb{H}^{\sigma})) so that (V~0,V~,𝕎~)\left(\widetilde{V}_{0},\widetilde{V},\widetilde{\mathbb{W}}\right) is the desired martingale solution.

First, we note that each solution sample path tV~(t)t\to\widetilde{V}(t) is weakly continuous in σ\mathbb{H}^{\sigma}. Indeed, since V~L2(Ω~;L([0,T],σ))\widetilde{V}\in L^{2}(\widetilde{\Omega};L^{\infty}([0,T],\mathbb{H}^{\sigma})), its L([0,T],σ)L^{\infty}([0,T],\mathbb{H}^{\sigma}) norm is almost surely finite. For any ϕσ\phi\in\mathbb{H}^{-\sigma}, by the density of C(𝕋3)C^{\infty}(\mathbb{T}^{3})\cap\mathbb{H} in σ\mathbb{H}^{-\sigma}, there exists a sequence ϕnC(𝕋3)\phi_{n}\in C^{\infty}(\mathbb{T}^{3})\cap\mathbb{H} such that ϕnϕ\phi_{n}\to\phi in σ\mathbb{H}^{-\sigma}. Therefore,

supt[0,T]|V~(t),ϕnV~(t),ϕ|V~L([0,T],σ)ϕnϕσ0.\sup_{t\in[0,T]}\left|\left\langle\widetilde{V}(t),\phi_{n}\right\rangle-\left\langle\widetilde{V}(t),\phi\right\rangle\right|\leq\|\widetilde{V}\|_{L^{\infty}([0,T],\mathbb{H}^{\sigma})}\|\phi_{n}-\phi\|_{-\sigma}\to 0.

Consequently, tV~(t),ϕt\to\left\langle\widetilde{V}(t),\phi\right\rangle is continuous for each ϕσ\phi\in\mathbb{H}^{-\sigma}. Therefore, to prove V~L2(Ω~;C([0,T],σ))\widetilde{V}\in L^{2}(\widetilde{\Omega};C([0,T],\mathbb{H}^{\sigma})), we only need to show that tV~(t)σt\to\|\widetilde{V}(t)\|_{\sigma} is continuous almost surely. Direct application of Itô’s formula is not possible for the norm σ\|\cdot\|_{\sigma}, due to the limited regularity of the solution and the Itô-Stratonovich corrector. We instead use a mollification argument, adapted from [39].

Let ε>0\varepsilon>0, and define the mollification operator JεJ_{\varepsilon} using the standard mollifier ϕε\phi_{\varepsilon} on the periodic torus with respect to the spatial variable:

Jεf(x)=(ϕεf)(x).J_{\varepsilon}f(x)=(\phi_{\varepsilon}\ast f)(x).

The operator is self-adjoint on HH and commutes with 𝒫\mathcal{P} and Λ\Lambda as a Fourier multiplier. Applying JεJ_{\varepsilon} to the equation satisfied by V~\widetilde{V}, we obtain

JεV~(t)+0tJε(θρ2𝒫Q(V~,V~)+𝒫F(V~)+ΛsV~)𝑑r=JεV~0+12k0tJε(𝒫Bk)2V~𝑑r+k0tJε𝒫BkV~𝑑W~rk.\displaystyle\begin{split}J_{\varepsilon}\widetilde{V}(t)+\int_{0}^{t}J_{\varepsilon}\left(\theta_{\rho}^{2}\mathcal{P}Q(\widetilde{V},\widetilde{V})+\mathcal{P}F(\widetilde{V})+\Lambda^{s}\widetilde{V}\right)dr&=J_{\varepsilon}\widetilde{V}_{0}+\frac{1}{2}\sum_{k}\int_{0}^{t}J_{\varepsilon}(\mathcal{P}B_{k})^{2}\widetilde{V}dr\\ &\qquad\qquad+\sum_{k}\int_{0}^{t}J_{\varepsilon}\mathcal{P}B_{k}\widetilde{V}d\widetilde{W}_{r}^{k}.\end{split}

Applying Itô’s formula to ΛσJεV~(t)2\|\Lambda^{\sigma}J_{\varepsilon}\widetilde{V}(t)\|^{2} yields

ΛσJεV~(t)2=ΛσJεV~020t2Jε(θρ2𝒫Q(V~,V~)+𝒫F(V~)+ΛsV~),Λ2σJεV~𝑑r+k0t(Jε(𝒫Bk)2V~,Λ2σJεV~+ΛσJε𝒫BkV~2)𝑑r+0t2kJε𝒫BkV~,Λ2σJεV~dW~rk.\displaystyle\begin{split}\|\Lambda^{\sigma}J_{\varepsilon}\widetilde{V}(t)\|^{2}&=\|\Lambda^{\sigma}J_{\varepsilon}\widetilde{V}_{0}\|^{2}-\int_{0}^{t}2\left\langle J_{\varepsilon}\left(\theta_{\rho}^{2}\mathcal{P}Q(\widetilde{V},\widetilde{V})+\mathcal{P}F(\widetilde{V})+\Lambda^{s}\widetilde{V}\right),\Lambda^{2\sigma}J_{\varepsilon}\widetilde{V}\right\rangle dr\\ &+\sum_{k}\int_{0}^{t}\left(\left\langle J_{\varepsilon}(\mathcal{P}B_{k})^{2}\widetilde{V},\Lambda^{2\sigma}J_{\varepsilon}\widetilde{V}\right\rangle+\|\Lambda^{\sigma}J_{\varepsilon}\mathcal{P}B_{k}\widetilde{V}\|^{2}\right)dr+\int_{0}^{t}2\sum_{k}\langle J_{\varepsilon}\mathcal{P}B_{k}\widetilde{V},\Lambda^{2\sigma}J_{\varepsilon}\widetilde{V}\rangle d\widetilde{W}_{r}^{k}.\end{split}

Estimate similarly as in the proof of Proposition 3.1, we have

|Jε(θρ2𝒫Q(V~,V~)+𝒫F(V~)+ΛsV~),Λ2σJεV~|(Cp,σ,ρ+1)V~σ+s22.\displaystyle\left|\left\langle J_{\varepsilon}\left(\theta_{\rho}^{2}\mathcal{P}Q(\widetilde{V},\widetilde{V})+\mathcal{P}F(\widetilde{V})+\Lambda^{s}\widetilde{V}\right),\Lambda^{2\sigma}J_{\varepsilon}\widetilde{V}\right\rangle\right|\leq(C_{p,\sigma,\rho}+1)\|\widetilde{V}\|_{\sigma+\frac{s}{2}}^{2}.

Denote by Aσ=ΛσJεA^{\sigma}=\Lambda^{\sigma}J_{\varepsilon} and perform estimates as in (3.4) gives

|Jε(𝒫Bk)2V~,Λ2σJεV~+ΛσJε𝒫BkV~|\displaystyle\left|\left\langle J_{\varepsilon}(\mathcal{P}B_{k})^{2}\widetilde{V},\Lambda^{2\sigma}J_{\varepsilon}\widetilde{V}\right\rangle+\|\Lambda^{\sigma}J_{\varepsilon}\mathcal{P}B_{k}\widetilde{V}\|\right|
=|[Aσ,Bk][𝒫,Bk]V~,AσV~+Aσ[𝒫,Bk]V~,[Aσ,Bk]V~\displaystyle=\Big{|}\langle[A^{\sigma},B_{k}][\mathcal{P},B_{k}]\widetilde{V},A^{\sigma}\widetilde{V}\rangle+\langle A^{\sigma}[\mathcal{P},B_{k}]\widetilde{V},[A^{\sigma},B_{k}]\widetilde{V}\rangle
+[[Aσ,Bk],Bk]V~,AσV~+[Aσ,Bk]V~,[Aσ,Bk]V~|.\displaystyle\qquad+\langle[[A^{\sigma},B_{k}],B_{k}]\widetilde{V},A^{\sigma}\widetilde{V}\rangle+\langle[A^{\sigma},B_{k}]\widetilde{V},[A^{\sigma},B_{k}]\widetilde{V}\rangle\Big{|}.

Similar to (3.9), under the assumption on the noise coefficients we have

|k0t(Jε(𝒫Bk)2V~,Λ2σJεV~+ΛσJε𝒫BkV~2)𝑑r|0tV~σ+s22𝑑r+C0tV~σ2𝑑r.\displaystyle\left|\sum_{k}\int_{0}^{t}\left(\left\langle J_{\varepsilon}(\mathcal{P}B_{k})^{2}\widetilde{V},\Lambda^{2\sigma}J_{\varepsilon}\widetilde{V}\right\rangle+\|\Lambda^{\sigma}J_{\varepsilon}\mathcal{P}B_{k}\widetilde{V}\|^{2}\right)dr\right|\leq\int_{0}^{t}\|\widetilde{V}\|_{\sigma+\frac{s}{2}}^{2}dr+C\int_{0}^{t}\|\widetilde{V}\|_{\sigma}^{2}dr.

On the other hand, since JεJ_{\varepsilon} is self-adjoint and each bkb_{k} in Bk=bkB_{k}=b_{k}\cdot\nabla is divergence-free, the Burkholder-Davis-Gundy inequality, Minkowski’s inequality, Lemma A.1, and Sobolev’s inequality yield:

𝔼supt[0,T]|0t2k(Jε𝒫BkV~,Λ2σJεV~𝒫BkV~,Λ2σV~)dW~rk|\displaystyle\mathbb{E}\sup_{t\in[0,T]}\left|\int_{0}^{t}2\sum_{k}\left(\left\langle J_{\varepsilon}\mathcal{P}B_{k}\widetilde{V},\Lambda^{2\sigma}J_{\varepsilon}\widetilde{V}\right\rangle-\left\langle\mathcal{P}B_{k}\widetilde{V},\Lambda^{2\sigma}\widetilde{V}\right\rangle\right)d\widetilde{W}_{r}^{k}\right|
=2𝔼supt[0,T]|0tkΛσBkV~,ΛσJε2V~ΛσV~dW~rk|\displaystyle=2\mathbb{E}\sup_{t\in[0,T]}\left|\int_{0}^{t}\sum_{k}\left\langle\Lambda^{\sigma}B_{k}\widetilde{V},\Lambda^{\sigma}J_{\varepsilon}^{2}\widetilde{V}-\Lambda^{\sigma}\widetilde{V}\right\rangle d\widetilde{W}_{r}^{k}\right|
C𝔼(0TkΛσBkV~,ΛσJε2V~ΛσV~2dr)12\displaystyle\leq C\mathbb{E}\left(\int_{0}^{T}\sum_{k}\left\langle\Lambda^{\sigma}B_{k}\widetilde{V},\Lambda^{\sigma}J_{\varepsilon}^{2}\widetilde{V}-\Lambda^{\sigma}\widetilde{V}\right\rangle^{2}dr\right)^{\frac{1}{2}}
C𝔼(0Tk[Λσ,Bk]V~,ΛσJε2V~ΛσV~2dr)12+C𝔼(0Tk[Jε,Bk]ΛσV~,ΛσJεV~2dr)12\displaystyle\leq C\mathbb{E}\left(\int_{0}^{T}\sum_{k}\left\langle[\Lambda^{\sigma},B_{k}]\widetilde{V},\Lambda^{\sigma}J_{\varepsilon}^{2}\widetilde{V}-\Lambda^{\sigma}\widetilde{V}\right\rangle^{2}dr\right)^{\frac{1}{2}}+C\mathbb{E}\left(\int_{0}^{T}\sum_{k}\left\langle[J_{\varepsilon},B_{k}]\Lambda^{\sigma}\widetilde{V},\Lambda^{\sigma}J_{\varepsilon}\widetilde{V}\right\rangle^{2}dr\right)^{\frac{1}{2}}
δ𝔼supt[0,T]V~σ2+Cδ𝔼(0TΛσJε2V~ΛσV~2𝑑r+0T[Jε,Bk]ΛσV~2𝑑r)0,\displaystyle\lesssim\delta\mathbb{E}\sup_{t\in[0,T]}\|\widetilde{V}\|_{\sigma}^{2}+C_{\delta}\mathbb{E}\left(\int_{0}^{T}\|\Lambda^{\sigma}J_{\varepsilon}^{2}\widetilde{V}-\Lambda^{\sigma}\widetilde{V}\|^{2}dr+\int_{0}^{T}\|[J_{\varepsilon},B_{k}]\Lambda^{\sigma}\widetilde{V}\|^{2}dr\right)\to 0,

by first letting ε0\varepsilon\to 0 then δ0\delta\to 0. Here, the convergence of the last integral follows from the Friedrichs’ lemma [34, 57]. Thus, there exists a sequence εn0\varepsilon_{n}\to 0 such that as nn\to\infty,

02kJεn𝒫BkV~,Λ2σJεnV~dW~rk02k𝒫BkV~,Λ2σV~dW~rk,\int_{0}^{\cdot}2\sum_{k}\langle J_{\varepsilon_{n}}\mathcal{P}B_{k}\widetilde{V},\Lambda^{2\sigma}J_{\varepsilon_{n}}\widetilde{V}\rangle d\widetilde{W}_{r}^{k}\to\int_{0}^{\cdot}2\sum_{k}\langle\mathcal{P}B_{k}\widetilde{V},\Lambda^{2\sigma}\widetilde{V}\rangle d\widetilde{W}_{r}^{k},

almost surely in C([0,T];)C([0,T];\mathbb{R}).

Consequently, for any 0t1<t2T0\leq t_{1}<t_{2}\leq T, we have almost surely

|V~(t2)σ2V~(t1)σ2|=limn|JεnV~(t2)σ2JεnV~(t1)σ2|t1t2(2(Cp,σ,ρ+2)V~σ+s22+CV~σ2)𝑑r+|t1t22k𝒫BkV~,Λ2σV~dW~rk|.\displaystyle\begin{split}\Big{|}\|\widetilde{V}(t_{2})\|_{\sigma}^{2}&-\|\widetilde{V}(t_{1})\|_{\sigma}^{2}\Big{|}=\lim_{n\to\infty}\left|\|J_{\varepsilon_{n}}\widetilde{V}(t_{2})\|_{\sigma}^{2}-\|J_{\varepsilon_{n}}\widetilde{V}(t_{1})\|_{\sigma}^{2}\right|\\ &\leq\int_{t_{1}}^{t_{2}}\left(2(C_{p,\sigma,\rho}+2)\|\widetilde{V}\|_{\sigma+\frac{s}{2}}^{2}+C\|\widetilde{V}\|_{\sigma}^{2}\right)dr+\left|\int_{t_{1}}^{t_{2}}2\sum_{k}\langle\mathcal{P}B_{k}\widetilde{V},\Lambda^{2\sigma}\widetilde{V}\rangle d\widetilde{W}_{r}^{k}\right|.\end{split} (3.16)

From (3.12), we know that V~L2(0,T;σ+s2)L(0,T;σ)\widetilde{V}\in L^{2}\left(0,T;\mathbb{H}^{\sigma+\frac{s}{2}}\right)\cap L^{\infty}\left(0,T;\mathbb{H}^{\sigma}\right) almost surely. Therefore (3.16) implies the continuity of tV~(t)σt\to\|\widetilde{V}(t)\|_{\sigma}. Combining this with the weak continuity proved earlier, we conclude that V~C([0,T],σ)\widetilde{V}\in C([0,T],\mathbb{H}^{\sigma}) almost surely, and hence V~L2(Ω;C([0,T],σ))\widetilde{V}\in L^{2}(\Omega;C([0,T],\mathbb{H}^{\sigma})). ∎

3.3. Pathwise uniqueness and local pathwise solutions

In this subsection, we establish the pathwise uniqueness of martingale solutions to the cutoff system (3.2). Combining this result with the existence of a martingale solution (Proposition 3.2), we then deduce the existence of a unique pathwise solution.

Proposition 3.3.

Assume the same conditions as in Proposition 3.1. Let (𝒮,𝕎,V1)(\mathcal{S},\mathbb{W},V_{1}) and (𝒮,𝕎,V2)(\mathcal{S},\mathbb{W},V_{2}) be two solutions to the cutoff system (3.2) on [0,T][0,T] with the same initial data V0V_{0}, over the same stochastic basis 𝒮\mathcal{S} and driven by the same noise 𝕎\mathbb{W}. Then

(V1=V2, for all t[0,T])=1.\mathbb{P}\left(V_{1}=V_{2},\text{ for all }t\in[0,T]\right)=1.
Proof.

Let V¯=V1V2\overline{V}=V_{1}-V_{2}. Denote 1,=W1,\|\cdot\|_{1,\infty}=\|\cdot\|_{W^{1,\infty}} for convenience. Note that V¯\overline{V} solves

dV¯+[θρ2(V11,)𝒫Q(V1,V1)θρ2(V21,)𝒫Q(V2,V2)+𝒫F(V¯)+ΛsV¯]dt=12k=1(𝒫Bk)2V¯+k=1𝒫BkV¯dWk,\displaystyle\begin{split}d\overline{V}+\big{[}\theta_{\rho}^{2}(\|V_{1}\|_{1,\infty})\mathcal{P}Q(V_{1},V_{1})-\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\mathcal{P}Q(V_{2},V_{2})&+\mathcal{P}F(\overline{V})+\Lambda^{s}\overline{V}\big{]}dt\\ &=\frac{1}{2}\sum_{k=1}^{\infty}(\mathcal{P}B_{k})^{2}\overline{V}+\sum_{k=1}^{\infty}\mathcal{P}B_{k}\overline{V}dW^{k},\end{split}

with initial data V¯0=0\overline{V}_{0}=0. By Itô’s formula, we have

12dV¯2+[θρ2(V11,)𝒫Q(V1,V1)\displaystyle\frac{1}{2}d\|\overline{V}\|^{2}+\Big{\langle}\big{[}\theta_{\rho}^{2}(\|V_{1}\|_{1,\infty})\mathcal{P}Q(V_{1},V_{1}) θρ2(V21,)𝒫Q(V2,V2)+ΛsV¯],V¯dt\displaystyle-\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\mathcal{P}Q(V_{2},V_{2})+\Lambda^{s}\overline{V}\big{]},\overline{V}\Big{\rangle}dt
=12k=1((𝒫Bk)2V¯,V¯+𝒫BkV¯2)dt+k=1𝒫BkV¯,V¯dWk.\displaystyle=\frac{1}{2}\sum_{k=1}^{\infty}\left(\left\langle(\mathcal{P}B_{k})^{2}\overline{V},\overline{V}\right\rangle+\|\mathcal{P}B_{k}\overline{V}\|^{2}\right)dt+\sum_{k=1}^{\infty}\langle\mathcal{P}B_{k}\overline{V},\overline{V}\rangle dW^{k}.

Using integration by parts, one has 𝒫Q(V2,V¯),V¯=𝒫BkV¯,V¯=0\langle\mathcal{P}Q(V_{2},\overline{V}),\overline{V}\rangle=\langle\mathcal{P}B_{k}\overline{V},\overline{V}\rangle=0, and (𝒫Bk)2V¯,V¯=𝒫BkV¯2\langle(\mathcal{P}B_{k})^{2}\overline{V},\overline{V}\rangle=-\|\mathcal{P}B_{k}\overline{V}\|^{2}. Thus, the above equation reduces to

12dV¯2+(θρ2(V11,)θρ2(V21,))𝒫Q(V1,V1)\displaystyle\frac{1}{2}d\|\overline{V}\|^{2}+\Big{\langle}\left(\theta_{\rho}^{2}(\|V_{1}\|_{1,\infty})-\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\right)\mathcal{P}Q(V_{1},V_{1}) +θρ2(V21,)𝒫Q(V¯,V1),V¯dt\displaystyle+\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\mathcal{P}Q(\overline{V},V_{1}),\overline{V}\Big{\rangle}dt
=Λs2V¯2dt.\displaystyle=-\|\Lambda^{\frac{s}{2}}\overline{V}\|^{2}dt.

For σ>3\sigma>3, using Lemma A.1 and Sobolev’s inequality f1,fσ12\|f\|_{1,\infty}\lesssim\|f\|_{\sigma-\frac{1}{2}}, we estimate

(θρ2(V11,)θρ2(V21,))𝒫Q(V1,V1),V¯CV¯1,V¯V11,V11CV¯σ122V1σ122,\displaystyle\Big{\langle}\left(\theta_{\rho}^{2}(\|V_{1}\|_{1,\infty})-\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\right)\mathcal{P}Q(V_{1},V_{1}),\overline{V}\Big{\rangle}\leq C\|\overline{V}\|_{1,\infty}\|\overline{V}\|\|V_{1}\|_{1,\infty}\|V_{1}\|_{1}\leq C\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}\|V_{1}\|_{\sigma-\frac{1}{2}}^{2},

and

θρ2(V21,)𝒫Q(V¯,V1),V¯C(V¯1,V11+V¯1V11,)V¯CV¯σ122V1σ12.\displaystyle\Big{\langle}\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\mathcal{P}Q(\overline{V},V_{1}),\overline{V}\Big{\rangle}\leq C\left(\|\overline{V}\|_{1,\infty}\|V_{1}\|_{1}+\|\overline{V}\|_{1}\|\|V_{1}\|_{1,\infty}\right)\|\overline{V}\|\leq C\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}\|V_{1}\|_{\sigma-\frac{1}{2}}.

Consequently, we obtain

12dV¯2+Λs2V¯2dtCV¯σ122(V1σ122+1)dt.\displaystyle\frac{1}{2}d\|\overline{V}\|^{2}+\|\Lambda^{\frac{s}{2}}\overline{V}\|^{2}dt\leq C\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}\left(\|V_{1}\|_{\sigma-\frac{1}{2}}^{2}+1\right)dt. (3.17)

Next, we estimate the σ12\sigma-\frac{1}{2} norm of V¯\overline{V}. By Itô’s formula we have

12dΛσ12V¯2\displaystyle\frac{1}{2}d\|\Lambda^{\sigma-\frac{1}{2}}\overline{V}\|^{2} +Λσ12[θρ2(V11,)𝒫Q(V1,V1)θρ2(V21,)𝒫Q(V2,V2)+ΛsV¯],Λσ12V¯dt\displaystyle+\Big{\langle}\Lambda^{\sigma-\frac{1}{2}}\big{[}\theta_{\rho}^{2}(\|V_{1}\|_{1,\infty})\mathcal{P}Q(V_{1},V_{1})-\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\mathcal{P}Q(V_{2},V_{2})+\Lambda^{s}\overline{V}\big{]},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\Big{\rangle}dt
=12k=1(Λσ12(𝒫Bk)2V¯,Λσ12V¯+Λσ12𝒫BkV¯2)dt+k=1Λσ12𝒫BkV¯,Λσ12V¯dWk.\displaystyle=\frac{1}{2}\sum_{k=1}^{\infty}\left(\left\langle\Lambda^{\sigma-\frac{1}{2}}(\mathcal{P}B_{k})^{2}\overline{V},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\right\rangle+\|\Lambda^{\sigma-\frac{1}{2}}\mathcal{P}B_{k}\overline{V}\|^{2}\right)dt+\sum_{k=1}^{\infty}\langle\Lambda^{\sigma-\frac{1}{2}}\mathcal{P}B_{k}\overline{V},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\rangle dW^{k}.

Following arguments similar to to (3.7) and (3.8), we have

12k=1(Λσ12(𝒫Bk)2V¯,Λσ12V¯+Λσ12𝒫BkV¯2)14Λσ12+s2V¯2+CV¯σ122.\displaystyle\frac{1}{2}\sum_{k=1}^{\infty}\left(\left\langle\Lambda^{\sigma-\frac{1}{2}}(\mathcal{P}B_{k})^{2}\overline{V},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\right\rangle+\|\Lambda^{\sigma-\frac{1}{2}}\mathcal{P}B_{k}\overline{V}\|^{2}\right)\leq\frac{1}{4}\|\Lambda^{\sigma-\frac{1}{2}+\frac{s}{2}}\overline{V}\|^{2}+C\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}.

For the nonlinear term, by utilizing the double cutoff design, we first rewrite

θρ2(V11,)𝒫Q(V1,V1)θρ2(V21,)𝒫Q(V2,V2)\displaystyle\theta_{\rho}^{2}(\|V_{1}\|_{1,\infty})\mathcal{P}Q(V_{1},V_{1})-\theta_{\rho}^{2}(\|V_{2}\|_{1,\infty})\mathcal{P}Q(V_{2},V_{2})
=(θρ(V11,)θρ(V21,))(θρ(V11,)𝒫Q(V1,V1)+θρ(V21,)𝒫Q(V2,V2))\displaystyle=\Big{(}\theta_{\rho}(\|V_{1}\|_{1,\infty})-\theta_{\rho}(\|V_{2}\|_{1,\infty})\Big{)}\Big{(}\theta_{\rho}(\|V_{1}\|_{1,\infty})\mathcal{P}Q(V_{1},V_{1})+\theta_{\rho}(\|V_{2}\|_{1,\infty})\mathcal{P}Q(V_{2},V_{2})\Big{)}
+θρ(V21,)θρ(V11,)𝒫(Q(V1,V1)Q(V2,V2)):=Q1+Q2,\displaystyle\qquad+\theta_{\rho}(\|V_{2}\|_{1,\infty})\theta_{\rho}(\|V_{1}\|_{1,\infty})\mathcal{P}\left(Q(V_{1},V_{1})-Q(V_{2},V_{2})\right):=Q_{1}+Q_{2},

where the identity a12b1a22b2=(a1a2)(a1b1+a2b2)+a1a2(b1b2)a_{1}^{2}b_{1}-a_{2}^{2}b_{2}=(a_{1}-a_{2})(a_{1}b_{1}+a_{2}b_{2})+a_{1}a_{2}(b_{1}-b_{2}) has been used. By the Lipschitz continuity of the cutoff function, Lemma A.1 and Sobolev’s inequality, we have

Λσ12Q1,Λσ12V¯Λσ12Q1Λσ12V¯V¯σ122(V1σ+122+V2σ+122).\displaystyle\Big{\langle}\Lambda^{\sigma-\frac{1}{2}}Q_{1},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\Big{\rangle}\leq\|\Lambda^{\sigma-\frac{1}{2}}Q_{1}\|\|\Lambda^{\sigma-\frac{1}{2}}\overline{V}\|\leq\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}\left(\|V_{1}\|_{\sigma+\frac{1}{2}}^{2}+\|V_{2}\|_{\sigma+\frac{1}{2}}^{2}\right).

Thanks to the property of the cutoff function, we deduce

Λσ12Q2,Λσ12V¯\displaystyle\Big{\langle}\Lambda^{\sigma-\frac{1}{2}}Q_{2},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\Big{\rangle} =θρ(V21,)θρ(V11,)Λσ1[(Q(V1,V¯)+Q(V¯,V2))],ΛσV¯\displaystyle=\theta_{\rho}(\|V_{2}\|_{1,\infty})\theta_{\rho}(\|V_{1}\|_{1,\infty})\left\langle\Lambda^{\sigma-1}\left[\left(Q(V_{1},\overline{V})+Q(\overline{V},V_{2})\right)\right],\Lambda^{\sigma}\overline{V}\right\rangle
Cσ(ρΛσV¯+(ΛσV1+ΛσV2)V¯1,)ΛσV¯\displaystyle\leq C_{\sigma}\left(\rho\|\Lambda^{\sigma}\overline{V}\|+\left(\|\Lambda^{\sigma}V_{1}\|+\|\Lambda^{\sigma}V_{2}\|\right)\|\overline{V}\|_{1,\infty}\right)\|\Lambda^{\sigma}\overline{V}\|
CσρΛσV¯2+Cρ(V1σ2+V2σ2)V¯σ122,\displaystyle\leq C_{\sigma}\rho\|\Lambda^{\sigma}\overline{V}\|^{2}+C_{\rho}\left(\|V_{1}\|_{\sigma}^{2}+\|V_{2}\|_{\sigma}^{2}\right)\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2},

where we used Young’s inequality for products at the final step.

Now we need to distinguish between s(1,2]s\in(1,2] and s=1s=1. For s>1s>1 we infer from the interpolation inequality ΛσV¯2εΛσ12+s2V¯2+CεΛσ12V¯2\|\Lambda^{\sigma}\overline{V}\|^{2}\leq\varepsilon\|\Lambda^{\sigma-\frac{1}{2}+\frac{s}{2}}\overline{V}\|^{2}+C_{\varepsilon}\|\Lambda^{\sigma-\frac{1}{2}}\overline{V}\|^{2} that

Λσ12Q2,Λσ12V¯12Λσ12+s2V¯2+C(1+V1σ2+V2σ2)V¯σ122,\displaystyle\Big{\langle}\Lambda^{\sigma-\frac{1}{2}}Q_{2},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\Big{\rangle}\leq\frac{1}{2}\|\Lambda^{\sigma-\frac{1}{2}+\frac{s}{2}}\overline{V}\|^{2}+C\left(1+\|V_{1}\|_{\sigma}^{2}+\|V_{2}\|_{\sigma}^{2}\right)\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}, (3.18)

by choosing appropriate ε\varepsilon. For s=1s=1, we impose the smallness condition ρCσ<12\rho C_{\sigma}<\frac{1}{2}, ensuring the same bound (3.18) holds. Combining the above estimates, we obtain

12dΛσ12V¯2\displaystyle\frac{1}{2}d\|\Lambda^{\sigma-\frac{1}{2}}\overline{V}\|^{2} +14Λσ12+s2V¯2dt\displaystyle+\frac{1}{4}\|\Lambda^{\sigma-\frac{1}{2}+\frac{s}{2}}\overline{V}\|^{2}dt
C(1+V1σ+122+V2σ+122)V¯σ122dt+k=1Λσ12𝒫BkV¯,Λσ12V¯dWtk.\displaystyle\leq C\left(1+\|V_{1}\|_{\sigma+\frac{1}{2}}^{2}+\|V_{2}\|_{\sigma+\frac{1}{2}}^{2}\right)\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}dt+\sum_{k=1}^{\infty}\langle\Lambda^{\sigma-\frac{1}{2}}\mathcal{P}B_{k}\overline{V},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\rangle dW_{t}^{k}.

This together with estimate (3.17) yields

dV¯σ122C(1+V1σ+122+V2σ+122)V¯σ122dt+2k=1Λσ12𝒫BkV¯,Λσ12V¯dWtk.\displaystyle d\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}\leq C\left(1+\|V_{1}\|_{\sigma+\frac{1}{2}}^{2}+\|V_{2}\|_{\sigma+\frac{1}{2}}^{2}\right)\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}dt+2\sum_{k=1}^{\infty}\langle\Lambda^{\sigma-\frac{1}{2}}\mathcal{P}B_{k}\overline{V},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\rangle dW_{t}^{k}.

Letting

Yt=exp(C0t(1+V1σ+122+V2σ+122)𝑑r),Y_{t}=\exp\left(-C\int_{0}^{t}\left(1+\|V_{1}\|_{\sigma+\frac{1}{2}}^{2}+\|V_{2}\|_{\sigma+\frac{1}{2}}^{2}\right)dr\right),

Itô’s formula gives

d(YtV¯σ122)2Ytk=1Λσ12𝒫BkV¯,Λσ12V¯dWtk.\displaystyle d\left(Y_{t}\|\overline{V}\|_{\sigma-\frac{1}{2}}^{2}\right)\leq 2Y_{t}\sum_{k=1}^{\infty}\langle\Lambda^{\sigma-\frac{1}{2}}\mathcal{P}B_{k}\overline{V},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\rangle dW_{t}^{k}.

In the integral form, this reads

YtV¯(t)σ122V¯0σ122+2k=10tYrΛσ12𝒫BkV¯,Λσ12V¯𝑑Wrk.\displaystyle Y_{t}\|\overline{V}(t)\|_{\sigma-\frac{1}{2}}^{2}\leq\|\overline{V}_{0}\|_{\sigma-\frac{1}{2}}^{2}+2\sum_{k=1}^{\infty}\int_{0}^{t}Y_{r}\langle\Lambda^{\sigma-\frac{1}{2}}\mathcal{P}B_{k}\overline{V},\Lambda^{\sigma-\frac{1}{2}}\overline{V}\rangle dW_{r}^{k}.

Since V0=0V_{0}=0 and the stochastic integral is a martingale, we obtain

𝔼[YtV¯(t)σ122]0.\displaystyle\mathbb{E}\left[Y_{t}\|\overline{V}(t)\|_{\sigma-\frac{1}{2}}^{2}\right]\leq 0.

As 0<Yt10<Y_{t}\leq 1, it follows that V¯(t)σ122=0\|\overline{V}(t)\|_{\sigma-\frac{1}{2}}^{2}=0 almost surely for all tt. Because V1V_{1} and V2V_{2} are modifications both with continuous sample paths, they are indistinguishable. This completes the proof of pathwise uniqueness. ∎

We are now ready to prove the main result, Theorem 2.3, concerning the existence of a unique maximal pathwise solution.

Proof of Theorem 2.3.

By the Yamada-Watanabe theorem [46], along with Proposition 3.2 and Proposition 3.3, we know that for any T>0T>0, the cutoff system (3.2) has a unique pathwise solution VL2(Ω;C([0,T;σ])L2(0,T;σ+s2))V\in L^{2}(\Omega;C([0,T;\mathbb{\mathbb{H}}^{\sigma}])\cap L^{2}(0,T;\mathbb{H}^{\sigma+\frac{s}{2}})). Define the stopping time

τ=inf{t0:Vσ>ρ}.\displaystyle\tau=\inf\{t\geq 0:\|V\|_{\sigma}>\rho\}. (3.19)

We first look at the case when s(1,2)s\in(1,2). Let C0C_{0} be the constant from the embedding HσW1,H^{\sigma}\subset W^{1,\infty}. Assume that for some deterministic M>0M>0, we have V0σM\|V_{0}\|_{\sigma}\leq M. Then, for any ρ>2C0M\rho>2C_{0}M, the stopping time τ\tau is positive. Hence, (V,τ)(V,\tau) is a local pathwise solution of (2.4). To generalize this result for V0L2(Ω,σ)V_{0}\in L^{2}(\Omega,\mathbb{H}^{\sigma}), we use a localization procedure. For any k0k\geq 0, set V0k=V0𝟏{kV0σk+1}V_{0}^{k}=V_{0}\mathbf{1}_{\{k\leq\|V_{0}\|_{\sigma}\leq k+1\}}. Then the above argument yields a local pathwise solution (Vk,τk)(V_{k},\tau_{k}) with ρ>2C0(k+1)\rho>2C_{0}(k+1). Then by defining

V=k0Vk𝟏{kV0σk+1},τ=k0τk𝟏{kV0σk+1},\displaystyle V=\sum_{k\geq 0}V_{k}\mathbf{1}_{\{k\leq\|V_{0}\|_{\sigma}\leq k+1\}},\quad\tau=\sum_{k\geq 0}\tau_{k}\mathbf{1}_{\{k\leq\|V_{0}\|_{\sigma}\leq k+1\}},

we obtain the local pathwise solution (V,τ)(V,\tau) of (2.4) with initial data V0L2(Ω,σ)V_{0}\in L^{2}(\Omega,\mathbb{H}^{\sigma}). To extend the solution to a maximal one, let 𝒯\mathcal{T} be the set of all stopping times corresponding to a local pathwise solution of (2.4) with initial data V0V_{0}. By [26, Chapter V, Section 18], there exists a stopping time ξ\xi such that ξ>τ\xi>\tau almost surely for any τ𝒯\tau\in\mathcal{T}, and there exists a sequence {τn}𝒯\{\tau_{n}\}\subset\mathcal{T} satisfying τnξ\tau_{n}\nearrow\xi almost surely. Let (Vn,τn)(V_{n},\tau_{n}) be the corresponding local pathwise solution and define

V(t,ω)=limnVn(tτn,ω)𝟏[0,ξ)(ω).\displaystyle V(t,\omega)=\lim_{n\to\infty}V_{n}(t\wedge\tau_{n},\omega)\mathbf{1}_{[0,\xi)}(\omega).

Then (V,(τn)n1,ξ)(V,(\tau_{n})_{n\geq 1},\xi) is the desired maximal pathwise solution in the sense of Definition 2.1.

For the case s=1s=1, we require the cutoff parameter ρ<12Cσ\rho<\frac{1}{2C_{\sigma}}, as stated in Proposition 3.1. Therefore for initial data V0V_{0} with V0σ<M:=14C0Cσ\|V_{0}\|_{\sigma}<M:=\frac{1}{4C_{0}C_{\sigma}}, we can always choose ρ\rho such that 2C0M<ρ<12Cσ2C_{0}M<\rho<\frac{1}{2C_{\sigma}} to ensure a unique local pathwise solution (V,τ)(V,\tau), where the stopping time τ>0\tau>0 corresponds to ρ\rho through (3.19). The extension to a maximal solution proceeds in the same manner as in the case s(1,2)s\in(1,2). This completes the proof. ∎

4. Remarks on the supercritical case

In this section, we discuss the supercritical case (s<1s<1). Two significant challenges arise in this setting. First, according to [32, Lemmas A.1 and A.3], the nonlinear term satisfies the following estimate:

|ΛσQ(V,V),ΛσV|CσVσΛσ+12V2.\begin{split}\Big{|}\Big{\langle}\Lambda^{\sigma}Q(V,V),\Lambda^{\sigma}V\Big{\rangle}\Big{|}\leq C_{\sigma}\|V\|_{\sigma}\|\Lambda^{\sigma+\frac{1}{2}}V\|^{2}.\end{split}

Thus for s<1s<1, the dissipation term Λσ+s2V2\|\Lambda^{\sigma+\frac{s}{2}}V\|^{2} is insufficient to control the nonlinearity. Indeed, as shown in [2], the PE system with fractional dissipation s<1s<1 is ill-posed in Sobolev spaces. Therefore, one must work in the analytic class with a decaying analytic radius, similar to the approach in [35].

The second difficulty arises when working with transport noise in the analytic class. Specifically, as shown in the example below, the cancellation of the highest-order terms involving the Itô-Stratonovich corrector, as in (3.6), is no longer valid in the analytic setting unless the noise coefficient bb is independent of the spatial variable. If bb is spatially constant, the computations do not involve significant additional difficulties; thus, we omit them, referring interested readers to [35]. Consequently, for s<1s<1, the method developed in this work can only establish the local existence of pathwise solutions to (1.1) in the analytic class when bb is spatially constant. The general case, where bb depends on spatial variables, requires alternative approaches and remains under investigation.

In the analytic setting, the cancellation terms involving the Itô-Stratonovich corrector analogous to (3.6) are given by

ΛσeτΛ(𝒫Bk)2V,ΛσeτΛV+ΛσeτΛ(𝒫Bk)V2,\big{\langle}\Lambda^{\sigma}e^{\tau\Lambda}(\mathcal{P}B_{k})^{2}V,\Lambda^{\sigma}e^{\tau\Lambda}V\big{\rangle}+\|\Lambda^{\sigma}e^{\tau\Lambda}(\mathcal{P}B_{k})V\|^{2},

where τ>0\tau>0 is the analytic radius and eτΛe^{\tau\Lambda} is defined in terms of the Fourier coefficients as:

(eτΛf^)k:=eτ|k|f^k,k2π3.(\widehat{e^{\tau\Lambda}f})_{k}:=e^{\tau|k|}\widehat{f}_{k},\quad k\in 2\pi\mathbb{Z}^{3}.

The following example, set on the 1D torus, demonstrates the cancellation of the terms with highest order Sobolev regularity, as in (3.6), generally does not hold when τ>0\tau>0.

Example 1.

Let b=2cos(x)=eix+eixb=2\cos(x)=e^{ix}+e^{-ix} and denote by aa^{*} the complex conjugate of aa\in\mathbb{C}. A direct computation gives

I:\displaystyle I: =ΛreτΛ(bx(bxf)),ΛreτΛf\displaystyle=\big{\langle}\Lambda^{r}e^{\tau\Lambda}(b\partial_{x}(b\partial_{x}f)),\Lambda^{r}e^{\tau\Lambda}f\big{\rangle}
=k2k2|k|2re2τ|k||fk^|2k(k1)(k2)|k|2re2τ|k|f^k2f^kk(k1)k|k2|2re2τ|k2|f^k2f^k,\displaystyle=-\sum_{k}2k^{2}|k|^{2r}e^{2\tau|k|}|\widehat{f_{k}}|^{2}-\sum_{k}(k-1)(k-2)|k|^{2r}e^{2\tau|k|}\widehat{f}_{k-2}\widehat{f}^{*}_{k}-\sum_{k}(k-1)k|k-2|^{2r}e^{2\tau|k-2|}\widehat{f}^{*}_{k-2}\widehat{f}_{k},

and

II:\displaystyle II: =ΛreτΛ(bxf),ΛreτΛ(bxf)\displaystyle=\big{\langle}\Lambda^{r}e^{\tau\Lambda}(b\partial_{x}f),\Lambda^{r}e^{\tau\Lambda}(b\partial_{x}f)\big{\rangle}
=k((k+1)2re2τ|k+1|+(k1)2re2τ|k1|)k2|f^k|2+kk(k2)|k1|2re2τ|k1|(f^k2f^k+f^k2f^k).\displaystyle=\sum_{k}((k+1)^{2r}e^{2\tau|k+1|}+(k-1)^{2r}e^{2\tau|k-1|})k^{2}|\widehat{f}_{k}|^{2}+\sum_{k}k(k-2)|k-1|^{2r}e^{2\tau|k-1|}(\widehat{f}^{*}_{k-2}\widehat{f}_{k}+\widehat{f}_{k-2}\widehat{f}^{*}_{k}).

Using f^k=f^k\widehat{f}_{k}=\widehat{f}^{*}_{-k}, we have

I+II=k(a(k)|f^k|2+b(k)𝔢(f^k2f^k)),\displaystyle I+II=\sum_{k}\left(a(k)|\widehat{f}_{k}|^{2}+b(k)\mathfrak{Re}(\widehat{f}_{k-2}\widehat{f}^{*}_{k})\right),

where

a(k)\displaystyle a(k) =((k+1)2re2τ|k+1|+(k1)2re2τ|k1|2|k|2re2τ|k|)k2,\displaystyle=((k+1)^{2r}e^{2\tau|k+1|}+(k-1)^{2r}e^{2\tau|k-1|}-2|k|^{2r}e^{2\tau|k|})k^{2},
b(k)\displaystyle b(k) =2k(k2)|k1|2re2τ|k1|(k1)(k2)|k|2re2τ|k|k(k1)|k2|2re2τ|k2|.\displaystyle=2k(k-2)|k-1|^{2r}e^{2\tau|k-1|}-(k-1)(k-2)|k|^{2r}e^{2\tau|k|}-k(k-1)|k-2|^{2r}e^{2\tau|k-2|}.

When τ=0\tau=0, all terms of degree greater than 2r2r cancel out. However, for τ>0\tau>0, such a cancellation breaks down due to the unequal exponential weights. This highlights a crucial difference between Sobolev-type estimates and analytic-type estimates.

Acknowledgments

R.H. was partially supported by a grant from the Simons Foundation (MP-TSM-00002783), and an ONR grant under #N00014-24-1-2432. Q.L. was partially supported by an AMS-Simons travel grant.

Appendix A Auxiliary lemmas

In this appendix, we summarize several lemmas that have been used repeatedly in our analysis.

Lemma A.1 (see [19]).

Let s0s\geq 0 and f,gHsW1,f,g\in H^{s}\cap W^{1,\infty}. We have

Λs(fg)L2fLgs+gLfs.\begin{split}\|\Lambda^{s}(fg)\|_{L^{2}}\lesssim\|f\|_{L^{\infty}}\|g\|_{s}+\|g\|_{L^{\infty}}\|f\|_{s}.\end{split}

For s>0s>0 we have

Λs(fg)fΛsgfLΛs1gL2+ΛsfL2gL.\displaystyle\|\Lambda^{s}(fg)-f\Lambda^{s}g\|\lesssim\|\nabla f\|_{L^{\infty}}\|\Lambda^{s-1}g\|_{L^{2}}+\|\Lambda^{s}f\|_{L^{2}}\|g\|_{L^{\infty}}.

We recall the following compactness result. For proofs, see [54, Theorem 5] and [29, Theorem 2.1], respectively.

Lemma A.2.

a) (Aubin-Lions-Simon Lemma). Let X2XX1X_{2}\subset X\subset X_{1} be Banach spaces such that the embedding X2XX_{2}\hookrightarrow\hookrightarrow X is compact and the embedding XX1X\hookrightarrow X_{1} is continuous. Let p(1,)p\in(1,\infty) and α(0,1)\alpha\in(0,1). Then, the following embedding is compact

Lp(0,t;X2)Wα,p(0,t;X1)Lp(0,t;X).L^{p}(0,t;X_{2})\cap W^{\alpha,p}(0,t;X_{1})\hookrightarrow\hookrightarrow L^{p}(0,t;X).

b) Let X2XX_{2}\subset X be Banach spaces such that X2X_{2} is reflexive and the embedding X2XX_{2}\hookrightarrow\hookrightarrow X is compact. Let α(0,1]\alpha\in(0,1] and p(1,)p\in(1,\infty) be such that αp>1\alpha p>1. Then, the following embedding is compact

Wα,p(0,t;X2)C([0,t],X).W^{\alpha,p}(0,t;X_{2})\hookrightarrow\hookrightarrow C([0,t],X).

We also recall the definitions of Sobolev spaces with fractional time derivative, see e.g. [55]. Let XX be a separable Hilbert space, t>0t>0, p>1p>1 and α(0,1)\alpha\in(0,1). The Sobolev space Wα,p(0,t;X)W^{\alpha,p}(0,t;X) is defined as:

Wα,p(0,t;X)={uLp(0,t;X)0t0t|u(s)u(r)|Xp|sr|1+αp𝑑r𝑑s<},W^{\alpha,p}(0,t;X)=\left\{u\in L^{p}(0,t;X)\mid\int_{0}^{t}\int_{0}^{t}\frac{|u(s)-u(r)|_{X}^{p}}{|s-r|^{1+\alpha p}}\,dr\,ds<\infty\right\},

with the norm:

uWα,p(0,t;X)p=0t|u(s)|Xp𝑑s+0t0t|u(s)u(r)|Xp|sr|1+αp𝑑r𝑑s.\|u\|^{p}_{W^{\alpha,p}(0,t;X)}=\int_{0}^{t}|u(s)|^{p}_{X}\,ds+\int_{0}^{t}\int_{0}^{t}\frac{|u(s)-u(r)|_{X}^{p}}{|s-r|^{1+\alpha p}}\,dr\,ds.

Finally, we recall the Burkholder-Davis-Gundy inequality. For a detailed proof, see [29, Lemma 2.1]. Here L2(𝒰,X)L_{2}\left(\mathscr{U},X\right) represents the Hilbert-Schmidt norm.

Lemma A.3.

Let XX be a separable Hilbert space. For ΦL2(Ω;L2(0,T;L2(𝒰,X)))\Phi\in L^{2}\left(\Omega;L^{2}\left(0,T;L_{2}\left(\mathscr{U},X\right)\right)\right), one has

𝔼supt[0,T]|0tΦ𝑑W|XrCr𝔼(0TΦL2(𝒰,X)2𝑑t)r/2.\mathbb{E}\sup_{t\in\left[0,T\right]}\left|\int_{0}^{t}\Phi\,dW\right|^{r}_{X}\leq C_{r}\,\mathbb{E}\left(\int_{0}^{T}\|\Phi\|_{L_{2}(\mathscr{U},X)}^{2}\,dt\right)^{r/2}. (A.1)

Moreover, if p2p\geq 2 and ΦLp(Ω;Lp(0,T;L2(𝒰,X)))\Phi\in L^{p}\left(\Omega;L^{p}\left(0,T;L_{2}\left(\mathscr{U},X\right)\right)\right), then

𝔼|0Φ𝑑W|Wα,p(0,T;X)pcp𝔼0TΦL2(𝒰,X)p𝑑t,\mathbb{E}\left|\int_{0}^{\cdot}\Phi\,dW\right|^{p}_{W^{\alpha,p}(0,T;X)}\leq c_{p}\,\mathbb{E}\int_{0}^{T}\|\Phi\|_{L_{2}(\mathscr{U},X)}^{p}\,dt, (A.2)

for α[0,1/2)\alpha\in[0,1/2).

Appendix B Commutator estimates

In this appendix, we omit the summation symbol for repeated indices. We first establish a result based on Lemma A.1 and the following estimate obtained in [16]: for any s,ε>0s\in\mathbb{R},\varepsilon>0, and smooth functions f,gf,g over 𝕋d\mathbb{T}^{d}, there exists a constant Cs,ε>0C_{s,\varepsilon}>0 such that (where j=xj\partial_{j}=\partial_{x_{j}})

[Λsj,g]fCs,ε(gHd2+1+εΛsf+gHd2+1+s+εf).\displaystyle\|[\Lambda^{s}\partial_{j},g]f\|\leq C_{s,\varepsilon}\left(\|g\|_{H^{\frac{d}{2}+1+\varepsilon}}\|\Lambda^{s}f\|+\|g\|_{H^{\frac{d}{2}+1+s+\varepsilon}}\|f\|\right). (B.1)

It is worth noting that this result was originally proved in [16] for the 2\mathbb{R}^{2} case, and the generalization to 𝕋d\mathbb{T}^{d} follows in the same manner.

Lemma B.1.

Let s>52s>\frac{5}{2} and α>s\alpha>-s. Then for any smooth divergence-free vector field bb and function VV on 𝕋3\mathbb{T}^{3}, we have

Λα[Λs,b]VbsVs+α+bα+sVs.\displaystyle\|\Lambda^{\alpha}[\Lambda^{s},b\cdot\nabla]V\|\lesssim\|b\|_{s}\|V\|_{s+\alpha}+\|b\|_{\alpha+s}\|V\|_{s}.
Proof.

By the divergence-free property of bb, we first rewrite

Λα[Λs,b]V\displaystyle\Lambda^{\alpha}[\Lambda^{s},b\cdot\nabla]V =Λα+s(bjjV)Λα(bjΛsjV)\displaystyle=\Lambda^{\alpha+s}(b^{j}\partial_{j}V)-\Lambda^{\alpha}(b^{j}\Lambda^{s}\partial_{j}V)
=[Λα+s,bj]jV[Λαj,bj]ΛsV.\displaystyle=[\Lambda^{\alpha+s},b_{j}]\partial_{j}V-[\Lambda^{\alpha}\partial_{j},b^{j}]\Lambda^{s}V.

For a fixed jj, since α+s>0\alpha+s>0 and s>52s>\frac{5}{2}, we apply Lemma A.1 to obtain:

[Λα+s,bj]jV\displaystyle\|[\Lambda^{\alpha+s},b_{j}]\partial_{j}V\| bjLΛα+s1jV+Λα+sbjjVL\displaystyle\lesssim\|\nabla b_{j}\|_{L^{\infty}}\|\Lambda^{\alpha+s-1}\partial_{j}V\|+\|\Lambda^{\alpha+s}b_{j}\|\|\partial_{j}V\|_{L^{\infty}}
bsVα+s+bα+sVs.\displaystyle\lesssim\|b\|_{s}\|V\|_{\alpha+s}+\|b\|_{\alpha+s}\|V\|_{s}.

Next, since s>32+1s>\frac{3}{2}+1, it follows from (B.1) that

[Λαj,bj]ΛsVbsΛα+sV+bα+sΛsV.\displaystyle\|[\Lambda^{\alpha}\partial_{j},b^{j}]\Lambda^{s}V\|\lesssim\|b\|_{s}\|\Lambda^{\alpha+s}V\|+\|b\|_{\alpha+s}\|\Lambda^{s}V\|.

Combining the above estimates, we obtain the desired estimate.

We next establish a lemma on a double commutator.

Lemma B.2.

Assume s>5/2s>5/2. Let bb be a smooth divergence-free vector field on 𝕋3\mathbb{T}^{3} with zero mean. We have

[[Λs,b],b]fbs+12fs,\displaystyle\|[[\Lambda^{s},b\cdot\nabla],b\cdot\nabla]f\|\lesssim\|b\|_{s+1}^{2}\|f\|_{s},

for any real-valued smooth function ff with zero mean.

Proof.

Since bb is divergence-free, we can rewrite:

[Λs,b]f=Λs(j(bjf))bjjΛsf=[Λsj,bj]f.\displaystyle[\Lambda^{s},b\cdot\nabla]f=\Lambda^{s}(\partial_{j}(b^{j}f))-b^{j}\partial_{j}\Lambda^{s}f=[\Lambda^{s}\partial_{j},b^{j}]f. (B.2)

Therefore, one has

[[Λs,b],b]f\displaystyle[[\Lambda^{s},b\cdot\nabla],b\cdot\nabla]f =[Λsj,bj]bfb[Λsj,bj]f\displaystyle=[\Lambda^{s}\partial_{j},b^{j}]b^{\ell}\partial_{\ell}f-b^{\ell}\partial_{\ell}[\Lambda^{s}\partial_{j},b^{j}]f
=[[Λsj,bj],b](f)+b[[Λsj,bj],]f:=I1+I2.\displaystyle=[[\Lambda^{s}\partial_{j},b^{j}],b^{\ell}](\partial_{\ell}f)+b^{\ell}[[\Lambda^{s}\partial_{j},b^{j}],\partial_{\ell}]f:=I_{1}+I_{2}.

Below we analyze them individually.

Let f=ff_{\ell}=\partial_{\ell}f. Since bb is divergence-free, one has

I1\displaystyle I_{1} =[[Λsj,bj],b](f)\displaystyle=[[\Lambda^{s}\partial_{j},b^{j}],b^{\ell}](f_{\ell})
=Λs(bjjbf)+Λs(bjbjf)bjΛs(jbf)bjΛs(bjf)bΛs(bjjf)+bbjΛs(jf)\displaystyle=\Lambda^{s}(b^{j}\partial_{j}b^{\ell}f_{\ell})+\Lambda^{s}(b^{j}b^{\ell}\partial_{j}f_{\ell})-b^{j}\Lambda^{s}(\partial_{j}b^{\ell}f_{\ell})-b^{j}\Lambda^{s}(b^{\ell}\partial_{j}f_{\ell})-b^{\ell}\Lambda^{s}(b^{j}\partial_{j}f_{\ell})+b^{\ell}b^{j}\Lambda^{s}(\partial_{j}f_{\ell})
=[Λs,bj]jbf+[Λs,bj](bjf)b[Λs,bj](jf)\displaystyle=[\Lambda^{s},b^{j}]\partial_{j}b^{\ell}f_{\ell}+[\Lambda^{s},b^{j}](b^{\ell}\partial_{j}f_{\ell})-b^{\ell}[\Lambda^{s},b^{j}](\partial_{j}f_{\ell})
=[Λs,bj]jbf+[[Λs,bj],b]jf.\displaystyle=[\Lambda^{s},b^{j}]\partial_{j}b^{\ell}f_{\ell}+[[\Lambda^{s},b^{j}],b^{\ell}]\partial_{j}f_{\ell}.

By Lemma A.1 and Sobolev embeddings, we estimate

[Λs,bj]jbfbjLΛs1(jbf)+ΛsbjjbfLbs2fs.\displaystyle\|[\Lambda^{s},b^{j}]\partial_{j}b^{\ell}f_{\ell}\|\lesssim\|\nabla b^{j}\|_{L^{\infty}}\|\Lambda^{s-1}(\partial_{j}b^{\ell}f_{\ell})\|+\|\Lambda^{s}b^{j}\|\|\partial_{j}b^{\ell}f_{\ell}\|_{L^{\infty}}\lesssim\|b\|_{s}^{2}\|f\|_{s}.

For the second term in I1I_{1}, let fj=jff_{j\ell}=\partial_{j}f_{\ell}. Then

[[Λs,bj],b]fj=Λs(bjbfj)bjΛs(bfj)bΛs(bjfj)+bjbΛsfj.\displaystyle[[\Lambda^{s},b^{j}],b^{\ell}]f_{j\ell}=\Lambda^{s}(b^{j}b^{\ell}f_{j\ell})-b^{j}\Lambda^{s}(b^{\ell}f_{j\ell})-b^{\ell}\Lambda^{s}(b^{j}f_{j\ell})+b^{j}b^{\ell}\Lambda^{s}f_{j\ell}. (B.3)

By a straightforward application of the fractional Leibniz rule as in [48, formula (1.6)] with s1=2,p=2,g=fj,s_{1}=2,p=2,g=f_{j\ell}, we obtain that

Λs(bjbfj)=|α|21α!α(bjb)Λs,αfj+|β|<s21β!βfjΛs,β(bjb)+E1,\displaystyle\Lambda^{s}(b^{j}b^{\ell}f_{j\ell})=\sum_{|\alpha|\leq 2}\frac{1}{\alpha!}\partial^{\alpha}(b^{j}b^{\ell})\Lambda^{s,\alpha}f_{j\ell}+\sum_{|\beta|<s-2}\frac{1}{\beta!}\partial^{\beta}f_{j\ell}\Lambda^{s,\beta}(b^{j}b^{\ell})+E_{1},

where Λs,α\Lambda^{s,\alpha} is a multiplier of order s|α|s-|\alpha| with Λs,0=Λs\Lambda^{s,0}=\Lambda^{s}, and E1Λ2(bjb)LΛs2fjbs+12fs\|E_{1}\|\lesssim\|\Lambda^{2}(b^{j}b^{\ell})\|_{L^{\infty}}\|\Lambda^{s-2}f_{j\ell}\|\lesssim\|b\|_{s+1}^{2}\|f\|_{s}. Similarly, we have

Λs(bfj)=|α|21α!αbΛs,αfj+|β|<s21β!βfjΛs,βb+E2,\displaystyle\Lambda^{s}(b^{\ell}f_{j\ell})=\sum_{|\alpha|\leq 2}\frac{1}{\alpha!}\partial^{\alpha}b^{\ell}\Lambda^{s,\alpha}f_{j\ell}+\sum_{|\beta|<s-2}\frac{1}{\beta!}\partial^{\beta}f_{j\ell}\Lambda^{s,\beta}b^{\ell}+E_{2},

and

Λs(bjfj)=|α|21α!αbjΛs,αfj+|β|<s21β!βfjΛs,βbj+E3,\displaystyle\Lambda^{s}(b^{j}f_{j\ell})=\sum_{|\alpha|\leq 2}\frac{1}{\alpha!}\partial^{\alpha}b^{j}\Lambda^{s,\alpha}f_{j\ell}+\sum_{|\beta|<s-2}\frac{1}{\beta!}\partial^{\beta}f_{j\ell}\Lambda^{s,\beta}b^{j}+E_{3},

with

E2Λ2bLΛs2fjbs+1fs, and E3Λ2bjLΛs2fjbs+1fs.\|E_{2}\|\lesssim\|\Lambda^{2}b^{\ell}\|_{L^{\infty}}\|\Lambda^{s-2}f_{j\ell}\|\lesssim\|b\|_{s+1}\|f\|_{s},\text{ and }\|E_{3}\|\lesssim\|\Lambda^{2}b^{j}\|_{L^{\infty}}\|\Lambda^{s-2}f_{j\ell}\|\lesssim\|b\|_{s+1}\|f\|_{s}.

In view of (B.3), the derivatives on fjf_{j\ell} of orders s1s-1 and ss cancel out. We therefore obtain by Sobolev embeddings that

[[Λs,bj],b]fjbs+12fs.\displaystyle\|[[\Lambda^{s},b^{j}],b^{\ell}]f_{j\ell}\|\lesssim\|b\|_{s+1}^{2}\|f\|_{s}.

For the commutator in I2I_{2}, we compute

[[Λsj,bj],]f\displaystyle[[\Lambda^{s}\partial_{j},b^{j}],\partial_{\ell}]f =Λs(bjjf)bjΛsjf(Λs(bjjf))+(bjΛsjf)\displaystyle=\Lambda^{s}(b^{j}\partial_{j}\partial_{\ell}f)-b^{j}\Lambda^{s}\partial_{j}\partial_{\ell}f-\partial_{\ell}(\Lambda^{s}(b^{j}\partial_{j}f))+\partial_{\ell}(b^{j}\Lambda^{s}\partial_{j}f)
=[bj,Λs]jf,\displaystyle=[\partial_{\ell}b^{j},\Lambda^{s}]\partial_{j}f,

which together with Lemma A.1 lead to

I2bL(bjLΛs1jf+ΛsbjjfL)bs+12fs.\displaystyle\|I_{2}\|\lesssim\|b^{\ell}\|_{L^{\infty}}\left(\|\nabla\partial_{\ell}b^{j}\|_{L^{\infty}}\|\Lambda^{s-1}\partial_{j}f\|+\|\Lambda^{s}\partial_{\ell}b^{j}\|\|\partial_{j}f\|_{L^{\infty}}\right)\lesssim\|b\|_{s+1}^{2}\|f\|_{s}.

This completes the proof.

Next, we provide a lemma concerning commutator estimates in Sobolev spaces with negative exponents. Since we could not locate a corresponding result in the literature, we provide the proof for completeness.

Lemma B.3.

Let s1,0αss\geq 1,0\leq\alpha\leq s and β>32\beta>\frac{3}{2}. For any smooth divergence-free vector field bb and function VV on 𝕋3\mathbb{T}^{3}, we have

Λα[Λs,b]Vs,α,βbs+α+βVsα.\displaystyle\|\Lambda^{-\alpha}[\Lambda^{s},b\cdot\nabla]V\|\lesssim_{s,\alpha,\beta}\|b\|_{s+\alpha+\beta}\|V\|_{s-\alpha}.
Proof.

Using the Minkowski inequality and the fact that

|asbs|s|ab|(bs1+|ab|s1),|a^{s}-b^{s}|\lesssim_{s}|a-b|(b^{s-1}+|a-b|^{s-1}),

for a,b0a,b\geq 0, we have

[Λs,b]Vα=(k|k|2α|jb^kjijV^j(|k|s|j|s)|2)12s(k|k|2α(j|b^kj||j||V^j||kj|(|j|s1+|kj|s1))2)12s(k|k|2α(j|b^kj||V^j||kj||j|s)2)12+(k|k|2α(j|b^kj||j||V^j||kj|s)2)12:=I1+I2.\displaystyle\begin{split}\|[\Lambda^{s},b\cdot\nabla]V\|_{-\alpha}&=\left(\sum_{k}|k|^{-2\alpha}\left|\sum_{j}\widehat{b}_{k-j}\cdot ij\widehat{V}_{j}\left(|k|^{s}-|j|^{s}\right)\right|^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s}\left(\sum_{k}|k|^{-2\alpha}\left(\sum_{j}|\widehat{b}_{k-j}||j||\widehat{V}_{j}||k-j|\left(|j|^{s-1}+|k-j|^{s-1}\right)\right)^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s}\left(\sum_{k}|k|^{-2\alpha}\left(\sum_{j}|\widehat{b}_{k-j}||\widehat{V}_{j}||k-j||j|^{s}\right)^{2}\right)^{\frac{1}{2}}\\ &\qquad+\left(\sum_{k}|k|^{-2\alpha}\left(\sum_{j}|\widehat{b}_{k-j}||j||\widehat{V}_{j}||k-j|^{s}\right)^{2}\right)^{\frac{1}{2}}:=I_{1}+I_{2}.\end{split}

From |j||k|+|kj||j|\leq|k|+|k-j| and Young’s convolution inequality, we deduce:

I1=Cs(k(j|b^kj||V^j||kj|α+1|j|sα|j|α|k|α|kj|α)2)12s,α(k(j|b^kj||V^j||kj|α+1|j|sα)2)12s,α(j|b^j||j|α+1)Vsαs,α,βbα+β+1Vsα.\displaystyle\begin{split}I_{1}&=C_{s}\left(\sum_{k}\left(\sum_{j}|\widehat{b}_{k-j}||\widehat{V}_{j}||k-j|^{\alpha+1}|j|^{s-\alpha}\frac{|j|^{\alpha}}{|k|^{\alpha}|k-j|^{\alpha}}\right)^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s,\alpha}\left(\sum_{k}\left(\sum_{j}|\widehat{b}_{k-j}||\widehat{V}_{j}||k-j|^{\alpha+1}|j|^{s-\alpha}\right)^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s,\alpha}\left(\sum_{j}|\widehat{b}_{j}||j|^{\alpha+1}\right)\|V\|_{s-\alpha}\lesssim_{s,\alpha,\beta}\|b\|_{\alpha+\beta+1}\|V\|_{s-\alpha}.\end{split} (B.4)

Using Minkowski’s inequality, we split I2I_{2} into three parts:

I2s(k|k|2α(|j|12|kj||b^kj||j||V^j||kj|s)2)12+(k|k|2α(|j|2|kj||b^kj||j||V^j||kj|s)2)12+(k|k|2α(12|kj|<|j|<2|kj||b^kj||j||V^j||kj|s)2)12:=I21+I22+I23.\displaystyle\begin{split}I_{2}&\lesssim_{s}\left(\sum_{k}|k|^{-2\alpha}\left(\sum_{|j|\leq\frac{1}{2}|k-j|}|\widehat{b}_{k-j}||j||\widehat{V}_{j}||k-j|^{s}\right)^{2}\right)^{\frac{1}{2}}+\left(\sum_{k}|k|^{-2\alpha}\left(\sum_{|j|\geq 2|k-j|}|\widehat{b}_{k-j}||j||\widehat{V}_{j}||k-j|^{s}\right)^{2}\right)^{\frac{1}{2}}\\ &\qquad+\left(\sum_{k}|k|^{-2\alpha}\left(\sum_{\frac{1}{2}|k-j|<|j|<2|k-j|}|\widehat{b}_{k-j}||j||\widehat{V}_{j}||k-j|^{s}\right)^{2}\right)^{\frac{1}{2}}:=I_{21}+I_{22}+I_{23}.\end{split}

For I21I_{21}, noting when |j|12|kj||j|\leq\frac{1}{2}|k-j|, there holds

12|kj||k|32|kj|.\frac{1}{2}|k-j|\leq|k|\leq\frac{3}{2}|k-j|.

Hence, by the assumption s1s\geq 1 and Young’s convolution inequality we have

I21s,α(k(|j|12|kj||b^kj||j||V^j||kj|sα)2)12s,α(k(|j|12|kj||b^kj||V^j||kj|s|j|sα)2)12s,αj|b^j||j|sVsαs,α,βbs+βVsα.\displaystyle\begin{split}I_{21}&\lesssim_{s,\alpha}\left(\sum_{k}\left(\sum_{|j|\leq\frac{1}{2}|k-j|}|\widehat{b}_{k-j}||j||\widehat{V}_{j}||k-j|^{s-\alpha}\right)^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s,\alpha}\left(\sum_{k}\left(\sum_{|j|\leq\frac{1}{2}|k-j|}|\widehat{b}_{k-j}||\widehat{V}_{j}||k-j|^{s}|j|^{s-\alpha}\right)^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s,\alpha}\sum_{j}|\widehat{b}_{j}||j|^{s}\|V\|_{s-\alpha}\lesssim_{s,\alpha,\beta}\|b\|_{s+\beta}\|V\|_{s-\alpha}.\end{split}

Similarly, we estimate

I22s,α,βbs+βVsα.I_{22}\lesssim_{s,\alpha,\beta}\|b\|_{s+\beta}\|V\|_{s-\alpha}.

For I23I_{23}, since |kj||j||k-j|\sim|j|, we have

I23s(k(12|kj|<|j|<2|kj||b^kj||kj||V^j||j|s)2)12s(k(12|kj|<|j|<2|kj||b^kj||kj|1+α|V^j||j|sα)2)12s,α(j|b^j||j|α+1)Vsαs,α,βbα+β+1Vsα.\displaystyle\begin{split}I_{23}&\lesssim_{s}\left(\sum_{k}\left(\sum_{\frac{1}{2}|k-j|<|j|<2|k-j|}|\widehat{b}_{k-j}||k-j||\widehat{V}_{j}||j|^{s}\right)^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s}\left(\sum_{k}\left(\sum_{\frac{1}{2}|k-j|<|j|<2|k-j|}|\widehat{b}_{k-j}||k-j|^{1+\alpha}|\widehat{V}_{j}||j|^{s-\alpha}\right)^{2}\right)^{\frac{1}{2}}\\ &\lesssim_{s,\alpha}\left(\sum_{j}|\widehat{b}_{j}||j|^{\alpha+1}\right)\|V\|_{s-\alpha}\lesssim_{s,\alpha,\beta}\|b\|_{\alpha+\beta+1}\|V\|_{s-\alpha}.\end{split}

Therefore

I2s,α,βbs+α+βVsα,I_{2}\lesssim_{s,\alpha,\beta}\|b\|_{s+\alpha+\beta}\|V\|_{s-\alpha},

which together with (B.4) completes the proof.

Finally, we give a lemma regarding commutators involving the hydrostatic Leray projector 𝒫\mathcal{P} defined in (2.1).

Lemma B.4.

Assume s>2s>2. Let b=(b1,b2,b3)C(𝕋3,3)b=(b^{1},b^{2},b^{3})\in C^{\infty}(\mathbb{T}^{3},\mathbb{R}^{3}) be divergence-free and φC(𝕋3,2)\varphi\in C^{\infty}(\mathbb{T}^{3},\mathbb{R}^{2}) such that 𝒫φ=φ\mathcal{P}\varphi=\varphi. Then we have

[𝒫,b]φsbs+1φs+z(b1,b2)s1φs+1.\|[\mathcal{P},b\cdot\nabla]\varphi\|_{s}\lesssim\|b\|_{s+1}\|\varphi\|_{s}+\|\partial_{z}(b^{1},b^{2})\|_{s-1}\|\varphi\|_{s+1}.

In addition, if s>52s>\frac{5}{2}, then one has

Λ12[Λs,b][𝒫,b]φbs+32φs12+bs+3z(b1,b2)s32φs+12.\displaystyle\|\Lambda^{-\frac{1}{2}}[\Lambda^{s},b\cdot\nabla][\mathcal{P},b\cdot\nabla]\varphi\|\lesssim\|b\|_{s+3}^{2}\|\varphi\|_{s-\frac{1}{2}}+\|b\|_{s+3}\|\partial_{z}(b^{1},b^{2})\|_{s-\frac{3}{2}}\|\varphi\|_{s+\frac{1}{2}}.
Proof.

To simplify the notation, let w=0zhφ(x,z~)𝑑z~w=-\int_{0}^{z}\nabla_{h}\cdot\varphi(x^{\prime},\tilde{z})d\tilde{z} so that hφ+zw=0\nabla_{h}\cdot\varphi+\partial_{z}w=0. From the definition in (2.1), we have

[𝒫,b]φ=𝒫(bφ)b𝒫φ=(I)(bφ)bφ=(bφ).\displaystyle[\mathcal{P},b\cdot\nabla]\varphi=\mathcal{P}(b\cdot\nabla\varphi)-b\cdot\nabla\mathcal{P}\varphi=(I-\mathbb{Q})(b\cdot\nabla\varphi)-b\cdot\nabla\varphi=-\mathbb{Q}(b\cdot\nabla\varphi).

Note that (recall from (2.1) that f¯\overline{f} is the average of ff with respect to the zz variable)

(bφ)=hΔh1h(bφ¯)\displaystyle\mathbb{Q}(b\cdot\nabla\varphi)=\nabla_{h}\Delta_{h}^{-1}\nabla_{h}\cdot(\overline{b\cdot\nabla\varphi})

depends only on the horizontal variables (x1,x2)(x_{1},x_{2}). Using integration by parts, the divergence-free property of b,(φ,w)b,(\varphi,w) and the periodicity, we compute

hbφ¯\displaystyle\nabla_{h}\cdot\overline{b\cdot\nabla\varphi} =01h((b1,b2)h)φ𝑑z+01(b1,b2)h(hφ)dz+h01b3zφdz\displaystyle=\int_{0}^{1}\nabla_{h}\left((b^{1},b^{2})\cdot\nabla_{h}\right)\cdot\varphi dz+\int_{0}^{1}(b^{1},b^{2})\cdot\nabla_{h}(\nabla_{h}\cdot\varphi)dz+\nabla_{h}\cdot\int_{0}^{1}b^{3}\partial_{z}\varphi dz
=01h((b1,b2)h)φ𝑑z+01z(b1,b2)hwdz+h01φh(b1,b2)𝑑z.\displaystyle=\int_{0}^{1}\nabla_{h}\left((b^{1},b^{2})\cdot\nabla_{h}\right)\cdot\varphi dz+\int_{0}^{1}\partial_{z}(b^{1},b^{2})\cdot\nabla_{h}wdz+\nabla_{h}\cdot\int_{0}^{1}\varphi\nabla_{h}\cdot(b^{1},b^{2})dz.

When s>2s>2, the space Hhs1H_{h}^{s-1} is a Banach algebra in the 2D2D case. Since (bφ)\mathbb{Q}(b\cdot\nabla\varphi) depends only on the horizontal variables (x1,x2)(x_{1},x_{2}), applying Minkowski’s inequality and Hölder inequality, we have

(bφ)s=(bφ)Hhs\displaystyle\|\mathbb{Q}(b\cdot\nabla\varphi)\|_{s}=\|\mathbb{Q}(b\cdot\nabla\varphi)\|_{H^{s}_{h}} hbφ¯Hhs1\displaystyle\lesssim\|\nabla_{h}\cdot\overline{b\cdot\nabla\varphi}\|_{H^{s-1}_{h}}
01(b(,z)Hhs+1φ(,z)Hhs+z(b1,b2)(,z)Hhs1w(,z)Hhs)𝑑z\displaystyle\lesssim\int_{0}^{1}\left(\|b(\cdot,z)\|_{H^{s+1}_{h}}\|\varphi(\cdot,z)\|_{H^{s}_{h}}+\|\partial_{z}(b^{1},b^{2})(\cdot,z)\|_{H^{s-1}_{h}}\|w(\cdot,z)\|_{H^{s}_{h}}\right)dz
bs+1φs+z(b1,b2)s1φs+1.\displaystyle\lesssim\|b\|_{s+1}\|\varphi\|_{s}+\|\partial_{z}(b^{1},b^{2})\|_{s-1}\|\varphi\|_{s+1}.

Combining this with Lemma B.3, when s>52s>\frac{5}{2} we conclude

Λ12[Λs,b][𝒫,b]φ\displaystyle\|\Lambda^{-\frac{1}{2}}[\Lambda^{s},b\cdot\nabla][\mathcal{P},b\cdot\nabla]\varphi\| bs+3[𝒫,b]φs12\displaystyle\lesssim\|b\|_{s+3}\|[\mathcal{P},b\cdot\nabla]\varphi\|_{s-\frac{1}{2}}
bs+3(bs+12φs12+z(b1,b2)s32φs+12)\displaystyle\lesssim\|b\|_{s+3}\left(\|b\|_{s+\frac{1}{2}}\|\varphi\|_{s-\frac{1}{2}}+\|\partial_{z}(b^{1},b^{2})\|_{s-\frac{3}{2}}\|\varphi\|_{s+\frac{1}{2}}\right)
bs+32φs12+bs+3z(b1,b2)s32φs+12.\displaystyle\lesssim\|b\|_{s+3}^{2}\|\varphi\|_{s-\frac{1}{2}}+\|b\|_{s+3}\|\partial_{z}(b^{1},b^{2})\|_{s-\frac{3}{2}}\|\varphi\|_{s+\frac{1}{2}}.

References

  • [1] Elie Abdo and Quyuan Lin. On the Dirichlet fractional Laplacian and applications to the SQG equation on bounded domains. arXiv preprint arXiv:2409.05209, 2024.
  • [2] Elie Abdo, Quyuan Lin, and Changhui Tan. Well-posedness and ill-posedness of the primitive equations with fractional horizontal dissipation. preprint, 2024.
  • [3] Robert A Adams and John JF Fournier. Sobolev spaces. Elsevier, 2003.
  • [4] Antonio Agresti. The primitive equations with rough transport noise: Global well-posedness and regularity. arXiv preprint arXiv:2310.01193, 2023.
  • [5] Antonio Agresti. Global smooth solutions by transport noise of 3D Navier-Stokes equations with small hyperviscosity. arXiv preprint arXiv:2406.09267, 2024.
  • [6] Antonio Agresti, Matthias Hieber, Amru Hussein, and Martin Saal. The stochastic primitive equations with non-isothermal turbulent pressure. arXiv preprint arXiv:2210.05973, 2022.
  • [7] Antonio Agresti, Matthias Hieber, Amru Hussein, and Martin Saal. The stochastic primitive equations with transport noise and turbulent pressure. Stochastics and Partial Differential Equations: Analysis and Computations, 12(1):53–133, 2024.
  • [8] Diego Alonso-Orán and Aythami Bethencourt de León. On the well-posedness of stochastic Boussinesq equations with transport noise. Journal of Nonlinear Science, 30(1):175–224, 2020.
  • [9] Pascal Azérad and Francisco Guillén. Mathematical justification of the hydrostatic approximation in the primitive equations of geophysical fluid dynamics. SIAM journal on mathematical analysis, 33(4):847–859, 2001.
  • [10] Zdzislaw Brzeźniak and Jakub Slavík. Well-posedness of the 3D stochastic primitive equations with multiplicative and transport noise. Journal of Differential Equations, 296:617–676, 2021.
  • [11] Luis A Caffarelli. Some nonlinear problems involving non-local diffusions. In ICIAM 07-6th International Congress on Industrial and Applied Mathematics, Eur. Math. Soc., Zürich, pages 43–56, 2009.
  • [12] Luis A Caffarelli and Alexis Vasseur. Drift diffusion equations with fractional diffusion and the quasi-geostrophic equation. Annals of Mathematics, pages 1903–1930, 2010.
  • [13] Chongsheng Cao, Slim Ibrahim, Kenji Nakanishi, and Edriss S Titi. Finite-time blowup for the inviscid primitive equations of oceanic and atmospheric dynamics. Communications in Mathematical Physics, 337(2):473–482, 2015.
  • [14] Chongsheng Cao and Edriss S Titi. Global well-posedness of the three-dimensional viscous primitive equations of large scale ocean and atmosphere dynamics. Annals of Mathematics, pages 245–267, 2007.
  • [15] José Antonio Carrillo, Manuel del Pino, Alessio Figalli, Giuseppe Mingione, Juan Luis Vázquez, and Juan Luis Vázquez. The mathematical theories of diffusion: nonlinear and fractional diffusion. Nonlocal and Nonlinear Diffusions and Interactions: New Methods and Directions: Cetraro, Italy 2016, pages 205–278, 2017.
  • [16] Dongho Chae, Peter Constantin, Diego Córdoba, Francisco Gancedo, and Jiahong Wu. Generalized surface quasi-geostrophic equations with singular velocities. Communications on Pure and Applied Mathematics, 65(8):1037–1066, 2012.
  • [17] Charles Collot, Slim Ibrahim, and Quyuan Lin. Stable singularity formation for the inviscid primitive equations. Annales de l’Institut Henri Poincaré C, 41(2):317–356, 2023.
  • [18] Peter Constantin, Nathan Glatt-Holtz, and Vlad Vicol. Unique ergodicity for fractionally dissipated, stochastically forced 2D Euler equations. Communications in Mathematical Physics, 330:819–857, 2014.
  • [19] Peter Constantin, Andrei Tarfulea, and Vlad Vicol. Long time dynamics of forced critical SQG. Communications in Mathematical Physics, 335:93–141, 2015.
  • [20] Peter Constantin and Vlad Vicol. Nonlinear maximum principles for dissipative linear nonlocal operators and applications. Geometric And Functional Analysis, 22:1289–1321, 2012.
  • [21] Dan Crisan, Franco Flandoli, and Darryl D Holm. Solution properties of a 3D stochastic Euler fluid equation. Journal of Nonlinear Science, 29(3):813–870, 2019.
  • [22] Yichen Dai, Zhong Tan, and Jiahong Wu. A class of global large solutions to the magnetohydrodynamic equations with fractional dissipation. Zeitschrift für angewandte Mathematik und Physik, 70:1–13, 2019.
  • [23] Arnaud Debussche, Nathan Glatt-Holtz, and Roger Temam. Local martingale and pathwise solutions for an abstract fluids model. Physica D: Nonlinear Phenomena, 240(14-15):1123–1144, 2011.
  • [24] Arnaud Debussche, Nathan Glatt-Holtz, Roger Temam, and Mohammed Ziane. Global existence and regularity for the 3D stochastic primitive equations of the ocean and atmosphere with multiplicative white noise. Nonlinearity, 25(7):2093, 2012.
  • [25] Arnaud Debussche and Umberto Pappalettera. Second order perturbation theory of two-scale systems in fluid dynamics. Journal of the European Mathematical Society, 2024.
  • [26] Joseph L Doob. Measure theory, volume 143. Springer Science & Business Media, 2012.
  • [27] Franco Flandoli. Random perturbation of PDEs and fluid dynamic models: École d’été de Probabilités de Saint-Flour XL–2010, volume 2015. Springer Science & Business Media, 2011.
  • [28] Franco Flandoli, Lucio Galeati, and Dejun Luo. Delayed blow-up by transport noise. Communications in Partial Differential Equations, 46(9):1757–1788, 2021.
  • [29] Franco Flandoli and Dariusz Gatarek. Martingale and stationary solutions for stochastic Navier-Stokes equations. Probability Theory and Related Fields, 102(3):367–391, 1995.
  • [30] Franco Flandoli and Dejun Luo. High mode transport noise improves vorticity blow-up control in 3D Navier–Stokes equations. Probability Theory and Related Fields, 180:309–363, 2021.
  • [31] Franco Flandoli and Umberto Pappalettera. From additive to transport noise in 2d fluid dynamics. Stochastics and Partial Differential Equations: Analysis and Computations, 10(3):964–1004, 2022.
  • [32] Tej Eddine Ghoul, Slim Ibrahim, Quyuan Lin, and Edriss S Titi. On the effect of rotation on the life-span of analytic solutions to the 3D inviscid primitive equations. Archive for rational mechanics and analysis, 243(2):747–806, 2022.
  • [33] Daniel Han-Kwan and Toan T Nguyen. Ill-posedness of the hydrostatic Euler and singular Vlasov equations. Archive for Rational Mechanics and Analysis, 221(3):1317–1344, 2016.
  • [34] Lars Hörmander. The analysis of linear partial differential operators III: Pseudo-differential operators. Springer Science & Business Media, 2007.
  • [35] Ruimeng Hu and Quyuan Lin. Local martingale solutions and pathwise uniqueness for the three-dimensional stochastic inviscid primitive equations. Stochastics and Partial Differential Equations: Analysis and Computations, 11(4):1470–1518, 2023.
  • [36] Ruimeng Hu and Quyuan Lin. Pathwise solutions for stochastic hydrostatic Euler equations and hydrostatic Navier-Stokes equations under the local Rayleigh condition. arXiv preprint arXiv:2301.07810, 2023.
  • [37] Ruimeng Hu, Quyuan Lin, and Rongchang Liu. Regularization by noise for the inviscid primitive equations. arXiv preprint arXiv:2407.21336, 2024.
  • [38] Slim Ibrahim, Quyuan Lin, and Edriss S Titi. Finite-time blowup and ill-posedness in Sobolev spaces of the inviscid primitive equations with rotation. Journal of Differential Equations, 286:557–577, 2021.
  • [39] Jong Uhn Kim. Existence of a local smooth solution in probability to the stochastic Euler equations in R3R^{3}. Journal of Functional Analysis, 256(11):3660–3687, 2009.
  • [40] Alexander Kiselev, Fedor Nazarov, and Alexander Volberg. Global well-posedness for the critical 2D dissipative quasi-geostrophic equation. Inventiones Mathematicae, 167(3):445–453, 2007.
  • [41] Georgij M Kobelkov. Existence of a solution ‘in the large’ for the 3D large-scale ocean dynamics equations. Comptes Rendus Mathematique, 343(4):283–286, 2006.
  • [42] Robert H Kraichnan. Small-scale structure of a scalar field convected by turbulence. The Physics of Fluids, 11(5):945–953, 1968.
  • [43] Robert H Kraichnan. Anomalous scaling of a randomly advected passive scalar. Physical review letters, 72(7):1016, 1994.
  • [44] Igor Kukavica, Roger Temam, Vlad C Vicol, and Mohammed Ziane. Local existence and uniqueness for the hydrostatic Euler equations on a bounded domain. Journal of Differential Equations, 250(3):1719–1746, 2011.
  • [45] Thomas Kurtz and Philip Protter. Weak convergence of stochastic integrals and differential equations II: Infinite dimensional case. Probabilistic Models for Nonlinear Partial Differential Equations, pages 197–285, 1996.
  • [46] Thomas G Kurtz. The Yamada-Watanabe-Engelbert theorem for general stochastic equations and inequalities. Electronic Journal of Probability [electronic only], 12:951–965, 2007.
  • [47] Oana Lang and Dan Crisan. Well-posedness for a stochastic 2D Euler equation with transport noise. Stochastics and Partial Differential Equations: Analysis and Computations, 11(2):433–480, 2023.
  • [48] Dong Li. On Kato–Ponce and fractional Leibniz. Revista matemática iberoamericana, 35(1):23–100, 2019.
  • [49] Jinkai Li and Edriss S Titi. The primitive equations as the small aspect ratio limit of the Navier–Stokes equations: Rigorous justification of the hydrostatic approximation. Journal de Mathématiques Pures et Appliquées, 124:30–58, 2019.
  • [50] Dejun Luo. Enhanced dissipation for stochastic Navier–Stokes equations with transport noise. Journal of Dynamics and Differential Equations, pages 1–36, 2023.
  • [51] Remigijus Mikulevicius and B Rozovskii. On equations of stochastic fluid mechanics. In Stochastics in Finite and Infinite Dimensions: In Honor of Gopinath Kallianpur, pages 285–302. Springer, 2001.
  • [52] Remigijus Mikulevicius and Boris L Rozovskii. Stochastic Navier–Stokes equations for turbulent flows. SIAM Journal on Mathematical Analysis, 35(5):1250–1310, 2004.
  • [53] Michael Renardy. Ill-posedness of the hydrostatic Euler and Navier–Stokes equations. Archive for rational mechanics and analysis, 194(3):877–886, 2009.
  • [54] Jacques Simon. Compact sets in the space Lp(0,T;B)L^{p}(0,T;B). Annali di Matematica pura ed applicata, 146(1):65–96, 1986.
  • [55] Jacques Simon. Sobolev, Besov and Nikolskii fractional spaces: imbeddings and comparisons for vector valued spaces on an interval. Annali di Matematica Pura ed Applicata, 157(1):117–148, 1990.
  • [56] Hao Tang and Feng-Yu Wang. A general framework for solving singular SPDEs with applications to fluid models driven by pseudo-differential noise. arXiv preprint arXiv:2208.08312, 2022.
  • [57] Michael Eugene Taylor. Pseudodifferential operators (PMS-34). In Pseudodifferential Operators (PMS-34). Princeton University Press, 2017.
  • [58] Tak Kwong Wong. Blowup of solutions of the hydrostatic Euler equations. Proceedings of the American Mathematical Society, 143(3):1119–1125, 2015.
  • [59] Jiahong Wu. The 2D magnetohydrodynamic equations with partial or fractional dissipation. Lectures on the analysis of nonlinear partial differential equations, Morningside Lectures on Mathematics, Part, 5:283–332, 2018.
  • [60] Jiahong Wu and Xiaojing Xu. Well-posedness and inviscid limits of the Boussinesq equations with fractional Laplacian dissipation. Nonlinearity, 27(9):2215, 2014.
  • [61] Wanrong Yang, Quansen Jiu, and Jiahong Wu. Global well-posedness for a class of 2D Boussinesq systems with fractional dissipation. Journal of Differential Equations, 257(11):4188–4213, 2014.