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

\calclayout\numberwithin

equationsection \allowdisplaybreaks\theoremstyleplain

\address

Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing 100190, China
and
LMAM, School of Mathematical Sciences, Peking University, Beijing, 100871, China \emailchenhp@amss.ac.cn

\address

School of Mathematics and Statistics, Jiangxi Normal University, Nanchang, Jiangxi 330022, China \emailzhishi@pku.org.cn; chenyong77@gmail.com

\address

LMAM, School of Mathematical Sciences, Peking University, Beijing, 100871, China \emailliuyong@math.pku.edu.cn

\subjclass

[2020]Primary 60H17; Secondary 37A25

\keywords

Complex Ginzburg-Landau equation, ergodicity, space-time white noise, stochastic heat equation with dispersion, complex Wiener-Itô integral
We thank Prof. Deng Zhang’s suggestion to pay attention to [24, 25, 34]. Y. Chen is supported by NSFC (No. 11961033). Y. Liu is supported by NSFC (No. 11731009, No. 12231002) and Center for Statistical Science, PKU

Ergodicity for Ginzburg-Landau equation with complex-valued space-time white noise on two-dimensional torus

Huiping Chen    Yong Chen    Yong Liu
Abstract

We investigate the ergodicity for the stochastic complex Ginzburg-Landau equation with a general non-linear term on the two-dimensional torus driven by a complex-valued space-time white noise. Due to the roughness of complex-valued space-time white noise, this equation is a singular stochastic partial differential equation and its solution is expected to be a distribution-valued stochastic process. For this reason, the non-linear term is ill-defined and needs to be renormalized. We first use the theory of complex multiple Wiener-Itô integral to renormalize this equation and then consider its global well-posedness. Further, we prove its ergodicity using an asymptotic coupling argument for a large dissipation coefficient.

1 Introduction

We are interested in the stochastic complex Ginzburg-Landau equation on the two-dimensional torus defined as

{tu=(i+μ)Δuν|u|2mu+τu+ξ,t>0,x𝕋2,u(0,)=u0,\begin{cases}\partial_{t}u=(\mathrm{i}+\mu)\Delta u-\nu|u|^{2m}u+\tau u+\xi,&t>0,x\in\mathbb{T}^{2},\\ u(0,\cdot)=u_{0},\end{cases} (1)

where μ>0\mu>0, ν,τ\nu,\tau\in\mathbb{C}, Reν>0\mathrm{Re}\,\nu>0, m1m\geq 1 is an integer and ξ\xi is the complex-valued space-time white noise as introduced in Section 2. The definition of ξ\xi can be found in (3), and a construction for the complex-valued space-time white noise is provided in (5). Here, we call iΔu\mathrm{i}\Delta u the dispersion term and μΔu\mu\Delta u the dissipation term. We aim to get its global well-posedness and ergodicity.

The complex Ginzburg-Landau equation is one of the most crucial nonlinear partial differential equations (PDEs) which describes various physical phenomena such as nonlinear waves and superconducting phase transition and superfluidity, see [1]. In addition to wide applications in physics, it has rich research background in mathematics as a parabolic equation. In [29], Parisi and Wu proposed to construct a quantum field theory based on the invariant measure of a random process, which is called stochastic quantisation. We establish the global well-posedness and ergodicity of (1), which is important to study the complex-valued Φ22(m+1)\Phi_{2}^{2(m+1)} measure in quantum field theory.

There has been a substantial amount of research on complex Ginzburg-Landau equations with random forcing. For non-white or multiplicative noise, see [3, 4, 21, 28] for example. In [15], Hairer studied the complex cubic Ginzburg-Landau equation driven by a real-valued space-time white noise in one spatial dimension. When the spatial dimension is larger than one, due to the roughness of complex-valued space-time white noise, complex Ginzburg-Landau equation driven by complex-valued space-time white noise is a singular stochastic partial differential equation and its solution is expected to be a distribution-valued stochastic process. For this reason, the non-linear term |u|2mu|u|^{2m}u is ill-defined and needs to be renormalized. The local and global well-posedness of the complex Ginzburg-Landau equation on the three-dimensional torus driven by complex-valued space-time white noise are obtained in [18, 19] using the theory of regularity structure (see [16]) and paracontrolled distribution (see [14]).

For two-dimensional case, in [34], Trenberth worked with the generalized Laguerre polynomials for the renormalization of (1) and obtained its global well-posedness. Different from the way of renormalization using generalized Laguerre polynomials in [34], we utilize complex multiple Wiener-Itô integral to renormalize (1) more clearly in Section 3. We show the regularity of stochastic heat equation with dispersion (see Proposition 3.1) combining the fact that a complex multiple Wiener-Itô integral can be viewed as a two-dimensional real multiple Wiener-Itô integral (see [9, Theorem 3.3]), with [27, Proposition 5], which concerning the regularity of real-valued random processes. This idea that utilizing the theory of two-dimensional real Wiener-Itô integral to solve the theoretical and practical problems involving complex Wiener-Itô integral, is what we have always wanted to express to the reader in [7] and [8]. Compared to [25] where the regularity is obtained using [26, Lemma 9], our proof is shorter and more concise. Motivated by [35], Matsuda in [25] proved the global well-posedness and strong Feller property of (1) for m=1m=1. Combining with the support theorem proved in [24], Matsuda got its exponential ergodicity in [25]. In our setting, we consider a general nonlinearity (m1m\geq 1) and prove the ergodicity of (1) using an alternative approach, namely an asymptotic coupling argument (see [17]), under the assumption that dissipation coefficient μ\mu is greater than some constant (see (68)).

This paper is organized as follows. Section 2 introduces some notions of Besov space and complex-valued space-time white noise. In Section 3, we show the regularity of Wick product associated with stochastic heat equation with dispersion (6). In Section 4, inspired by the Da Prato-Debussche method (see [10]) and combining with the priori estimates in Section 3, we obtain the global well-posedness of the stochastic complex Ginzburg-Landau equation (1) in the sense of renormalization. In Section 5, by the asymptotic (or generalized) coupling approach developed in [17, 22], we show the existence and uniqueness of the invariant measure of the renormalization solution to (1).

2 Preliminaries

To characterize the regularity of the stationary solution of the stochastic heat equation with dispersion (6), we introduce the Besov space. We refer readers to [2, Chapter 2] and [13, Chapter 3] for more details about the Besov space and Fourier analysis on the dd-dimensional torus 𝕋d\mathbb{T}^{d}, where d1d\geq 1 is an integer. We denote by C(𝕋d,)C^{\infty}\left(\mathbb{T}^{d},\mathbb{C}\right) the space of all complex-valued smooth functions defined over 𝕋d\mathbb{T}^{d}. For p1p\geq 1, let Lp:=Lp(𝕋d,)L^{p}:=L^{p}\left(\mathbb{T}^{d},\mathbb{C}\right) and Lp(×𝕋d,)L^{p}\left(\mathbb{R}\times\mathbb{T}^{d},\mathbb{C}\right) be the space of all complex-valued functions that are pp-integrable over 𝕋d\mathbb{T}^{d} and ×𝕋d\mathbb{R}\times\mathbb{T}^{d}, respectively. For f,gL2(G,)f,g\in L^{2}(G,\mathbb{C}), where G=𝕋d or ×𝕋dG=\mathbb{T}^{d}\mbox{ or }\mathbb{R}\times\mathbb{T}^{d}, we set

f,g=Gf(x)g(x)dx\left\langle f,g\right\rangle=\int_{G}f(x)g(x)\mathrm{d}x

in this chapter. Note that we do not take complex conjugate for gg.

For any fL1f\in L^{1} and kdk\in\mathbb{Z}^{d}, we write

f(k)=f^(k):=𝕋df(x)e2πikxdx,\mathscr{F}f(k)=\hat{f}(k):=\int_{\mathbb{T}^{d}}f(x)e^{-2\pi\mathrm{i}k\cdot x}\mathrm{d}x,

for the Fourier coefficient of ff with frequency kk.

We now introduce distributions on the torus. The set of test functions on the torus is defined as

𝒮(𝕋d,):={fC(𝕋d,):Mα<, s.t., αfLMα<,αd},\mathscr{S}\left(\mathbb{T}^{d},\mathbb{C}\right):=\left\{f\in{C}^{\infty}\left(\mathbb{T}^{d},\mathbb{C}\right):\exists M_{\alpha}<\infty,\text{ s.t., }\left\|\partial^{\alpha}f\right\|_{L^{\infty}}\leq M_{\alpha}<\infty,\alpha\in\mathbb{N}^{d}\right\},

where MαM_{\alpha} is a constant depending on the multi-index α\alpha. We define the semi-norms as

fm:=sup|α|mαfL,m,\left\|f\right\|_{m}:=\sup_{|\alpha|\leq m}\left\|\partial^{\alpha}f\right\|_{L^{\infty}},\quad m\in\mathbb{N},

where for a multi-index α=(α1,,αd)d\alpha=\left(\alpha_{1},\ldots,\alpha_{d}\right)\in\mathbb{N}^{d}, |α|=α1++αd|\alpha|=\alpha_{1}+\cdots+\alpha_{d}. The space of all complex-valued linear and continuous functionals on 𝒮(𝕋d,)\mathscr{S}\left(\mathbb{T}^{d},\mathbb{C}\right) is denoted by 𝒮(𝕋d,)\mathscr{S}^{\prime}\left(\mathbb{T}^{d},\mathbb{C}\right) and is called the space of distributions. For u𝒮(𝕋d,)u\in\mathscr{S}^{\prime}\left(\mathbb{T}^{d},\mathbb{C}\right), we define the Fourier coefficient with frequency kdk\in\mathbb{Z}^{d} as

u^(k):=u,e2πik(),\hat{u}(k):=\left\langle u,e^{-2\pi\mathrm{i}k\cdot(\cdot)}\right\rangle,

where u,\left\langle u,\cdot\right\rangle stands for the action of uu on 𝒮(𝕋d,)\mathscr{S}\left(\mathbb{T}^{d},\mathbb{C}\right). This is well-defined since e2πik()𝒮(𝕋d,)e^{-2\pi\mathrm{i}k\cdot(\cdot)}\in\mathscr{S}\left(\mathbb{T}^{d},\mathbb{C}\right).

We denote by {χj}j=1\left\{\chi_{j}\right\}_{j=-1}^{\infty} a dyadic partition of unity satisfying the following properties

  • χ1,χ0:d[0,1]\chi_{-1},\chi_{0}:\mathbb{R}^{d}\rightarrow[0,1] are radial and infinitely differentiable;

  • supp(χ1)B(0,43)\mathrm{supp}(\chi_{-1})\subseteq B(0,\frac{4}{3}) and supp(χ0)B(0,83)B(0,34)\mathrm{supp}(\chi_{0})\subseteq B(0,\frac{8}{3})\setminus B(0,\frac{3}{4}), where B(0,r):={xd:|x|<r}B(0,r):=\left\{x\in\mathbb{R}^{d}:|x|<r\right\};

  • for j0j\geq 0, χj()=χ0(/2j)\chi_{j}(\cdot)=\chi_{0}(\cdot/2^{j});

  • for any xdx\in\mathbb{R}^{d}, j=1χj(x)=1\sum_{j=-1}^{\infty}\chi_{j}(x)=1.

The existence of dyadic partitions of unity can be found in [2, Proposition 2.10].

For any fC(𝕋d,)f\in C^{\infty}(\mathbb{T}^{d},\mathbb{C}) and j1j\geq-1, we define the jj-th Littlewood-Paley block as

δjf:=kdχj(k)f^(k)e2πikx.\delta_{j}f:=\sum_{k\in\mathbb{Z}^{d}}\chi_{j}(k)\hat{f}(k)e^{2\pi\mathrm{i}k\cdot x}.

The space of p,qα\mathcal{B}_{p,q}^{\alpha} for α\alpha\in\mathbb{R} and p,q[1,]p,q\in[1,\infty] is defined as the completion of C(𝕋d,)C^{\infty}\left(\mathbb{T}^{d},\mathbb{C}\right) with respect to the norm

fp,qα:=(supj12αjqδjfLpq)1q.\|f\|_{\mathcal{B}_{p,q}^{\alpha}}:=\left(\sup_{j\geq-1}2^{\alpha jq}\left\|\delta_{j}f\right\|_{L^{p}}^{q}\right)^{\frac{1}{q}}.

When p=q=p=q=\infty, we denote by 𝒞α\mathcal{C}^{\alpha} for α\alpha\in\mathbb{R} the space ,α\mathcal{B}_{\infty,\infty}^{\alpha}. That is, the space 𝒞α\mathcal{C}^{\alpha} for α\alpha\in\mathbb{R} is defined as the completion of C(𝕋d,)C^{\infty}\left(\mathbb{T}^{d},\mathbb{C}\right) with respect to the norm

f𝒞α:=supj12αjδjfL.\|f\|_{\mathcal{C}^{\alpha}}:=\sup_{j\geq-1}2^{\alpha j}\left\|\delta_{j}f\right\|_{L^{\infty}}.

This space is a subspace of 𝒮(𝕋d,)\mathscr{S}^{\prime}\left(\mathbb{T}^{d},\mathbb{C}\right). We set 𝒞α:=ϵ>0𝒞αϵ\mathcal{C}^{\alpha-}:=\bigcap_{\epsilon>0}\mathcal{C}^{\alpha-\epsilon} for α\alpha\in\mathbb{R}.

We refer reader to [2, 26, 33] for the following properties of the Besov space.

Proposition 2.1

Let α1,α2,p1,p2,q1,q2[1,]\alpha_{1},\alpha_{2}\in\mathbb{R},p_{1},p_{2},q_{1},q_{2}\in[1,\infty]. Then,

fp1,q1α1\displaystyle\|f\|_{\mathcal{B}_{p_{1},q_{1}}^{\alpha_{1}}} Cfp1,q1α2, whenever α1α2,\displaystyle\leq C\|f\|_{\mathcal{B}_{p_{1},q_{1}}^{\alpha_{2}}},\quad\text{ whenever }\alpha_{1}\leq\alpha_{2},
fp1,q1α1\displaystyle\|f\|_{\mathcal{B}_{p_{1},q_{1}}^{\alpha_{1}}} fp1,q2α1, whenever q1q2,\displaystyle\leq\|f\|_{\mathcal{B}_{p_{1},q_{2}}^{\alpha_{1}}},\quad\text{ whenever }q_{1}\geq q_{2},
fp1,q1α1\displaystyle\|f\|_{\mathcal{B}_{p_{1},q_{1}}^{\alpha_{1}}} Cfp2,q1α1, whenever p1p2,\displaystyle\leq C\|f\|_{\mathcal{B}_{p_{2},q_{1}}^{\alpha_{1}}},\quad\text{ whenever }p_{1}\leq p_{2},
fp1,q1α1\displaystyle\|f\|_{\mathcal{B}_{p_{1},q_{1}}^{\alpha_{1}}} Cfp1,q2α2, whenever α1<α2.\displaystyle\leq C\|f\|_{\mathcal{B}_{p_{1},q_{2}}^{\alpha_{2}}},\quad\text{ whenever }\alpha_{1}<\alpha_{2}.
Proposition 2.2

Let p[1,]p\in[1,\infty]. Then the space p,10\mathcal{B}_{p,1}^{0} is continuously embedded in LpL^{p} and

fLpfp,10.\|f\|_{L^{p}}\leq\|f\|_{\mathcal{B}_{p,1}^{0}}.

On the other hand, LpL^{p} is continuously embedded in p,0\mathcal{B}_{p,\infty}^{0} and

fp,0CfLp.\|f\|_{\mathcal{B}_{p,\infty}^{0}}\leq C\|f\|_{L^{p}}.
Proposition 2.3

Let α\alpha\in\mathbb{R}, r[1,]r\in[1,\infty], 1qp1\leq q\leq p\leq\infty and β=α+d(1q1p)\beta=\alpha+d\left(\frac{1}{q}-\frac{1}{p}\right). Then

fp,rαCfq,rα+d(1q1p).\|f\|_{\mathcal{B}_{p,r}^{\alpha}}\leq C\|f\|_{\mathcal{B}_{q,r}^{\alpha+d\left(\frac{1}{q}-\frac{1}{p}\right)}}.

The smoothing effect of the heat semigroup (Pt=eAt)t0\left(P_{t}=\mathrm{e}^{At}\right)_{t\geq 0} with generator A:=(i+μ)Δ1A:=\left(\mathrm{i}+\mu\right)\Delta-1 is shown in the following proposition.

Proposition 2.4

Let fp,qαf\in\mathcal{B}_{p,q}^{\alpha} for some α\alpha\in\mathbb{R} and p,q[1,]p,q\in[1,\infty]. Then, for all β\beta\in\mathbb{R} and βα\beta\geq\alpha,

Ptfp,qβCtαβ2fp,qα,\left\|P_{t}f\right\|_{\mathcal{B}_{p,q}^{\beta}}\leq Ct^{\frac{\alpha-\beta}{2}}\|f\|_{\mathcal{B}_{p,q}^{\alpha}},

for every t>0t>0.

Proposition 2.5

Let α,β\alpha,\beta\in\mathbb{R} be such that 0βα20\leq\beta-\alpha\leq 2 and p,q[1,]p,q\in[1,\infty]. Then, there exists a constant C(0,)C\in(0,\infty) such that for t0t\geq 0,

(1Pt)fp,qαCtβα2fp,qβ.\left\|\left(1-P_{t}\right)f\right\|_{\mathcal{B}_{p,q}^{\alpha}}\leq Ct^{\frac{\beta-\alpha}{2}}\|f\|_{\mathcal{B}_{p,q}^{\beta}}.

The above proposition shows that, if 0<βα20<\beta-\alpha\leq 2 and fp,qβf\in\mathcal{B}_{p,q}^{\beta}, then the mapping tPtft\mapsto P_{t}f is continuous as a mapping from [0,)[0,\infty) to p,qα\mathcal{B}_{p,q}^{\alpha}. When β=α\beta=\alpha, this is still valid. That is, if α\alpha\in\mathbb{R} and fp,qαf\in\mathcal{B}_{p,q}^{\alpha}, then the mapping tPtft\mapsto P_{t}f is continuous as a mapping from [0,)[0,\infty) to p,qα\mathcal{B}_{p,q}^{\alpha}.

The multiplicative structure and duality property of Besov space are presented in the following two propositions.

Proposition 2.6

Let α,β\alpha,\beta\in\mathbb{R} satisfy α+β>0\alpha+\beta>0 and p,p1,p2,q[1,]p,p_{1},p_{2},q\in[1,\infty] be such that 1p=1p1+1p2\frac{1}{p}=\frac{1}{p_{1}}+\frac{1}{p_{2}}. The mapping (f,g)fg(f,g)\mapsto fg can be extended to a continuous linear map from p1,qα×p2,qβ\mathcal{B}_{p_{1},q}^{\alpha}\times\mathcal{B}_{p_{2},q}^{\beta} to p,qαβ\mathcal{B}_{p,q}^{\alpha\wedge\beta} and for (f,g)p1,qα×p2,qβ(f,g)\in\mathcal{B}_{p_{1},q}^{\alpha}\times\mathcal{B}_{p_{2},q}^{\beta},

fgp,qαβCfp1,qαgp2,qβ.\left\|fg\right\|_{\mathcal{B}_{p,q}^{\alpha\wedge\beta}}\leq C\left\|f\right\|_{\mathcal{B}_{p_{1},q}^{\alpha}}\left\|g\right\|_{\mathcal{B}_{p_{2},q}^{\beta}}.
Proposition 2.7

Let α[0,1)\alpha\in[0,1) and p,q,p,q[1,]p,q,p^{\prime},q^{\prime}\in[1,\infty] be such that 1=1p+1p=1q+1q1=\frac{1}{p}+\frac{1}{p^{\prime}}=\frac{1}{q}+\frac{1}{q^{\prime}}. Then the mapping (f,g)f,g(f,g)\mapsto\left\langle f,g\right\rangle extends to a continuous bilinear form on p,qα×p,qα\mathcal{B}_{p,q}^{\alpha}\times\mathcal{B}_{p^{\prime},q^{\prime}}^{-\alpha} and for all (f,g)p,qα×p,qα(f,g)\in\mathcal{B}_{p,q}^{\alpha}\times\mathcal{B}_{p^{\prime},q^{\prime}}^{-\alpha}

|f,g|Cfp,qαgp,qα.|\langle f,g\rangle|\leq C\|f\|_{\mathcal{B}_{p,q}^{\alpha}}\|g\|_{\mathcal{B}_{p^{\prime},q^{\prime}}^{-\alpha}}.

For every α(0,1]\alpha\in(0,1], the following gradient estimate shows that the Besov norm 1,1α\left\|\cdot\right\|_{\mathcal{B}_{1,1}^{\alpha}} can be controlled by the norms L1\|\cdot\|_{L^{1}} and ()L1\|\nabla(\cdot)\|_{L^{1}}.

Proposition 2.8

Let α(0,1)\alpha\in(0,1) and f1,1αf\in\mathcal{B}_{1,1}^{\alpha}. Then

f1,1αC(fL11αfL1α+fL1).\|f\|_{\mathcal{B}_{1,1}^{\alpha}}\leq C\left(\|f\|_{L^{1}}^{1-\alpha}\|\nabla f\|_{L^{1}}^{\alpha}+\|f\|_{L^{1}}\right).

In particular, by [26, Remark 19], there exists C<C<\infty such that

f1,1αC(fL1+fL1).\|f\|_{\mathcal{B}_{1,1}^{\alpha}}\leq C\left(\|\nabla f\|_{L^{1}}+\|f\|_{L^{1}}\right). (2)

Set Λ=(A)12\Lambda=(-A)^{\frac{1}{2}}. For s0s\geq 0 and 1p1\leq p\leq\infty, we denote by HpsH_{p}^{s} the subspace of LpL^{p}, consisting of all ff that can be written as f=Λsgf=\Lambda^{-s}g, where gLpg\in L^{p}. The HpsH_{p}^{s} norm of ff is defined as the LpL^{p} norm of gg, that is, fHps=ΛsgLp\left\|f\right\|_{H_{p}^{s}}=\left\|\Lambda^{s}g\right\|_{L^{p}}.

Recall the following Sobolev embedding results for the Sobolev space HpsH_{p}^{s} (see [5, Theorem 6.4.4] and [33, Theorem 2.13]).

Proposition 2.9
  1. (i)

    Let s0s\geq 0, 1<p<1<p<\infty and ϵ>0\epsilon>0. Then Hps+ϵp,1s1,1sH_{p}^{s+\epsilon}\subset\mathcal{B}_{p,1}^{s}\subset\mathcal{B}_{1,1}^{s} and the embedding is continuous and dense.

  2. (ii)

    Let α\alpha\in\mathbb{R} and 1p1p21\leq p_{1}\leq p_{2}\leq\infty. Then Hp1αH_{p_{1}}^{\alpha} is continuously embedded in Hp2αd(1p11p2)H_{p_{2}}^{\alpha-d(\frac{1}{p_{1}}-\frac{1}{p_{2}})}.

The following interpolation inequality and multiplicative inequality for the Sobolev space HpsH_{p}^{s} can be found in [5, Theorem 6.4.5 ], [30, Lemma A.4] and [31, Lemma 2.1].

Proposition 2.10
  1. (i)

    Let s(0,1)s\in(0,1) and p(1,)p\in(1,\infty). Then for uHp1u\in H_{p}^{1},

    uHpsuLp1suHp1s.\|u\|_{H_{p}^{s}}\lesssim\|u\|_{L^{p}}^{1-s}\|u\|_{H_{p}^{1}}^{s}.
  2. (ii)

    Let s>0s>0 and p(1,)p\in(1,\infty). If uLp1Hp4su\in L^{p_{1}}\cap H^{s}_{p_{4}} and vLp3Hp2sv\in L^{p_{3}}\cap H^{s}_{p_{2}}, where pi(1,],i=1,,4p_{i}\in(1,\infty],i=1,\ldots,4, such that 1p=1p1+1p2=1p3+1p4\frac{1}{p}=\frac{1}{p_{1}}+\frac{1}{p_{2}}=\frac{1}{p_{3}}+\frac{1}{p_{4}}, then

    uvHpsuLp1vHp2s+vLp3uHp4s.\left\|uv\right\|_{H^{s}_{p}}\lesssim\|u\|_{L^{p_{1}}}\left\|v\right\|_{H_{p_{2}}^{s}}+\|v\|_{L^{p_{3}}}\left\|u\right\|_{H^{s}_{p_{4}}}.

Next, we introduce the definition of complex-valued space-time white noise over ×𝕋d\mathbb{R}\times\mathbb{T}^{d}, where d1d\geq 1 is an integer. A complex-valued space-time white noise defined on the probability space (Ω,,P)(\Omega,\mathcal{F},P) over ×𝕋d\mathbb{R}\times\mathbb{T}^{d} is a family of centered complex Gaussian random variables {ξ(φ):φL2(×𝕋d,)}\left\{\xi(\varphi):\varphi\in L^{2}\left(\mathbb{R}\times\mathbb{T}^{d},\mathbb{C}\right)\right\} such that for any φ,ψL2(×𝕋d,)\varphi,\psi\in L^{2}\left(\mathbb{R}\times\mathbb{T}^{d},\mathbb{C}\right),

E[ξ(φ)ξ(ψ)]=0,E[ξ(φ)ξ(ψ)¯]=φ,ψ¯L2(×𝕋d,).\mathrm{E}\left[\xi(\varphi)\xi(\psi)\right]=0,\quad\mathrm{E}\left[\xi(\varphi)\overline{\xi(\psi)}\right]=\left\langle\varphi,\overline{\psi}\right\rangle_{L^{2}\left(\mathbb{R}\times\mathbb{T}^{d},\mathbb{C}\right)}. (3)

Let (ζ(,k)=(B1(,k)+iB2(,k))/2)kd\left(\zeta(\cdot,k)=\left(B_{1}(\cdot,k)+\mathrm{i}B_{2}(\cdot,k)\right)/\sqrt{2}\right)_{k\in\mathbb{Z}^{d}} be a family of independent complex-valued Brownian motions over \mathbb{R}, where (B1(,k),B2(,k))t\left(B_{1}(\cdot,k),B_{2}(\cdot,k)\right)_{t\in\mathbb{R}} is a two-dimensional standard Brownian motion for a fixed kdk\in\mathbb{Z}^{d}. We define ξ\xi as the time derivative of the cylindrical Wiener process given by

(t,x)W(t,x):=kdζ(t,k)e2πikx.(t,x)\mapsto W(t,x):=\sum_{k\in\mathbb{Z}^{d}}\zeta(t,k)e^{2\pi\mathrm{i}k\cdot x}. (4)

To be precise, for every φL2(×𝕋d,)\varphi\in L^{2}\left(\mathbb{R}\times\mathbb{T}^{d},\mathbb{C}\right), we define

ξ(φ):=kdφ^(t,k)dζ(t,k),\xi(\varphi):=\sum_{k\in\mathbb{Z}^{d}}\int_{-\infty}^{\infty}\hat{\varphi}(t,k)\mathrm{d}\zeta(t,k), (5)

where the integral is interpreted in the sense of Itô, and φ^(t,k)\hat{\varphi}(t,k) denotes φ(t,)^(k)\widehat{\varphi(t,\cdot)}(k). According to Itô’s and Fourier’s isometries, ξ(φ)\xi(\varphi), φL2(×𝕋d,)\varphi\in L^{2}\left(\mathbb{R}\times\mathbb{T}^{d},\mathbb{C}\right), is well defined and (3) is satisfied. Since ξ(φ)\xi(\varphi) is also a centered complex Gaussian random variable, this provides us with a construction of complex-valued space-time white noise. For a given φL2(×𝕋d,)\varphi\in L^{2}\left(\mathbb{R}\times\mathbb{T}^{d},\mathbb{C}\right), we also use somewhat informal notation

ξ(φ):=×𝕋dφ(t,x)ξ(dt,dx),\xi(\varphi):=\int_{\mathbb{R}\times\mathbb{T}^{d}}\varphi(t,x)\xi(\mathrm{d}t,\mathrm{d}x),

although ξ\xi is almost surely not a measure and ξ(φ)\xi(\varphi) is only defined on a set of measure one which depends on the choice of φ\varphi.

3 Stochastic heat equation with dispersion

In this section, we concentrate on the stochastic heat equation with dispersion defined on the two-dimensional torus 𝕋2\mathbb{T}^{2} as follows

tZ=[(i+μ)Δ1]Z+ξ,t>0,x𝕋2.\partial_{t}Z=\left[(\mathrm{i}+\mu)\Delta-1\right]Z+\xi,\;t>0,\;x\in\mathbb{T}^{2}. (6)

We derive the regularity of the stationary solution of (6) and its Wick powers.

We set A:=(i+μ)Δ1A:=(\mathrm{i}+\mu)\Delta-1 and let {Pt=etA}t0\left\{P_{t}=e^{tA}\right\}_{t\geq 0} be the semigroup associated to the generator AA. For k=(k1,k2)2k=(k_{1},k_{2})\in\mathbb{Z}^{2}, we define |k|=k12+k22|k|=\sqrt{k_{1}^{2}+k_{2}^{2}} and set ρk=1+4π2μ|k|2\rho_{k}=1+4\pi^{2}\mu|k|^{2}, θk=4π2|k|2\theta_{k}=4\pi^{2}|k|^{2}. The distribution-valued stochastic process (Z,t)t>\left(Z_{-\infty,t}\right)_{t>-\infty} is defined as a stationary solution to (6) by using Duhamel’s principle and has a formal expression as follows. For ϕL2(𝕋2,)\phi\in L^{2}(\mathbb{T}^{2},\mathbb{C}),

Z,t(ϕ)\displaystyle Z_{-\infty,t}(\phi) =t𝕋2[e(ts)Aξ(s,)](x)ϕ(x)dxds\displaystyle=\int_{-\infty}^{t}\int_{\mathbb{T}^{2}}\left[e^{(t-s)A}\xi(s,\cdot)\right](x)\phi(x)\mathrm{d}x\mathrm{d}s (7)
=t𝕋2H(ts,y),ϕξ(ds,dy),\displaystyle=\int_{-\infty}^{t}\int_{\mathbb{T}^{2}}\left\langle H(t-s,\cdot-y),\phi\right\rangle\xi(\mathrm{d}s,\mathrm{d}y), (8)

where H(t,z)H(t,z) is the heat kernel defined as

H(t,z)=k2e(ρk+iθk)te2πikz=et4π(i+μ)te|z|24(i+μ)t,t{0},z𝕋2.H(t,z)=\sum_{k\in\mathbb{Z}^{2}}e^{-\left(\rho_{k}+\mathrm{i}\theta_{k}\right)t}e^{2\pi\mathrm{i}k\cdot z}=\frac{e^{-t}}{4\pi(\mathrm{i}+\mu)t}e^{-\frac{|z|^{2}}{4(\mathrm{i}+\mu)t}},\quad t\in\mathbb{R}\setminus\{0\},\quad z\in\mathbb{T}^{2}.

For ϕL2(𝕋2,)\phi\in L^{2}(\mathbb{T}^{2},\mathbb{C}),

t𝕋2|H(ts,y),ϕ|2dyds\displaystyle\int_{-\infty}^{t}\int_{\mathbb{T}^{2}}\left|\left\langle H(t-s,\cdot-y),\phi\right\rangle\right|^{2}\mathrm{d}y\mathrm{d}s (9)
=\displaystyle= 𝕋2𝕋2t𝕋2H(ts,xy)H(ts,zy)¯dydsϕ(x)ϕ(z)¯dxdz\displaystyle\,\int_{\mathbb{T}^{2}}\int_{\mathbb{T}^{2}}\int_{-\infty}^{t}\int_{\mathbb{T}^{2}}H(t-s,x-y)\overline{H(t-s,z-y)}\mathrm{d}y\mathrm{d}s\phi(x)\overline{\phi(z)}\mathrm{d}x\mathrm{d}z (10)
=\displaystyle= k2𝕋2𝕋2te2ρk(ts)e2πik(xz)dsϕ(x)ϕ(z)¯dxdz\displaystyle\,\sum_{k\in\mathbb{Z}^{2}}\int_{\mathbb{T}^{2}}\int_{\mathbb{T}^{2}}\int_{-\infty}^{t}e^{-2\rho_{k}\left(t-s\right)}e^{2\pi\mathrm{i}k\cdot(x-z)}\mathrm{d}s\phi(x)\overline{\phi(z)}\mathrm{d}x\mathrm{d}z (11)
=\displaystyle= k212ρk𝕋2𝕋2e2πik(xz)ϕ(x)ϕ(z)¯dxdz\displaystyle\,\sum_{k\in\mathbb{Z}^{2}}\frac{1}{2\rho_{k}}\int_{\mathbb{T}^{2}}\int_{\mathbb{T}^{2}}e^{2\pi\mathrm{i}k\cdot(x-z)}\phi(x)\overline{\phi(z)}\mathrm{d}x\mathrm{d}z (12)
=\displaystyle= k212ρk|ϕ^(k)|2ϕL22<,\displaystyle\,\sum_{k\in\mathbb{Z}^{2}}\frac{1}{2\rho_{k}}\left|\hat{\phi}(k)\right|^{2}\leq\left\|\phi\right\|_{L^{2}}^{2}<\infty, (13)

Therefore, Z,t(ϕ)Z_{-\infty,t}(\phi) for ϕL2(𝕋2,)\phi\in L^{2}(\mathbb{T}^{2},\mathbb{C}) is a well-defined Gaussian variable with mean zero and variance k212ρk|ϕ^(k)|2\sum_{k\in\mathbb{Z}^{2}}\frac{1}{2\rho_{k}}\left|\hat{\phi}(k)\right|^{2}.

We set Z,t:0,0:(x)1Z^{:0,0:}_{-\infty,t}(x)\equiv 1 for x𝕋2x\in\mathbb{T}^{2}. For k,l0k,l\geq 0, k+l>0k+l>0, we define (k,l)(k,l)-th Wick power of Z,t(ϕ)Z_{-\infty,t}(\phi), where ϕL2(𝕋2,)\phi\in L^{2}(\mathbb{T}^{2},\mathbb{C}), as the complex (k,l)(k,l)-th Wiener-Itô integral with respect to ξ\xi

Z,t:k,l:(ϕ)=\displaystyle Z^{:k,l:}_{-\infty,t}(\phi)= {(,t])×𝕋2}k+lj=1kH(tsj,yj)j=k+1k+lH¯(tsj,yj),ϕ\displaystyle\int_{\left\{(-\infty,t])\times\mathbb{T}^{2}\right\}^{k+l}}\left\langle\prod_{j=1}^{k}H(t-s_{j},\cdot-y_{j})\prod_{j=k+1}^{k+l}\overline{H}(t-s_{j},\cdot-y_{j}),\phi\right\rangle (14)
ξ(ds1,dy1)ξ(dsk,dyk)ξ¯(dsk+1,dyk+1)ξ¯(dsk+l,dyk+l).\displaystyle\xi(\mathrm{d}s_{1},\mathrm{d}y_{1})\cdots\xi(\mathrm{d}s_{k},\mathrm{d}y_{k})\overline{\xi}(\mathrm{d}s_{k+1},\mathrm{d}y_{k+1})\cdots\overline{\xi}(\mathrm{d}s_{k+l},\mathrm{d}y_{k+l}). (15)

By the similar argument as (9), we know that for ϕL2(𝕋2,)\phi\in L^{2}(\mathbb{T}^{2},\mathbb{C}),

k!l!{(,t])×𝕋2}k+l|j=1kH(tsj,yj)j=k+1k+lH¯(tsj,yj),ϕ|2\displaystyle k!l!\int_{\left\{(-\infty,t])\times\mathbb{T}^{2}\right\}^{k+l}}\left|\left\langle\prod_{j=1}^{k}H(t-s_{j},\cdot-y_{j})\prod_{j=k+1}^{k+l}\overline{H}(t-s_{j},\cdot-y_{j}),\phi\right\rangle\right|^{2} (16)
ds1dy1dsk+ldyk+l\displaystyle\qquad\qquad\qquad\qquad\quad\mathrm{d}s_{1}\mathrm{d}y_{1}\cdots\mathrm{d}s_{k+l}\mathrm{d}y_{k+l} (17)
=\displaystyle= k!l!j1,,jk+l2(i=1k+l12ρji)|ϕ^(i=1k+lji)|2<.\displaystyle\,k!l!\sum_{j_{1},\ldots,j_{k+l}\in\mathbb{Z}^{2}}\left(\prod_{i=1}^{k+l}\frac{1}{2\rho_{j_{i}}}\right)\left|\hat{\phi}\left(\sum_{i=1}^{k+l}j_{i}\right)\right|^{2}<\infty. (18)

Therefore, Z,t:k,l:(ϕ)k,l(ξ)Z^{:k,l:}_{-\infty,t}(\phi)\in\mathscr{H}_{k,l}(\xi) with variance (16), where ϕL2(𝕋2,)\phi\in L^{2}(\mathbb{T}^{2},\mathbb{C}) and k,l(ξ)\mathscr{H}_{k,l}(\xi) is the complex (k,l)(k,l)-th Wiener-Itô chaos of ξ\xi.

From the definition (14), we get that

Z,t:l,k:(x)=Z,t:k,l:(x)¯,x𝕋2,k,l.Z^{:l,k:}_{-\infty,t}(x)=\overline{Z^{:k,l:}_{-\infty,t}(x)},\quad x\in\mathbb{T}^{2},\quad k,l\in\mathbb{N}. (19)
Proposition 3.1

For each k,lk,l\in\mathbb{N}, any α(0,1)\alpha\in(0,1) and every p[1,)p\in[1,\infty),

suptE[Z,t:k,l:𝒞αp]<.\sup_{t\in\mathbb{R}}\mathrm{E}\left[\left\|Z^{:k,l:}_{-\infty,t}\right\|^{p}_{\mathcal{C}^{-\alpha}}\right]<\infty. (20)

Moreover, Z,t:k,l:(x)C(;𝒞α)Z^{:k,l:}_{-\infty,t}(x)\in C(\mathbb{R};\mathcal{C}^{-\alpha}) almost surely and for every p[1,)p\in[1,\infty), there exists θ0(0,1)\theta_{0}\in(0,1) depending on α\alpha such that

sups<tE[Z,t:k,l:Z,s:k,l:𝒞αp]|ts|θ0p<.\sup_{s<t}\frac{\mathrm{E}\left[\left\|Z^{:k,l:}_{-\infty,t}-Z^{:k,l:}_{-\infty,s}\right\|^{p}_{\mathcal{C}^{-\alpha}}\right]}{|t-s|^{\theta_{0}p}}<\infty. (21)
{proof}

Combining [27, Proposition 5] with the fact that Z,:k,l:Z^{:k,l:}_{-\infty,\cdot} is spatially and temporally stationary, to prove (20), it suffices to show that there exists 0<C<0<C<\infty such that for every ω2\omega\in\mathbb{Z}^{2},

E[|Z^,t:k,l:(ω)|2]C(1+|ω|)2,\mathrm{E}\left[\left|\hat{Z}^{:k,l:}_{-\infty,t}(\omega)\right|^{2}\right]\leq C(1+|\omega|)^{-2}, (22)

To prove (21), it is sufficient to prove that for any λ(0,1)\lambda\in(0,1),

E[|Z^,t:k,l:(ω)Z^,s:k,l:(ω)|2]C|ts|λ(1+|ω|)2+2λ,\mathrm{E}\left[\left|\hat{Z}^{:k,l:}_{-\infty,t}(\omega)-\hat{Z}^{:k,l:}_{-\infty,s}(\omega)\right|^{2}\right]\leq C|t-s|^{\lambda}(1+|\omega|)^{-2+2\lambda}, (23)

uniformly in 0<|ts|<10<|t-s|<1 and ω2\omega\in\mathbb{Z}^{2}. Then by [27, Proposition 5], we have that for every β<λ\beta<-\lambda, Z,t:k,l:(x)C(,𝒞β)Z^{:k,l:}_{-\infty,t}(x)\in C(\mathbb{R},\mathcal{C}^{\beta}) and

sups<tE[Z,t:k,l:Z,s:k,l:𝒞βp]|ts|λp/2<.\sup_{s<t}\frac{\mathrm{E}\left[\left\|Z^{:k,l:}_{-\infty,t}-Z^{:k,l:}_{-\infty,s}\right\|^{p}_{\mathcal{C}^{\beta}}\right]}{|t-s|^{\lambda p/2}}<\infty. (24)

Then for any α(0,1)\alpha\in(0,1), we take λ=α/2(0,1)\lambda=\alpha/2\in(0,1) and then α<λ-\alpha<-\lambda. Replacing β\beta with α-\alpha in (24), we can get that Z,t:k,l:(x)C(;𝒞α)Z^{:k,l:}_{-\infty,t}(x)\in C(\mathbb{R};\mathcal{C}^{-\alpha}) and (21) with θ0=α/4\theta_{0}=\alpha/4.

We first prove (22). For a fixed ω2\omega\in\mathbb{Z}^{2}, let ϕ=e2πiωxL2(𝕋2,)\phi=e^{-2\pi\mathrm{i}\omega\cdot x}\in L^{2}(\mathbb{T}^{2},\mathbb{C}), where x𝕋2x\in\mathbb{T}^{2}.

E[Z^,t:k,l:(ω)Z^,s:k,l:(ω)¯]=E[Z,t:k,l:,ϕZ,s:k,l:,ϕ¯]\displaystyle\mathrm{E}\left[\hat{Z}^{:k,l:}_{-\infty,t}(\omega)\overline{\hat{Z}^{:k,l:}_{-\infty,s}(\omega)}\right]=\mathrm{E}\left[\left\langle Z^{:k,l:}_{-\infty,t},\phi\right\rangle\overline{\left\langle Z^{:k,l:}_{-\infty,s},\phi\right\rangle}\right]
=\displaystyle= k!l!{(,ts])×𝕋2}k+lj=1kH(tsj,x1yj)j=k+1k+lH¯(tsj,x1yj),ϕ\displaystyle\,k!l!\int_{\left\{(-\infty,t\wedge s])\times\mathbb{T}^{2}\right\}^{k+l}}\left\langle\prod_{j=1}^{k}H(t-s_{j},x_{1}-y_{j})\prod_{j=k+1}^{k+l}\overline{H}(t-s_{j},x_{1}-y_{j}),\phi\right\rangle
j=1kH¯(ssj,x2yj)j=k+1k+lH(ssj,x2yj),ϕ¯ds1dy1dsk+ldyk+l\displaystyle\left\langle\prod_{j=1}^{k}\overline{H}(s-s_{j},x_{2}-y_{j})\prod_{j=k+1}^{k+l}H(s-s_{j},x_{2}-y_{j}),\overline{\phi}\right\rangle\mathrm{d}s_{1}\mathrm{d}y_{1}\cdots\mathrm{d}s_{k+l}\mathrm{d}y_{k+l}
=\displaystyle= k!l!𝕋2𝕋2(ts𝕋2H(tsj,x1yj)H¯(ssj,x2yj)dyjdsj)k\displaystyle\,k!l!\int_{\mathbb{T}^{2}}\int_{\mathbb{T}^{2}}\left(\int_{-\infty}^{t\wedge s}\int_{\mathbb{T}^{2}}H(t-s_{j},x_{1}-y_{j})\overline{H}(s-s_{j},x_{2}-y_{j})\mathrm{d}y_{j}\mathrm{d}s_{j}\right)^{k}
(ts𝕋2H¯(tsj,x1yj)H(ssj,x2yj)yjdsj)lϕ(x1)ϕ(x2)¯dx1dx2\displaystyle\left(\int_{-\infty}^{t\wedge s}\int_{\mathbb{T}^{2}}\overline{H}(t-s_{j},x_{1}-y_{j})H(s-s_{j},x_{2}-y_{j})y_{j}\mathrm{d}s_{j}\right)^{l}\phi(x_{1})\overline{\phi(x_{2})}\mathrm{d}x_{1}\mathrm{d}x_{2}
=\displaystyle= k!l!𝕋2𝕋2(j2eρj|ts|iθj(ts)2ρje2πij(x1x2))k\displaystyle\,k!l!\int_{\mathbb{T}^{2}}\int_{\mathbb{T}^{2}}\left(\sum_{j\in\mathbb{Z}^{2}}\frac{e^{-\rho_{j}|t-s|-\mathrm{i}\theta_{j}(t-s)}}{2\rho_{j}}e^{2\pi\mathrm{i}j\cdot(x_{1}-x_{2})}\right)^{k}
(j2eρj|ts|+iθj(ts)2ρje2πij(x1x2))lϕ(x1)ϕ(x2)¯dx1dx2\displaystyle\left(\sum_{j\in\mathbb{Z}^{2}}\frac{e^{-\rho_{j}|t-s|+\mathrm{i}\theta_{j}(t-s)}}{2\rho_{j}}e^{-2\pi\mathrm{i}j\cdot(x_{1}-x_{2})}\right)^{l}\phi(x_{1})\overline{\phi(x_{2})}\mathrm{d}x_{1}\mathrm{d}x_{2}
=k!l!j1++jkjk+1jk+l=ωi=1k+leρji|ts|iaiθji(ts)2ρji,\displaystyle=\,k!l!\sum_{j_{1}+\cdots+j_{k}-j_{k+1}-\cdots-j_{k+l}=\omega}\prod_{i=1}^{k+l}\frac{e^{-\rho_{j_{i}}|t-s|-\mathrm{i}a_{i}\theta_{j_{i}}(t-s)}}{2\rho_{j_{i}}},

where ai=1a_{i}=1 for 1ik1\leq i\leq k and ai=1a_{i}=-1 for k+1ik+lk+1\leq i\leq k+l. Therefore, by [16, Lemma 10.14] or [35, Corollary C.3], we get that for every ϵ(0,1)\epsilon\in(0,1),

E[|Z^,t:k,l:(ω)|2]=k!l!j1++jkjk+1jk+l=ωi=1k+l12ρjiC1(1+|ω|2)1ϵ.\mathrm{E}\left[\left|\hat{Z}^{:k,l:}_{-\infty,t}(\omega)\right|^{2}\right]=k!l!\sum_{j_{1}+\cdots+j_{k}-j_{k+1}-\cdots-j_{k+l}=\omega}\prod_{i=1}^{k+l}\frac{1}{2\rho_{j_{i}}}\leq C\frac{1}{(1+|\omega|^{2})^{1-\epsilon}}.

Since ϵ(0,1)\epsilon\in(0,1) is arbitrary, we finish the proof of (22).

Next, we show (23).

E[|Z^,t:k,l:(ω)Z^,s:k,l:(ω)|2]\displaystyle\mathrm{E}\left[\left|\hat{Z}^{:k,l:}_{-\infty,t}(\omega)-\hat{Z}^{:k,l:}_{-\infty,s}(\omega)\right|^{2}\right]
=\displaystyle= E[|Z^,t:k,l:(ω)|2+|Z^,s:k,l:(ω)|2Z^,t:k,l:(ω)Z^,s:k,l:(ω)¯Z^,t:k,l:(ω)¯Z^,s:k,l:(ω)]\displaystyle\mathrm{E}\left[\left|\hat{Z}^{:k,l:}_{-\infty,t}(\omega)\right|^{2}+\left|\hat{Z}^{:k,l:}_{-\infty,s}(\omega)\right|^{2}-\hat{Z}^{:k,l:}_{-\infty,t}(\omega)\overline{\hat{Z}^{:k,l:}_{-\infty,s}(\omega)}-\overline{\hat{Z}^{:k,l:}_{-\infty,t}(\omega)}\hat{Z}^{:k,l:}_{-\infty,s}(\omega)\right]
=\displaystyle= 2k!l!j1++jkjk+1jk+l=ω(i=1k+l12ρji)(1e|ts|i=1k+lρjicos(i=1k+laiθji(ts)))\displaystyle 2k!l!\sum_{j_{1}+\cdots+j_{k}-j_{k+1}-\cdots-j_{k+l}=\omega}\left(\prod_{i=1}^{k+l}\frac{1}{2\rho_{j_{i}}}\right)\left(1-e^{-|t-s|\sum_{i=1}^{k+l}\rho_{j_{i}}}\cos\left(\sum_{i=1}^{k+l}a_{i}\theta_{j_{i}}(t-s)\right)\right)
\displaystyle\leq C|ts|λj1++jkjk+1jk+l=ω(i=1k+l1ρji)i=1k+lρjiλ\displaystyle C|t-s|^{\lambda}\sum_{j_{1}+\cdots+j_{k}-j_{k+1}-\cdots-j_{k+l}=\omega}\left(\prod_{i=1}^{k+l}\frac{1}{\rho_{j_{i}}}\right)\sum_{i=1}^{k+l}\rho_{j_{i}}^{\lambda}
\displaystyle\leq C|ts|λj1++jkjk+1jk+l=ω1ρj11λi=2k+l1ρji,\displaystyle C|t-s|^{\lambda}\sum_{j_{1}+\cdots+j_{k}-j_{k+1}-\cdots-j_{k+l}=\omega}\frac{1}{\rho_{j_{1}}^{1-\lambda}}\prod_{i=2}^{k+l}\frac{1}{\rho_{j_{i}}},

where in the third inequality, we use the fact that for any λ(0,1)\lambda\in(0,1), a,ba,b\in\mathbb{R},

1cos(b|ts|)ea|ts|C(a|ts|)λ,1-\cos\left(b|t-s|\right)e^{-a|t-s|}\leq C\left(a|t-s|\right)^{\lambda}, (25)

which holds uniformly for 0<|ts|<10<|t-s|<1, and (i=1k+lρji)λi=1k+lρjiλ\left(\sum_{i=1}^{k+l}\rho_{j_{i}}\right)^{\lambda}\leq\sum_{i=1}^{k+l}\rho_{j_{i}}^{\lambda}. Then by [16, Lemma 10.14] or [35, Lemma C.2, Corollary C.3], we know that for every ϵ(0,1)\epsilon\in(0,1),

E[|Z^,t:k,l:(ω)Z^,s:k,l:(ω)|2]C|ts|λ1(1+|ω|2)1λϵ,\mathrm{E}\left[\left|\hat{Z}^{:k,l:}_{-\infty,t}(\omega)-\hat{Z}^{:k,l:}_{-\infty,s}(\omega)\right|^{2}\right]\leq C|t-s|^{\lambda}\frac{1}{(1+|\omega|^{2})^{1-\lambda-\epsilon}},

Since ϵ(0,1)\epsilon\in(0,1) is arbitrary, we prove (23).

Now we consider the regularity of the solution Zs,tZ_{s,t} of (6) with zero initial condition at time ss\in\mathbb{R}. Note that

Zs,t(x)=Z,t(x)(PtsZ,s)(x).Z_{s,t}(x)=Z_{-\infty,t}(x)-\left(P_{t-s}Z_{-\infty,s}\right)(x).

Then we define the (k,l)(k,l)-th Wick power of Zs,t(x)Z_{s,t}(x) as

Zs,t:k,l:=i=0kj=0l(ki)(lj)(1)i+j(PtsZ,s)i(PtsZ,s)¯jZ,t:ki,lj:,Z_{s,t}^{:k,l:}=\sum_{i=0}^{k}\sum_{j=0}^{l}\binom{k}{i}\binom{l}{j}(-1)^{i+j}\left(P_{t-s}Z_{-\infty,s}\right)^{i}\overline{\left(P_{t-s}Z_{-\infty,s}\right)}^{j}Z_{-\infty,t}^{:k-i,l-j:}, (26)

and by (19), we have

Zs,t:l,k:=Zs,t:k,l:¯,k,l.Z_{s,t}^{:l,k:}=\overline{Z_{s,t}^{:k,l:}},\quad k,l\in\mathbb{N}.
Proposition 3.2

Let p1,α,α(0,1),T>0p\geq 1,\alpha,\alpha^{\prime}\in(0,1),T>0 and ss\in\mathbb{R}. Then for each k,lk,l\in\mathbb{N}, there exists θ>0\theta>0 depending on α,α\alpha,\alpha^{\prime} and CC depending on α,α,k,l,p\alpha,\alpha^{\prime},k,l,p such that

E[sup0tTt(k+l1)αpZs,s+t:k,l:𝒞αp]CTpθ.\mathrm{E}\left[\sup_{0\leq t\leq T}t^{(k+l-1)\alpha^{\prime}p}\left\|Z_{s,s+t}^{:k,l:}\right\|_{\mathcal{C}^{-\alpha}}^{p}\right]\leq CT^{p\theta}. (27)
{proof}

By the fact that for any s,ts,t\in\mathbb{R}, Z,sZ_{-\infty,s} and Z,tZ_{-\infty,t} have the same distribution and the definition (26), we know that for any s0,s,ts_{0},s,t\in\mathbb{R}, Zs0,s0+t:k,l:Z_{s_{0},s_{0}+t}^{:k,l:} and Zs,s+t:k,l:Z_{s,s+t}^{:k,l:} have the same distribution. Therefore, without loss of generality, we prove (27) for s=0s=0. Let α¯=α23α\bar{\alpha}=\alpha\wedge\frac{2}{3}\alpha^{\prime} and V(t)=PtZ,0V(t)=P_{t}Z_{-\infty,0}. We first consider the case for k+l=1k+l=1. By Proposition 2.5,

Z0,t:k,l:𝒞α\displaystyle\left\|Z_{0,t}^{:k,l:}\right\|_{\mathcal{C}^{-\alpha}} Z,t:k,l:Z,0:k,l:𝒞α+(1Pt)Z,0:k,l:𝒞α\displaystyle\leq\left\|Z_{-\infty,t}^{:k,l:}-Z_{-\infty,0}^{:k,l:}\right\|_{\mathcal{C}^{-\alpha}}+\left\|(1-P_{t})Z_{-\infty,0}^{:k,l:}\right\|_{\mathcal{C}^{-\alpha}}
Z,t:k,l:Z,0:k,l:𝒞α+tαα¯2Z,0:k,l:𝒞α¯\displaystyle\lesssim\left\|Z_{-\infty,t}^{:k,l:}-Z_{-\infty,0}^{:k,l:}\right\|_{\mathcal{C}^{-\alpha}}+t^{\frac{\alpha-\bar{\alpha}}{2}}\left\|Z_{-\infty,0}^{:k,l:}\right\|_{\mathcal{C}^{-\bar{\alpha}}}

Then by Proposition 3.1,

E[sup0tTZ0,t:k,l:𝒞αp]\displaystyle\mathrm{E}\left[\sup_{0\leq t\leq T}\left\|Z_{0,t}^{:k,l:}\right\|_{\mathcal{C}^{-\alpha}}^{p}\right] E[sup0tTZ,t:k,l:Z,0:k,l:𝒞αp]+Tαα¯2pE[Z,0:k,l:𝒞α¯p]\displaystyle\lesssim\mathrm{E}\left[\sup_{0\leq t\leq T}\left\|Z_{-\infty,t}^{:k,l:}-Z_{-\infty,0}^{:k,l:}\right\|^{p}_{\mathcal{C}^{-\alpha}}\right]+T^{\frac{\alpha-\bar{\alpha}}{2}p}\mathrm{E}\left[\left\|Z_{-\infty,0}^{:k,l:}\right\|^{p}_{\mathcal{C}^{-\bar{\alpha}}}\right]
Tθ0p+Tαα¯2p.\displaystyle\lesssim T^{\theta_{0}p}+T^{\frac{\alpha-\bar{\alpha}}{2}p}.

Then there exists θ>0\theta>0 depending on α,α\alpha,\alpha^{\prime} such that (27) holds for k+l=1k+l=1.

Now we consider the case for k+l>1k+l>1. By Proposition 2.1, 2.4 and 2.6, we get that

V(t)kV(t)¯l𝒞α\displaystyle\left\|V(t)^{k}\overline{V(t)}^{l}\right\|_{\mathcal{C}^{-\alpha}} V(t)kV(t)¯l𝒞α¯V(t)𝒞2α¯k+l1V(t)𝒞α¯\displaystyle\lesssim\left\|V(t)^{k}\overline{V(t)}^{l}\right\|_{\mathcal{C}^{-\bar{\alpha}}}\lesssim\|V(t)\|_{\mathcal{C}^{2\bar{\alpha}}}^{k+l-1}\|V(t)\|_{\mathcal{C}^{-\bar{\alpha}}}
t32α¯(k+l1)Z,0𝒞α¯k+l.\displaystyle\lesssim t^{-\frac{3}{2}\bar{\alpha}(k+l-1)}\left\|Z_{-\infty,0}\right\|_{\mathcal{C}^{-\bar{\alpha}}}^{k+l}.

Similarly, for (i,j)(k,l)(i,j)\neq(k,l), we have that

V(t)iV(t)¯jZ,t:ki,lj:𝒞α\displaystyle\left\|V(t)^{i}\overline{V(t)}^{j}Z_{-\infty,t}^{:k-i,l-j:}\right\|_{\mathcal{C}^{-\alpha}} V(t)iV(t)¯jZ,t:ki,lj:𝒞α¯V(t)𝒞2α¯i+jZ,t:ki,lj:𝒞α¯\displaystyle\lesssim\left\|V(t)^{i}\overline{V(t)}^{j}Z_{-\infty,t}^{:k-i,l-j:}\right\|_{\mathcal{C}^{-\bar{\alpha}}}\lesssim\|V(t)\|_{\mathcal{C}^{2\bar{\alpha}}}^{i+j}\|Z_{-\infty,t}^{:k-i,l-j:}\|_{\mathcal{C}^{-\bar{\alpha}}}
t32α¯(i+j)Z,0𝒞α¯i+jZ,t:ki,lj:𝒞α¯.\displaystyle\lesssim t^{-\frac{3}{2}\bar{\alpha}(i+j)}\left\|Z_{-\infty,0}\right\|_{\mathcal{C}^{-\bar{\alpha}}}^{i+j}\left\|Z_{-\infty,t}^{:k-i,l-j:}\right\|_{\mathcal{C}^{-\bar{\alpha}}}.

Therefore,

E[sup0tTt(k+l1)αpZ0,t:k,l:𝒞αp]\displaystyle\mathrm{E}\left[\sup_{0\leq t\leq T}t^{(k+l-1)\alpha^{\prime}p}\left\|Z_{0,t}^{:k,l:}\right\|_{\mathcal{C}^{-\alpha}}^{p}\right]
\displaystyle\lesssim T(α32α¯)(k+l1)pE[Z,0𝒞α¯(k+l)p]\displaystyle\,T^{(\alpha^{\prime}-\frac{3}{2}\bar{\alpha})(k+l-1)p}\mathrm{E}\left[\left\|Z_{-\infty,0}\right\|_{\mathcal{C}^{-\bar{\alpha}}}^{(k+l)p}\right]
+0i+j<k+lT((k+l1)α32α¯(i+j))p(E[Z,0𝒞α¯2(i+j)p])12(E[sup0tTZ,t:ki,lj:𝒞α¯2p])12,\displaystyle+\sum_{0\leq i+j<k+l}T^{\left((k+l-1)\alpha^{\prime}-\frac{3}{2}\bar{\alpha}(i+j)\right)p}\left(\mathrm{E}\left[\left\|Z_{-\infty,0}\right\|_{\mathcal{C}^{-\bar{\alpha}}}^{2(i+j)p}\right]\right)^{\frac{1}{2}}\left(\mathrm{E}\left[\sup_{0\leq t\leq T}\left\|Z_{-\infty,t}^{:k-i,l-j:}\right\|^{2p}_{\mathcal{C}^{-\bar{\alpha}}}\right]\right)^{\frac{1}{2}},

where we use Cauchy-Schwarz inequality in the last inequality. Combining with Proposition 3.1, we have (27) for k+l>1k+l>1.

Let M:=m(m+1)2+2m+1M:=\frac{m(m+1)}{2}+2m+1 and define a set

L:={i,j:0im+1,0jm,ij,i+j>1}.L:=\left\{i,j\in\mathbb{N}:0\leq i\leq m+1,0\leq j\leq m,i\geq j,i+j>1\right\}. (28)

Let α,α(0,1)\alpha,\alpha^{\prime}\in(0,1) and T>0T>0, we define

CM,α(0,T):=C([0,T];𝒞α)×C((0,T];𝒞α)M1,C^{M,-\alpha}(0,T):=C([0,T];\mathcal{C}^{-\alpha})\times C((0,T];\mathcal{C}^{-\alpha})^{M-1}, (29)

and for a vector A={At,At(i,j):i,jL,t[0,T]}CM,α(0,T)A=\left\{A_{t},A_{t}^{(i,j)}:i,j\in L,t\in[0,T]\right\}\in C^{M,-\alpha}(0,T), we define the norm

Aα,α,T:=\displaystyle\left\|A\right\|_{\alpha,\alpha^{\prime},T}:= sup0tT(At𝒞αsupi,jLt(i+j1)αAt(i,j)𝒞α).\displaystyle\sup_{0\leq t\leq T}\left(\left\|A_{t}\right\|_{\mathcal{C}^{-\alpha}}\vee\sup_{i,j\in L}t^{(i+j-1)\alpha^{\prime}}\left\|A_{t}^{(i,j)}\right\|_{\mathcal{C}^{-\alpha}}\right).

We denote by \mathcal{L} the set

:={ACM,α(0,):α,α(0,1),T>0,Aα,α,T<}\mathcal{L}:=\left\{A\in C^{M,-\alpha}(0,\infty):\forall\alpha,\alpha^{\prime}\in(0,1),\forall T>0,\left\|A\right\|_{\alpha,\alpha^{\prime},T}<\infty\right\} (30)

Then by Proposition 3.2, we know that

Z¯:={Z0,t,Z0,t:i,j::i,jL,t>0}\underline{Z}:=\left\{Z_{0,t},Z_{0,t}^{:i,j:}:i,j\in L,t>0\right\}\in\mathcal{L}

almost surely and for every p[1,)p\in[1,\infty), there exists θ>0\theta^{\prime}>0 depending on α,α\alpha,\alpha^{\prime} and CC depending on α,α,p,m\alpha,\alpha^{\prime},p,m such that

E[Z¯α,α,Tp]CTpθ.\mathrm{E}\left[\left\|\underline{Z}\right\|^{p}_{\alpha,\alpha^{\prime},T}\right]\leq CT^{p\theta^{\prime}}. (31)

4 Global well-posedness

4.1 Local existence and uniqueness

We fix α,α(0,1)\alpha,\alpha^{\prime}\in(0,1) small enough and Z¯={Z0,t,Z0,t:i,j::i,jL,t>0}\underline{Z}=\left\{Z_{0,t},Z_{0,t}^{:i,j:}:i,j\in L,t>0\right\}\in\mathcal{L} with finite norm Z¯α,α,T\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},T} for a given T>0T>0 from now on. For vC((0,T];𝒞α+ϵ)v\in C((0,T];\mathcal{C}^{\alpha+\epsilon}), where ϵ>0\epsilon>0 is arbitrary, we define

Ψ(vt,Z¯t)\displaystyle\Psi\left(v_{t},\underline{Z}_{t}\right) :=νi=0m+1j=0m(m+1i)(mj)vtivt¯jZ0,t:m+1i,mj:+(τ+1)(vt+Z0,t)\displaystyle:=-\nu\sum_{i=0}^{m+1}\sum_{j=0}^{m}\binom{m+1}{i}\binom{m}{j}v_{t}^{i}\overline{v_{t}}^{j}Z_{0,t}^{:m+1-i,m-j:}+(\tau+1)(v_{t}+Z_{0,t})
=νvtm+1vt¯m+Ψ(vt,Z¯t),\displaystyle=-\nu v_{t}^{m+1}\overline{v_{t}}^{m}+\Psi^{\prime}\left(v_{t},\underline{Z}_{t}\right),

where

Ψ(vt,Z¯t):=νi+j<2m+1(m+1i)(mj)vtivt¯jZ0,t:m+1i,mj:+(τ+1)(vt+Z0,t).\Psi^{\prime}\left(v_{t},\underline{Z}_{t}\right):=-\nu\sum_{i+j<2m+1}\binom{m+1}{i}\binom{m}{j}v_{t}^{i}\overline{v_{t}}^{j}Z_{0,t}^{:m+1-i,m-j:}+(\tau+1)(v_{t}+Z_{0,t}). (32)

By Proposition 2.6 and Z¯CM,α(0,T)\underline{Z}\in C^{M,-\alpha}(0,T), we know that Ψ(vt,Z¯t)\Psi\left(v_{t},\underline{Z}_{t}\right) and Ψ(vt,Z¯t)\Psi^{\prime}\left(v_{t},\underline{Z}_{t}\right) are well-defined.

Definition 4.1

Let T>0T>0 and α0\alpha_{0}, β\beta and γ\gamma satisfy

α0>0,β>0,α0+β2<γ<12m+1.\alpha_{0}>0,\quad\beta>0,\quad\frac{\alpha_{0}+\beta}{2}<\gamma<\frac{1}{2m+1}.

We say that a function vv is a mild solution of

{tv=[(i+μ)Δ1]v+Ψ(v,Z¯),t>0,x𝕋2,v(0,)=u0𝒞α0,\begin{cases}\partial_{t}v=\left[(\mathrm{i}+\mu)\Delta-1\right]v+\Psi(v,\underline{Z}),&t>0,x\in\mathbb{T}^{2},\\ v(0,\cdot)=u_{0}\in\mathcal{C}^{-\alpha_{0}},\end{cases} (33)

up the time TT if vC((0,T];𝒞β)v\in C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) with the norm sup0<tTtγvt𝒞β<\sup_{0<t\leq T}t^{\gamma}\left\|v_{t}\right\|_{\mathcal{C}^{\beta}}<\infty and

vt=Ptu0+0tPtsΨ(vs,Z¯s)ds,v_{t}=P_{t}u_{0}+\int_{0}^{t}P_{t-s}\Psi\left(v_{s},\underline{Z}_{s}\right)\mathrm{d}s, (34)

for every tTt\leq T.

Motivated by the Da Prato-Debussche method (see [10]), we say that uu solves the equation (1) in the sense of renormalization with initial data u0𝒞α0u_{0}\in\mathcal{C}^{-\alpha_{0}} if u=v+Z0,u=v+Z_{0,\cdot}, where the remainder vv is a mild solution of (33). That is, u=v+Z0,u=v+Z_{0,\cdot} solves the equation

{tu=[(i+μ)Δ1]u+Ψ(uZ0,t,Z¯)+ξ,t>0,x𝕋2,u(0,)=u0𝒞α0.\begin{cases}\partial_{t}u=\left[(\mathrm{i}+\mu)\Delta-1\right]u+\Psi(u-Z_{0,t},\underline{Z})+\xi,&t>0,x\in\mathbb{T}^{2},\\ u(0,\cdot)=u_{0}\in\mathcal{C}^{-\alpha_{0}}.\end{cases} (35)
Theorem 4.2 (Local existence and uniqueness)

Let u0𝒞α0u_{0}\in\mathcal{C}^{-\alpha_{0}} with α0>0\alpha_{0}>0. Assume that there exists R>0R>0 such that u0𝒞α0R\left\|u_{0}\right\|_{\mathcal{C}^{-\alpha_{0}}}\leq R. Then for β\beta and γ\gamma satisfying

β>0,α0+β2<γ<12m+1,\beta>0,\quad\frac{\alpha_{0}+\beta}{2}<\gamma<\frac{1}{2m+1},

and T>0T>0, there exists TTT^{*}\leq T depending on RR and Z¯α,α,T\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},T} such that (33) has a unique mild solution on [0,T][0,T^{*}].

{proof}

We solve the integral equation

vt=Ptu0+0tPtsΨ(vs,Z¯s)ds,v_{t}=P_{t}u_{0}+\int_{0}^{t}P_{t-s}\Psi\left(v_{s},\underline{Z}_{s}\right)\mathrm{d}s,

by using a fixed point argument in the space C((0,T];𝒞β)C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) with the norm sup0<tTtγvt𝒞β<\sup_{0<t\leq T}t^{\gamma}\left\|v_{t}\right\|_{\mathcal{C}^{\beta}}<\infty. We denote by T\mathcal{M}_{T} the solution map

(Tv)t=Ptu0+0tPtsΨ(vs,Z¯s)ds,0tT.\left(\mathcal{M}_{T}v\right)_{t}=P_{t}u_{0}+\int_{0}^{t}P_{t-s}\Psi\left(v_{s},\underline{Z}_{s}\right)\mathrm{d}s,\quad 0\leq t\leq T.

It suffices to prove that there exists TT^{*} small enough such that T\mathcal{M}_{T^{*}} is a contraction from

T:={vC((0,T];𝒞β):sup0<tTtγvt𝒞β1}\mathcal{B}_{T^{*}}:=\left\{v\in C\left(\left(0,T^{*}\right];\mathcal{C}^{\beta}\right):\sup_{0<t\leq T^{*}}t^{\gamma}\left\|v_{t}\right\|_{\mathcal{C}^{\beta}}\leq 1\right\} (36)

into itself. We first prove that there exists TT^{*} small enough such that T\mathcal{M}_{T^{*}} maps T\mathcal{B}_{T^{*}} into itself. Without loss of generality, we assume that T1TT^{*}\leq 1\wedge T. By Proposition 2.4 and 2.6, for 0<tT0<t\leq T^{*}, we have that

(Tv)t𝒞β\displaystyle\left\|\left(\mathcal{M}_{T}v\right)_{t}\right\|_{\mathcal{C}^{\beta}} Ptu0𝒞β+0tPtsΨ(vs,Z¯s)𝒞βds\displaystyle\lesssim\left\|P_{t}u_{0}\right\|_{\mathcal{C}^{\beta}}+\int_{0}^{t}\left\|P_{t-s}\Psi\left(v_{s},\underline{Z}_{s}\right)\right\|_{\mathcal{C}^{\beta}}\mathrm{d}s
tα0+β2u0𝒞α0+0tvsm+1vs¯m𝒞βds+0tvs𝒞βds\displaystyle\lesssim t^{-\frac{\alpha_{0}+\beta}{2}}\left\|u_{0}\right\|_{\mathcal{C}^{-\alpha_{0}}}+\int_{0}^{t}\left\|v_{s}^{m+1}\overline{v_{s}}^{m}\right\|_{\mathcal{C}^{\beta}}\mathrm{d}s+\int_{0}^{t}\left\|v_{s}\right\|_{\mathcal{C}^{\beta}}\mathrm{d}s
+i+j<2m+10t(ts)α+β2vsivs¯jZ0,s:m+1i,mj:𝒞αds+0tZ0,s𝒞αds\displaystyle\quad+\sum_{i+j<2m+1}\int_{0}^{t}(t-s)^{-\frac{\alpha+\beta}{2}}\left\|v_{s}^{i}\overline{v_{s}}^{j}Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}\mathrm{d}s+\int_{0}^{t}\left\|Z_{0,s}\right\|_{\mathcal{C}^{-\alpha}}\mathrm{d}s
tα0+β2u0𝒞α0+0ts(2m+1)γds\displaystyle\lesssim t^{-\frac{\alpha_{0}+\beta}{2}}\left\|u_{0}\right\|_{\mathcal{C}^{-\alpha_{0}}}+\int_{0}^{t}s^{-(2m+1)\gamma}\mathrm{d}s
+i+j<2m+10t(ts)α+β2s(i+j)γ(2m(i+j))αds\displaystyle\quad+\sum_{i+j<2m+1}\int_{0}^{t}(t-s)^{-\frac{\alpha+\beta}{2}}s^{-(i+j)\gamma-(2m-(i+j))\alpha^{\prime}}\mathrm{d}s
tα0+β2u0𝒞α0+0ts(2m+1)γds+0t(ts)α+β2s2mγds\displaystyle\lesssim t^{-\frac{\alpha_{0}+\beta}{2}}\left\|u_{0}\right\|_{\mathcal{C}^{-\alpha_{0}}}+\int_{0}^{t}s^{-(2m+1)\gamma}\mathrm{d}s+\int_{0}^{t}(t-s)^{-\frac{\alpha+\beta}{2}}s^{-2m\gamma}\mathrm{d}s
tα0+β2u0𝒞α0+t1(2m+1)γ+t1α+β22mγ,\displaystyle\lesssim t^{-\frac{\alpha_{0}+\beta}{2}}\left\|u_{0}\right\|_{\mathcal{C}^{-\alpha_{0}}}+t^{1-(2m+1)\gamma}+t^{1-\frac{\alpha+\beta}{2}-2m\gamma},

where we choose α,α>0\alpha,\alpha^{\prime}>0 small enough such that α<β\alpha<\beta, α+β2+2mγ<1\frac{\alpha+\beta}{2}+2m\gamma<1 and α<γ\alpha^{\prime}<\gamma. Multiplying both sides by tγt^{\gamma}, we obtain that

tγ(Tv)t𝒞βC(tγα0+β2R+t12mγ+t1α+β2(2m1)γ)C(R+1)tθ,\displaystyle t^{\gamma}\left\|\left(\mathcal{M}_{T}v\right)_{t}\right\|_{\mathcal{C}^{\beta}}\leq C\left(t^{\gamma-\frac{\alpha_{0}+\beta}{2}}R+t^{1-2m\gamma}+t^{1-\frac{\alpha+\beta}{2}-(2m-1)\gamma}\right)\leq C(R+1)t^{\theta},

where θ=min{γα0+β2,12mγ,1α+β2(2m1)γ}>0\theta=\min\left\{\gamma-\frac{\alpha_{0}+\beta}{2},1-2m\gamma,1-\frac{\alpha+\beta}{2}-(2m-1)\gamma\right\}>0 and CC is a constant depending on Z¯α,α,T\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},T}. Let

T=(1C(R+1))1θ,T^{*}=\left(\frac{1}{C(R+1)}\right)^{\frac{1}{\theta}}, (37)

then

sup0<tTtγ(Tv)t𝒞β1.\sup_{0<t\leq T^{*}}t^{\gamma}\left\|\left(\mathcal{M}_{T^{*}}v\right)_{t}\right\|_{\mathcal{C}^{\beta}}\leq 1.

Then we finish the proof that there exists TT^{*} small enough such that T\mathcal{M}_{T^{*}} maps T\mathcal{B}_{T^{*}} into itself. The contraction property of T\mathcal{M}_{T^{*}} can be proved by using the similar argument.

Next we show the uniqueness of solutions. For TT^{*} defined as above, let v,vC((0,T];𝒞β)v,v^{\prime}\in C\left(\left(0,T^{*}\right];\mathcal{C}^{\beta}\right) with finite norm sup0<tTtγ()t𝒞β\sup_{0<t\leq T^{*}}t^{\gamma}\left\|(\cdot)_{t}\right\|_{\mathcal{C}^{\beta}} both be the mild solutions of (33). Define

T:=sup{t[0,T]:st,vs=vs}.T^{**}:=\sup\left\{t\in[0,T^{*}]:\forall s\leq t,v_{s}=v^{\prime}_{s}\right\}.

Assume that T<TT^{**}<T^{*}. Taking smaller β(0,1)\beta\in(0,1) and larger γ((α0+β)/2,1/(2m+1))\gamma\in\left((\alpha_{0}+\beta)/2,1/(2m+1)\right), we have that there exists δ(0,TT)\delta\in\left(0,T^{*}-T^{**}\right) such that (vT+t)0tδ,(vT+t)0tδδ\left(v_{T^{**}+t}\right)_{0\leq t\leq\delta},\left(v_{T^{**}+t}^{\prime}\right)_{0\leq t\leq\delta}\in\mathcal{B}_{\delta} (see (36) for the definition of δ\mathcal{B}_{\delta}). By using the fixed point argument as above, we can show that δ:δδ\mathcal{M}^{\prime}_{\delta}:\mathcal{B}_{\delta}\rightarrow\mathcal{B}_{\delta} is a contraction, where

(δy)t:=PtyT+0tPtsΨ(yT+s,Z¯T,T+s)𝑑s,yT+δ,0tδ,\left(\mathcal{M}^{\prime}_{\delta}y\right)_{t}:=P_{t}y_{T^{**}}+\int_{0}^{t}P_{t-s}\Psi\left(y_{T^{**}+s},\underline{Z}_{T^{**},T^{**}+s}\right)ds,\quad y_{T^{**}+\cdot}\in\mathcal{B}_{\delta},\quad 0\leq t\leq\delta,

and Z¯T,T+s={ZT,T+s,ZT,T+s:i,j::i,jL}\underline{Z}_{T^{**},T^{**}+s}=\left\{Z_{T^{**},T^{**}+s},Z_{T^{**},T^{**}+s}^{:i,j:}:i,j\in L\right\}. Since both (vT+t)0tδ\left(v_{T^{**}+t}\right)_{0\leq t\leq\delta} and (vT+t)0tδ\left(v_{T^{**}+t}^{\prime}\right)_{0\leq t\leq\delta} are fixed points of δ\mathcal{M}^{\prime}_{\delta}, we know that vT+t=vT+tv_{T^{**}+t}=v_{T^{**}+t}^{\prime} for 0tδ0\leq t\leq\delta. This contradicts the definition of TT^{**}. Therefore T=TT^{**}=T^{*} and v=vv=v^{\prime}.

4.2 A priori estimate

Testing the equation (33) with |v|2p2v¯|v|^{2p-2}\overline{v}, we get the following expression of vtL2p2p\|v_{t}\|_{L^{2p}}^{2p} (see [25, Proposition 3.3], [34, Proposition 6.2] and [18, Proposition 5.2] for the similar argument).

Proposition 4.3

Let vC((0,T];𝒞β)v\in C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) be a mild solution to (33). Then for all s0>0s_{0}>0 and every p1p\geq 1,

12p(vtL2p2pvs0L2p2p)=\displaystyle\frac{1}{2p}\left(\|v_{t}\|_{L^{2p}}^{2p}-\|v_{s_{0}}\|_{L^{2p}}^{2p}\right)= Re[(i+μ)s0t{(vsvs¯)p1vs¯},vsds]s0tvsL2p2pds\displaystyle-\mathrm{Re}\left[(\mathrm{i}+\mu)\int_{s_{0}}^{t}\left\langle\nabla\left\{(v_{s}\overline{v_{s}})^{p-1}\overline{v_{s}}\right\},\nabla v_{s}\right\rangle\mathrm{d}s\right]-\int_{s_{0}}^{t}\|v_{s}\|_{L^{2p}}^{2p}\mathrm{d}s (38)
Reνs0tvsL2p+2m2p+2mds+s0t|vs|2p2,Re(vs¯Ψ(vs,Z¯s))ds,\displaystyle-\mathrm{Re}\nu\int_{s_{0}}^{t}\|v_{s}\|_{L^{2p+2m}}^{2p+2m}\mathrm{d}s+\int_{s_{0}}^{t}\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}\Psi^{\prime}\left(v_{s},\underline{Z}_{s}\right)\right)\right\rangle\mathrm{d}s,

for all s0tTs_{0}\leq t\leq T. In particular, if we differentiate with respect to tt, then

12ptvtL2p2p=Re[(i+μ){(vtvt¯)p1vt¯},vt]vtL2p2pReνvtL2p+2m2p+2m+|vt|2p2,Re(vt¯Ψ(vt,Z¯t)),,\begin{aligned} \frac{1}{2p}\partial_{t}\|v_{t}\|_{L^{2p}}^{2p}=&-\mathrm{Re}\left[(\mathrm{i}+\mu)\left\langle\nabla\left\{(v_{t}\overline{v_{t}})^{p-1}\overline{v_{t}}\right\},\nabla v_{t}\right\rangle\right]-\|v_{t}\|_{L^{2p}}^{2p}\\ &-\mathrm{Re}\nu\|v_{t}\|_{L^{2p+2m}}^{2p+2m}+\left\langle|v_{t}|^{2p-2},\mathrm{Re}\left(\overline{v_{t}}\Psi^{\prime}\left(v_{t},\underline{Z}_{t}\right)\right)\right\rangle,\end{aligned}, (39)

for every t(0,T]t\in(0,T].

Remark 4.4

Theorem 4.2 implies that vC((0,T];𝒞β)v\in C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) for some β<1\beta<1 since the initial data u0u_{0} of (33) belongs to 𝒞α0\mathcal{C}^{-\alpha_{0}} with α0>0\alpha_{0}>0. However, Proposition 4.3 involves the first order partial derivatives of vv and its proof requires some time regularity on vv. One can prove that for fixed t>0t>0, vtv_{t} is almost a Hölder continuous function from (0,T](0,T] to 𝒞2\mathcal{C}^{2-} with some strictly positive exponent (see [25, Proposition 3.1] for the proof in case of m=1m=1).

Proposition 4.5

Let vC((0,T];𝒞β)v\in C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) be a mild solution to (33). Let 1p1+μ(μ+1+μ2)1\leq p\leq 1+\mu\left(\mu+\sqrt{1+\mu^{2}}\right). For every 0δ<10\leq\delta<1 such that

p1μ(μ+1+μ2)1δ,\frac{p-1}{\mu\left(\mu+\sqrt{1+\mu^{2}}\right)}\leq 1-\delta, (40)

we have

12ptvtL2p2p+δμ|vt|2|vt|2p2L1+ReνvtL2p+2m2p+2m+vtL2p2p\displaystyle\frac{1}{2p}\partial_{t}\|v_{t}\|_{L^{2p}}^{2p}+\delta\mu\left\||\nabla v_{t}|^{2}|v_{t}|^{2p-2}\right\|_{L^{1}}+\mathrm{Re}\nu\|v_{t}\|_{L^{2p+2m}}^{2p+2m}+\|v_{t}\|_{L^{2p}}^{2p}
\displaystyle\leq |vt|2p2,Re(vt¯Ψ(vt,Z¯t)),\displaystyle\,\left\langle|v_{t}|^{2p-2},\mathrm{Re}\left(\overline{v_{t}}\Psi^{\prime}\left(v_{t},\underline{Z}_{t}\right)\right)\right\rangle,

for every t(0,T]t\in(0,T]. In particular, if we integrate with respect to tt, then we have for all 0<s0tT0<s_{0}\leq t\leq T,

12p(vtL2p2pvs0L2p2p)+δμs0t|vs|2|vs|2p2L1ds\displaystyle\frac{1}{2p}\left(\|v_{t}\|_{L^{2p}}^{2p}-\left\|v_{s_{0}}\right\|_{L^{2p}}^{2p}\right)+\delta\mu\int_{s_{0}}^{t}\left\||\nabla v_{s}|^{2}|v_{s}|^{2p-2}\right\|_{L^{1}}\mathrm{d}s
+\displaystyle+ Reνs0tvsL2p+2m2p+2mds+s0tvsL2p2pdss0t|vs|2p2,Re(vs¯Ψ(vs,Z¯s))ds.\displaystyle\,\mathrm{Re}\nu\int_{s_{0}}^{t}\|v_{s}\|_{L^{2p+2m}}^{2p+2m}\mathrm{d}s+\int_{s_{0}}^{t}\|v_{s}\|_{L^{2p}}^{2p}\mathrm{d}s\leq\int_{s_{0}}^{t}\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}\Psi^{\prime}\left(v_{s},\underline{Z}_{s}\right)\right)\right\rangle\mathrm{d}s.
{proof}

By Proposition 4.3, it suffices to show that for every 0δ<10\leq\delta<1 satisfying (40),

δμ|vt|2|vt|2p2L1Re[(i+μ){(vtvt¯)p1vt¯},vt],t(0,T].\delta\mu\left\||\nabla v_{t}|^{2}|v_{t}|^{2p-2}\right\|_{L^{1}}\leq\mathrm{Re}\left[(\mathrm{i}+\mu)\left\langle\nabla\left\{(v_{t}\overline{v_{t}})^{p-1}\overline{v_{t}}\right\},\nabla v_{t}\right\rangle\right],\quad t\in(0,T]. (41)

Note that

2Re[(i+μ){(vtvt¯)p1vt¯},vt]\displaystyle 2\mathrm{Re}\left[(\mathrm{i}+\mu)\left\langle\nabla\left\{(v_{t}\overline{v_{t}})^{p-1}\overline{v_{t}}\right\},\nabla v_{t}\right\rangle\right]
=\displaystyle= (i+μ)[p|vt|2p2,|vt|2+(p1)|vt|2p4,vt¯2(vt)2]\displaystyle\,(\mathrm{i}+\mu)\left[p\left\langle\left|v_{t}\right|^{2p-2},\left|\nabla v_{t}\right|^{2}\right\rangle+(p-1)\left\langle\left|v_{t}\right|^{2p-4},\overline{v_{t}}^{2}\left(\nabla v_{t}\right)^{2}\right\rangle\right]
+(i+μ)[p|vt|2p2,|vt|2+(p1)|vt|2p4,vt2(vt¯)2]\displaystyle\,+(-\mathrm{i}+\mu)\left[p\left\langle\left|v_{t}\right|^{2p-2},\left|\nabla v_{t}\right|^{2}\right\rangle+(p-1)\left\langle\left|v_{t}\right|^{2p-4},{v_{t}}^{2}\left(\nabla\overline{v_{t}}\right)^{2}\right\rangle\right]
=\displaystyle=  2pμ|vt|2p4,|vt|2|vt|2+2(p1)|vt|2p4,μxt2μyt22xtyt\displaystyle\,2p\mu\left\langle\left|v_{t}\right|^{2p-4},\left|v_{t}\right|^{2}\left|\nabla v_{t}\right|^{2}\right\rangle+2(p-1)\left\langle\left|v_{t}\right|^{2p-4},\mu x_{t}^{2}-\mu y_{t}^{2}-2x_{t}y_{t}\right\rangle
=\displaystyle=  2pμ|vt|2p4,xt2+yt2+2(p1)|vt|2p4,μxt2μyt22xtyt\displaystyle\,2p\mu\left\langle\left|v_{t}\right|^{2p-4},x_{t}^{2}+y_{t}^{2}\right\rangle+2(p-1)\left\langle\left|v_{t}\right|^{2p-4},\mu x_{t}^{2}-\mu y_{t}^{2}-2x_{t}y_{t}\right\rangle
=\displaystyle=  2|vt|2p4,μ(2p1)xt2+μyt22(p1)xtyt,\displaystyle\,2\left\langle|v_{t}|^{2p-4},\mu(2p-1)x_{t}^{2}+\mu y_{t}^{2}-2(p-1)x_{t}y_{t}\right\rangle,

where xt=12(vtvt¯)=12|vt|2x_{t}=\frac{1}{2}\nabla(v_{t}\overline{v_{t}})=\frac{1}{2}\nabla\left|v_{t}\right|^{2} and yt=i2(vtvt¯vt¯vt)y_{t}=\frac{\mathrm{i}}{2}(v_{t}\nabla\overline{v_{t}}-\overline{v_{t}}\nabla v_{t}) are the real and imaginary parts of vt¯vt\overline{v_{t}}\nabla v_{t} respectively. Then

2Re[(i+μ){(vtvt¯)p1vt¯},vt]2δμ|vt|2|vt|2p2L1\displaystyle 2\mathrm{Re}\left[(\mathrm{i}+\mu)\left\langle\nabla\left\{(v_{t}\overline{v_{t}})^{p-1}\overline{v_{t}}\right\},\nabla v_{t}\right\rangle\right]-2\delta\mu\left\||\nabla v_{t}|^{2}|v_{t}|^{2p-2}\right\|_{L^{1}}
=\displaystyle=  2Re[(i+μ){(vtvt¯)p1vt¯},vt]2δμ|vt|2p4,xt2+yt2\displaystyle\,2\mathrm{Re}\left[(\mathrm{i}+\mu)\left\langle\nabla\left\{(v_{t}\overline{v_{t}})^{p-1}\overline{v_{t}}\right\},\nabla v_{t}\right\rangle\right]-2\delta\mu\left\langle|v_{t}|^{2p-4},x_{t}^{2}+y_{t}^{2}\right\rangle
=\displaystyle=  2|vt|2p4,μ(2p1δ)xt2+μ(1δ)yt22(p1)xtyt.\displaystyle\,2\left\langle|v_{t}|^{2p-4},\mu(2p-1-\delta)x_{t}^{2}+\mu(1-\delta)y_{t}^{2}-2(p-1)x_{t}y_{t}\right\rangle.

To prove (41), it suffices to prove that the quadratic form

μ(2p1δ)xt2+μ(1δ)yt22(p1)xtyt\mu(2p-1-\delta)x_{t}^{2}+\mu(1-\delta)y_{t}^{2}-2(p-1)x_{t}y_{t}

is nonnegative. The matrix corresponding to this quadratic form is

(μ(2p1δ)(p1)(p1)μ(1δ)).\begin{pmatrix}\mu(2p-1-\delta)&-(p-1)\\ -(p-1)&\mu(1-\delta)\end{pmatrix}.

Since μ(2p1δ)>0\mu(2p-1-\delta)>0 and when δ\delta satisfies (40), the determinant of this matrix is nonnegative, we complete the proof.

Proposition 4.6

Let vC((0,T];𝒞β)v\in C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) be a mild solution to (33) and 1p1+μ(μ+1+μ2)1\leq p\leq 1+\mu\left(\mu+\sqrt{1+\mu^{2}}\right). Then for every 0<tT0<t\leq T,

vtL2p2pC[tpm(1+t2αmp~Z¯α,α,Tp~)pp+m].\left\|v_{t}\right\|_{L^{2p}}^{2p}\leq C\left[t^{-\frac{p}{m}}\vee\left(1+t^{-2\alpha^{\prime}m\tilde{p}}\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},T}^{\tilde{p}}\right)^{\frac{p}{p+m}}\right].

where the constant p~>0\tilde{p}>0 and α,α(0,1)\alpha,\alpha^{\prime}\in(0,1) are small enough.

{proof}

By Proposition 4.5, to estimate vsL2p2p\left\|v_{s}\right\|_{L^{2p}}^{2p} for 0<st0<s\leq t, we need to estimate |vs|2p2,Re(vs¯Ψ(vs,Z¯s))\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}\Psi^{\prime}\left(v_{s},\underline{Z}_{s}\right)\right)\right\rangle. By the definition of Ψ(vs,Z¯s)\Psi^{\prime}\left(v_{s},\underline{Z}_{s}\right) (see (32)),

|vs|2p2,Re(vs¯Ψ(vs,Z¯s))\displaystyle\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}\Psi^{\prime}\left(v_{s},\underline{Z}_{s}\right)\right)\right\rangle (42)
=\displaystyle= |vs|2p2,Re(νi+j<2m+1(m+1i)(mj)vsivs¯j+1Z0,s:m+1i,mj:)\displaystyle\,\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(-\nu\sum_{i+j<2m+1}\binom{m+1}{i}\binom{m}{j}v_{s}^{i}\overline{v_{s}}^{j+1}Z_{0,s}^{:m+1-i,m-j:}\right)\right\rangle
+|vs|2p2,Re((τ+1)(|vs|2+vs¯Z0,s)).\displaystyle\,+\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left((\tau+1)(|v_{s}|^{2}+\overline{v_{s}}Z_{0,s})\right)\right\rangle.

Let

As=vsL2p+2m2p+2m,Bs=|vs|2|vs|2p2L1.A_{s}=\left\|v_{s}\right\|_{L^{2p+2m}}^{2p+2m},\quad B_{s}=\left\||\nabla v_{s}|^{2}|v_{s}|^{2p-2}\right\|_{L^{1}}. (43)

We show that for 0<st0<s\leq t, each term of |vs|2p2,Re(vs¯Ψ(vs,Z¯s))\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}\Psi^{\prime}\left(v_{s},\underline{Z}_{s}\right)\right)\right\rangle, namely vsL2p2p\left\|v_{s}\right\|_{L^{2p}}^{2p}, |vs|2p2,Re(vs¯Z0,s)\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}Z_{0,s}\right)\right\rangle and

|vs|2p2,Re(vsivs¯j+1Z0,s:m+1i,mj:),0im+1,0jm,(i,j)(m+1,m),\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(v_{s}^{i}\overline{v_{s}}^{j+1}Z_{0,s}^{:m+1-i,m-j:}\right)\right\rangle,\quad 0\leq i\leq m+1,0\leq j\leq m,(i,j)\neq(m+1,m),

can be controlled by AsA_{s} and BsB_{s}. By Proposition 2.7, for α(0,1)\alpha\in(0,1),

||vs|2p2,Re(vsivs¯j+1Z0,s:m+1i,mj:)|\displaystyle\left|\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(v_{s}^{i}\overline{v_{s}}^{j+1}Z_{0,s}^{:m+1-i,m-j:}\right)\right\rangle\right| (44)
\displaystyle\leq |vs|2p+i+j1,|Z0,s:m+1i,mj:|\displaystyle\,\left\langle|v_{s}|^{2p+i+j-1},\left|Z_{0,s}^{:m+1-i,m-j:}\right|\right\rangle
\displaystyle\lesssim |vs|2p+i+j11,1αZ0,s:m+1i,mj:𝒞α.\displaystyle\,\left\||v_{s}|^{2p+i+j-1}\right\|_{\mathcal{B}_{1,1}^{\alpha}}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}.

By Proposition 2.8,

|vs|2p+i+j11,1α|vs|2p+i+j1L11α|vs|2p+i+j1L1α+|vs|2p+i+j1L1.\left\||v_{s}|^{2p+i+j-1}\right\|_{\mathcal{B}_{1,1}^{\alpha}}\lesssim\left\||v_{s}|^{2p+i+j-1}\right\|_{L^{1}}^{1-\alpha}\left\|\nabla|v_{s}|^{2p+i+j-1}\right\|_{L^{1}}^{\alpha}+\left\||v_{s}|^{2p+i+j-1}\right\|_{L^{1}}. (45)

By the Cauchy-Schwarz inequality,

|vs|2p+i+j1L1Bs12|vs|2(p+i+j1)L112.\left\|\nabla|v_{s}|^{2p+i+j-1}\right\|_{L^{1}}\lesssim B_{s}^{\frac{1}{2}}\left\||v_{s}|^{2(p+i+j-1)}\right\|_{L^{1}}^{\frac{1}{2}}. (46)

By Sobolev inequality fLq(fL22+fL22)12\left\|f\right\|_{L^{q}}\lesssim\left(\left\|f\right\|_{L^{2}}^{2}+\left\|\nabla f\right\|_{L^{2}}^{2}\right)^{\frac{1}{2}} for every q<q<\infty (see [5, Theorem 6.5.1] or [33, Theorem 2.13]), taking q=2(p+i+j1)pq=\frac{2(p+i+j-1)}{p} specifically and combining with Jensen’s inequality, we have that

|vs|2(p+i+j1)L112|vs|2pL1p+i+j12p+Bsp+i+j12pAsp+i+j12p+2m+Bsp+i+j12p\left\||v_{s}|^{2(p+i+j-1)}\right\|_{L^{1}}^{\frac{1}{2}}\lesssim\left\||v_{s}|^{2p}\right\|_{L^{1}}^{\frac{p+i+j-1}{2p}}+B_{s}^{\frac{p+i+j-1}{2p}}\lesssim A_{s}^{\frac{p+i+j-1}{2p+2m}}+B_{s}^{\frac{p+i+j-1}{2p}} (47)

Moreover, by Jensen’s inequality again,

|vs|2p+i+j1L1As2p+i+j12p+2m.\left\||v_{s}|^{2p+i+j-1}\right\|_{L^{1}}\lesssim A_{s}^{\frac{2p+i+j-1}{2p+2m}}. (48)

Therefore, combining with (45), (46), (47) and (48),

|vs|2p+i+j11,1αAs2p+i+j1αp2p+2mBsα2+As2p+i+j12p+2m(1α)Bs2p+i+j12pα+As2p+i+j12p+2m.\left\||v_{s}|^{2p+i+j-1}\right\|_{\mathcal{B}_{1,1}^{\alpha}}\lesssim A_{s}^{\frac{2p+i+j-1-\alpha p}{2p+2m}}B_{s}^{\frac{\alpha}{2}}+A_{s}^{\frac{2p+i+j-1}{2p+2m}(1-\alpha)}B_{s}^{\frac{2p+i+j-1}{2p}\alpha}+A_{s}^{\frac{2p+i+j-1}{2p+2m}}. (49)

Here we take 0<α<1mpm(2m1)0<\alpha<\frac{1}{m}\wedge\frac{p}{m(2m-1)} such that

2p+2m1αp2p+2m+α2<1,2p+2m12p+2m(1α)+2p+2m12pα<1.\frac{2p+2m-1-\alpha p}{2p+2m}+\frac{\alpha}{2}<1,\quad\frac{2p+2m-1}{2p+2m}(1-\alpha)+\frac{2p+2m-1}{2p}\alpha<1.

Then we can fine exponents 0<γi,jk<10<\gamma_{i,j}^{k}<1, 1k41\leq k\leq 4, such that

1γi,j12p+2m1αp2p+2m+α2γi,j2=1,1γi,j32p+2m12p+2m(1α)+2p+2m12pγi,j4α=1.\frac{1}{\gamma_{i,j}^{1}}\frac{2p+2m-1-\alpha p}{2p+2m}+\frac{\alpha}{2\gamma_{i,j}^{2}}=1,\quad\frac{1}{\gamma_{i,j}^{3}}\frac{2p+2m-1}{2p+2m}(1-\alpha)+\frac{2p+2m-1}{2p\gamma_{i,j}^{4}}\alpha=1.

Then by (44), (49) and Young’s inequality in the form abaq1q1+bq2q2ab\leq\frac{a^{q_{1}}}{q_{1}}+\frac{b^{q_{2}}}{q_{2}}, where a,b0a,b\geq 0, q1,q2>1q_{1},q_{2}>1 satisfy 1q1+1q2=1\frac{1}{q_{1}}+\frac{1}{q_{2}}=1, we can obtain that

||vs|2p2,Re(vsivs¯j+1Z0,s:m+1i,mj:)|\displaystyle\left|\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(v_{s}^{i}\overline{v_{s}}^{j+1}Z_{0,s}^{:m+1-i,m-j:}\right)\right\rangle\right|
\displaystyle\lesssim (Asγi,j1+Bsγi,j2+Asγi,j3+Bsγi,j4+Asγi,j5)Z0,s:m+1i,mj:𝒞α,\displaystyle\,\left(A_{s}^{\gamma_{i,j}^{1}}+B_{s}^{\gamma_{i,j}^{2}}+A_{s}^{\gamma_{i,j}^{3}}+B_{s}^{\gamma_{i,j}^{4}}+A_{s}^{\gamma_{i,j}^{5}}\right)\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}},

where γi,j5=2p+i+j12p+2m\gamma_{i,j}^{5}=\frac{2p+i+j-1}{2p+2m}. Using Young’s inequality, now in the form

abϵaq1+C(ϵ)bq2,ab\leq\epsilon a^{q_{1}}+C(\epsilon)b^{q_{2}}, (50)

where ϵ>0\epsilon>0 is arbitrary, a,b0a,b\geq 0, q1,q2>1q_{1},q_{2}>1 satisfy 1q1+1q2=1\frac{1}{q_{1}}+\frac{1}{q_{2}}=1 and C(ϵ)=1q2(q1ϵ)q2q1C(\epsilon)=\frac{1}{q_{2}}(q_{1}\epsilon)^{-\frac{q_{2}}{q_{1}}}, we obtain the bound

|ν|(m+1i)(mj)||vs|2p2,Re(vsivs¯j+1Z0,s:m+1i,mj:)|\displaystyle|\nu|\binom{m+1}{i}\binom{m}{j}\left|\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(v_{s}^{i}\overline{v_{s}}^{j+1}Z_{0,s}^{:m+1-i,m-j:}\right)\right\rangle\right| (51)
\displaystyle\leq 1(m+1)(m+2)(Reν8As+δμ4Bs)+Ck=15Z0,s:m+1i,mj:𝒞α11γi,jk,\displaystyle\,\frac{1}{(m+1)(m+2)}\left(\frac{\mathrm{Re}\nu}{8}A_{s}+\frac{\delta\mu}{4}B_{s}\right)+C\sum_{k=1}^{5}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\frac{1}{1-\gamma_{i,j}^{k}}},

where 0δ<10\leq\delta<1 satisfy (40) and CC is a constant depending only on mm, ν\nu and γi,jk\gamma_{i,j}^{k}, 1k51\leq k\leq 5.

Now we estimate |vs|2p2,Re(vs¯Z0,s)\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}Z_{0,s}\right)\right\rangle. By the similar argument as above, we have that

(|τ|+1)||vs|2p2,Re(vs¯Z0,s)|\displaystyle(|\tau|+1)\left|\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}Z_{0,s}\right)\right\rangle\right| (52)
\displaystyle\leq 1(m+1)(m+2)(Reν8As+δμ4Bs)+Ck=15Z0,s𝒞α11γ0,0k,\displaystyle\,\frac{1}{(m+1)(m+2)}\left(\frac{\mathrm{Re}\nu}{8}A_{s}+\frac{\delta\mu}{4}B_{s}\right)+C\sum_{k=1}^{5}\left\|Z_{0,s}\right\|_{\mathcal{C}^{-\alpha}}^{\frac{1}{1-\gamma_{0,0}^{k}}},

where CC is a constant depending only on mm, ν\nu and τ\tau.

Next we estimate vsL2p2p\left\|v_{s}\right\|_{L^{2p}}^{2p}. By Jensen’s inequality and Young’s inequality in the form (50), we have that

(|τ|+1)vsL2p2pC(|τ|+1)Aspp+mReν4As+C,(|\tau|+1)\left\|v_{s}\right\|_{L^{2p}}^{2p}\leq C(|\tau|+1)A_{s}^{\frac{p}{p+m}}\leq\frac{\mathrm{Re}\nu}{4}A_{s}+C, (53)

where C>0C>0 is a constant depending on ν,p,m\nu,p,m and τ\tau.

Let pm,mk=11γ0,0k11γm,mkp_{m,m}^{k}=\frac{1}{1-\gamma_{0,0}^{k}}\vee\frac{1}{1-\gamma_{m,m}^{k}}, pm+1,mk=1p_{m+1,m}^{k}=1 and pi,jk=11γi,jkp_{i,j}^{k}=\frac{1}{1-\gamma_{i,j}^{k}} for 1k51\leq k\leq 5, 0im+10\leq i\leq m+1, 0jm0\leq j\leq m and (i,j)(m+1,m),(m,m)(i,j)\neq(m+1,m),(m,m). Then combining with (42), (51), (52) and (53), we get that

||vs|2p2,Re(vs¯Ψ(vs,Z¯s))|Reν2As+δμ2Bs+Ci=0m+1j=0mk=15Z0,s:m+1i,mj:𝒞αpi,jk,\left|\left\langle|v_{s}|^{2p-2},\mathrm{Re}\left(\overline{v_{s}}\Psi^{\prime}\left(v_{s},\underline{Z}_{s}\right)\right)\right\rangle\right|\leq\frac{\mathrm{Re}\nu}{2}A_{s}+\frac{\delta\mu}{2}B_{s}+C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\sum_{k=1}^{5}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{p_{i,j}^{k}}, (54)

where C>0C>0 is a constant depending on ν,p,m,τ\nu,p,m,\tau and all pi,jkp_{i,j}^{k}. Finally, by Proposition 4.5, we have that

12psvsL2p2p+δμ2|vs|2|vs|2p2L1+Reν2vsL2p+2m2p+2m+vsL2p2p\displaystyle\frac{1}{2p}\partial_{s}\|v_{s}\|_{L^{2p}}^{2p}+\frac{\delta\mu}{2}\left\||\nabla v_{s}|^{2}|v_{s}|^{2p-2}\right\|_{L^{1}}+\frac{\mathrm{Re}\nu}{2}\|v_{s}\|_{L^{2p+2m}}^{2p+2m}+\|v_{s}\|_{L^{2p}}^{2p} (55)
\displaystyle\leq Ci=0m+1j=0mk=15Z0,s:m+1i,mj:𝒞αpi,jk\displaystyle C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\sum_{k=1}^{5}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{p_{i,j}^{k}}
\displaystyle\leq Ci=0m+1j=0mZ0,s:m+1i,mj:𝒞αp~,\displaystyle C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\tilde{p}},

where

p~=max{pi,jk:1k5,0im+1,0jm,}.\tilde{p}=\max\left\{p_{i,j}^{k}:1\leq k\leq 5,0\leq i\leq m+1,0\leq j\leq m,\right\}. (56)

Let s<ts<t and α(0,1)\alpha^{\prime}\in(0,1) be small enough. Thus for r[s,t]r\in[s,t],

12prvrL2p2p+Reν2vrL2p+2m2p+2m\displaystyle\frac{1}{2p}\partial_{r}\left\|v_{r}\right\|_{L^{2p}}^{2p}+\frac{\mathrm{Re}\nu}{2}\|v_{r}\|_{L^{2p+2m}}^{2p+2m}
\displaystyle\leq C+Ci+j<2m+1sα(2mij)p~supsrtrα(2mij)p~Z0,r:m+1i,mj:𝒞αp~\displaystyle\,C+C\sum_{i+j<2m+1}s^{-\alpha^{\prime}(2m-i-j)\tilde{p}}\sup_{s\leq r\leq t}r^{\alpha^{\prime}(2m-i-j)\tilde{p}}\left\|Z_{0,r}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\tilde{p}}
\displaystyle\leq C+Cs2αmp~Z¯α,α,tp~.\displaystyle\,C+Cs^{-2\alpha^{\prime}m\tilde{p}}\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},t}^{\tilde{p}}.

By Jensen’s inequality,

rvrL2p2p+C1(vrL2p2p)p+mp\displaystyle\partial_{r}\left\|v_{r}\right\|_{L^{2p}}^{2p}+C_{1}\left(\left\|v_{r}\right\|_{L^{2p}}^{2p}\right)^{\frac{p+m}{p}} rvrL2p2p+C1vrL2p+2m2p+2m\displaystyle\leq\partial_{r}\left\|v_{r}\right\|_{L^{2p}}^{2p}+C_{1}\left\|v_{r}\right\|_{L^{2p+2m}}^{2p+2m}
C2(1+s2αmp~Z¯α,α,tp~).\displaystyle\leq C_{2}\left(1+s^{-2\alpha^{\prime}m\tilde{p}}\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},t}^{\tilde{p}}\right).

Let f(r)=vrL2p2pf(r)=\left\|v_{r}\right\|_{L^{2p}}^{2p} for r[s,t]r\in[s,t]. Then by [35, Lemma 3.8],

f(r)\displaystyle f(r) f(s)(1+m2pC1(rs)f(s)mp)pm[2C2C1(1+s2αmp~Z¯α,α,tp~)]pp+m\displaystyle\leq\frac{f(s)}{\left(1+\frac{m}{2p}C_{1}(r-s)f(s)^{\frac{m}{p}}\right)^{\frac{p}{m}}}\vee\left[\frac{2C_{2}}{C_{1}}\left(1+s^{-2\alpha^{\prime}m\tilde{p}}\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},t}^{\tilde{p}}\right)\right]^{\frac{p}{p+m}}
C3(rs)pm[2C2C1(1+s2αmp~Z¯α,α,tp~)]pp+m\displaystyle\leq C_{3}(r-s)^{-\frac{p}{m}}\vee\left[\frac{2C_{2}}{C_{1}}\left(1+s^{-2\alpha^{\prime}m\tilde{p}}\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},t}^{\tilde{p}}\right)\right]^{\frac{p}{p+m}}

Let r=tr=t, s=t/2s=t/2 and then

vtL2p2p\displaystyle\left\|v_{t}\right\|_{L^{2p}}^{2p} C[tpm(1+t2αmp~Z¯α,α,tp~)pp+m]\displaystyle\leq C\left[t^{-\frac{p}{m}}\vee\left(1+t^{-2\alpha^{\prime}m\tilde{p}}\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},t}^{\tilde{p}}\right)^{\frac{p}{p+m}}\right]
C[tpm(1+t2αmp~Z¯α,α,Tp~)pp+m].\displaystyle\leq C\left[t^{-\frac{p}{m}}\vee\left(1+t^{-2\alpha^{\prime}m\tilde{p}}\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},T}^{\tilde{p}}\right)^{\frac{p}{p+m}}\right].

Combining Proposition 4.6 with (55), we can get the following corollary.

Corollary 4.7

Let vC((0,T];𝒞β)v\in C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) be a mild solution to (33). Let 1p1+μ(μ+1+μ2)1\leq p\leq 1+\mu\left(\mu+\sqrt{1+\mu^{2}}\right) and 0δ<10\leq\delta<1 satisfying (40). Then for every 1tT1\leq t\leq T, there exists γ(p)>1\gamma(p)>1, C(Z¯α,α,1)>0C(\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})>0 depending on Z¯α,α,1\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1} and C>0C>0 such that

12pvtL2p2p+δμ21t|vs|2|vs|2p2L1ds+Reν21tvsL2p+2m2p+2mds+1tvsL2p2pds\displaystyle\frac{1}{2p}\|v_{t}\|_{L^{2p}}^{2p}+\frac{\delta\mu}{2}\int_{1}^{t}\left\||\nabla v_{s}|^{2}|v_{s}|^{2p-2}\right\|_{L^{1}}\mathrm{d}s+\frac{\mathrm{Re}\nu}{2}\int_{1}^{t}\|v_{s}\|_{L^{2p+2m}}^{2p+2m}\mathrm{d}s+\int_{1}^{t}\|v_{s}\|_{L^{2p}}^{2p}\mathrm{d}s
\displaystyle\leq C(Z¯α,α,1)+Ci=0m+1j=0m1tZ0,s:m+1i,mj:𝒞αγ(p)ds.\displaystyle\,C(\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})+C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{1}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma(p)}\mathrm{d}s.
{proof}

By (55), we obtain that

12pvtL2p2p+δμ21t|vs|2|vs|2p2L1ds+Reν21tvsL2p+2m2p+2mds+1tvsL2p2pds\displaystyle\frac{1}{2p}\|v_{t}\|_{L^{2p}}^{2p}+\frac{\delta\mu}{2}\int_{1}^{t}\left\||\nabla v_{s}|^{2}|v_{s}|^{2p-2}\right\|_{L^{1}}\mathrm{d}s+\frac{\mathrm{Re}\nu}{2}\int_{1}^{t}\|v_{s}\|_{L^{2p+2m}}^{2p+2m}\mathrm{d}s+\int_{1}^{t}\|v_{s}\|_{L^{2p}}^{2p}\mathrm{d}s
\displaystyle\leq 12pv1L2p2p+Ci=0m+1j=0m1tZ0,s:m+1i,mj:𝒞αp~ds,\displaystyle\,\frac{1}{2p}\|v_{1}\|_{L^{2p}}^{2p}+C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{1}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\tilde{p}}\mathrm{d}s,

where p~\tilde{p} is defined as (56). By Proposition 4.6

12pv1L2p2pC(1+Z¯α,α,1p~)pp+m.\frac{1}{2p}\|v_{1}\|_{L^{2p}}^{2p}\leq C\left(1+\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1}^{\tilde{p}}\right)^{\frac{p}{p+m}}.

Taking γ(p)=p~\gamma(p)=\tilde{p} and C(Z¯α,α,1)=C(1+Z¯α,α,1p~)pp+mC(\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})=C\left(1+\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1}^{\tilde{p}}\right)^{\frac{p}{p+m}}, we complete the proof.

4.3 Global existence and uniqueness

Theorem 4.8 (Global existence and uniqueness)

Let μ>0\mu>0 satisfy 2+2μ(μ+1+μ2)>2m+12+2\mu(\mu+\sqrt{1+\mu^{2}})>2m+1. For every u0𝒞α0u_{0}\in\mathcal{C}^{-\alpha_{0}} and α0,β\alpha_{0},\beta, γ\gamma satisfying

11+μ(μ+1+μ2)α022m+1,β>0,α0+β2<γ<12m+1,\frac{1}{1+\mu(\mu+\sqrt{1+\mu^{2}})}\leq\alpha_{0}\leq\frac{2}{2m+1},\quad\beta>0,\quad\frac{\alpha_{0}+\beta}{2}<\gamma<\frac{1}{2m+1},

there exists a unique mild solution vC((0,);𝒞β)v\in C\left(\left(0,\infty\right);\mathcal{C}^{\beta}\right) of (33).

{proof}

Let T>0T>0. By Theorem 4.2, there exists TTT^{*}\leq T and a unique mild solution up to time TT^{*} of (33). Under the assumptions that μ\mu satisfies 2m+1<2+2μ(μ+1+μ2)2m+1<2+2\mu(\mu+\sqrt{1+\mu^{2}}) and α0[1/(1+μ(μ+1+μ2)),2/(2m+1)]\alpha_{0}\in\left[1/\left(1+\mu(\mu+\sqrt{1+\mu^{2}})\right),2/\left(2m+1\right)\right], by Proposition 2.2 and 2.3, we can find 1p1+μ(μ+1+μ2)1\leq p\leq 1+\mu(\mu+\sqrt{1+\mu^{2}}) such that L2p𝒞α0L^{2p}\hookrightarrow\mathcal{C}^{-\alpha_{0}}. Then Proposition 4.6 provides a priori estimate of vT𝒞α0\left\|v_{T^{*}}\right\|_{\mathcal{C}^{-\alpha_{0}}} depending only on Z¯α,α,T\|\underline{Z}\|_{\alpha,\alpha^{\prime},T}. Using Theorem 4.2 again, there exists TT^{***} bounded from below by a constant depending on the priori estimate of vT𝒞α0\left\|v_{T^{*}}\right\|_{\mathcal{C}^{-\alpha_{0}}} (see (37)), such that (33) has a unique mild solution on [T,(T+T)T][T^{*},(T^{*}+T^{***})\wedge T] with initial condition vTv_{T^{*}}. We then continue this process until the whole interval [0,T][0,T] is covered.

Corollary 4.9

Let μ>0\mu>0 satisfy 2+2μ(μ+1+μ2)>2m+12+2\mu(\mu+\sqrt{1+\mu^{2}})>2m+1. Let α0,α[1/(1+μ(μ+1+μ2)),2/(2m+1)]\alpha_{0},\alpha\in\left[1/\left(1+\mu(\mu+\sqrt{1+\mu^{2}})\right),2/\left(2m+1\right)\right] and 1p1+μ(μ+1+μ2)1\leq p\leq 1+\mu(\mu+\sqrt{1+\mu^{2}}). For every u0𝒞α0u_{0}\in\mathcal{C}^{-\alpha_{0}}, let u(;u0)=Z0,+vu(\cdot;u_{0})=Z_{0,\cdot}+v_{\cdot}, where vv is a mild solution to (33). Then

supu0𝒞α0supt0(tpm1)E[u(t;u0)𝒞α2p]<.\sup_{u_{0}\in\mathcal{C}^{-\alpha_{0}}}\sup_{t\geq 0}\left(t^{\frac{p}{m}}\wedge 1\right)\mathrm{E}\left[\left\|u(t;u_{0})\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]<\infty.
{proof}

Let t>1t>1 and vt1,t:=u(t;u0)Zt1,tv_{t-1,t}:=u(t;u_{0})-Z_{t-1,t}. Under the assumptions that 2m+1<2+2μ(μ+1+μ2)2m+1<2+2\mu(\mu+\sqrt{1+\mu^{2}}) and α[1/(1+μ(μ+1+μ2)),2/(2m+1)]\alpha\in\left[1/\left(1+\mu(\mu+\sqrt{1+\mu^{2}})\right),2/\left(2m+1\right)\right], we can find 1p1+μ(μ+1+μ2)1\leq p\leq 1+\mu(\mu+\sqrt{1+\mu^{2}}) such that L2p𝒞αL^{2p}\hookrightarrow\mathcal{C}^{-\alpha}. Then by Proposition 3.2 and 4.6,

E[u(t;u0)𝒞α2p]\displaystyle\mathrm{E}\left[\left\|u(t;u_{0})\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]
\displaystyle\leq C(E[vt1,t𝒞α2p]+E[Zt1,t𝒞α2p])\displaystyle\,C\left(\mathrm{E}\left[\left\|v_{t-1,t}\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]+\mathrm{E}\left[\left\|Z_{t-1,t}\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]\right)
\displaystyle\leq C(E[vt1,tL2p2p]+E[Zt1,t𝒞α2p])\displaystyle\,C\left(\mathrm{E}\left[\left\|v_{t-1,t}\right\|_{L^{2p}}^{2p}\right]+\mathrm{E}\left[\left\|Z_{t-1,t}\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]\right)
\displaystyle\leq C+Csup0s1(Zt1,t1+s𝒞αpp~p+msupi,jLs(i+j1)αpp~p+mZt1,t1+s:i,j:𝒞αpp~p+m)\displaystyle\,C+C\sup_{0\leq s\leq 1}\left(\left\|Z_{t-1,t-1+s}\right\|_{\mathcal{C}^{-\alpha}}^{\frac{p\tilde{p}}{p+m}}\vee\sup_{i,j\in L}s^{(i+j-1)\alpha^{\prime}\frac{p\tilde{p}}{p+m}}\left\|Z_{t-1,t-1+s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{\frac{p\tilde{p}}{p+m}}\right)
\displaystyle\leq C,\displaystyle\,C,

where CC is a constant not depending on tt and the set LL is defined in (28). Then we get that

supt1E[u(t;u0)𝒞α2p]<.\sup_{t\geq 1}\mathrm{E}\left[\left\|u(t;u_{0})\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]<\infty.

For t1t\leq 1, by Proposition 4.6, we have that

E[u(t;u0)𝒞α2p]E[vtL2p2p]+E[Z0,t𝒞α2p]1+tpm.\mathrm{E}\left[\left\|u(t;u_{0})\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]\lesssim\mathrm{E}\left[\left\|v_{t}\right\|_{L^{2p}}^{2p}\right]+\mathrm{E}\left[\left\|Z_{0,t}\right\|_{\mathcal{C}^{-\alpha}}^{2p}\right]\lesssim 1+t^{-\frac{p}{m}}.

Then the proof is completed.

5 Ergodicity

5.1 Existence of invariant measure

In this section, we fix μ>0\mu>0 satisfying 2+2μ(μ+1+μ2)>2m+12+2\mu(\mu+\sqrt{1+\mu^{2}})>2m+1 and α0,β\alpha_{0},\beta, γ\gamma satisfying

11+μ(μ+1+μ2)α022m+1,β>0,α0+β2<γ<12m+1.\frac{1}{1+\mu(\mu+\sqrt{1+\mu^{2}})}\leq\alpha_{0}\leq\frac{2}{2m+1},\quad\beta>0,\quad\frac{\alpha_{0}+\beta}{2}<\gamma<\frac{1}{2m+1}.

Let u(;u0)=Z0,+vu(\cdot;u_{0})=Z_{0,\cdot}+v_{\cdot}, where vv is a mild solution to (33) with initial data u0𝒞α0u_{0}\in\mathcal{C}^{-\alpha_{0}}. We denote by b(𝒞α0)\mathcal{B}_{b}(\mathcal{C}^{-\alpha_{0}}) the space of bounded measurable functions from 𝒞α0\mathcal{C}^{-\alpha_{0}} to \mathbb{R}. For every Φb(𝒞α0)\Phi\in\mathcal{B}_{b}(\mathcal{C}^{-\alpha_{0}}) and t0t\geq 0, we define the map St:ΦStΦS_{t}:\Phi\rightarrow S_{t}\Phi by

StΦ(x)=E[Φ(u(t;x)],x𝒞α0.S_{t}\Phi(x)=\mathrm{E}\left[\Phi(u(t;x)\right],x\in\mathcal{C}^{-\alpha_{0}}.

Let (St)t0\left(S_{t}^{*}\right)_{t\geq 0} be the dual semigroup of (St)t0\left(S_{t}\right)_{t\geq 0}, that is,

Stη(A)=𝒞α0P(u(t;x)A)η(dx),S_{t}^{*}\eta(A)=\int_{\mathcal{C}^{-\alpha_{0}}}P(u(t;x)\in A)\eta(\mathrm{d}x),

where AA is a Borel subset of 𝒞α0\mathcal{C}^{-\alpha_{0}} and η\eta is probability measure on 𝒞α0\mathcal{C}^{-\alpha_{0}}. We set

𝒢t=σ({ξ(φ):φ|(t,)×𝕋20,φL2(×𝕋2,)})\mathcal{G}_{t}=\sigma\left(\left\{\xi(\varphi):\varphi|_{(t,\infty)\times\mathbb{T}^{2}}\equiv 0,\varphi\in L^{2}(\mathbb{R}\times\mathbb{T}^{2},\mathbb{C})\right\}\right)

for t>t>-\infty and let (t)t>\left(\mathcal{F}_{t}\right)_{t>-\infty} be the usual augmentation of the filtration (𝒢t)t>\left(\mathcal{G}_{t}\right)_{t>-\infty}. Since the solution of (33) is globally well-posed and depends continuously on the initial data, using the similar argument in [25, Section 4], we can prove that (u(t;u0))t0\left(u(t;u_{0})\right)_{t\geq 0} is a Feller Markov process on 𝒞α0\mathcal{C}^{-\alpha_{0}} with transition semigroup (St)t0\left(S_{t}\right)_{t\geq 0} with respect to the filtration (t)t0\left(\mathcal{F}_{t}\right)_{t\geq 0}. Then in the spirit of [35, Proposition 4.4], by Krylov-Bogoliubov Theorem, we prove that there exists a invariant measure of transition semigroup (St)t0\left(S_{t}\right)_{t\geq 0}.

Proposition 5.1

Let α0[1/(1+μ(μ+1+μ2)),2/(2m+1)]\alpha_{0}\in\left[1/\left(1+\mu(\mu+\sqrt{1+\mu^{2}})\right),2/\left(2m+1\right)\right]. For every x𝒞α0x\in\mathcal{C}^{-\alpha_{0}}, there exists a probability measure ηx\eta_{x} on 𝒞α0\mathcal{C}^{-\alpha_{0}} such that Stηx=ηxS_{t}^{*}\eta_{x}=\eta_{x} for all t0t\geq 0.

{proof}

Let t>0t>0, α0,α[1/(1+μ(μ+1+μ2)),2/(2m+1)]\alpha_{0},\alpha\in\left[1/\left(1+\mu(\mu+\sqrt{1+\mu^{2}})\right),2/\left(2m+1\right)\right], α<α0\alpha<\alpha_{0} and p[1,1+μ(μ+1+μ2)]p\in[1,1+\mu(\mu+\sqrt{1+\mu^{2}})]. By Markov’s and Jensen’s inequality, there exists a positive constant CC such that for K>0K>0,

P(u(t;x)𝒞α>K)CK(Eu(t;x)𝒞α2p)12p,P\left(\|u(t;x)\|_{\mathcal{C}^{-\alpha}}>K\right)\leq\frac{C}{K}\left(\mathrm{E}\|u(t;x)\|_{\mathcal{C}^{-\alpha}}^{2p}\right)^{\frac{1}{2p}},

Therefore, by Corollary 4.9,

0tP(u(s;x)𝒞α>K)ds\displaystyle\int_{0}^{t}P\left(\|u(s;x)\|_{\mathcal{C}^{-\alpha}}>K\right)\mathrm{d}s CK0t(Eu(s;x)𝒞α2p)12pds\displaystyle\leq\frac{C}{K}\int_{0}^{t}\left(\mathrm{E}\|u(s;x)\|_{\mathcal{C}^{-\alpha}}^{2p}\right)^{\frac{1}{2p}}\mathrm{d}s
CK[01s12mds+1t1ds]\displaystyle\leq\frac{C}{K}\left[\int_{0}^{1}s^{-\frac{1}{2m}}\mathrm{d}s+\int_{1}^{t}1\mathrm{d}s\right]
=CKt.\displaystyle=\frac{C}{K}t.

Let Qt=1t0tSsδxdsQ_{t}=\frac{1}{t}\int_{0}^{t}S_{s}^{*}\delta_{x}\mathrm{d}s, where δx\delta_{x} is the Dirac measure at x𝒞α0x\in\mathcal{C}^{-\alpha_{0}}. Then for Kε=CεK_{\varepsilon}=\frac{C}{\varepsilon}, we have that

Qt({f𝒞α:f𝒞α>Kε})=1t0tP(u(s;x)𝒞α>Kϵ)dsε.Q_{t}\left(\left\{f\in\mathcal{C}^{-\alpha}:\|f\|_{\mathcal{C}^{-\alpha}}>K_{\varepsilon}\right\}\right)=\frac{1}{t}\int_{0}^{t}P\left(\|u(s;x)\|_{\mathcal{C}^{-\alpha}}>K_{\epsilon}\right)\mathrm{d}s\leq\varepsilon.

Since the embedding 𝒞α𝒞α0\mathcal{C}^{-\alpha}\hookrightarrow\mathcal{C}^{-\alpha_{0}} is compact for α<α0\alpha<\alpha_{0}, {f𝒞α:f𝒞αKε}\left\{f\in\mathcal{C}^{-\alpha}:\|f\|_{\mathcal{C}^{-\alpha}}\leq K_{\varepsilon}\right\} is a compact subset of 𝒞α0\mathcal{C}^{-\alpha_{0}}. This implies that {Qt}t0\left\{Q_{t}\right\}_{t\geq 0} is tight in 𝒞α0\mathcal{C}^{-\alpha_{0}} and by the Krylov-Bogoliubov Theorem (see [11, Corollary 3.1.2]), there exists a probability measure ηx\eta_{x} on 𝒞α0\mathcal{C}^{-\alpha_{0}} such that Stηx=ηxS_{t}^{*}\eta_{x}=\eta_{x} for all t0t\geq 0.

5.2 Uniqueness of invariant measure

Using an asymptotic coupling argument (see [17, Corollary 2.2]), we prove the uniqueness of invariant measure of transition semigroup (St)t0\left(S_{t}\right)_{t\geq 0}. We first introduce some notations. Let 𝒫\mathcal{P} be a Markov transition kernel on a Polish space (𝕏,ρ)(\mathbb{X},\rho) and let 𝕏=𝕏\mathbb{X}_{\infty}=\mathbb{X}^{\mathbb{N}} be the space of one-sided infinite sequences with product topology. Denote by (𝕏)\mathcal{M}(\mathbb{X}), (𝕏)\mathcal{M}\left(\mathbb{X}_{\infty}\right) and (𝕏×𝕏)\mathcal{M}\left(\mathbb{X}_{\infty}\times\mathbb{X}_{\infty}\right) the space of all Borel probability measures on 𝕏\mathbb{X}, 𝕏×𝕏\mathbb{X}_{\infty}\times\mathbb{X}_{\infty} and 𝕏\mathbb{X}_{\infty} respectively. Let 𝒫:𝕏(𝕏)\mathcal{P}_{\infty}:\mathbb{X}\rightarrow\mathcal{M}\left(\mathbb{X}_{\infty}\right) be the probability kernel defined by stepping with the Markov kernel 𝒫\mathcal{P}. For η(𝕏)\eta\in\mathcal{M}(\mathbb{X}), we define η𝒫(𝕏)\eta\mathcal{P}_{\infty}\in\mathcal{M}\left(\mathbb{X}_{\infty}\right) as η𝒫=𝕏𝒫(x,)dη(x)\eta\mathcal{P}_{\infty}=\int_{\mathbb{X}}\mathcal{P}_{\infty}(x,\cdot)\mathrm{d}\eta(x). Given η1,η2(𝕏)\eta_{1},\eta_{2}\in\mathcal{M}(\mathbb{X}), we define the set of all generalized couplings as

𝒞~(η1𝒫,η2𝒫):={Γ(𝕏×𝕏):ΓΠi1ηi𝒫,i=1,2},\tilde{\mathcal{C}}\left(\eta_{1}\mathcal{P}_{\infty},\eta_{2}\mathcal{P}_{\infty}\right):=\left\{\Gamma\in\mathcal{M}\left(\mathbb{X}_{\infty}\times\mathbb{X}_{\infty}\right):\Gamma\circ\Pi_{i}^{-1}\ll\eta_{i}\mathcal{P}_{\infty},i=1,2\right\},

where Πi\Pi_{i} is the projection onto the ii-th coordinate. We define the diagonal at infinity as

D:={(x,y)𝕏×𝕏:limnρ(xn,yn)=0}.D:=\left\{(x,y)\in\mathbb{X}_{\infty}\times\mathbb{X}_{\infty}:\lim_{n\rightarrow\infty}\rho\left(x_{n},y_{n}\right)=0\right\}.

We recall an abstract result based on asymptotic argument in [17].

Theorem 5.2

[17, Corollary 2.2] If there exists a Borel measurable set B𝕏B\subset\mathbb{X} such that

  1. (i)

    η(B)>0\eta(B)>0 for any invariant probability measure η\eta of 𝒫\mathcal{P},

  2. (ii)

    there exists a measurable map B×B(x,y)Γx,y(𝕏×𝕏)B\times B\ni(x,y)\mapsto\Gamma_{x,y}\in\mathcal{M}\left(\mathbb{X}_{\infty}\times\mathbb{X}_{\infty}\right) such that Γx,y𝒞~(δx𝒫,δy𝒫)\Gamma_{x,y}\in\tilde{\mathcal{C}}\left(\delta_{x}\mathcal{P}_{\infty},\delta_{y}\mathcal{P}_{\infty}\right) and Γx,y(D)>0\Gamma_{x,y}(D)>0 for every x,yBx,y\in B.

Then there exists at most one invariant probability measure for 𝒫\mathcal{P}.

In order to construct an asymptotic coupling, we first consider the following auxiliary system

{tv~=[(i+μ)Δ1]v~+λ(vv~)+Ψ(v~,Z¯),t>0,x𝕋2,v(0,)=u1𝒞α0,\begin{cases}\partial_{t}\tilde{v}=\left[(\mathrm{i}+\mu)\Delta-1\right]\tilde{v}+\lambda\left(v-\tilde{v}\right)+\Psi(\tilde{v},\underline{Z}),&t>0,x\in\mathbb{T}^{2},\\ v(0,\cdot)=u_{1}\in\mathcal{C}^{-\alpha_{0}},\end{cases} (57)

where λ\lambda will be taken sufficiently large later. Compared to (33), there is a new dissipation term in (57) and the initial data is different. By the similar argument of the proof of Theorem 4.8, we can show that (57) is globally well-posed.

Theorem 5.3

Let μ>0\mu>0 satisfy 2+2μ(μ+1+μ2)>2m+12+2\mu(\mu+\sqrt{1+\mu^{2}})>2m+1. For every u1𝒞α0u_{1}\in\mathcal{C}^{-\alpha_{0}} and α0,β\alpha_{0},\beta, γ\gamma satisfying

11+μ(μ+1+μ2)α022m+1,β>0,α0+β2<γ<12m+1,\frac{1}{1+\mu(\mu+\sqrt{1+\mu^{2}})}\leq\alpha_{0}\leq\frac{2}{2m+1},\quad\beta>0,\quad\frac{\alpha_{0}+\beta}{2}<\gamma<\frac{1}{2m+1},

there exists a unique mild solution v~C((0,);𝒞β)\tilde{v}\in C\left(\left(0,\infty\right);\mathcal{C}^{\beta}\right) of (57).

Let u~:=v~+Z0,\tilde{u}:=\tilde{v}+Z_{0,\cdot}, then u~\tilde{u} solves the equation

{tu~=[(i+μ)Δ1]u~+λ(uu~)+Ψ(u~Z0,t,Z¯)+ξ,t>0,x𝕋2,u~(0,)=u1𝒞α0.\begin{cases}\partial_{t}\tilde{u}=\left[(\mathrm{i}+\mu)\Delta-1\right]\tilde{u}+\lambda\left(u-\tilde{u}\right)+\Psi(\tilde{u}-Z_{0,t},\underline{Z})+\xi,&t>0,x\in\mathbb{T}^{2},\\ \tilde{u}(0,\cdot)=u_{1}\in\mathcal{C}^{-\alpha_{0}}.\end{cases} (58)

Analogue to Corollary 4.7, we can prove the following proposition.

Proposition 5.4

Let v~C((0,T];𝒞β)\tilde{v}\in C\left(\left(0,T\right];\mathcal{C}^{\beta}\right) be a mild solution to (57). Let 1p1+μ(μ+1+μ2)1\leq p\leq 1+\mu\left(\mu+\sqrt{1+\mu^{2}}\right) and 0δ<10\leq\delta<1 satisfying (40). Then for every 1tT1\leq t\leq T, there exists γ(p)>1\gamma(p)>1, C(λ,Z¯α,α,1)>0C(\lambda,\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})>0 depending on λ\lambda, Z¯α,α,1\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1} and C>0C>0 such that

12pv~tL2p2p+δμ21t|v~s|2|v~s|2p2L1ds+Reν21tv~sL2p+2m2p+2mds+1tv~sL2p2pds\displaystyle\frac{1}{2p}\|\tilde{v}_{t}\|_{L^{2p}}^{2p}+\frac{\delta\mu}{2}\int_{1}^{t}\left\||\nabla\tilde{v}_{s}|^{2}|\tilde{v}_{s}|^{2p-2}\right\|_{L^{1}}\mathrm{d}s+\frac{\mathrm{Re}\nu}{2}\int_{1}^{t}\|\tilde{v}_{s}\|_{L^{2p+2m}}^{2p+2m}\mathrm{d}s+\int_{1}^{t}\|\tilde{v}_{s}\|_{L^{2p}}^{2p}\mathrm{d}s
\displaystyle\leq C(Z¯α,α,1)+Ci=0m+1j=0m1tZ0,s:m+1i,mj:𝒞αγ(p)ds.\displaystyle\,C(\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})+C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{1}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma(p)}\mathrm{d}s.
{proof}

Analogue to Proposition 4.3 and 4.5, we have that for every 0<stt0<s\leq t\leq t,

12psv~sL2p2p+δμ|v~s|2|v~s|2p2L1+Reνv~sL2p+2m2p+2m+(λ+1)v~sL2p2p\displaystyle\frac{1}{2p}\partial_{s}\|\tilde{v}_{s}\|_{L^{2p}}^{2p}+\delta\mu\left\||\nabla\tilde{v}_{s}|^{2}|\tilde{v}_{s}|^{2p-2}\right\|_{L^{1}}+\mathrm{Re}\nu\|\tilde{v}_{s}\|_{L^{2p+2m}}^{2p+2m}+(\lambda+1)\|\tilde{v}_{s}\|_{L^{2p}}^{2p} (59)
\displaystyle\leq |v~s|2p2,Re(v~s¯Ψ(v~s,Z¯s))+λ|v~s|2p2,Re(v~s¯vs).\displaystyle\,\left\langle|\tilde{v}_{s}|^{2p-2},\mathrm{Re}\left(\overline{\tilde{v}_{s}}\Psi^{\prime}\left(\tilde{v}_{s},\underline{Z}_{s}\right)\right)\right\rangle+\lambda\left\langle|\tilde{v}_{s}|^{2p-2},\mathrm{Re}\left(\overline{\tilde{v}_{s}}v_{s}\right)\right\rangle.

Using the similar argument to (54), we know that there exists γ(p)>1\gamma(p)>1 such that

||v~s|2p2,Re(v~s¯Ψ(v~s,Z¯s))|\displaystyle\left|\left\langle|\tilde{v}_{s}|^{2p-2},\mathrm{Re}\left(\overline{\tilde{v}_{s}}\Psi^{\prime}\left(\tilde{v}_{s},\underline{Z}_{s}\right)\right)\right\rangle\right| δμ2|v~s|2|v~s|2p2L1+Reν2v~sL2p+2m2p+2m\displaystyle\leq\frac{\delta\mu}{2}\left\||\nabla\tilde{v}_{s}|^{2}|\tilde{v}_{s}|^{2p-2}\right\|_{L^{1}}+\frac{\mathrm{Re}\nu}{2}\|\tilde{v}_{s}\|_{L^{2p+2m}}^{2p+2m} (60)
+Ci=0m+1j=0mZ0,s:m+1i,mj:𝒞αγ(p),\displaystyle\quad+C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma(p)},

where C>0C>0 is a constant depending on ν,p,m\nu,p,m, and τ\tau. For the term |v~s|2p2,Re(v~s¯vs)\left\langle|\tilde{v}_{s}|^{2p-2},\mathrm{Re}\left(\overline{\tilde{v}_{s}}v_{s}\right)\right\rangle, by Hölder’s and Young’s inequality,

||v~s|2p2,Re(v~s¯vs)|\displaystyle\left|\left\langle|\tilde{v}_{s}|^{2p-2},\mathrm{Re}\left(\overline{\tilde{v}_{s}}v_{s}\right)\right\rangle\right| |v~s|2p1,|vs|vsL2pv~sL2p2p1\displaystyle\leq\left\langle|\tilde{v}_{s}|^{2p-1},|v_{s}|\right\rangle\leq\left\|v_{s}\right\|_{L^{2p}}\left\|\tilde{v}_{s}\right\|_{L^{2p}}^{2p-1} (61)
12pvsL2p2p+2p12pv~sL2p2p12pvsL2p2p+v~sL2p2p.\displaystyle\leq\frac{1}{2p}\left\|v_{s}\right\|_{L^{2p}}^{2p}+\frac{2p-1}{2p}\left\|\tilde{v}_{s}\right\|_{L^{2p}}^{2p}\leq\frac{1}{2p}\left\|v_{s}\right\|_{L^{2p}}^{2p}+\left\|\tilde{v}_{s}\right\|_{L^{2p}}^{2p}.

Combining (59) (60), (61) and Corollary 4.7, we get that there exists C(λ,Z¯α,α,1)>0C(\lambda,\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})>0 depending on λ\lambda, Z¯α,α,1\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1} and C>0C>0 depending on ν,p,m,τ,λ\nu,p,m,\tau,\lambda such that

12psv~sL2p2p+δμ2|v~s|2|v~s|2p2L1+Reν2v~sL2p+2m2p+2m+v~sL2p2p\displaystyle\frac{1}{2p}\partial_{s}\|\tilde{v}_{s}\|_{L^{2p}}^{2p}+\frac{\delta\mu}{2}\left\||\nabla\tilde{v}_{s}|^{2}|\tilde{v}_{s}|^{2p-2}\right\|_{L^{1}}+\frac{\mathrm{Re}\nu}{2}\|\tilde{v}_{s}\|_{L^{2p+2m}}^{2p+2m}+\|\tilde{v}_{s}\|_{L^{2p}}^{2p} (62)
\displaystyle\leq C(λ,Z¯α,α,1)+Ci=0m+1j=0mZ0,s:m+1i,mj:𝒞αγ(p).\displaystyle\,C(\lambda,\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})+C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma(p)}.

By a similar argument to Proposition 4.6, we derive that

12pv~1L2p2pC(λ,Z¯α,α,1).\frac{1}{2p}\|\tilde{v}_{1}\|_{L^{2p}}^{2p}\leq C(\lambda,\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1}). (63)

Combining (62) with (63), we finish the proof.

For γ>0\gamma>0 and K>0K>0, we define the set

EK,γ:={Z¯α,α,1K,i=0m+1j=0m1TZ0,s:i,j:𝒞αγdsK(1+T),T1}.E_{K,\gamma}:=\left\{\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1}\leq K,\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{1}^{T}\left\|Z_{0,s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma}\mathrm{d}s\leq K(1+T),\forall T\geq 1\right\}. (64)

In the spirit of [32, Lemma 3.6], we prove the following Lemma.

Lemma 5.5

For every γ>0\gamma>0 and ε>0\varepsilon>0, there exists K>0K>0 such that P(EK,γ)1εP(E_{K,\gamma})\geq 1-\varepsilon.

{proof}

Note that

(Z,t:i,j:)t0,0im+1,0jm,\left(Z_{-\infty,t}^{:i,j:}\right)_{t\geq 0},\quad 0\leq i\leq m+1,\quad 0\leq j\leq m,

are stationary Markov processes. By [11, Theorem 3.3.1], we have that for every q1q\geq 1, there exists a random variable XL2(Ω,,P)X\in L^{2}(\Omega,\mathcal{F},P) such that

YT:=1Ti=0m+1j=0m0TZ,s:i,j:𝒞αqdsX,T,P-a.s.Y_{T}:=\frac{1}{T}\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{0}^{T}\left\|Z_{-\infty,s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{q}\mathrm{d}s\rightarrow X,\quad T\rightarrow\infty,\quad P\mbox{-a.s.}

This implies that for every ϵ>0\epsilon>0, there exists Ω1Ω\Omega_{1}\subset\Omega such that P(Ω1)<ϵ/2P(\Omega_{1})<\epsilon/2 and

supωΩ1|YT(ω)X(ω)|0,T.\sup_{\omega\notin\Omega_{1}}\left|Y_{T}(\omega)-X(\omega)\right|\rightarrow 0,\quad T\rightarrow\infty.

Then there exists T0T_{0} independent of ω\omega such that for all TT0T\geq T_{0},

YT(ω)X(ω)+1,ωΩ1.Y_{T}(\omega)\leq X(\omega)+1,\quad\forall\omega\notin\Omega_{1}.

Since XL2(Ω,,P)X\in L^{2}(\Omega,\mathcal{F},P), we have that for ϵ>0\epsilon>0, there exists Kϵ=2EX2/ϵ>0K_{\epsilon}=2\mathrm{E}X^{2}/\epsilon>0 such that by Markov’s inequality,

P(X>Kϵ)EX2Kϵ=ϵ2.P(X>K_{\epsilon})\leq\frac{\mathrm{E}X^{2}}{K_{\epsilon}}=\frac{\epsilon}{2}.

On Ω1c{XKϵ}\Omega_{1}^{c}\cap\left\{X\leq K_{\epsilon}\right\}, we have that

YT(ω)X(ω)+1Kϵ+1,TT0.Y_{T}(\omega)\leq X(\omega)+1\leq K_{\epsilon}+1,\quad\forall T\geq T_{0}.

Therefore, let K1=Kϵ+1K_{1}=K_{\epsilon}+1, we get that for any γ>0\gamma>0,

P(i=0m+1j=0m0TZ,s:i,j:𝒞α2γdsK1T,TT0)\displaystyle P\left(\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{0}^{T}\left\|Z_{-\infty,s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{2\gamma}\mathrm{d}s\leq K_{1}T,\forall T\geq T_{0}\right)
\displaystyle\geq P(Ω1c{XKϵ})P(Ω1c)+P({XKϵ})11ϵ.\displaystyle\,P\left(\Omega_{1}^{c}\cap\left\{X\leq K_{\epsilon}\right\}\right)\geq P\left(\Omega_{1}^{c}\right)+P\left(\left\{X\leq K_{\epsilon}\right\}\right)-1\geq 1-\epsilon.

By (20), we know that there exists K2>0K_{2}>0 such that

P(i=0m+1j=0m0TZ,s:i,j:𝒞α2γdsK2(T+1),T0)1ϵ.P\left(\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{0}^{T}\left\|Z_{-\infty,s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{2\gamma}\mathrm{d}s\leq K_{2}(T+1),\forall T\geq 0\right)\geq 1-\epsilon.

Using (26), the similar argument in the proof of Proposition 3.2 and the fact that s>1s>1, we have that for any γ>0\gamma>0,

i=0m+1j=0mZ0,s:i,j:𝒞αγi=0m+1j=0m(Z,0:i,j:𝒞α2γ(i+j)+Z,s:i,j:𝒞α2γ).\sum_{i=0}^{m+1}\sum_{j=0}^{m}\left\|Z_{0,s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma}\lesssim\sum_{i=0}^{m+1}\sum_{j=0}^{m}\left(\left\|Z_{-\infty,0}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{2\gamma(i+j)}+\left\|Z_{-\infty,s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{2\gamma}\right).

Combining with (20), we derive that there exists K3>0K_{3}>0 such that

P(i=0m+1j=0m1TZ0,s:i,j:𝒞αγdsK3(T+1),T1)1ϵ.P\left(\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{1}^{T}\left\|Z_{0,s}^{:i,j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma}\mathrm{d}s\leq K_{3}(T+1),\forall T\geq 1\right)\geq 1-\epsilon.

By (31), there exists K4K_{4} such that Z¯α,α,1K4\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1}\leq K_{4} almost surely. Then we complete the proof. After these preparations, we demonstrate the uniqueness of invariant probability measure by using Theorem 5.2 and a similar argument to the proof of [32, Theorem 1.1].

Theorem 5.6

Let μ>0\mu>0 satisfy (68), which implies 2+2μ(μ+1+μ2)>2m+12+2\mu(\mu+\sqrt{1+\mu^{2}})>2m+1, and α0[1/(1+μ(μ+1+μ2)),2/(2m+1)]\alpha_{0}\in\left[1/\left(1+\mu(\mu+\sqrt{1+\mu^{2}})\right),2/\left(2m+1\right)\right]. For every x𝒞α0x\in\mathcal{C}^{-\alpha_{0}}, there exists a unique probability measure ηx\eta_{x} on 𝒞α0\mathcal{C}^{-\alpha_{0}} such that Stηx=ηxS_{t}^{*}\eta_{x}=\eta_{x} for all t0t\geq 0.

{proof}

Let u(t;x)=Z0,t+v(t;u0)u(t;x)=Z_{0,t}+v\left(t;u_{0}\right) and u~(t;y)=Z0,t+v~(t;u1)\tilde{u}(t;y)=Z_{0,t}+\tilde{v}\left(t;u_{1}\right), where v(t;u0)v\left(t;u_{0}\right), v~(t;u1)\tilde{v}\left(t;u_{1}\right) is the unique mild solution to (33) and (57) with initial data x,y𝒞α0x,y\in\mathcal{C}^{-\alpha_{0}} respectively. We apply Theorem 5.2 to B=𝕏=𝒞α0B=\mathbb{X}=\mathcal{C}^{-\alpha_{0}} and 𝒫=P1(x,dy)\mathcal{P}=P_{1}(x,\mathrm{d}y), where P1(x,)P_{1}(x,\cdot) denotes the marginal distribution of u(t;x)u(t;x) at time t=1t=1.

Let wt:=u~(t;u1)u(t;u0)=v~(t;u1)v(t;u0)w_{t}:=\tilde{u}\left(t;u_{1}\right)-u\left(t;u_{0}\right)=\tilde{v}\left(t;u_{1}\right)-v\left(t;u_{0}\right) and W~(t)=W(t)0tτRλwsds\tilde{W}(t)=W(t)-\int_{0}^{t\wedge\tau_{R}}\lambda w_{s}\mathrm{d}s, where initial datas u0,u1𝒞α0u_{0},u_{1}\in\mathcal{C}^{-\alpha_{0}}, WW is a cylindrical Wiener process defined as (4) and

τR:=inf{t>0,0twsL22dsR}.\tau_{R}:=\inf\left\{t>0,\int_{0}^{t}\left\|w_{s}\right\|_{L^{2}}^{2}ds\geq R\right\}.

Note that

E[exp(120τRλwsL22𝑑s)]e12Rλ2,\mathrm{E}\left[\exp\left(\frac{1}{2}\int_{0}^{\tau_{R}}\|\lambda w_{s}\|_{L^{2}}^{2}ds\right)\right]\leq e^{\frac{1}{2}R\lambda^{2}},

then by Girsanov theorem (see [6, 12] or [20, Theorem 5.1]), there exists a probability measure QQ on (Ω,,(t)t0)\left(\Omega,\mathcal{F},\left(\mathcal{F}_{t}\right)_{t\geq 0}\right) such that under QQ, W~\tilde{W} is a cylindrical Wiener process. Moreover, it holds that PQP\sim Q on =σ(t0t)\mathcal{F}_{\infty}=\sigma\left(\cup_{t\geq 0}\mathcal{F}_{t}\right).

Let Z^\hat{Z} be the solution to the following linear equation

dZ^(t)=AZ^(t)dt+dW~(t),Z^(0)=0,d\hat{Z}(t)=A\hat{Z}(t)dt+d\tilde{W}(t),\quad\hat{Z}(0)=0,

and Z^0,t:k,l:\hat{Z}_{0,t}^{:k,l:} can be defined similarly as Z0,t:k,l:{Z}_{0,t}^{:k,l:} in Section 3. Since Z^0,t=Z0,t+at\hat{Z}_{0,t}=Z_{0,t}+a_{t}, where at=0te(ts)Aλws1{sτR}dsC([0,);𝒞α′′)a_{t}=-\int_{0}^{t}e^{(t-s)A}\lambda w_{s}\textbf{1}_{\left\{s\leq\tau_{R}\right\}}\mathrm{d}s\in C\left([0,\infty);\mathcal{C}^{\alpha^{\prime\prime}}\right) for some α′′>α\alpha^{\prime\prime}>\alpha, we have that under PP,

Z^0,t:k,l:=i=0kj=0l(ki)(lj)atiat¯jZ0,t:ki,lj:,\hat{Z}_{0,t}^{:k,l:}=\sum_{i=0}^{k}\sum_{j=0}^{l}\binom{k}{i}\binom{l}{j}a_{t}^{i}\overline{a_{t}}^{j}Z_{0,t}^{:k-i,l-j:}, (65)

almost surely. Thus

Z¯^={Z^0,,Z^0,:i,j:,i,jL},\underline{\hat{Z}}=\left\{\hat{Z}_{0,\cdot},\hat{Z}_{0,\cdot}^{:i,j:},i,j\in L\right\}\in\mathcal{L},

where sets LL and \mathcal{L} are defined in (28) and (30) respectively. Then by Theorem 4.8, there exists a unique mild solution v^(;u1)C((0,),𝒞β)\hat{v}(\cdot;u_{1})\in C\left((0,\infty),\mathcal{C}^{\beta}\right) to the following equation

{tv^=[(i+μ)Δ1]v^+Ψ(v^,Z¯^),t>0,x𝕋2,v^(0,)=u1𝒞α0.\begin{cases}\partial_{t}\hat{v}=\left[(\mathrm{i}+\mu)\Delta-1\right]\hat{v}+\Psi(\hat{v},\underline{\hat{Z}}),&t>0,x\in\mathbb{T}^{2},\\ \hat{v}(0,\cdot)=u_{1}\in\mathcal{C}^{-\alpha_{0}}.\end{cases}

Let u^(t;u1)=v^(t;u1)+Z^0,t\hat{u}(t;u_{1})=\hat{v}(t;u_{1})+\hat{Z}_{0,t}. Then we know that under QQ, u^(t;u1)\hat{u}(t;u_{1}) is also a mild solution to (35) with initial data u1u_{1} and ξ\xi replaced by ξλwt1{tτR}\xi-\lambda w_{t}1_{\left\{t\leq\tau_{R}\right\}}. Combining Theorem 4.8 with Yamada-Watanabe Theorem (see [23]), we have that under QQ, u^(t;u1)\hat{u}(t;u_{1}) has the same law as the solution u(t;u1)u(t;u_{1}) to (35) starting from u1u_{1} under PP. Since PQP\sim Q, we have that under PP, the marginal distributions of the pair (u(;u0),u^(;u1))\left(u\left(\cdot;u_{0}\right),\hat{u}\left(\cdot;u_{1}\right)\right) are equivalent to these of (u(;u0),u(;u1))\left(u\left(\cdot;u_{0}\right),u\left(\cdot;u_{1}\right)\right). For u0,u1𝒞α0u_{0},u_{1}\in\mathcal{C}^{-\alpha_{0}}, we set

Γu0,u1:= law of (u(;u0),u^(;u1)).\Gamma_{u_{0},u_{1}}:=\mbox{ law of }\left(u\left(\cdot;u_{0}\right),\hat{u}\left(\cdot;u_{1}\right)\right).

Then Γu0,u1𝒞~(δu0𝒫,δu1𝒫)\Gamma_{u_{0},u_{1}}\in\tilde{\mathcal{C}}\left(\delta_{u_{0}}\mathcal{P}_{\infty},\delta_{u_{1}}\mathcal{P}_{\infty}\right). It remains to prove that Γu0,u1(D)>0\Gamma_{u_{0},u_{1}}(D)>0.

Since u^(t;u1)=v^(t;u1)+Z^0,t\hat{u}(t;u_{1})=\hat{v}(t;u_{1})+\hat{Z}_{0,t}, we have that under PP, u^(t;u1)\hat{u}\left(t;u_{1}\right) satisfies the following equation with initial data u1𝒞α0u_{1}\in\mathcal{C}^{-\alpha_{0}},

tu^\displaystyle\partial_{t}\hat{u} =[(i+μ)Δ1]u^+Ψ(u^Z^0,t,Z¯^)+ξλw𝟏{tτR}\displaystyle=\left[(\mathrm{i}+\mu)\Delta-1\right]\hat{u}+\Psi(\hat{u}-\hat{Z}_{0,t},\underline{\hat{Z}})+\xi-\lambda w\mathbf{1}_{\left\{t\leq\tau_{R}\right\}}
=[(i+μ)Δ1]u^+Ψ(u^Z0,t,Z¯)+ξλw𝟏{tτR},t>0,x𝕋2.\displaystyle=\left[(\mathrm{i}+\mu)\Delta-1\right]\hat{u}+\Psi(\hat{u}-Z_{0,t},\underline{Z})+\xi-\lambda w\mathbf{1}_{\left\{t\leq\tau_{R}\right\}},\quad t>0,x\in\mathbb{T}^{2}.

where we use (65) to get that Ψ(u^Z^0,t,Z¯^)=Ψ(u^Z0,t,Z¯)\Psi(\hat{u}-\hat{Z}_{0,t},\underline{\hat{Z}})=\Psi(\hat{u}-Z_{0,t},\underline{Z}). Then on {τR=}\left\{\tau_{R}=\infty\right\}, u^(t;u1)Z0,t\hat{u}(t;u_{1})-Z_{0,t} also satisfies (57). By Theorem 5.3, we obtain that on {τR=}\left\{\tau_{R}=\infty\right\}, u^(t;u1)Z0,t=v~(t;u1)=u~(t;u1)Z0,t\hat{u}(t;u_{1})-Z_{0,t}=\tilde{v}(t;u_{1})=\tilde{u}(t;u_{1})-Z_{0,t}, which implies that u^(t;u1)=u~(t;u1)\hat{u}(t;u_{1})=\tilde{u}(t;u_{1}) on {τR=}\left\{\tau_{R}=\infty\right\}. To prove Γu0,u1(D)>0\Gamma_{u_{0},u_{1}}(D)>0, it is sufficient to estimate wt=u~(t;u1)u(t;u0)=v~(t;u1)v(t;u0)w_{t}=\tilde{u}\left(t;u_{1}\right)-u\left(t;u_{0}\right)=\tilde{v}\left(t;u_{1}\right)-v\left(t;u_{0}\right) on {τR=}\left\{\tau_{R}=\infty\right\}.

By the definition of wtw_{t}, we know that ww is the mild solution to the following equation

{tw=[(i+μ)Δ1]wλw+Ψ(v~,Z¯)Ψ(v,Z¯),t>0,x𝕋2,w(0,)=u0u1.\begin{cases}\partial_{t}w=\left[(\mathrm{i}+\mu)\Delta-1\right]w-\lambda w+\Psi\left(\tilde{v},\underline{Z}\right)-\Psi\left(v,\underline{Z}\right),&t>0,x\in\mathbb{T}^{2},\\ w(0,\cdot)=u_{0}-u_{1}.\end{cases}

Testing against w¯t\overline{w}_{t} and using a similar argument in Proposition 4.5, we obtain that for every 0δ<10\leq\delta<1 (see (40) for p=1p=1),

12twtL22+δμwtL22+(λ+1)wtL22Re[wt¯,Ψ(v~t,Z¯t)Ψ(vt,Z¯t)],\frac{1}{2}\partial_{t}\|w_{t}\|_{L^{2}}^{2}+\delta\mu\|\nabla w_{t}\|_{L^{2}}^{2}+(\lambda+1)\|w_{t}\|_{L^{2}}^{2}\leq\mathrm{Re}\left[\left\langle\overline{w_{t}},\Psi\left(\tilde{v}_{t},\underline{Z}_{t}\right)-\Psi\left(v_{t},\underline{Z}_{t}\right)\right\rangle\right],

where

Ψ(v~t,Z¯t)Ψ(vt,Z¯t)\displaystyle\Psi\left(\tilde{v}_{t},\underline{Z}_{t}\right)-\Psi\left(v_{t},\underline{Z}_{t}\right)
=\displaystyle= i=0m+1j=0m(m+1i)(mj)𝟏{(i,j)(0,0)}Z0,t:m+1i,mj:(v~tiv~t¯jvtivt¯j)+(τ+1)(v~tvt)\displaystyle\,\sum_{i=0}^{m+1}\sum_{j=0}^{m}\binom{m+1}{i}\binom{m}{j}\mathbf{1}_{\left\{(i,j)\neq(0,0)\right\}}Z_{0,t}^{:m+1-i,m-j:}\left(\tilde{v}_{t}^{i}\overline{\tilde{v}_{t}}^{j}-v_{t}^{i}\overline{v_{t}}^{j}\right)+(\tau+1)\left(\tilde{v}_{t}-v_{t}\right)
=\displaystyle= i+j>0(m+1i)(mj)Z0,t:m+1i,mj:[(v~tvt)vt¯jl=0i1vti1lv~tl+(v~t¯vt¯)v~tis=0j1vt¯j1sv~t¯s]\displaystyle\,\sum_{i+j>0}\binom{m+1}{i}\binom{m}{j}Z_{0,t}^{:m+1-i,m-j:}\left[\left(\tilde{v}_{t}-v_{t}\right)\overline{v_{t}}^{j}\sum_{l=0}^{i-1}v_{t}^{i-1-l}\tilde{v}_{t}^{l}+\left(\overline{\tilde{v}_{t}}-\overline{v_{t}}\right)\tilde{v}_{t}^{i}\sum_{s=0}^{j-1}\overline{v_{t}}^{j-1-s}\overline{\tilde{v}_{t}}^{s}\right]
+(τ+1)(v~tvt).\displaystyle+(\tau+1)\left(\tilde{v}_{t}-v_{t}\right).

Then

|Re[wt¯,Ψ(v~t,Z¯t)Ψ(vt,Z¯t)]||wt|,|Ψ(v~t,Z¯t)Ψ(vt,Z¯t)|\displaystyle\left|\mathrm{Re}\left[\left\langle\overline{w_{t}},\Psi\left(\tilde{v}_{t},\underline{Z}_{t}\right)-\Psi\left(v_{t},\underline{Z}_{t}\right)\right\rangle\right]\right|\leq\left\langle\left|w_{t}\right|,\left|\Psi\left(\tilde{v}_{t},\underline{Z}_{t}\right)-\Psi\left(v_{t},\underline{Z}_{t}\right)\right|\right\rangle
\displaystyle\leq Ci+j>0|wt|2,|Z0,t:m+1i,mj:|(l=0i1|vt|i+j1l|v~t|l+s=0j1|v~t|i+s|vt|j1s)\displaystyle\,C\sum_{i+j>0}\left\langle\left|w_{t}\right|^{2},\left|Z_{0,t}^{:m+1-i,m-j:}\right|\left(\sum_{l=0}^{i-1}\left|v_{t}\right|^{i+j-1-l}\left|\tilde{v}_{t}\right|^{l}+\sum_{s=0}^{j-1}\left|\tilde{v}_{t}\right|^{i+s}\left|v_{t}\right|^{j-1-s}\right)\right\rangle
+(|τ|+1)wtL22.\displaystyle+(|\tau|+1)\|w_{t}\|_{L^{2}}^{2}.

We now estimate each term |wt|2,|vt|k|v~t|i+j1k|Z0,t:m+1i,mj:|\left\langle\left|w_{t}\right|^{2},\left|v_{t}\right|^{k}\left|\tilde{v}_{t}\right|^{i+j-1-k}\left|Z_{0,t}^{:m+1-i,m-j:}\right|\right\rangle for (i,j,k)(i,j,k) belongs to the set

I:={(i,j,k)3:0im+1,0jm,(i,j)(0,0),0ki+j1}.I:=\left\{(i,j,k)\in\mathbb{N}^{3}:0\leq i\leq m+1,0\leq j\leq m,(i,j)\neq(0,0),0\leq k\leq i+j-1\right\}. (66)

By Proposition 2.9 and 2.10, we have that for α>0\alpha>0 sufficiently small and α<s<12\alpha<s<\frac{1}{2},

wt243,1αwt2H43swtH2swtL4wtH2122wtL2(wtL2+wtL2).\left\|w_{t}^{2}\right\|_{\mathcal{B}_{\frac{4}{3},1}^{\alpha}}\lesssim\left\|w_{t}^{2}\right\|_{H_{\frac{4}{3}}^{s}}\lesssim\left\|w_{t}\right\|_{H_{2}^{s}}\|w_{t}\|_{L^{4}}\lesssim\left\|w_{t}\right\|_{H_{2}^{\frac{1}{2}}}^{2}\lesssim\|w_{t}\|_{L^{2}}\left(\|\nabla w_{t}\|_{L^{2}}+\|w_{t}\|_{L^{2}}\right).

Then by Proposition 2.7 and Young’s inequality,

|wt|2,|vt|k|v~t|i+j1k|Z0,t:m+1i,mj:|\displaystyle\left\langle\left|w_{t}\right|^{2},\left|v_{t}\right|^{k}\left|\tilde{v}_{t}\right|^{i+j-1-k}\left|Z_{0,t}^{:m+1-i,m-j:}\right|\right\rangle
\displaystyle\leq Cwt243,1αvtkv~ti+j1kZ0,t:m+1i,mj:4,α\displaystyle\,C\left\|w_{t}^{2}\right\|_{\mathcal{B}_{\frac{4}{3},1}^{\alpha}}\left\|v_{t}^{k}\tilde{v}_{t}^{i+j-1-k}Z_{0,t}^{:m+1-i,m-j:}\right\|_{\mathcal{B}_{4,\infty}^{-\alpha}}
\displaystyle\leq CwtL2(wtL2+wtL2)vtkv~ti+j1kZ0,t:m+1i,mj:4,α\displaystyle\,C\|w_{t}\|_{L^{2}}\left(\|\nabla w_{t}\|_{L^{2}}+\|w_{t}\|_{L^{2}}\right)\left\|v_{t}^{k}\tilde{v}_{t}^{i+j-1-k}Z_{0,t}^{:m+1-i,m-j:}\right\|_{\mathcal{B}_{4,\infty}^{-\alpha}}
\displaystyle\leq CwtL22(1+vtkv~ti+j1kZ0,t:m+1i,mj:4,α2)+2δμ(m+1)(m+2)(2m+1)wtL22,\displaystyle\,C\|w_{t}\|_{L^{2}}^{2}\left(1+\left\|v_{t}^{k}\tilde{v}_{t}^{i+j-1-k}Z_{0,t}^{:m+1-i,m-j:}\right\|_{\mathcal{B}_{4,\infty}^{-\alpha}}^{2}\right)+\frac{2\delta\mu}{(m+1)(m+2)(2m+1)}\|\nabla w_{t}\|_{L^{2}}^{2},

where (m+1)(m+2)(2m+1)/2(m+1)(m+2)(2m+1)/2 is the cardinality of the set II (see (66)). Therefore,

12twtL22+λwtL22\displaystyle\frac{1}{2}\partial_{t}\|w_{t}\|_{L^{2}}^{2}+\lambda\|w_{t}\|_{L^{2}}^{2}
\displaystyle\leq CwtL22(1+(i,j,k)Ivtkv~ti+j1kZ0,t:m+1i,mj:4,α2):=wtL22Lt.\displaystyle\,C\|w_{t}\|_{L^{2}}^{2}\left(1+\sum_{(i,j,k)\in I}\left\|v_{t}^{k}\tilde{v}_{t}^{i+j-1-k}Z_{0,t}^{:m+1-i,m-j:}\right\|_{\mathcal{B}_{4,\infty}^{-\alpha}}^{2}\right):=\|w_{t}\|_{L^{2}}^{2}L_{t}.

By Gronwall’s inequality, for t1t\geq 1,

wtL22w1L22exp[1t2(λ+Ls)ds].\|w_{t}\|_{L^{2}}^{2}\leq\|w_{1}\|_{L^{2}}^{2}\exp\left[\int_{1}^{t}2(-\lambda+L_{s})\mathrm{d}s\right]. (67)

Recall that for every γ>0\gamma>0, there exists K>0K>0 such that P(EK,γ)>0P(E_{K,\gamma})>0, where EK,γE_{K,\gamma} is defined in (64). We estimate each term of 1tLsds\int_{1}^{t}L_{s}\mathrm{d}s on EK,γE_{K,\gamma} with γ>0\gamma>0 to be determined later. Note that for 1st1\leq s\leq t, by Proposition 2.1, 2.3, 2.9 and 2.10, for any β>α\beta^{\prime}>\alpha and any k1k\geq 1,

vsk4,β2\displaystyle\left\|v_{s}^{k}\right\|_{\mathcal{B}_{4,\infty}^{\beta^{\prime}}}^{2} vsk2,β+122vsk2,1β+122vskH2β02vskL22(1β0)vskH212β0\displaystyle\lesssim\left\|v_{s}^{k}\right\|_{\mathcal{B}_{2,\infty}^{\beta^{\prime}+\frac{1}{2}}}^{2}\lesssim\left\|v_{s}^{k}\right\|_{\mathcal{B}_{2,1}^{\beta^{\prime}+\frac{1}{2}}}^{2}\lesssim\left\|v_{s}^{k}\right\|_{H_{2}^{\beta_{0}}}^{2}\lesssim\left\|v_{s}^{k}\right\|_{L^{2}}^{2(1-\beta_{0})}\left\|v_{s}^{k}\right\|_{H^{1}_{2}}^{2\beta_{0}}
vskL22(1β0)(vskL22β0+vsk1vsL22β0)\displaystyle\lesssim\left\|v_{s}^{k}\right\|_{L^{2}}^{2(1-\beta_{0})}\left(\left\|v_{s}^{k}\right\|_{L^{2}}^{2\beta_{0}}+\left\|v_{s}^{k-1}\nabla v_{s}\right\|_{L^{2}}^{2\beta_{0}}\right)
vsL2k2k+vsL2k2k(1β0)vs2k2|vs|2L1β0,\displaystyle\lesssim\left\|v_{s}\right\|_{L^{2k}}^{2k}+\left\|v_{s}\right\|_{L^{2k}}^{2k(1-\beta_{0})}\left\|v_{s}^{2k-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}},

where β0=β+12+ϵ\beta_{0}=\beta^{\prime}+\frac{1}{2}+\epsilon and ϵ>0\epsilon>0 is arbitrary. This inequality is also valid for v~s\tilde{v}_{s}. Then for (i,j,k)I(i,j,k)\in I (see (66) for the definition of the set II), by Proposition 2.6,

1tvskv~si+j1kZ0,s:m+1i,mj:4,α2ds\displaystyle\int_{1}^{t}\left\|v_{s}^{k}\tilde{v}_{s}^{i+j-1-k}Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{B}_{4,\infty}^{-\alpha}}^{2}\mathrm{d}s
\displaystyle\lesssim 1tvsk4,β2v~si+j1k4,β2Z0,s:m+1i,mj:𝒞α2ds\displaystyle\,\int_{1}^{t}\left\|v_{s}^{k}\right\|_{\mathcal{B}_{4,\infty}^{\beta^{\prime}}}^{2}\left\|\tilde{v}_{s}^{i+j-1-k}\right\|_{\mathcal{B}_{4,\infty}^{\beta^{\prime}}}^{2}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2}\mathrm{d}s
\displaystyle\lesssim 1t(vsL2k2k+vsL2k2k(1β0)vs2k2|vs|2L1β0)Z0,s:m+1i,mj:𝒞α2\displaystyle\,\int_{1}^{t}\left(\left\|v_{s}\right\|_{L^{2k}}^{2k}+\left\|v_{s}\right\|_{L^{2k}}^{2k(1-\beta_{0})}\left\|v_{s}^{2k-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}}\right)\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2}
(v~sL2(i+j1k)2(i+j1k)+v~sL2(i+j1k)2(i+j1k)(1β0)v~s2(i+j1k)2|v~s|2L1β0)ds.\displaystyle\quad\quad\left(\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)}}^{2(i+j-1-k)}+\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)}}^{2(i+j-1-k)(1-\beta_{0})}\left\|\tilde{v}_{s}^{2(i+j-1-k)-2}\left|\nabla\tilde{v}_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}}\right)\mathrm{d}s.

Then we combine Hölder’s inequality, Corollary 4.7 and Proposition 5.4 to estimate each term in the right side of the inequality above.

We show the estimation of the following term as an example. Taking pi4>1p^{4}_{i}>1, 1i51\leq i\leq 5, which satisfy i=151pi4=1\sum_{i=1}^{5}\frac{1}{p^{4}_{i}}=1, (1β0)pj41(1-\beta_{0})p^{4}_{j}\geq 1 for j=1,3j=1,3 and β0pk41\beta_{0}p^{4}_{k}\leq 1 for k=2,4k=2,4, and by Hölder’s inequality, we get that

1tvsL2k2k(1β0)vs2k2|vs|2L1β0v~sL2(i+j1k)2(i+j1k)(1β0)\displaystyle\int_{1}^{t}\left\|v_{s}\right\|_{L^{2k}}^{2k(1-\beta_{0})}\left\|v_{s}^{2k-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)}}^{2(i+j-1-k)(1-\beta_{0})}
v~s2(i+j1k)2|v~s|2L1β0Z0,s:m+1i,mj:𝒞α2ds\displaystyle\quad\quad\left\|\tilde{v}_{s}^{2(i+j-1-k)-2}\left|\nabla\tilde{v}_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2}\mathrm{d}s
\displaystyle\lesssim (0tvsL2k2k(1β0)p14ds)1p14(vs2k2|vs|2L1β0p24ds)1p24\displaystyle\,\left(\int_{0}^{t}\left\|v_{s}\right\|_{L^{2k}}^{2k(1-\beta_{0})p^{4}_{1}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{1}}}\left(\left\|v_{s}^{2k-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}p^{4}_{2}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{2}}}
(0tv~sL2(i+j1k)2(i+j1k)(1β0)p34ds)1p34(0tv~s2(i+j1k)2|v~s|2L1β0p44ds)1p44\displaystyle\left(\int_{0}^{t}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)}}^{2(i+j-1-k)(1-\beta_{0})p^{4}_{3}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{3}}}\left(\int_{0}^{t}\left\|\tilde{v}_{s}^{2(i+j-1-k)-2}\left|\nabla\tilde{v}_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}p^{4}_{4}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{4}}}
(0tZ0,s:m+1i,mj:𝒞α2p54ds)1p54\displaystyle\left(\int_{0}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2p^{4}_{5}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{5}}}
\displaystyle\lesssim (0tvsL2k(1β0)p142k(1β0)p14ds)1p14(vs2k2|vs|2L1ds+t)1p24\displaystyle\,\left(\int_{0}^{t}\left\|v_{s}\right\|_{L^{2k(1-\beta_{0})p^{4}_{1}}}^{2k(1-\beta_{0})p^{4}_{1}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{1}}}\left(\left\|v_{s}^{2k-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}\mathrm{d}s+t\right)^{\frac{1}{p^{4}_{2}}}
(0tv~sL2(i+j1k)(1β0)p342(i+j1k)(1β0)p34ds)1p34(0tv~s2(i+j1k)2|v~s|2L1ds+t)1p44\displaystyle\left(\int_{0}^{t}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)(1-\beta_{0})p^{4}_{3}}}^{2(i+j-1-k)(1-\beta_{0})p^{4}_{3}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{3}}}\left(\int_{0}^{t}\left\|\tilde{v}_{s}^{2(i+j-1-k)-2}\left|\nabla\tilde{v}_{s}\right|^{2}\right\|_{L^{1}}\mathrm{d}s+t\right)^{\frac{1}{p^{4}_{4}}}
(0tZ0,s:m+1i,mj:𝒞α2p54ds)1p54.\displaystyle\left(\int_{0}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2p^{4}_{5}}\mathrm{d}s\right)^{\frac{1}{p^{4}_{5}}}.

where we use Young’s inequality in the last inequality. Similarly, taking the constants pi1>1p^{1}_{i}>1 for 1i31\leq i\leq 3, pi2>1p^{2}_{i}>1 for 1i41\leq i\leq 4 and pi3>1p^{3}_{i}>1 for 1i41\leq i\leq 4, which satisfy 1=i=131pi1=i=141pi2=i=141pi31=\sum_{i=1}^{3}\frac{1}{p^{1}_{i}}=\sum_{i=1}^{4}\frac{1}{p^{2}_{i}}=\sum_{i=1}^{4}\frac{1}{p^{3}_{i}}, (1β0)pj21(1-\beta_{0})p^{2}_{j}\geq 1 for j=2,3j=2,3, (1β0)p131(1-\beta_{0})p^{3}_{1}\geq 1 and β0p231\beta_{0}p^{3}_{2}\leq 1, we have that

1tvsL2k2kv~sL2(i+j1k)2(i+j1k)Z0,s:m+1i,mj:𝒞α2ds\displaystyle\int_{1}^{t}\left\|v_{s}\right\|_{L^{2k}}^{2k}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)}}^{2(i+j-1-k)}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2}\mathrm{d}s
\displaystyle\lesssim (0tvsL2kp112kp11ds)1p11(0tv~sL2(i+j1k)p212(i+j1k)p21ds)1p21(0tZ0,s:m+1i,mj:𝒞α2p31ds)1p31,\displaystyle\,\left(\int_{0}^{t}\left\|v_{s}\right\|_{L^{2kp^{1}_{1}}}^{2kp^{1}_{1}}\mathrm{d}s\right)^{\frac{1}{p^{1}_{1}}}\left(\int_{0}^{t}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)p^{1}_{2}}}^{2(i+j-1-k)p^{1}_{2}}\mathrm{d}s\right)^{\frac{1}{p^{1}_{2}}}\left(\int_{0}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2p^{1}_{3}}\mathrm{d}s\right)^{\frac{1}{p^{1}_{3}}},

and

1tvsL2k2kv~sL2(i+j1k)2(i+j1k)(1β0)v~s2(i+j1k)2|v~s|2L1β0Z0,s:m+1i,mj:𝒞α2ds\displaystyle\int_{1}^{t}\left\|v_{s}\right\|_{L^{2k}}^{2k}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)}}^{2(i+j-1-k)(1-\beta_{0})}\left\|\tilde{v}_{s}^{2(i+j-1-k)-2}\left|\nabla\tilde{v}_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2}\mathrm{d}s
\displaystyle\lesssim (0tvsL2kp122kp12ds)1p12(0tv~sL2(i+j1k)(1β0)p222(i+j1k)(1β0)p22ds)1p22\displaystyle\,\left(\int_{0}^{t}\left\|v_{s}\right\|_{L^{2kp^{2}_{1}}}^{2kp^{2}_{1}}\mathrm{d}s\right)^{\frac{1}{p^{2}_{1}}}\left(\int_{0}^{t}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)(1-\beta_{0})p^{2}_{2}}}^{2(i+j-1-k)(1-\beta_{0})p^{2}_{2}}\mathrm{d}s\right)^{\frac{1}{p^{2}_{2}}}
(0tv~s2(i+j1k)2|v~s|2L1ds+t)1p32(0tZ0,s:m+1i,mj:𝒞α2p42ds)1p42,\displaystyle\left(\int_{0}^{t}\left\|\tilde{v}_{s}^{2(i+j-1-k)-2}\left|\nabla\tilde{v}_{s}\right|^{2}\right\|_{L^{1}}\mathrm{d}s+t\right)^{\frac{1}{p^{2}_{3}}}\left(\int_{0}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2p^{2}_{4}}\mathrm{d}s\right)^{\frac{1}{p^{2}_{4}}},

and

1tvsL2k2k(1β0)vs2k2|vs|2L1β0v~sL2(i+j1k)2(i+j1k)Z0,s:m+1i,mj:𝒞α2ds\displaystyle\int_{1}^{t}\left\|v_{s}\right\|_{L^{2k}}^{2k(1-\beta_{0})}\left\|v_{s}^{2k-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}^{\beta_{0}}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)}}^{2(i+j-1-k)}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2}\mathrm{d}s
\displaystyle\lesssim (0tvsL2k(1β0)p132k(1β0)p13ds)1p13(0tvs2k2|vs|2L1ds+t)1p23\displaystyle\,\left(\int_{0}^{t}\left\|v_{s}\right\|_{L^{2k(1-\beta_{0})p^{3}_{1}}}^{2k(1-\beta_{0})p^{3}_{1}}\mathrm{d}s\right)^{\frac{1}{p^{3}_{1}}}\left(\int_{0}^{t}\left\|v_{s}^{2k-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}\mathrm{d}s+t\right)^{\frac{1}{p^{3}_{2}}}
(0tv~sL2(i+j1k)p332(i+j1k)p33ds)1p33(0tZ0,s:m+1i,mj:𝒞α2p43ds)1p43.\displaystyle\left(\int_{0}^{t}\left\|\tilde{v}_{s}\right\|_{L^{2(i+j-1-k)p^{3}_{3}}}^{2(i+j-1-k)p^{3}_{3}}\mathrm{d}s\right)^{\frac{1}{p^{3}_{3}}}\left(\int_{0}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{2p^{3}_{4}}\mathrm{d}s\right)^{\frac{1}{p^{3}_{4}}}.

Let μ>0\mu>0 be such that

1+μ(μ+1+μ2)\displaystyle 1+\mu\left(\mu+\sqrt{1+\mu^{2}}\right) (68)
\displaystyle\geq 2mmax{p11,p21,p12,(1β0)p22,(1β0)p13,p33,(1β0)p14,(1β0)p34}.\displaystyle 2m\max\left\{p^{1}_{1},p^{1}_{2},p^{2}_{1},(1-\beta_{0})p^{2}_{2},(1-\beta_{0})p^{3}_{1},p^{3}_{3},(1-\beta_{0})p^{4}_{1},(1-\beta_{0})p^{4}_{3}\right\}. (69)

This also implies that 2+2μ(μ+1+μ2)>2m+12+2\mu(\mu+\sqrt{1+\mu^{2}})>2m+1. Then we can use Corollary 4.7 and Proposition 5.4 to control terms like

0tvsL2q2q,0tv~sL2q2qds,0tvs2q2|vs|2L1ds,0tv~s2q2|v~s|2L1ds,\int_{0}^{t}\left\|v_{s}\right\|^{2q}_{L^{2q}},\quad\int_{0}^{t}\left\|\tilde{v}_{s}\right\|^{2q}_{L^{2q}}\mathrm{d}s,\quad\int_{0}^{t}\left\|v_{s}^{2q-2}\left|\nabla v_{s}\right|^{2}\right\|_{L^{1}}\mathrm{d}s,\quad\int_{0}^{t}\left\|\tilde{v}_{s}^{2q-2}\left|\nabla\tilde{v}_{s}\right|^{2}\right\|_{L^{1}}\mathrm{d}s,

for some q[1,1+μ(μ+1+μ2)]q\in\left[1,1+\mu\left(\mu+\sqrt{1+\mu^{2}}\right)\right]. Finally, we obtain that there exists a large γ\gamma such that on EK,γE_{K,\gamma},

1tLsds\displaystyle\int_{1}^{t}L_{s}\mathrm{d}s C(λ,Z¯α,α,1)+Ct+Ci=0m+1j=0m1tZ0,s:m+1i,mj:𝒞αγds\displaystyle\leq C(\lambda,\left\|\underline{Z}\right\|_{\alpha,\alpha^{\prime},1})+Ct+C\sum_{i=0}^{m+1}\sum_{j=0}^{m}\int_{1}^{t}\left\|Z_{0,s}^{:m+1-i,m-j:}\right\|_{\mathcal{C}^{-\alpha}}^{\gamma}\mathrm{d}s
C(K,λ)(1+t).\displaystyle\leq C(K,\lambda)(1+t).

Combining with (67), we obtain that on EK,γE_{K,\gamma}, we can take λ\lambda large enough such that there exist constants C1,C2>0C_{1},C_{2}>0 such that

wtL22C1eC2t0,t0.\|w_{t}\|_{L^{2}}^{2}\leq C_{1}e^{-C_{2}t}\rightarrow 0,\quad t\rightarrow 0.

This implies that for fixed K>0K>0, there exists R>0R>0 such that τR=\tau_{R}=\infty on EK,γE_{K,\gamma}.

Moreover, under the assumptions that 2m+1<2+2μ(μ+1+μ2)2m+1<2+2\mu(\mu+\sqrt{1+\mu^{2}}) and α0[1/(1+μ(μ+1+μ2)),2/(2m+1)]\alpha_{0}\in\left[1/\left(1+\mu(\mu+\sqrt{1+\mu^{2}})\right),2/\left(2m+1\right)\right], we can find 1p1+12μ(μ+1+μ2)1\leq p\leq 1+\frac{1}{2}\mu(\mu+\sqrt{1+\mu^{2}}) such that L2p𝒞α0L^{2p}\hookrightarrow\mathcal{C}^{-\alpha_{0}}. Thus on EK,γE_{K,\gamma}, by Hölder’s inequality, Corollary 4.7 and Proposition 5.4, we know that

wt𝒞α0wtL2pwtL2(vtL2(2p1)2p1+v~tL2(2p1)2p1)eCt(1+t)0,t.\left\|w_{t}\right\|_{\mathcal{C}^{-\alpha_{0}}}\lesssim\left\|w_{t}\right\|_{L^{2p}}\lesssim\left\|w_{t}\right\|_{L^{2}}\left(\left\|v_{t}\right\|_{L^{2(2p-1)}}^{2p-1}+\left\|\tilde{v}_{t}\right\|_{L^{2(2p-1)}}^{2p-1}\right)\lesssim e^{-Ct}(1+t)\rightarrow 0,\quad t\rightarrow\infty.

Therefore, we obtain that

Γu0,u1(D)P(EK,γ)>0.\Gamma_{u_{0},u_{1}}(D)\geq P(E_{K,\gamma})>0.

By [17, Corollary 2.2] (or Theorem 5.2), we get the uniqueness of the invariant probability measure.

References

  • [1] I. S. Aranson and L. Kramer. The world of the complex Ginzburg-Landau equation. Rev. Modern Phys., 74(1):99–143, 2002. MR1895097.
  • [2] H. Bahouri, J. Y. Chemin, and R. Danchin. Fourier analysis and nonlinear partial differential equations, volume 343 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer, Heidelberg, 2011. MR2768550.
  • [3] M. Barton-Smith. Global solution for a stochastic Ginzburg-Landau equation with multiplicative noise. Stochastic Anal. Appl., 22(1):1–18, 2004. MR2033986.
  • [4] M. Barton-Smith. Invariant measure for the stochastic Ginzburg Landau equation. NoDEA Nonlinear Differential Equations Appl., 11(1):29–52, 2004. MR2035365.
  • [5] J. Bergh and J. Löfström. Interpolation spaces. An introduction, volume 223 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin-New York, 1976. MR0482275.
  • [6] R. H. Cameron and W. T. Martin. Transformations of Wiener integrals under translations. Ann. of Math. (2), 45:386–396, 1944. MR0010346.
  • [7] H. Chen, Y. Chen, and Y. Liu. Berry-esséen bound for complex wiener-itô integral, 2024. Preprint at https://arxiv.org/pdf/2407.05353v1.
  • [8] H. Chen, Y. Chen, and Y. Liu. Kernel representation formula: From complex to real Wiener–Itô integrals and vice versa. Stochastic Process. Appl., 167:104241, 2024. MR4661207.
  • [9] Y. Chen and Y. Liu. On the fourth moment theorem for complex multiple Wiener-Itô integrals. Infin. Dimens. Anal. Quantum Probab. Relat. Top., 20(1):1750005, 2017. MR3623878.
  • [10] G. Da Prato and A. Debussche. Strong solutions to the stochastic quantization equations. Ann. Probab., 31(4):1900–1916, 2003. MR2016604.
  • [11] G. Da Prato and J. Zabczyk. Ergodicity for infinite-dimensional systems, volume 229 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1996. MR1417491.
  • [12] I. V. Girsanov. On transforming a class of stochastic processes by absolutely continuous substitution of measures. Teor. Verojatnost. i Primenen., 5:314–330, 1960. MR0133152.
  • [13] L. Grafakos. Classical Fourier analysis, volume 249 of Graduate Texts in Mathematics. Springer, New York, second edition, 2008. MR2445437.
  • [14] M. Gubinelli and N. Perkowski. Lectures on singular stochastic PDEs, volume 29 of Ensaios Matemáticos [Mathematical Surveys]. Sociedade Brasileira de Matemática, Rio de Janeiro, 2015. MR3445609.
  • [15] M. Hairer. Exponential mixing properties of stochastic PDEs through asymptotic coupling. Probab. Theory Related Fields, 124(3):345–380, 2002. MR1939651.
  • [16] M. Hairer. A theory of regularity structures. Invent. Math., 198(2):269–504, 2014. MR3274562.
  • [17] M. Hairer, J. C. Mattingly, and M. Scheutzow. Asymptotic coupling and a general form of Harris’ theorem with applications to stochastic delay equations. Probab. Theory Related Fields, 149(1-2):223–259, 2011. MR2773030.
  • [18] M. Hoshino. Global well-posedness of complex Ginzburg-Landau equation with a space-time white noise. Ann. Inst. Henri Poincaré Probab. Stat., 54(4):1969–2001, 2018. MR3865664.
  • [19] M. Hoshino, Y. Inahama, and N. Naganuma. Stochastic complex Ginzburg-Landau equation with space-time white noise. Electron. J. Probab., 22(104):1–68, 2017. MR3742401.
  • [20] I. Karatzas and S. E. Shreve. Brownian motion and stochastic calculus, volume 113 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1991. MR1121940.
  • [21] S. Kuksin and A. Shirikyan. Randomly forced CGL equation: stationary measures and the inviscid limit. J. Phys. A, 37(12):3805–3822, 2004. MR2039838.
  • [22] A. Kulik and M. Scheutzow. Generalized couplings and convergence of transition probabilities. Probab. Theory Related Fields, 171(1-2):333–376, 2018. MR3800835.
  • [23] T. G. Kurtz. The Yamada-Watanabe-Engelbert theorem for general stochastic equations and inequalities. Electron. J. Probab., 12:951–965, 2007. MR2336594.
  • [24] T. Matsuda. Characterization of the support for wick powers of the additive stochastic heat equation, 2020. Preprint at https://arxiv.org/abs/2001.11705.
  • [25] T. Matsuda. Global well-posedness of the two-dimensional stochastic complex Ginzburg-Landau equation with cubic nonlinearity, 2020. Preprint at https://arxiv.org/abs/2003.01569.
  • [26] J. C. Mourrat and H. Weber. Global well-posedness of the dynamic Φ4\Phi^{4} model in the plane. Ann. Probab., 45(4):2398–2476, 2017. MR3693966.
  • [27] J. C. Mourrat, H. Weber, and W. Xu. Construction of Φ34\Phi^{4}_{3} diagrams for pedestrians. In From particle systems to partial differential equations, volume 209 of Springer Proc. Math. Stat., pages 1–46. Springer, Cham, 2017. MR3746744.
  • [28] C. Odasso. Ergodicity for the stochastic complex Ginzburg-Landau equations. Ann. Inst. H. Poincaré Probab. Statist., 42(4):417–454, 2006. MR2242955.
  • [29] G. Parisi and Y. S. Wu. Perturbation theory without gauge fixing. Sci. Sinica, 24(4):483–496, 1981. MR0626795.
  • [30] S. G. Resnick. Dynamical problems in non-linear advective partial differential equations. ProQuest LLC, Ann Arbor, MI, 1995. Thesis (Ph.D.)–The University of Chicago. MR2716577.
  • [31] M. Röckner, R. Zhu, and X. Zhu. Sub and supercritical stochastic quasi-geostrophic equation. Ann. Probab., 43(3):1202–1273, 2015. MR3342662.
  • [32] M. Röckner, R. Zhu, and X. Zhu. Ergodicity for the stochastic quantization problems on the 2D-torus. Comm. Math. Phys., 352(3):1061–1090, 2017. MR3631399.
  • [33] Y. Sawano. Theory of Besov spaces, volume 56 of Developments in Mathematics. Springer, Singapore, 2018. MR3839617.
  • [34] W. J. Trenberth. Global well-posedness for the two-dimensional stochastic complex Ginzburg-Landau equation, 2020. Preprint at https://arxiv.org/abs/1911.09246.
  • [35] P. Tsatsoulis and H. Weber. Spectral gap for the stochastic quantization equation on the 2-dimensional torus. Ann. Inst. Henri Poincaré Probab. Stat., 54(3):1204–1249, 2018. MR3825880.