1 Introduction

In this work we study a damped/driven nonlinear Schrödinger equation

$$\begin{aligned} u_t-\nu \Delta u+i|u|^2u=\sqrt{\nu }\eta (t,x), \quad x\in {\mathbb {R}}^{n}, \end{aligned}$$
(1.1)

i.e. a CGL equation without linear dispersion, with cubic Hamiltonian nonlinearity and a random forcing. The dimension n is any, \(0<\nu \le 1\) is the viscosity constant and the random force \(\eta \) is white in time t and regular in x. The equation is considered under the odd periodic boundary conditions,

$$\begin{aligned} u(t,\dots ,x_j,\dots )=u(t,\dots ,x_j+2\pi ,\dots )=-u(t,\dots ,x_j+\pi ,\dots ),\quad j=1,\dots ,n. \end{aligned}$$

The latter implies that u vanishes on the boundary of the cube of half-periods \(K^{n}= [0, \pi ]^n\),

$$\begin{aligned} u\mid _{\partial K^{n}} =0. \end{aligned}$$

We denote by \(\{\varphi _d(\cdot ),\; d=(d_1,\dots ,d_n)\in {\mathbb {N}}^n\}\) the trigonometric basis in the space of odd periodic functions,

$$\begin{aligned} \varphi _d(x)=(\tfrac{2}{\pi })^{\frac{n}{2}}\sin (d_1x_1)\cdots \sin (d_nx_n). \end{aligned}$$

The basis is orthonormal with respect to the normalised scalar product \(\langle \!\langle \cdot ,\cdot \rangle \!\rangle \) in \(L_2(K^{n}, \pi ^{-n} dx)\),

$$\begin{aligned} \langle \!\langle u,v\rangle \!\rangle = \int _{K^n} \langle u(x), v(x)\rangle \pi ^{-n} dx, \end{aligned}$$
(1.2)

where \(\langle \cdot , \cdot \rangle \) is the real scalar product in \({\mathbb {C}}\), \(\langle u, v \rangle =\mathfrak {R}u{\bar{v}}\). It is formed by eigenfunctions of the Laplacian:

$$\begin{aligned} (-\Delta )\varphi _d=|d|^2\varphi _d. \end{aligned}$$

The force \(\eta (t,x)\) is a random field of the form

$$\begin{aligned} \eta (t,x)=\frac{\partial }{\partial t}\xi (t,x),\quad \xi (t,x)=\sum _{d\in {\mathbb {N}}^n}b_d\beta _d(t)\varphi _d(x). \end{aligned}$$
(1.3)

Here \(\beta _d(t)=\beta ^R_d(t)+i\beta ^I_d(t)\), where \(\beta _d^R(t)\), \(\beta _d^I(t)\) are independent real-valued standard Brownian motions, defined on a complete probability space \((\Omega ,{\mathcal {F}},{\mathbb {P}})\) with a filtration \(\{{\mathcal {F}}_t;t\geqslant 0\}\). The set of real numbers \(\{b_d,\;d\in {\mathbb {N}}^n\}\) is assumed to form a non-zero sequence, satisfying

$$\begin{aligned} 0<B_{m_*}<\infty , \quad m_*= \min \{ m\in {\mathbb {Z}}: m>n/2\}, \end{aligned}$$
(1.4)

where for a real number k we set

$$\begin{aligned} B_k:=\sum _{d\in {\mathbb {N}}^n}|d|^{2k} |b_d|^2\le \infty . \end{aligned}$$

For \(m\ge 0\) we denote by \(H^m\) the Sobolev space of order m, formed by complex odd periodic functions, equipped with the homogeneous norm,

$$\begin{aligned}\Vert u\Vert _m=\Vert (-\Delta )^{\frac{m}{2}}u\Vert _0,\end{aligned}$$

where \(\Vert \cdot \Vert _0\) is the \(L^2\)-norm on \(K^{n}\), \( \Vert u\Vert ^2_0 = \langle \!\langle u,u\rangle \!\rangle \) (see (1.2)). If we write \(u\in H^m\) as Fourier series, \(u(x)=\sum _{d\in {\mathbb {N}}^n}u_d\varphi _d(x),\) then \( \Vert u\Vert _m^2=\sum _{d\in {\mathbb {N}}^n}|d|^{2m}| u_d|^2. \)

Equation (1.1) with small \(\nu \) belongs to a group of equations, describing turbulence in the CGL equations. These equations have got quite a lot of attention in physical literature as models for turbulence in various media, e.g. see [3, Chapter 5]. In particular – as a natural model for hydrodynamical turbulence since Eq. (1.1) is obtained from the Navier-Stokes system by replacing the Euler term \((u\cdot \nabla ) u \), which is a quadratic Hamiltonian nonlinearity, by \(i|u|^2u\), which is a cubic Hamiltonian nonlinearity, see [13].

The global solvability of Eq. (1.1) for any space dimension n is established in [8, 10]. It is proved there that if

$$\begin{aligned} u(0,x)=u_0(x), \end{aligned}$$
(1.5)

where \(u_0\in H^m\cap C(K^n)\), \(m\in {\mathbb {N}}\), and if \(B_m<\infty \), then the problem (1.1), (1.5) has a unique strong solution u(tx) in \(H^m\) which we write as \(u(t,x;u_0)\), or \(u(t;u_0)\), or \(u_\nu (t;u_0)\). Its norm satisfies

$$\begin{aligned} {\mathbb {E}}\Vert u(t;u_0)\Vert _m^2 \le C_m \nu ^{-m},\quad t\ge 0, \end{aligned}$$

where \(C_m\) depends on \(\Vert u_0\Vert _m, |u_0|_\infty \) and \(B_m, B_{m_*}\). Furthermore, denoting by \(C_0(K^{n})\) the space of continuous complex functions on \(K^{n}\), vanishing at \(\partial K^{n}\), we have that the solutions u(tx) define a Markov process in \(C_0(K^{n})\). Moreover, if the noise \(\eta (t,\cdot )\) is non-degenerate in the sense that in (1.3) all coefficients \(b_d\) are non-zero, then this process is mixing.Footnote 1

Our goal is to study the growth of higher Sobolev norms for solutions of Eq. (1.1) as \(\nu \rightarrow 0\) on time intervals of order \({\mathcal {O}}(\frac{1}{\nu })\). The main result of this work is the following.

Theorem 1

For any real number \(m>2\), in addition to (1.4), assume that \(B_m<\infty \). Then there exists \(\kappa _{n,m}>0\) such that for every fixed quadruple \((\delta ,\kappa ,{\mathscr {K}}, T_0)\), where

$$\begin{aligned} \kappa \in (0,\kappa _{n,m}),\quad \delta \in (0,\tfrac{1}{8}), \quad {\mathscr {K}}, T_0>0, \end{aligned}$$

there exists a \(\nu _0>0\) with the property that if \(0<\nu \le \nu _0\), then for every \(u_0\in H^m \cap C_0(K^{n})\), satisfying

$$\begin{aligned} |u_0|_\infty \leqslant {\mathscr {K}}, \quad \Vert u_0\Vert _m \le \nu ^{-\kappa m}, \end{aligned}$$
(1.6)

the solution \(u(t,x;u_0)\) is such that

  1. (1)
    $$\begin{aligned} {\mathbb {P}}\big \{ \sup _{t \in [t_0, t_0+ T_0\nu ^{-1}]} \Vert u_\nu ^\omega (t)\Vert _m> \nu ^{-m\kappa }\big \} \ge 1-\delta , \quad \forall \, t_0\ge 0. \end{aligned}$$
  2. (2)

    If m is an integer, \(m\ge 3\), then a possible choice of \(\kappa _{n,m}\) is \(\kappa _{n,m}= \tfrac{1}{35}\), and there exists \(C\ge 1\), depending on \(\kappa < \tfrac{1}{35}\), \({\mathscr {K}}, m,B_{m_*}\) and \(B_m\), such that

    $$\begin{aligned} C^{-1} \nu ^{-2m\kappa +1}\leqslant {\mathbb {E}}\left( \nu \int _{t_0}^{t_0+\nu ^{-1}}\Vert u_\nu (s)\Vert _m^2ds\right) \leqslant C \nu ^{-m}, \quad \forall \, t_0\ge 0. \end{aligned}$$
    (1.7)

A similar result holds for the classical \(C^k\)-norms of solutions:

Proposition 2

For any integer \(m \ge 2\) in addition to (1.4) assume that \(B_m<\infty \). Then for every fixed triplet \(K, {\mathcal {K}}, T_0>0\) and any \(0<\kappa <1/16\) we have

$$\begin{aligned} {\mathbb {P}}\left\{ \sup _{t \in [t_0, t_0+ T_0\nu ^{-1}]} |u_\nu ^\omega (t;u_0)|_{C^m} > K \nu ^{-m\kappa }\right\} \rightarrow 1 \quad \text {as} \quad \nu \rightarrow 0, \end{aligned}$$
(1.8)

for each \(t_0\ge 0\), if \(u_0\) satisfies \(|u_0|_\infty \le {\mathcal {K}}\), \(|u_0|_{C^m} \le \nu ^{-\kappa m}\). The rate of convergence depends only on the triplet and \(\kappa \).

For a proof of this result see the extended version of our work [6]. Due to (1.8), for any \(m>2+n/2\) we have

$$\begin{aligned} {\mathbb {P}}\left\{ \sup _{T_0 \le t \le t_0+{T_0}{\nu ^{-1}}} \Vert u(t)\Vert _{m} \ge K \nu ^{- \lfloor m-\frac{n}{2} \rfloor \kappa } \right\} \rightarrow 1\quad \text {as} \quad \nu \rightarrow 0, \end{aligned}$$

for every \(K>0\) and \(0<\kappa <1/16\), where for \(a\in {\mathbb {R}}\) we denote \( \lfloor a \rfloor = \max \{ n\in {\mathbb {Z}}: n<a\}. \) This improves the first assertion of Theorem 1 for large m.

We have the following two corollaries from Theorem 1, valid if the Markov process defined by the Eq. (1.1) is mixing:

Corollary 3

Assume that \(B_m<\infty \) for all m and \(b_d\ne 0\) for all d. Then Eq. (1.1) is mixing and for any \(\kappa <1/35\) and \(0<\nu \le \nu _0\) its unique stationary measure \(\mu _\nu \) satisfies

$$\begin{aligned} C^{-1} \nu ^{-2m\kappa +1} \le \int \Vert u\Vert _m^2\mu _\nu (du) \le C\nu ^{-m}, \quad 3\le m\in {\mathbb {N}}. \end{aligned}$$
(1.9)

Here C and \(\nu _0\) are as in Theorem 1.

Corollary 4

Under the assumptions of Corollary 3, for any \(u_0\in C^\infty \) we have

$$\begin{aligned} \tfrac{1}{2} C^{-1} \nu ^{-2m\kappa +1} \le {\mathbb {E}}\Vert u (s;u_0)\Vert _m^2 \le 2C\nu ^{-m}, \quad 3\le m\in {\mathbb {N}}, \end{aligned}$$

if \(s\ge T(\nu , u_0, \kappa , B_m, B_{m_*})\), where C is the same as in (1.9).

Theorem 1 rigorously establishes the energy cascade to high frequencies for solutions of Eq. (1.1) with small \(\nu \). Indeed, if \(u_0(x)\) and \(\eta (t,x)\) are smooth functions of x (or even trigonometric polynomials of x), then in view of (1.7) for \(0<\nu \ll 1\) and \(t > rsim \nu ^{-1}\) a substantial part of the energy \( \frac{1}{2} \sum |u_d(t)|^2 \) of a solution \(u(t,x;u_0)\) is carried by high modes \(u_d\) with \(|d|\gg 1\). Relation (1.7) (valid for all integer \(m\ge 3\)) also means that the averaged in time space-scale \(l_x\) of solutions for (1.1) satisfies \( l_x \in [\nu ^{1/2}, \nu ^{1/35}], \) and goes to zero with \(\nu \) (see [1, 9]). We recall that the energy cascade to high frequencies and formation of short space-scale is the driving force of the Kolmogorov theory of turbulence, see [5].

We mention that in the work [12] the stochastic CGL equation

$$\begin{aligned} u_t-(\nu +i)\Delta u+i|u|^2u=\sqrt{\nu }\eta (t,x),\quad 0<\nu \le 1, \end{aligned}$$
(1.10)

with linear dispersion and white in time random force \(\eta \) as in (1.3) was considered under the odd periodic boundary conditions, and the inviscid limiting dynamics as \(\nu \rightarrow 0\) was examined. However, since the limiting Eq. (1.10)\({}_{\nu =0}\) is a regular PDE in difference with the Eq. (1.1)\({}_{\nu =0}\), the results on the inviscid limit in [12] differ in spirit from those in our work, and we do not discuss them now.

Deterministic versions of the result of Theorem 1 for Eq. (1.1) with \(\eta =0\), where \(\nu \) is a small non-zero complex number such that \(\mathfrak {R}\nu \ge 0\) and \(\mathfrak {I}\nu \le 0\) are known, see [9]. In particular, if \(\nu \) is a positive real number and \(u_0\) is a smooth function of order one, then for any integer \(m\ge 4\) a solution \(u_\nu (t,x;u_0)\) satisfies estimates (1.7) with the averaging \( \nu {\mathbb {E}}\int _t^{t+\nu ^{-1}}\! \!\dots ds \) replaced by \( \nu ^{1/3} \int _0^{\nu ^{-1/3}}\! \!\dots ds, \) with the same upper bound and with the lower bound \(C_m \nu ^{-\kappa _m m}\), where \(\kappa _m\rightarrow 1/3\) as \(m\rightarrow \infty \). Moreover, it was then shown in [2] that the lower bounds remain true with \(\kappa =1/3\), and that the estimates \( \sup _{t\in [0, |\nu |^{-1/3}]} \Vert u(t)\Vert _{C^m} \ge C_m |\nu |^{-m/3}, m\ge 2, \) hold for smooth solutions of Eq. (1.1) with \(\eta =0\) and any non-zero complex “viscosity” \(\nu \).

The better quality of the lower bounds for solutions of the deterministic equations is due to an extra difficulty which occurs in the stochastic case: when time grows, simultaneously with increasing of high Sobolev norms of a solution, its \(L_2\)-norm may decrease, which accordingly would weaken the mechanism, responding for the energy transfer to high modes. Significant part of the proof of Theorem 1 is devoted to demonstration that the \(L_2\)-norm of a solution cannot go down without sending up the second Sobolev norm.

If \(\eta =0\) and \(\nu =i\delta \in i{\mathbb {R}}\), then (1.1) is a Hamiltonian PDE (the defocusing Schrödinger equation), and the \(L_2\)-norm is its integral of motion. If this integral is of order one, then the results of [9] (see there Appendix 3) imply that at some point of each time-interval of order \(\delta ^{-1/3}\) the \(C^m\)-norm of a corresponding solution will become \(\ > rsim \delta ^{-m\kappa }\) if \(m\ge 2\), for any \(\kappa <1/3\). Furthermore, if \(n=2\) and \(\delta =1\), then due to [4] for \(m>1\) and any \(M>1\) there exists a \(T=T(m,M)\) and a smooth \(u_0(x)\) such that \(\Vert u_0\Vert _m< M^{-1}\) and \(\Vert u(T;u_0)\Vert _m>M\).

The paper is organized as follows. In Sect. 2, we recall the results from [8, 10] on solutions of the Eq. (1.1). Next we show in Sect. 3 that if the noise \(\eta \) is non-degenerate, the \(L^2\)-norm of a solution of Eq. (1.1) cannot stay too small on time intervals of order \({\mathcal {O}}(\frac{1}{\nu })\) with high probability, unless its \(H^2\)-norm gets very large

(see Lemma 12). Then in Sect. 4 we derive from this fact the assertion (1) of Theorem 1. We prove assertion (2) and both corollaries in Sect. 5.

Constants in estimates never depend on \(\nu \), unless otherwise stated. For a metric space M we denote by \({\mathcal {B}}(M)\) the Borel \(\sigma \)-algebra on M, and by \({{\mathcal {P}}}(M)\) – the space of probability Borel measures on M. By \({\mathcal {D}}(\xi )\) we denote the law of a r.v. \(\xi \), and by \(|\cdot |_p\) – the norm in \(L_p(K^n)\).

2 Solutions and estimates

Strong solutions for the Eq. (1.1) are defined in the usual way:

Definition 5

Let \((\Omega , {\mathcal {F}},\{{\mathcal {F}}_t\}_{t\geqslant 0}, {\mathbb {P}})\) be the filtered probability space as in the introduction. Let \(u_0\) in (1.5) be a r.v., measurable in \({\mathcal {F}}_0\) and independent from the Wiener process \(\xi \) (e.g., \(u_0(x)\) may be a non-random function). Then a random process \(u(t)=u(t;u_0)\in C_0(K^{n}) \), \(t\in [0,T]\), adapted to the filtration, is called a strong solution of (1.1), (1.5), if

  1. (1)

    a.s. its trajectories u(t) belong to the space

    $$\begin{aligned} {\mathcal {H}}([0,T]):=C([0,T],C_0(K^{n}))\cap L^2([0,T], H^1); \end{aligned}$$
  2. (2)

    we have

    $$\begin{aligned} u(t)=u_0+\int _0^t(\nu \Delta u-i|u|^2u)ds+\sqrt{\nu }\,\xi (t),\quad \forall t\in [0,T], \quad a.s., \end{aligned}$$

    where both sides are regarded as elements of \(H^{-1}\).

If (1)-(2) hold for every \(T<\infty \), then u(t) is a strong solution for \(t\in [0,\infty )\). In this case a.s. \(u\in C([0,\infty ),C_0(K^{n}))\cap L_{loc}^2([0,\infty ), H^1). \)

Everywhere below when we talk about solutions for the problem (1.1), (1.5) we assume that the r.v. \(u_0\) is as in the definition above.

The global well-posedness of Eq. (1.1) was established in [8, 10]:

Theorem 6

For any \(u_0\in C_0(K^{n})\) the problem (1.1), (1.5) has a unique strong solution \(u^\omega (t,x;u_0)\), \(t\ge 0\). The family of solutions \(\{ u^\omega (t;u_0)\}\) defines in the space \(C_0(K^{n})\) a Fellerian Markov process.

In [8, 10] the theorem above is proved when (1.4) is replaced by the weaker assumption \(B_*<\infty \), where \(\ B_* =\sum |b_d| \) (note that \(B_*\le C_n B_{m_*}^{1/2}\)).

The transition probability for the obtained Markov process in \(C_0(K^{n})\) is

$$\begin{aligned} P_t(u,\Gamma )={\mathbb {P}}\{u(t;u)\in \Gamma \}, \quad u\in C_0(K^{n}),\; \Gamma \in {\mathscr {B}}(C_0(K^{n})), \end{aligned}$$

and the corresponding Markov semigroup in the space \({\mathscr {P}}(C_0(K^{n}))\) of Borel measures on \(C_0(K^{n})\) is formed by the operators \(\{ {\mathcal {B}}_t^*, t\ge 0\}\),

$$\begin{aligned} {\mathcal {B}}_t^*\mu (\Gamma )=\int _{C_0(K^{n})}P_t(u,\Gamma )\mu (du),\quad t\in {\mathbb {R}}. \end{aligned}$$

Then \({\mathcal {B}}_t^*\mu = {\mathcal {D}}u(t;u_0)\) if \(u_0\) is a r.v., independent from \(\xi \) and such that \({\mathcal {D}}(u_0)=\mu \).

Introducing the slow time \(\tau =\nu t\) and denoting \(v(\tau ,x)=u(\frac{\tau }{\nu },x)\), we rewrite Eq. (1.1) in the following form, more convenient for some calculations:

$$\begin{aligned} \frac{\partial v}{\partial \tau }-\Delta v+i\nu ^{-1}|v|^2v={\tilde{\eta }}(\tau ,x), \end{aligned}$$
(2.1)

where

$$\begin{aligned}{\tilde{\eta }}(\tau ,x)=\frac{\partial }{\partial \tau }{\tilde{\xi }}(\tau ,x),\quad {\tilde{\xi }}(\tau ,x)=\sum _{d\in {\mathbb {N}}^n}b_d{\tilde{\beta }}_d(\tau )\varphi _d(x),\end{aligned}$$

and \({\tilde{\beta }}_d(\tau ):=\nu ^{1/2}\beta _d(\tau \nu ^{-1})\), \(d\in {\mathbb {N}}^d\), is another set of independent standard complex Brownian motions.

Let \(\Upsilon \in C^\infty ({\mathbb {R}})\) be any smooth function such

$$\begin{aligned}\Upsilon (r)={\left\{ \begin{array}{ll}0 ,&{}\text { for }r\leqslant \frac{1}{4};\\ r, &{}\text { for }r\geqslant \frac{1}{2}. \end{array}\right. }\end{aligned}$$

Writing \(v\in {\mathbb {C}}\) in the polar form \(v=re^{i\Phi }\), where \(r=|v|\), and recalling that \(\langle \cdot ,\cdot \rangle \) stands for the real scalar product in \({\mathbb {C}}\), we apply Itô’s formula to \(\Upsilon (|v|)\) and obtain that the process \(\Upsilon (\tau ):=\Upsilon (|v(\tau )|)\) satisfies

$$\begin{aligned} \begin{aligned} \Upsilon (\tau )&=\Upsilon _0+\int _0^\tau \Big [\Upsilon '(r)(\nabla r-r|\nabla \Phi |^2)\\&\quad +\frac{1}{2}\sum _{d\in {\mathbb {N}}^n}b_d^2\Big (\Upsilon ''(r)\langle e^{i\Phi },\varphi _d\rangle ^2+ \Upsilon '(r)\frac{1}{r}(|\varphi _d|^2-\langle e^{i\Phi },\varphi _d\rangle ^2)\Big )\Big ]ds+{\mathbb {W}}(\tau ),\end{aligned} \end{aligned}$$
(2.2)

where \(\Upsilon _0=\Upsilon (|v(0)|)\) and \({\mathbb {W}}(\tau )\) is the stochastic integral

$$\begin{aligned}{\mathbb {W}}(\tau )=\sum _{d\in {\mathbb {N}}^n}\int _0^\tau \Upsilon '(r)b_d\varphi _d\langle e^{i\Phi }, d{\tilde{\beta }}_d(s)\rangle .\end{aligned}$$

In [10] Eq. (2.1) is considered with \(\nu =1\) and, following [8], the norm \(|v(t)|_\infty \) of a solution v is estimated via \(\Upsilon (t)\) (since \(|v| \le \Upsilon +1/2\)). But the nonlinear term \(i\nu ^{-1}|v|^2v\) does not contribute to Eq. (2.2), which is the same as the \(\Upsilon \)-equation (2.3) in [10] (and as the corresponding equation in [8, Section 3.1]). So the estimates on \(|\Upsilon (t)|_\infty \) and the resulting estimates on \(|v(t)|_\infty \), obtained in [10], remain true for solutions of (2.1) with any \(\nu \). Thus we get the following upper bound for quadratic exponential moments of the \(L_\infty \)-norms of solutions:Footnote 2

Theorem 7

For any \(T>0\) there are constants \(c_*>0\) and \(C>0\), depending only on \(B_*\) and T, such that for any r.v. \(v_0^\omega \in C_0(K^{n})\) as in Definition 5, any \(\tau \geqslant 0\) and any \(c\in (0,c_*]\), a solution \(v(\tau ; v_0)\) of Eq. (2.1) satisfies

$$\begin{aligned} {\mathbb {E}}\exp (c \sup _{\tau \leqslant s\leqslant \tau +T}|v(s)|_\infty ^2)\leqslant C\, {\mathbb {E}}\exp (5c\ |v_0|^2_\infty )\le \infty . \end{aligned}$$
(2.3)

In [10] the result above is proved for a deterministic initial data \(v_0\). The theorem’s assertion follows by averaging the result of [10] in \(v_0^\omega \).

The estimate (2.3) is crucial for derivation of further properties of solutions, including the given below upper bounds for their Sobolev norms, obtained in the work [8]. Since the scaling of the equation in [8] differs from that in (2.1) and the result there is a bit less general than in the theorem below, a sketch of the proof is given in Appendix B.

Theorem 8

Assume that \(B_m<\infty \) for some \( m\in {\mathbb {N}}\), and \(v_0=v_0^\nu \in H^m\cap C_0(K^n)\) satisfies

$$\begin{aligned} |v_0|_\infty \le M, \quad \Vert v_0\Vert _m \le M_m \nu ^{-m}, \quad 0<\nu \le 1. \end{aligned}$$

Then

$$\begin{aligned} {\mathbb {E}}\Vert v(\tau ;v_0)\Vert _m^2\leqslant C_{m}\nu ^{-m},\quad \forall \tau \in [0,\infty ), \end{aligned}$$
(2.4)

where \(C_{M,m}\) also depends on M, \(M_m\) and \(B_m\), \(B_{m_*}\).

Neglecting the dependence on \(\nu \), we have that if \(B_m<\infty \), \(m\in {\mathbb {N}}\), and a r.v. \(v_0^\omega \in H^m\cap C_0(K^n)\) satisfies \({\mathbb {E}}\Vert v_0\Vert _m^2<\infty \) and \({\mathbb {E}}\exp (c\ |v_0|^2_\infty )<\infty \) for some \(c>0\), then Eq. (2.1) has a solution, equal \(v_0\) at \(t=0\), such that

$$\begin{aligned}&{\mathbb {E}}\Vert v(\tau ;v_0)\Vert _m^2 \le e^{- t} {\mathbb {E}}\Vert v_0\Vert _m^2 +C, \quad \tau \ge 0, \end{aligned}$$
(2.5)
$$\begin{aligned}&{\mathbb {E}}\sup _{0\le \tau \le T} \Vert v(\tau ;v_0)\Vert _m^2 \le C' , \end{aligned}$$
(2.6)

where \( C>0\) depend on \(c, \nu , {\mathbb {E}}\exp (c\ |v_0|^2_\infty ), B_{m_*}\) and \(B_m\), while \(C'\) also depends on \({\mathbb {E}}\Vert v_0\Vert _m^2<\infty \) and T. See Appendix B.

As it is shown in [10], the estimate (2.3) jointly with an abstract theorem from [11], imply that under a mild nondegeneracy assumption on the random force the Markov process in the space \(C_0(K^{n})\), constructed in Theorem 6, is mixing:

Theorem 9

For each \(\nu >0\), there is an integer \(N=N(B_*,\nu )> 0\) such that if \(b_d\ne 0\) for \(|d|\leqslant N\), then the Eq. (1.1) is mixing. I.e. it has a unique stationary measure \(\mu _\nu \in {\mathscr {P}}(C_0(K^{n}))\), and for any probability measure \(\lambda \in {\mathscr {P}}(C_0(K^{n}))\) we have \({\mathcal {B}}^*_t\lambda \rightharpoonup \mu _\nu \) as \(t\rightarrow \infty \).

Under the assumption of Theorem 8, for any \(u_0\in H^m\) the law \({\mathcal {D}}u(t;u_0)\) of a solution \(u(t;u_0)\) is a measure in \(H^m\). The mixing property in Theorem 9 and (2.4) easily imply

Corollary 10

If under the assumptions of Theorem 9\(B_m<\infty \) for some \(m\in {\mathbb {N}}\) and \(u_0\in H^m\), then \({\mathcal {D}}(u(t;u_0)) \rightharpoonup \mu _\nu \) in \({\mathcal {P}}(H^m)\).

In view of Theorems  78 with \(v_0=0\) and the established mixing, we have:

Corollary 11

Under the assumptions of Theorem 9, if \(v^{st}(\tau )\) is the stationary solution of the equation, then

$$\begin{aligned} {\mathbb {E}}\exp (c_*\sup _{\tau \leqslant s\leqslant \tau +T}|v^{st}(s)|_\infty ^2)\leqslant {\mathcal {C}}, \end{aligned}$$

where the constant \({\mathcal {C}}>0\) depends only on T and \(B_*\). If in addition \(B_m<\infty \) for some \(m\in {\mathbb {N}}\cup \{0\}\), then \( {\mathbb {E}}\Vert v^{st}(\tau )\Vert _m^2\leqslant C_{m}\nu ^{-m}, \) where \(C_m\) depends on \(B_*\) and \(B_m\).

Finally we note that applying Itô’s formula to \(\Vert v^{st}(\tau )\Vert _0^2\), where \(v^{st}\) is a stationary solution of (2.1), and taking the expectation we get the balance relation

$$\begin{aligned} {\mathbb {E}}\Vert v^{st}(\tau )\Vert ^2_1 = B_0. \end{aligned}$$
(2.7)

We cannot prove that \({\mathbb {E}}\Vert v^{st}(\tau )\Vert _0^2 \ge B'>0\) for some \(\nu \)-independent constant \(B'\), and cannot bound from below the energy \(\tfrac{1}{2} {\mathbb {E}}\Vert v(\tau ; v_0)\Vert ^2_0\) of a solution v by a positive \(\nu \)-independent quantity. Instead in next section we get a weaker conditional lower bound on the energies of solutions.

3 Conditional lower bound for the \(L^2\)-norm of solutions

In this section we prove the following result:

Lemma 12

Let \(B_2<\infty \) and \(u(\tau ;u_0)\), where \(u_0\in H^2\cap C_0(K^n)\) is non-random, be a solution of Eq. (2.1). Take any constants \( \chi >0, \Gamma \ge 1, \tau _0\ge 0, \) and define the stopping time

$$\begin{aligned} \tau _\Gamma :=\inf \{\tau \ge \tau _0 : \Vert u(\tau )\Vert _2\geqslant \Gamma \} \end{aligned}$$

(as usual, \(\tau _\Gamma =\infty \) if the set under the \(\inf \)-sign is empty). Then

$$\begin{aligned} {\mathbb {E}} \int _{\tau _0}^{\tau \wedge \tau _\Gamma } {\mathbb {I}}_{[0,\chi ]} (\Vert u(s)\Vert _0)ds\leqslant 2(1+\tau -\tau _0)B_0^{-1} \chi \Gamma , \end{aligned}$$
(3.1)

for any \(\tau >\tau _0\).

Proof

We establish the result by adapting the proof from [16] (also see [11, Theorem 5.2.12]) to non-stationary solutions. The argument relies on the concept of local time for semi-martingales (see e.g. [14, Chapter VI.1] for details of the concept). By \([\cdot ]_b\) we denote the quasinorm \( [u]_b^2 = \sum _d |u_d|^2 b_d^2. \)

Without loss of generality we assume \(\tau _0=0\). Otherwise we just need to replace \(u( \tau ,x)\) by the process \({\tilde{u}}(\tau ,x):=u(\tau +\tau _0,x)\), apply the lemma with \(\tau _0=0\) and with \(u_0\) replaced by the initial data \({\tilde{u}}_0^\omega =u^\omega (\tau _0; u_0)\), and then average the estimate in the random \({\tilde{u}}_0^\omega \).

Let us write the solution \(u(\tau ;u_0)\) as \(u(\tau )=\) \(\sum _{d\in {\mathbb {N}}^n}u_d(\tau )\varphi _d\). For any fixed function \(g\in C^2({\mathbb {R}})\), consider the process

$$\begin{aligned} f(\tau )=g(\Vert u(\tau \wedge \tau _\Gamma )\Vert _0^2). \end{aligned}$$

Since

$$\begin{aligned} \partial _ug(\Vert u\Vert _0^2)= & {} 2g'(\Vert u\Vert _0^2)\langle \!\langle u,\cdot \rangle \!\rangle , \quad \partial _{uu}g(\Vert u\Vert _0^2)=4g''(\Vert u\Vert _0^2)\langle \!\langle u,\cdot \rangle \!\rangle \langle \!\langle u,\cdot \rangle \!\rangle \\&+2g'(\Vert u\Vert ^2_0)\langle \!\langle \cdot ,\cdot \rangle \!\rangle , \end{aligned}$$

then by Itô’s formula we have

$$\begin{aligned} f(\tau )=f(0)+ \int _0^{\tau \wedge \tau _\Gamma }A(s)ds+\sum _{d\in {\mathbb {N}}^n}b_d\int _0^{\tau \wedge \tau _\Gamma } 2g'(\Vert u(s)\Vert _0^2)\langle u_d(s), d\beta _d(s)\rangle ,\end{aligned}$$
(3.2)
$$\begin{aligned} \begin{aligned} A(s)&=2g'(\Vert u\Vert _0^2)\langle \!\langle u, \Delta u-\frac{1}{\nu }i|u|^2u\rangle \!\rangle +2 \sum _d b_d^2 \big ( g''(\Vert u\Vert _0^2 ) |u_d|^2 +g'(\Vert u\Vert _0^2)\big ) \\&=- 2g'(\Vert u\Vert ^2_0)\Vert u\Vert _1^2+ 2g''(\Vert u\Vert _0^2) [u]_b^2 +2g'(\Vert u\Vert _0^2) B_0, \quad u=u(s). \end{aligned} \end{aligned}$$
(3.3)

Step 1: We firstly show that for any bounded measurable set \({G}\subset {\mathbb {R}}\), denoting by \({\mathbb {I}}_G\) its indicator function, we have the following equality

$$\begin{aligned} \begin{aligned} 2{\mathbb {E}} \int _0^{\tau \wedge \tau _\Gamma }&{\mathbb {I}}_G(f(s)) \,\big (g'(\Vert u(s)\Vert _0^2)\big )^2 \! [u(s)]_b^2 ds =\int _{-\infty }^\infty {\mathbb {I}}_G(a)\\ {}&\Big [{\mathbb {E}}(f(\tau )-a)_+-{\mathbb {E}}(f(0)-a)_+-{\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a+\infty )}(f(s))A(s)ds\Big ]da. \end{aligned}\end{aligned}$$
(3.4)

Let \(L(\tau ,a)\), \((\tau ,a)\in [0,\infty )\times {\mathbb {R}}\), be the local time for the semi martingale \(f(\tau )\) (see e.g. [14, Chapter VI.1]). Since in view of (3.2) the quadratic variation of the process \(f(\tau )\) is

$$\begin{aligned} d \langle f,f \rangle _s = \sum _d (2g'(\Vert u\Vert _0^2) |u_d| b_d)^2 =4\big ( g'(\Vert u\Vert _0^2)\big )^2 [u]_b^2, \end{aligned}$$

then for any bounded measurable set \(G\subset {\mathbb {R}}\), we have the following equality (known as the occupation time formula, see [14, Corollary VI.1.6]),

$$\begin{aligned} \int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_G(f(s)) 4\big (g'(\Vert u(s)\Vert _0^2)\big )^2\, [u(s)]_b^2 ds=\int _{-\infty }^\infty {\mathbb {I}}_{{G}}(a)L(\tau ,a)da.\end{aligned}$$
(3.5)

For the local time \(L(\tau ,a)\), due to Tanaka’s formula (see [14, Theorem VI.1.2]) we have

$$\begin{aligned} \begin{aligned}(f(\tau )-a)_+=&(f(0)-a)_+\\&+\sum _{d\in {\mathbb {N}}^n} b_d\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,+\infty )}\big (f(s)\big )2g'(\Vert u(s)\Vert _0^2) \langle u_d (s), d\beta _d(s)\rangle \\&+\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,+\infty )}(f(s))A(s)ds+\frac{1}{2}L(\tau ,a). \end{aligned}\end{aligned}$$
(3.6)

Taking expectation of both sides of (3.5) and (3.6) we obtain the required equality (3.4).

Step 2: Let us choose \({G}=[\rho _0,\rho _1]\) with \(\rho _1>\rho _0>0\), and \(g(x) =g_{\rho _0} (x) \in C^2({\mathbb {R}})\) such that \(g'(x)\geqslant 0\), \(g(x)=\sqrt{x}\) for \(x\geqslant \rho _0\) and \(g(x)=0\) for \(x\leqslant 0\). Then due to the factors \({\mathbb {I}}_G(f)\) and \({\mathbb {I}}_G(a)\) in (3.4), we may there replace g(x) by \(\sqrt{x}\), and accordingly replace \( g(\Vert u\Vert ^2_0), g'(\Vert u\Vert ^2_0)\) and \( g''(\Vert u\Vert ^2_0) \) by \(\Vert u\Vert _0\), \( \tfrac{1}{2}\Vert u\Vert _0^{-1}\) and \( -\tfrac{1}{4}\Vert u\Vert _0^{-3}\). So the relation (3.4) takes the form

$$\begin{aligned} \begin{aligned}&{\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma } {\mathbb {I}}_G(f(s)) \Vert u(s)\Vert _0^{-2} [u(s)]_b^2 =2\int _{\rho _0}^{\rho _1}\Big [{\mathbb {E}}(f(\tau )-a)_+-{\mathbb {E}}(f(0)-a)_+\Big ]da\\&\quad -2\int _{\rho _0}^{\rho _1}\Big \{{\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,+\infty )}\Big (f(s)\Big )\Big [\frac{2}{2\Vert u(s)\Vert _0}(B_0-\Vert u(s)\Vert _1^2)\\&\quad -\frac{2}{4\Vert u(s)\Vert _0^3} [u(s)]_b^2 \Big ]ds\Big \}da. \end{aligned}\end{aligned}$$

Since the l.h.s. of the above equality is non-negative, we have

$$\begin{aligned} \begin{aligned}&\int _{\rho _0}^{\rho _1}\Big [{\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,+\infty )}\Big (f(s)\Big )\frac{1}{\Vert u(s)\Vert _0^3}\Big (B_0\Vert u(s)\Vert _0^2-\tfrac{1}{2} [u(s)]_b^2 \Big )ds\Big ]da\\&\quad \leqslant \int _{\rho _0}^{\rho _1} {\mathbb {E}} \Big [\big ( (f(\tau )-a)_+ - (f(0)-a)_+\big ) + \int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,+\infty )}\big (f(s)\big )\frac{\Vert u(s)\Vert _1^2}{\Vert u(s)\Vert _0} ds\Big ]da. \end{aligned} \end{aligned}$$
(3.7)

Noting that

$$\begin{aligned} B_0\Vert u\Vert _0^2-\tfrac{1}{2} [u(s)]_b^2 =\sum _{d\in {\mathbb {N}}^n} (B_0-\frac{1}{2}b_d^2)|u_d|^2\geqslant \frac{B_0}{2} \Vert u\Vert _0^2, \end{aligned}$$

that by the definition of the stopping time \(\tau _\Gamma \)

$$\begin{aligned} (f(\tau )-a)_+-(f(0)-a)_+\leqslant \Gamma , \end{aligned}$$

and that by interpolation,

$$\begin{aligned}\int _0^{\tau \wedge \tau _\Gamma }\frac{\Vert u(s)\Vert _1^2}{\Vert u(s)\Vert _0}ds\leqslant \int _0^{\tau \wedge \tau _\Gamma }\Vert u(s)\Vert _2ds\leqslant (\tau \wedge \tau _\Gamma ) \Gamma , \end{aligned}$$

we derive from (3.7) the relation

$$\begin{aligned} \frac{B_0}{2}\! \int _{\rho _0}^{\rho _1} \big ( {\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,+\infty )}\big (f(s)\big ) \Vert u(s)\Vert _0^{-1} ds\big ) da \le (\rho _1-\rho _0) \Gamma \big (1 \,+ \, \tau ) . \end{aligned}$$

When \(\rho _0 \rightarrow 0\), we have \(g(x) \rightarrow \sqrt{x}\) and \(f(\tau ) \rightarrow \Vert u(\tau \wedge \tau _\Gamma )\Vert _0\). So sending \(\rho _0\) to 0 and using Fatou’s lemma we get from the last estimate that

$$\begin{aligned} \int _0^{\rho _1}{\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,\infty )}\Big (\Vert u(s)\Vert _0\Big )\Vert u(s)\Vert _0^{-1}dsda\leqslant 2\rho _1(1+\tau )B_0^{-1}\Gamma . \end{aligned}$$

As the l.h.s. above is not smaller than

$$\begin{aligned}\begin{aligned} \frac{1}{\chi }\int _0^{\rho _1}{\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,\chi ]}(\Vert u(s)\Vert _0)dsda, \end{aligned}\end{aligned}$$

then

$$\begin{aligned} \frac{1}{\rho _1}\int _0^{\rho _1} {\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,\chi ]}(\Vert u(s)\Vert _0)dsda\leqslant 2(1+\tau )B_0^{-1}\Gamma \,\chi . \end{aligned}$$
(3.8)

By the monotone convergence theorem

$$\begin{aligned} \lim _{a\rightarrow 0} {\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(a,\chi ]}(\Vert u(s)\Vert _0)ds = {\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(0,\chi ]}(\Vert u(s)\Vert _0)ds, \end{aligned}$$

so we get from (3.8) that

$$\begin{aligned} {\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{(0,\chi ]}(\Vert u(s)\Vert _0)ds\leqslant 2(1+\tau )B_0^{-1}\Gamma \chi . \end{aligned}$$
(3.9)

Step 3: We continue to verify that

$$\begin{aligned} {\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{\{0\}}(\Vert u(s)\Vert _0)ds=0. \end{aligned}$$
(3.10)

To do this let us fix any index \(d\in {\mathbb {N}}^n\) such that \(b_d\ne 0\). The process \(u_d(\tau )\) is a semimartingale, \(\ du_d = v_d ds +b_d d\beta _d, \, \) where \(v_d(s)\) is the d-th Fourier coefficient of \(\, \Delta u+\frac{1}{\nu }i|u|^2u\,\) for the solution \(u(\tau )=\sum _d u_d(\tau ) \varphi _d\) which we discuss. Consider the stopping time

$$\begin{aligned} \tau _R = \inf \{ s\le {\tau \wedge \tau _\Gamma }: |u(s)|_\infty \ge R\}. \end{aligned}$$

Due to (2.3) and (2.6), \( {\mathbb {P}} (\tau _R= \tau \wedge \tau _\Gamma ) \rightarrow 1\) as \(R\rightarrow \infty \). Let us denote \( u_d^R(\tau ) = u_d(\tau \wedge \tau _R). \) To prove (3.10) it suffices to verify that

$$\begin{aligned} \pi (\delta ) := {\mathbb {E}}\int _0^{\tau \wedge \tau _\Gamma }{\mathbb {I}}_{\{ |u_d(s)| <\delta \}} ds \rightarrow 0\quad \text {as} \quad \delta \rightarrow 0. \end{aligned}$$

If we replace above \(u_d\) by \(u_d^R\), then the obtained new quantity \( \pi ^R(\delta )\) differs from \( \pi (\delta )\) at most by \( {\mathbb {P}} (\tau _R< \tau \wedge \tau _\Gamma )\). The process \(u_d^R\) is an Ito process with a bounded drift. So by [7, Theorem 2.2.2, p. 52], \( \pi ^R(\delta )\) goes to zero with \(\delta \). Thus, given any \(\varepsilon >0\), we firstly choose R sufficiently big and then \(\delta \) sufficiently small to achieve \(\pi (\delta ) <\varepsilon \), for a suitable \(\delta (\varepsilon )>0\). So (3.10) is verified. Jointly with (3.9) this proves (3.1). \(\square \)

4 Lower bounds for Sobolev norms of solutions

In this section we work with Eq. (1.1) in the original time scale t and provide lower bounds for the \(H^m\)-norms of its solutions with \(m>2\). This will prove the assertion (1) of Theorem 1. As always, the constants do not depend on \(\nu \), unless otherwise stated.

Theorem 13

For any integer \(m\geqslant 3\), if \(B_m<\infty \) and

$$\begin{aligned} 0<\kappa <\tfrac{1}{35},\quad T_0\ge 0, \quad T_1>0, \end{aligned}$$

then for any r.v. \(u_0(x)\in H^m\cap C_0(K^n)\), satisfying

$$\begin{aligned} {\mathbb {E}}\Vert u_0\Vert _m^2<\infty , \quad {\mathbb {E}}\exp (c\ |u_0|^2_\infty ) \le C <\infty \end{aligned}$$
(4.1)

for some \(c, C >0\), we have

$$\begin{aligned} {\mathbb {P}}\left\{ \sup _{T_0 \le t \le T_0+{T_1}{\nu ^{-1}}} \Vert u(t;u_0)\Vert _{m} \ge K \nu ^{-m\kappa } \right\} \rightarrow 1\quad \text {as} \quad \nu \rightarrow 0, \end{aligned}$$
(4.2)

for every \(K>0\).

Proof

Consider the complement to the event in (4.2):

$$\begin{aligned} Q^\nu =\left\{ \sup _{T_0\leqslant t \leqslant T_0+\frac{T_1}{\nu }}\Vert u(t )\Vert _{m}<K\nu ^{-m\kappa }\right\} . \end{aligned}$$

We will prove the assertion (4.2) by contradiction. Namely, we assume that there exists a \(\gamma >0\) and a sequence \(\nu _j\rightarrow 0\) such that

$$\begin{aligned} {\mathbb {P}}( Q^{\nu _j})\geqslant 5\gamma \quad \text {for}\quad j=1,2,\dots , \end{aligned}$$
(4.3)

and will derive a contradiction. Below we write \( Q^{\nu _j}\) as Q and always suppose that

$$\begin{aligned} \nu \in \{\nu _1, \nu _2, \dots \}. \end{aligned}$$

The constants in the proof may depend on \({\mathcal {K}}, K, \gamma \), \(B_{\,m\vee m_*}\), but not on \(\nu \).

Without lost of generality we assume that \(T_1=1\). For any \(T_0>0\), due to (2.5) and (2.3) the r.v. \( {\tilde{u}}_0 := u(T_1) \) satisfies (4.1) with c replaced by c/5. So considering \({\tilde{u}}(t,x) = u(t+T_0,x)\) we may assume that \(T_0=0\).

Let us denote \(J_1=[0,\frac{1}{\nu }]\). Due to Theorem 7,

$$\begin{aligned} {\mathbb {P}}(Q_1)\geqslant 1-\gamma , \quad Q_1=\{\sup _{t\in J_1}|u(t)|_\infty \leqslant C_1(\gamma )\}, \end{aligned}$$

uniformly in \(\nu \), for a suitable \(C_1(\gamma )\). Then, by the definition of Q and Sobolev’s interpolation,

$$\begin{aligned} \Vert u^\omega (t)\Vert _l\leqslant C_{l,\gamma }\nu ^{-l\kappa }, \quad \omega \in Q\cap Q_1, \; t\in J_1, \end{aligned}$$
(4.4)

for \(l\in [0,m]\) (and any \(\nu \in \{\nu _1,\nu _2,\dots \}\)).

Denote \(J_2=[0,\frac{1}{2\nu }]\) and consider the stopping time

$$\begin{aligned} \tau _1=\inf \{t\in J_2:\Vert u(t)\Vert _2\geqslant C_{2,\gamma }\nu ^{-2\kappa }\} \le \tfrac{1}{2\nu }. \end{aligned}$$

Then \(\tau _1=\frac{1}{2\nu }\) for \(\omega \in Q\cap Q_1\). So due to (3.1) with \(\Gamma =C_{2,\gamma }\nu ^{-2\kappa }\), for any \(\chi >0\), we have

$$\begin{aligned} \begin{aligned}{\mathbb {E}}\big (\nu \int _{J_2}{\mathbb {I}}_{[0,\chi ]}(\Vert u(s)\Vert _0)ds{\mathbb {I}}_{Q\cap Q_1}(\omega )\big )&={\mathbb {E}}\big (\nu \int _0^{\frac{1}{2\nu }\wedge \tau _1}{\mathbb {I}}_{[0,\chi ]}(\Vert u(s)\Vert _0)ds{\mathbb {I}}_{Q\cap Q_1}(\omega )\big )\\&\leqslant {\mathbb {E}}\big (\nu \int _0^{\frac{1}{2\nu }\wedge \tau _1}{\mathbb {I}}_{[0,\chi ]}(\Vert u(s)\Vert _0)ds\big )\leqslant C \nu ^{-2\kappa }\chi . \end{aligned} \end{aligned}$$

Consider the event

$$\begin{aligned} \Lambda =\{\omega \in Q\cap Q_1:\Vert u(s)\Vert _0\leqslant \chi , \;\forall s\in J_2\}. \end{aligned}$$

Due to the above, we have,

$$\begin{aligned} {\mathbb {P}}(\Lambda )\leqslant 2 {\mathbb {E}}\big (\nu \int _{J_2}{\mathbb {I}}_{[0,\chi ]}(\Vert u(s)\Vert _0)ds{\mathbb {I}}_{Q\cap Q_1}(\omega )\big )\leqslant 2{C}\nu ^{-2\kappa }\chi . \end{aligned}$$

So \({\mathbb {P}}(\Lambda )\leqslant \gamma \) if we choose

$$\begin{aligned} \chi =c_3(\gamma )\nu ^{2\kappa }, \quad c_3(\gamma )=\gamma (2{C})^{-1}. \end{aligned}$$
(4.5)

Let us set

$$\begin{aligned} Q_2=(Q\cap Q_1)\setminus \Lambda , \quad {\mathbb {P}}(Q_2)\geqslant 3\gamma , \end{aligned}$$
(4.6)

and for \(\chi \) as in (4.5), consider the stopping time

$$\begin{aligned} {{\tilde{\tau }}}_1=\inf \{t\in J_2:\Vert u(t)\Vert _0\geqslant \chi \}. \end{aligned}$$

Then \({{\tilde{\tau }}}_1\leqslant \frac{1}{2\nu }\) for all \(\omega \in Q_2\). Consider the function

$$\begin{aligned} v(t,x) := u ({{\tilde{\tau }}}_1+t,x), \quad t \in [0,\tfrac{1}{2\nu }]. \end{aligned}$$

It solves Eq. (1.1) with modified Wiener processes and with initial data \(v_0(x) = u^\omega ({{\tilde{\tau }}}_1,x)\), satisfying

$$\begin{aligned} \Vert v_0^\omega \Vert _0 \ge \chi =c\nu ^{2\kappa } \quad \text {if}\quad \omega \in Q_2. \end{aligned}$$
(4.7)

Now we introduce another stopping time, in terms of v(tx):

$$\begin{aligned}\tau _2=\inf \{t\in [0,\tfrac{1}{2\nu }]: \Vert v(t)\Vert _m\geqslant K\nu ^{-m\kappa }\}\le \tfrac{1}{2\nu }. \end{aligned}$$

For \(\omega \in Q_2\), \(\tau _2=\tfrac{1}{2\nu }\) and in view of (4.4)

$$\begin{aligned} \Vert v^\omega (t)\Vert _l\leqslant C_3(\gamma )\nu ^{-l\kappa }, \quad t\in [0,\tfrac{1}{2\nu }],\; l\in [0,m], \quad \forall \, \omega \in Q_2. \end{aligned}$$
(4.8)

Step 1: Let us estimate from above the increment \({\mathscr {E}}(t,x)=|v(t\wedge \tau _2,x)|^2-|v_0(x)|^2\). Due to Itô’s formula, we have that

$$\begin{aligned} \begin{aligned} {\mathscr {E}}(t,x)&=2\nu \int _{0}^{t\wedge \tau _2}\Big (\langle v(s,x),\Delta v(s,x)\rangle +\sum _{d\in {\mathbb {N}}^n}b_d^2\varphi ^2_d(x)\Big )ds + \sqrt{\nu } \, M(t,x), \\ M(t,x)&= \int _{0}^{t\wedge \tau _2}\sum _{d\in {\mathbb {N}}^n}b_d\varphi _d(x)\langle v(s,x),d\beta _d(s)\rangle . \, \end{aligned} \end{aligned}$$

We treat M as a martingale M(t) in the space \(H^1\). Since in view of (A.3) for \(0\le s<\tau _2\) we have

$$\begin{aligned} \Vert v(s) \varphi _d\Vert _1 \le C\big ( |v(s)|_\infty \Vert \varphi _d\Vert _1 + |v(s)\Vert _1 |\varphi _d|_\infty \big ) \le C(\zeta d+\zeta ^{(m-1)/m} \nu ^{-\kappa }), \end{aligned}$$

where \(\zeta = \sup _{0\leqslant s\leqslant \frac{1}{\nu }}|u(s)|_\infty \) (the assertion is empty if \(\tau _2=0\)), then for any \(0<T_*\le \tfrac{1}{2\nu }\)

$$\begin{aligned} {\mathbb {E}}\Vert M(T_*)\Vert _1^2 \le \int _0^{T_*} {\mathbb {E}}\, \sum _d b_d^2 \Vert \varphi _d v(s)\Vert _1^2 ds \le C T_* \nu ^{-2\kappa }, \end{aligned}$$
(4.9)

where we used that \(B_1<\infty \). So by Doob’s inequality

$$\begin{aligned} {\mathbb {P}}\Big (\sup _{0\leqslant s\leqslant T_*} \Vert M(s)\Vert _1^2\geqslant r^2\Big )\leqslant C T_*r^{-2} \nu ^{-2\kappa }, \quad \forall r>0. \end{aligned}$$
(4.10)

Let us choose

$$\begin{aligned} T_*=\nu ^{-b}, \quad b\in (0, 1), \end{aligned}$$

where b will be specified later. Then \( 1\le T_*\le \tfrac{1}{2\nu } \) if \(\nu \) is sufficiently small, so due to (4.10)

$$\begin{aligned} {\mathbb {P}}( Q_3)\geqslant 1-\gamma , \quad Q_3=\{\sup _{0\leqslant \tau \leqslant T_*}\Vert M(\tau )\Vert _1\leqslant C_4(\gamma )\nu ^{-\kappa }\sqrt{T_*}\}, \end{aligned}$$

for a suitable \(C_4(\gamma )\) (and for \(\nu \ll 1\)); thus \( {\mathbb {P}} (Q_2\cap Q_3) \ge 2\gamma \). Since \( \Vert \langle v , \Delta v\rangle \Vert _1 \le C|v|_\infty \Vert v\Vert _3\) by (A.2) and \( \Vert \sum _d b_d \varphi _d\Vert _1 \le C, \) then in view of (4.8) and the definition of \(Q_3\),

$$\begin{aligned} \Vert {\mathscr {E}}^\omega (\tau )\Vert _1\leqslant C(\gamma )(\nu ^{1-3\kappa }T_*+\nu ^{\frac{1}{2}-\kappa } T_*^{1/2}), \qquad \forall \, \tau \in [0,T_*] , \;\; \forall \,\omega \in Q_2\cap Q_3. \end{aligned}$$
(4.11)

Step 2: For any \(x\in K^n\), denoting \( R(t) = |v(t,x)|^2, \, a(t) = \Delta v(t,x) \) and \(\xi (t) = \xi (t,x)\), we write the equation for \(v(t) := v(t,x)\) as an Itô process:

$$\begin{aligned} dv(t) = (-i R v +\nu a) dt + \sqrt{\nu }\, d\xi (t). \end{aligned}$$
(4.12)

Setting \(w(t) = e^{i\int _0^t R(s)ds} v(t)\), we observe that w also is an Itô process, \( w(0) =v_0\) and \( dv = e^{-i\int _0^t R(s)ds} dw -i Rv\, dt. \) From here and (4.12),

$$\begin{aligned} w(t) = v_0 +\nu \int _0^t e^{i\int _0^s R(s')ds'} a(s) ds + \sqrt{\nu }\, \int _0^t e^{i\int _0^s R(s')ds'} d \xi (s). \end{aligned}$$

So \(v(t\wedge \tau _2)= v(t\wedge \tau _2,x)\) can be written as

$$\begin{aligned} v(t\wedge \tau _2,x)= I_1(t\wedge \tau _2,x) + I_2(t\wedge \tau _2,x)+ I_3(t\wedge \tau _2,x), \end{aligned}$$
(4.13)

where

$$\begin{aligned}\begin{aligned}&I_1(t,x)= e^{-i\int _0^{t}|v(s,x)|^2ds}v_0,\quad I_2(t,x)=\nu \int _0^{t}e^{-i\int _{s}^{t}|v(s',x)|^2ds'}\Delta v(s,x)ds, \\&I_3(t,x)= \sqrt{\nu } e^{-i\int _0^{t}|v(s',x)|^2ds'} \int _0^{t}e^{i\int _0^{s}|v(s',x)|^2ds'}d\xi (s,x) . \end{aligned} \end{aligned}$$

Our next goal is to obtain a lower bound for \( \Vert v(T_*)\Vert _1\) when \(\omega \in Q_2\cap Q_3\), using the above decomposition (4.13).

Step 3: We first deal with the stochastic term \(I_3(t)\). For \(0 \leqslant s\leqslant s_1\leqslant T_*\wedge \tau _2\) we set

$$\begin{aligned} \begin{aligned} W(s,s_1, x):= \exp ( i \int _{s}^{s_1}|v(s',x)|^2ds'), \quad F(s,s_1,x) := \int _{s}^{s_1}|v(s',x)|^2ds'; \end{aligned} \end{aligned}$$
(4.14)

then \( W(s,s_1, x) = \exp \big (i F(s,s_1, x)\big )\). The functions F and W are periodic in x, but not odd. Speaking about them we understand \(\Vert \cdot \Vert _m\) as the non-homogeneous Sobolev norm, so \( \Vert F\Vert _m^2 = \Vert F\Vert _0^2 + \Vert ( -\Delta )^{m/2} F\Vert _0^2, \) etc. We write \(I_3\) as

$$\begin{aligned} I_3(t) =\sqrt{\nu }\ {\overline{W}} ( 0,t \wedge \tau _2 ,x) \int _0^{t\wedge \tau _2} W(0,s,x) d\xi (s,x). \end{aligned}$$
(4.15)

In view of (A.1),

$$\begin{aligned} \Vert \exp (iF(s,s_1 \cdot ))\Vert _{k} \le C _{k}(1+ |F(s,s_1, \cdot )|_\infty )^{k-1} \Vert F(s,s_1,\cdot )\Vert _{k}, \quad k \in {\mathbb {N}}. \end{aligned}$$
(4.16)

For any \(s\in J =[ 0, T_*\wedge \tau _2)\), by (A.3) and the definition of \( \tau _2\), we have that \(v:=v(s)\) satisfies

$$\begin{aligned} \Vert |v|^2\Vert _{1} \le C |v|_\infty \Vert v\Vert _{1} \le C |v|_\infty \Vert v\Vert _0^{1- 1/m} \Vert v\Vert _m^{1/m} \le C' |v|_\infty ^{2-1/m} \nu ^{-\kappa } \end{aligned}$$
(4.17)

(this assertion is empty if \( \tau _2=0\) since then \(J=\emptyset \)). So for \(s,s_1 \in J\),

$$\begin{aligned}&|F(s,s_1,\cdot )|_\infty \leqslant |s_1-s| \sup _{s'\in J}|v(s')|^2_\infty ,\quad \Vert F (s,s_1,\cdot )\Vert _{k} \\&\quad \le C \nu ^{ -\kappa k} |s_1-s|\big (\sup _{s' \in J} | v(s')|_\infty \big )^{2-k/m}\;\; \end{aligned}$$

for \(k\le m\). Then, due to (4.16),

$$\begin{aligned} \Vert W(0,s\wedge \tau _2,\cdot )\Vert _{1} \leqslant C' T_* \nu ^{-\kappa } (1+ \sup _{s \in J} |v(s)|_\infty ^{2} ). \end{aligned}$$
(4.18)

Consider the stochastic integral in (4.15),

$$\begin{aligned} N(t,x)=\int _0^{t}W(0,s,x) d\xi (s,x). \end{aligned}$$

The process \( t \mapsto W(0,t,x) \) is adapted to the filtration \(\{ {\mathcal {F}}_t\}\), and

$$\begin{aligned}dW(0,t,x)=i|v(t,x)|^2W(0,t,x)dt.\end{aligned}$$

So integrating by parts (see, e.g., [14, Proposition IV.3.1]) we re-write N as

$$\begin{aligned}N(t,x)=W(0,t,x)\xi (t,x)- i \int _0^t\xi (s,x) |v(s,x)|^2W(0,s,x)ds, \end{aligned}$$

and we see from (4.15) that

$$\begin{aligned} I_3(t)=\sqrt{\nu }\xi (t\wedge \tau _2,x)+ i\sqrt{\nu }\int _0^{t\wedge \tau _2}\xi (s,x) |v(s,x)|^2W(s,t\wedge \tau _2,x)ds.\nonumber \\ \end{aligned}$$
(4.19)

Due to (1.4) and since \(B_m<\infty \), the Wiener process \(\xi (t,x)\) satisfies

$$\begin{aligned}{\mathbb {E}}\Vert \xi (T_*,x)\Vert _1^2\leqslant CB_1T_*,\end{aligned}$$

and

$$\begin{aligned}{\mathbb {E}}\sup _{0\leqslant t\leqslant T_*}|\xi (t,\cdot )|_\infty \leqslant \sum _{d\in {\mathbb {N}}^n}b_d({\mathbb {E}}\sup _{0\leqslant t\leqslant T_*}|\beta _d(t)\varphi _d|_\infty )\leqslant CB_*\sqrt{T_*},\end{aligned}$$

(we recall that \(B_*=\sum _{d\in {\mathbb {N}}^n}|b_d|<\infty \)). Therefore,

$$\begin{aligned} {\mathbb {P}}(Q_4)\geqslant 1-\gamma , \quad Q_4=\{ \sup _{0\leqslant t\leqslant T_*} ( \Vert \xi (t)\Vert _1\vee |\xi (t)|_\infty ) \leqslant CT_*^{1/2}\}, \end{aligned}$$

with a suitable \(C=C(\gamma )\). Let

$$\begin{aligned} {\tilde{Q}}=\bigcap _{i=1}^4 Q_i, \end{aligned}$$

then \({\mathbb {P}}({\tilde{Q}})\geqslant \gamma \). As \(\tau _2 = T_*\) for \(\omega \in {\tilde{Q}}\), then due to (4.17), (4.18), (4.19) and (A.3), for \(\omega \in {\tilde{Q}}\) we have

$$\begin{aligned} \begin{aligned} \sup _{0\leqslant t\leqslant T_*} \Vert I^\omega _3(t)\Vert _1&\le \sqrt{\nu }\, \sup _{0\leqslant t\leqslant T_*} \Big (\Vert \xi ^\omega (t)\Vert _1+\int _0^t\Vert \xi ^\omega (s)|v^\omega (s)|^2W^\omega (s,t)\Vert _1ds\Big )\\&\leqslant {C} T_*^{5/2}\nu ^{\frac{1}{2}-\kappa }. \end{aligned} \end{aligned}$$
(4.20)

Setp 4: We then consider the term \( I_{2}= \nu \int _0^{t\wedge \tau _2} {\bar{W}}(s, t\wedge \tau _2,x) \Delta v(s,x)ds \). To bound its \(H^1\)-norm we need to estimate \(\Vert W\Delta v\Vert _1\). Since

$$\begin{aligned} \Vert \partial _x^a W\partial _x^bv\Vert _0\leqslant C \Vert W\Vert _3^{1/3}\Vert v\Vert _3^{2/3}|v|_\infty ^{1/3}\quad \text {if} \;\; |a|=1,|b|=2, \end{aligned}$$

(see [17, Proposition 3.6]), we have

$$\begin{aligned} \Vert W\Delta v\Vert _1\leqslant C(\Vert v\Vert _3+ \Vert W\Vert _3^{1/3}\Vert v\Vert _3^{2/3}|v|_\infty ^{1/3}).\end{aligned}$$

Then in view of (4.16) and (4.8), for \(\omega \in {\tilde{Q}}\)

$$\begin{aligned} \Vert W\Delta v\Vert _1 \le C\big ( \nu ^{-3\kappa } +(T_*^3 \nu ^{-3\kappa })^{1/3} \nu ^{-2\kappa } \big ) \le C \nu ^{-3\kappa } T_*, \end{aligned}$$

and accordingly

$$\begin{aligned} \sup _{0\leqslant t\leqslant T_*}\Vert I^\omega _2(t)\Vert _1\leqslant \nu \sup _{0\leqslant t\leqslant T_*}\!\int _0^{t}\Vert W^\omega (s,T_*)\Delta v^\omega (s)\Vert _1ds \leqslant C \nu ^{1-3\kappa }T_*^{2},\;\; \; \forall \,\omega \in {\tilde{Q}}. \end{aligned}$$
(4.21)

Step 5: Now we estimate from below the \(H^1\)-norm of the term \(I^\omega _{1}(T_*,x)\), \(\omega \in {\tilde{Q}}\). Writing it as \( I^\omega _1(T_*,x) = e^{-i T_*|v_0(x)|^2} e^{ -i \int _0^{T_*} {\mathscr {E}} (s,x) ds} v_0(x) \) wee see that

$$\begin{aligned} \Vert I^\omega _1(T_*) \Vert _1\geqslant \Vert \nabla (\exp (-iT_*|v_0|^2)v_0\Vert _0-\Vert \nabla (\exp (-i\int _0^{T_*} {\mathscr {E}} (s)ds))v_0\Vert _0-\Vert v_0\Vert _1. \end{aligned}$$

This first term on the r.h.s is

$$\begin{aligned} T_*\Vert v_0\nabla (|v_0|^2)\Vert _0=T_*\tfrac{2}{3}\Vert \nabla |v_0|^3\Vert _0\geqslant CT_* \Vert |v_0|^3\Vert _0\geqslant CT_*\Vert v_0\Vert _0^3\ge C T_* \nu ^{6\kappa },\quad C>0, \end{aligned}$$

where we have used the fact that \(u|_{\partial K^{n}}=0\), Poincaré’s inequality and (4.7).

For \(\omega \in {\tilde{Q}}\) and \(0\le s\le T_*\), in view of (4.11), the second term is bounded by

$$\begin{aligned} \begin{aligned} \left\| \left( \int _0^{T_*}\nabla {\mathscr {E}}(s)ds\right) v_0\right\| _0 \leqslant CT_*|v_0|_\infty \sup _{0\leqslant s\leqslant T_*}\Vert {\mathscr {E}}(s)\Vert _1 \le {C} T_*( \nu ^{1-3\kappa } T_* + \nu ^{\frac{1}{2}-\kappa } T_*^{1/2} ). \end{aligned} \end{aligned}$$

Therefore, using (4.11), we get for the term \(I^\omega _1(T_*)\) the following lower bound:

$$\begin{aligned} \Vert I_{1}^\omega (T_*)\Vert _1\geqslant C \Big (\nu ^{6\kappa }T_*- T_*\big (\nu ^{1-3\kappa }T_* +\nu ^{\frac{1}{2}-\kappa } T_*^{1/2}\big ) -\nu ^{-\kappa } \Big ). \end{aligned}$$

Recalling \(T_*=\nu ^{-b}\) we see that if we assume that

$$\begin{aligned} {\left\{ \begin{array}{ll} 6\kappa -b<-\kappa ,\\ 6\kappa -b< 1-3\kappa -2b, \\ 6\kappa -b < 1/2 -\kappa -\frac{3}{2}b , \end{array}\right. } \end{aligned}$$
(4.22)

then for \(\omega \in {\tilde{Q}}\),

$$\begin{aligned} \Vert I_{1}^\omega (T_*)\Vert _1\geqslant C \nu ^{6\kappa }T_*, \quad C >0, \end{aligned}$$
(4.23)

provided that \(\nu \) is sufficiently small.

Step 6: Finally, remembering that \(\tau _2=T_*\) for \(\omega \in {\tilde{Q}}\) and combining the relations (4.20), (4.21) and (4.23) to estimate the terms of (4.13), we see that for \(\omega \in {\tilde{Q}}\) we have

$$\begin{aligned} \Vert v^\omega (T_*)\Vert _1\geqslant \Vert I_{1}^\omega (T_*)\Vert _1-\Vert I^\omega _2(T_*)\Vert _1-\Vert I^\omega _3(\tau _*)\Vert _1\geqslant \tfrac{1}{2}C_1 \nu ^{6\kappa -b}, \quad C_1>0, \end{aligned}$$
(4.24)

if we assume in addition to (4.22) that

$$\begin{aligned} 6\kappa -b<\frac{1}{2}-\kappa -\frac{5}{2}b, \end{aligned}$$
(4.25)

and \(\nu \) is small. Note that this relation implies the last two in (4.22).

Combining (4.8) and (4.24) we get that

$$\begin{aligned} \nu ^{-b+7\kappa }\leqslant C^{-1}_2, \end{aligned}$$
(4.26)

for all sufficiently small \(\nu \). Thus we have obtained a contradiction with the existence of the sets \(Q^{\nu _j}\) as at the beginning of the proof if (for a chosen \(\kappa \)) we can find a \(b\in (0,1)\) which meets (4.22), (4.25) and

$$\begin{aligned} -b+7\kappa <0. \end{aligned}$$

Noting that this is nothing but the first relation in (4.22), we see that we have obtained a contradiction if

$$\begin{aligned} \kappa< \tfrac{1}{7} b, \quad \kappa < \tfrac{1}{14} -\tfrac{3}{14} b, \end{aligned}$$

for some \(b\in (0,1)\). We see immediately that such a b exists if and only if \(\kappa < \tfrac{1}{35} \).

\(\square \)

Amplification If we replace the condition \(m\geqslant 3\) with the weaker assumption

$$\begin{aligned} {\mathbb {R}}\ni m>2, \end{aligned}$$

then the statement (4.2) remains true for \(0<\kappa <\kappa (n,m)\) with a suitable (less explicit) constant \(\kappa (n,m)>0\). In this case we obtain a contradiction with the assumption (4.3) by deriving a lower bound for \(\Vert v(T_*)\Vert _\alpha \), where \(\alpha =\min \{1,m-2\}\in (0,1]\), using the decomposition (4.13). The proof remains almost identical except that now, firstly, we bound \(\Vert I_{2}\Vert _\alpha \) (\(\alpha <1\)) from above using the following estimate from [15, Theorem 5, p. 206] (also see there p. 14):

$$\begin{aligned}\Vert W\Delta u\Vert _\alpha \leqslant C\Vert u\Vert _{2+\alpha }(|W|_\infty +|W|_\infty ^{1-\frac{2\alpha }{n}}\Vert W\Vert _2^{\frac{2\alpha }{n}});\end{aligned}$$

and, secondly, estimate \(\Vert I_{1}^\omega (T_*)\Vert _\alpha \) (\(\alpha <1\)) from below as

$$\begin{aligned} \Vert I_{1}^\omega (T_*)\Vert _\alpha \geqslant {\Vert I_{1}^\omega (T_*)\Vert _1^{2-\alpha }} \,{\Vert I^\omega _2(T_*)\Vert _2^{-1+\alpha }}, \end{aligned}$$

which directly follows from Sobolev’s interpolation. See [6] for more details.

5 Lower bounds for time-averaged Sobolev norms

In this section we prove the assertion (2) of Theorem 1. We provide each space \(H^r\), \(r\ge 0\), with the scalar product

$$\begin{aligned} \langle \!\langle u,v\rangle \!\rangle _r:=\langle \!\langle (-\Delta )^{\frac{r}{2}}u,(-\Delta )^{\frac{r}{2}}v\rangle \!\rangle , \end{aligned}$$

corresponding to the norm \(\Vert u\Vert _r\). Let \( u(t) =\sum u_d(t) \varphi _d \) be a solution of Eq. (1.1). Applying Itô’s formula to the functional \(\Vert u\Vert _m^2\), we have for any \(0\leqslant t<t'<\infty \) the relation

$$\begin{aligned} \begin{aligned} \Vert u(t')\Vert _m^2=&\Vert u(t)\Vert _m^2+2\int _{t}^{t'}\langle \!\langle u(s), \nu \Delta u(s) - i|u(s)|^2u(s)\rangle \!\rangle _mds\\&+2\nu B_m(t'-t)+2\sqrt{\nu } M(t,t'), \end{aligned}\end{aligned}$$
(5.1)

where M stands for the real scalar productis the stochastic integral

$$\begin{aligned} M(t,t'):= \int _t^{t'}\sum _{d\in {\mathbb {N}}^n} b_d|d|^{2m} \langle u_d(s), d\beta _d(s)\rangle . \end{aligned}$$

Let us fix a \(\gamma \in (0,\frac{1}{8})\). Due to Theorems 7 and  13, for small enough \(\nu \) there exists an event \(\Omega _1\subset \Omega \), \( {\mathbb {P}}(\Omega _1)\geqslant 1-\gamma /2, \) such that for all \(\omega \in \Omega _1\) we have:

  1. a)

    \(\sup _{0\leqslant t\leqslant \frac{1}{\nu }}|u^\omega (t)|_\infty \leqslant C(\gamma )\), for a suitable \(C(\gamma )>0\);

  2. b)

    there exist \(t_\omega \in [0,\frac{1}{3\nu }]\) and \(t_\omega '\in [\frac{2}{3\nu },\frac{1}{\nu }]\) satisfying

    $$\begin{aligned} \Vert u^\omega (t_\omega )\Vert _m,\;\Vert u^\omega (t_\omega ')\Vert _m\geqslant \nu ^{-m\kappa }. \end{aligned}$$
    (5.2)

Since for the martingale M(0, t) we have that

$$\begin{aligned} {\mathbb {E}}|M(0,\tfrac{1}{\nu })|^2\leqslant B_m {\mathbb {E}}\int _0^\frac{1}{\nu } \Vert u(s)\Vert _m^2ds =: X_m, \end{aligned}$$

then by Doob’s inequality

$$\begin{aligned} {\mathbb {P}}(\Omega _2)\geqslant 1-\frac{\gamma }{2}, \qquad \Omega _2=\Big \{ \sup _{0\leqslant t\leqslant \frac{1}{\nu }}|M(0,t)|\leqslant c(\gamma ) X_m^{1/2} \Big \}. \end{aligned}$$

Now let us set \(\ {\hat{\Omega }}=\Omega _1\cap \Omega _2.\) Then \({\mathbb {P}}({\hat{\Omega }})\geqslant 1-\gamma \) for small enough \(\nu \), and for any \(\omega \in {\hat{\Omega }}\) there are two alternatives:

(i) there exists a \(t_\omega ^0\in [0,\frac{1}{3\nu }]\) such that \(\Vert u^\omega (t^0_\omega )\Vert _m=\frac{1}{3}\nu ^{-\kappa m}\). Then from (5.1) and (5.2) in view of (A.4) we get

$$\begin{aligned}\begin{aligned}&\frac{8}{9}\nu ^{-2m\kappa }+2\nu \int _{t_\omega ^0}^{t'_\omega }\Vert u^\omega (s)\Vert _{m+1}^2ds \leqslant C(m,\gamma ) \int _0^{\frac{1}{\nu }} \Vert u^\omega (s)\Vert _m^2ds+2B_m\\&\quad +2\sqrt{\nu }c(\gamma ) X_m^{1/2}. \end{aligned} \end{aligned}$$

(ii) There exists no \(t\in [0,\frac{1}{3\nu }]\) with \(\Vert u^\omega (t)\Vert _m=\frac{1}{3}\nu ^{-\kappa m}\). In this case, since \(\Vert u^\omega (t)\Vert _m\) is continuous with respect to t, then due to (5.2) \( \Vert u^\omega (t)\Vert _m > \frac{1}{3}\nu ^{-m\kappa } \) for all \( t\in [0,\frac{1}{3\nu }]\). This leads to the relation

$$\begin{aligned} \tfrac{1}{27}\nu ^{-2m\kappa -1}\leqslant \int _0^{\frac{1}{\nu }}\Vert u^\omega (s)\Vert _m^2ds. \end{aligned}$$

In both cases for \(\omega \in {\hat{\Omega }}\) we have:

$$\begin{aligned}\frac{1}{27}\nu ^{-2m\kappa }\leqslant C'(m,\gamma ) \int _0^{\frac{1}{\nu }} \Vert u(s)\Vert _m^2ds+2B_m +{\nu } c(\gamma )^2 + X_m. \end{aligned}$$

It implies that

$$\begin{aligned} {\mathbb {E}}\nu \int _0^{\frac{1}{\nu }}\Vert u(\tau )\Vert _m^2d\tau \geqslant C\nu ^{-2m\kappa +1} \end{aligned}$$

(for small enough \(\nu \)), and gives the lower bound in (1.7).

The upper bound follows directly from Theorem 8.

Proof of Corollaries 3 and 4

Since \(B_k<\infty \) for each k and all coefficients \(b_d\) are non-zero, then Eq. (1.1) is mixing in the spaces \(H^m\), \(m\in {\mathbb {N}}\), see Corollary 10. As the stationary solution \(v^{st} \) satisfies Corollary 11 with any m, then for each \(\mu \in {\mathbb {N}}\) and \(M>0\), interpolating the norm \(\Vert u\Vert _\mu \) via \(\Vert u\Vert _0\) and \(\Vert u\Vert _m\) with m sufficiently large we get that the stationary measure \(\mu _\nu \) satisfies

$$\begin{aligned} \int \Vert u\Vert _\mu ^M \mu _\nu (du) < \infty \quad \forall \, \mu \in {\mathbb {N}}, \;\forall M>0. \end{aligned}$$
(5.3)

Similar, in view of (2.5) and Theorem  7,

$$\begin{aligned} {\mathbb {E}}\Vert u(t;u_0)\Vert _\mu ^M \le C_\nu (u_0) \quad \forall t\ge 0, \end{aligned}$$
(5.4)

for each \(u_0\in C^\infty \) and every \(\mu \) and M as in (5.3). Now let us consider the integral in (1.7) and write it as

$$\begin{aligned} J_t := \nu \int _t^{t+\nu ^{-1}} {\mathbb {E}}\Vert u(s)\Vert _m^2ds. \end{aligned}$$

Replacing the integrand in \(J_t\) with \({\mathbb {E}}( \Vert u_\nu (s)\Vert _m \wedge N)^2\), \( N\ge 1\), using the convergence

$$\begin{aligned} {\mathbb {E}}\big ( \Vert u(s;v_0) \Vert _m \wedge N\big )^2 \rightarrow \int \big ( \Vert u \Vert _m \wedge N\big )^2 \mu _\nu (du) \quad \text {as} \quad s\rightarrow \infty \quad \forall \,N, \end{aligned}$$
(5.5)

which follows from Corollary 10, and the estimates (5.3), (5.4) we get that

$$\begin{aligned} J_t \, \rightarrow \, \int \Vert u\Vert _m^2 \,\mu _\nu (du)\quad \text {as} \quad t\rightarrow \infty . \end{aligned}$$
(5.6)

This convergence and (1.7) imply the assertion of Corollary 3.

Now the convergence (5.5) jointly with estimates (5.3), (5.4) and (1.8) imply Corollary 4. \(\square \)