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

Strong Approximation of Stochastic Semiclassical Schrödinger Equation with Multiplicative Noise

Lihai JI Institute of Applied Physics and Computational Mathematics, Beijing 100094, China, and Shanghai Zhangjiang Institute of Mathematics, Shanghai 201203, China [email protected]  and  Zhihui LIU Department of Mathematics & National Center for Applied Mathematics Shenzhen (NCAMS) & Shenzhen International Center for Mathematics, Southern University of Science and Technology, Shenzhen 518055, China [email protected]
Abstract.

We consider the stochastic nonlinear Schrödinger equation driven by a multiplicative noise in a semiclassical regime, where the Plank constant is small. In this regime, the solution of the equation exhibits high-frequency oscillations. We design an efficient numerical method combining the spectral Galerkin approximation and the midpoint scheme. This accurately approximates the solution, or at least of the associated physical observables. Furthermore, the strong convergence rate for the proposed scheme is derived, which explicitly depends on the Planck constant. This conclusion implies the semiclassical regime’s admissible meshing strategies for obtaining “correct” physical observables.

Key words and phrases:
stochastic semiclassical Schrödinger equation, spectral Galerkin method, midpoint scheme, strong convergence rate
2010 Mathematics Subject Classification:
Primary 60H35; 60H15, 65L60
The first author is supported by the National Natural Science Foundation of China (NNSFC), No. 12171047. The second author is supported by NNSFC, No. 12101296, Basic and Applied Basic Research Foundation of Guangdong Province, No. 2024A1515012348, and Shenzhen Basic Research Special Project (Natural Science Foundation) Basic Research (General Project), No. JCYJ20220530112814033.

1. Introduction

Many problems of solid state physics require the solution of the following semiclassical nonlinear Schrödinger (NLS) equation:

(1) 𝐢εtutε+ε22Δutε+F(utε)=0,t>0,xd,\displaystyle{\bf i}\varepsilon\partial_{t}u_{t}^{\varepsilon}+\frac{\varepsilon^{2}}{2}\Delta u_{t}^{\varepsilon}+F(u_{t}^{\varepsilon})=0,\quad t>0,\quad x\in\mathbb{R}^{d},

where 0<ε10<\varepsilon\ll 1 is the scaled Planck constant describing the microscopic and macroscopic scale ratio. Here the solution {utε:t0}\{u_{t}^{\varepsilon}:t\geq 0\} is the electron wave function. It is well known that Eq. (1) propagates oscillations with frequency 1/ε1/\varepsilon in space and time. In addition, these oscillations pose a huge challenge for the numerical computation. Recently, much progress has been made in this area, such as the time splitting spectral method [BJM02, BJM03, JZ13], the Gaussian beam method [JWY11, QY10], the Hagedorn wave packet approach [FGL09, Zho14], and the Frozen Gaussian beam method [Kay06, LZ18]. Besides, some efficient, conservative methods are particularly designed for Eq. (1), see, e.g., [CL22, CZ23]. We refer to the comprehensive review [JMS11, LL20] and references therein for more recent studies.

In Eq. (1), no exterior influence is considered. When it comes to quantum dynamics and studying wave propagation in random media, noise has to be introduced. For instance, in [CMZ20, JLRZ20, WL24], the semiclassical NLS equations with random initial data and potentials are investigated. In this work, we aim to deepen the understanding of the highly oscillatory behavior of a semiclassical NLS equation with a Wiener process perturbation. To our knowledge, this is the first result of investigating the stochastic semiclassical NLS equation.

Consider the following stochastic semiclassical NLS equation with multiplicative noise on the dd-dimensional torus 𝕋:=[0,1]d\mathbb{T}:=[0,1]^{d}:

(2) 𝐢εdutε+(ε22Δutε+F(utε))dt=G(utε)dWt,(t,x)(0,T]×𝕋,\displaystyle{\bf i}\varepsilon du_{t}^{\varepsilon}+(\frac{\varepsilon^{2}}{2}\Delta u_{t}^{\varepsilon}+F(u_{t}^{\varepsilon}))dt=G(u_{t}^{\varepsilon})dW_{t},\quad(t,x)\in(0,T]\times\mathbb{T},

with periodic boundary condition and initial datum u0εu_{0}^{\varepsilon}. Here WW is a L2(𝕋;)L^{2}(\mathbb{T};\mathbb{R})-valued QQ-Wiener process defined on a complete filtered probability space (see Section 2). Note that in the case of ε1\varepsilon\equiv 1, Eq. (2) becomes the classical stochastic NLS equation, which has been studied both theoretically and numerically, see, e.g., [DBD04, CHJ17, CHLZ17b, CHLZ19, CLZ23, CH16, CHL17, CHP16, CHLZ17a, HW19, Liu13] and references therein.

Designing efficient numerical methods that accurately approximate the solution of Eq. (2), or at least of the associated physical observables, is a formidable mathematical challenge. This paper investigates the spectral Galerkin approximation and its full discretization via a midpoint scheme for Eq. (2). It is well known that for the deterministic case, i.e., Eq. (1), [MPP99] proved that, for the best combination of the spatial and temporal discretizations, one needs the spatial mesh size O(ε)O(\varepsilon) and the time step size O(ε)O(\varepsilon) to guarantee good approximations to all (smooth) observables for sufficiently small ε\varepsilon. Failure to satisfy these conditions leads to wrong numerical observables. This paper attempts to understand the spectral Galerkin approximation’s resolution capacity and mesh strategies for the stochastic semiclassical NLS equation (2). By analyzing the strong convergence rates for the proposed numerical approximation, we obtain the admissible meshing strategies for obtaining “correct” physical observables in the semiclassical regime.

The rest of the paper is organized as follows. Section 2 introduces proper notations, assumptions, and preliminaries. In Section 3, we give the well-posedness and regularity estimates for the solution of Eq. (2). Rigorous strong error estimations for the aforementioned spectral Galerkin approximation and its temporal midpoint full discretization are performed in Sections 4 and 5, respectively.

2. Preliminaries

We first introduce some frequently used notations. Let T+=(0,)T\in\mathbb{R}^{*}_{+}=(0,\infty) be a fixed terminal time. For an integer M+={1,2,}M\in\mathbb{N}_{+}=\{1,2,\cdots\}, we denote by M:={0,1,,M}\mathbb{Z}_{M}:=\{0,1,\cdots,M\} and M={1,,M}\mathbb{Z}_{M}^{*}=\{1,\cdots,M\}. Throughout the paper, we use the notations Lx2:=L2(𝕋;)L_{x}^{2}:=L^{2}(\mathbb{T};\mathbb{C}), Ltp:=Lp([0,T];)L_{t}^{p}:=L^{p}([0,T];\mathbb{C}), and Lωp:=Lp(Ω;)L_{\omega}^{p}:=L^{p}(\Omega;\mathbb{C}), with p[2,]p\in[2,\infty], to denote the usual Lebesgue spaces over \mathbb{C}, and H:=L2(𝕋;)H:=L^{2}(\mathbb{T};\mathbb{R}). Denote by A:Dom(A)HHA:{\rm Dom}(A)\subset H\rightarrow H the periodic Laplacian operator on HH with an eigensystem {(λk,ek)}k=1\{(\lambda_{k},e_{k})\}_{k=1}^{\infty} where the sequence of eigenvalues {λk}k=1\{\lambda_{k}\}_{k=1}^{\infty} is listed in an increasing order. Let (A)θ(-A)^{\theta} be the fractional powers of A-A, and H˙θ\dot{H}^{\theta} be the domain of (A)θ/2(-A)^{\theta/2}, with θ+\theta\in\mathbb{R}_{+}. We use ((H;H˙θ),(H;H˙θ))(\mathcal{L}(H;\dot{H}^{\theta}),\|\cdot\|_{\mathcal{L}(H;\dot{H}^{\theta})}) to denote the space of bounded linear operators from HH to H˙θ\dot{H}^{\theta} for θ+\theta\in\mathbb{R}_{+} and (H):=(H;H)\mathcal{L}(H):=\mathcal{L}(H;H). For convenience, sometimes we use the temporally, sample path, and spatially mixed norm LωpLtrH˙θ\|\cdot\|_{L_{\omega}^{p}L_{t}^{r}\dot{H}^{\theta}} in different orders, such as

XLωpLtrH˙θ:=(Ω(0TX(t,ω)θrdt)pr(dω))1p\displaystyle\|X\|_{L_{\omega}^{p}L_{t}^{r}\dot{H}^{\theta}}:=(\int_{\Omega}(\int_{0}^{T}\|X(t,\omega)\|_{\theta}^{r}{\rm d}t)^{\frac{p}{r}}\mathbb{P}({\rm d}\omega))^{\frac{1}{p}}

for XLωpLtrH˙θX\in L_{\omega}^{p}L_{t}^{r}\dot{H}^{\theta}, with the usual modification for r=r=\infty or q=q=\infty.

Then, we introduce the driving stochastic process to study the stochastic semiclassical NLS equation (2). Let (Ω,,)(\Omega,\mathscr{F},\mathbb{P}) be a probability space with a normal filtration {t}0tT\{\mathscr{F}_{t}\}_{0\leq t\leq T}. Let 𝐐(H){\bf Q}\in\mathcal{L}(H) be a self-adjoint and nonnegative definite operator on HH. Denote by U0:=𝐐1/2(H)U_{0}:={\bf Q}^{1/2}(H) and (2θ:=HS(U0;H˙θ),2θ)(\mathcal{L}_{2}^{\theta}:=HS(U_{0};\dot{H}^{\theta}),\|\cdot\|_{\mathcal{L}_{2}^{\theta}}) the space of Hilbert–Schmidt operators from U0U_{0} to H˙θ\dot{H}^{\theta} for θ+\theta\in\mathbb{R}_{+}. Let {W(t):t[0,T]}\{W(t):t\in[0,T]\} be an HH-valued 𝐐{\bf Q}-Wiener process in the stochastic basis (Ω,,{t}0tT,)(\Omega,\mathscr{F},\{\mathscr{F}_{t}\}_{0\leq t\leq T},\mathbb{P}), i.e., there exists an orthonormal basis {gk}k=1\{g_{k}\}_{k=1}^{\infty} of HH which forms the eigenvectors of 𝐐\bf Q subject to the eigenvalues {qk}k=1\{q_{k}\}_{k=1}^{\infty} and a sequence of mutually independent Brownian motions {βk}k=1\{\beta_{k}\}_{k=1}^{\infty} such that (see [LR15, Proposition 2.1.10])

W(t)=k+𝐐1/2gkβk(t)=k+qkgkβk(t),0tT.\displaystyle W(t)=\sum_{k\in\mathbb{N}_{+}}{\bf Q}^{1/2}g_{k}\beta_{k}(t)=\sum_{k\in\mathbb{N}_{+}}\sqrt{q_{k}}g_{k}\beta_{k}(t),\quad 0\leq t\leq T.

We assume that the initial datum u0ϵu_{0}^{\epsilon} is an 0\mathscr{F}_{0}-measurable random variable. Our main assumptions on the coefficients of Eq. (2) are the following Lipschitz continuity and linear growth conditions.

Assumption 2.1.

There exist positive constants μ\mu, L1L_{1} and L2L_{2} such that

(3) F(u)F(v)L1uv,\displaystyle\|F(u)-F(v)\|\leq L_{1}\|u-v\|,\quad F(z)μL1(1+zμ),\displaystyle\|F(z)\|_{\mu}\leq L_{1}(1+\|z\|_{\mu}),
(4) G(u)G(v)20L2uv,\displaystyle\|G(u)-G(v)\|_{\mathcal{L}_{2}^{0}}\leq L_{2}\|u-v\|,\quad G(z)2μL2(1+zμ),\displaystyle\|G(z)\|_{\mathcal{L}_{2}^{\mu}}\leq L_{2}(1+\|z\|_{\mu}),

for all u,vHu,v\in H and zH˙μz\in\dot{H}^{\mu}.

Remark 2.1.

Let FF and GG denote the Nymiskii operators associated to some functions f:𝕋×f:\mathbb{T}\times\mathbb{C}\to\mathbb{C} and g:g:\mathbb{C}\to\mathbb{C}, respectively:

F(u)(x):\displaystyle F(u)(x): =f(x,u(x)),uLx2,x𝕋,\displaystyle=f(x,u(x)),\quad u\in L_{x}^{2},\ x\in\mathbb{T},
G(u)v(x):\displaystyle G(u)v(x): =g(u(x))v(x),uLx2,vU0,x𝕋.\displaystyle=g(u(x))v(x),\quad u\in L_{x}^{2},\ v\in U_{0},\ x\in\mathbb{T}.

A sufficient condition such that the condition (3) holds with μ=1\mu=1 is that ff possesses bounded partial derivatives. Indeed, suppose that |1f(x,y)|L1f|\partial_{1}f(x,y)|\leq L_{1}^{f} and |2f(x,y)|L2f|\partial_{2}f(x,y)|\leq L_{2}^{f} for some constants L1f,L2f0L_{1}^{f},L^{f}_{2}\geq 0, then for any u,vHu,v\in H and zH˙1z\in\dot{H}^{1},

F(u)F(v)2=𝕋|f(x,u(x))f(x,v(x))|2𝑑x(L2f)2uv2,\displaystyle\|F(u)-F(v)\|^{2}=\int_{\mathbb{T}}|f(x,u(x))-f(x,v(x))|^{2}dx\leq(L_{2}^{f})^{2}\|u-v\|^{2},

and

F(z)12\displaystyle\|F(z)\|_{1}^{2} =F(z)2+DF(z)2\displaystyle=\|F(z)\|^{2}+\|DF(z)\|^{2}
(F(z)F(0)+F(0))2+𝕋|1f(x,z(x))+2f(x,z(x))z(x)|2𝑑x\displaystyle\leq(\|F(z)-F(0)\|+\|F(0)\|)^{2}+\int_{\mathbb{T}}|\partial_{1}f(x,z(x))+\partial_{2}f(x,z(x))z^{\prime}(x)|^{2}dx
C(1+L1f)2+2(L2f)2z12,\displaystyle\leq C(1+L^{f}_{1})^{2}+2(L_{2}^{f})^{2}\|z\|_{1}^{2},

which imply the condition (3) with μ=1\mu=1 and certain L1>0L_{1}>0 depending on L1fL^{f}_{1} and L2fL^{f}_{2}. Here and in the rest of the paper, we use CC to denote a generic positive constant independent of the discrete parameters that would differ in each appearance.

A sufficient condition such that the condition (4) holds with μ=1\mu=1 is that gg is continuously differentiable with bounded derivative and gkW1,g_{k}\in W^{1,\infty} (i.e., gkg_{k} and its derivative are essentially bounded on 𝕋\mathbb{T}) for each k+k\in\mathbb{N}_{+} such that k+qkgkW1,2<\sum_{k\in\mathbb{N}_{+}}q_{k}\|g_{k}\|^{2}_{W^{1,\infty}}<\infty (see [LQ21, Example 5.1]). In fact, for any u,vHu,v\in H and zH˙1z\in\dot{H}^{1}, we have

G(u)G(v)202\displaystyle\|G(u)-G(v)\|^{2}_{\mathcal{L}_{2}^{0}} =k+qk(g(u)g(v))gk2g2k+qkgkLx2uv2,\displaystyle=\sum_{k\in\mathbb{N}_{+}}q_{k}\|(g(u)-g(v))g_{k}\|^{2}\leq\|g^{\prime}\|_{\infty}^{2}\sum_{k\in\mathbb{N}_{+}}q_{k}\|g_{k}\|_{L_{x}^{\infty}}^{2}\|u-v\|^{2},

and

G(z)212\displaystyle\|G(z)\|^{2}_{\mathcal{L}_{2}^{1}} Ck+qk(g(z)gk2+g(z)zgk+g(z)gk2)\displaystyle\leq C\sum_{k\in\mathbb{N}_{+}}q_{k}(\|g(z)g_{k}\|^{2}+\|g^{\prime}(z)z^{\prime}g_{k}+g(z)g^{\prime}_{k}\|^{2})
C(1+g2)k+qkgkWx1,2(1+z1)2,\displaystyle\leq C(1+\|g^{\prime}\|_{\infty}^{2})\sum_{k\in\mathbb{N}_{+}}q_{k}\|g_{k}\|_{W_{x}^{1,\infty}}^{2}(1+\|z\|_{1})^{2},

which show the condition (4) with μ=1\mu=1 and certain L2>0L_{2}>0 depending on g\|g^{\prime}\|_{\infty} and k+qkgkW1,2\sum_{k\in\mathbb{N}_{+}}q_{k}\|g_{k}\|^{2}_{W^{1,\infty}}.

Remark 2.2.

In the case of additive noise, e.g., for G=IdG={\rm Id}, the identity operator on Lx2L_{x}^{2}, the condition (4) is equivalent to

(5) L2:=(Δ)μ/2Q1/2HS<.\displaystyle L_{2}:=\|(-\Delta)^{\mu/2}Q^{1/2}\|_{HS}<\infty.

3. Well-posedness and Moment’s Estimates

Under Assumption 2.1, as the coefficients of Eq. (2) is Lipschitz continuous in HH and grow linearly in H˙μ\dot{H}^{\mu}, we have seen that its LωpH˙μL_{\omega}^{p}\dot{H}^{\mu}-well-posedness is straightforward (see [HL19, Theorem 2.3]). More precisely, Eq. (2) has a unique mild solution given by

(6) utε=Stεu0ε+𝐢ε10tStrεF(urε)𝑑r𝐢ε10tStrεG(urε)𝑑Wr,\displaystyle u_{t}^{\varepsilon}=S_{t}^{\varepsilon}u^{\varepsilon}_{0}+{\bf i}\varepsilon^{-1}\int_{0}^{t}S_{t-r}^{\varepsilon}F(u_{r}^{\varepsilon})dr-{\bf i}\varepsilon^{-1}\int_{0}^{t}S_{t-r}^{\varepsilon}G(u_{r}^{\varepsilon})dW_{r},

where Stε:=exp(𝐢ε2tΔ)S_{t}^{\varepsilon}:=\exp(\frac{{\bf i}\varepsilon}{2}t\Delta), t[0,T]t\in[0,T], denotes the semigroup generated by 𝐢εΔ/2{\bf i}\varepsilon\Delta/2.

In the following, we will present the moment’s estimates of the exact solution to Eq. (2) or Eq. (6), which would be valuable in the derivation of error estimations between the exact solution of Eq. (2) and the numerical ones in Sections 4 and 5. If there is a damping term, this moment’s estimate could be time-uniform.

Theorem 3.1.

Let p2p\geq 2, μ>0\mu>0, u0ϵLp(Ω;H˙μ)u_{0}^{\epsilon}\in L^{p}(\Omega;\dot{H}^{\mu}), and Assumption 2.1 hold. Then Eq. (2) possesses a unique solution {u(t):t0}\{u(t):t\geq 0\} and there exists a positive constant CC such that

(7) supt[0,T]utεLωpH˙μCeCTε2(1+u0εLωpH˙μ),\displaystyle\sup_{t\in[0,T]}\|u_{t}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}}\leq Ce^{CT\varepsilon^{-2}}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}),

and for any t,s[0,T]t,s\in[0,T],

(8) utεusεLωpLx2CeCTε2(1+u0εLωpH˙μ)|ts|1/2.\displaystyle\|u_{t}^{\varepsilon}-u_{s}^{\varepsilon}\|_{L_{\omega}^{p}L_{x}^{2}}\leq Ce^{CT\varepsilon^{-2}}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})|t-s|^{1/2}.

In particular if F(u)=𝐢αuF(u)={\bf i}\alpha u with α>0\alpha>0 and G=IdG={\rm Id} such that (5) holds, then there exists a positive constant CC such that for any t[0,T]t\in[0,T],

(9) utεLωpH˙μ\displaystyle\|u_{t}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}} eαtε1u0εLωpH˙μ+C(αε)1/2,\displaystyle\leq e^{-\alpha t\varepsilon^{-1}}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+C(\alpha\varepsilon)^{-1/2},

and for any t,s[0,T]t,s\in[0,T],

(10) utεusεLωpLx2C(eαtε1u0εLωpH˙μ+(αε)1/2)|ts|1/2.\displaystyle\|u_{t}^{\varepsilon}-u_{s}^{\varepsilon}\|_{L_{\omega}^{p}L_{x}^{2}}\leq C(e^{-\alpha t\varepsilon^{-1}}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+(\alpha\varepsilon)^{-1/2})|t-s|^{1/2}.
Proof.

By the Minkowski inequality and the Burkholder–Davis–Gundy inequality, we have

utεLωpH˙μStεu0εLωpH˙μ+ε10tStrεF(urε)LωpH˙μ𝑑r+ε1(0tStrεG(urε)LωpH˙μ2𝑑r)1/2.\begin{split}\|u_{t}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}}&\leq\|S_{t}^{\varepsilon}u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+\varepsilon^{-1}\int_{0}^{t}\|S_{t-r}^{\varepsilon}F(u_{r}^{\varepsilon})\|_{L_{\omega}^{p}\dot{H}^{\mu}}dr\\[4.2679pt] &\quad+\varepsilon^{-1}(\int_{0}^{t}\|S_{t-r}^{\varepsilon}G(u_{r}^{\varepsilon})\|^{2}_{L_{\omega}^{p}\dot{H}^{\mu}}dr)^{1/2}.\end{split}

Using the equality (32) in Appendix with μ0\mu\equiv 0 and Assumption 2.1, we obtain

utεLωpH˙μ\displaystyle\|u_{t}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}} u0εLωpH˙μ+L1ε10t(1+urεLωpH˙μ)𝑑r\displaystyle\leq\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+L_{1}\varepsilon^{-1}\int_{0}^{t}(1+\|u_{r}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})dr
+L2ε1(0t(1+urεLωpH˙μ)2𝑑r)1/2,\displaystyle\quad+L_{2}\varepsilon^{-1}(\int_{0}^{t}(1+\|u_{r}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})^{2}dr)^{1/2},

which shows that

utεLωpH˙μ2\displaystyle\|u_{t}^{\varepsilon}\|^{2}_{L_{\omega}^{p}\dot{H}^{\mu}} 3u0εLωpH˙μ2+3ε2(L12t+L22)0t(1+urεLωpH˙μ)2𝑑r\displaystyle\leq 3\|u^{\varepsilon}_{0}\|^{2}_{L_{\omega}^{p}\dot{H}^{\mu}}+3\varepsilon^{-2}(L^{2}_{1}t+L_{2}^{2})\int_{0}^{t}(1+\|u_{r}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})^{2}dr
CTε2u0εLωpH˙μ2+CTε20turεLωpH˙μ2𝑑r.\displaystyle\leq CT\varepsilon^{-2}\|u^{\varepsilon}_{0}\|^{2}_{L_{\omega}^{p}\dot{H}^{\mu}}+CT\varepsilon^{-2}\int_{0}^{t}\|u_{r}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}}^{2}dr.

We conclude (7) by the Gronwall inequality.

To show the Hölder continuous of the solution utεu_{t}^{\varepsilon}, we suppose that 0s<tT0\leq s<t\leq T without loss of generality. According to the mild formulation (6) and the Minkowski inequality, we get

utεusεLωpLx2I+II+III,\displaystyle\|u_{t}^{\varepsilon}-u_{s}^{\varepsilon}\|_{L_{\omega}^{p}L_{x}^{2}}\leq I+II+III,

where

I\displaystyle I :=(StsεId)usεLωpLx2,\displaystyle:=\|(S^{\varepsilon}_{t-s}-{\rm Id})u^{\varepsilon}_{s}\|_{L_{\omega}^{p}L_{x}^{2}},
II\displaystyle II :=ε1stStrεF(urε)𝑑rLωpLx2,\displaystyle:=\varepsilon^{-1}\|\int_{s}^{t}S^{\varepsilon}_{t-r}F(u^{\varepsilon}_{r})dr\|_{L_{\omega}^{p}L_{x}^{2}},
III\displaystyle III :=ε1stStrεG(urε)𝑑W(r)LωpLx2.\displaystyle:=\varepsilon^{-1}\|\int_{s}^{t}S^{\varepsilon}_{t-r}G(u^{\varepsilon}_{r})dW(r)\|_{L_{\omega}^{p}L_{x}^{2}}.

The smoothing property (33) in Appendix and the estimation (7) yield that

(11) I\displaystyle I Cεμ/2tμ/2eCtε2(1+u0εLωpH˙μ)(ts)μ/2.\displaystyle\leq C\varepsilon^{\mu/2}t^{\mu/2}e^{Ct\varepsilon^{-2}}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})(t-s)^{\mu/2}.

For the second term IIII, by the Minkowski inequality, the isometry property (32) with α=0\alpha=0, the condition (3), and the estimation (7), we have

II\displaystyle II stF(ur)LωpLx2𝑑r\displaystyle\leq\int_{s}^{t}\|F(u_{r})\|_{L_{\omega}^{p}L_{x}^{2}}dr
L1st(1+urLωpH˙μ)𝑑r\displaystyle\leq L_{1}\int_{s}^{t}(1+\|u_{r}\|_{L_{\omega}^{p}\dot{H}^{\mu}})dr
(12) CeCTε2(1+u0εLωpH˙μ)(ts).\displaystyle\leq Ce^{CT\varepsilon^{-2}}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})(t-s).

For the last term IIIIII, by the Burkholder–Davis–Gundy inequality, the isometry property (32) with α=0\alpha=0, the condition (4), and the estimation (7),

(13) III\displaystyle III (stG(ur)Lωp202dr)12CeCTε2(1+u0εLωpH˙μ)(ts)12.\displaystyle\leq(\int_{s}^{t}\|G(u_{r})\|^{2}_{L_{\omega}^{p}\mathcal{L}_{2}^{0}}{\rm d}r)^{\frac{1}{2}}\leq Ce^{CT\varepsilon^{-2}}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})(t-s)^{\frac{1}{2}}.

Combining the above estimations (11), (3), and (13), we conclude (8).

Now assume that F(u)=𝐢αuF(u)={\bf i}\alpha u and G=IdG={\rm Id}. Due to (32) and the condition (5), we obtain

utεLωpH˙μ\displaystyle\|u_{t}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}} eαtε1u0εLωpH˙μ+Cε1(0te2αr/εSr2μ2𝑑r)1/2\displaystyle\leq e^{-\alpha t\varepsilon^{-1}}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+C\varepsilon^{-1}(\int_{0}^{t}e^{-2\alpha r/\varepsilon}\|S_{r}\|^{2}_{\mathcal{L}_{2}^{\mu}}dr)^{1/2}
=eαtε1u0εLωpH˙μ+CL2ε1(0te2αr/ε𝑑r)1/2\displaystyle=e^{-\alpha t\varepsilon^{-1}}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+CL_{2}\varepsilon^{-1}(\int_{0}^{t}e^{-2\alpha r/\varepsilon}dr)^{1/2}
=eαtε1u0εLωpH˙μ+C(αε)1/2.\displaystyle=e^{-\alpha t\varepsilon^{-1}}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+C(\alpha\varepsilon)^{-1/2}.

This shows (9), and we can conclude (10) by the previous argument. The proof is complete. ∎

4. Spatial Semi-discretization

In the previous section, we derive the uniform boundedness and the Hölder continuous of the solution to Eq. (2). In this section, we apply the spectral Galerkin method to discretize Eq. (2) spatially to obtain a valid approximation.

In the sequel, let N+N\in\mathbb{N}_{+} and define the finite dimensional subspace VN:=span{e1,e2,,eN}V_{N}:={\rm span}\{e_{1},e_{2},\cdots,e_{N}\}. Then the corresponding numerical solution {uN:=uε,N}\{u^{N}:=u^{\varepsilon,N}\}, NN\in\mathbb{N}, of Eq. (2) satisfies

(14) 𝐢εdutε,N+(ε22Δutε,N+PNF(utε,N))dt=PNG(utε,N)dWt,t(0,T],u0ε,N=PNu0ε.\displaystyle\begin{split}&{\bf i}\varepsilon du_{t}^{\varepsilon,N}+(\frac{\varepsilon^{2}}{2}\Delta u_{t}^{\varepsilon,N}+P_{N}F(u_{t}^{\varepsilon,N}))dt=P_{N}G(u_{t}^{\varepsilon,N})dW_{t},\quad t\in(0,T],\\[4.2679pt] &u^{\varepsilon,N}_{0}=P_{N}u^{\varepsilon}_{0}.\end{split}

Similarly to Eq. (6), we have

(15) utε,N=StNu0ε+𝐢ε10tStrε,NF(urε,N)𝑑r𝐢ε10tStrε,NG(urε,N)𝑑Wr,\displaystyle u_{t}^{\varepsilon,N}=S_{t}^{N}u^{\varepsilon}_{0}+{\bf i}\varepsilon^{-1}\int_{0}^{t}S_{t-r}^{\varepsilon,N}F(u_{r}^{\varepsilon,N})dr-{\bf i}\varepsilon^{-1}\int_{0}^{t}S_{t-r}^{\varepsilon,N}G(u_{r}^{\varepsilon,N})dW_{r},

where StN:=PNStε=exp(𝐢ε2tΔN)S_{t}^{N}:=P_{N}S_{t}^{\varepsilon}=\exp(\frac{{\bf i}\varepsilon}{2}t\Delta_{N}), t[0,T]t\in[0,T].

Denote by etN:=utεutε,Ne^{N}_{t}:=u_{t}^{\varepsilon}-u_{t}^{\varepsilon,N}, t[0,T]t\in[0,T]. Then subtracting Eq. (15) from Eq. (6) yields that

etN=(StεStN)u0ε+𝐢ε10t[StrεF(urε)StrεPNF(urε,N)]𝑑r𝐢ε10t[StrεG(urε)StrεPNG(urε,N)]𝑑Wr:=IN+IIN+IIIN.\displaystyle\begin{split}e^{N}_{t}=&(S_{t}^{\varepsilon}-S_{t}^{N})u^{\varepsilon}_{0}+{\bf i}\varepsilon^{-1}\int_{0}^{t}[S_{t-r}^{\varepsilon}F(u_{r}^{\varepsilon})-S_{t-r}^{\varepsilon}P_{N}F(u_{r}^{\varepsilon,N})]dr\\[2.84526pt] &-{\bf i}\varepsilon^{-1}\int_{0}^{t}[S_{t-r}^{\varepsilon}G(u_{r}^{\varepsilon})-S_{t-r}^{\varepsilon}P_{N}G(u^{\varepsilon,N}_{r})]dW_{r}\\[4.2679pt] &:=I^{N}+II^{N}+III^{N}.\end{split}

For any t[0,T]t\in[0,T] and α>0\alpha>0, we define the following three operators:

Stα,ε:=\displaystyle S_{t}^{\alpha,\varepsilon}:= exp((𝐢ε2ΔαεId)t)=eα/εtStε,\displaystyle\exp((\frac{{\bf i}\varepsilon}{2}\Delta-\frac{\alpha}{\varepsilon}{\rm Id})t)=e^{-\alpha/\varepsilon t}S_{t}^{\varepsilon},
Stα,ε,N:=\displaystyle S_{t}^{\alpha,\varepsilon,N}:= PNStα,ε=exp((𝐢ε2ΔNαεId)t),\displaystyle P_{N}S_{t}^{\alpha,\varepsilon}=\exp((\frac{{\bf i}\varepsilon}{2}\Delta_{N}-\frac{\alpha}{\varepsilon}{\rm Id})t),
etα,N:=\displaystyle e^{\alpha,N}_{t}:= utα,εutα,ε,N.\displaystyle u_{t}^{\alpha,\varepsilon}-u_{t}^{\alpha,\varepsilon,N}.

Using the semigroup theory, we can derive the following error estimate between the spectral Galerkin approximation (14) and the exact solution of Eq. (2).

Theorem 4.1.

Let p2p\geq 2, μ>0\mu>0, u0ϵLp(Ω;H˙μ)u_{0}^{\epsilon}\in L^{p}(\Omega;\dot{H}^{\mu}), and Assumption 2.1 hold. Then there exists a positive constant CC such that

(16) sup0tTetNLωpLx2CeCTε2(1+u0εLωpH˙μ)Nμ/d.\displaystyle\sup_{0\leq t\leq T}\|e^{N}_{t}\|_{L_{\omega}^{p}L_{x}^{2}}\leq Ce^{CT\varepsilon^{-2}}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})N^{-\mu/d}.

Particularly, if F(u)=𝐢αuF(u)={\bf i}\alpha u with α>0\alpha>0 and G=IdG={\rm Id} such that (5) holds, then

(17) etNLωpLx2(eαtε1u0εLωpH˙μ+C(αε)1/2)Nμ/d.\displaystyle\|e^{N}_{t}\|_{L_{\omega}^{p}L_{x}^{2}}\leq(e^{-\alpha t\varepsilon^{-1}}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+C(\alpha\varepsilon)^{-1/2})N^{-\mu/d}.
Proof.

For the first term INI^{N}, we use the inequality (36) in Appendix to get

(StεStε,N)u0ε2λNμu0εμ2,\displaystyle\|(S_{t}^{\varepsilon}-S_{t}^{\varepsilon,N})u^{\varepsilon}_{0}\|^{2}\leq\lambda_{N}^{-\mu}\|u^{\varepsilon}_{0}\|_{\mu}^{2},

which implies

(18) INLωpLx2CλNμ/2u0εLωpH˙μ.\displaystyle\|I^{N}\|_{L_{\omega}^{p}L_{x}^{2}}\leq C\lambda_{N}^{-\mu/2}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}.

For the second term IINII^{N}, by the Minkovski inequality, the estimate (36), the boundedness of StεPNS_{t}^{\varepsilon}P_{N}, and the condition (3), we have

(19) IINLωpLx2ε10t(StrεStrε,N)F(urε)LωpLx2𝑑r+ε10tStrε,N(F(urε)F(urε,N))LωpLx2𝑑rCε1λNμ/20tF(urε)Lp(Ω;H˙μ)𝑑r+Cε10tF(urε)F(urε,N)LωpLx2𝑑rCε1λNμ/2(1+u0εLωpH˙μ)+Cε10terNLωpLx2𝑑r.\begin{split}\|II^{N}\|_{L_{\omega}^{p}L_{x}^{2}}&\leq\varepsilon^{-1}\int_{0}^{t}\|(S_{t-r}^{\varepsilon}-S_{t-r}^{\varepsilon,N})F(u_{r}^{\varepsilon})\|_{L_{\omega}^{p}L_{x}^{2}}dr\\[4.2679pt] &\quad+\varepsilon^{-1}\int_{0}^{t}\|S_{t-r}^{\varepsilon,N}(F(u_{r}^{\varepsilon})-F(u_{r}^{\varepsilon,N}))\|_{L_{\omega}^{p}L_{x}^{2}}dr\\[4.2679pt] &\leq C\varepsilon^{-1}\lambda_{N}^{-\mu/2}\int_{0}^{t}\|F(u_{r}^{\varepsilon})\|_{L^{p}(\Omega;\dot{H}^{\mu})}dr\\[4.2679pt] &\quad+C\varepsilon^{-1}\int_{0}^{t}\|F(u_{r}^{\varepsilon})-F(u_{r}^{\varepsilon,N})\|_{L_{\omega}^{p}L_{x}^{2}}dr\\[4.2679pt] &\leq C\varepsilon^{-1}\lambda_{N}^{-\mu/2}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})+C\varepsilon^{-1}\int_{0}^{t}\|e^{N}_{r}\|_{L_{\omega}^{p}L_{x}^{2}}dr.\end{split}

For the last term IIINIII^{N}, it follows from the Minkovski inequality, the Burkholder–Davis–Gundy inequality, and the condition (3) that

(20) IIINLωpLx2ε10t(StrεStrεPN)G(urε)𝑑WrLωpLx2+ε10tStrεPNG(urε)G(urε,N))dWrLωpLx2Cε1λNμ/2(0t(1+urεLp(Ω;H˙μ)2)𝑑r)1/2+Cε1(0terNLωpLx22𝑑r)1/2Cε1λNμ/2(1+u0εLωpH˙μ)+Cε1(0terNLωpLx22𝑑r)1/2.\begin{split}\|III^{N}\|_{L_{\omega}^{p}L_{x}^{2}}&\leq\varepsilon^{-1}\left\|\int_{0}^{t}(S_{t-r}^{\varepsilon}-S_{t-r}^{\varepsilon}P_{N})G(u_{r}^{\varepsilon})dW_{r}\right\|_{L_{\omega}^{p}L_{x}^{2}}\\[4.2679pt] &\quad+\varepsilon^{-1}\left\|\int_{0}^{t}S_{t-r}^{\varepsilon}P_{N}G(u_{r}^{\varepsilon})-G(u_{r}^{\varepsilon,N}))dW_{r}\right\|_{L_{\omega}^{p}L_{x}^{2}}\\[4.2679pt] &\leq C\varepsilon^{-1}\lambda_{N}^{-\mu/2}(\int_{0}^{t}(1+\|u_{r}^{\varepsilon}\|^{2}_{L^{p}(\Omega;\dot{H}^{\mu})})dr)^{1/2}\\[4.2679pt] &\quad+C\varepsilon^{-1}(\int_{0}^{t}\|e^{N}_{r}\|^{2}_{L_{\omega}^{p}L_{x}^{2}}dr)^{1/2}\\[4.2679pt] &\leq C\varepsilon^{-1}\lambda_{N}^{-\mu/2}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})+C\varepsilon^{-1}(\int_{0}^{t}\|e^{N}_{r}\|^{2}_{L_{\omega}^{p}L_{x}^{2}}dr)^{1/2}.\end{split}

Combing (18), (19), and (20), we obtain

etNLωpLx2\displaystyle\|e^{N}_{t}\|_{L_{\omega}^{p}L_{x}^{2}} Cε1λNμ/2(1+u0εLωpH˙μ)+Cε10terNLωpLx2𝑑r\displaystyle\leq C\varepsilon^{-1}\lambda_{N}^{-\mu/2}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}})+C\varepsilon^{-1}\int_{0}^{t}\|e^{N}_{r}\|_{L_{\omega}^{p}L_{x}^{2}}dr
+Cε1(0terNLωpLx22𝑑r)1/2,\displaystyle\quad+C\varepsilon^{-1}(\int_{0}^{t}\|e^{N}_{r}\|^{2}_{L_{\omega}^{p}L_{x}^{2}}dr)^{1/2},

which implies that

etNLωpLx22Cε2λNμ(1+u0εLωpH˙μ2)+Cε2(1+t)0terNLωpLx22𝑑r.\displaystyle\begin{split}\|e^{N}_{t}\|^{2}_{L_{\omega}^{p}L_{x}^{2}}&\leq C\varepsilon^{-2}\lambda_{N}^{-\mu}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}^{2})+C\varepsilon^{-2}(1+t)\int_{0}^{t}\|e^{N}_{r}\|^{2}_{L_{\omega}^{p}L_{x}^{2}}dr.\end{split}

Then, the Gronwall inequality, combined with the Weyl law, yields the assertion (16).

Now assume that F(u)=𝐢αuF(u)={\bf i}\alpha u and G=IdG={\rm Id}. Then

etN=\displaystyle e^{N}_{t}= (Stε,αStε,α,N)u0ε𝐢ε10t(Strε,αStrε,α,N)𝑑Wr.\displaystyle(S_{t}^{\varepsilon,\alpha}-S_{t}^{\varepsilon,\alpha,N})u^{\varepsilon}_{0}-{\bf i}\varepsilon^{-1}\int_{0}^{t}(S_{t-r}^{\varepsilon,\alpha}-S_{t-r}^{\varepsilon,\alpha,N})dW_{r}.

By the inequality (36) and the condition (5),

etNLωpLx2\displaystyle\|e^{N}_{t}\|_{L_{\omega}^{p}L_{x}^{2}} (Stε,αStε,α,N)u0εLωpLx2+ε10t(Strε,αStrε,α,N)𝑑WrLωpLx2\displaystyle\leq\|(S_{t}^{\varepsilon,\alpha}-S_{t}^{\varepsilon,\alpha,N})u^{\varepsilon}_{0}\|_{L_{\omega}^{p}L_{x}^{2}}+\varepsilon^{-1}\|\int_{0}^{t}(S_{t-r}^{\varepsilon,\alpha}-S_{t-r}^{\varepsilon,\alpha,N})dW_{r}\|_{L_{\omega}^{p}L_{x}^{2}}
eαtε1λNμ/2u0εLωpH˙μ+Cε1(0tSrε,αSrε,α,N2μ2𝑑r)1/2\displaystyle\leq e^{-\alpha t\varepsilon^{-1}}\lambda_{N}^{-\mu/2}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+C\varepsilon^{-1}(\int_{0}^{t}\|S_{r}^{\varepsilon,\alpha}-S_{r}^{\varepsilon,\alpha,N}\|^{2}_{\mathcal{L}_{2}^{\mu}}dr)^{1/2}
eαtε1λNμ/2u0εLωpH˙μ+Cε1λNμ/2(0te2αr/ε𝑑r)1/2,\displaystyle\leq e^{-\alpha t\varepsilon^{-1}}\lambda_{N}^{-\mu/2}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+C\varepsilon^{-1}\lambda_{N}^{-\mu/2}(\int_{0}^{t}e^{-2\alpha r/\varepsilon}dr)^{1/2},

which shows (17). The completes the proof. ∎

5. Spatio-temporal Full Discretization

In this section, we investigate the full discretization of Eq. (2), spatially by the spectral Galerkin method in Section 3 and temporally by the midpoint scheme. Let T<M+T<M\in\mathbb{N}_{+} and τ=T/M(0,1)\tau=T/M\in(0,1) be the temporal step size. Then, the midpoint-spectral Galerkin approximation of Eq. (2) is to find

(21) um+1ε,N=umε,N+(𝐢ε2ΔNum+1/2ε,N+𝐢ε1PNF(um+1/2ε,N))τ𝐢ε1PNG(umε,N)δmW,\begin{split}u_{m+1}^{\varepsilon,N}=u_{m}^{\varepsilon,N}+&(\frac{{\bf i}\varepsilon}{2}\Delta_{N}u_{m+1/2}^{\varepsilon,N}+{\bf i}\varepsilon^{-1}P_{N}F(u_{m+1/2}^{\varepsilon,N}))\tau-{\bf i}\varepsilon^{-1}P_{N}G(u_{m}^{\varepsilon,N})\delta_{m}W,\end{split}

where um+1/2:=(um+um+1)/2u_{m+1/2}:=(u_{m}+u_{m+1})/2 and δmW:=W(tm+1)W(tm)\delta_{m}W:=W(t_{m+1})-W(t_{m}), mMm\in\mathbb{Z}_{M}. It is straightforward to see that

(I𝐢ε4ΔNτ)um+1ε,N=(I+𝐢ε4ΔNτ)umε,N+𝐢ε1PNF(um+1/2ε,N)τ𝐢ε1PNG(umε,N)δmW.\begin{split}(I-\frac{{\bf i}\varepsilon}{4}\Delta_{N}\tau)u_{m+1}^{\varepsilon,N}=&(I+\frac{{\bf i}\varepsilon}{4}\Delta_{N}\tau)u_{m}^{\varepsilon,N}+{\bf i}\varepsilon^{-1}P_{N}F(u_{m+1/2}^{\varepsilon,N})\tau\\[4.2679pt] &-{\bf i}\varepsilon^{-1}P_{N}G(u_{m}^{\varepsilon,N})\delta_{m}W.\end{split}

Denote by

Sτ:=Sτε,N:=(I𝐢ε4ΔNτ)1(I+𝐢ε4ΔNτ),Tτ:=Tτε,N:=(I𝐢ε4ΔNτ)1.S_{\tau}:=S^{\varepsilon,N}_{\tau}:=(I-\frac{{\bf i}\varepsilon}{4}\Delta_{N}\tau)^{-1}(I+\frac{{\bf i}\varepsilon}{4}\Delta_{N}\tau),\quad T_{\tau}:=T^{\varepsilon,N}_{\tau}:=(I-\frac{{\bf i}\varepsilon}{4}\Delta_{N}\tau)^{-1}.

Then we have the full discrete mild formulation

(22) umε,N=SτmPNu0ε+𝐢ε1j=0m1SτmjTτPNF(uj+1/2ε,N)τ𝐢ε1j=0m1SτmjTτPNG(ujε,N)δjW.\begin{split}u_{m}^{\varepsilon,N}=S_{\tau}^{m}P_{N}u_{0}^{\varepsilon}&+{\bf i}\varepsilon^{-1}\sum_{j=0}^{m-1}S_{\tau}^{m-j}T_{\tau}P_{N}F(u_{j+1/2}^{\varepsilon,N})\tau\\ &-{\bf i}\varepsilon^{-1}\sum_{j=0}^{m-1}S_{\tau}^{m-j}T_{\tau}P_{N}G(u_{j}^{\varepsilon,N})\delta_{j}W.\end{split}

It is not difficult to show the well-posedness and moment’s estimate of the midpoint-spectral Galerkin approximation (21) under Assumption 2.1.

Lemma 5.1.

Let p2p\geq 2, μ>0\mu>0, u0εLp(Ω;H˙μ)u^{\varepsilon}_{0}\in L^{p}(\Omega;\dot{H}^{\mu}), and Assumption 2.1 hold. Then there exists τ0(0,1)\tau_{0}\in(0,1) such that for any τ(0,τ0)\tau\in(0,\tau_{0}), the midpoint-spectral Galerkin approximation (21) is uniquely solved and there exists a positive constant CC such that

(23) supmMumε,NLωpH˙μCeCTε2(1+u0εLωpH˙μ).\displaystyle\sup_{m\in\mathbb{Z}_{M}}\|u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}\dot{H}^{\mu}}\leq Ce^{CT\varepsilon^{-2}}(1+\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}).

In particular, if F(u)=𝐢αuF(u)={\bf i}\alpha u with α>0\alpha>0 and G=IdG={\rm Id} such that (5) holds, then for any mMm\in\mathbb{Z}_{M},

(24) umε,NLωpH˙μ\displaystyle\|u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}\dot{H}^{\mu}} eαtmε1u0εLωpH˙μ+C(αε)1/2.\displaystyle\leq e^{-\alpha t_{m}\varepsilon^{-1}}\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+C(\alpha\varepsilon)^{-1/2}.

The proof of Lemma 5.1 is similar to the same as in the proof of Theorem 3.1 and thus is omitted here. Now, we are in the position to show the approximate error between uε,N(tm)u^{\varepsilon,N}(t_{m}) and umε,Nu_{m}^{\varepsilon,N}.

Theorem 5.1.

Let p2p\geq 2, μ>0\mu>0, u0εLp(Ω;H˙μ)u^{\varepsilon}_{0}\in L^{p}(\Omega;\dot{H}^{\mu}), and Assumption 2.1 hold. Then there exists a positive constant CC such that

(25) supmMuε,N(tm)umε,NLωpLx2CeCTε2(1+u0εLωpH˙μ)τ(μ/31/2)1/2.\displaystyle\sup_{m\in\mathbb{Z}_{M}}\|u^{\varepsilon,N}(t_{m})-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\leq Ce^{CT\varepsilon^{-2}}(1+\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})\tau^{(\mu/3-1/2)\wedge 1/2}.

Moreover, if F(u)=𝐢αuF(u)={\bf i}\alpha u with α>0\alpha>0 and G=IdG={\rm Id} such that (5) holds, then

(26) supmuε,N(tm)umε,NLωpLx2C(u0εLωpH˙μ+(αε)1/2)τ(μ/31/2)1/2.\displaystyle\sup_{m\in\mathbb{N}}\|u^{\varepsilon,N}(t_{m})-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\leq C(\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+(\alpha\varepsilon)^{-1/2})\tau^{(\mu/3-1/2)\wedge 1/2}.
Proof.

Note that subtracting Eq. (22) from Eq. (15) with t=tmt=t_{m} implies that

uε,N(tm)umε,N\displaystyle u^{\varepsilon,N}(t_{m})-u_{m}^{\varepsilon,N} =(StmSτm)PNu0ε\displaystyle=(S_{t_{m}}-S_{\tau}^{m})P_{N}u_{0}^{\varepsilon}
+𝐢ε1j=0m1tjtj+1[StmrPNF(ur)SτmjTτPNF(uj+1/2ε,N)]𝑑r\displaystyle\quad+{\bf i}\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}[S_{t_{m}-r}P_{N}F(u_{r})-S_{\tau}^{m-j}T_{\tau}P_{N}F(u_{j+1/2}^{\varepsilon,N})]dr
𝐢ε1j=0m1tjtj+1[StmrPNG(ur)SτmjTτPNG(ujε,N)]𝑑Wr\displaystyle\quad-{\bf i}\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}[S_{t_{m}-r}P_{N}G(u_{r})-S_{\tau}^{m-j}T_{\tau}P_{N}G(u_{j}^{\varepsilon,N})]dW_{r}
:=Iτ+IIτ+IIIτ.\displaystyle:=I_{\tau}+II_{\tau}+III_{\tau}.

For the first term IτI_{\tau}, we use the inequality (37) in Appendix to derive

(27) IτLωpLx2\displaystyle\|I_{\tau}\|_{L_{\omega}^{p}L_{x}^{2}} Ctmμ/6εμ/2τμ/3u0εLωpH˙μ.\displaystyle\leq Ct_{m}^{\mu/6}\varepsilon^{\mu/2}\tau^{\mu/3}\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}}.

For the second term IIτII_{\tau}, using the Minkovski inequality, the inequality (39) with α=0\alpha=0, and the condition (3), we obtain

IIτLωpLx2\displaystyle\|II_{\tau}\|_{L_{\omega}^{p}L_{x}^{2}} ε1j=0m1tjtj+1(StmrSτmjTτ)PNF(uj+1/2ε,N)LωpLx2𝑑r\displaystyle\leq\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|(S_{t_{m}-r}-S_{\tau}^{m-j}T_{\tau})P_{N}F(u_{j+1/2}^{\varepsilon,N})\|_{L_{\omega}^{p}L_{x}^{2}}dr
+ε1j=0m1tjtj+1(StmrPN[F(ur)F(uj+1/2ε,N)]LωpLx2dr\displaystyle\quad+\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|(S_{t_{m}-r}P_{N}[F(u_{r})-F(u_{j+1/2}^{\varepsilon,N})]\|_{L_{\omega}^{p}L_{x}^{2}}dr
ε1j=0m1tjtj+1(StmrSτmjTτ)PN(H˙μ,L2)F(uj+1/2ε,N)LωpH˙μ𝑑r\displaystyle\leq\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|(S_{t_{m}-r}-S_{\tau}^{m-j}T_{\tau})P_{N}\|_{\mathcal{L}(\dot{H}^{\mu},L^{2})}\|F(u_{j+1/2}^{\varepsilon,N})\|_{L_{\omega}^{p}\dot{H}^{\mu}}dr
+ε1j=0m1tjtj+1F(ur)F(uj+1/2ε,N)LωpLx2𝑑r\displaystyle\quad+\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|F(u_{r})-F(u_{j+1/2}^{\varepsilon,N})\|_{L_{\omega}^{p}L_{x}^{2}}dr
Ctmβ(1+ujε,NLωpH˙μ)τ(μ/3)1+ε1j=0m1tjtj+1uruj+1/2ε,NLωpLx2𝑑r.\displaystyle\leq Ct_{m}^{\beta}(1+\|u_{j}^{\varepsilon,N}\|_{L_{\omega}^{p}\dot{H}^{\mu}})\tau^{(\mu/3)\wedge 1}+\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|u_{r}-u_{j+1/2}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}dr.

By the Hölder continuous of utεu_{t}^{\varepsilon} in (8), we get

ε1j=0m1tjtj+1uruj+1/2ε,NLωpLx2𝑑r\displaystyle\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|u_{r}-u_{j+1/2}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}dr
Cε1j=0m1tjtj+1ur(utj+utj+1)/2LωpLx2𝑑r\displaystyle\leq C\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|u_{r}-(u_{t_{j}}+u_{t_{j+1}})/2\|_{L_{\omega}^{p}L_{x}^{2}}dr
+Cε1j=0m1utjujε,N2+utj+1uj+1ε,N2LωpLx2τ\displaystyle\quad+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|\frac{u_{t_{j}}-u_{j}^{\varepsilon,N}}{2}+\frac{u_{t_{j+1}}-u_{j+1}^{\varepsilon,N}}{2}\|_{L_{\omega}^{p}L_{x}^{2}}\tau
Cε1j=0m1tjtj+1ur(utj+utj+1)/2LωpLx2𝑑r\displaystyle\leq C\varepsilon^{-1}\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|u_{r}-(u_{t_{j}}+u_{t_{j+1}})/2\|_{L_{\omega}^{p}L_{x}^{2}}dr
+Cε1j=0m1utjujε,NLωpLx2τ+Cε1j=0m1utj+1uj+1ε,NLωpLx2τ\displaystyle\quad+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|u_{t_{j}}-u_{j}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|u_{t_{j+1}}-u_{j+1}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau
Cε1eCtmε2(1+u0εLωpH˙μ)τ1/2\displaystyle\leq C\varepsilon^{-1}e^{Ct_{m}\varepsilon^{-2}}(1+\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})\tau^{1/2}
+Cε1j=0m1utjujε,NLωpLx2τ+Cε1utmumε,NLωpLx2τ,\displaystyle\quad+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|u_{t_{j}}-u_{j}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau+C\varepsilon^{-1}\|u_{t_{m}}-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau,

which by combining with (23), yields that

(28) IIτLωpLx2CeCtmε2(1+u0εLωpH˙μ)τ1/2+Cε1j=0m1utjujε,NLωpLx2τ+Cε1utmumε,NLωpLx2τ.\begin{split}\|II_{\tau}\|_{L_{\omega}^{p}L_{x}^{2}}&\leq Ce^{Ct_{m}\varepsilon^{-2}}(1+\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})\tau^{1/2}\\[4.2679pt] &\quad+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|u_{t_{j}}-u_{j}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau+C\varepsilon^{-1}\|u_{t_{m}}-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau.\end{split}

Similarly, using the Minkovski inequality, the discrete Burkholder–Davis–Gundy inequality, the inequality (39) with α=0\alpha=0, and the condition (4), we have

(29) IIIτLωpLx2ε1(j=0m1tjtj+1(StmrSτmjTτ)PNG(ujε,N)Lωp202𝑑r)12+ε1(j=0m1tjtj+1StmrPN[G(ur)G(ujε,N)]Lωp202𝑑r)12ε1j=0m1(tjtj+1(StmrSτmjTτ)PN(H˙μ,L2)2G(uj+1/2ε,N)Lωp2μ2𝑑r)1/2+ε1(j=0m1tjtj+1G(ur)G(uj+1/2ε,N)Lωp202𝑑r)1/2CeCtmε2(1+u0εLωp2μ)τ(μ/31/2)1/2+Cε1j=0m1utjujε,NLωpLx2τ+Cε1utmumε,NLωpLx2τ.\begin{split}&\|III_{\tau}\|_{L_{\omega}^{p}L_{x}^{2}}\\[4.2679pt] &\leq\varepsilon^{-1}(\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|(S_{t_{m}-r}-S_{\tau}^{m-j}T_{\tau})P_{N}G(u_{j}^{\varepsilon,N})\|^{2}_{L_{\omega}^{p}\mathcal{L}_{2}^{0}}dr)^{\frac{1}{2}}\\[4.2679pt] &\quad+\varepsilon^{-1}(\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|S_{t_{m}-r}P_{N}[G(u_{r})-G(u_{j}^{\varepsilon,N})]\|^{2}_{L_{\omega}^{p}\mathcal{L}_{2}^{0}}dr)^{\frac{1}{2}}\\[4.2679pt] &\leq\varepsilon^{-1}\sum_{j=0}^{m-1}(\int_{t_{j}}^{t_{j+1}}\|(S_{t_{m}-r}-S_{\tau}^{m-j}T_{\tau})P_{N}\|^{2}_{\mathcal{L}(\dot{H}^{\mu},L^{2})}\|G(u_{j+1/2}^{\varepsilon,N})\|^{2}_{L_{\omega}^{p}\mathcal{L}_{2}^{\mu}}dr)^{1/2}\\[4.2679pt] &\quad+\varepsilon^{-1}(\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|G(u_{r})-G(u_{j+1/2}^{\varepsilon,N})\|^{2}_{L_{\omega}^{p}\mathcal{L}_{2}^{0}}dr)^{1/2}\\[4.2679pt] &\leq Ce^{Ct_{m}\varepsilon^{-2}}(1+\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\mathcal{L}_{2}^{\mu}})\tau^{(\mu/3-1/2)\wedge 1/2}\\[4.2679pt] &\quad+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|u_{t_{j}}-u_{j}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau+C\varepsilon^{-1}\|u_{t_{m}}-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau.\end{split}

Combing (27), (28) and (29), we find

uε,N(tm)umε,NLωpLx2\displaystyle\|u^{\varepsilon,N}(t_{m})-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}} CeCtmε2(1+u0εLωpH˙μ)τ(μ/31/2)1/2\displaystyle\leq Ce^{Ct_{m}\varepsilon^{-2}}(1+\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})\tau^{(\mu/3-1/2)\wedge 1/2}
+Cε1j=0m1utjujε,NLωpLx2τ\displaystyle\quad+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|u_{t_{j}}-u_{j}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau
+Cε1(j=0m1utjujε,NLωpLx22τ)12\displaystyle\quad+C\varepsilon^{-1}(\sum_{j=0}^{m-1}\|u_{t_{j}}-u_{j}^{\varepsilon,N}\|^{2}_{L_{\omega}^{p}L_{x}^{2}}\tau)^{\frac{1}{2}}
+Cε1utmumε,NLωpLx2τ.\displaystyle\quad+C\varepsilon^{-1}\|u_{t_{m}}-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\tau.

Therefore, for τ<Cε1\tau<C\varepsilon^{-1} which always holds for sufficiently small ε\varepsilon, we get

uε,N(tm)umε,NLωpLx22\displaystyle\|u^{\varepsilon,N}(t_{m})-u_{m}^{\varepsilon,N}\|^{2}_{L_{\omega}^{p}L_{x}^{2}} CeCtmε2(1+u0εLωpH˙μ)2τ2[(μ/31/2)1/2]\displaystyle\leq Ce^{Ct_{m}\varepsilon^{-2}}(1+\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})^{2}\tau^{2[(\mu/3-1/2)\wedge 1/2]}
+Cε1j=0m1utjujε,NLωpLx22τ.\displaystyle\quad+C\varepsilon^{-1}\sum_{j=0}^{m-1}\|u_{t_{j}}-u_{j}^{\varepsilon,N}\|^{2}_{L_{\omega}^{p}L_{x}^{2}}\tau.

The desired result (25) is then obtained by the Gronwall inequality.

Finally, to show (26), it suffices to use the estimate (37), (39), and (40) with α>0\alpha>0, in addition to the previous arguments. For instance, let us consider the stochastic convolution in the additive noise case where we apply (39) with α>0\alpha>0:

IIIτLωpLx2\displaystyle\|III_{\tau}\|_{L_{\omega}^{p}L_{x}^{2}} ε1(j=0m1tjtj+1StmrSτmjTτLωp202𝑑r)12\displaystyle\leq\varepsilon^{-1}(\sum_{j=0}^{m-1}\int_{t_{j}}^{t_{j+1}}\|S_{t_{m}-r}-S_{\tau}^{m-j}T_{\tau}\|^{2}_{L_{\omega}^{p}\mathcal{L}_{2}^{0}}dr)^{\frac{1}{2}}
Ceαtmε1τ(μ/31/2)1/2.\displaystyle\leq Ce^{-\alpha t_{m}\varepsilon^{-1}}\tau^{(\mu/3-1/2)\wedge 1/2}.

The situation is similar to the estimates of the other terms. This finishes the proof of (26). ∎

Combining Theorems 4.1 and 5.1 will then give the following strong convergence rate between the exact solution of Eq. (2) and its full discretization (21).

Theorem 5.2.

Let p2p\geq 2, μ0\mu\geq 0, u0εLp(Ω;H˙μ)u^{\varepsilon}_{0}\in L^{p}(\Omega;\dot{H}^{\mu}), and Assumption 2.1 hold. Then

(30) supmMuε(tm)umε,NLωpLx2CeCTε2(1+u0εLωpH˙μ)(Nμ/d+τ(μ/31/2)1/2).\displaystyle\sup_{m\in\mathbb{Z}_{M}}\|u^{\varepsilon}(t_{m})-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\leq Ce^{CT\varepsilon^{-2}}(1+\|u_{0}^{\varepsilon}\|_{L_{\omega}^{p}\dot{H}^{\mu}})(N^{-\mu/d}+\tau^{(\mu/3-1/2)\wedge 1/2}).

If F(u)=𝐢αuF(u)={\bf i}\alpha u with α>0\alpha>0 and G=IdG={\rm Id} such that (5) holds, then there exists a positive constant CC such that

(31) supmuε(tm)umε,NLωpLx2C(u0εLωpH˙μ+(αε)1/2)(Nμ/d+τ(μ/31/2)1/2).\displaystyle\sup_{m\in\mathbb{N}}\|u^{\varepsilon}(t_{m})-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\leq C(\|u^{\varepsilon}_{0}\|_{L_{\omega}^{p}\dot{H}^{\mu}}+(\alpha\varepsilon)^{-1/2})(N^{-\mu/d}+\tau^{(\mu/3-1/2)\wedge 1/2}).
Remark 5.1.

We can formulate the following meshing strategy based on (30) and (31). For instance, let F(u)=𝐢αuF(u)={\bf i}\alpha u with α>0\alpha>0, G=IdG={\rm Id} such that (5) holds, and δ>0\delta>0 be the desired error bound. Then supmuε(tm)umε,NLωpLx2δ\sup_{m\in\mathbb{N}}\|u^{\varepsilon}(t_{m})-u_{m}^{\varepsilon,N}\|_{L_{\omega}^{p}L_{x}^{2}}\leq\delta holds provided τ=N2μd\tau=N^{-\frac{2\mu}{d}} with μ3\mu\geq 3 or τ=N6μd(2s3)\tau=N^{-\frac{6\mu}{d(2s-3)}} with μ[2,3)\mu\in[2,3) and

N2μd/εCδ2.\displaystyle N^{-\frac{2\mu}{d}}/\varepsilon\leq C\delta^{2}.

Appendix

In this appendix, we collect several regularity results and error estimates of the Schrödinger semigroup and its discretizations.

Lemma 5.2.

Let μ0,ρ[0,1]\mu\geq 0,\rho\in[0,1], uH˙μu\in\dot{H}^{\mu}, vH˙ρv\in\dot{H}^{\rho}, α\alpha\in\mathbb{R}, and t0t\geq 0. Then

(32) Stα,εuμ\displaystyle\|S_{t}^{\alpha,\varepsilon}u\|_{\mu} =eαtε1uμ,\displaystyle=e^{-\alpha t\varepsilon^{-1}}\|u\|_{\mu},
(33) (StεId)v\displaystyle\|(S^{\varepsilon}_{t}-{\rm Id})v\| Cερtρv2ρ,\displaystyle\leq C\varepsilon^{\rho}t^{\rho}\|v\|_{2\rho},
(34) (Stα,εId)v\displaystyle\|(S^{\alpha,\varepsilon}_{t}-{\rm Id})v\| Cερtρeαtε1v2ρ+v2.\displaystyle\leq C\varepsilon^{\rho}t^{\rho}e^{-\alpha t\varepsilon^{-1}}\|v\|_{2\rho}+\|v\|^{2}.
Proof.

By the definitions of Stα,εS_{t}^{\alpha,\varepsilon} and μ\|\cdot\|_{\mu}-norm, in combination with the increasing property of λk\lambda_{k} with respect to k+k\in\mathbb{N}_{+}, we have

Stα,εvμ2\displaystyle\|S_{t}^{\alpha,\varepsilon}v\|_{\mu}^{2} =k=1λkμek,v2|e(𝐢ε2λkαε)t|2=e2αεtk=1λkμek,v2=e2αεtvμ2,\displaystyle=\sum_{k=1}^{\infty}\lambda_{k}^{\mu}\langle e_{k},v\rangle^{2}|e^{(\frac{{\bf i}\varepsilon}{2}\lambda_{k}-\frac{\alpha}{\varepsilon})t}|^{2}=e^{-\frac{2\alpha}{\varepsilon}t}\sum_{k=1}^{\infty}\lambda_{k}^{\mu}\langle e_{k},v\rangle^{2}=e^{-\frac{2\alpha}{\varepsilon}t}\|v\|_{\mu}^{2},

which shows (32).

Similarly,

(Stα,εId)v2\displaystyle\|(S^{\alpha,\varepsilon}_{t}-{\rm Id})v\|^{2} =k=1ek,v2|e(𝐢ε2λkαε)t1|2.\displaystyle=\sum_{k=1}^{\infty}\langle e_{k},v\rangle^{2}|e^{(\frac{{\bf i}\varepsilon}{2}\lambda_{k}-\frac{\alpha}{\varepsilon})t}-1|^{2}.

It is clear that for any β[0,1]\beta\in[0,1],

(35) |e𝐢xe𝐢y|=2|sin[(xy)/2]|2|xy|β,x,y.\displaystyle|e^{{\bf i}x}-e^{{\bf i}y}|=2|\sin[(x-y)/2]|\leq 2|x-y|^{\beta},\quad\forall~{}x,y\in\mathbb{R}.

Then

(StεId)v2\displaystyle\|(S^{\varepsilon}_{t}-{\rm Id})v\|^{2} |ε2t|2ρk=1|λk|2ρek,v2Cε2ρt2ρv2ρ2,\displaystyle\leq\left|\frac{\varepsilon}{2}t\right|^{2\rho}\sum_{k=1}^{\infty}|\lambda_{k}|^{2\rho}\langle e_{k},v\rangle^{2}\leq C\varepsilon^{2\rho}t^{2\rho}\|v\|_{2\rho}^{2},

which shows (33). To show (34), we just need to note that

|e(𝐢ε2λkαε)t1|eαεt|e𝐢ε2λkt1|+|1eαεt|2eαεt|ε2tλk|ρ+1,\displaystyle|e^{(\frac{{\bf i}\varepsilon}{2}\lambda_{k}-\frac{\alpha}{\varepsilon})t}-1|\leq e^{-\frac{\alpha}{\varepsilon}t}|e^{\frac{{\bf i}\varepsilon}{2}\lambda_{k}t}-1|+|1-e^{-\frac{\alpha}{\varepsilon}t}|\leq 2e^{-\frac{\alpha}{\varepsilon}t}\left|\frac{\varepsilon}{2}t\lambda_{k}\right|^{\rho}+1,

and then use the previous argument. ∎

Lemma 5.3.

Let μ0\mu\geq 0, vH˙μv\in\dot{H}^{\mu}, α\alpha\in\mathbb{R}, and t0t\geq 0. Then

(36) (Stα,εStα,ε,N)veαtε1λNμ/2vμ.\displaystyle\|(S_{t}^{\alpha,\varepsilon}-S_{t}^{\alpha,\varepsilon,N})v\|\leq e^{-\alpha t\varepsilon^{-1}}\lambda_{N}^{-\mu/2}\|v\|_{\mu}.
Proof.

By the definitions of Stα,εS_{t}^{\alpha,\varepsilon}, Stα,ε,NS_{t}^{\alpha,\varepsilon,N}, and μ\|\cdot\|_{\mu}-norm, in combination with the increasing property of λk\lambda_{k} with respect to k+k\in\mathbb{N}_{+}, we have

(Stα,εStα,ε,N)v2\displaystyle\|(S_{t}^{\alpha,\varepsilon}-S_{t}^{\alpha,\varepsilon,N})v\|^{2} =k=N+1ek,v2|e(𝐢ε2λkαε)t|2=e2αεtk=N+1ek,v2\displaystyle=\sum_{k=N+1}^{\infty}\langle e_{k},v\rangle^{2}|e^{(\frac{{\bf i}\varepsilon}{2}\lambda_{k}-\frac{\alpha}{\varepsilon})t}|^{2}=e^{-\frac{2\alpha}{\varepsilon}t}\sum_{k=N+1}^{\infty}\langle e_{k},v\rangle^{2}
e2αεtλNμk=N+1|λk|μek,v2e2αεtλNμvμ2,\displaystyle\leq e^{-\frac{2\alpha}{\varepsilon}t}\lambda_{N}^{-\mu}\sum_{k=N+1}^{\infty}|\lambda_{k}|^{\mu}\langle e_{k},v\rangle^{2}\leq e^{-\frac{2\alpha}{\varepsilon}t}\lambda_{N}^{-\mu}\|v\|_{\mu}^{2},

which implies the assertion. ∎

Lemma 5.4.

Let μ(0,6]\mu\in(0,6], α\alpha\in\mathbb{R}, and k+k\in\mathbb{N}_{+}. Then

(37) (Sα,ε(tk)(Sτα,ε)k)(H˙μ,L2)Ceαtk/εtkμ/6εμ/2τμ/3.\displaystyle\|(S^{\alpha,\varepsilon}(t_{k})-(S_{\tau}^{\alpha,\varepsilon})^{k})\|_{\mathcal{L}(\dot{H}^{\mu},L^{2})}\leq Ce^{-\alpha t_{k}/\varepsilon}t_{k}^{\mu/6}\varepsilon^{\mu/2}\tau^{\mu/3}.
Proof.

By the definitions of Stα,εS_{t}^{\alpha,\varepsilon}, Stα,ε,NS_{t}^{\alpha,\varepsilon,N}, and μ\|\cdot\|_{\mu}-norm, in combination with the increasing property of λk\lambda_{k} with respect to k+k\in\mathbb{N}_{+}, we have

(Sα,ε(tk)(Sτα,ε)k)(H˙μ,L2)2\displaystyle\|(S^{\alpha,\varepsilon}(t_{k})-(S_{\tau}^{\alpha,\varepsilon})^{k})\|_{\mathcal{L}(\dot{H}^{\mu},L^{2})}^{2}
=supvμ=1(Sα,ε(tk)(Sτα,ε)k)v2\displaystyle=\sup_{\|v\|_{\mu}=1}\|(S^{\alpha,\varepsilon}(t_{k})-(S_{\tau}^{\alpha,\varepsilon})^{k})v\|^{2}
=supvμ=1m=1(Sα,ε(tk)(Sτα,ε)k)v,em2\displaystyle=\sup_{\|v\|_{\mu}=1}\sum_{m=1}^{\infty}\langle(S^{\alpha,\varepsilon}(t_{k})-(S_{\tau}^{\alpha,\varepsilon})^{k})v,e_{m}\rangle^{2}
=e2αtk/εsupvμ=1m=1|e𝐢εtkλm2(1+𝐢ετλm/41𝐢ετλm/4)k|2v,em2\displaystyle=e^{-2\alpha t_{k}/\varepsilon}\sup_{\|v\|_{\mu}=1}\sum_{m=1}^{\infty}|e^{\frac{{\bf i}\varepsilon t_{k}\lambda_{m}}{2}}-(\frac{1+{\bf i}\varepsilon\tau\lambda_{m}/4}{1-{\bf i}\varepsilon\tau\lambda_{m}/4})^{k}|^{2}\langle v,e_{m}\rangle^{2}
e2αtk/ε(supm+|e𝐢εtkλm2(1+𝐢ετλm/41𝐢ετλm/4)k|2λmμ)(supvμ=1k=1λkμv,ek2)\displaystyle\leq e^{-2\alpha t_{k}/\varepsilon}(\sup_{m\in\mathbb{N}_{+}}|e^{\frac{{\bf i}\varepsilon t_{k}\lambda_{m}}{2}}-(\frac{1+{\bf i}\varepsilon\tau\lambda_{m}/4}{1-{\bf i}\varepsilon\tau\lambda_{m}/4})^{k}|^{2}\lambda_{m}^{-\mu})(\sup_{\|v\|_{\mu}=1}\sum_{k=1}^{\infty}\lambda_{k}^{\mu}\langle v,e_{k}\rangle^{2})
=e2αtk/εsupm+|e𝐢εtkλm/2e2𝐢karctan(ετλm/4)|2λmμ.\displaystyle=e^{-2\alpha t_{k}/\varepsilon}\sup_{m\in\mathbb{N}_{+}}\left|e^{{\bf i}\varepsilon t_{k}\lambda_{m}/2}-e^{2{\bf i}k\arctan(\varepsilon\tau\lambda_{m}/4)}\right|^{2}\lambda_{m}^{-\mu}.

Using the inequality (35), we have

(Sα,ε(tk)(Sτα,ε)k)(H˙μ,L2)2e2αtk/εsupk+(2k)2β|ετλk/4arctan(ετλk/4)|2βλkμ.\begin{split}&\|(S^{\alpha,\varepsilon}(t_{k})-(S_{\tau}^{\alpha,\varepsilon})^{k})\|_{\mathcal{L}(\dot{H}^{\mu},L^{2})}^{2}\\[4.2679pt] &\leq e^{-2\alpha t_{k}/\varepsilon}\sup_{k\in\mathbb{N}_{+}}(2k)^{2\beta}|\varepsilon\tau\lambda_{k}/4-\arctan(\varepsilon\tau\lambda_{k}/4)|^{2\beta}\lambda_{k}^{-\mu}.\end{split}

By the elementary inequality |xarctan(x)|C|x|3|x-\arctan(x)|\leq C|x|^{3}, xx\in\mathbb{R}, we get

(Sα,ε(tk)(Sτα,ε)k)(H˙μ,L2)\displaystyle\|(S^{\alpha,\varepsilon}(t_{k})-(S_{\tau}^{\alpha,\varepsilon})^{k})\|_{\mathcal{L}(\dot{H}^{\mu},L^{2})} Ceαtk/εtkβτ2βε3βsupk+λk(μ/23β),\displaystyle\leq Ce^{-\alpha t_{k}/\varepsilon}t_{k}^{\beta}\tau^{2\beta}\varepsilon^{3\beta}\sup_{k\in\mathbb{N}_{+}}\lambda_{k}^{-(\mu/2-3\beta)},

which shows (37) with β=μ/6\beta=\mu/6. ∎

Lemma 5.5.

Let β[1/3,1]\beta\in[1/3,1], j+j\in\mathbb{N}_{+}, and r[tj1,tj]r\in[t_{j-1},t_{j}]. Then

(38) SrSτjTτ(H˙6β;L2)Ctjβτ(2β)1ε.\displaystyle\|S_{-r}-S_{\tau}^{-j}T_{\tau}\|_{\mathcal{L}(\dot{H}^{6\beta};L^{2})}\leq Ct_{j}^{\beta}\tau^{(2\beta)\wedge 1}\varepsilon.

Let s2s\geq 2, uLp(Ω;H1)u\in L^{p}(\Omega;H^{1}), α\alpha\in\mathbb{R}, and t0t\geq 0. Then

(39) tjtj+1(StmrSτmjTτ)PN(H˙μ;L2)𝑑rCeαtmε1tms/6ετ1+(μ/3)1,\displaystyle\int_{t_{j}}^{t_{j+1}}\|(S_{t_{m}-r}-S_{\tau}^{m-j}T_{\tau})P_{N}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}dr\leq Ce^{-\alpha t_{m}\varepsilon^{-1}}t_{m}^{s/6}\varepsilon\tau^{1+(\mu/3)\wedge 1},
(40) (tjtj+1(StmrSτmjTτ)PN(H˙μ;L2)2𝑑r)12Ceαtmε1tms/6ετ1/2+(μ/3)1.\displaystyle(\int_{t_{j}}^{t_{j+1}}\|(S_{t_{m}-r}-S_{\tau}^{m-j}T_{\tau})P_{N}\|^{2}_{\mathcal{L}(\dot{H}^{\mu};L^{2})}dr)^{\frac{1}{2}}\leq Ce^{-\alpha t_{m}\varepsilon^{-1}}t_{m}^{s/6}\varepsilon\tau^{1/2+(\mu/3)\wedge 1}.
Proof.

Since SτS_{\tau} is isometric in L2L^{2}, we have

SrSτjTτ(H˙μ;L2)\displaystyle\|S_{-r}-S_{\tau}^{-j}T_{\tau}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}
SrStj(H˙μ;L2)+StjSτj(H˙μ;L2)+Sτj(IdTτ)(H˙μ;L2)\displaystyle\leq\|S_{-r}-S_{-t_{j}}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}+\|S_{-t_{j}}-S_{\tau}^{-j}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}+\|S_{\tau}^{-j}({\rm Id}-T_{\tau})\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}
=SrStj(H˙μ;L2)+StjSτj(H˙μ;L2)+IdTτ(H˙μ;L2).\displaystyle=\|S_{-r}-S_{-t_{j}}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}+\|S_{-t_{j}}-S_{\tau}^{-j}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}+\|{\rm Id}-T_{\tau}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}.

As in the proof of Lemma 5.4, it holds that

SrStj(H˙μ;L2)\displaystyle\|S_{-r}-S_{-t_{j}}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}
supk+|e𝐢εrλk/2e2𝐢jarctan(ετλk/4)|λkμ/2\displaystyle\leq\sup_{k\in\mathbb{N}_{+}}|e^{{\bf i}\varepsilon r\lambda_{k}/2}-e^{2{\bf i}j\arctan(\varepsilon\tau\lambda_{k}/4)}|\lambda_{k}^{-\mu/2}
supk+|e𝐢εrλk/2e𝐢εtjλk/2|λkμ/2+supk+|e𝐢εtjλk/2e2𝐢jarctan(ετλk/4)|λkμ/2\displaystyle\leq\sup_{k\in\mathbb{N}_{+}}|e^{{\bf i}\varepsilon r\lambda_{k}/2}-e^{{\bf i}\varepsilon t_{j}\lambda_{k}/2}|\lambda_{k}^{-\mu/2}+\sup_{k\in\mathbb{N}_{+}}|e^{{\bf i}\varepsilon t_{j}\lambda_{k}/2}-e^{2{\bf i}j\arctan(\varepsilon\tau\lambda_{k}/4)}|\lambda_{k}^{-\mu/2}
supk+|ελk(rtj)|λkμ/2+supk+(2j)β|ετλk/4arctan(ετλk/4)|βλkμ/2\displaystyle\leq\sup_{k\in\mathbb{N}_{+}}|\varepsilon\lambda_{k}(r-t_{j})|\lambda_{k}^{-\mu/2}+\sup_{k\in\mathbb{N}_{+}}(2j)^{\beta}|\varepsilon\tau\lambda_{k}/4-\arctan(\varepsilon\tau\lambda_{k}/4)|^{\beta}\lambda_{k}^{-\mu/2}
τεsupk+λk(μ/21)+Ctjβτ2βε3β.\displaystyle\leq\tau\varepsilon\sup_{k\in\mathbb{N}_{+}}\lambda_{k}^{-(\mu/2-1)}+Ct_{j}^{\beta}\tau^{2\beta}\varepsilon^{3\beta}.

Similarly,

StjSτj(H˙μ;L2)Ctjβτ2βε3β,\displaystyle\|S_{-t_{j}}-S_{\tau}^{-j}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}\leq Ct_{j}^{\beta}\tau^{2\beta}\varepsilon^{3\beta},

and

IdTτ(H˙μ;L2)\displaystyle\|{\rm Id}-T_{\tau}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})} supk+|1(1𝐢ετλk/4)1|λkμ/214τεsupk+λk(μ/21).\displaystyle\leq\sup_{k\in\mathbb{N}_{+}}|1-(1-{\bf i}\varepsilon\tau\lambda_{k}/4)^{-1}|\lambda_{k}^{-\mu/2}\leq\frac{1}{4}\tau\varepsilon\sup_{k\in\mathbb{N}_{+}}\lambda_{k}^{-(\mu/2-1)}.

Combining the above estimates, equality can be derived from the equality (38).

Finally, as StS_{t}, tt\in\mathbb{R}, is isometric in L2L^{2}, we have

(StmrαSτα,mjTτ)PN(H˙μ;L2)\displaystyle\|(S_{t_{m}-r}^{\alpha}-S_{\tau}^{\alpha,m-j}T_{\tau})P_{N}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})} =eαtmε1SrSτjTτ(H˙μ;L2)\displaystyle=e^{-\alpha t_{m}\varepsilon^{-1}}\|S_{-r}-S_{\tau}^{-j}T_{\tau}\|_{\mathcal{L}(\dot{H}^{\mu};L^{2})}
Ceαtmε1tjβτ(2β)1ε.\displaystyle\leq Ce^{-\alpha t_{m}\varepsilon^{-1}}t_{j}^{\beta}\tau^{(2\beta)\wedge 1}\varepsilon.

which shows (39) and (40).

References

  • [BJM02] W. Bao, S. Jin, and P. A. Markowich, On time-splitting spectral approximations for the Schrödinger equation in the semiclassical regime, J. Comput. Phys. 175 (2002), 487–524.
  • [BJM03] W. Bao, S. Jin, and P. Markowich, Numerical study of time-splitting spectral discretizations of nonlinear Schrödinger equations in the semiclassical regimes, SIAM J. Sci. Comput. 25 (2003), 27–64.
  • [CH16] C. Chen and J. Hong, Symplectic Runge–Kutta semidiscretization for stochastic Schrödinger equation, SIAM J. Numer. Anal. 54 (2016), 2569–2593.
  • [CHJ17] C. Chen, J. Hong, and L. Ji, Mean-square convergence of a symplectic local discontinuous Galerkin method applied to stochastic linear Schrödinger equation, IMA J. Numer. Anal. 37 (2017), 1041–1065.
  • [CHL17] J. Cui, J. Hong, and Z. Liu, Strong convergence rate of finite difference approximations for stochastic cubic Schrödinger equations, J. Differ. Equ. 263 (2017), 3687–3713.
  • [CHLZ17a] J. Cui, J. Hong, Z. Liu, and W. Zhou, Numerical analysis on ergodic limit of approximations for stochastic NLS equation via multi-symplectic scheme, SIAM. J. Numer. Anal. 55 (2017), 305–327.
  • [CHLZ17b] by same author, Stochastic symplectic and multi-symplectic methods for nonlinear Schrödinger equation with white noise dispersion, J. Comput. Phys. 342 (2017), 267–285.
  • [CHLZ19] by same author, Strong convergence rate of splitting schemes for stochastic nonlinear Schrödinger equations, J. Differ. Equ. 266 (2019), 5625–5663.
  • [CHP16] C. Chen, J. Hong, and A. Prohl, Convergence of a θ\theta-scheme to solve the stochastic nonlinear Schrödinger equation with stratonovich noise, Stoch. Partial Differ. Equ. Anal. Comput. 4 (2016), 274–318.
  • [CL22] J. Cai and H. Liang, Compact exponential conservative approaches for the Schrödinger equation in the semiclassical regimes, SIAM J. Sci. Comput. 44 (2022), B585–B604.
  • [CLZ23] J. Cui, S. Liu, and H. Zhou, Optimal control for stochastic nonlinear Schrödinger equation on graph, SIAM J. Control Optim. 61 (2023), 2021–2042.
  • [CMZ20] J. Chen, D. Ma, and Z. Zhang, A multiscale reduced basis method for the Schrödinger equation with multiscale and random potentials, Multiscale Model. Simul. 18 (2020), 1409–1434.
  • [CZ23] J. Cai and H. Zhang, High-order conservative schemes for the nonlinear Schrödinger equation in the semiclassical limit, Appl. Math. Lett. 144 (2023), Paper No. 108703, 10. MR 4589150
  • [DBD04] A. De Bouard and A. Debussche, A semi-discrete scheme for the stochastic nonlinear Schrödinger equation, Numer. Math. 96 (2004), 733–770.
  • [FGL09] E. Faou, V. Gradinaru, and C. Lubich, Computing semiclassical quantum dynamics with Hagedorn wavepackets, SIAM J. Sci. Comput. 31 (2009), 3027–3041.
  • [HL19] J. Hong and Z. Liu, Well-posedness and optimal regularity of stochastic evolution equations with multiplicative noises, J. Diff. Equ. 266 (2019), no. 8, 4712–4745.
  • [HW19] J. Hong and X. Wang, Invariant Measures for Stochastic Nonlinear Schrödinger Equations: Numerical Approximations and Symplectic Structures, Lecture Notes in Mathematics 2251, Springer Singapore, 2019.
  • [JLRZ20] S. Jin, L. Liu, G. Russo, and Z. Zhou, Gaussian wave packet transform based numerical scheme for the semi-classical Schrödinger equation with random inputs, J. Comput. Phys. 401 (2020), 109015.
  • [JMS11] S. Jin, P. Markowich, and C. Sparber, Mathematical and computational methods for semiclassical Schrödinger equations, Acta Numer. 20 (2011), 121–209.
  • [JWY11] S. Jin, H. Wu, and X. Yang, Semi-Eulerian and high order Gaussian beam methods for the Schrödinger equation in the semiclassical regime, Commun. Comput. Phys. 9 (2011), 668–687.
  • [JZ13] S. Jin and Z. Zhou, A semi-Lagrangian time splitting method for the Schrödinger equation with vector potentials, Commun. Inf. Syst. 13 (2013), 247–289.
  • [Kay06] K. Kay, The Herman-Kluk approximation: derivation and semiclassical corrections, Chem. Phys. 322 (2006), 3–12.
  • [Liu13] J. Liu, Order of convergence of splitting schemes for both deterministic and stochastic nonlinear Schrödinger equations, SIAM J. Numer. Anal. 51 (2013), 1911–1932.
  • [LL20] C. Lasser and C. Lubich, Computing quantum dynamics in the semiclassical regime, Acta Numer. 29 (2020), 229–401.
  • [LQ21] Z. Liu and Z. Qiao, Strong approximation of monotone stochastic partial differential equations driven by multiplicative noise, Stoch. Partial Differ. Equ. Anal. Comput. 9 (2021), no. 3, 559–602.
  • [LR15] W. Liu and M. Röckner, Stochastic Partial Differential Eequations: An Introduction, Springer Singapore, 2015.
  • [LZ18] J. Lu and Z. Zhou, Frozen Gaussian approximation with surface hopping for mixed quantum-classical dynamics: a mathematical justification of fewest switches surface hopping algorithms, Math. Comput. 87 (2018), 2189–2232.
  • [MPP99] P. Markowich, P. Pietra, and C. Pohl, Numerical approximation of quadratic observables of Schrödinger-type equations in the semi-classical limit, Numer. Math. 81 (1999), 595.
  • [QY10] J. Qian and L. Ying, Fast multiscale Gaussian wavepacket transforms and multiscale Gaussian beams for the wave equation, Multiscale Model. Simul. 8 (2010), 1803–1837.
  • [WL24] Y. Wang and L. Liu, On a neural network approach for solving potential control problem of the semiclassical Schrödinger equation, J. Comput. Appl. Math. 438 (2024), 115504.
  • [Zho14] Z. Zhou, Numerical approximation of the Schrödinger equation with the electromagnetic field by the Hagedorn wave packets, J. Comput. Phys. 272 (2014), 386–407.