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

Asymptotic behavior of discrete Schrödinger equation on the hexagonal triangulation

Huabin Ge, Bobo Hua, Longsong Jia, Puchun Zhou
Abstract

In this article, we prove the decay estimate for the discrete Schrödinger equation (DS) on the hexagonal triangulation. The l1ll^{1}\rightarrow l^{\infty} dispersive decay rate is t34\left\langle t\right\rangle^{-\frac{3}{4}}, which is faster than the decay rate of DS on the 2-dimensional lattice 2\mathbb{Z}^{2}, which is t23\left\langle t\right\rangle^{-\frac{2}{3}}, see [32]. The proof relies on the detailed analysis of singularities of the corresponding phase function and the theory of uniform estimates on oscillatory integrals developed by Karpushkin [15]. Moreover, we prove the Strichartz estimate and give an application to the discrete nonlinear Schrödinger equation (DNLS) on the hexagonal triangulation.

1 Introduction

1.1 discrete Schrödinger equation on the hexagonal triangulation

Discrete Schrödinger equations are fundamental discrete dynamical models, which has many applications in physics, see e.g. [26, 27, 19, 8, 9, 20]. Let G=(V,E)G=(V,E) be an undirect graph with weights {we}eE\{w_{e}\}_{e\in E} on edges. We write uvu\sim v if uu and vv are connected by an edge in EE. A Schrödinger equation on GG describes the continuous-time random walk (CTQW) on GG, see e.g. [30], written as

{itu(v,t)+ΔGu(v,t)=0,(v,t)V×[0,),u(v,0)=ψ(v),vV.\displaystyle\left\{\begin{aligned} &i\partial_{t}u(v,t)+\Delta_{G}u(v,t)=0,~{}~{}\forall(v,t)\in V\times[0,\infty),\\ &u(v,0)=\psi(v),~{}~{}\forall v\in V.\end{aligned}\right. (1.1)

where the discrete Laplacian ΔG\Delta_{G} is given by

ΔGfu=v:vwωuv(fvfw).\displaystyle\Delta_{G}f_{u}=\sum_{v:v\sim w}\omega_{uv}(f_{v}-f_{w}). (1.2)

The Strichartz estimate is a powerful tool for analyzing the asymptotic behavior of dispersive equations, which was first introduced in [34]. To establish the Strichartz estimate, one usually need an l1ll^{1}\rightarrow l^{\infty} estimate (or L1LL^{1}\rightarrow L^{\infty} estimate in the continuous case) with the form

u(v,t)lCtσψl1,\displaystyle\|u(v,t)\|_{l^{\infty}}\leq C\left\langle t\right\rangle^{-\sigma}\|\psi\|_{l^{1}}, (1.3)

where t\left\langle t\right\rangle is the Japanese bracket given by t:=1+|t|,\left\langle t\right\rangle:=1+|t|, and CC is a constant independent of t.t. In the continuous case, it is well-known that the L1LL^{1}\rightarrow L^{\infty} estimate for linear Schrödinger equations in d\mathbb{R}^{d} holds with σ=d2.\sigma=\frac{d}{2}. As for the discrete Schrödinger equations on d\mathbb{Z}^{d}, the l1ll^{1}\rightarrow l^{\infty} estimate holds for σ=d3\sigma=\frac{d}{3}, see [32]. Application of those dispersive estimates can be found in [28, 18, 25, 4], where the spectral problem and stability of breathers are discussed. Dispersive estimates for the discrete wave equation (DW) and the discrete Klein-Gorden equation (DKG) on d\mathbb{Z}^{d} are also studied in [32, 10, 12, 11, 7].

Dispersive estimates on metric graphs has also been studied in the past 20 years. In [1, 6, 5, 13, 22, 23] the dispersive estimates for star-shaped networks and tadpole graph are discussed. Recently, dispersive estimates on regular trees and Cartesian product of the integer lattice and a finite graph were studied by Ammari and Sabri [2, 3]. After that finite metric graphs with infinite ends are also considered, see [24]. We should also remark that the dispersive estimate of quantum walk in discrete time on 11-dimensional lattice is established in [21].

Those works inspired us to explore dispersive estimates on more general graphs, especially the triangulations of surfaces. In this article, we mainly focus on the discrete Schrödinger equation (DS) on the 11-skeleton of the hexagonal triangulation 𝒯H\mathcal{T}_{H}, which is a regular triangulation of 2\mathbb{R}^{2} with degree 66, as shown in Figure 1.1. The weights we focus on here are the simplest weights with ω1.\omega\equiv 1. We denote by ΔGH\Delta_{G_{H}} the discrete Laplacian on 𝒯H\mathcal{T}_{H} .

[Uncaptioned image]
\captionof

figureHexagonal triangulation 𝒯H\mathcal{T}_{H} of the plane.

Since the hexagonal triangulation can be viewed as a Cayley graph of the two-dimensional lattice 2\mathbb{Z}^{2}, with the help of discrete Fourier analysis on the lattice, we can obtain the explicit formula of the solution to (LABEL:LS) in C1([0,+);l2(2))C^{1}([0,+\infty);l^{2}(\mathbb{Z}^{2})).

u(n,t)=2k2u(k,0)[0,2π]2eitg(x)+i(kn),xdx,n2,\displaystyle u(n,t)=2\sum_{k\in\mathbb{Z}^{2}}u(k,0)\int_{[0,2\pi]^{2}}e^{-itg(x)+i\left\langle(k-n),x\right\rangle}\mathrm{d}x,~{}~{}\forall n\in\mathbb{Z}^{2}, (1.4)

where x=(x1,x2)tx=(x_{1},x_{2})^{t} and

g(x)=62cosx12cosx22cos(x1+x2).\displaystyle g(x)=6-2\cos x_{1}-2\cos x_{2}-2\cos(x_{1}+x_{2}). (1.5)

To obtain the l1ll^{1}\rightarrow l^{\infty} estimate of the solution to (LABEL:LS), it is sufficient to obtain a uniform estimate of the oscillatory integral G(l,t)=[0,2π]2eitg(x)+il,xdxG(l,t)=\int_{[0,2\pi]^{2}}e^{-itg(x)+i\left\langle l,x\right\rangle}\mathrm{d}x, i.e.

G(l,t)Ctσ,l2,\displaystyle G(l,t)\leq C\left\langle t\right\rangle^{-\sigma},~{}~{}\forall~{}l\in\mathbb{Z}^{2}, (1.6)

where CC is a positive constant independent of the parameter ll.

The estimate of oscillatory integrals required for the Schrödinger equation on the hexagonal triangulation differs from the case on 2\mathbb{Z}^{2} in [32], as the inability to apply the method of separation of variables poses significant difficulties for the estimate. In this article, we will apply the method in [31], which established the dispersive estimate of the discrete wave equation on 2\mathbb{Z}^{2}, and reduce the estimate (1.6) to a uniform estimate of a oscillatory integral with parameters.

Let v=(v1,v2)=l/tv=(v_{1},v_{2})=l/t, then we only need to estimate the oscillatory integral

I(t,v)=G(tv,t)=[0,2π]2eitϕ(v,x)dx,(t,v)×2,I(t,v)=G(tv,t)=\int_{[0,2\pi]^{2}}e^{-it\phi(v,x)}\mathrm{d}x,~{}~{}(t,v)\in\mathbb{R}\times\mathbb{R}^{2},

where the phase function ϕ(v,x)\phi(v,x) is given by

ϕ(v,x)=g(x)v,x.\displaystyle\phi(v,x)=g(x)-\left\langle v,x\right\rangle. (1.7)

The problem then is reduced to find a uniform estimate of I(t,v).I(t,v). First, we transform I(x,t)I(x,t) to an oscillatory integral on 2.\mathbb{R}^{2}. Let 𝐚(x)\mathbf{a}(x) be a function in Cc(2)C_{c}^{\infty}(\mathbb{R}^{2}), whose support is contained in [4π,4π]2.[-4\pi,4\pi]^{2}. Moreover, 𝐚(x)\mathbf{a}(x) is chosen to have the following properties.

  1. 1.

    There exists a neighborhood WW of [0,2π]2[0,2\pi]^{2} such that

    𝐚(x)0,xW.\mathbf{a}(x)\neq 0,~{}~{}\forall x\in W.
  2. 2.

    l2𝐚(x+2πl)=1,x2.\sum_{l\in\mathbb{Z}^{2}}\mathbf{a}(x+2\pi l)=1,~{}~{}\forall x\in\mathbb{R}^{2}.

Then we have

I(t,v)\displaystyle I(t,v) =l2[0,2π]2eitϕ(x)𝐚(x+2πl)dx\displaystyle=\sum_{l\in\mathbb{Z}^{2}}\int_{[0,2\pi]^{2}}e^{-it\phi(x)}\mathbf{a}(x+2\pi l)\mathrm{d}x
=2eitϕ(x)𝐚(x)dx\displaystyle=\int_{\mathbb{R}^{2}}e^{-it\phi(x)}\mathbf{a}(x)\mathrm{d}x (1.8)

With the help of uniform estimates of oscillatory integrals in Section 2 and the analysis of singularities of the phase function ϕ(x)\phi(x) in Section 4, we get the following lemma.

Lemma 1.1.

Let uu be the solution to (LABEL:LS), then we have the following l1ll^{1}\rightarrow l^{\infty} decay estimate.

u(,t)lCt34ψl1\displaystyle\|u(\cdot,t)\|_{l^{\infty}}\leq C\left\langle t\right\rangle^{-\frac{3}{4}}\|\psi\|_{l^{1}} (1.9)

Then by the theory of Keel and Tao [17], we can finally obtain the desired Strichartz estimate.

Let us first introduce the notion of admissible pairs, see [35, Chapter 2].

Definition 1.2.

Fix σ>0\sigma>0, we call a pair (q,r)(q,r) of exponents σ\sigma-admissible if 2q,r2\leq q,r\leq\infty, 1q+σr=σ2\frac{1}{q}+\frac{\sigma}{r}=\frac{\sigma}{2}, and (q,r,σ)(2,,1).(q,r,\sigma)\neq(2,\infty,1).

With the notion of the admissible pair, we have the following theorem.

Theorem 1.3.

Let u(v,t)C1([0,+);l2(2))u(v,t)\in C^{1}([0,+\infty);l^{2}(\mathbb{Z}^{2})) be a solution to (LABEL:LS)\eqref{LS} on the hexagonal triangulation 𝒯H\mathcal{T}_{H}. Then one has the energy identity

u(,t)l2=ψl2,\displaystyle\|u(\cdot,t)\|_{l^{2}}=\|\psi\|_{l^{2}}, (1.10)

and the sharp decay estimate

u(,t)lCt34ψl1.\displaystyle\|u(\cdot,t)\|_{l^{\infty}}\leq C\left\langle t\right\rangle^{-\frac{3}{4}}\|\psi\|_{l^{1}}. (1.11)

And for the inhomogeneous discrete Schrödinger equation

itu(v,t)+ΔGHu(v,t)+F(v,t)=0,(v,t)V×[0,),i\partial_{t}u(v,t)+\Delta_{G_{H}}u(v,t)+F(v,t)=0,~{}~{}\forall(v,t)\in V\times[0,\infty),

there is the Strichartz estimates with σ\sigma-admissible pairs (q,r)(q,r) and (q~,r~)(\tilde{q},\tilde{r}), where σ=34\sigma=\frac{3}{4}.

uLqlrC(ψl2+FLq~(0,T)lr~),\displaystyle\|u\|_{L^{q}l^{r}}\leq C\left(\|\psi\|_{l^{2}}+\|F\|_{L^{\tilde{q}^{\prime}}(0,T)l^{\tilde{r}^{\prime}}}\right), (1.12)

where T(0,]T\in(0,\infty] and CC is a constant independent of TT.

From Theorem 1.11 we see that the decay rate of the discrete Schrödinger equation on the hexagonal triangulation is faster than that on 2\mathbb{Z}^{2}. This phenomenon could be attributed to two possible reasons. Firstly, on the hexagonal triangulation, in addition to edges in the horizontal and vertical directions, there are edges in an additional direction, resulting in a faster decay rate of the solution. Secondly, compared to 2\mathbb{Z}^{2}, geodesic balls on the hexagonal triangulation are closer to those in Euclidean space. Therefore, the decay rate is more similar to that of the Schrödinger equation on a plane.

It is interesting to consider the estimate of Schrödinger equation on other cellular decompositions of 2\mathbb{R}^{2}, such as the hexagonal tiling.

2 Uniform estimates of oscillatory integrals

In this section, we introduce some results of uniform estimates of oscillatory integrals, which are used to prove Lemma 1.1. The results we use come from the pioneer work of uniform estimates of phase functions with two variables introduced by Karpushkin [16].

Definition 2.1.

An oscillatory integral with phase ϕ\phi and amplitude 𝐚\mathbf{a} is an integral given by

J(t,ϕ,𝐚)=n𝐚(x)eitϕ(x)dx,\displaystyle J(t,\phi,\mathbf{a})=\int_{\mathbb{R}^{n}}\mathbf{a}(x)e^{it\phi(x)}\mathrm{d}x, (2.1)

where 𝐚C0(n)\mathbf{a}\in C_{0}^{\infty}(\mathbb{R}^{n}) and t.t\in\mathbb{R}. If the support of 𝐚\mathbf{a} is sufficiently small which is contained in a small neighborhood of x=0x=0, and ϕ\phi is an analytic function at the origin, then as tt\rightarrow\infty, one can obtain the following expansion, see [16].

J(t,ϕ,𝐚)sk=0n1bs,k(𝐚)ts(lnt)k,\displaystyle J(t,\phi,\mathbf{a})\approx\sum_{s}\sum_{k=0}^{n-1}b_{s,k}(\mathbf{a})t^{s}(\ln t)^{k}, (2.2)

where ss is a negative rational number independent of 𝐚.\mathbf{a}.

Definition 2.2.

Let (β(ϕ),p(ϕ))(\beta(\phi),p(\phi)) be the maximum over all pairs (s,k)(s,k) in (2.2) under the lexicographic ordering with the following property.

For any neighborhood WW of 0, there exists a function 𝐚(x)Cc(W)\mathbf{a}(x)\in C_{c}^{\infty}(W), such that bβ(ϕ),p(ϕ)(𝐚)0.b_{\beta(\phi),p(\phi)}(\mathbf{a})\neq 0.

We denote by O(ϕ)=(β(ϕ),p(ϕ))O(\phi)=(\beta(\phi),p(\phi)) the maximal exponential pair. β(ϕ)\beta(\phi) and p(ϕ)p(\phi) are called the oscillatory index of ϕ\phi and the multiplicity of the oscillation of ϕ\phi at 0, see [36, 14].

Now we assume that f:nf:\mathbb{R}^{n}\rightarrow\mathbb{R} is analytic at the point x0.x_{0}. For a positive number r>0r>0, we denote by Hr,x0H_{r,x_{0}} the set of real analytic functions on B(x0,r)B(x_{0},r) which has a unique holomorphic extension on the ball Bn(x0,r).B_{\mathbb{C}^{n}}(x_{0},r).

By Hr,x0(ϵ)H_{r,x_{0}}(\epsilon), we denote the set given by

Hr,x0(ϵ)={fHr,x0:|f|ϵ.}H_{r,x_{0}}(\epsilon)=\{f\in H_{r,x_{0}}:|f|\leq\epsilon.\}
Definition 2.3.

An oscillatory integral with phase ϕ\phi is said to have a uniform estimate at the point x0x_{0} with exponent (β,p)(\beta,p) if for any sufficiently small number r>0r>0, there exists ϵ=ϵ(r)>0\epsilon=\epsilon(r)>0, C=C(r)>0C=C(r)>0 and a neighborhood WB(x0,r)W\subset B(x_{0},r) such that

J(t,ϕ+P,𝐚)Ctβlnp(|t|+2)𝐚CN(W),t.J(t,\phi+P,\mathbf{a})\leq C\left\langle t\right\rangle^{\beta}\ln^{p}(|t|+2)\|\mathbf{a}\|_{C^{N}(W)},~{}~{}\forall t\in\mathbb{R}.

given that 𝐚Cc(W)\mathbf{a}\in C_{c}^{\infty}(W) and PHr,x0(ϵ).P\in H_{r,x_{0}}(\epsilon). And the constant NN only depends on the dimension n.n. For simplicity, we denote by

M(ϕ,x0)(β,p),M(\phi,x_{0})\curlyeqprec(\beta,p),

that ϕ\phi has a uniform estimate at the point x0x_{0} with exponent (β,p).(\beta,p).

In [16], it was proved that for the function ϕ:n\phi:\mathbb{R}^{n}\rightarrow\mathbb{R} is analytic at x0nx_{0}\in\mathbb{R}^{n} with |ϕ(x0)|=0|\nabla\phi(x_{0})|=0, and Hessϕ(x0)\mathrm{Hess}\phi(x_{0}) has corank 2, then one has M(ϕ,x0)O(ϕ).M(\phi,x_{0})\curlyeqprec O(\phi). In particular, when n=2n=2, the following uniform estimates hold.

M(ϕ,x0)O(ϕx0),M(\phi,x_{0})\curlyeqprec O(\phi_{x_{0}}),

where ϕx0=ϕ(x+x0).\phi_{x_{0}}=\phi(x+x_{0}).

Theorem 2.4 ([16]).

Let f:2f:\mathbb{R}^{2}\rightarrow\mathbb{R} be a function that analytic at 0,0, with f(0)=0\nabla f(0)=0 and det(Hessf(0))=0.\mathrm{det}(\mathrm{Hess}f(0))=0. Then

M(f,0)O(ϕ).M(f,0)\curlyeqprec O(\phi).

And by the method of stationary phase, one has the following lemma.

Lemma 2.5.

Let m,n1m,n\geq 1, and

h(ξ,y)=h0(ξ)+P(y),(ξ,y)n×m,h(\xi,y)=h_{0}(\xi)+P(y),~{}\forall(\xi,y)\in\mathbb{R}^{n}\times\mathbb{R}^{m},

where P(y)=j=1mcjyj2P(y)=\sum_{j=1}^{m}c_{j}y_{j}^{2} with cj0.c_{j}\neq 0. If M(h0,0)(β,p)M(h_{0},0)\curlyeqprec(\beta,p), then we have

M(h)(βm2,p).M(h)\curlyeqprec(\beta-\frac{m}{2},p).

3 Newton Polyhedra

In this section, we introduce some basics of Newton polyhedra, see [29]. Let ϕ\phi be a function on n\mathbb{R}^{n} which is real analytic at 0. Assume that

ϕ(0)=|ϕ(0)|=0,\displaystyle\phi(0)=|\nabla\phi(0)|=0, (3.1)

and the Taylor series of ϕ\phi at 0 is

ϕ(x)=αnaαxα,\displaystyle\phi(x)=\sum_{\alpha\in\mathbb{Z}^{n}}a_{\alpha}x^{\alpha},

set 𝒯(ϕ):={αn:aα0}\mathcal{T}(\phi):=\{\alpha\in\mathbb{Z}^{n}:a_{\alpha}\neq 0\}, which is called the Taylor support of ϕ\phi at 0. The Newton polyhedron N(ϕ)N(\phi) is the convex hull of the set

α𝒯(ϕ)(α+(+)n).\displaystyle\cup_{\alpha\in\mathcal{T}(\phi)}(\alpha+(\mathbb{R}_{+})^{n}).
Definition 3.1.

Let 𝒫\mathcal{P} be a face of 𝒩(ϕ).\mathcal{N}(\phi). We denote by ϕ𝒫(x)\phi_{\mathcal{P}}(x) the function defined as

ϕ𝒫(x)=α𝒫aαxα.\displaystyle\phi_{\mathcal{P}}(x)=\sum_{\alpha\in\mathcal{P}}a_{\alpha}x^{\alpha}.

We call ϕ\phi an \mathbb{R}-nondegenerate function if for any compact face 𝒫\mathcal{P} of 𝒩(ϕ)\mathcal{N}(\phi), the following relationship holds.

i=1n{x:iϕ𝒫(x)=0}i=1n{x:xi=0}.\displaystyle\cap_{i=1}^{n}\{x:\partial_{i}\phi_{\mathcal{P}}(x)=0\}\subset\cup_{i=1}^{n}\{x:x_{i}=0\}. (3.2)

Otherwise, we call ϕ\phi is \mathbb{R}-degenerate.

The Newton distance dϕd_{\phi} is given by

dϕ=inf{ρ>0:(ρ,,ρ)𝒩(ϕ)}.\displaystyle d_{\phi}=\inf\{\rho>0:(\rho,...,\rho)\in\mathcal{N}(\phi)\}.

Set dϕ=(dϕ,,dϕ)\textbf{d}_{\phi}=(d_{\phi},...,d_{\phi}). Let FϕF_{\phi} be the principal face of 𝒩ϕ\mathcal{N}_{\phi}, which is the face of minimal dimension containing dϕ.\textbf{d}_{\phi}. Set kϕ=ndim(Fϕ).k_{\phi}=n-\mathrm{dim}(F_{\phi}). Then Varchenko proved the following result in [36].

Theorem 3.2.

Let ϕ\phi be nondegenerate\mathbb{R}-nondegenerate and ϕ(0)=|ϕ(0)|=0.\phi(0)=|\nabla\phi(0)|=0. Assume that (β(ϕ),p(ϕ))(\beta(\phi),p(\phi)) is the maximal exponential pair. Then one has

β(ϕ)1dϕ,p(ϕ)kϕ1.\displaystyle\beta(\phi)\leq-\frac{1}{d_{\phi}},~{}p(\phi)\leq k_{\phi}-1.

4 Singularities of the phase function and the l1ll^{1}\rightarrow l^{\infty} estimate

Now let us analyze the oscillatory integral I(t,v)I(t,v) defined in (1.8). Since the phase function ϕ(v,x)=g(x)v,x\phi(v,x)=g(x)-\left\langle v,x\right\rangle, and

g(x)\displaystyle\nabla g(x) =(2sin(x1)+2sin(x1+x2),2sin(x2)+2sin(x1+x2))t,\displaystyle=(2\sin(x_{1})+2\sin(x_{1}+x_{2}),2\sin(x_{2})+2\sin(x_{1}+x_{2}))^{t},

it is easy to see that g(x)42.\|\nabla g(x)\|\leq 4\sqrt{2}. Then when v=lt42+1\|v\|=\|\frac{l}{t}\|\geq 4\sqrt{2}+1, we have ϕ(v,x)1\|\nabla\phi(v,x)\|\geq 1. Therefore, with the help of the method of stationary phase, see [33], we can easily obtain that the oscillatory integral I(v,t)I(v,t) decays faster than any algebraic function of tt uniformly. So it is sufficient to consider the decay estimate when the parameter vv is contained inside the ball B(0,42+1).B(0,4\sqrt{2}+1).

Let us first consider the Hessian of the phase function Hessxϕ(v,x)\mathrm{Hess}_{x}\phi(v,x). It is simply to verify that

Hessxϕ(v,x)=2(cos(x1+x2)+cosx1cos(x1+x2)cos(x1+x2)cos(x1+x2)+cosx2).\displaystyle\mathrm{Hess}_{x}\phi(v,x)=2\begin{pmatrix}\cos(x_{1}+x_{2})+\cos x_{1}&\cos(x_{1}+x_{2})\\ \cos(x_{1}+x_{2})&\cos(x_{1}+x_{2})+\cos{x_{2}}\end{pmatrix}. (4.1)

We observe that the corank of the Hessian of ϕ(v,x)\phi(v,x) is at most 11. Otherwise, we have cos(x1+x2),cosx1\cos(x_{1}+x_{2}),\cos{x_{1}} and cosx2\cos{x_{2}} are all equal to 0. Therefore, xi=kiπ+π2x_{i}=k_{i}\pi+\frac{\pi}{2}, for some integers ki,i=1,2k_{i},~{}i=1,2. Then x1+x2=(k1+k2+1)πx_{1}+x_{2}=(k_{1}+k_{2}+1)\pi, which leads to a contradiction.

Since the Hessian of ϕ(v,x)\phi(v,x) is independent of the parameter vv, we can define a set Γ\Gamma consisting of points xspt(𝐚)x\in\mathrm{spt}(\mathbf{a}) where Hessxϕ(v,x)\mathrm{Hess}_{x}\phi(v,x) degenerate, where spt(𝐚)\mathrm{spt}(\mathbf{a}) is the support of function aa in the definition (1.8).

Γ={xspt(𝐚):det(Hessx)=0}.\displaystyle\Gamma=\{x\in\mathrm{spt}(\mathbf{a}):\mathrm{det}(\mathrm{Hess}_{x})=0\}.

For simplicity, from now on we denote by Ω\Omega the support of the function 𝐚\mathbf{a}. For each vg(Γ),v\in\nabla g(\Gamma), there exists a set

Γ(v)={xΩ:xϕ(v,x)=0,det(Hessxϕ(v,x))=0.}\Gamma(v)=\{x\in\Omega:\nabla_{x}\phi(v,x)=0,\mathrm{det}(\mathrm{Hess}_{x}\phi(v,x))=0.\}

It is easy to see that for each fixed vg(Γ)v\in\nabla g(\Gamma), Γ(v)\Gamma(v) is a nonempty compact set. Then by the argument in Section 2, for each ξΓ(v)\xi\in\Gamma(v), there exists an exponential pair (β(ϕ(v,),ξ),p(ϕ(v,),ξ))(\beta(\phi(v,\cdot),\xi),p(\phi(v,\cdot),\xi)) such that

M(ϕ(v,),ξ)(β(ϕ(v,),ξ),p(ϕ(v,),ξ)).M(\phi(v,\cdot),\xi)\curlyeqprec(\beta(\phi(v,\cdot),\xi),p(\phi(v,\cdot),\xi)).

Then by the definition of uniform estimates for oscillatory integrals, for each vg(Γ)v\in\nabla g(\Gamma), there exists an exponential pair (β(v),p(v))(\beta(v),p(v)) and positive constants ϵ(v)\epsilon(v) and C(v)C(v) such that

I(v,t)C(v)tβ(v)log|t|p(v),vB(v,ϵ(v)).I(v^{\prime},t)\leq C(v)\left\langle t\right\rangle^{\beta(v)}\log|t|^{p(v)},~{}~{}\forall v^{\prime}\in B(v,\epsilon(v)).

Since g(Γ)\nabla g(\Gamma) is a compact subset of 2,\mathbb{R}^{2}, and vB(v,ϵ(v))g(Γ)\cup_{v}B(v,\epsilon(v))\supset\nabla g(\Gamma), there exists a finite collection {vi}i=1Ng(Γ)\{v_{i}\}_{i=1}^{N}\subset\nabla g(\Gamma) such that

i=1NB(vi,ϵ(v))g(Γ).\cup_{i=1}^{N}B(v_{i},\epsilon(v))\supset\nabla g(\Gamma).

Therefore one can get a uniform estimate of I(v,t)I(v,t) when vi=1NB(vi,ϵ(v))v\in\cup_{i=1}^{N}B(v_{i},\epsilon(v)). For rest the parameters vB(0,42+1)\i=1NB(vi,ϵ(v))v\in B(0,4\sqrt{2}+1)\backslash\cup_{i=1}^{N}B(v_{i},\epsilon(v)), it is easy to find that det(Hessx)ϕ(v,x)C0\mathrm{det}(\mathrm{Hess}_{x})\phi(v,x)\geq C_{0} for some positive constant C0,C_{0}, therefore by the stationary phase method, one can obtain that

I(v,t)C1t1,vB(0,42+1)\i=1NB(vi,ϵ(v)),I(v,t)\leq C_{1}\left\langle t\right\rangle^{-1},~{}~{}\forall v\in B(0,4\sqrt{2}+1)\backslash\cup_{i=1}^{N}B(v_{i},\epsilon(v)),

where C1C_{1} is a positive constant independent of v.v.

Before analyzing singularities of the phase function ϕ(v,x)\phi(v,x), let us first introduce a useful lemma. We call αn\alpha\in\mathbb{R}^{n} a weight if α\alpha is a vector that consists of positive elements. By kαxk^{\alpha}x we denote (kα1x1,,kαnxn)(k^{\alpha_{1}}x_{1},...,k^{\alpha_{n}}x_{n}) for every k>0k>0 and xn.x\in\mathbb{R}^{n}.

Definition 4.1.

For a polynomial h:nh:\mathbb{R}^{n}\rightarrow\mathbb{R}, we call hh quasi-homogenous of degree ρ\rho with respect to the weight α\alpha, if and only if h(kαx)=kρh(x).h(k^{\alpha}x)=k^{\rho}h(x). Let α,n\mathcal{E}_{\alpha,n} be the sets of all quasi-homogenous polynomials of degree 11 with respect to α\alpha. And by α,n\mathcal{H}_{\alpha,n} we denote the set of functions which is real-analytic at the origin, whose Taylor series consists of monomials that are quasi-homogeneous of degree greater than 11 with respect to α.\alpha.

Given a positive weight α\alpha and a function hh that is real-analytic at the origin, then the following lemma, which was first proved in [15], is crucial for our analysis.

Lemma 4.2.

Let h:nh:\mathbb{R}^{n}\rightarrow\mathbb{R} be a function which is real-analytic at the origin. Assuming that there exists a polynomial h0(x)α,nh_{0}(x)\in\mathcal{E}_{\alpha,n} and an analytic function h1(x)α,nh_{1}(x)\in\mathcal{H}_{\alpha,n} such that

h(x)=h0(x)+h1(x),h(x)=h_{0}(x)+h_{1}(x),

then M(h0,0)(β,p)M(h_{0},0)\curlyeqprec(\beta,p) implies M(h,0)(β,p).M(h,0)\curlyeqprec(\beta,p).

With the help of Lemma 4.2, we can reduce the problem of estimating the decay rate of some oscillatory with some polynomial phases. In particular, we have the following Lemma for estimating the oscillatory integral I(t,v).I(t,v).

Lemma 4.3.

Let ϕ(v,x)\phi(v,x) be the phase function (1.5). Set vg(Γ)v\in\nabla g(\Gamma) and xΓ(v).x\in\Gamma(v). Let ϕx0(v,x)=ϕ(v,x+x0).\phi_{x_{0}}(v,x)=\phi(v,x+x_{0}). Then there exists an invertible linear transform u=T(x)u=T(x) on n\mathbb{R}^{n}, the Taylor expansion of the phase function ϕ(v,u)\phi(v,u) with respect to uu under the transformation TT has one of the following forms,

ϕx0(v,u)\displaystyle\phi_{x_{0}}(v,u) =ϕx0(v,0)+α2,0u12+α1,2u1u22+α0,4u24+h1(u1,u2),\displaystyle=\phi_{x_{0}}(v,0)+\alpha_{2,0}u_{1}^{2}+\alpha_{1,2}u_{1}u_{2}^{2}+\alpha_{0,4}u_{2}^{4}+h_{1}(u_{1},u_{2}), (4.2)
ϕx0(v,u)\displaystyle\phi_{x_{0}}(v,u) =ϕx0(v,0)+β2,0u12+β0,3u23+h2(u1,u2),\displaystyle=\phi_{x_{0}}(v,0)+\beta_{2,0}u_{1}^{2}+\beta_{0,3}u_{2}^{3}+h_{2}(u_{1},u_{2}), (4.3)

where u=(u1,u2)u=(u_{1},u_{2}), h1(12,14),2h_{1}\in\mathcal{H}_{(\frac{1}{2},\frac{1}{4}),2}, h2(12,13),2h_{2}\in\mathcal{H}_{(\frac{1}{2},\frac{1}{3}),2}. Moreover, α2,0u12+α1,2u1u22+α0,4u24\alpha_{2,0}u_{1}^{2}+\alpha_{1,2}u_{1}u_{2}^{2}+\alpha_{0,4}u_{2}^{4} is \mathbb{R}-nondegenerate in (4.2), and β2,0,β0,30\beta_{2,0},\beta_{0,3}\neq 0 in (4.3)

By the definition (4.1), α2,0u12+α1,2u1u22+α0,4u24(12,14),2\alpha_{2,0}u_{1}^{2}+\alpha_{1,2}u_{1}u_{2}^{2}+\alpha_{0,4}u_{2}^{4}\in\mathcal{E}_{(\frac{1}{2},\frac{1}{4}),2} and β2,0u12+β0,3u23(12,13),2\beta_{2,0}u_{1}^{2}+\beta_{0,3}u_{2}^{3}\in\mathcal{E}_{(\frac{1}{2},\frac{1}{3}),2}. Therefore, we only need to estimate the oscillatory integrals whose phase functions are α2,0u12+α1,2u1u22+α0,4u24\alpha_{2,0}u_{1}^{2}+\alpha_{1,2}u_{1}u_{2}^{2}+\alpha_{0,4}u_{2}^{4} or β2,0u12+β0,3u23\beta_{2,0}u_{1}^{2}+\beta_{0,3}u_{2}^{3}. By Lemma 2.5 and 3.2, the maximal exponential pairs of the \mathbb{R}-nondegenerate polynomials are listed as below.

ff (β(f),p(f))(\beta(f),p(f))
α2,0u12+α1,2u1u22+α0,4u24\alpha_{2,0}u_{1}^{2}+\alpha_{1,2}u_{1}u_{2}^{2}+\alpha_{0,4}u_{2}^{4} (34,0)(-\frac{3}{4},0)
β2,0u12+β0,3u23\beta_{2,0}u_{1}^{2}+\beta_{0,3}u_{2}^{3} (56,0)(-\frac{5}{6},0)
Lemma 4.4.

Let 0 be an singularity of phase functions ϕ1\phi_{1} and ϕ2\phi_{2} with the forms (4.2) and (4.3) respectively, then the following uniform estimates hold.

M(ϕ1,0)\displaystyle M(\phi_{1},0) (34,0),\displaystyle\curlyeqprec(-\frac{3}{4},0), (4.4)
M(ϕ2,0)\displaystyle M(\phi_{2},0) (56,0).\displaystyle\curlyeqprec(-\frac{5}{6},0). (4.5)
Proof of Lemma 1.1.

This Lemma can be obtained by Lemma 4.2, 4.3 and 4.4 immediately. ∎

Now we begin to proof Lemma 4.3. Since we only need to analyze all possible singularities, in later arguments, we always assume vg(Γ)v\in\nabla g(\Gamma) and xΓ(v).x\in\Gamma(v).

Since the Hessian matrix Hessxϕ(v,x)\mathrm{Hess}_{x}\phi(v,x) is degenerate at xx, by the previous arguments, its rank is 1.1. Therefore, without loss of generality, we can assume the following equation holds

cos(x1+x2)=λ(cos(x1+x2)+cos(x1)),cos(x1+x2)+cos(x2)=λcos(x1+x2).\cos(x_{1}+x_{2})=\lambda(\cos(x_{1}+x_{2})+\cos(x_{1})),\cos(x_{1}+x_{2})+\cos(x_{2})=\lambda\cos(x_{1}+x_{2}).

where λ.\lambda\in\mathbb{R}. And a direct calculation shows that the degeneration of the Hessian is equivalent to

det(Hessxϕ(v,x))=(cosx1+cosx2)cos(x1+x2)+cosx1cosx2=0.\displaystyle\mathrm{det}(\mathrm{Hess}_{x}\phi(v,x))=(\cos x_{1}+\cos x_{2})\cos(x_{1}+x_{2})+\cos x_{1}\cos x_{2}=0. (4.6)
Proof of Lemma 4.3.

We divide our argument into four cases.

(A)cosx1=cosx2=0,(B)λ=0,(C)λ>0and(D)λ<0.(\mathrm{A})\cos{x_{1}}=\cos{x_{2}}=0,~{}(\mathrm{B})\lambda=0,~{}(\mathrm{C})\lambda>0~{}\text{and}~{}(\mathrm{D})\lambda<0.

Here in (B)(D)(\mathrm{B})\sim(\mathrm{D}) we assume that cosx10.\cos x_{1}\neq 0.

  1. (A)

    When cosx1=cosx2=0,\cos{x_{1}}=\cos{x_{2}}=0, then sin(x1+x2)=0.\sin(x_{1}+x_{2})=0. Then the function ϕx(v,x)=ϕ(v,x+x)\phi_{x}(v,x^{\prime})=\phi(v,x+x^{\prime}) has the Taylor expansion

    ϕx(v,x)\displaystyle\phi_{x}(v,x^{\prime}) ϕx(v,0)+cos(x1+x2)(x1+x2)213(sinx1(x1)3+sinx2(x2)3)\displaystyle\sim\phi_{x}(v,0)+\cos(x_{1}+x_{2})(x_{1}^{\prime}+x_{2}^{\prime})^{2}-\frac{1}{3}(\sin x_{1}(x_{1}^{\prime})^{3}+\sin x_{2}(x_{2}^{\prime})^{3})
    112cos(x1+x2)(x1+x2)4+\displaystyle-\frac{1}{12}\cos(x_{1}+x_{2})(x_{1}^{\prime}+x_{2}^{\prime})^{4}+\cdots

    Set u=(u1,u2)u=(u_{1},u_{2}), where u1=x1+x2u_{1}=x_{1}^{\prime}+x_{2}^{\prime} and u2=x2.u_{2}=x_{2}^{\prime}. Then

    ϕx(v,u)\displaystyle\phi_{x}(v,u) ϕx(v,0)+cos(x1+x2)u12\displaystyle\sim\phi_{x}(v,0)+\cos(x_{1}+x_{2})u_{1}^{2}
    13(sinx1u133sinx1u12u2+3sinx1u1u22+(sinx2sinx1)u23)\displaystyle-\frac{1}{3}(\sin x_{1}u_{1}^{3}-3\sin x_{1}u_{1}^{2}u_{2}+3\sin x_{1}u_{1}u_{2}^{2}+(\sin x_{2}-\sin x_{1})u_{2}^{3})
    112cos(x1+x2)u14+\displaystyle-\frac{1}{12}\cos(x_{1}+x_{2})u_{1}^{4}+\cdots

    When sinx1sinx2\sin x_{1}\neq\sin x_{2}, 0 is the singularity with the form (4.2). When sinx1=sinx2\sin x_{1}=\sin x_{2}, 0 is the singularity with the form (4.3), where α2,0=±1,α1,2=±1\alpha_{2,0}=\pm 1,~{}\alpha_{1,2}=\pm 1 and α0,4=0.\alpha_{0,4}=0. Then it is a \mathbb{R}-nondegenerate polynomial.

  2. (B)

    Now we consider the second case where λ=0.\lambda=0. In this case, we have cosx2=cos(x1+x2)=0,\cos x_{2}=\cos(x_{1}+x_{2})=0, and

    Hessxϕ(v,x)=2(cosx1000).\displaystyle\mathrm{Hess}_{x}\phi(v,x)=2\begin{pmatrix}\cos x_{1}&0\\ 0&0\end{pmatrix}.

    In this case,

    ϕx(v,x)\displaystyle\phi_{x}(v,x^{\prime}) ϕx(v,0)+cosx1(x1)213(sinx2(x2)3+sin(x1+x2)(x1+x2)3)\displaystyle\sim\phi_{x}(v,0)+\cos x_{1}(x_{1}^{\prime})^{2}-\frac{1}{3}(\sin{x_{2}}(x_{2}^{\prime})^{3}+\sin{(x_{1}+x_{2})}(x_{1}^{\prime}+x_{2}^{\prime})^{3})
    112cosx1(x1)4+,\displaystyle-\frac{1}{12}\cos x_{1}(x_{1}^{\prime})^{4}+\cdots,

    where |cosx1|=1.|\cos{x_{1}}|=1. So when sinx2+sin(x1+x2)0\sin x_{2}+\sin(x_{1}+x_{2})\neq 0, 0 is the singularity with the form (4.2), otherwise 0 is the singularity with the form (4.3).

  3. (C)

    Now let us come to the third case, where λ>0\lambda>0. In this case, we have the following equations.

    cos(x1)\displaystyle\cos(x_{1}) =1λλcos(x1+x2),\displaystyle=\frac{1-\lambda}{\lambda}\cos(x_{1}+x_{2}), (4.7)
    cos(x2)\displaystyle\cos(x_{2}) =(λ1)cos(x1+x2),\displaystyle=(\lambda-1)\cos(x_{1}+x_{2}), (4.8)
    λ\displaystyle\lambda =cos(x2)cos(x1).\displaystyle=-\frac{\cos(x_{2})}{\cos(x_{1})}. (4.9)

    Similarly for the function ϕx(v,x)=ϕ(v,x+x)\phi_{x}(v,x^{\prime})=\phi(v,x+x^{\prime}), one have the following Taylor expansion.

    ϕx(v,x)ϕx(v,0)+cos(x1+x2)(1λx1+λx2)2\displaystyle\phi_{x}(v,x^{\prime})\sim\phi_{x}(v,0)+\cos{(x_{1}+x_{2})}(\frac{1}{\sqrt{\lambda}}x_{1}^{\prime}+\sqrt{\lambda}x_{2}^{\prime})^{2}
    13[sin(x1+x2)(x1+x2)3+sinx1(x1)3+sinx2(x2)3]\displaystyle-\frac{1}{3}\big{[}\sin(x_{1}+x_{2})(x_{1}^{\prime}+x_{2}^{\prime})^{3}+\sin{x_{1}}(x_{1}^{\prime})^{3}+\sin{x_{2}}(x_{2}^{\prime})^{3}\big{]}
    112[cos(x1+x2)(x1+x2)4+cosx1(x1)4+cosx2(x2)4]+\displaystyle-\frac{1}{12}\big{[}\cos{(x_{1}+x_{2})}(x_{1}^{\prime}+x_{2}^{\prime})^{4}+\cos{x_{1}}(x_{1}^{\prime})^{4}+\cos{x_{2}}(x_{2}^{\prime})^{4}\big{]}+\cdots

    By change of variables, set u=(u1,u2)u=(u_{1},u_{2}), where u1=1λx1+λx2u_{1}=\frac{1}{\sqrt{\lambda}}x_{1}^{\prime}+\sqrt{\lambda}x_{2}^{\prime} and u2=x2u_{2}=x_{2}^{\prime}, we obtain

    ϕx(v,u)ϕx(v,0)+cos(x1+x2)u12\displaystyle\phi_{x}(v,u)\sim\phi_{x}(v,0)+\cos{(x_{1}+x_{2})}u_{1}^{2}
    13[sin(x1+x2)(λu1+(1λ)u2)3+sinx1(λu1λu2)3+(sinx2)u23]\displaystyle-\frac{1}{3}\big{[}\sin{(x_{1}+x_{2})}(\sqrt{\lambda}u_{1}+(1-\lambda)u_{2})^{3}+\sin{x_{1}}(\sqrt{\lambda}u_{1}-\lambda u_{2})^{3}+(\sin{x_{2}})u_{2}^{3}\big{]}
    112[cos(x1+x2)(λu1+(1λ)u2)4+cosx1(λu1λu2)4+(cosx2)u24]\displaystyle-\frac{1}{12}\big{[}\cos{(x_{1}+x_{2})}(\sqrt{\lambda}u_{1}+(1-\lambda)u_{2})^{4}+\cos{x_{1}}(\sqrt{\lambda}u_{1}-\lambda u_{2})^{4}+(\cos{x_{2}})u_{2}^{4}\big{]}
    +\displaystyle+\cdots (4.10)

    We denote by α2,0,α1,2,α0,3\alpha_{2,0},\alpha_{1,2},\alpha_{0,3} and α0,4\alpha_{0,4} the coefficients of monomials u12,u1u22,u23u_{1}^{2},u_{1}u_{2}^{2},u_{2}^{3} and u24u_{2}^{4} in (4.10) respectively. Then we have

    α2,0=cos(x1+x2),\displaystyle\alpha_{2,0}=\cos{(x_{1}+x_{2})},
    α1,2=sin(x1+x2)λ(1λ)2sinx1λλ2,\displaystyle\alpha_{1,2}=-\sin(x_{1}+x_{2})\sqrt{\lambda}(1-\lambda)^{2}-\sin{x_{1}}\sqrt{\lambda}\lambda^{2},
    α0,3=13[sin(x1+x2)(1λ)3sinx1λ3+sinx2],\displaystyle\alpha_{0,3}=-\frac{1}{3}\big{[}\sin{(x_{1}+x_{2})}(1-\lambda)^{3}-\sin{x_{1}}\lambda^{3}+\sin{x_{2}}\big{]},
    α0,4=112[cos(x1+x2)(1λ)4+cosx1λ4+cosx2].\displaystyle\alpha_{0,4}=-\frac{1}{12}\big{[}\cos{(x_{1}+x_{2})}(1-\lambda)^{4}+\cos{x_{1}}\lambda^{4}+\cos{x_{2}}\big{]}.

    Let Σ0\Sigma_{0} be solutions to the equation (4.6). Moreover, we denote by Σ1\Sigma_{1} and Σ2\Sigma_{2} subsets of 2\mathbb{R}^{2} where α0,3=0\alpha_{0,3}=0 and α2,0u12+α1,2u1u22+α0,4u24\alpha_{2,0}u_{1}^{2}+\alpha_{1,2}u_{1}u_{2}^{2}+\alpha_{0,4}u_{2}^{4} is \mathbb{R}-degenerate respectively. Then the assertions in Lemma 4.4 in Case (C) hold if Σ0Σ1Σ2=\Sigma_{0}\cap\Sigma_{1}\cap\Sigma_{2}=\emptyset for λ>0.\lambda>0. Firstly, we consider the set Σ1\Sigma_{1}. Suppose that the coefficient α0,3=0\alpha_{0,3}=0. Then we have

    (1λ)3sin(x1+x2)λ3sinx1+sinx2=0.\displaystyle(1-\lambda)^{3}\sin{(x_{1}+x_{2})}-\lambda^{3}\sin{x_{1}}+\sin{x_{2}}=0. (4.11)

    By equations (4.9) and (4.11), we have

    (cosx1+cosx2)3sin(x1+x2)+cos3x2sinx1+cos3x1sinx2=0,\displaystyle(\cos{x_{1}}+\cos{x_{2}})^{3}\sin{(x_{1}+x_{2})}+\cos^{3}x_{2}\sin x_{1}+\cos^{3}x_{1}\sin x_{2}=0,

    which leads to

    sin(x1+x2)[(cosx1+cosx2)3+cos2x1+cos2x2cosx1cosx2cos(x1x2)]=0.\displaystyle\sin{(x_{1}+x_{2})}\big{[}(\cos{x_{1}}+\cos{x_{2}})^{3}+\cos^{2}x_{1}+\cos^{2}x_{2}-\cos{x_{1}}\cos{x_{2}}\cos(x_{1}-x_{2})\big{]}=0. (4.12)

    Therefore, Σ1=Σ11Σ12\Sigma_{1}=\Sigma_{1}^{1}\cup\Sigma_{1}^{2}, where

    Σ11={(x1,x2)Γ(v):x1+x2=kπ,for somek},\Sigma_{1}^{1}=\{(x_{1},x_{2})\in\Gamma(v):x_{1}+x_{2}=k\pi,~{}\text{for some}~{}k\in\mathbb{Z}\},

    and Σ12\Sigma_{1}^{2} is the solution to

    (cosx1+cosx2)3+cos2x1+cos2x2cosx1cosx2cos(x1x2)=0.\displaystyle(\cos{x_{1}}+\cos{x_{2}})^{3}+\cos^{2}x_{1}+\cos^{2}x_{2}-\cos{x_{1}}\cos{x_{2}}\cos(x_{1}-x_{2})=0. (4.13)

    So if xΣ0Σ11x\in\Sigma_{0}\cap\Sigma_{1}^{1}, by equation (4.6) we have cosx2=0,\cos{x_{2}}=0, which will be attributed to the cases (A) or (B). Therefore, we only need to analyze the set Σ0Σ12Σ2.\Sigma_{0}\cap\Sigma_{1}^{2}\cap\Sigma_{2}.

    Suppose that the polynomial α2,0u12+α1,2u1u22+α0,4u24\alpha_{2,0}u_{1}^{2}+\alpha_{1,2}u_{1}u_{2}^{2}+\alpha_{0,4}u_{2}^{4} is \mathbb{R}-degenerate, then its discriminant is equal to 0, that is

    α1,224α2,0α0,4=0.\displaystyle\alpha_{1,2}^{2}-4\alpha_{2,0}\alpha_{0,4}=0. (4.14)

    Since

    α2,0α0,4=cos(x1+x2)12[cos(x1+x2)(1λ)4+cosx1λ4+cosx2],\displaystyle\alpha_{2,0}\alpha_{0,4}=-\frac{\cos{(x_{1}+x_{2})}}{12}\big{[}\cos{(x_{1}+x_{2})}(1-\lambda)^{4}+\cos{x_{1}}\lambda^{4}+\cos{x_{2}}\ \big{]},

    by equations (4.7) and (4.8),

    α2,0α0,4=cos2(x1+x2)4λ(1λ)2.\displaystyle\alpha_{2,0}\alpha_{0,4}=\frac{\cos^{2}(x_{1}+x_{2})}{4}\lambda(1-\lambda)^{2}. (4.15)

    Therefore, by (4.14) and (4.15) we have

    cos2(x1+x2)4λ(1λ)2=14[sin(x1+x2)λ(1λ)2+sinx1λ52]2.\displaystyle\frac{\cos^{2}(x_{1}+x_{2})}{4}\lambda(1-\lambda)^{2}=\frac{1}{4}\big{[}\sin(x_{1}+x_{2})\sqrt{\lambda}(1-\lambda)^{2}+\sin{x_{1}}\lambda^{\frac{5}{2}}\big{]}^{2}. (4.16)

    By equations (4.6), (4.9) and (4.16), we have

    cos2x1cos2x2=[sin(x1+x2)(cosx1+cosx2)2cosx1+sinx1cos2x2cosx1]2.\displaystyle\cos^{2}{x_{1}}\cos^{2}x_{2}=\big{[}\sin{(x_{1}+x_{2})}\frac{(\cos x_{1}+\cos{x_{2}})^{2}}{\cos{x_{1}}}+\sin{x_{1}}\frac{\cos^{2}{x_{2}}}{\cos{x_{1}}}\big{]}^{2}. (4.17)

    Let xΣ0Σ2x\in\Sigma_{0}\cap\Sigma_{2}, then by (4.12) and (4.17), we have

    cos2x1cos2x2(cosx1+cosx2)2=(cos2x2sinx1cos2x1sinx2)2.\displaystyle\cos^{2}{x_{1}}\cos^{2}{x_{2}}(\cos{x_{1}}+\cos{x_{2}})^{2}=(\cos^{2}x_{2}\sin{x_{1}}-\cos^{2}{x_{1}}\sin{x_{2}})^{2}. (4.18)

    By direct calculation, (4.18) is equivalent to the equation

    cos2x1+x22[cos2x1cos2x2cos2x1x22sin2x1x22(cos2x1x22+sin2x1+x22)2]=0.\displaystyle\cos^{2}{\frac{x_{1}+x_{2}}{2}}[\cos^{2}{x_{1}}\cos^{2}{x_{2}}\cos^{2}{\frac{x_{1}-x_{2}}{2}}-\sin^{2}{\frac{x_{1}-x_{2}}{2}}(\cos^{2}{\frac{x_{1}-x_{2}}{2}}+\sin^{2}{\frac{x_{1}+x_{2}}{2}})^{2}]=0.

    Since in case (C), cosx1+x220,\cos{\frac{x_{1}+x_{2}}{2}}\neq 0, we denote by by Σ2\Sigma_{2}^{\prime} the solution to

    cos2x1cos2x2cos2x1x22sin2x1x22(cos2x1x22+sin2x1+x22)2=0.\displaystyle\cos^{2}{x_{1}}\cos^{2}{x_{2}}\cos^{2}{\frac{x_{1}-x_{2}}{2}}-\sin^{2}{\frac{x_{1}-x_{2}}{2}}(\cos^{2}{\frac{x_{1}-x_{2}}{2}}+\sin^{2}{\frac{x_{1}+x_{2}}{2}})^{2}=0. (4.19)

    Then sets Σ0,Σ12\Sigma_{0},\Sigma_{1}^{2} and Σ2\Sigma_{2}^{\prime} are shown in Figure C, with curves colored in red, green and black respectively.

    [Uncaptioned image]\captionof

    figureCurves Σ0,Σ12\Sigma_{0},\Sigma_{1}^{2} and Σ2\Sigma_{2}^{\prime}.

    The intersection of three curves are empty whenever λ=cosx2cosx1>0\lambda=\frac{\cos{x_{2}}}{\cos{x_{1}}}>0 as shown in Figure C. Since the proof consists of lengthy calculations, we will leave it to the Proposition 6.1 in Appendix 6.

  4. (D)

    Finally, we consider the case when λ<0.\lambda<0. Similarly, let ϕx(v,x)=ϕ(v,x+x)\phi_{x}(v,x^{\prime})=\phi(v,x+x^{\prime}), we have

    ϕx(v,x)ϕx(v,0)cos(x1+x2)(1λx1λx2)2\displaystyle\phi_{x}(v,x^{\prime})\sim\phi_{x}(v,0)-\cos{(x_{1}+x_{2})}(\frac{1}{\sqrt{-\lambda}}x_{1}^{\prime}-\sqrt{-\lambda}x_{2}^{\prime})^{2}
    13[sin(x1+x2)(x1+x2)3+sinx1(x1)3+sinx2(x2)3]\displaystyle-\frac{1}{3}\big{[}\sin{(x_{1}+x_{2})}(x_{1}^{\prime}+x_{2}^{\prime})^{3}+\sin{x_{1}}(x_{1}^{\prime})^{3}+\sin{x_{2}}(x_{2}^{\prime})^{3}\big{]}
    112[cos(x1+x2)(x1+x2)4+cosx1(x1)4+cosx2(x2)4+\displaystyle-\frac{1}{12}[\cos(x_{1}+x_{2})(x_{1}^{\prime}+x_{2}^{\prime})^{4}+\cos{x_{1}}(x_{1}^{\prime})^{4}+\cos_{x_{2}}(x_{2}^{\prime})^{4}+\cdots
    =ϕx(v,0)cos(x1+x2)u12\displaystyle=\phi_{x}(v,0)-\cos{(x_{1}+x_{2})}u_{1}^{2}
    13[sin(x1+x2)(λu1+(1λ)u2)3+sinx1(λu1λu2)3+sinx2u23]\displaystyle-\frac{1}{3}\big{[}\sin{(x_{1}+x_{2})}(\sqrt{-\lambda}u_{1}+(1-\lambda)u_{2})^{3}+\sin{x_{1}}(\sqrt{-\lambda}u_{1}-\lambda u_{2})^{3}+\sin{x_{2}}u_{2}^{3}\big{]}
    112[cos(x1+x2)(λu1+(1λ)u2)4+cosx1(λu1λu2)4+cosx2u24]+\displaystyle-\frac{1}{12}[\cos(x_{1}+x_{2})(\sqrt{-\lambda}u_{1}+(1-\lambda)u_{2})^{4}+\cos{x_{1}}(\sqrt{-\lambda}u_{1}-\lambda u_{2})^{4}+\cos{x_{2}}u_{2}^{4}]+\cdots~{}

    where u1=1λx1λx2u_{1}=\frac{1}{\sqrt{-\lambda}}x_{1}^{\prime}-\sqrt{-\lambda}x_{2}^{\prime} and u2=x2u_{2}=x_{2}^{\prime}. Since all the analyses are similar to the case (C), we omit the detailed proof here.

5 Strichartz estimate on hexagonal triangulation and its application

To prove the Strichartz estimate on the hexagonal triangulation, we need the well-known result of Keel and Tao.

Theorem 5.1 (Keel and Tao [17]).

Let HH be a Hilbert space, (X,dx)(X,\mathrm{d}x) be a measure space, and U(t):HL2(X)U(t):H\to L^{2}(X) be a one-parameter family of mappings, which has the energy estimate

U(t)fLx2CfH,\|U(t)f\|_{L_{x}^{2}}\leq C\|f\|_{H},

and the decay estimate

U(t)(U(S))gL(X)CtsσgL1(X),\|U(t)(U(S))^{*}g\|_{L^{\infty}(X)}\leq C\left\langle t-s\right\rangle^{-\sigma}\|g\|_{L^{1}(X)},

for some σ>0\sigma>0. Then,

U(t)fLqLrCfL2,\displaystyle\|U(t)f\|_{L^{q}L^{r}}\leq C\|f\|_{L^{2}},
(U(t))F(t,)dtL2CFLtqLxr,\displaystyle\|\int(U(t))^{*}F(t,\cdot)\mathrm{d}t\|_{L^{2}}\leq C\|F\|_{L_{t}^{q^{\prime}}L_{x}^{r^{\prime}}},
0tU(t)(U(s))F(s,)dsL2CFLtq~Lxr~,\displaystyle\|\int_{0}^{t}U(t)(U(s))^{*}F(s,\cdot)\mathrm{d}s\|_{L^{2}}\leq C\|F\|_{L_{t}^{\tilde{q}^{\prime}}L_{x}^{\tilde{r}^{\prime}}},

where exponent pairs (q,r,σ)(q,r,\sigma) and (q~,r~,σ)(\tilde{q},\tilde{r},\sigma) do not equal to (2,,1)(2,\infty,1), and

1q+σrσ2.\frac{1}{q}+\frac{\sigma}{r}\leq\frac{\sigma}{2}.

Now we prove the decay estimate (1.11) of (LABEL:LS).

Proof of Theorem 1.3.

With the help of Theorem 5.1 and Lemma 1.3, we can prove the Theorem 1.3 immediately. ∎

With the help of Theorem 1.3, we can establish the global existence of the solution to discrete nonlinear Schrödinger equations (DNLS) with a small initial value on the hexagonal triangulation 𝒯H.\mathcal{T}_{H}.

{itu(v,t)+ΔGHu(v,t)+|u(v,t)|2σu(v,t)=0,(v,t)2×[0,)u(,0)=ψl2(2).\displaystyle\left\{\begin{aligned} &i\partial_{t}u(v,t)+\Delta_{G_{H}}u(v,t)+|u(v,t)|^{2\sigma}u(v,t)=0,~{}~{}\forall(v,t)\in\mathbb{Z}^{2}\times[0,\infty)\\ &u(\cdot,0)=\psi\in l^{2}(\mathbb{Z}^{2}).\end{aligned}\right. (5.1)
Theorem 5.2.

Set σ116\sigma\geq\frac{11}{6}. Then there exists ϵ>0\epsilon>0 and a positive constant CC such that if ψl2ϵ\|\psi\|_{l^{2}}\leq\epsilon, then (5.1) has a global solution uC1((0,];l2(2))u\in C^{1}((0,\infty];l^{2}(\mathbb{Z}^{2})). Moreover,

ψLqlrCϵ,\|\psi\|_{L^{q}l^{r}}\leq C\epsilon,

given that

1q+34r38.\frac{1}{q}+\frac{3}{4r}\leq\frac{3}{8}.

As a corollary, we have

limtu(t)lr=0,r>2.\lim_{t\rightarrow\infty}\|u(t)\|_{l^{r}}=0,~{}~{}\forall r>2.
Proof.

Define a metric space

={u(t):supq,ruLqlr<2Cψl2,where1q+34r38},\mathcal{M}=\{u(t):\sup_{q,r}\|u\|_{L^{q}l^{r}}<2C\|\psi\|_{l^{2}},~{}\text{where}~{}\frac{1}{q}+\frac{3}{4r}\leq\frac{3}{8}\},

where CC is the constant in Strichartz estimate 1.3. Let LL be an operator such that

L(u)=eitΔGHψ+i0tei(ts)ΔGH|u|2σu(s)ds.L(u)=e^{it\Delta_{G_{H}}}\psi+i\int_{0}^{t}e^{i(t-s)\Delta_{G_{H}}}|u|^{2\sigma}u(s)\mathrm{d}s.

Let the exponent pair (q~,r~)=(,2),(\tilde{q},\tilde{r})=(\infty,2), by (1.12) we have

sup(q,r):1q+34r38Lu(t)LqlrCψl2+Cu2σ+1L1l2,\displaystyle\sup_{(q,r):\frac{1}{q}+\frac{3}{4r}\leq\frac{3}{8}}\|Lu(t)\|_{L^{q}l^{r}}\leq C\|\psi\|_{l^{2}}+C\|u^{2\sigma+1}\|_{L^{1}l^{2}},
Cψl2+CuL2σ+1l2(2σ+1)2σ+1.\displaystyle\leq C\|\psi\|_{l^{2}}+C\|u\|^{2\sigma+1}_{L^{2\sigma+1}l^{2(2\sigma+1)}}.

Since σ116\sigma\geq\frac{11}{6}, we have

12σ+1+34(2σ+1)38.\frac{1}{2\sigma+1}+\frac{3}{4(2\sigma+1)}\leq\frac{3}{8}.

Then we obtain

sup(q,r):1q+34r38Lu(t)Lqlr\displaystyle\sup_{(q,r):\frac{1}{q}+\frac{3}{4r}\leq\frac{3}{8}}\|Lu(t)\|_{L^{q}l^{r}} Cψl2+Cu2σ+1\displaystyle\leq C\|\psi\|_{l^{2}}+C\|u\|_{\mathcal{M}}^{2\sigma+1}
Cψl2+C(2Cϵ)2σ+1.\displaystyle\leq C\|\psi\|_{l^{2}}+C(2C\epsilon)^{2\sigma+1}.

Let ϵ\epsilon be small enough, then we have

sup(q,r):1q+34r38Lu(t)Lqlr2Cψl2.\sup_{(q,r):\frac{1}{q}+\frac{3}{4r}\leq\frac{3}{8}}\|Lu(t)\|_{L^{q}l^{r}}\leq 2C\|\psi\|_{l^{2}}.

Therefore, the image L()L(\mathcal{M}) is contained in \mathcal{M} given that ψ\|\psi\| is small enough. Let u,u^u,\hat{u}\in\mathcal{M}, then with the same initial data u(,0)=u^(,0)=ψ,u(\cdot,0)=\hat{u}(\cdot,0)=\psi,

sup(q,r):1q+34r38LuLu^\displaystyle\sup_{(q,r):\frac{1}{q}+\frac{3}{4r}\leq\frac{3}{8}}\|Lu-L\hat{u}\| C0uu^L2σ+1l2(2σ+1)(uL2σ+1l2(2σ+1)+u^L2σ+1l2(2σ+1)),\displaystyle\leq C_{0}\|u-\hat{u}\|_{L^{2\sigma+1}l^{2(2\sigma+1)}}(\|u\|_{L^{2\sigma+1}l^{2(2\sigma+1)}}+\|\hat{u}\|_{L^{2\sigma+1}l^{2(2\sigma+1)}}),
2CC1uu^L2σ+1l2(2σ+1)ψl22σ.\displaystyle\leq 2CC_{1}\|u-\hat{u}\|_{L^{2\sigma+1}l^{2(2\sigma+1)}}\|\psi\|_{l^{2}}^{2\sigma}.

Therefore LL is a contraction map given that ψ\psi is sufficiently small. Thus we obtain a global solution of (5.1). ∎

6 Appendix

Proposition 6.1.

Let Σ12\Sigma_{1}^{2} and Σ2\Sigma_{2}^{\prime} be the curves defined in Section 4, then Σ12Σ2={(k1π+π2,k2π+π2):k1,k2}\Sigma_{1}^{2}\cap\Sigma_{2}^{\prime}=\{(k_{1}\pi+\frac{\pi}{2},k_{2}\pi+\frac{\pi}{2}):k_{1},k_{2}\in\mathbb{Z}\}.

Proof.

Let A=x1x22A=\frac{x_{1}-x_{2}}{2} and B=x1+x22B=\frac{x_{1}+x_{2}}{2}. Then by the change of variables, we can rewrite (4.13) and (4.19) as

8cos3Acos3B6cos2Acos2B2cos4Bcos2A+cos2B+1=0,\displaystyle 8\cos^{3}A\cos^{3}B-6\cos^{2}A\cos^{2}B-2\cos^{4}B-\cos^{2}A+\cos^{2}B+1=0, (6.1)

and

cos2B(cos2A+cos2B1)2=sin2B(cos2Bcos2A+1)2,\displaystyle\cos^{2}B(\cos^{2}A+\cos^{2}B-1)^{2}=\sin^{2}B(\cos^{2}B-\cos^{2}A+1)^{2}, (6.2)

respectively. When cos2B=12,\cos^{2}B=\frac{1}{2}, by equation (6.2) we have cos2A=1,\cos^{2}A=1, which means (x1,x2)Σ0Σ11.(x_{1},x_{2})\in\Sigma_{0}\cap\Sigma_{1}^{1}. Then by equation (4.6), the case will be attributed to Case (A) or Case (B). So without loss of generality, we assume that cos2B12\cos^{2}B\neq\frac{1}{2}. By direct calculation, we have

sin2A=cos2B(1±|sin2B|)2cos2B1\displaystyle\sin^{2}A=\frac{\cos^{2}B(1\pm|\sin 2B|)}{2\cos^{2}B-1}

Since sin2A[0,1]\sin^{2}A\in[0,1],

sin2A=cos2B(1|sin2B|)2cos2B1\displaystyle\sin^{2}A=\frac{\cos^{2}B(1-|\sin 2B|)}{2\cos^{2}B-1} (6.3)

Due to periodicity, we assume that (A,B)(π2,3π2)×(π2,π2).(A,B)\in(-\frac{\pi}{2},\frac{3\pi}{2})\times(-\frac{\pi}{2},\frac{\pi}{2}). For (A,B)(π2,π2)2(A,B)\in(-\frac{\pi}{2},\frac{\pi}{2})^{2}, we have (x1,x2)(π,π)2.(x_{1},x_{2})\in(-\pi,\pi)^{2}. Since A(π2,π2)A\in(-\frac{\pi}{2},\frac{\pi}{2}), cosx1+cosx2>0.\cos{x_{1}}+\cos{x_{2}}>0. Therefore, cos2x1+cos2x2cosx1cosx2cos(x1x2)>0,\cos^{2}x_{1}+\cos^{2}x_{2}-\cos{x_{1}}\cos{x_{2}}\cos(x_{1}-x_{2})>0, so the equation (4.13) does not have a solution in this case. Therefore, we only need to consider the case when (A,B)(π2,3π2)×(π2,π2).(A,B)\in(\frac{\pi}{2},\frac{3\pi}{2})\times(-\frac{\pi}{2},\frac{\pi}{2}). Now we prove that if the equation (6.3) holds, then

8cos3Acos3B6cos2Acos2B2cos4Bcos2A+cos2B+1<0.8\cos^{3}A\cos^{3}B-6\cos^{2}A\cos^{2}B-2\cos^{4}B-\cos^{2}A+\cos^{2}B+1<0.

We divide our proof into two parts.

  1. (I)

    Firstly, we consider the case when cos2B>12.\cos^{2}B>\frac{1}{2}. Since cosA<0,\cos A<0, we only need to prove

    2cos4Bcos2A+cos2B+1=cos2B(12cos2B)+sin2A<0\displaystyle-2\cos^{4}B-\cos^{2}A+\cos^{2}B+1=\cos^{2}B(1-2\cos^{2}B)+\sin^{2}A<0 (6.4)

    By (6.3), we have

    cos2B(12cos2B)+sin2A=cos2B(sin22B|sin2B|)2cos2B1<0.\displaystyle\cos^{2}B(1-2\cos^{2}B)+\sin^{2}A=\frac{\cos^{2}B(\sin^{2}2B-|\sin 2B|)}{2\cos^{2}B-1}<0.

    Therefore, we finish the proof of case (I).

  2. (II)

    When cos2B<12,\cos^{2}B<\frac{1}{2}, by direct calculation,

    2cos2Acos2B2cos4Bcos2A+cos2B+1\displaystyle-2\cos^{2}A\cos^{2}B-2\cos^{4}B-\cos^{2}A+\cos^{2}B+1
    =cos2B(sin22B|sin2B|+2sin2B2cos2B|sin2B|)2cos2B1.\displaystyle=\frac{\cos^{2}B(\sin^{2}2B-|\sin 2B|+2\sin^{2}B-2\cos^{2}B|\sin 2B|)}{2\cos^{2}B-1}.

    Since cos2B<12\cos^{2}B<\frac{1}{2}, we have

    2sin2B|sin2B|=2|sinB|(|sinB||cosB|)>0.2\sin^{2}B-|\sin 2B|=2|\sin B|(|\sin B|-|\cos B|)>0.

    Since sin2B>12,\sin^{2}B>\frac{1}{2}, we have

    sin22B2cos2B|sin2B|=2cos22B(2sin2B|sin2B|)>0.\sin^{2}2B-2\cos^{2}B|\sin 2B|=2\cos^{2}2B(2\sin^{2}B-|\sin 2B|)>0.

    Therefore, we have

    2cos2Acos2B2cos4Bcos2A+cos2B+1<0.-2\cos^{2}A\cos^{2}B-2\cos^{4}B-\cos^{2}A+\cos^{2}B+1<0.

    Thus we finish the proof.

Acknowledgements. H. Ge is supported by NSFC, no.12341102, no.12122119. B. Hua is supported by NSFC, no.12371056, and by Shanghai Science and Technology Program [Project No. 22JC1400100]. The authors would like to thank Jiawei Cheng for helpful advice.

References

  • [1] R. Adami, C. Cacciapuoti, D. Finco, and D. Noja. Fast solitons on star graphs. Rev. Math. Phys., 23(4):409–451, 2011.
  • [2] K. Ammari and H. Bouzidi. Exact boundary controllability of the linear biharmonic Schrödinger equation with variable coefficients. J. Dyn. Control Syst., 29(3):703–719, 2023.
  • [3] K. Ammari and M. Sabri. Dispersion on certain Cartesian products of graphs. In Control and inverse problems, Trends Math., pages 217–222. Birkhäuser/Springer, Cham, [2023] ©2023.
  • [4] D. Bambusi. Asymptotic stability of breathers in some Hamiltonian networks of weakly coupled oscillators. Comm. Math. Phys., 324(2):515–547, 2013.
  • [5] V. Banica and L. Ignat. Dispersion for the schrödinger equation on the line with multiple dirac delta potentials and on delta trees. Analysis & PDE, 7(4):903–927, 2014.
  • [6] V. Banica and L. I. Ignat. Dispersion for the Schrödinger equation on networks. Journal of mathematical physics, 52(8), 2011.
  • [7] C. Bi, J. Cheng, and B. Hua. The wave equation on lattices and oscillatory integrals. arXiv e-prints, pages arXiv–2312, 2023.
  • [8] S. Burger, F. S. Cataliotti, C. Fort, P. Maddaloni, F. Minardi, and M. Inguscio. Quasi-2d Bose-Einstein condensation in an optical lattice. Europhysics Letters, 57(1):1, 2002.
  • [9] D. N. Christodoulides, F. Lederer, and Y. Silberberg. Discretizing light behaviour in linear and nonlinear waveguide lattices. Nature, 424(6950):817–823, 2003.
  • [10] S. Cuccagna and M. Tarulli. On asymptotic stability of standing waves of discrete Schrödinger equation in \mathbb{Z}. SIAM J. Math. Anal., 41(3):861–885, 2009.
  • [11] Jean-Claude Cuenin and Isroil A. Ikromov. Sharp time decay estimates for the discrete Klein-Gordon equation. Nonlinearity, 34(11):7938–7962, 2021.
  • [12] I. Egorova, E. Kopylova, and G. Teschl. Dispersion estimates for one-dimensional discrete Schrödinger and wave equations. J. Spectr. Theory, 5(4):663–696, 2015.
  • [13] L. I. Ignat. Strichartz estimates for the schrödinger equation on a tree and applications. SIAM journal on mathematical analysis, 42(5):2041–2057, 2010.
  • [14] V. N. Karpushkin. Uniform estimates of oscillating integrals in 2\mathbb{R}^{2}. Dokl. Akad. Nauk, 254(1):28–31, 1980.
  • [15] V. N. Karpushkin. Uniform estimates of oscillating integrals with a parabolic or a hyperbolic phase. Trudy Sem. Petrovsk., (9):3–39, 1983.
  • [16] V. N. Karpushkin. A theorem on uniform estimates for oscillatory integrals with a phase depending on two variables. Trudy Sem. Petrovsk., (10):150–169, 238, 1984.
  • [17] M. Keel and T. Tao. Endpoint strichartz estimates. American Journal of Mathematics, 120(5):955–980, 1998.
  • [18] P. G. Kevrekidis, D. E. Pelinovsky, and A. Stefanov. Asymptotic stability of small bound states in the discrete nonlinear Schrödinger equation. SIAM J. Math. Anal., 41(5):2010–2030, 2009.
  • [19] G. Kopidakis, S. Aubry, and G.P. Tsironis. Targeted energy transfer through discrete breathers in nonlinear systems. Physical Review Letters, 87(16):165501, 2001.
  • [20] R. Livi, R. Franzosi, and G.-L. Oppo. Self-localization of Bose-Einstein condensates in optical lattices via boundary dissipation. Physical review letters, 97(6):060401, 2006.
  • [21] M. Maeda, H. Sasaki, E. Segawa, A. Suzuki, and K. Suzuki. Dispersive estimates for quantum walks on 1d lattice. Journal of the Mathematical Society of Japan, 74(1):217–246, 2022.
  • [22] F. A. Mehmeti, K. Ammari, and S. Nicaise. Dispersive effects and high frequency behaviour for the Schrödinger equation in star-shaped networks. Portugaliae Mathematica, 72(4):309–355, 2015.
  • [23] F. A. Mehmeti, K. Ammari, and S. Nicaise. Dispersive effects for the Schrödinger equation on the tadpole graph. Journal of Mathematical Analysis and Applications, 448(1):262–280, 2017.
  • [24] F. A. Mehmeti, K. Ammari, and S. Nicaise. Dispersive effects for the Schrödinger equation on finite metric graphs with infinite ends. arXiv preprint arXiv:2310.16628, 2023.
  • [25] T. Mizumachi and D. Pelinovsky. On the asymptotic stability of localized modes in the discrete nonlinear Schrödinger equation. Discrete Contin. Dyn. Syst. Ser. S, 5(5):971–987, 2012.
  • [26] R. Morandotti, U. Peschel, J.S. Aitchison, H.S. Eisenberg, and Y. Silberberg. Dynamics of discrete solitons in optical waveguide arrays. Physical Review Letters, 83(14):2726, 1999.
  • [27] M. Onorato, A. R. Osborne, M. Serio, and S. Bertone. Freak waves in random oceanic sea states. Physical review letters, 86(25):5831, 2001.
  • [28] D. E. Pelinovsky and A. Stefanov. On the spectral theory and dispersive estimates for a discrete Schrödinger equation in one dimension. J. Math. Phys., 49(11):113501, 17, 2008.
  • [29] D. H. Phong and E. M. Stein. The Newton polyhedron and oscillatory integral operators. Acta Math., 179(1):105–152, 1997.
  • [30] S. Salimi. Study of continuous-time quantum walks on quotient graphs via quantum probability theory. International Journal of Quantum Information, 6(04):945–957, 2008.
  • [31] P. Schultz. The wave equation on the lattice in two and three dimensions. Comm. Pure Appl. Math., 51(6):663–695, 1998.
  • [32] A. Stefanov and P. G. Kevrekidis. Asymptotic behaviour of small solutions for the discrete nonlinear Schrödinger and Klein–Gordon equations. Nonlinearity, 18(4):1841, 2005.
  • [33] E. M. Stein and T. S. Murphy. Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, volume 3. Princeton University Press, 1993.
  • [34] R.S. Strichartz. Restrictions of fourier transforms to quadratic surfaces and decay of solutions of wave equations. 1977.
  • [35] T. Tao. Nonlinear dispersive equations: local and global analysis. Number 106. American Mathematical Soc., 2006.
  • [36] A.N. Varchenko. Newton polyhedra and estimates of oscillatory integrals. Funktsional. Anal. Prilozhen, 10(3):13–38, 1976.

Huabin Ge, [email protected]
School of Mathematics, Renmin University of China, Beijing, 100872, P.R. China

Bobo Hua, [email protected]
School of Mathematical Sciences, LMNS, Fudan University, Shanghai, 200433, P.R. China

Longsong Jia, [email protected]
School of Mathematical Sciences, Peking University, Beijing, 100871, P.R. China

Puchun Zhou, [email protected]
School of Mathematical Sciences, Fudan University, Shanghai, 200433, P.R. China