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

Cutoff Resolvent Estimates and the Semilinear Schrödinger Equation

Hans Christianson Department of Mathematics, University of California, Berkeley, CA 94720 USA hans@math.berkeley.edu
Abstract.

This paper shows how abstract resolvent estimates imply local smoothing for solutions to the Schrödinger equation. If the resolvent estimate has a loss when compared to the optimal, non-trapping estimate, there is a corresponding loss in regularity in the local smoothing estimate. As an application, we apply well-known techniques to obtain well-posedness results for the semi-linear Schrödinger equation.

1. Introduction

In this short note we show how cutoff semiclassical resolvent estimates for the Laplacian on a non-compact manifold, with spectral parameter on the real axis, lead to well-posedness results for the semilinear Schrödinger equation. Motivated by the requirements of [Chr3] and [BGT2], and the microlocal inverse estimates of [Chr1, Chr2], we first prove a general theorem for a large class of resolvents. Following the recent work of Nonnenmacher-Zworski [NoZw], we apply the general theorem in the case there is a hyperbolic fractal trapped set.

Let (M,g)(M,g) be a Riemannian manifold of dimension nn without boundary, with (non-negative) Laplace-Beltrami operator Δ-\Delta acting on functions. The Laplace-Beltrami operator is an unbounded, essentially self-adjoint operator on L2(M)L^{2}(M) with domain H2(M)H^{2}(M). We assume (M,g)(M,g) is asymptotically Euclidean in the sense of [NoZw, (3.7)-(3.9)] and that the classical resolvent (Δ(λ2+iϵ))1(-\Delta-(\lambda^{2}+i\epsilon))^{-1} obeys a limiting absorption principle as ϵ0+\epsilon\to 0+, λ0\lambda\neq 0.

Our first result is that if we have cutoff semiclassical resolvent estimates with a sufficiently small loss, then we have weighted smoothing for the Schrödinger propagator with a loss. Let ρs\rho_{s} be a smooth, non-vanishing weight function satisfying

(1.1) ρs(x)dg(x,x0)s,\displaystyle\rho_{s}(x)\equiv\left\langle d_{g}(x,x_{0})\right\rangle^{-s},

for some fixed x0x_{0} and xx outside a compact set.

Theorem 1.

Suppose for each compactly supported function χ𝒞c(M)\chi\in{\mathcal{C}}^{\infty}_{c}(M) with sufficiently small support, there is h0>0h_{0}>0 such that the semi-classical Laplace-Beltrami operator satisfies

(1.2) χ(h2ΔE)1χuL2(M)g(h)huL2(M),E>0\displaystyle\|\chi(-h^{2}\Delta-E)^{-1}\chi u\|_{L^{2}(M)}\leq\frac{g(h)}{h}\|u\|_{L^{2}(M)},\,\,\,E>0

uniformly in 0<hh00<h\leq h_{0}, where g(h)c0>0g(h)\geq c_{0}>0, g(h)=o(h1)g(h)=o(h^{-1}). Then for each T>0T>0 and s>1/2s>1/2, there is a constant C=CT,s>0C=C_{T,s}>0 such that

(1.3) 0TρseitΔu0H1/2η(M)2𝑑tCu0L2(M)2,\displaystyle\int_{0}^{T}\left\|\rho_{s}e^{it\Delta}u_{0}\right\|_{H^{1/2-\eta}(M)}^{2}dt\leq C\|u_{0}\|_{L^{2}(M)}^{2},

where η0\eta\geq 0 satisfies

(1.4) g(h)h2η=𝒪(1),\displaystyle g(h)h^{2\eta}={\mathcal{O}}(1),

and ρs\rho_{s} is given by (1.1).

The assumption that (M,g)(M,g) is asymptotically Euclidean is that there exists R0>0R_{0}>0 sufficiently large that, on each infinite branch of MB(0,R0)M\setminus B(0,R_{0}), the semiclassical Laplacian h2Δ-h^{2}\Delta takes the form

h2Δ|MB(0,R0)=|α|2aα(x,h)(hDx)α,\displaystyle-h^{2}\Delta|_{M\setminus B(0,R_{0})}=\sum_{|\alpha|\leq 2}a_{\alpha}(x,h)(hD_{x})^{\alpha},

with aα(x,h)a_{\alpha}(x,h) independent of hh for |α|=2|\alpha|=2,

|α|=2aα(x,h)(hDx)αC1|ξ|2,   0<C<, and\displaystyle\sum_{|\alpha|=2}a_{\alpha}(x,h)(hD_{x})^{\alpha}\geq C^{-1}|\xi|^{2},\,\,\,0<C<\infty,\text{ and}
|α|2aα(x,h)(hDx)α|ξ|2,as |x| uniformly in h.\displaystyle\sum_{|\alpha|\leq 2}a_{\alpha}(x,h)(hD_{x})^{\alpha}\to|\xi|^{2},\,\,\,\text{as }|x|\to\infty\text{ uniformly in }h.

In order to quote the results of [NoZw] we also need the following analyticity assumption: θ0[0,π)\exists\theta_{0}\in[0,\pi) such that the aα(x,h)a_{\alpha}(x,h) are extend holomorphically to

{rω:ωn,dist(ω,SSn)<ϵ,r,|r|R0,argr[ϵ,θ0+ϵ)}.\displaystyle\{r\omega:\omega\in{\mathbb{C}}^{n},\,\,\text{dist}\,(\omega,\SS^{n})<\epsilon,\,\,r\in{\mathbb{C}},\,\,|r|\geq R_{0},\,\,\arg r\in[-\epsilon,\theta_{0}+\epsilon)\}.

As in [NoZw], the analyticity assumption immediately implies

xβ(|α|2aα(x,h)ξα|ξ|2)=o(|x||β|)ξ2,|x|.\displaystyle\partial_{x}^{\beta}\left(\sum_{|\alpha|\leq 2}a_{\alpha}(x,h)\xi^{\alpha}-|\xi|^{2}\right)=o(|x|^{-|\beta|})\left\langle\xi\right\rangle^{2},\,\,|x|\to\infty.

Recall the free Laplacian (Δ0λ2)1(-\Delta_{0}-\lambda^{2})^{-1} on n{\mathbb{R}}^{n} has a holomorphic continuation from Imλ>0\,\mathrm{Im}\,\lambda>0 to λ\lambda\in{\mathbb{C}} for n3n\geq 3 odd, and to the logarithmic covering space for nn even. This motivates the limiting absorption assumption, that

limϵ0+,λ0ρs(Δ(λ2+iϵ))1ρs\displaystyle\lim_{\epsilon\to 0+,\,\,\,\lambda\neq 0}\rho_{s}(-\Delta-(\lambda^{2}+i\epsilon))^{-1}\rho_{s}

exists as a bounded operator

L2(M,dvolg)L2(M,dvolg),\displaystyle L^{2}(M,d\text{vol}_{g})\to L^{2}(M,d\text{vol}_{g}),

provided s>1/2s>1/2. As in the free case, we allow a possible logarithmic singularity at λ=0\lambda=0.

The problem of “local smoothing” estimates for the Schrödinger equation has a long history. The sharpest results to date are those of Doi [Doi] and Burq [Bur]. Doi proved if MM is asymptotically Euclidean, then one has the estimate

(1.5) 0TχeitΔu0H1/2(M)2𝑑tCu0L2(M)2\displaystyle\int_{0}^{T}\left\|\chi e^{it\Delta}u_{0}\right\|_{H^{1/2}(M)}^{2}dt\leq C\|u_{0}\|_{L^{2}(M)}^{2}

for χ𝒞c(M)\chi\in{\mathcal{C}}^{\infty}_{c}(M) if and only if there are no trapped sets. Burq’s paper showed if there is trapping due to the presence of several convex obstacles in n{\mathbb{R}}^{n} satisfying certain assumptions, then one has the estimate (1.5) with the H1/2H^{1/2} norm replaced by H1/2ηH^{1/2-\eta} for η>0\eta>0. In [Chr3], the author considered an arbitary, single trapped hyperbolic orbit. One of the goals of this paper is to use estimates obtained by Nonnenmacher-Zworski [NoZw] for fractal hyperbolic trapped sets to obtain similar results to [Chr3] for the semilinear Schrödinger equation. To that end we have the following corollary to Theorem 1.

Corollary 1.1.

Assume (M,g)(M,g) admits a hyperbolic fractal trapped set, KEK_{E}, in the energy level E>0E>0 and that the topological pressure PE(1/2)<0P_{E}(1/2)<0. Then h2ΔE-h^{2}\Delta-E satisfies (1.2) for some E>0E>0 with g(h)=Clog(1/h)g(h)=C\log(1/h), and for every η>0\eta>0, T>0T>0, and s>1/2s>1/2, there exists a constant C=CPE,η,T,s>0C=C_{P_{E},\eta,T,s}>0 such that

0TρseitΔu0H1/2η(M)2𝑑tCu0L2(M)2.\displaystyle\int_{0}^{T}\left\|\rho_{s}e^{it\Delta}u_{0}\right\|_{H^{1/2-\eta}(M)}^{2}dt\leq C\|u_{0}\|_{L^{2}(M)}^{2}.

We remark that the assumption PE(1/2)<0P_{E}(1/2)<0 implies the trapped set KEK_{E} is filamentary or “thin” (see [NoZw] for definitions).

We consider the following semilinear Schrödinger equation problem:

(1.8) {itu+Δu=F(u)on I×M;u(0,x)=u0(x),\displaystyle\left\{\begin{array}[]{cc}i\partial_{t}u+\Delta u=F(u)\,\,\text{on }I\times M;\\ u(0,x)=u_{0}(x),\end{array}\right.

where II\subset{\mathbb{R}} is an interval containing 0. Here the nonlinearity FF satisfies

F(u)=G(|u|2)u,\displaystyle F(u)=G^{\prime}(|u|^{2})u,

and G:G:{\mathbb{R}}\to{\mathbb{R}} is at least C3C^{3} and satisfies

|G(k)(r)|Ckrβk,\displaystyle|G^{(k)}(r)|\leq C_{k}\langle r\rangle^{\beta-k},

for some β12\beta\geq{\frac{1}{2}}.

In §3 we prove a family of Strichartz-type estimates which will result in the following well-posedness theorem.

Theorem 2.

Suppose (M,g)(M,g) satisfies the assumptions of the introduction, and set

(1.9) δ=4η2η+10.\displaystyle\delta=\frac{4\eta}{2\eta+1}\geq 0.

Then for each

(1.10) s>n22max{2β2,2}+δ\displaystyle s>\frac{n}{2}-\frac{2}{\max\{2\beta-2,2\}}+\delta

and each u0Hs(M)u_{0}\in H^{s}(M) there exists p>max{2β2,2}p>\max\{2\beta-2,2\} and 0<T10<T\leq 1 such that (1.8) has a unique solution

(1.11) uC([T,T];Hs(M))Lp([T,T];L(M)).\displaystyle u\in C([-T,T];H^{s}(M))\cap L^{p}([-T,T];L^{\infty}(M)).

Moreover, the map u0(x)u(t,x)C([T,T];Hs(M))u_{0}(x)\mapsto u(t,x)\in C([-T,T];H^{s}(M)) is Lipschitz continuous on bounded sets of Hs(M)H^{s}(M), and if u0Hs\|u_{0}\|_{H^{s}} is bounded, TT is bounded from below.

If, in addition, (M,g)(M,g) satisfies the assumptions of Corollary 1.1, n3n\leq 3, β<3\beta<3, and G(r)+G(r)\to+\infty as r+r\to+\infty, then uu in (1.11) extends to a solution

uC((,);H1(M))Lp((,);L(M)).\displaystyle u\in C((-\infty,\infty);H^{1}(M))\cap L^{p}((-\infty,\infty);L^{\infty}(M)).
Remark 1.2.

In particular, the cubic defocusing non-linear Schrödinger equation is globally H1H^{1}-well-posed in three dimensions with a fractal trapped hyperbolic set which is sufficiently filamentary. Of course other nonlinearities can be considered, but for simplicity we consider only these in this work.

Acknowledgments. This research was partially conducted during the period the author was employed by the Clay Mathematics Institute as a Liftoff Fellow.

2. Proof of Theorem 1

Since we are assuming (Δz)1(-\Delta-z)^{-1} obeys a limiting absorption principle, we have

ρs(Δ(τiϵ))1ρsL2L2Cϵ\displaystyle\|\rho_{s}(-\Delta-(\tau-i\epsilon))^{-1}\rho_{s}\|_{L^{2}\to L^{2}}\leq C_{\epsilon}

for 0<ϵ0|τ|C0<\epsilon_{0}\leq|\tau|\leq C. For |σ|C|\sigma|\geq C for some C>0C>0, σ\sigma\in{\mathbb{C}} in a neighbourhood of the real axis, write

Δσ\displaystyle-\Delta-\sigma =\displaystyle= Δzh2\displaystyle-\Delta-\frac{z}{h^{2}}
=\displaystyle= h2(h2Δz),\displaystyle h^{-2}(-h^{2}\Delta-z),

for

z[Eα,E+α]+i[c0h,c0h].\displaystyle z\in[E-\alpha,E+\alpha]+i[-c_{0}h,c_{0}h].

Now

(h2Δz)\displaystyle(-h^{2}\Delta-z)

is a Fredholm operator for zz in the specified range, and hence the “gluing” techniques from [Vod] and [Chr3, §2] can be used to conclude for s>1/2s>1/2,

ρs(h2Δz)1ρs\displaystyle\rho_{s}(-h^{2}\Delta-z)^{-1}\rho_{s}

has a holomorphic extension to a slightly smaller neighbourhood in zz, and in particular,

ρs(h2ΔE)1ρsL2L2Cg(h)h.\displaystyle\|\rho_{s}(-h^{2}\Delta-E)^{-1}\rho_{s}\|_{L^{2}\to L^{2}}\leq C\frac{g(h)}{h}.

Rescaling, we have

(2.1) ρs(Δτ)1ρsL2L2Cg(τ1/2)τ1/2,τ𝒞±ϵ,\displaystyle\left\|\rho_{s}(-\Delta-\tau)^{-1}\rho_{s}\right\|_{L^{2}\to L^{2}}\leq C\frac{g(\left\langle\tau\right\rangle^{1/2})}{\left\langle\tau\right\rangle^{1/2}},\,\,\,\tau\in\mathcal{C}_{\pm\epsilon},

where (see Figure 1)

𝒞±ϵ={τ:|τ|ϵ}{τ:|τ|=ϵ,±Imτ0}.\displaystyle\mathcal{C}_{\pm\epsilon}=\{\tau\in{\mathbb{R}}:|\tau|\geq\epsilon\}\cup\{\tau\in{\mathbb{C}}:|\tau|=\epsilon,\,\pm\,\mathrm{Im}\,\tau\geq 0\}.

Refer to captionϵ\epsilonCϵC_{-\epsilon}

Figure 1. The curve 𝒞ϵ\mathcal{C}_{-\epsilon} in the complex plane.

As in [Chr3] and [Bur], the following lemma follows from integration by parts and interpolation, together with the condition on η\eta, (1.4).

Lemma 2.1.

With the notation and assumptions above, we have

ρs(Δτ)1ρsL2H1Cg(τ1/2),τ𝒞±ϵ,\displaystyle\|\rho_{s}(-\Delta-\tau)^{-1}\rho_{s}\|_{L^{2}\to H^{1}}\leq Cg(\left\langle\tau\right\rangle^{1/2}),\,\,\,\tau\in\mathcal{C}_{\pm\epsilon},

and for every r[1,1]r\in[-1,1],

ρs(Δτ)1ρsHrH1+rη/2C,τ𝒞±ϵ.\displaystyle\|\rho_{s}(-\Delta-\tau)^{-1}\rho_{s}\|_{H^{r}\to H^{1+r-\eta/2}}\leq C,\,\,\,\tau\in\mathcal{C}_{\pm\epsilon}.

Theorem 1 now follows from the standard “TTTT^{*}” argument, letting ϵ0\epsilon\to 0 in (2.1) (see [BGT2], the references cited therein, and [Chr3]).

The following Corollary uses interpolation with an H2H^{2} estimate to replace the H1/2ηH^{1/2-\eta} norm on the left hand side of (1.3) with H1/2H^{1/2}, and will be of use in §3. See [Chr3] for the details of the proof.

Corollary 2.2.

Suppose (M,g)(M,g) satisfies the assumptions of Theorem 1. For each T>0T>0 and s>1/2s>1/2, there is a constant C>0C>0 such that

(2.2) 0TρseitΔu0H1/2(M)2𝑑tCu0Hδ(M)2,\displaystyle\int_{0}^{T}\left\|\rho_{s}e^{it\Delta}u_{0}\right\|_{H^{1/2}(M)}^{2}dt\leq C\|u_{0}\|_{H^{\delta}(M)}^{2},

where δ0\delta\geq 0 is given by (1.9).

In particular, if (M,g)(M,g) satisfies the assumptions of Corollary 1.1, then for any δ>0\delta>0, there is C=Cδ>0C=C_{\delta}>0 such that (2.2) holds.

3. Strichartz-type Inequalities

In this section we give several families of Strichartz-type inequalities and prove Theorem 2. The statements and proofs are mostly adaptations of similar inequalities in [BGT2], so we leave out the proofs of these in the interest of space.

If we view MUM\setminus U, where UU is a neighbourhood of KEK_{E}, as a manifold with non-trapping geometry, we may apply the results of [HTW] or [BoTz] to a solution of the Schrödinger equation away from the trapping region, resulting in perfect Strichartz estimates. For this section we need (1.3) only with a compact cutoff χ\chi instead of with the more general weight ρs\rho_{s}.

Proposition 3.1.

For every 0<T10<T\leq 1 and each χ𝒞c(M)\chi\in{\mathcal{C}}^{\infty}_{c}(M) satisfying χ1\chi\equiv 1 near UU , there is a constant C>0C>0 such that

(3.1) (1χ)uLp([0,T])Ws,q(M)Cu0Hs(M),\displaystyle\|(1-\chi)u\|_{L^{p}([0,T])W^{s,q}(M)}\leq C\|u_{0}\|_{H^{s}(M)},

where u=eitΔu0u=e^{it\Delta}u_{0}, s[0,1]s\in[0,1], and (p,q)(p,q), p>2p>2 satisfy

2p+nq=n2.\displaystyle\frac{2}{p}+\frac{n}{q}=\frac{n}{2}.
Remark 3.2.

In the sequel, wherever unambiguous, we will write

LTpWs,q:=Lp([0,T])Ws,q(M)\displaystyle L_{T}^{p}W^{s,q}:=L^{p}([0,T])W^{s,q}(M)

and

Hs:=Hs(M).\displaystyle H^{s}:=H^{s}(M).
Proposition 3.3.

Suppose (M,g)(M,g) satisfies the assumptions of the Introduction, u=eitΔu0u=e^{it\Delta}u_{0}, and

v=0tei(tτ)Δf(τ)𝑑τ.\displaystyle v=\int_{0}^{t}e^{i(t-\tau)\Delta}f(\tau)d\tau.

Then for each 0<T10<T\leq 1 and δ0\delta\geq 0 satisfying (1.9), we have the estimates

(3.2) uLTpWsδ,qCu0Hs\displaystyle\|u\|_{L^{p}_{T}W^{s-\delta,q}}\leq C\|u_{0}\|_{H^{s}}

and

(3.3) vLTpWsδ,qCfLT1Hs,\displaystyle\|v\|_{L_{T}^{p}W^{s-\delta,q}}\leq C\|f\|_{L_{T}^{1}H^{s}},

where s[0,1]s\in[0,1] and (p,q)(p,q), p>2p>2 satisfy the Euclidean scaling

(3.4) 2p+nq=n2.\displaystyle\frac{2}{p}+\frac{n}{q}=\frac{n}{2}.

The proof uses a local WKB expansion localized also in time to the scale of inverse frequency, followed by summing over frequency bands (see [Chr3] and [BGT1]). The only difference here is the explicit dependence of δ\delta on η\eta, which is related to the growth of the function g(h)g(h).

Proof of Theorem 2.

The proof of Theorem 2 is a slight modification of the proof of Proposition 3.1 in [BGT1], but we include it here in the interest of completeness. Fix ss satisfying 1.10 and choose p>max{2β2,2}p>\max\{2\beta-2,2\} satisfying

s>n22p+δn21max{2β2,2}\displaystyle s>\frac{n}{2}-\frac{2}{p}+\delta\geq\frac{n}{2}-\frac{1}{\max\{2\beta-2,2\}}

where δ0\delta\geq 0 satisfies (1.9). Set σ=sδ\sigma=s-\delta and

YT=C([T,T];Hs(M))Lp([T,T];Wσ,q(M))\displaystyle Y_{T}=C([-T,T];H^{s}(M))\cap L^{p}([-T,T];W^{\sigma,q}(M))

for

2p+nq=n2,\displaystyle\frac{2}{p}+\frac{n}{q}=\frac{n}{2},

equipped with the norm

uYT=max|t|Tu(t)Hs(M)+uLTpWσ,q.\displaystyle\|u\|_{Y_{T}}=\max_{|t|\leq T}\|u(t)\|_{H^{s}(M)}+\|u\|_{L^{p}_{T}W^{\sigma,q}}.

Let Φ\Phi be the nonlinear functional

Φ(u)=eitΔu0i0tei(tτ)ΔF(u(τ))𝑑τ.\displaystyle\Phi(u)=e^{it\Delta}u_{0}-i\int_{0}^{t}e^{i(t-\tau)\Delta}F(u(\tau))d\tau.

If we can show that Φ:YTYT\Phi:Y_{T}\to Y_{T} and is a contraction on a ball in YTY_{T} centered at 0 for sufficiently small T>0T>0, this will prove the first assertion of the Proposition, along with the Sobolev embedding

(3.5) Wσ,q(M)L(M),\displaystyle W^{\sigma,q}(M)\subset L^{\infty}(M),

since σ>n/q\sigma>n/q. From Proposition 3.3, we bound the WσW^{\sigma} part of the YTY_{T} norm by the HsH^{s} norm, giving

Φ(u)YT\displaystyle\|\Phi(u)\|_{Y_{T}} \displaystyle\leq C(u0Hs+TTF(u(τ))Hs𝑑τ)\displaystyle C\left(\|u_{0}\|_{H^{s}}+\int_{-T}^{T}\|F(u(\tau))\|_{H^{s}}d\tau\right)
\displaystyle\leq C(u0Hs+TT(1+|u(τ)|)L2β2)u(τ)Hsdτ),\displaystyle C\left(\|u_{0}\|_{H^{s}}+\int_{-T}^{T}\|(1+|u(\tau)|)\|_{L^{\infty}}^{2\beta-2})\|u(\tau)\|_{H^{s}}d\tau\right),

where the last inequality follows by our assumptions on the structure of FF. Applying Hölder’s inequality in time with p~=p/(2β2)\tilde{p}=p/(2\beta-2) and q~\tilde{q} satisfying

1q~+1p~=1\displaystyle\frac{1}{\tilde{q}}+\frac{1}{\tilde{p}}=1

gives

Φ(u)YTC(u0Hs+TγuLTHs(1+|u|)LTpL2β2))\displaystyle\|\Phi(u)\|_{Y_{T}}\leq C\left(\|u_{0}\|_{H^{s}}+T^{\gamma}\|u\|_{L^{\infty}_{T}H^{s}}\|(1+|u|)\|_{L^{p}_{T}L^{\infty}}^{2\beta-2})\right)

where γ=1/q~>0\gamma=1/\tilde{q}>0. Thus

Φ(u)YTC(u0Hs+Tγ(uYT+uYT2β)).\displaystyle\|\Phi(u)\|_{Y_{T}}\leq C\left(\|u_{0}\|_{H^{s}}+T^{\gamma}(\|u\|_{Y_{T}}+\|u\|_{Y_{T}}^{2\beta})\right).

Similarly, we have for u,vYTu,v\in Y_{T},

(3.6) Φ(u)Φ(v)YT\displaystyle\|\Phi(u)-\Phi(v)\|_{Y_{T}}\leq
\displaystyle\leq CTγuvLTHs(1+|u|)LTpL2β2+(1+|v|)LTpL2β2)\displaystyle CT^{\gamma}\|u-v\|_{L^{\infty}_{T}H^{s}}\|(1+|u|)\|_{L^{p}_{T}L^{\infty}}^{2\beta-2}+\|(1+|v|)\|_{L^{p}_{T}L^{\infty}}^{2\beta-2})
\displaystyle\leq CTγuvYT(1+|u|)YT2β2+(1+|v|)YT2β2),\displaystyle CT^{\gamma}\|u-v\|_{Y_{T}}\|(1+|u|)\|_{Y_{T}}^{2\beta-2}+\|(1+|v|)\|_{Y_{T}}^{2\beta-2}),

which is a contraction for sufficiently small TT. This concludes the proof of the first assertion in the Proposition.

To get the second assertion, we observe from 3.6 and the definition of YTY_{T}, if uu and vv are two solutions to (1.8) with initial data u0u_{0} and u1u_{1} respectively, so

Φ~(v)=eitΔu1i0tei(tτ)ΔF(v(τ))𝑑τ,\displaystyle\widetilde{\Phi}(v)=e^{it\Delta}u_{1}-i\int_{0}^{t}e^{i(t-\tau)\Delta}F(v(\tau))d\tau,

we have

max|t|Tu(t)v(t)Hs\displaystyle\max_{|t|\leq T}\|u(t)-v(t)\|_{H^{s}}
=\displaystyle= max|t|TΦ(u)(t)Φ~(v)(t)Hs\displaystyle\max_{|t|\leq T}\|\Phi(u)(t)-\widetilde{\Phi}(v)(t)\|_{H^{s}}
C\displaystyle\leq C (u0u1Hs\displaystyle\Bigg{(}\|u_{0}-u_{1}\|_{H^{s}}
+Tγmax|t|Tu(t)v(t)Hs(1+|u|)LTpL2β2+(1+|v|)LTpL2β2)),\displaystyle\quad+T^{\gamma}\max_{|t|\leq T}\|u(t)-v(t)\|_{H^{s}}\|(1+|u|)\|_{L^{p}_{T}L^{\infty}}^{2\beta-2}+\|(1+|v|)\|_{L^{p}_{T}L^{\infty}}^{2\beta-2})\Bigg{)},

which, for T>0T>0 sufficiently small gives the Lipschitz continuity.

If (M,g)(M,g) satisfies the assumptions of Corollary 1.1, n3n\leq 3, β<3\beta<3, and G(r)+G(r)\to+\infty as r+r\to+\infty, we can take ss and pp satisfying p>max{2β2,2}p>\max\{2\beta-2,2\} and

s>n22p+δn22max{2β2,2}\displaystyle s>\frac{n}{2}-\frac{2}{p}+\delta\geq\frac{n}{2}-\frac{2}{\max\{2\beta-2,2\}}

for any δ>0\delta>0. Then σ=sδ>q/n\sigma=s-\delta>q/n and the preceding argument holds. Finally, the proof of the global well-posedness now follows from the standard global well-posedness arguments from, for example, [Caz, Chapter 6]. ∎

References

  • [BoTz] Bouclet, J-M. and Tzvetkov, N. Strichartz Estimates for Long Range Perturbations. preprint.
    http://arxiv.org/pdf/math/0509489
  • [Bur] Burq, N. Smoothing Effect for Schrödinger Boundary Value Problems. Duke Math. Journal. 123, No. 2, 2004, p. 403-427.
  • [BGT1] Burq, N. Gérard, P., and Tzvetkov, N. Strichartz Inequalities and the Nonlinear Schrödinger Equation on Compact Manifolds. Amer. J. Math. 126, No. 3, 2004, p. 569-605.
  • [BGT2] Burq, N., Gérard, P., and Tzvetkov, N. On Nonlinear Schrödinger Equations in Exterior Domains. Ann. I H. Poincaré. 21, 2004, p. 295-318.
  • [Caz] Cazenave, T. Semilinear Schrödinger Equations. Courant Lecture Notes in Mathematics, AMS, 2003.
  • [Chr1] Christianson, H. Semiclassical Non-concentration near Hyperbolic Orbits. J. Funct. Anal. 262, 2007, no. 2, p. 145-195.
  • [Chr2] Christianson, H. Quantum Monodromy and Non-concentration near Semi-hyperbolic Orbits. in preparation.
  • [Chr3] Christianson, H. Dispersive Estimates for Manifolds with one Trapped Orbit. preprint. 2006.
    http://www.math.berkeley.edu/\simhans/papers/sm.pdf
  • [Doi] Doi, S.-I. Smoothing effects of Schrödinger Evolution Groups on Riemannian Manifolds. Duke Mathematical Journal. 82, No. 3, 1996, p. 679-706.
  • [HTW] Hassell, A., Tao, T., and Wunsch, J. Sharp Strichartz Estimates on Non-trapping Asymptotically Conic Manifolds. preprint. 2004.
    http://www.arxiv.org/pdf/math.AP/0408273
  • [NoZw] Nonnenmacher, S. and Zworski, M. Quantum decay rates in chaotic scattering. preprint. 2007.
    http://math.berkeley.edu/\simzworski/nz3.ps.gz
  • [Vod] Vodev, G. Exponential Bounds of the Resolvent for a Class of Noncompactly Supported Perturbations of the Laplacian. Math. Res. Lett. 7 (2000), no. 2-3, 287–298.