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

Resonances as Viscosity Limits for Exponentially Decaying Potentials

Haoren Xiong [email protected] Department of Mathematics, University of California, Berkeley, CA 94720, USA
Abstract

We show that the complex absorbing potential (CAP) method for computing scattering resonances applies to the case of exponentially decaying potentials. That means that the eigenvalues of Δ+Viεx2-\Delta+V-i\varepsilon x^{2}, |V(x)|Ce2γ|x||V(x)|\leq Ce^{-2\gamma|x|} converge, as ε0+\varepsilon\to 0+, to the poles of the meromorphic continuation of (Δ+Vλ2)1(-\Delta+V-\lambda^{2})^{-1} uniformly on compact subsets of Reλ>0\operatorname{Re}\lambda>0, Imλ>γ\operatorname{Im}\lambda>-\gamma, argλ>π/8\arg\lambda>-\pi/8.

1 Introduction

The complex absorbing potential (CAP) method has been used as a computational tool for finding scattering resonances – see Riss–Meyer RiMe and Seideman–Miller semi for an early treatment and Jagau et al Jag for some recent developments. For potentials VLcompV\in L^{\infty}_{\operatorname{comp}} the method was justified by Zworski Zw-vis . In xiong2020 the author extended it to potentials which are dilation analytic near infinity. In this paper we show that the CAP method is also valid for potentials which are exponentially decaying. While the key component of Zw-vis and xiong2020 was the method of complex scaling (see Hunziker hunziker1986 , Sjöstrand–Zworski sjostrand1991 for an account and references), here we use complex scaling on the Fourier transform side following Nakamura nakamura1990 and Kameoka–Nakamura Nakamura .

Thus, we consider the Schrödinger operator P:=Δ+VP:=-\Delta+V acting on L2(n)L^{2}({\mathbb{R}}^{n}) whose potential is exponentially decaying, this means that there exist C,γ>0C,\gamma>0 such that

|V(x)|Ce2γ|x|.|V(x)|\leq Ce^{-2\gamma|x|}. (1.1)

Let RV(λ)=(Pλ2)1R_{V}(\lambda)=(P-\lambda^{2})^{-1} be the resolvent of PP, initially defined for Imλ>0\operatorname{Im}\lambda>0. The exponentially weighted resolvent VRV(λ)V\sqrt{V}R_{V}(\lambda)\sqrt{V} can be meromorphically continued to the strip Imλ>γ\operatorname{Im}\lambda>-\gamma, see Froese froese , Gannot gannot and a review in §2. Resonances of PP are the poles in this meromorphic continuation.

We now introduce a regularized operator,

Pε:=Δiεx2+V,ε>0.P_{\varepsilon}:=-\Delta-i\varepsilon x^{2}+V,\ \ \varepsilon>0. (1.2)

(We write x2:=x12++xn2x^{2}:=x_{1}^{2}+\cdots+x_{n}^{2}.) It is easy to see, with details reviewed in §4, that PεP_{\varepsilon} is a non-normal unbounded operator on L2(n)L^{2}({\mathbb{R}}^{n}) with a discrete spectrum. When V0V\equiv 0, PεP_{\varepsilon} is reduced to the rescaled Davies harmonic oscillator – see §3, whose spectrum is given by

εeiπ/4(2|α|+n),α0n,|α|:=α1++αn,\sqrt{\varepsilon}\,e^{-i\pi/4}(2|\alpha|+n),\quad\alpha\in{\mathbb{N}}_{0}^{n},\quad|\alpha|:=\alpha_{1}+\cdots+\alpha_{n},

where 0{\mathbb{N}}_{0} denotes the set of nonnegative integers. Thus we will restrict our attentions to argz>π/4\arg z>-\pi/4. Suppose that

σ(Pε)eiπ/4[0,)={λj(ε)2}j=1,π/8<argλj(ε)<7π/8.\sigma(P_{\varepsilon})\cap{\mathbb{C}}\setminus e^{-i\pi/4}[0,\infty)=\{\lambda_{j}(\varepsilon)^{2}\}_{j=1}^{\infty},\quad-\pi/8<\arg\lambda_{j}(\varepsilon)<7\pi/8. (1.3)

Zworski Zw-vis proved that resonances can be defined as the limit points of {λj(ε)}j=1\{\lambda_{j}(\varepsilon)\}_{j=1}^{\infty} as ε0+\varepsilon\to 0+, in the case of compactly supported potentials. We generalize this result to the case of exponentially decaying potentials. More precisely, we have

Theorem 1.

For any 0<a<a<b0<a^{\prime}<a<b and γ<γ\gamma^{\prime}<\gamma such that the rectangle

Ω:=(a,a)+i(γ,b){λ:π/8<argλ<7π/8},\Omega:=(a^{\prime},a)+i(-\gamma^{\prime},b)\Subset\{\lambda\in{\mathbb{C}}:-\pi/8<\arg\lambda<7\pi/8\}, (1.4)

we have, uniformly on Ω\Omega,

λj(ε)λj,ε0+,\lambda_{j}(\varepsilon)\to\lambda_{j},\quad\varepsilon\to 0+,

where λj\lambda_{j} are the resonances of PP.

Notation. We use the following notation: f=𝒪(g)Hf=\mathcal{O}_{\ell}(g)_{H} means that fHCg\|f\|_{H}\leq C_{\ell}g where the norm (or any seminorm) is in the space HH, and the constant CC_{\ell} depends on \ell. When either \ell or HH are absent then the constant is universal or the estimate is scalar, respectively. When G=𝒪(g):H1H2G=\mathcal{O}_{\ell}(g):{H_{1}\to H_{2}} then the operator G:H1H2G:H_{1}\to H_{2} has its norm bounded by CgC_{\ell}g. Also when no confusion is likely to result, we denote the operator fgff\mapsto gf where gg is a function by gg.

2 meromorphic continuation

In this section we will introduce a meromorphic continuation of the weighted resolvent VRV(λ)V\sqrt{V}R_{V}(\lambda)\sqrt{V} from Imλ>0\operatorname{Im}\lambda>0 to the strip Imλ>γ\operatorname{Im}\lambda>-\gamma under the assumption (1.1). As in froese , we define the resonances of PP as the poles of this meromorphic continuation, with agreement of multiplicities. For a detailed presentation, we refer to froese .

Let R0(λ):=(Δλ2)1R_{0}(\lambda):=(-\Delta-\lambda^{2})^{-1} be the free resolvent. For Imλ>0\operatorname{Im}\lambda>0, the resolvent equation

R0(λ)RV(λ)RV(λ)VR0(λ)=0R_{0}(\lambda)-R_{V}(\lambda)-R_{V}(\lambda)VR_{0}(\lambda)=0

implies

(IVRV(λ)V)(I+VR0(λ)V)=I.(I-\sqrt{V}R_{V}(\lambda)\sqrt{V})(I+\sqrt{V}R_{0}(\lambda)\sqrt{V})=I.

Since R0(λ)=𝒪(|Imλ|1):L2L2R_{0}(\lambda)=\mathcal{O}(|\operatorname{Im}\lambda|^{-1}):L^{2}\to L^{2}, then for Imλ\operatorname{Im}\lambda large, I+VR0(λ)VI+\sqrt{V}R_{0}(\lambda)\sqrt{V} is invertible by a Neumann series argument and

IVRV(λ)V=(I+VR0(λ)V)1.I-\sqrt{V}R_{V}(\lambda)\sqrt{V}=(I+\sqrt{V}R_{0}(\lambda)\sqrt{V})^{-1}. (2.1)

We will show that the right side of (2.1)\eqref{eqn:meroCont} has a meromorphic continuation. For that, we recall some bounds of the free resolvent with exponential weights, see gannot for details, to prove the following lemma:

Lemma 1.

For any a>0a>0 and γ<γ\gamma^{\prime}<\gamma,

λ(I+VR0(λ)V)1,Reλ>a,Imλ>γ,\lambda\mapsto(I+\sqrt{V}R_{0}(\lambda)\sqrt{V})^{-1},\quad\operatorname{Re}\lambda>a,\ \operatorname{Im}\lambda>-\gamma^{\prime},

is a meromorphic family of operators on L2(n)L^{2}({\mathbb{R}}^{n}) with poles of finite rank.

Proof.

Choose φ𝒞(n)\varphi\in{{\mathcal{C}}^{\infty}}({\mathbb{R}}^{n}) satisfying φ(x)=|x|\varphi(x)=|x| for large |x||x|, it is well known that for each c>0c>0, the weighted resolvent:

ecφR0(λ)ecφ:L2(n)L2(n)e^{-c\varphi}R_{0}(\lambda)e^{-c\varphi}:L^{2}({\mathbb{R}}^{n})\to L^{2}({\mathbb{R}}^{n})

extends analytically across Reλ>0\operatorname{Re}\lambda>0 to the strip Imλ>c\operatorname{Im}\lambda>-c, see (gannot, , §1) and references given there. Moreover, Gannot (gannot, , §1) proved that for any a,c,ε>0a,c,\varepsilon>0 and αn,|α|2\alpha\in{\mathbb{N}}^{n},\ |\alpha|\leq 2 there exists Cα=Cα(a,c,ε)C_{\alpha}=C_{\alpha}(a,c,\varepsilon) such that

Dα(ecφR0(λ)ecφ)L2L2Cα|λ||α|1,for Imλ>c+ε,Reλ>a.\|D^{\alpha}(e^{-c\varphi}R_{0}(\lambda)e^{-c\varphi})\|_{L^{2}\to L^{2}}\leq C_{\alpha}|\lambda|^{|\alpha|-1},\quad\textrm{for }\operatorname{Im}\lambda>-c+\varepsilon,\ \operatorname{Re}\lambda>a. (2.2)

In particular, for Reλ>a\operatorname{Re}\lambda>a and Imλ>γ\operatorname{Im}\lambda>-\gamma^{\prime},

λeγφR0(λ)eγφ\lambda\mapsto e^{-\gamma^{\prime}\varphi}R_{0}(\lambda)e^{-\gamma^{\prime}\varphi}

is an analytic family of operators L2H2L^{2}\to H^{2}. Since lim|x|V(x)eγφ(x)=0\lim_{|x|\to\infty}\sqrt{V(x)}e^{\gamma^{\prime}\varphi(x)}=0 by (1.1), it is easy to see that Veγφ:H2L2\sqrt{V}e^{\gamma^{\prime}\varphi}:H^{2}\to L^{2} is compact. Hence,

λVR0(λ)V=Veγφ(eγφR0(λ)eγφ)Veγφ\lambda\mapsto\sqrt{V}R_{0}(\lambda)\sqrt{V}=\sqrt{V}e^{\gamma^{\prime}\varphi}(e^{-\gamma^{\prime}\varphi}R_{0}(\lambda)e^{-\gamma^{\prime}\varphi})\sqrt{V}e^{\gamma^{\prime}\varphi}

is an analytic family of compact operators L2L2L^{2}\to L^{2} for Reλ>a,Imλ>γ\operatorname{Re}\lambda>a,\ \operatorname{Im}\lambda>-\gamma^{\prime}. Recalling that I+VR0(λ)VI+\sqrt{V}R_{0}(\lambda)\sqrt{V} is invertible for Imλ1\operatorname{Im}\lambda\gg 1, then by the analytic Fredholm theory – see (res, , §C.4), λ(I+VR0(λ)V)1\lambda\mapsto(I+\sqrt{V}R_{0}(\lambda)\sqrt{V})^{-1} is a meromorphic family of operators in the same range of λ\lambda. ∎

From now on, we identify the resonances λj\lambda_{j}, in Ω\Omega given in (1.4), with the poles of (I+VR0(λ)V)1(I+\sqrt{V}R_{0}(\lambda)\sqrt{V})^{-1}, with agreement of multiplicities. More precisely, the multiplicity of resonance λ\lambda is given by

m(λ):=12πitrλ(I+VR0(ζ)V)1ζ(VR0(ζ)V)dζ,m(\lambda):=\frac{1}{2\pi i}\operatorname{tr}\oint_{\lambda}(I+\sqrt{V}R_{0}(\zeta)\sqrt{V})^{-1}\partial_{\zeta}(\sqrt{V}R_{0}(\zeta)\sqrt{V})\,d\zeta, (2.3)

where the integral is over a positively oriented circle enclosing λ\lambda and containing no poles other than λ\lambda.

3 resolvent estimates for the Davies harmonic oscillator

The operator Hc:=Δ+cx2,π<argc0H_{c}:=-\Delta+cx^{2},\ -\pi<\arg c\leq 0, was used by Davies Dav to illustrate properties of non-normal differential operators. We recall some known facts about HcH_{c} and its resolvent. As established in Dav , HcH_{c} is an unbounded operator on L2(n)L^{2}({\mathbb{R}}^{n}) with the discrete spectrum given by

σ(Hc)={c1/2(n+2|α|):α0n}.\sigma(H_{c})=\{c^{1/2}(n+2|\alpha|):\alpha\in{\mathbb{N}}_{0}^{n}\}. (3.1)

In particular σ(Hiε)eiπ/4[0,)\sigma(H_{-i\varepsilon})\subset e^{-i\pi/4}[0,\infty), then one can study the resolvent of HiεH_{-i\varepsilon} outside eiπ/4[0,)e^{-i\pi/4}[0,\infty). Unlike the normal operators, there does not exist any constant CC such that (Δiεx2z)1L2L2Cdist(z,eiπ/4[0,))1\|(-\Delta-i\varepsilon x^{2}-z)^{-1}\|_{L^{2}\to L^{2}}\leq C\operatorname{dist}(z,e^{-i\pi/4}[0,\infty))^{-1}. Instead, according to Hitrik–Sjöstrand–Viola HSV , (Zw-vis, , §3) and references given there, for Ω{z:π/2<argz<0}eiπ/4[0,)\Omega\Subset\{z:-\pi/2<\arg z<0\}\setminus e^{-i\pi/4}[0,\infty), there exists C=C(Ω)C=C(\Omega) such that

1Ceε12/C(Δiεx2z)1L2L2CeCε12,zΩ.\frac{1}{C}e^{\varepsilon^{-\frac{1}{2}}/C}\leq\|(-\Delta-i\varepsilon x^{2}-z)^{-1}\|_{L^{2}\to L^{2}}\leq Ce^{C\varepsilon^{-\frac{1}{2}}},\quad z\in\Omega. (3.2)

In this section we will show how exponential weights dramatically improve the bound (3.2) for (Δiεx2λ2)1(-\Delta-i\varepsilon x^{2}-\lambda^{2})^{-1} in the rectangle Ω\Omega given by (1.4), which will be crucial in the proof of Theorem 1.

First, note that Δxiεx2=1(ξ2+iεΔξ)-\Delta_{x}-i\varepsilon x^{2}={\mathcal{F}}^{-1}(\xi^{2}+i\varepsilon\Delta_{\xi}){\mathcal{F}}, where {\mathcal{F}} denotes the Fourier transform u(ξ)=u^(ξ)=(2π)n/2eixξu(x)𝑑x{\mathcal{F}}u(\xi)=\hat{u}(\xi)=(2\pi)^{-n/2}\int e^{-ix\cdot\xi}u(x)\,dx. Inspired by nakamura1990 and Nakamura , we introduce a family of spectral deformations in the Fourier space as follows.

For any fixed Ω\Omega given in (1.4), we choose ρ𝒞([0,);)\rho\in{{\mathcal{C}}^{\infty}}([0,\infty);{\mathbb{R}}) with ρ0\rho\equiv 0 near 0 and ρ(t)1\rho(t)\equiv 1 for t1t\gg 1 such that

0ρ(t)<γ1tanπ8,t0;Ω{x+iy:x>0,y>γρ(x)},0\leq\rho^{\prime}(t)<\gamma^{-1}\tan\frac{\pi}{8},\ \forall\,t\geq 0;\quad\Omega\Subset\{x+iy:x>0,\ y>-\gamma\rho(x)\}, (3.3)

and define the map

ψ:nn,ψ(ξ)=|ξ|1ρ(|ξ|)ξ,\psi:{\mathbb{R}}^{n}\to{\mathbb{R}}^{n},\quad\psi(\xi)=|\xi|^{-1}\rho(|\xi|)\,\xi, (3.4)

then ψ\psi is smooth with the Jacobian:

Dψ(ξ)=|ξ|1ρ(|ξ|)I+(|ξ|2ρ(|ξ|)|ξ|3ρ(|ξ|))ξξT.D\psi(\xi)=|\xi|^{-1}{\rho(|\xi|)}I+(|\xi|^{-2}\rho^{\prime}(|\xi|)-|\xi|^{-3}\rho(|\xi|))\,\xi\cdot\xi^{T}. (3.5)

Let AA be an orthogonal matrix with nn-th column |ξ|1ξ|\xi|^{-1}\xi, then we have

ATDψ(ξ)A=diag[|ξ|1ρ(|ξ|),,|ξ|1ρ(|ξ|),ρ(|ξ|)].A^{T}D\psi(\xi)\,A=\textrm{diag}[\,|\xi|^{-1}{\rho(|\xi|)},\cdots,|\xi|^{-1}{\rho(|\xi|)},\,\rho^{\prime}(|\xi|)\,]. (3.6)

For θ\theta\in{\mathbb{R}}, we consider a family of deformations:

ϕθ(ξ)=ξ+θψ(ξ),\phi_{\theta}(\xi)=\xi+\theta\psi(\xi), (3.7)

and the corresponding unitary operators Uθ,θU_{\theta},\ \theta\in{\mathbb{R}} defined by

Uθu(ξ):=(detDϕθ(ξ))12u(ϕθ(ξ)).U_{\theta}u(\xi):=(\operatorname{det}D\phi_{\theta}(\xi))^{\frac{1}{2}}u(\phi_{\theta}(\xi)). (3.8)

Using (3.6), we can compute detDϕθ(ξ)\operatorname{det}D\phi_{\theta}(\xi) explicitly, i.e.

Jθ(ξ)detDϕθ(ξ)=det(I+θDψ(ξ))=(1+θρ(|ξ|))(1+θ|ξ|1ρ(|ξ|))n1,J_{\theta}(\xi)\equiv\operatorname{det}D\phi_{\theta}(\xi)=\operatorname{det}(I+\theta D\psi(\xi))=(1+\theta\rho^{\prime}(|\xi|)\,)\,(1+\theta|\xi|^{-1}{\rho(|\xi|)}\,)^{n-1}, (3.9)

then by (3.3), UθU_{\theta} is invertible as detDϕθ(ξ)0\operatorname{det}D\phi_{\theta}(\xi)\neq 0 for θ\theta\in{\mathbb{R}}, |θ|<γ|\theta|<\gamma, the inverse is given by

Uθ1v(ξ)=(detDϕθ(ϕθ1(ξ)))12v(ϕθ1(ξ)).U_{\theta}^{-1}v(\xi)=(\operatorname{det}D\phi_{\theta}(\phi_{\theta}^{-1}(\xi)))^{-\frac{1}{2}}v(\phi_{\theta}^{-1}(\xi)). (3.10)

Now we consider the deformed operators of ξ2+iεΔξ\xi^{2}+i\varepsilon\Delta_{\xi}:

Qε,θ:=Uθ(ξ2+iεΔξ)Uθ1=ϕθ(ξ)2iεJθ(ξ)12DξlJlj(ξ)Jθ(ξ)Jkj(ξ)DξkJθ(ξ)12\begin{split}Q_{\varepsilon,\theta}&:=U_{\theta}(\xi^{2}+i\varepsilon\Delta_{\xi})U_{\theta}^{-1}\\ &=\phi_{\theta}(\xi)^{2}-i\varepsilon J_{\theta}(\xi)^{-\frac{1}{2}}D_{\xi_{l}}J^{lj}(\xi)J_{\theta}(\xi)J^{kj}(\xi)D_{\xi_{k}}J_{\theta}(\xi)^{-\frac{1}{2}}\end{split} (3.11)

where Dξk=iξkD_{\xi_{k}}=-i\partial_{\xi_{k}}, Jθ(ξ)=detDϕθ(ξ)J_{\theta}(\xi)=\operatorname{det}D\phi_{\theta}(\xi), Jlj(ξ)=[Dϕθ(ξ)1]jlJ^{lj}(\xi)=[D\phi_{\theta}(\xi)^{-1}]_{jl}. To extend Qε,θQ_{\varepsilon,\theta} to θ\theta\in{\mathbb{C}}, we define

Dγ:={θ:|Reθ|+|Imθ|<γ}.D_{\gamma}:=\{\theta\in{\mathbb{C}}:|\operatorname{Re}\theta|+|\operatorname{Im}\theta|<\gamma\}. (3.12)

In view of (3.3) and (3.9), Dϕθ1D\phi_{\theta}^{-1} and detDϕθ\operatorname{det}D\phi_{\theta} extend analytically to θDγ\theta\in D_{\gamma}. Therefore, we obtain that Qε,θQ_{\varepsilon,\theta}, given by the second equation in (3.11), extends analytically to θDγ\theta\in D_{\gamma}.

Then we introduce some preliminary results about the spectrum of Qε,θQ_{\varepsilon,\theta} :

Proposition 1.

There exists constant ε0=ε0(Ω,γ)\varepsilon_{0}=\varepsilon_{0}(\Omega,\gamma) such that for all 0<ε<ε00<\varepsilon<\varepsilon_{0} and θDγ\theta\in D_{\gamma},

σ(Qε,θ){z:|z|>1,π/2<argz<π}=.\sigma(Q_{\varepsilon,\theta})\cap\{z\in{\mathbb{C}}:|z|>1,\pi/2<\arg z<\pi\}=\emptyset.
Proof.

We note that for θDγ\theta\in D_{\gamma}, by (3.3),

1tanπ8<1|θ||ρ(t)||1+θρ(t)|1+|θ||ρ(t)|<1+tanπ8,t0.1-\tan\frac{\pi}{8}<1-|\theta||\rho^{\prime}(t)|\leq|1+\theta\rho^{\prime}(t)|\leq 1+|\theta||\rho^{\prime}(t)|<1+\tan\frac{\pi}{8},\quad\forall\,t\geq 0.

Thus, (3.9) implies that C1<|Jθ(ξ)|<CC^{-1}<|J_{\theta}(\xi)|<C for some constant C>0C>0. Since

[Dϕθ(ξ)]jl=(1+θρ(|ξ|)|ξ|)δjl+θ|ξ|ρ(|ξ|)θρ(|ξ|)|ξ|3ξjξl[D\phi_{\theta}(\xi)]_{jl}=\left(1+\theta\frac{\rho(|\xi|)}{|\xi|}\right)\delta_{jl}+\frac{\theta|\xi|\rho^{\prime}(|\xi|)-\theta\rho(|\xi|)}{|\xi|^{3}}\xi_{j}\xi_{l}

by (3.5), and ρ𝒞c((0,))\rho^{\prime}\in{{\mathcal{C}}^{\infty}_{\rm{c}}}((0,\infty)), together with (3.9), we conclude that

Jθ,Jθ1,Jlj𝒞b(n), 1j,ln.J_{\theta},J_{\theta}^{-1},J^{lj}\in{{\mathcal{C}}^{\infty}_{b}}({\mathbb{R}}^{n}),\ 1\leq j,l\leq n. (3.13)

Here 𝒞b(n):={u𝒞(n):|αu|Cα for all α0n}{{\mathcal{C}}^{\infty}_{b}}({\mathbb{R}}^{n}):=\{u\in{{\mathcal{C}}^{\infty}}({\mathbb{R}}^{n}):|\partial^{\alpha}u|\leq C_{\alpha}\textrm{ for all }\alpha\in{\mathbb{N}}_{0}^{n}\}. Hence we have

Qε,θ=ϕθ(ξ)2iεJkj(ξ)Jlj(ξ)DξkDξl+εaj(ξ)Dξj+εb(ξ),Q_{\varepsilon,\theta}=\phi_{\theta}(\xi)^{2}-i\varepsilon J^{kj}(\xi)J^{lj}(\xi)D_{\xi_{k}}D_{\xi_{l}}+\varepsilon a_{j}(\xi)D_{\xi_{j}}+\varepsilon b(\xi), (3.14)

where aj,b𝒞b(n)a_{j},b\in{{\mathcal{C}}^{\infty}_{b}}({\mathbb{R}}^{n}). Let h=εh=\sqrt{\varepsilon}, then Qε,θ=qθ(ξ,hDξ;h)Q_{\varepsilon,\theta}=q_{\theta}(\xi,hD_{\xi};h) is a semiclassical differential operator – see Zworski (Zw, , §4), with the symbol

qθ(ξ,ξ;h)=ϕθ(ξ)2i(Dϕθ(ξ)2ξ)ξ+haj(ξ)ξj+h2b(ξ),q_{\theta}(\xi,\xi^{\ast};h)=\phi_{\theta}(\xi)^{2}-i(D\phi_{\theta}(\xi)^{-2}\xi^{\ast})\cdot\xi^{\ast}+ha_{j}(\xi)\xi_{j}^{\ast}+h^{2}b(\xi), (3.15)

where (ξ,ξ)(\xi,\xi^{\ast}) are coordinates of Tn\textrm{T}^{\ast}{\mathbb{R}}^{n}, Dϕθ(ξ)2=(Dϕθ(ξ)1)T(Dϕθ(ξ)1)D\phi_{\theta}(\xi)^{-2}=(D\phi_{\theta}(\xi)^{-1})^{T}(D\phi_{\theta}(\xi)^{-1}) since Dϕθ(ξ)D\phi_{\theta}(\xi) is a symmetric matrix. Choose m(ξ,ξ)=1+ξ2+ξ2m(\xi,\xi^{\ast})=1+\xi^{2}+{\xi^{\ast}}^{2} as an order function, we recall the symbol class S(m)S(m) from (Zw, , §4.4),

S(m):={a𝒞:|αa|Cαmfor α02n}.S(m):=\{a\in{{\mathcal{C}}^{\infty}}:|\partial^{\alpha}a|\leq C_{\alpha}m\quad\textrm{for }\forall\,\alpha\in{\mathbb{N}}_{0}^{2n}\}. (3.16)

Then by (3.3), (3.7) and (3.13), we have qθS(m)q_{\theta}\in S(m). Hence it suffices to show that there exists constant h0>0h_{0}>0 such that for h<h0h<h_{0},

qθz is elliptic in S(m) for |z|>1,π/2<argz<π.q_{\theta}-z\textrm{ is elliptic in }S(m)\textrm{ for }|z|>1,\ \pi/2<\arg z<\pi.

For a detailed introduction of general elliptic theory, we refer to (Zw, , §4).

Using (3.4) we calculate:

ϕθ(ξ)2=(ξ+θψ(ξ))(ξ+θψ(ξ))=(|ξ|+θρ(|ξ|))2.\phi_{\theta}(\xi)^{2}=(\xi+\theta\psi(\xi))\cdot(\xi+\theta\psi(\xi))=(|\xi|+\theta\rho(|\xi|))^{2}. (3.17)

Then for θDγ\theta\in D_{\gamma}, by (3.3), we have

π/4<argϕθ(ξ)2<π/4,|ϕθ(ξ)2|>(1tanπ8)2|ξ|2.-\pi/4<\arg\phi_{\theta}(\xi)^{2}<\pi/4,\quad|\phi_{\theta}(\xi)^{2}|>\left(1-\tan\frac{\pi}{8}\right)^{2}|\xi|^{2}. (3.18)

To obtain similar bounds for the argument and modulus of (Dϕθ(ξ)2ξ)ξ(D\phi_{\theta}(\xi)^{-2}\xi^{\ast})\cdot\xi^{\ast}, we recall (3.6) to compute

(Dϕθ2ξ)ξ=(1+θρ(|ξ|)|ξ|1)2(η12++ηn12)+(1+θρ(|ξ|))2ηn2,(D\phi_{\theta}^{-2}\xi^{\ast})\cdot\xi^{\ast}=(1+\theta{\rho(|\xi|)}{|\xi|}^{-1})^{-2}({\eta_{1}^{\ast}}^{2}+\cdots+{\eta_{n-1}^{\ast}}^{2})+(1+\theta\rho^{\prime}(|\xi|))^{-2}{\eta_{n}^{\ast}}^{2}, (3.19)

where η=ATξn\eta^{\ast}=A^{T}\xi^{\ast}\in{\mathbb{R}}^{n} with the same orthogonal matrix AA as in (3.6). By (3.3), for θDγ\theta\in D_{\gamma}, we have

±Imθ00±arg(1+θρ(|ξ|)|ξ|1),±arg(1+θρ(|ξ|))<π/8,\pm\operatorname{Im}\theta\geq 0\implies 0\leq\pm\arg(1+\theta{\rho(|\xi|)}{|\xi|}^{-1}),\,\pm\arg(1+\theta\rho^{\prime}(|\xi|))<\pi/8,

Hence, for all θDγ\theta\in D_{\gamma},

±Imθ00arg(Dϕθ2ξ)ξ<π/4,\pm\operatorname{Im}\theta\geq 0\implies 0\leq\mp\arg\,(D\phi_{\theta}^{-2}\xi^{\ast})\cdot\xi^{\ast}<\pi/4, (3.20)

and by applying the following basic inequality with (3.3) to (3.19),

|r1eiθ1+r2eiθ2|2=r12+r22+2r1r2cos(θ1θ2)1|cos(θ1θ2)|2(r1+r2)2,|r_{1}e^{i\theta_{1}}+r_{2}e^{i\theta_{2}}|^{2}=r_{1}^{2}+r_{2}^{2}+2r_{1}r_{2}\cos(\theta_{1}-\theta_{2})\geq\frac{1-|\cos(\theta_{1}-\theta_{2})|}{2}(r_{1}+r_{2})^{2}, (3.21)

we also obtain that for all θDγ\theta\in D_{\gamma},

|(Dϕθ2ξ)ξ|C|η|2=C|ξ|2.|(D\phi_{\theta}^{-2}\xi^{\ast})\cdot\xi^{\ast}|\geq C|\eta^{\ast}|^{2}=C|\xi^{\ast}|^{2}. (3.22)

Since arg(ϕθ(ξ)2z)(π/2,π/4)\arg(\phi_{\theta}(\xi)^{2}-z)\in(-\pi/2,\pi/4) for π/2<argz<π\pi/2<\arg z<\pi and argi(Dϕθ2ξ)ξ(3π/4,π/4)\arg-i(D\phi_{\theta}^{-2}\xi^{\ast})\cdot\xi^{\ast}\in(-3\pi/4,-\pi/4) by (3.20), using (3.21) together with (3.18) and (3.22), we have

|ϕθ(ξ)2zi(Dϕθ2ξ)ξ|C|ϕθ(ξ)2z|+C|i(Dϕθ2ξ)ξ|C|ϕθ(ξ)2|+C|z|+C|ξ|2C(1+|ξ|2+|ξ|2)=Cm.\begin{split}|\phi_{\theta}(\xi)^{2}-z-i(D\phi_{\theta}^{-2}\xi^{\ast})\cdot\xi^{\ast}|&\geq C|\phi_{\theta}(\xi)^{2}-z|+C|-i(D\phi_{\theta}^{-2}\xi^{\ast})\cdot\xi^{\ast}|\\ &\geq C|\phi_{\theta}(\xi)^{2}|+C|z|+C|\xi^{\ast}|^{2}\\ &\geq C(1+|\xi|^{2}+|\xi^{\ast}|^{2})=Cm.\end{split} (3.23)

Then by (3.15), we conclude that there exists h0>0h_{0}>0 such that for all h<h0h<h_{0}, |qθz|Cm|q_{\theta}-z|\geq Cm, which completes the proof. ∎

Refer to caption
Figure 1: An illustration of the results of Proposition 2 in the case of dim=1\dim=1, β=0.4\beta=0.4, which shows that the numerical range of the principal symbol of Qε,0.4iQ_{\varepsilon,-0.4i} avoids the region {λ2:λΩ}\{\lambda^{2}:\lambda\in\Omega\}. We choose ρ()=0.4tanh()\rho(\cdot)=0.4\tanh(\cdot) to compute the numerical range of (ϕ0.4i(ξ)2i(ϕ0.4i(ξ))2ξ2)1/2(\phi_{-0.4i}(\xi)^{2}-i(\phi_{-0.4i}^{\prime}(\xi))^{-2}{\xi^{\ast}}^{2})^{1/2}.
Proposition 2.

For any β(γ,γ)\beta\in(\gamma^{\prime},\gamma) satisfying

Ω{x+iy:x>0,y>βρ(x)},\Omega\Subset\{x+iy:x>0,y>-\beta\rho(x)\}, (3.24)

there exists ε0=ε0(Ω,γ,β)\varepsilon_{0}=\varepsilon_{0}(\Omega,\gamma,\beta) such that for all 0<ε<ε00<\varepsilon<\varepsilon_{0},

σ(Qε,iβ){λ2:λΩ}=.\sigma(Q_{\varepsilon,-i\beta})\cap\{\lambda^{2}:\lambda\in\Omega\}=\emptyset.
Proof.

As in the proof of Proposition 1, it suffices to show that there exists h0=h0(Ω,γ,β)h_{0}=h_{0}(\Omega,\gamma,\beta) such that for 0<h<h00<h<h_{0},

qiβ(ξ,ξ;h)λ2 is elliptic in S(m)for λΩ.q_{-i\beta}(\xi,\xi^{\ast};h)-\lambda^{2}\textrm{ is elliptic in }S(m)\ \textrm{for }\lambda\in\Omega.

Recalling argi(Dϕiβ2ξ)ξ[π/2,π/4)\arg-i(D\phi_{-i\beta}^{-2}\xi^{\ast})\cdot\xi^{\ast}\in[-\pi/2,-\pi/4) by (3.20), in order to apply (3.21), we claim that

δ>0 s.t. arg(ϕiβ(ξ)2λ2)π/2δ or 3π/4+δ,for all λΩ,ξn.\exists\,\delta>0\textrm{ s.t. }\arg(\phi_{-i\beta}(\xi)^{2}-\lambda^{2})\leq\pi/2-\delta\textrm{ or }\geq 3\pi/4+\delta,\ \textrm{for all }\lambda\in\Omega,\,\xi\in{\mathbb{R}}^{n}. (3.25)

We notice that for |ξ|1|\xi|\gg 1, ϕiβ(ξ)2=(|ξ|iβ)2\phi_{-i\beta}(\xi)^{2}=(|\xi|-i\beta)^{2} by (3.17), thus arg(ϕiβ(ξ)2λ2)(π/4,0)\arg(\phi_{-i\beta}(\xi)^{2}-\lambda^{2})\in(-\pi/4,0), in other words, there exists some large RR such that (3.25) holds for |ξ|>R|\xi|>R with δ=π/2\delta=\pi/2. It remains to show that (3.25) holds for all |ξ|R|\xi|\leq R and λΩ\lambda\in\Omega. We argue by contradiction: if it does not hold, there must exist λΩ¯\lambda\in\overline{\Omega}, ξn\xi\in{\mathbb{R}}^{n} such that arg(ϕiβ(ξ)2λ2)[π/2,3π/4]\arg(\phi_{-i\beta}(\xi)^{2}-\lambda^{2})\in[\pi/2,3\pi/4], i.e.

0Re((|ξ|iβρ(|ξ|))2λ2)Im((|ξ|iβρ(|ξ|))2λ2),0\leq-\operatorname{Re}\,((|\xi|-i\beta\rho(|\xi|))^{2}-\lambda^{2})\leq\operatorname{Im}\,((|\xi|-i\beta\rho(|\xi|))^{2}-\lambda^{2}),

which immediately implies Imλ0\operatorname{Im}\lambda\leq 0. Let t=|ξ|t=|\xi| and write λ=xiy\lambda=x-iy, then we have

x2y2t2+β2ρ(t)2\displaystyle x^{2}-y^{2}-t^{2}+\beta^{2}\rho(t)^{2} 2xy2βtρ(t)\displaystyle\leq 2xy-2\beta t\rho(t) (3.26)
βtρ(t)\displaystyle\beta t\rho(t) xy\displaystyle\leq xy (3.27)

Since x>0x>0 and 0y<βρ(x)0\leq y<\beta\rho(x) by (3.24), then (3.26) implies that

x22βxρ(x)β2ρ(x)2<t22βtρ(t)β2ρ(t)2.x^{2}-2\beta x\rho(x)-\beta^{2}\rho(x)^{2}<t^{2}-2\beta t\rho(t)-\beta^{2}\rho(t)^{2}.

Let S(x)=x22βxρ(x)β2ρ(x)2S(x)=x^{2}-2\beta x\rho(x)-\beta^{2}\rho(x)^{2}, by (3.3),

S(x)=2x(1βρ(x)xβρ(x)βρ(x)xβρ(x))>2x(12tanπ8tan2π8)=0,\begin{split}S^{\prime}(x)&=2x\left(1-\beta\frac{\rho(x)}{x}-\beta\rho^{\prime}(x)-\beta\frac{\rho(x)}{x}\cdot\beta\rho^{\prime}(x)\right)\\ &>2x\left(1-2\tan\frac{\pi}{8}-\tan^{2}\frac{\pi}{8}\right)=0,\end{split}

thus S(x)<S(t)x<tS(x)<S(t)\implies x<t. Recalling that ρ\rho is non-decreasing, we have βtρ(t)βxρ(x)>xy\beta t\rho(t)\geq\beta x\rho(x)>xy, which contradicts (3.27). Hence (3.25) holds, using (3.21) and (3.22), we obtain that

|ϕiβ(ξ)2λ2i(Dϕiβ2ξ)ξ|C(δ)(|(|ξ|iβρ(|ξ|))2λ2|+|ξ|2).|\phi_{-i\beta}(\xi)^{2}-\lambda^{2}-i(D\phi_{-i\beta}^{-2}\xi^{\ast})\cdot\xi^{\ast}|\geq C(\delta)(|(|\xi|-i\beta\rho(|\xi|))^{2}-\lambda^{2}|+|\xi^{\ast}|^{2}).

Since for |ξ|1|\xi|\gg 1,

|(|ξ|iβρ(|ξ|))2λ2|=|(|ξ|iβ)2λ2||ξ|2β2|λ|2,|(|\xi|-i\beta\rho(|\xi|))^{2}-\lambda^{2}|=|(|\xi|-i\beta)^{2}-\lambda^{2}|\geq|\xi|^{2}-\beta^{2}-|\lambda|^{2},

there exists R=R(Ω,β)>0R=R(\Omega,\beta)>0 such that |(|ξ|iβρ(|ξ|))2λ2|(1+|ξ|2)/2|(|\xi|-i\beta\rho(|\xi|))^{2}-\lambda^{2}|\geq(1+|\xi|^{2})/2 whenever |ξ|>R|\xi|>R. We also note that, by (3.24),

dist({tiβρ(t):t0},±Ω)C=C(Ω,γ,β)>0,\textrm{dist}\,(\{t-i\beta\rho(t):t\geq 0\},\,\pm\Omega)\geq C=C(\Omega,\gamma,\beta)>0,

thus |(|ξ|iβρ(|ξ|))2λ2|C2C2(1+R2)1(1+|ξ|2)|(|\xi|-i\beta\rho(|\xi|))^{2}-\lambda^{2}|\geq C^{2}\geq C^{2}(1+R^{2})^{-1}(1+|\xi|^{2}) for |ξ|R|\xi|\leq R. Hence |ϕiβ(ξ)2λ2i(Dϕiβ2ξ)ξ|C(1+|ξ|2+|ξ|2)|\phi_{-i\beta}(\xi)^{2}-\lambda^{2}-i(D\phi_{-i\beta}^{-2}\xi^{\ast})\cdot\xi^{\ast}|\geq C(1+|\xi|^{2}+|\xi^{\ast}|^{2}), where CC determined by Ω,γ,β\Omega,\gamma,\beta. Then by (3.15), we conclude that there exist h0=h0(Ω,γ,β)h_{0}=h_{0}(\Omega,\gamma,\beta) and C=C(Ω,γ,β)>0C=C(\Omega,\gamma,\beta)>0 such that

for all 0<h<h0,λΩ,|qiβ(ξ,ξ;h)λ2|Cm(ξ,ξ),\textrm{for all }0<h<h_{0},\,\lambda\in\Omega,\quad|q_{-i\beta}(\xi,\xi^{\ast};h)-\lambda^{2}|\geq Cm(\xi,\xi^{\ast}), (3.28)

which completes the proof. ∎

Now we state the main result of this section:

Lemma 2.

For any 0<a<a<b0<a^{\prime}<a<b and γ<γ\gamma^{\prime}<\gamma such that the rectangle

Ω:=(a,a)+i(γ,b){λ:π/8<argλ<7π/8},\Omega:=(a^{\prime},a)+i(-\gamma^{\prime},b)\Subset\{\lambda\in{\mathbb{C}}:-\pi/8<\arg\lambda<7\pi/8\},

there exist constant C=C(Ω,γ)>0C=C(\Omega,\gamma)>0 and ε0=ε0(Ω,γ)>0\varepsilon_{0}=\varepsilon_{0}(\Omega,\gamma)>0 such that uniformly for 0<ε<ε00<\varepsilon<\varepsilon_{0},

eγ|x|(Δiεx2λ2)1eγ|x|L2L2C,λΩ.\|e^{-\gamma|x|}(-\Delta-i\varepsilon x^{2}-\lambda^{2})^{-1}e^{-\gamma|x|}\|_{L^{2}\to L^{2}}\leq C,\quad\forall\,\lambda\in\Omega.
Proof.

We consider the matrix element

Bf,gε(λ):=eγ|x|(Δiεx2λ2)1eγ|x|f,gLx2,for f,gL2(n),B^{\varepsilon}_{f,g}(\lambda):=\langle e^{-\gamma|x|}(-\Delta-i\varepsilon x^{2}-\lambda^{2})^{-1}e^{-\gamma|x|}f,g\rangle_{L_{x}^{2}},\quad\textrm{for }f,g\in L^{2}({\mathbb{R}}^{n}),

where u,vLx2=nuv¯𝑑x\langle u,v\rangle_{L_{x}^{2}}=\int_{{\mathbb{R}}^{n}}u\bar{v}\,dx is the standard L2L^{2} inner product. It suffices to show that there exist C,ε0C,\varepsilon_{0} such that uniformly for 0<ε<ε00<\varepsilon<\varepsilon_{0},

|Bf,gε(λ)|CfL2gL2,for all f,gL2,λΩ.|B^{\varepsilon}_{f,g}(\lambda)|\leq C\|f\|_{L^{2}}\|g\|_{L^{2}},\quad\textrm{for all }f,g\in L^{2},\ \lambda\in\Omega. (3.29)

Recalling (3.1), both Δxiεx2λ2-\Delta_{x}-i\varepsilon x^{2}-\lambda^{2} and ξ2+iεΔξλ2\xi^{2}+i\varepsilon\Delta_{\xi}-\lambda^{2} are invertible for λΩ\lambda\in\Omega. Then we have

Bf,gε(λ)=(Δxiεx2λ2)1eγ|x|f,eγ|x|gLx2=1(ξ2+iεΔξλ2)1eγ|x|f,eγ|x|gLx2=(ξ2+iεΔξλ2)1(eγ|x|f)(ξ),(eγ|x|g)(ξ)Lξ2.\begin{split}B^{\varepsilon}_{f,g}(\lambda)&=\langle\,(-\Delta_{x}-i\varepsilon x^{2}-\lambda^{2})^{-1}e^{-\gamma|x|}f,\,e^{-\gamma|x|}g\rangle_{L_{x}^{2}}\\ &=\langle{\mathcal{F}}^{-1}(\xi^{2}+i\varepsilon\Delta_{\xi}-\lambda^{2})^{-1}{\mathcal{F}}e^{-\gamma|x|}f,\,e^{-\gamma|x|}g\rangle_{L_{x}^{2}}\\ &=\langle\,(\xi^{2}+i\varepsilon\Delta_{\xi}-\lambda^{2})^{-1}{\mathcal{F}}(e^{-\gamma|x|}f)(\xi),\,{\mathcal{F}}(e^{-\gamma|x|}g)(\xi)\rangle_{L_{\xi}^{2}}.\end{split} (3.30)

Let Fγ(ξ):=(eγ|x|f)(ξ)F_{\gamma}(\xi):={\mathcal{F}}(e^{-\gamma|x|}f)(\xi) and Gγ(ξ):=(eγ|x|g)(ξ)G_{\gamma}(\xi):={\mathcal{F}}(e^{-\gamma|x|}g)(\xi), recalling the formula

(e|x|)(ξ)=cn(1+ξ2)n+12,cn=(2π)n2Γ((n+1)/2)πn+12,{\mathcal{F}}(e^{-|x|})(\xi)=c_{n}(1+\xi^{2})^{-\frac{n+1}{2}},\quad c_{n}=(2\pi)^{\frac{n}{2}}\Gamma((n+1)/2)\pi^{-\frac{n+1}{2}},

then Fγ=Kγf^F_{\gamma}=K_{\gamma}\ast\hat{f} and Gγ=Kγg^G_{\gamma}=K_{\gamma}\ast\hat{g}, where Kγ(ξ)=cnγ(γ2+ξ2)n+12K_{\gamma}(\xi)=c_{n}\gamma\,(\gamma^{2}+\xi^{2})^{-\frac{n+1}{2}}.

First we consider, for θ\theta\in{\mathbb{R}}, |θ|<γ|\theta|<\gamma and UθU_{\theta} defined by (3.8), the integral kernel of the map Uθ(Kγ)U_{\theta}\circ(K_{\gamma}\,\ast\ ):

K(ξ,η;θ):=(detDϕθ(ξ))12Kγ(ϕθ(ξ)η),ξ,ηn.K(\xi,\eta;\theta):=(\operatorname{det}D\phi_{\theta}(\xi))^{\frac{1}{2}}K_{\gamma}(\phi_{\theta}(\xi)-\eta),\quad\xi,\eta\in{\mathbb{R}}^{n}.

We claim that K(ξ,η;θ)K(\xi,\eta;\theta) has an analytic extension to θDγ\theta\in D_{\gamma}. Since KγK_{\gamma} extends analytically to the strip {ξn:|Imξ|<γ}\{\xi\in{\mathbb{C}}^{n}:|\operatorname{Im}\xi|<\gamma\}, it suffices to show that |Im(ϕθ(ξ)η)|=|Imθψ(ξ)|<γ|\operatorname{Im}(\phi_{\theta}(\xi)-\eta)|=|\operatorname{Im}\theta\psi(\xi)|<\gamma, which is a direct consequence of θDγ\theta\in D_{\gamma} and |ψ(ξ)|1|\psi(\xi)|\leq 1 by (3.4). Then for θDγ\theta\in D_{\gamma}, using (3.3) and (3.9), we can estimate K(ξ,η;θ)K(\xi,\eta;\theta) as follows:

|K(ξ,η;θ)|Cγ|γ2+(ξ+θψ(ξ)η)2|n+12Cγ|γ2|Imθ|2|ψ(ξ)|2+(ξη+Reθψ(ξ))2|n+12Cγ(γ2|Imθ|2+(|ξη||Reθ|)2)n+12\begin{split}|K(\xi,\eta;\theta)|&\leq C\gamma\,|\gamma^{2}+(\xi+\theta\psi(\xi)-\eta)^{2}|^{-\frac{n+1}{2}}\\ &\leq C\gamma\,|\gamma^{2}-|\operatorname{Im}\theta|^{2}|\psi(\xi)|^{2}+(\xi-\eta+\operatorname{Re}\theta\psi(\xi))^{2}|^{-\frac{n+1}{2}}\\ &\leq C\gamma\,(\gamma^{2}-|\operatorname{Im}\theta|^{2}+(|\xi-\eta|-|\operatorname{Re}\theta|)^{2})^{-\frac{n+1}{2}}\end{split}

thus

max{supξnn|K(ξ,η;θ)|𝑑η,supηnn|K(ξ,η;θ)|𝑑ξ}Cγxn(γ2|Imθ|2+(|x||Reθ|)2)n+12𝑑xC(γ,θ).\begin{split}{}&\quad\max\,\{\,\sup_{\xi\in{\mathbb{R}}^{n}}\int_{{\mathbb{R}}^{n}}|K(\xi,\eta;\theta)|d\eta,\ \sup_{\eta\in{\mathbb{R}}^{n}}\int_{{\mathbb{R}}^{n}}|K(\xi,\eta;\theta)|d\xi\,\}\\ &\leq C\gamma\int_{x\in{\mathbb{R}}^{n}}(\gamma^{2}-|\operatorname{Im}\theta|^{2}+(|x|-|\operatorname{Re}\theta|)^{2})^{-\frac{n+1}{2}}dx\leq C(\gamma,\theta).\end{split} (3.31)

Hence, by Schur’s criterion, Uθ(Kγ)U_{\theta}\circ(K_{\gamma}\,\ast\ ), first defined for θDγ\theta\in D_{\gamma}\cap{\mathbb{R}}, with the integral kernel K(ξ,η;θ)K(\xi,\eta;\theta), extends to θDγ\theta\in D_{\gamma} as an analytic family of operators L2L2L^{2}\to L^{2}. In particular,

DγθUθFγ=Uθ(Kγf^) and UθGγ=Uθ(Kγg^),D_{\gamma}\owns\theta\mapsto U_{\theta}F_{\gamma}=U_{\theta}(K_{\gamma}\ast\hat{f})\textrm{ and }U_{\theta}G_{\gamma}=U_{\theta}(K_{\gamma}\ast\hat{g}),

are two analytic families of functions in L2(n)L^{2}({\mathbb{R}}^{n}).

Now we define

Bf,gε(λ;θ)=(Qε,θλ2)1UθFγ,Uθ¯GγB^{\varepsilon}_{f,g}(\lambda;\theta)=\langle\,(Q_{\varepsilon,\theta}-\lambda^{2})^{-1}U_{\theta}F_{\gamma},U_{\bar{\theta}}G_{\gamma}\rangle

for θDγ\theta\in D_{\gamma}, with Qε,θQ_{\varepsilon,\theta} given by (3.11), where we write Uθ¯GγU_{\bar{\theta}}G_{\gamma} instead of UθGγU_{\theta}G_{\gamma}. Then by Proposition 1, there exists ε0=ε0(Ω,γ)\varepsilon_{0}=\varepsilon_{0}(\Omega,\gamma) such that for all 0<ε<ε00<\varepsilon<\varepsilon_{0}, and |λ|>1|\lambda|>1 with π/4<argλ<π/2\pi/4<\arg\lambda<\pi/2,

DγθBf,gε(λ;θ) is analytic. D_{\gamma}\owns\theta\mapsto B^{\varepsilon}_{f,g}(\lambda;\theta)\textrm{ is analytic. }

However, for θDγ\theta\in D_{\gamma}\cap{\mathbb{R}}, since UθU_{\theta} is unitary, by (3.30) we have

Bf,gε(λ;θ)=Uθ(ξ2+iεΔξλ2)1Uθ1UθFγ,UθGγ=Uθ(ξ2+iεΔξλ2)1Fγ,UθGγ=(ξ2+iεΔξλ2)1Fγ,Gγ=Bf,gε(λ).\begin{split}B_{f,g}^{\varepsilon}(\lambda;\theta)&=\langle U_{\theta}(\xi^{2}+i\varepsilon\Delta_{\xi}-\lambda^{2})^{-1}U_{\theta}^{-1}U_{\theta}F_{\gamma},\,U_{\theta}G_{\gamma}\rangle\\ &=\langle U_{\theta}(\xi^{2}+i\varepsilon\Delta_{\xi}-\lambda^{2})^{-1}F_{\gamma},\,U_{\theta}G_{\gamma}\rangle\\ &=\langle\,(\xi^{2}+i\varepsilon\Delta_{\xi}-\lambda^{2})^{-1}F_{\gamma},\,G_{\gamma}\rangle=B_{f,g}^{\varepsilon}(\lambda).\end{split}

Thus by analyticity, Bf,gε(λ;θ)Bf,gε(λ),θDγB_{f,g}^{\varepsilon}(\lambda;\theta)\equiv B_{f,g}^{\varepsilon}(\lambda),\ \forall\,\theta\in D_{\gamma} whenever |λ|>1|\lambda|>1, π/4<argλ<π/2\pi/4<\arg\lambda<\pi/2. In particular, for fixed β(γ,γ)\beta\in(\gamma^{\prime},\gamma) satisfying (3.24),

Bf,gε(λ)=Bf,gε(λ;iβ) whenever |λ|>1,π/4<argλ<π/2.B_{f,g}^{\varepsilon}(\lambda)=B_{f,g}^{\varepsilon}(\lambda;-i\beta)\textrm{ whenever }|\lambda|>1,\,\pi/4<\arg\lambda<\pi/2.

In view of Proposition 2 and (3.1), both Bf,gε(λ)B_{f,g}^{\varepsilon}(\lambda) and Bf,gε(λ;iβ)B_{f,g}^{\varepsilon}(\lambda;-i\beta) are analytic in Ω\Omega. Without loss of generality, we may assume that a>1a>1 in (1.4), then

Ω{λ:|λ|>1,π/4<argλ<π/2},\Omega\cap\{\lambda:|\lambda|>1,\pi/4<\arg\lambda<\pi/2\}\neq\emptyset,

where Bf,gε(λ)B_{f,g}^{\varepsilon}(\lambda) and Bf,gε(λ;iβ)B_{f,g}^{\varepsilon}(\lambda;-i\beta) coincide. Hence by analyticity, we conclude that for each 0<ε<ε00<\varepsilon<\varepsilon_{0},

Bf,gε(λ)=Bf,gε(λ;iβ)=(Qε,iβλ2)1UiβFγ,UiβGγ,λΩ.B_{f,g}^{\varepsilon}(\lambda)=B_{f,g}^{\varepsilon}(\lambda;-i\beta)=\langle\,(Q_{\varepsilon,-i\beta}-\lambda^{2})^{-1}U_{-i\beta}F_{\gamma},\,U_{i\beta}G_{\gamma}\,\rangle,\quad\forall\,\lambda\in\Omega. (3.32)

By the elliptic theory of semiclassical differential operators – see (Zw, , §4.7), (3.28) implies that there exists ε0=ε0(Ω,γ,β)\varepsilon_{0}=\varepsilon_{0}(\Omega,\gamma,\beta) such that for all 0<ε<ε00<\varepsilon<\varepsilon_{0},

(Qε,iβλ2)1L2L2C(Ω,γ,β),λΩ.\|(Q_{\varepsilon,-i\beta}-\lambda^{2})^{-1}\|_{L^{2}\to L^{2}}\leq C(\Omega,\gamma,\beta),\quad\forall\,\lambda\in\Omega. (3.33)

Recalling (3.31), by Schur’s criterion, we obtain that

UiβFγL2=Uiβ(Kγf^)L2C(γ,β)f^L2=C(γ,β)fL2UiβGγL2=Uiβ(Kγg^)L2C(γ,β)g^L2=C(γ,β)gL2\begin{split}\|U_{-i\beta}F_{\gamma}\|_{L^{2}}&=\|U_{-i\beta}\circ(K_{\gamma}\ast\hat{f})\|_{L^{2}}\leq C(\gamma,\beta)\|\hat{f}\|_{L^{2}}=C(\gamma,\beta)\|f\|_{L^{2}}\\ \|U_{i\beta}G_{\gamma}\|_{L^{2}}&=\|U_{i\beta}\circ(K_{\gamma}\ast\hat{g})\|_{L^{2}}\leq C(\gamma,\beta)\|\hat{g}\|_{L^{2}}=C(\gamma,\beta)\|g\|_{L^{2}}\end{split} (3.34)

Combining (3.32), (3.33) and (3.34), also noticing that β\beta can be determined by Ω,γ\Omega,\gamma, we obtain (3.29) with C=C(Ω,γ)C=C(\Omega,\gamma), which completes the proof. ∎

4 eigenvalues of the regularized operator

In this section we will review the meromorphy of the resolvent

RV,ε(λ):=(Pελ2)1,ε>0,R_{V,\varepsilon}(\lambda):=(P_{\varepsilon}-\lambda^{2})^{-1},\quad\varepsilon>0,

with PεP_{\varepsilon} in (1.2), in a similar form to the meromorphic continuation of the weighted resolvent VRV(λ)V\sqrt{V}R_{V}(\lambda)\sqrt{V} given by (2.1).

First we write Rε(λ):=(Δiεx2λ2)1R_{\varepsilon}(\lambda):=(-\Delta-i\varepsilon x^{2}-\lambda^{2})^{-1} and recall

Rε(λ)=𝒪δ(1/|λ|):L2L2,δ<argλ<3π/4δ,|λ|>δ,R_{\varepsilon}(\lambda)=\mathcal{O}_{\delta}(1/|\lambda|):L^{2}\to L^{2},\quad\delta<\arg\lambda<3\pi/4-\delta,\ |\lambda|>\delta, (4.1)

which follows from (semiclassical) ellipticity. Then

(Pελ2)Rε(λ)=I+VRε(λ),π/8<argλ<7π/8.(P_{\varepsilon}-\lambda^{2})R_{\varepsilon}(\lambda)=I+VR_{\varepsilon}(\lambda),\quad-\pi/8<\arg\lambda<7\pi/8. (4.2)

In view of (4.1), I+VRε(λ)I+VR_{\varepsilon}(\lambda) is invertible for π/4<argλ<π/2\pi/4<\arg\lambda<\pi/2, |λ|1|\lambda|\gg 1. Since Rε(λ):L2H2R_{\varepsilon}(\lambda):L^{2}\to H^{2} is analytic in {λ:π/8<argλ<7π/8}\{\lambda:-\pi/8<\arg\lambda<7\pi/8\}, see (3.1), V:H2L2V:H^{2}\to L^{2} is compact by (1.1), we have λVRε(λ)\lambda\mapsto VR_{\varepsilon}(\lambda) is an analytic family of compact operators for π/8<argλ<7π/8-\pi/8<\arg\lambda<7\pi/8. Hence λ(I+VRε(λ))1\lambda\mapsto(I+VR_{\varepsilon}(\lambda))^{-1} is a meromorphic family of operators in the same range of λ\lambda. Using (4.2), we conclude that RV,ε(λ)=Rε(λ)(I+VRε(λ))1R_{V,\varepsilon}(\lambda)=R_{\varepsilon}(\lambda)(I+VR_{\varepsilon}(\lambda))^{-1} is meromorphic for π/8<argλ<7π/8-\pi/8<\arg\lambda<7\pi/8 (in fact RV,ε(λ)R_{V,\varepsilon}(\lambda) is meromorphic for λ\lambda\in{\mathbb{C}} by the Gohberg–Sigal factorization theorem - see (res, , §C.4)), with poles {λj(ε)}j=1\{\lambda_{j}(\varepsilon)\}_{j=1}^{\infty}, i.e. {λj(ε)2}j=1\{\lambda_{j}(\varepsilon)^{2}\}_{j=1}^{\infty} are the eigenvalues of PεP_{\varepsilon} in {z:argzπ/4}\{z\in{\mathbb{C}}:\arg z\neq-\pi/4\}. Then we have

Lemma 3.

For each ε>0\varepsilon>0,

λ(I+VRε(λ)V)1,π/8<argλ<7π/8,\lambda\mapsto(I+\sqrt{V}R_{\varepsilon}(\lambda)\sqrt{V})^{-1},\quad-\pi/8<\arg\lambda<7\pi/8,

is a meromorphic family of operators on L2(n)L^{2}({\mathbb{R}}^{n}) with poles of finite rank. Moreover,

mε(λ):=12πitrλ(I+VRε(ζ)V)1ζ(VRε(ζ)V)dζ,m_{\varepsilon}(\lambda):=\frac{1}{2\pi i}\operatorname{tr}\oint_{\lambda}(I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})^{-1}\partial_{\zeta}(\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})\,d\zeta, (4.3)

where the integral is over a positively oriented circle enclosing λ\lambda and containing no poles other than possibly λ\lambda, satisfies

mε(λ)=12πitrλ(ζ2Pε)12ζ𝑑ζ.m_{\varepsilon}(\lambda)=\frac{1}{2\pi i}\operatorname{tr}\oint_{\lambda}(\zeta^{2}-P_{\varepsilon})^{-1}2\zeta\,d\zeta. (4.4)

Remark. The multiplicity of an eigenvalue λ2\lambda^{2} of PεP_{\varepsilon} can be defined by the right side of (4.4), thus Lemma 3 implies that the poles of (I+VRε(λ)V)1(I+\sqrt{V}R_{\varepsilon}(\lambda)\sqrt{V})^{-1} coincide with {λj(ε)}j=1\{\lambda_{j}(\varepsilon)\}_{j=1}^{\infty} given in (1.3), with agreement of multiplicities.

Proof.

Following the above argument, it easy to see that λVRε(λ)V\lambda\mapsto\sqrt{V}R_{\varepsilon}(\lambda)\sqrt{V} is an analytic family of compact operators for π/8<argλ<7π/8-\pi/8<\arg\lambda<7\pi/8. Then

λ(I+VRε(λ)V)1,π/8<argλ<7π/8,\lambda\mapsto(I+\sqrt{V}R_{\varepsilon}(\lambda)\sqrt{V})^{-1},\quad-\pi/8<\arg\lambda<7\pi/8,

is a meromorphic family of operators, since I+VRε(λ)VI+\sqrt{V}R_{\varepsilon}(\lambda)\sqrt{V} is invertible for π/4<argλ<π/2\pi/4<\arg\lambda<\pi/2, |λ|1|\lambda|\gg 1 by (4.1). In this range of λ\lambda, I+VRε(λ)I+VR_{\varepsilon}(\lambda) is also invertible by the Neumann series argument, thus we have

(Pελ2)1=Rε(λ)(I+VRε(λ))1=Rε(λ)j=0(1)j(VRε(λ))j=Rε(λ)(IVj=0(1)j(VRε(λ)V)jVRε(λ))=Rε(λ)[IV(I+VRε(λ)V)1VRε(λ)].\begin{split}(P_{\varepsilon}-\lambda^{2})^{-1}&=R_{\varepsilon}(\lambda)(I+VR_{\varepsilon}(\lambda))^{-1}\\ &=R_{\varepsilon}(\lambda)\sum_{j=0}^{\infty}(-1)^{j}(VR_{\varepsilon}(\lambda))^{j}\\ &=R_{\varepsilon}(\lambda)(I-\sqrt{V}\,\sum_{j=0}^{\infty}(-1)^{j}(\sqrt{V}R_{\varepsilon}(\lambda)\sqrt{V})^{j}\,\sqrt{V}R_{\varepsilon}(\lambda))\\ &=R_{\varepsilon}(\lambda)[\,I-\sqrt{V}(I+\sqrt{V}R_{\varepsilon}(\lambda)\sqrt{V})^{-1}\sqrt{V}R_{\varepsilon}(\lambda)\,].\end{split} (4.5)

Since both sides of (4.5) are meromorphic for π/8<argλ<7π/8-\pi/8<\arg\lambda<7\pi/8, by meromorphy, we conclude that (4.5) holds for all π/8<argλ<7π/8-\pi/8<\arg\lambda<7\pi/8, as an identity between meromorphic families of operators.

To obtain the multiplicity formula, we fix any λ\lambda with π/8<argλ<7π/8-\pi/8<\arg\lambda<7\pi/8, then there exists a neighborhood λU\lambda\in U in this half plane and finite rank operators AjA_{j}, 1jJ1\leq j\leq J such that (I+VRε(ζ)V)1j=1JAj(ζλ)j(I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})^{-1}-\sum_{j=1}^{J}\frac{A_{j}}{(\zeta-\lambda)^{j}} is analytic in ζU\zeta\in U. Let 𝒞λU\mathcal{C}_{\lambda}\subset U be a positively oriented circle enclosing λ\lambda and containing no poles of (I+VRε(ζ)V)1(I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})^{-1} other than possibly λ\lambda, thus it also contains no poles of (ζ2Pε)1(\zeta^{2}-P_{\varepsilon})^{-1} other than possibly λ\lambda as a consequence of (4.5). On the one hand, we can compute

mε(λ)=12πitr𝒞λ(I+VRε(ζ)V)1VRε(ζ)2V 2ζ𝑑ζ=12πitr𝒞λj=1JAjVRε(ζ)22ζV(ζλ)jdζ=j=1Jk=0j11k!(j1k)!trAjVζkRε(ζ)ζj1k(Rε(ζ)2ζ)V.\begin{split}m_{\varepsilon}(\lambda)&=\frac{1}{2\pi i}\operatorname{tr}\int_{\mathcal{C}_{\lambda}}(I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})^{-1}\sqrt{V}R_{\varepsilon}(\zeta)^{2}\sqrt{V}\,2\zeta d\zeta\\ &=\frac{1}{2\pi i}\operatorname{tr}\int_{\mathcal{C}_{\lambda}}\sum_{j=1}^{J}\frac{A_{j}\sqrt{V}R_{\varepsilon}(\zeta)^{2}2\zeta\sqrt{V}}{(\zeta-\lambda)^{j}}d\zeta\\ &=\sum_{j=1}^{J}\sum_{k=0}^{j-1}\frac{1}{k!(j-1-k)!}\operatorname{tr}A_{j}\sqrt{V}\,\partial_{\zeta}^{k}R_{\varepsilon}(\zeta)\,\partial_{\zeta}^{j-1-k}(R_{\varepsilon}(\zeta)2\zeta)\sqrt{V}.\end{split} (4.6)

On the other hand, by (4.5), we have

12πitrλ(ζ2Pε))12ζdζ=12πitr𝒞λj=1JRε(ζ)2ζVAjVRε(ζ)(ζλ)jdζ=j=1Jk=0j11k!(j1k)!trζj1k(Rε(ζ)2ζ)VAjVζkRε(ζ).\begin{split}{}&\quad\ \frac{1}{2\pi i}\operatorname{tr}\oint_{\lambda}(\zeta^{2}-P_{\varepsilon}))^{-1}2\zeta d\zeta\\ &=\frac{1}{2\pi i}\operatorname{tr}\int_{\mathcal{C}_{\lambda}}\sum_{j=1}^{J}\frac{R_{\varepsilon}(\zeta)2\zeta\sqrt{V}A_{j}\sqrt{V}R_{\varepsilon}(\zeta)}{(\zeta-\lambda)^{j}}d\zeta\\ &=\sum_{j=1}^{J}\sum_{k=0}^{j-1}\frac{1}{k!(j-1-k)!}\operatorname{tr}\partial_{\zeta}^{j-1-k}(R_{\varepsilon}(\zeta)2\zeta)\sqrt{V}A_{j}\sqrt{V}\,\partial_{\zeta}^{k}R_{\varepsilon}(\zeta).\end{split} (4.7)

Now we compare (4.6) and (4.7), since each AjA_{j} has finite rank, we can apply cyclicity of the trace to obtain the multiplicity formula (4.4). ∎

5 Proof of convergence

The proof of convergence is based on Lemma 1, Lemma 3, with an application of the Gohberg–Sigal–Rouché theorem, see Gohberg–Sigal gohberg1971operator and (res, , Appendix C.).

We now state a more precise version of Theorem 1 involving the multiplicities given in (2.3) and (4.3) as follows:

Theorem 2.

For any Ω\Omega given in (1.4), there exists δ0=δ0(Ω)\delta_{0}=\delta_{0}(\Omega) satisfying the following: for any 0<δ<δ00<\delta<\delta_{0}, there exists εδ>0\varepsilon_{\delta}>0 such that for any λΩ\lambda\in\Omega with m(λ)>0m(\lambda)>0,

#{λj(ε)}j=1B(λ,δ)=m(λ),for all 0<ε<εδ,\#\,\{\lambda_{j}(\varepsilon)\}_{j=1}^{\infty}\cap B(\lambda,\delta)=m(\lambda),\quad\textrm{for all }0<\varepsilon<\varepsilon_{\delta},

where {λj(ε)}j=1\{\lambda_{j}(\varepsilon)\}_{j=1}^{\infty} given in (1.3) is counted with multiplicity, B(λ,δ):={z:|zλ|<δ}B(\lambda,\delta):=\{z\in{\mathbb{C}}:|z-\lambda|<\delta\}.

Proof.

In view of Lemma 1, the poles of (I+VR0(λ)V)1(I+\sqrt{V}R_{0}(\lambda)\sqrt{V})^{-1} are isolated in the region {λ:Reλ>0,Imλ>γ}\{\lambda\in{\mathbb{C}}:\operatorname{Re}\lambda>0,\operatorname{Im}\lambda>-\gamma\}, thus there are finitely many λΩ\lambda\in\Omega with m(λ)>0m(\lambda)>0, denoted by λ1,,λJ\lambda_{1},\ldots,\lambda_{J}. We choose δ0>0\delta_{0}>0 such that B(λj,δ0)B(\lambda_{j},\delta_{0}), j=1,,Jj=1,\ldots,J are disjoint discs in Ω\Omega, then for any fixed 0<δ<δ00<\delta<\delta_{0} and each λΩ\lambda\in\Omega with m(λ)>0m(\lambda)>0, we have

(I+VR0(ζ)V)1L2L2<C(δ),ζB(λ,δ),\|(I+\sqrt{V}R_{0}(\zeta)\sqrt{V})^{-1}\|_{L^{2}\to L^{2}}<C(\delta),\quad\forall\,\zeta\in\partial B(\lambda,\delta),

for some constant C(δ)>0C(\delta)>0.

In order to apply the Gohberg–Sigal–Rouché theorem, we need to estimate :

I+VRε(ζ)V(I+VR0(ζ)V)L2L2,for any ζΩ.\|I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}-(I+\sqrt{V}R_{0}(\zeta)\sqrt{V})\|_{L^{2}\to L^{2}},\quad\textrm{for any }\zeta\in\Omega.

1. Choose χ𝒞c(n)\chi\in{{\mathcal{C}}^{\infty}_{\rm{c}}}({\mathbb{R}}^{n}) satisfying χ1\chi\equiv 1 in Bn(0,1)B_{{\mathbb{R}}^{n}}(0,1) and suppχBn(0,2)\operatorname{supp}\chi\subset B_{{\mathbb{R}}^{n}}(0,2), here Bn(0,r):={xn:|x|<r}B_{{\mathbb{R}}^{n}}(0,r):=\{x\in{\mathbb{R}}^{n}:|x|<r\}, we define χR(x)=χ(R1x)\chi_{R}(x)=\chi(R^{-1}x) and calculate:

I+VRε(ζ)V(I+VR0(ζ)V)=VRε(ζ)VχRVRε(ζ)χRV+VχR(Rε(ζ)R0(ζ))χRV(VR0(ζ)VχRVR0(ζ)χRV).\begin{split}{}&\quad\ I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}-(I+\sqrt{V}R_{0}(\zeta)\sqrt{V})\\ &=\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}-\chi_{R}\sqrt{V}R_{\varepsilon}(\zeta)\chi_{R}\sqrt{V}+\sqrt{V}\chi_{R}(R_{\varepsilon}(\zeta)-R_{0}(\zeta))\chi_{R}\sqrt{V}\\ &\quad-(\sqrt{V}R_{0}(\zeta)\sqrt{V}-\chi_{R}\sqrt{V}R_{0}(\zeta)\chi_{R}\sqrt{V}).\end{split} (5.1)

2. The first term can be written as (1χR)VRε(ζ)V+χRVRε(ζ)(1χR)V(1-\chi_{R})\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}+\chi_{R}\sqrt{V}R_{\varepsilon}(\zeta)(1-\chi_{R})\sqrt{V}. Let γ~=(γ+γ)/2\tilde{\gamma}=(\gamma+\gamma^{\prime})/2, then

(1χR)VRε(ζ)V=(1χR)Veγ~|x|(eγ~|x|Rε(ζ))eγ~|x|)Veγ~|x|,(1-\chi_{R})\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}=(1-\chi_{R})\sqrt{V}e^{\tilde{\gamma}|x|}(e^{-\tilde{\gamma}|x|}R_{\varepsilon}(\zeta))e^{-\tilde{\gamma}|x|})\sqrt{V}e^{\tilde{\gamma}|x|},

where |V(x)eγ~|x||Ce(γ~γ)|x|=Ce(γγ)|x|/2|\sqrt{V(x)}e^{\tilde{\gamma}|x|}|\leq Ce^{(\tilde{\gamma}-\gamma)|x|}=Ce^{-(\gamma-\gamma^{\prime})|x|/2}. By Lemma 2, there exists ε0=ε0(Ω,γ~)\varepsilon_{0}=\varepsilon_{0}(\Omega,\tilde{\gamma}) such that for any 0<ε<ε00<\varepsilon<\varepsilon_{0}, eγ~|x|Rε(ζ))eγ~|x|L2L2C(Ω,γ~)\|e^{-\tilde{\gamma}|x|}R_{\varepsilon}(\zeta))e^{-\tilde{\gamma}|x|}\|_{L^{2}\to L^{2}}\leq C(\Omega,\tilde{\gamma}). Thus,

(1χR)VRε(ζ)VL2L2C(Ω,γ)e(γγ)R/2,for any 0<ε<ε0.\|(1-\chi_{R})\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}\|_{L^{2}\to L^{2}}\leq C(\Omega,\gamma)e^{-(\gamma-\gamma^{\prime})R/2},\quad\textrm{for any }0<\varepsilon<\varepsilon_{0}.

Similarly, we can bound χRVRε(ζ)(1χR)VL2L2\|\chi_{R}\sqrt{V}R_{\varepsilon}(\zeta)(1-\chi_{R})\sqrt{V}\|_{L^{2}\to L^{2}} by the right side above. Hence for any 0<ε<ε00<\varepsilon<\varepsilon_{0},

VRε(ζ)VχRVRε(ζ)χRVL2L2Ce(γγ)R/2,ζΩ.\|\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}-\chi_{R}\sqrt{V}R_{\varepsilon}(\zeta)\chi_{R}\sqrt{V}\|_{L^{2}\to L^{2}}\leq Ce^{-(\gamma-\gamma^{\prime})R/2},\quad\forall\,\zeta\in\Omega. (5.2)

3. We can estimate the third term in (5.1) by a similar argument. (2.2) implies that

eγ~|x|R0(ζ)eγ~|x|L2L2C(Ω,γ),ζΩ.\|e^{-\tilde{\gamma}|x|}R_{0}(\zeta)e^{-\tilde{\gamma}|x|}\|_{L^{2}\to L^{2}}\leq C(\Omega,\gamma),\quad\forall\,\zeta\in\Omega.

Hence, arguing as above, we obtain that

VR0(ζ)VχRVR0(ζ)χRVL2L2Ce(γγ)R/2,ζΩ.\|\sqrt{V}R_{0}(\zeta)\sqrt{V}-\chi_{R}\sqrt{V}R_{0}(\zeta)\chi_{R}\sqrt{V}\|_{L^{2}\to L^{2}}\leq Ce^{-(\gamma-\gamma^{\prime})R/2},\quad\forall\,\zeta\in\Omega. (5.3)

4. We note that

χR(Rε(ζ)R0(ζ))χR=iεχR(Δiεx2ζ2)1x2(Δζ2)1χR,\chi_{R}(R_{\varepsilon}(\zeta)-R_{0}(\zeta))\chi_{R}=i\varepsilon\,\chi_{R}(-\Delta-i\varepsilon x^{2}-\zeta^{2})^{-1}x^{2}(\Delta-\zeta^{2})^{-1}\chi_{R},

and recall Zw-vis that there exists C=C(Ω,χR)C=C(\Omega,\chi_{R}) (independent of ε\varepsilon) such that

χR(Δiεx2ζ2)1x2(Δζ2)1χRL2L2C,ζΩ,ε>0,\|\chi_{R}(-\Delta-i\varepsilon x^{2}-\zeta^{2})^{-1}x^{2}(\Delta-\zeta^{2})^{-1}\chi_{R}\|_{L^{2}\to L^{2}}\leq C,\quad\forall\,\zeta\in\Omega,\,\varepsilon>0,

which is proved using the method of complex scaling, see (Zw-vis, , §5) for details. Hence

VχR(Rε(ζ)R0(ζ))χRVL2L2C(Ω,χR)ε,ζΩ,ε>0.\|\sqrt{V}\chi_{R}(R_{\varepsilon}(\zeta)-R_{0}(\zeta))\chi_{R}\sqrt{V}\|_{L^{2}\to L^{2}}\leq C(\Omega,\chi_{R})\,\varepsilon,\quad\forall\,\zeta\in\Omega,\,\varepsilon>0. (5.4)

By (5.2) and (5.3), we can first fix RR sufficiently large such that

VRε(ζ)VχRVRε(ζ)χRVL2L21/(3C(δ)),ζΩ, 0ε<ε0.\|\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}-\chi_{R}\sqrt{V}R_{\varepsilon}(\zeta)\chi_{R}\sqrt{V}\|_{L^{2}\to L^{2}}\leq 1/(3C(\delta)),\quad\forall\,\zeta\in\Omega,\ 0\leq\varepsilon<\varepsilon_{0}.

Then by (5.4), there exists εδ>0\varepsilon_{\delta}>0 such that for all 0<ε<εδ0<\varepsilon<\varepsilon_{\delta},

VχR(Rε(ζ)R0(ζ))χRVL2L21/(3C(δ)),ζΩ.\|\sqrt{V}\chi_{R}(R_{\varepsilon}(\zeta)-R_{0}(\zeta))\chi_{R}\sqrt{V}\|_{L^{2}\to L^{2}}\leq 1/(3C(\delta)),\quad\forall\,\zeta\in\Omega.

We may assume that εδ<ε0\varepsilon_{\delta}<\varepsilon_{0}, thus by (5.1), we conclude that for each 0<ε<εδ0<\varepsilon<\varepsilon_{\delta},

(I+VR0(ζ)V)1(I+VRε(ζ)V(I+VR0(ζ)V))L2L2<1,\|(I+\sqrt{V}R_{0}(\zeta)\sqrt{V})^{-1}(\,I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V}-(I+\sqrt{V}R_{0}(\zeta)\sqrt{V})\,)\|_{L^{2}\to L^{2}}<1,

on B(λ,δ)\partial B(\lambda,\delta).

Now we apply the Gohberg–Sigal–Rouché theorem to obtain that

m(λ)=12πitrB(λ,δ)(I+VR0(ζ)V)1ζ(VR0(ζ)V)dζ=12πitrB(λ,δ)(I+VRε(ζ)V)1ζ(VRε(ζ)V)dζ,\begin{split}m(\lambda)&=\frac{1}{2\pi i}\operatorname{tr}\int_{\partial B(\lambda,\delta)}(I+\sqrt{V}R_{0}(\zeta)\sqrt{V})^{-1}\partial_{\zeta}(\sqrt{V}R_{0}(\zeta)\sqrt{V})\,d\zeta\\ &=\frac{1}{2\pi i}\operatorname{tr}\int_{\partial B(\lambda,\delta)}(I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})^{-1}\partial_{\zeta}(\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})\,d\zeta,\end{split}

for each 0<ε<εδ0<\varepsilon<\varepsilon_{\delta}. Let λ1(ε),,λK(ε)\lambda_{1}(\varepsilon),\ldots,\lambda_{K}(\varepsilon) be the distinct poles of (I+VRε(ζ)V)1(I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})^{-1} in B(λ,δ)B(\lambda,\delta), then

m(λ)=k=1K12πitrλk(ε)(I+VRε(ζ)V)1ζ(VRε(ζ)V)dζ=k=1Kmε(λk(ε)),m(\lambda)=\sum_{k=1}^{K}\frac{1}{2\pi i}\operatorname{tr}\oint_{\lambda_{k}(\varepsilon)}(I+\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})^{-1}\partial_{\zeta}(\sqrt{V}R_{\varepsilon}(\zeta)\sqrt{V})\,d\zeta=\sum_{k=1}^{K}m_{\varepsilon}(\lambda_{k}(\varepsilon)),

Therefore, with Lemma 3 and (4.4), we obtain that

#{λj(ε)}j=1B(λ,δ)=m(λ), 0<ε<εδ,\#\,\{\lambda_{j}(\varepsilon)\}_{j=1}^{\infty}\cap B(\lambda,\delta)=m(\lambda),\quad\forall\,0<\varepsilon<\varepsilon_{\delta},

which completes the proof. ∎

Acknowledgements.
The author would like to thank Maciej Zworski for helpful discussions. I am also grateful to the anonymous referee for the careful reading of the first version and for the valuable comments. This project was supported in part by the National Science Foundation grant 1500852.

DATA AVAILABILITY STATEMENT
The data that supports the findings of this study are available within the article.

References

  • (1) UV Riss and H-D Meyer. Reflection-free complex absorbing potentials. Journal of Physics B: Atomic, Molecular and Optical Physics, 28:1475, (1995).
  • (2) Tamar Seideman and William H Miller. Calculation of the cumulative reaction probability via a discrete variable representation with absorbing boundary conditions. The Journal of chemical physics, 96:4412, (1992).
  • (3) Thomas-C Jagau, Dmitry Zuev, Ksenia B Bravaya, Evgeny Epifanovsky, and Anna I Krylov. A fresh look at resonances and complex absorbing potentials: Density matrix-based approach. The journal of physical chemistry letters, 5:310, (2014).
  • (4) Maciej Zworski. Scattering resonances as viscosity limits. In Algebraic and Analytic Microlocal Analysis, page 635. Springer, (2013).
  • (5) Haoren Xiong. Resonances as viscosity limits for exterior dilation analytic potentials. arXiv preprint arXiv:2002.12490, (2020).
  • (6) Walter Hunziker. Distortion analyticity and molecular resonance curves. In Annales de l’IHP Physique théorique, volume 45, page 339, (1986).
  • (7) Johannes Sjöstrand and Maciej Zworski. Complex scaling and the distribution of scattering poles. Journal of the American Mathematical Society, 4:729, (1991).
  • (8) Shu Nakamura. Distortion analyticity for two-body schrödinger operators. In Annales de l’IHP Physique théorique, volume 53, page 149, (1990).
  • (9) Kentaro Kameoka and Shu Nakamura. Resonances and viscosity limit for the wigner-von neumann type hamiltonian. arXiv preprint arXiv:2003.07001, (2020).
  • (10) Richard Froese. Upper bounds for the resonance counting function of schrödinger operators in odd dimensions. Canadian Journal of Mathematics, 50:538, (1998).
  • (11) Oran Gannot. From quasimodes to resonances: exponentially decaying perturbations. Pacific Journal of Mathematics, 277:77, (2015).
  • (12) Semyon Dyatlov and Maciej Zworski. Mathematical theory of scattering resonances, volume 200. American Mathematical Soc., (2019).
  • (13) E Brian Davies. Pseudo–spectra, the harmonic oscillator and complex resonances. Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences, 455:585, (1999).
  • (14) Michael Hitrik, Johannes Sjöstrand, and Joe Viola. Resolvent estimates for elliptic quadratic differential operators. Analysis & PDE, 6:181, (2013).
  • (15) Maciej Zworski. Semiclassical analysis, volume 138. American Mathematical Soc., (2012).
  • (16) IC U. Gohberg and E. I. Sigal. An operator generalization of the logarithmic residue theorem and the theorem of rouché. Mathematics of the USSR-Sbornik, 13:603, (1971).