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

Pseudomodes for biharmonic operators with complex potentials

Tho Nguyen Duc Department of Mathematics, Faculty of Nuclear Sciences and Physical Engineering, Czech Technical University in Prague, Trojanova 13, 12000 Prague 2, Czech Republic nguyed16@fjfi.cvut.cz
Abstract.

This article is devoted to the construction of pseudomodes of one-dimensional biharmonic operators with the complex-valued potentials via the WKB method. As a by-product, the shape of pseudospectrum near infinity can be described. This is a newly discovered systematic method that goes beyond the standard semi-classical setting which is a direct consequence. This approach can cover a wide class of previously inaccessible potentials, from logarithmic to superexponential ones.

1. Introduction

1.1. Context and motivations

In the well-known self-adjoint theory, the norm of the resolvent associated with a self-adjoint operator is very large if and only if the spectral parameter comes close to the spectrum. The picture is very different for the non-self-adjoint operator: the resolvent may blow up when the spectral parameter is far from the spectrum. It leads to the instability of the spectrum under a small perturbation and reveals that, in this case, the numerical methods will fail to compute the eigenvalues. In order to describe this pathological property of the non-self-adjoint operators, the notion of pseudospectra was regarded [27, 8, 12]. That is, given a positive number ε\varepsilon, the ε\varepsilon-pseudospectrum σε(H)\sigma_{\varepsilon}(H) of an operator HH in a complex Hilbert space is defined as its spectrum σ(H)\sigma(H) along with those resolvent points whose norm of the resolvent larger than ε1\varepsilon^{-1}. This definition is equivalent to the description of the set σε(H)\sigma_{\varepsilon}(H) as the spectrum σ(H)\sigma(H) enlarged by the complex points λ\lambda (called pseudoeigenvalues) for which there exists a vector Ψ𝒟(H)\Psi\in\mathcal{D}(H) such that

(Hλ)Ψ<εΨ.\|(H-\lambda)\Psi\|<\varepsilon\,\|\Psi\|\,. (1.1)

Any Ψ\Psi satisfies (1.1) is called a pseudomode (or its other names: pseudoeigenfunction, pseudoeigenvector, quasimode).

The analysis of pseudospectra of the non-self-adjoint Schrödinger operators has given rise to many investigations in twenty years [7, 16, 13, 14, 25, 22, 15, 21, 6]. By change of scale to transform the original Schrödinger operator to its semi-classical form, which has the small parameter h2h^{2} in front of the second derivative, Davies’ pioneering work [7] has constructed the pseudomode for the semi-classical Schrödinger operator as h0h\to 0, then the pseudomode of the original operator with large λ\lambda is achieved (i.e. those corresponding in (1.1) to |λ|+|\lambda|\to+\infty with ελ0\varepsilon_{\lambda}\to 0). However, this method seems merely effective for a certain class of polynomial potentials in which the scaling is able to be performed, while it is inapplicable, for example, for logarithmic or exponential potentials. Moreover, the semi-classical approach is even more inaccessible to the class of discontinuous potentials since the semi-classical construction of pseudomode requires that the potential is smooth (or at least continuous). In [15], Raphaël and Krejčiřík had studied the imaginary sign potential by constructing the resolvent kernel of the Schrödinger operator. Up to twenty years after the earlier Davies’ work [7], Krejčiřík and Siegl in [21] developed a direct construction of large-energy pseudomodes for Schrödinger operator, which does not require the passage through semi-classical setting and can cover all mentioned above potentials. Recently, this technique is applicable to other models such as the damped wave equation [1] and Dirac operator [20]. The purpose of the present paper is to extend the method developed in [21, 20] to higher differential operator by considering biharmomic instead of Schrödinger operators and to discover the more universal shape of the potentials such that this method works.

More precisely, this document is devoted to construct a λ\lambda-dependent family of pseudomodes fλf_{\lambda} such that

(Vλ)fλ=o(1)fλ, as λ in Ω.\|\left(\mathscr{L}_{V}-\lambda\right)f_{\lambda}\|=o(1)\|f_{\lambda}\|,\qquad\text{ as }\lambda\to\infty\text{ in }\Omega\subset\mathbb{C}. (1.2)

Here V\mathscr{L}_{V} is the one-dimensional biharmonic operator which is defined as a forth derivative perturbed by a complex-valued potential VV:

V=d4dx4+V(x).\mathscr{L}_{V}=\frac{\mathrm{d}^{4}}{\mathrm{d}x^{4}}+V(x).

In [4], when the author work with the non-self-adjoint harmonic oscillator, it has shown us that we do not always obtain the decay in (1.2) by just letting λ\lambda\to\infty in \mathbb{C}. Indeed, he proves in [4] that the norm of the resolvent is bounded above when λ\lambda\to\infty along some half-lines parallel to the positive semi-axis +\mathbb{R}^{+} and blows up when λ\lambda\to\infty in a region bounded by two certain curves. The set Ω\Omega is a region in the neighborhood of infinity in \mathbb{C} which contains large λ\lambda allowing the decay in (1.2) to happen and thus the norm of the resolvent will go up in this region.

There are three main theorems in this paper. The first theorem will provide the answer to the question: What is the behaviour of VV such that we obtain the decay in (1.2) when λ\lambda\to\infty in the region parallel to the positive semi-axis? We will see that the behaviour at infinity of the imaginary part of VV, denoted by ImV\mathrm{Im}\,V, plays the decisive role in this mission, that is

(limxImV(x))(limx+ImV(x))<0.\left(\lim_{x\to-\infty}\mathrm{Im}\,V(x)\right)\left(\lim_{x\to+\infty}\mathrm{Im}\,V(x)\right)<0.

This condition is essential to ensure the “significantly non-normality” of HVH_{V}. Theorem 2.1 also generalizes the results in [21, Thm. 3.7] and [20, Thm. 3.10, Thm. 3.11] in which only λ+\lambda\to+\infty on the positive semi-axis is considered. Furthermore, the class of the potential VV is also widened in our paper by controlling the derivative of VV by general functions τ±\tau_{\pm} (see Cond. (2.2)) instead of some polynomial functions in [21, Cond. (3.2)]. This allows us to cover the super-exponential function (see Example 4) that could not be covered in [21].

Theorem 2.2 addresses the question: What is the shape of Ω\Omega for each type of the potential VV? Once again, our assumption can cover a larger class of potentials than [21, Asm. III] and [20, Asm. II]. By applying it for the functions from growing slowly at ++\infty such as logarithmic one V(x)=iln(x)V(x)=i\ln(x) or root-type V(x)=ixγV(x)=ix^{\gamma} with γ(0,1)\gamma\in(0,1) to growing faster at ++\infty such as polynomial V(x)=ixγV(x)=ix^{\gamma} with γ1\gamma\geq 1 and super-exponential functions V(x)=ieexV(x)=ie^{e^{x}}, the region Ω\Omega for each type of them is described differently (see Subsection 2.2.2 in which many picture for illustrations are added). Finally, Theorem 2.3 shows us that the semi-classical setting is actually a special consequence of our pseudomode construction. In all of the theorems mentioned above, the regularity of the potentials are assumed as mild as possible, which is a small plus point compared with the previous semi-classical setting that the smoothness of VV always assumed highly.

The technique that we employed to find the pseudomodes is the (J)WKB method (also known as the Liouville–Green approximation). The WKB method is not only seen as a tool to approximate the eigenfunction for some differential operators, but it also reveals some information of the eigenvalues. For instance, in [3, 11], the asymptotic expansions for the eigenvalues of the self-adjoint magnetic Laplacian were found in the process of doing the WKB analysis. Now we could ask a similar question for the non-self-adjoint operator

“Can we describe Ω\Omega (pseudo-spectrum near the infinity) by the WKB method?”

Continuing the work of [21, 20], we would like to provide a positive answer by considering the higher differential operator with more general potentials. We refer [20, Remark 2.3] to explain why our approach goes beyond the standard semi-classical settings. It is our belief that the study in this article is a necessary step allowing us to approach in the future with more general differential operators such as dndxn+V(x)\frac{\mathrm{d}^{n}}{\mathrm{d}x^{n}}+V(x) for arbitrary n1n\in\mathbb{N}_{1} and a complex electric potential VV. The extension of this analysis to the magnetic Laplacian with a complex magnetic field (recently motivated in [19]) also constitutes a challenging open problem.

Biharmonic operator has its own application in continuum mechanics and in linear elasticity theory. In recent years, the biharmonic operator attracted considerable attention in particular in the context of spectral theory. For instance, in [9], Enblom study the bound of the eigenvalues for the non-self-adjoint polyharmonic operator in which the biharmonic operator is a special case. The sharp confinement of the eigenvalues in a closed disk of the biharmonic operators has been produced in [18] for low dimension. We also list here some recent studies related to the spectral properties of the biharmonic operators [2, 17, 5, 10].

1.2. Handy notations and conventions

Here we summarise some special notations and conventions which we use regularly in the paper:

  1. 1)

    k\mathbb{N}_{k}, with a non-negative integer kk, is the set of integers starting from kk;

  2. 2)

    For the semi real axes, we denote 0(,0]\mathbb{R}_{0}^{-}\coloneqq(-\infty,0] and 0+[0,)\mathbb{R}_{0}^{+}\coloneqq[0,\infty), and for strictly positive or negative axes, we denote (,0)\mathbb{R}^{-}\coloneqq(-\infty,0) and +(0,+)\mathbb{R}^{+}\coloneqq(0,+\infty);

  3. 3)

    For the list of integer numbers from mm to nn, where m,nm,n\in\mathbb{Z} and m<nm<n, we denote [[m,n]][[m,n]], i.e. [[m,n]]:={k:mkn}[[m,n]]:=\left\{k\in\mathbb{Z}:m\leq k\leq n\right\} ;

  4. 4)

    fnf^{n} and f(n)f^{(n)} denotes respectively the power nn and the nn-th derivative of a function f:f:\mathbb{R}\to\mathbb{C} with n0n\in\mathbb{N}_{0};

  5. 5)

    We use the same symbol :=L2()\|\cdot\|:=\left\|\cdot\right\|_{L^{2}(\mathbb{R})} for L2L^{2}-norms of complex-valued functions defined on \mathbb{R};

  6. 6)

    We often write λ=α+iβ\lambda=\alpha+i\beta where α,β\alpha,\beta\in\mathbb{R} and denote VλλVV_{\lambda}\coloneqq\lambda-V;

  7. 7)

    For two real-valued functions aa and bb, we write aba\lesssim b (respectively, aba\gtrsim b) if there exists a constant C>0C>0, independent of λ\lambda and xx (or any other relevant parameter such as α\alpha and β\beta), such that aCba\leq Cb (respectively, aCba\geq Cb); and we write aba\approx b if aba\lesssim b and aba\gtrsim b.

1.3. Structure of the paper

The rest of this article is organized as follows: In Section 2, we establish the main conditions for the admissible class of the potentials VV and the main theorems related to the central problem (1.2) come shortly after. Many nontrivial and illuminating examples are contained in this section. We reserve Section 3 to describe the WKB method for the biharmonic operators, which is the main tool to construct the pseudomodes. Section 4 is devoted to the large real pseudoeigenvalues in the region parallel to the positive semi-axis +\mathbb{R}^{+} while the pseudoeigenvalues corresponding to the large imaginary part are dealt with in Section 5. The method in Section 5 is also used to prove the semi-classical result.

2. Statements, results and applications

2.1. Statements and results

In this article, we consider the maximal forth-order differential operator perturbed by a multiplication operator, where we denote by the same symbol VV, as follows

Dom(V)={uL2():(d4dx4+V(x))uL2()},\displaystyle\mathrm{Dom}(\mathscr{L}_{V})=\left\{u\in L^{2}(\mathbb{R}):\left(\frac{\mathrm{d}^{4}}{\mathrm{d}x^{4}}+V(x)\right)u\in L^{2}(\mathbb{R})\right\},
V=d4dx4+V(x),\displaystyle\mathscr{L}_{V}=\frac{\mathrm{d}^{4}}{\mathrm{d}x^{4}}+V(x),

where V:V:\mathbb{R}\to\mathbb{C} is assumed to be locally L2L^{2}-integrable, i.e. VLloc2()V\in L^{2}_{\mathrm{loc}}(\mathbb{R}). This condition ensures that all the action of V\mathscr{L}_{V} is well-defined in the sense of distributions. It follows that V\mathscr{L}_{V} is a closed operator. However, the closedness of V\mathscr{L}_{V} is inessential for our construction of pseudomode. Since the pseudomode constructed in this document has a compact support, our method can be modified to adapt with any closed extension of the operator initially defined on 𝒞0()\mathcal{C}_{0}^{\infty}(\mathbb{R}).

Since λ=α+iβ\lambda=\alpha+i\beta, there are two obvious ways to make λ\lambda become largely, that is increasing α\alpha or increasing β\beta. Therefore, depending on the decisive role of the real part or imaginary part of the pseudoeigenvalue in the decaying estimation of (1.2), the pseudomodes are accordingly constructed in different ways.

2.1.1. Large real pseudoeigenvalues

Let us state the main assumptions on the admissible class of the potential VV.

Assumption I.

Let N0N\in\mathbb{N}_{0}, assume that VWlocN+3,()V\in W^{N+3,\infty}_{\mathrm{loc}}(\mathbb{R}) satisfy the following conditions:

  1. 1)

    ImV\mathrm{Im}\,V has a different asymptotic behaviour at ±\pm\infty:

    lim supxImV(x)<0<lim infx+ImV(x);\limsup_{x\to-\infty}\mathrm{Im}\,V(x)<0<\liminf_{x\to+\infty}\mathrm{Im}\,V(x); (2.1)
  2. 2)

    There exist continuous functions τ±:0±+\tau_{\pm}:\mathbb{R}_{0}^{\pm}\to\mathbb{R}^{+} such that, for all n[[1,N+3]]n\in[[1,N+3]],

    |V(n)(x)|=𝒪(τ±(x)n|V(x)|),x±;\displaystyle\left|V^{(n)}(x)\right|=\mathcal{O}\left(\tau_{\pm}(x)^{n}\left|V(x)\right|\right),\qquad x\to\pm\infty; (2.2)
  3. 3)

    Additional assumptions for τ±\tau_{\pm} and VV in each of the following cases:

    1. a)

      If VV is unbounded at ±\pm\infty, assume that τ±(1)\tau_{\pm}^{(1)} exist for |x|1|x|\gtrsim 1 and there exist ν±1\nu_{\pm}\geq-1 for which

      |x|ν±=𝒪(τ±(x)),τ±(1)(x)=𝒪(|x|ν±τ±(x)),x±|x|^{\nu_{\pm}}=\mathcal{O}\left(\tau_{\pm}(x)\right),\qquad\tau_{\pm}^{(1)}(x)=\mathcal{O}\left(|x|^{\nu_{\pm}}\tau_{\pm}(x)\right),\qquad x\to\pm\infty (2.3)

      such that

      |ImV(1)(x)|=𝒪(τ±(x)|ImV(x)|),x±,|\mathrm{Im}\,V^{(1)}(x)|=\mathcal{O}\left(\tau_{\pm}(x)\left|\mathrm{Im}\,V(x)\right|\right),\qquad x\to\pm\infty, (2.4)

      and assume further that: ε1>0\exists\varepsilon_{1}>0,

      τ±4(x)(τ±12+4ε1(x)+|ReV(x)|3+ε1)=𝒪(|ImV(x)|4),x±;\tau_{\pm}^{4}(x)\left(\tau_{\pm}^{12+4\varepsilon_{1}}(x)+\left|\mathrm{Re}\,V(x)\right|^{3+\varepsilon_{1}}\right)=\mathcal{O}\left(\left|\mathrm{Im}\,V(x)\right|^{4}\right),\qquad x\to\pm\infty; (2.5)
    2. b)

      If VV is bounded at ±\pm\infty, then assume that: ε2(0,13)\exists\varepsilon_{2}\in\left(0,\frac{1}{3}\right),

      τ±(x)=𝒪(|x|13ε2),x±.\tau_{\pm}(x)=\mathcal{O}\left(|x|^{\frac{1}{3}-\varepsilon_{2}}\right),\qquad x\to\pm\infty. (2.6)

Notice that the constant ε1\varepsilon_{1} in (2.5) shall be considered sufficiently small. Since optimization this constant is not interesting, we are not trying to do that in our work.

Next lines are some comments on Assumption I. By comparing with the same assumption for the potential in Schrödinger operator [21, Asm. I], the regularity of VV in the biharmonic operator requires to be higher, this can be seen clearly in the formula of the remainders λ,0\mathcal{R}_{\lambda,0} in (3.12) and in [21, Eqn. 2.18]. These remainders are the amount left over after performing WKB construction (see Section 3). As mentioned in the introduction, the assumption (2.1) makes the operator highly non-self-adjoint. Indeed, if (V)\left(\mathscr{L}_{V}\right)^{*} is the formal adjoint of HVH_{V}, i.e. (V)=V¯\left(\mathscr{L}_{V}\right)^{*}=\mathscr{L}_{\overline{V}}, it is straightforward to verify (at least algebraically) that the normality relation V(V)=(V)V\mathscr{L}_{V}\left(\mathscr{L}_{V}\right)^{*}=\left(\mathscr{L}_{V}\right)^{*}\mathscr{L}_{V} holds iff ImV(1)=0\mathrm{Im}\,V^{(1)}=0, i.e. ImV\mathrm{Im}\,V is a constant. The sign of ImV\mathrm{Im}\,V in (2.1) will determine the sign for the decay of the pseudomode. The larger ImV\mathrm{Im}\,V is, the faster the pseudomode decreases at infinity, see the estimate (4.17). Furthermore, we can invert the sign of ImV\mathrm{Im}\,V at infinity and the construction of pseudomode does not change more, see Remark 4.6. The condition (2.2) is designed exclusively for the very special shapes of transport solutions and the remainder that will be described in the next section. To control the too large ReV\mathrm{Re}\,V and any wild behaviour of the derivatives of VV, the conditions (2.5) and (2.6) are employed, furthermore, the natural imposition of the assumption (2.5) on ReV\mathrm{Re}\,V is also discussed in Remark 4.4, which shows that this condition is optimal in the polynomial case. Finally, two assumptions (2.3) and (2.4) are technical tools to guarantee that the values of τ±\tau_{\pm} and ImV\mathrm{Im}\,V on some suitable interval can be comparable up to a constant (see (4.12)). This technique has been lately used very much, for instance, in [21, 1, 20, 24], however, in these papers, they fix the functions τ±(x)|x|ν\tau_{\pm}(x)\coloneqq|x|^{\nu} for some ν1\nu\geq-1. Here, we provide an improvement of this technique by controlling by general functions τ±\tau_{\pm} satisfying (2.3). For example, ImV(x)=eex\mathrm{Im}\,V(x)=e^{e^{x}} can be covered by our assumption with τ(x)=ex\tau(x)=e^{x}, but it is not allowed by the assumption in the papers mentioned above.

In order to state our first theorem, we denote the interval B:=[β,β+]B:=[-\beta_{-},\beta_{+}] where β±\beta_{\pm} are non-negative constants satisfying

lim supxImV(x)<β and β+<lim infx+ImV(x).\limsup_{x\to-\infty}\mathrm{Im}\,V(x)<-\beta_{-}\qquad\text{ and }\qquad\beta_{+}<\liminf_{x\to+\infty}\mathrm{Im}\,V(x). (2.7)
Theorem 2.1.

Let Assumption I hold for some N0N\in\mathbb{N}_{0}. Then there exists a λ\lambda-dependent family (Ψλ,N)Dom(HV)\left(\Psi_{\lambda,N}\right)\subset\textup{Dom}(H_{V}) such that for all α1\alpha\gtrsim 1 and βB\beta\in B, we have

(HVλ)Ψλ,NΨλ,NαN+14+σ(N)(α)+σ+(N)(α),\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}\lesssim\alpha^{-\frac{N+1}{4}}+\sigma_{-}^{(N)}(\alpha)+\sigma_{+}^{(N)}(\alpha), (2.8)

where

σ(N)(α)αN+14supx[δα,0]τ(x)N+1|V(x)|,σ+(N)(α)αN+14supx[0,δα+]τ+(x)N+1|V(x)|,\sigma_{-}^{(N)}(\alpha)\coloneqq\alpha^{-\frac{N+1}{4}}\displaystyle\sup_{x\in[-\delta_{\alpha}^{-},0]}\tau_{-}(x)^{N+1}|V(x)|,\qquad\sigma_{+}^{(N)}(\alpha)\coloneqq\alpha^{-\frac{N+1}{4}}\displaystyle\sup_{x\in[0,\delta_{\alpha}^{+}]}\tau_{+}(x)^{N+1}|V(x)|,

in which δα±\delta_{\alpha}^{\pm} are defined as follows:

  1. a)

    If VV is unbounded at ±\pm\infty, δα±\delta_{\alpha}^{\pm} are the smallest positive solutions of the equations

    |ImV(±x)|τ±(±x)=α3+ε14+ε1.\frac{\left|\mathrm{Im}\,V(\pm x)\right|}{\tau_{\pm}(\pm x)}=\alpha^{\frac{3+\varepsilon_{1}}{4+\varepsilon_{1}}}.
  2. b)

    If VV is bounded at ±\pm\infty, δα±=α343ε2\delta_{\alpha}^{\pm}=\alpha^{\frac{3}{4-3\varepsilon_{2}}} .

In particular, if VV and τ±\tau_{\pm} is bounded at ±\pm\infty, then

σ±(N)(α)αN+14.\sigma_{\pm}^{(N)}(\alpha)\lesssim\alpha^{-\frac{N+1}{4}}. (2.9)

Although the right hand side of (2.8) does not show us the decay obviously, Theorem 2.1 are very workable in many elementary cases such as logarithmic functions, polynomials, exponential functions and even super-exponential (see its application in Subsection 2.2.1). This theorem shows us that the regularity of the potentials has a direct influence on the decay rates of the problem (1.2): the more regular the potential is, the stronger the rate of decay in (1.2) is obtained. Furthermore, the shape of Ω\Omega corresponding to these large pseudoeigenvalues can be described generally as follows

Ω:={α+iβ:α1 and βB},\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\alpha\gtrsim 1\text{ and }\beta\in B\right\},

in which the wide of the interval BB can be any size as long as it is contained in the interval (lim supxImV(x),lim infx+ImV(x))\displaystyle\left(\limsup_{x\to-\infty}\mathrm{Im}\,V(x),\liminf_{x\to+\infty}\mathrm{Im}\,V(x)\right). Furthermore, the method can also be applied for the decaying but not integrable potential

V(x)=isgn(x)|x|γ,|x|1,γ(0,1),V(x)=i\frac{\textup{sgn}(x)}{|x|^{\gamma}},\qquad|x|\gtrsim 1,\,\gamma\in(0,1),

in which the Assumption (2.1) is broken (see Example 5 and Subsection 4.4).

2.1.2. Large imaginary pseudoeigenvalues

Concerning the pseudoeigenvalues whose imaginary part β\beta play the main role in making the right hand side of (1.2) decaying, the pseudomodes will be constructed such that their supports live completely in 0+\mathbb{R}_{0}^{+}. Therefore, it will be more convenient to consider the operators on L2(0+)L^{2}(\mathbb{R}_{0}^{+}) instead of L2()L^{2}(\mathbb{R}). Then the application for the class of operators on L2()L^{2}(\mathbb{R}) is easily obtained by the trivial extension of pseudomodes from 0+\mathbb{R}_{0}^{+} to \mathbb{R}. We assume that ImV\mathrm{Im}\,V is strictly increasing for sufficiently large xx and unbounded at ++\infty such that we can determine a unique turning point xβ>0x_{\beta}>0 of the equation

ImV(xβ)=β.\mathrm{Im}\,V(x_{\beta})=\beta.

The WKB analysis will be performed around xβx_{\beta} and the support of pseudomode will be inside some appropriate neighborhood of this point. Here are our assumptions for this construction:

Assumption II.

Let N0N\in\mathbb{N}_{0}, and let VWlocN+3,2(0+)V\in W^{N+3,2}_{\mathrm{loc}}(\mathbb{R}_{0}^{+}) satisfy all conditions of Assumption I and further the followings:

  1. 1)

    ImV\mathrm{Im}\,V goes to ++\infty as x+x\to+\infty

    limx+ImV(x)=+;\lim_{x\to+\infty}\mathrm{Im}\,V(x)=+\infty; (2.10)
  2. 2)

    There exists (t1,t2)[0,ε120(3+ε1)]2\left(t_{1},t_{2}\right)\in\left[0,\frac{\varepsilon_{1}}{20(3+\varepsilon_{1})}\right]^{2} satisfying

    t114+ε1t2<ε120(4+ε1)t_{1}-\frac{1}{4+\varepsilon_{1}}t_{2}<\frac{\varepsilon_{1}}{20(4+\varepsilon_{1})} (2.11)

    such that, for all x1x\gtrsim 1,

    ImV(x)1t1τ(x)1+t2\displaystyle\mathrm{Im}\,V(x)^{1-t_{1}}\tau(x)^{1+t_{2}} ImV(1)(x),\displaystyle\lesssim\mathrm{Im}\,V^{(1)}(x), (2.12)
    |ImV(2)(x)|\displaystyle\left|\mathrm{Im}\,V^{(2)}(x)\right| ImV(1)(x)τ(x),\displaystyle\lesssim\mathrm{Im}\,V^{(1)}(x)\tau(x), (2.13)

    where ττ+\tau\coloneqq\tau_{+}.

Although there are more conditions for the imaginary part ImV\mathrm{Im}\,V in Assumption II, the class of admissible potentials is still very large. In [21, Cond. 5.2] for the Schrödinger operator, the authors set up the condition

ImV(x)xνImV(1)(x),x1,\mathrm{Im}\,V(x)x^{\nu}\lesssim\mathrm{Im}\,V^{(1)}(x),\qquad x\gtrsim 1, (2.14)

and this is a particular case of (2.12) with τ(x)=xν\tau(x)=x^{\nu} and t1=t2=0t_{1}=t_{2}=0. However, (2.14) can not treat the potential V(x)=iln(x)V(x)=i\ln(x) (here ν=1\nu=-1), it is because

ImV(x)τ(x)=ln(x)x1 and ImV(1)(x)=x1.\mathrm{Im}\,V(x)\tau(x)=\ln(x)x^{-1}\qquad\text{ and }\qquad\mathrm{Im}\,V^{(1)}(x)=x^{-1}.

By allowing t1t_{1} and t2t_{2} to be flexible satisfying (2.11), this potential can be covered in our assumption (see Example 6). Let us define a neighborhood of xβx_{\beta} in which the pseudomode lives, that is

Jβ=(xβ2Δβ,xβ+2Δβ).J_{\beta}=\left(x_{\beta}-2\Delta_{\beta},x_{\beta}+2\Delta_{\beta}\right).

Here, Δβ\Delta_{\beta} is defined as follows: From the assumption (2.3), there exists a constant η\eta such that, for sufficiently large x>0x>0,

ητ(x)xν4,\frac{\eta}{\tau(x)}\leq\frac{x^{-\nu}}{4},

where ν:=ν+\nu:=\nu_{+} is the number defined in (2.3) and then we define

Δβητ(xβ).\Delta_{\beta}\coloneqq\frac{\eta}{\tau(x_{\beta})}. (2.15)

Since ν1\nu\geq-1, then xβ2Δβxβxβν2xβ2x_{\beta}-2\Delta_{\beta}\geq x_{\beta}-\frac{x_{\beta}^{\nu}}{2}\geq\frac{x_{\beta}}{2}, we deduce that Jβ+J_{\beta}\subset\mathbb{R}^{+} for xβ>0x_{\beta}>0. We see that if τ(x)=x1\tau(x)=x^{-1} (see Examples 6 and 7), the support of the pseudomode is able to be extended on +\mathbb{R}^{+}, i.e. |Jβ|=4Δβ=4ηxβ+|J_{\beta}|=4\Delta_{\beta}=4\eta x_{\beta}\to+\infty as β+\beta\to+\infty. This is one of the strengths of this direct construction which makes it going beyond the semi-classical construction (see [20, Remark 2.3]). Now we can state our second theorem:

Theorem 2.2.

Let Assumption II hold for some N0N\in\mathbb{N}_{0}. Assume that there exists a (β\beta-dependent) α\alpha such that the following holds as β+\beta\to+\infty, for all xJβx\in J_{\beta},

αReV(x)|α|,\displaystyle\alpha-\mathrm{Re}\,V(x)\approx|\alpha|, (2.16)
[βτ(xβ)]45|α|[βτ(xβ)1]4+ε13+ε1.\displaystyle\left[\beta\tau(x_{\beta})\right]^{\frac{4}{5}}\lesssim|\alpha|\lesssim\left[\beta\tau(x_{\beta})^{-1}\right]^{\frac{4+\varepsilon_{1}}{3+\varepsilon_{1}}}. (2.17)

Then, there exist c>0c>0, β0>0\beta_{0}>0 and a family (Ψλ,N)Dom(HV)\left(\Psi_{\lambda,N}\right)\subset\textup{Dom}(H_{V}) such that for all ββ0\beta\geq\beta_{0}, we have

(HVλ)Ψλ,NL2(+)Ψλ,NL2(+)κ(β)+σ(N)(β),\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|_{L^{2}(\mathbb{R}^{+})}}{\|\Psi_{\lambda,N}\|_{L^{2}(\mathbb{R}^{+})}}\lesssim\kappa(\beta)+\sigma^{(N)}(\beta),

in which

  • κ(β)exp(cImV(1)(xβ)τ(xβ)2|α|34+β34)\displaystyle\kappa(\beta)\coloneqq\exp\left(-c\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{|\alpha|^{\frac{3}{4}}+\beta^{\frac{3}{4}}}\right),

  • σ(0)(β):=k=02j=1k+1τ(xβ)k+1(|ReV(xβ)|j+βj)|α|k34+j\displaystyle\sigma^{(0)}(\beta):=\sum_{k=0}^{2}\sum_{j=1}^{k+1}\frac{\tau(x_{\beta})^{k+1}\left(\left|\mathrm{Re}\,V(x_{\beta})\right|^{j}+\beta^{j}\right)}{\left|\alpha\right|^{\frac{k-3}{4}+j}},

  • σ(N)(β):=k=03N1j=1k+N+1τ(xβ)k+N+1(|ReV(xβ)|j+βj)|α|k+N34+j\displaystyle\sigma^{(N)}(\beta):=\sum_{k=0}^{3N-1}\sum_{j=1}^{k+N+1}\frac{\tau(x_{\beta})^{k+N+1}\left(\left|\mathrm{Re}\,V(x_{\beta})\right|^{j}+\beta^{j}\right)}{\left|\alpha\right|^{\frac{k+N-3}{4}+j}},   N1N\geq 1.

The same result for Schrödinger operator given in [21, Cond. (5.5)] is known to be optimal when considering the case ReV=0\mathrm{Re}\,V=0, we believe that in this case our bound curves for α\alpha in (2.17) is also optimal. The study of optimality of our estimates on these bounds constitutes an interesting open problem.

2.1.3. Semi-classical setting

Let us consider the semi-classical biharmonic operator on \mathbb{R}

Hh=h4d4dx4+W(x),H_{h}=h^{4}\frac{\mathrm{d}^{4}}{\mathrm{d}x^{4}}+W(x),

where hh is the positive semi-classical parameter. By using the same construction as for Theorem 2.2, we can establish a pseudomode for this operator as h0h\to 0:

Theorem 2.3.

Let N0N\in\mathbb{N}_{0} and let WWlocN+3,()W\in W^{N+3,\infty}_{\textup{loc}}(\mathbb{R}), μ>0\mu>0, x0x_{0}\in\mathbb{R}. Assume that there exists a neighborhood II of x0x_{0} such that the function ImW(x)ImW(x0)\mathrm{Im}\,W(x)-\mathrm{Im}\,W(x_{0}) changes its sign at the point x0x_{0} on II. By fixing z=μ+W(x0)z=\mu+W(x_{0}), then there exist h0>0h_{0}>0 and a family Ψh,NDom(Hh)\Psi_{h,N}\in\textup{Dom}(H_{h}) such that for all h(0,h0)h\in(0,h_{0}),

(Hhz)Ψh,NΨh,NhN+1.\frac{\left\|\left(H_{h}-z\right)\Psi_{h,N}\right\|}{\|\Psi_{h,N}\|}\lesssim h^{N+1}.

2.2. Applications

Our goal in this section is to give some examples which are direct or indirect (Example 5) application of Theorem 2.1, Theorem 2.2 and Theorem 2.3.

2.2.1. Application of Theorem 2.1

Example 1.

Let us list some smooth potentials VV defined on \mathbb{R} such that the Assumption I holds.

  1. 1)

    VV is bounded at both -\infty and ++\infty: Consider two smooth bounded potentials on \mathbb{R}

    V1(x)=iarctan(x) and V2(x)=ixx2+1.V_{1}(x)=i\arctan(x)\text{ and }V_{2}(x)=i\frac{x}{\sqrt{x^{2}+1}}. (2.18)

    They satisfy Assumption I with τ±(x)=(x2+1)12\tau_{\pm}(x)=\left(x^{2}+1\right)^{-\frac{1}{2}} and

    limxImVj(x)=1,limx+ImVj(x)=1, for j=1,2.\lim_{x\to-\infty}\mathrm{Im}\,V_{j}(x)=-1,\qquad\lim_{x\to+\infty}\mathrm{Im}\,V_{j}(x)=1,\qquad\text{ for }j=1,2.

    Since both potentials are smooth, we can achieve the arbitrary fast decay in (2.9) by taking any large NN. More precisely, Theorem 2.1 states that: For any N0N\in\mathbb{N}_{0} and for any β±[0,1)\beta_{\pm}\in[0,1), there exists a family (Ψλ,N)(\Psi_{\lambda,N}) such that

    (HVjλ)Ψλ,NΨλ,NαN+14, for j=1,2,\frac{\|(H_{V_{j}}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}\lesssim\alpha^{-\frac{N+1}{4}},\qquad\text{ for }j=1,2,

    for all λ\lambda belonging to Ω:={α+iβ:α1 and β[β,β+](1,1)}\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\alpha\gtrsim 1\text{ and }\beta\in[-\beta_{-},\beta_{+}]\subset(-1,1)\right\} whose picture is given in Figure 1.

    Refer to caption
    Figure 1. Illustration of the shape of Ω\Omega (in cyan color) associated with the potentials V1(x)=iarctan(x)V_{1}(x)=i\arctan(x) and V2(x)=ixx2+1V_{2}(x)=i\frac{x}{\sqrt{x^{2}+1}} given in (2.18).
  2. 2)

    VV is bounded at -\infty and unbounded at ++\infty: A simple choice is

    V(x)=i(ex1).V(x)=i\left(e^{x}-1\right).

    It meets the condition (2.1) since

    limxImV(x)=1,limx+ImV(x)=+,\lim_{x\to-\infty}\mathrm{Im}\,V(x)=-1,\qquad\lim_{x\to+\infty}\mathrm{Im}\,V(x)=+\infty,

    and it satisfies the other conditions with τ±=1\tau_{\pm}=1 and ν+=0\nu_{+}=0. Depending on the behaviour of VV at ±\pm\infty, σ(N)(α)\sigma_{-}^{(N)}(\alpha) and σ+(N)(α)\sigma_{+}^{(N)}(\alpha) have different decaying:

    σ(N)(α)αN+14,σ+(N)(α)=α2N4+ε.\sigma_{-}^{(N)}(\alpha)\lesssim\alpha^{-\frac{N+1}{4}},\qquad\sigma_{+}^{(N)}(\alpha)=\alpha^{\frac{2-N}{4}+\varepsilon}.

    Here, to estimate σ+(N)\sigma_{+}^{(N)}, we just notice that δα+\delta_{\alpha}^{+} is the solution of the equation

    ImV(x)=α3+ε14+ε1,\mathrm{Im}\,V(x)=\alpha^{\frac{3+\varepsilon_{1}}{4+\varepsilon_{1}}},

    therefore, by writing 3+ε14+ε134+ε\frac{3+\varepsilon_{1}}{4+\varepsilon_{1}}\eqqcolon\frac{3}{4}+\varepsilon, we get the above estimate, since

    σ+(N)(α)=αN+14supx[0,δα+]|V(x)|=αN+14ImV(δα+)=α3+ε14+ε1N+14.\sigma_{+}^{(N)}(\alpha)=\alpha^{-\frac{N+1}{4}}\displaystyle\sup_{x\in[0,\delta_{\alpha}^{+}]}|V(x)|=\alpha^{-\frac{N+1}{4}}\mathrm{Im}\,V\left(\delta_{\alpha}^{+}\right)=\alpha^{\frac{3+\varepsilon_{1}}{4+\varepsilon_{1}}-\frac{N+1}{4}}.

    Then, for any N3N\in\mathbb{N}_{3} and for any β[0,1)\beta_{-}\in[0,1) and β+0+\beta_{+}\in\mathbb{R}_{0}^{+}, there exists a family (Ψλ,N)(\Psi_{\lambda,N}) such that

    (HVλ)Ψλ,NΨλ,Nα2N4+ε,\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}\lesssim\alpha^{\frac{2-N}{4}+\varepsilon},

    for all λΩ:={α+iβ:α1 and β[β,β+](1,+)}\displaystyle\lambda\in\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\alpha\gtrsim 1\text{ and }\beta\in[-\beta_{-},\beta_{+}]\subset(-1,+\infty)\right\}.

Example 2 (Potentials with logarithmic imaginary).

Consider VWlocN+3,()V\in W^{N+3,\infty}_{\textup{loc}}(\mathbb{R}), with N 0N\geq\leavevmode\nobreak\ 0, satisfying Assumption I with τ±(x)=(x2+1)12\tau_{\pm}(x)=\left(x^{2}+1\right)^{-\frac{1}{2}}, which satisfy (2.3) with ν±=1\nu_{\pm}=-1, that has the form

V(x)=ReV(x)+iln(x+x2+1),V(x)=\mathrm{Re}\,V(x)+i\ln\left(x+\sqrt{x^{2}+1}\right),

where

|ReV(x)||x|ρln(|x|)4,|x|1, with some ρ<43.\left|\mathrm{Re}\,V(x)\right|\lesssim|x|^{\rho}\ln\left(|x|\right)^{4},\qquad|x|\gtrsim 1,\text{ with some }\rho<\frac{4}{3}.

The condition ρ<43\rho<\frac{4}{3} is sufficient to guarantee (2.5). For instance, ReV(x)\mathrm{Re}\,V(x) is a polynomial of degree ρ\rho. The range such that the constants β±\beta_{\pm} in Theorem 2.1 can be taken is 0+\mathbb{R}_{0}^{+} since

limxImV(x)=,limx+ImV(x)=+.\lim_{x\to-\infty}\mathrm{Im}\,V(x)=-\infty,\qquad\lim_{x\to+\infty}\mathrm{Im}\,V(x)=+\infty.

Concerning σ±(N)(α)\sigma_{\pm}^{(N)}(\alpha), since the functions τ±(x)N+1|V(x)|\tau_{\pm}(x)^{N+1}|V(x)| are bounded on \mathbb{R} for any N1N\geq 1 (or for any N0N\geq 0 if ρ<1\rho<1), we do not need to compute δα±\delta_{\alpha}^{\pm} in this situation. Accordingly, for any N1N\in\mathbb{N}_{1} and for any β±0+\beta_{\pm}\in\mathbb{R}_{0}^{+}, there exists a family (Ψλ,N)(\Psi_{\lambda,N}) such that

(HVλ)Ψλ,NΨλ,NαN+14,\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}\lesssim\alpha^{-\frac{N+1}{4}},

for all λΩ:={α+iβ:α1 and β[β,β+]}\displaystyle\lambda\in\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\alpha\gtrsim 1\text{ and }\beta\in[-\beta_{-},\beta_{+}]\subset\mathbb{R}\right\}.

Example 3 (Polynomial-like potentials).

Let us take a look at the potential VWlocN+3,()V\in W^{N+3,\infty}_{\textup{loc}}(\mathbb{R}), with N0N\geq 0, satisfying Assumption I with τ±(x)=(x2+1)12\tau_{\pm}(x)=(x^{2}+1)^{-\frac{1}{2}} and having the form

|ReV(x)||x|ρ,|ImV(x)||x|γ,|x|1,|\mathrm{Re}\,V(x)|\lesssim|x|^{\rho},\qquad|\mathrm{Im}\,V(x)|\approx|x|^{\gamma},\qquad|x|\gtrsim 1,

with ρ\rho\in\mathbb{R} and γ0\gamma\geq 0. For examples, ReV\mathrm{Re}\,V and ImV\mathrm{Im}\,V are, respectively, the polynomials of degree ρ\rho and γ\gamma. It is necessary to assume that γ0\gamma\geq 0 in order to meet the condition (2.1) and we assume further that γ>3ρ44\gamma>\frac{3\rho-4}{4} such that (2.5) is satisfied. Accordingly, the fast growth of |ReV(x)||\mathrm{Re}\,V(x)| require the fast growth of ImV(x)\mathrm{Im}\,V(x). In particular if ρ<43\rho<\frac{4}{3} (i.e. |ReV(x)||\mathrm{Re}\,V(x)| grows slower that |x|43|x|^{\frac{4}{3}}) even a bounded ImV\mathrm{Im}\,V fits. In order to apply Theorem 2.1, let us denote the quantity

ω:=max{ρ,γ}.\omega:=\max\{\rho,\gamma\}.

Clearly, ω0\omega\geq 0 and VV is bounded if and only if ω=0\omega=0. When |V||V| is unbounded, δα±\delta_{\alpha}^{\pm} is the solution of the equation

|x|γ(x2+1)12=α3+ε14+ε1.|x|^{\gamma}\left(x^{2}+1\right)^{\frac{1}{2}}=\alpha^{\frac{3+\varepsilon_{1}}{4+\varepsilon_{1}}}.

When xx is large enough, since |x|γ(x2+1)12|x|γ+1|x|^{\gamma}\left(x^{2}+1\right)^{\frac{1}{2}}\approx|x|^{\gamma+1}, we can approximate the solution of the above equation with the notation \approx introduced in Subsection 1.2 as follows

δα=δα+=δα34(γ+1)+ε.\delta^{-}_{\alpha}=\delta^{+}_{\alpha}=\delta\approx\alpha^{\frac{3}{4(\gamma+1)}+\varepsilon}.

Here ε>0\varepsilon>0 can be made arbitrary small by an appropriate choice of small ε1>0\varepsilon_{1}>0. Hence Theorem 2.1 results that

(HVλ)Ψλ,NΨλ,N={𝒪(αN+14),ωN+1,𝒪(αN+14+34ωN1γ+1+ε(ωN1)),ω>N+1,\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}=\left\{\begin{aligned} &\mathcal{O}\left(\alpha^{-\frac{N+1}{4}}\right),\qquad&\omega\leq N+1,\\ &\mathcal{O}\left(\alpha^{-\frac{N+1}{4}+\frac{3}{4}\frac{\omega-N-1}{\gamma+1}+\varepsilon(\omega-N-1)}\right),\qquad&\omega>N+1,\end{aligned}\right. (2.19)

as α+\alpha\to+\infty and β[β,β+]\beta\in[-\beta_{-},\beta_{+}] which is mentioned in (2.7). When VV is bounded, the decay is also included in the case ωN+1\omega\leq N+1. To improve the decay rate in the second case, we let ε\varepsilon be small enough and notice that from the condition γ>3ρ44\gamma>\frac{3\rho-4}{4}, we can control the other term by

34ωN1γ+1<{34if γρ,1if γ<ρ.\frac{3}{4}\frac{\omega-N-1}{\gamma+1}<\left\{\begin{aligned} &\frac{3}{4}\qquad&&\text{if }\gamma\geq\rho,\\ &1\qquad&&\text{if }\gamma<\rho.\end{aligned}\right.

By considering ε\varepsilon very small in (2.19), we see that the pseudomode with N=3N=3 (i.e. we require at least VWloc6,()V\in W^{6,\infty}_{\textup{loc}}(\mathbb{R})) is sufficient to treat all polynomial-like potentials. Comparing this with the same results for Schrödinger operators in [21, Ex. 3.8] and Dirac operators in [20, Ex. 2], more terms in the pseudomode expansion are needed in the higher order differential operators.

Example 4 (Super-exponential potential).

We devote the other application of Theorem 2.1 for the potential that is smooth and grow very fast at infinity, that is

V(x)=cosh(sinh(x))+isinh(sinh(x)).V(x)=\cosh(\sinh(x))+i\sinh(\sinh(x)).

The Assumption I is satisfied with τ±(x)=cosh(x)\tau_{\pm}(x)=\cosh(x) and ν±=0\nu_{\pm}=0. We emphasize here that [21, Asm. I] can not cover this potential, more precisely [21, Cond. 3.2] can not be satisfied. The solution δ\delta of the equation |sinh(sinh(±δ))|cosh(δ)=α3+ε14+ε1\frac{|\sinh(\sinh(\pm\delta))|}{\cosh(\delta)}=\alpha^{\frac{3+\varepsilon_{1}}{4+\varepsilon_{1}}} can be estimated as follows

δα=δα+=δln(ln(α34+ε)),\delta_{\alpha}^{-}=\delta_{\alpha}^{+}=\delta\approx\ln\left(\ln\left(\alpha^{\frac{3}{4}+\varepsilon}\right)\right),

where ε\varepsilon can be made arbitrary small by an appropriate choice of small ε1\varepsilon_{1}. Then, for arbitrary β±0+\beta_{\pm}\in\mathbb{R}_{0}^{+} and for all β[β,β+]\beta\in[-\beta_{-},\beta_{+}]\subset\mathbb{R}, the pseudomodes with N3N\geq 3 leads to a decay

(HVλ)Ψλ,NΨλ,N=𝒪(α2N4+2ε), as α+.\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}=\mathcal{O}\left(\alpha^{\frac{2-N}{4}+2\varepsilon}\right),\qquad\text{ as }\alpha\to+\infty.
Example 5 (Decaying potentials).

Consider a class of potentials with the asymptotic behaviour

V(x)=isgn(x)|x|γ,|x|1, 0<γ<1,V(x)=i\frac{\textup{sgn}(x)}{|x|^{\gamma}},\qquad|x|\gtrsim 1,\ 0<\gamma<1, (2.20)

which spoils the assumption (2.1). However, the analysis in Subsection 4.4 shows us that we can apply the same construction as for Theorem 2.1 to set up pseudomodes for large pseudoeigenvalues λ=α+iβ\lambda=\alpha+i\beta satisfying

|β|α34γ1γ=o(1), as α+,|β|0.|\beta|\alpha^{\frac{3}{4}\frac{\gamma}{1-\gamma}}=o(1),\qquad\text{ as }\alpha\to+\infty,\,|\beta|\to 0. (2.21)

Furthermore, if we make the restriction (2.21) stronger by considering

|β|α34γ1γε, as α+,|\beta|\lesssim\alpha^{-\frac{3}{4}\frac{\gamma}{1-\gamma}-\varepsilon},\qquad\text{ as }\alpha\to+\infty,

we will have a decay

(HVλ)Ψλ,NΨλ,N=𝒪(αN+14), as α+.\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}=\mathcal{O}\left(\alpha^{-\frac{N+1}{4}}\right),\qquad\text{ as }\alpha\to+\infty.

In other words, the Ω\Omega in the main problem (1.2) can be described by (see Figure 2)

Ω={α+iβ:α1,|β|α34γ1γε}.\Omega=\left\{\alpha+i\beta\in\mathbb{C}:\alpha\gtrsim 1,\ |\beta|\lesssim\alpha^{-\frac{3}{4}\frac{\gamma}{1-\gamma}-\varepsilon}\right\}.

In [21, Eqn. 3.24], Krejčiřík and Siegl used this type of decaying potential to give a natural Laptev-Safronov eigenvalues bounds for the Schrödinger operator with LpL^{p}-potentials (p>1p>1) which appears in [23, Theorem 5]. Here, in the same manner, by observing that VLpV\in L^{p} if γp>1\gamma p>1, it yields that

34γ1γ(p1)=34γpγ1γ>34.\frac{3}{4}\frac{\gamma}{1-\gamma}(p-1)=\frac{3}{4}\frac{\gamma p-\gamma}{1-\gamma}>\frac{3}{4}.

From this, we obtain a bound for Ω\Omega, which is also a bound for the distribution of the eigenvalues of the biharmonic operator

Ω{α+iβ:|β|p1=o(α34),α+}.\Omega\subset\left\{\alpha+i\beta\in\mathbb{C}:|\beta|^{p-1}=o\left(\alpha^{-\frac{3}{4}}\right),\alpha\to+\infty\right\}.

If the power of α\alpha is 12-\frac{1}{2} in the Schrödinger case, this power is replaced by 34-\frac{3}{4} for the biharmonic one.

Refer to caption
Figure 2. Illustrations of the shapes of Ω\Omega (in cyan color) associated with the potential V(x)=isgn(x)|x|γV(x)=i\frac{\textup{sgn}(x)}{|x|^{\gamma}} given in (2.20) with γ=12\gamma=\frac{1}{2}. The curves are the graphs of β=±α341100\beta=\pm\alpha^{-\frac{3}{4}-\frac{1}{100}}.

2.2.2. Application of Theorem 2.2

Now, we would like to apply Theorem 2.2 to study the elementary potentials considered in Subsection 2.2.1. It is worthwhile to mention here that the below Examples 6 and 8 can not be covered by [21, Cond. 5.2] or [20, Cond. 4.3].

Example 6.

Let us consider again the logarithmic potential VWlocN,2(0+)V\in W_{\mathrm{loc}}^{N,2}(\mathbb{R}_{0}^{+}) that has the following behaviour

V(x):=iln(x),x1.V(x):=i\ln(x),\qquad x\gtrsim 1. (2.22)

All conditions of Assumption II are satisfied with τ(x)=(x2+1)12x1\tau(x)=\left(x^{2}+1\right)^{-\frac{1}{2}}\approx x^{-1} (ν=1\nu=-1 in (2.3)) as x1x\gtrsim 1, and for instance t1=0t_{1}=0 and some t2>0t_{2}>0. Given β>0\beta>0, then xβ>0x_{\beta}>0 is determined by the relation xβ=eβx_{\beta}=e^{\beta}. Since ReV(x)=0\mathrm{Re}\,V(x)=0, the conditions (2.16) and (2.17) are assured iff

β45exp(45β)αβ43εexp((43ε)β),\beta^{\frac{4}{5}}\exp\left(-\frac{4}{5}\beta\right)\lesssim\alpha\lesssim\beta^{\frac{4}{3}-\varepsilon}\exp\left(\left(\frac{4}{3}-\varepsilon\right)\beta\right), (2.23)

where ε>0\varepsilon>0 can be chosen arbitrarily small by an appropriate choice of small ε1\varepsilon_{1}. Then, thanks to the second inequality in (2.23), we have

κ(β){exp(cβ3ε41exp(β3ε4))if α>β,exp(cβ34exp(β))if αβ.\displaystyle\kappa(\beta)\lesssim\left\{\begin{aligned} &\exp\left(-c\beta^{\frac{3\varepsilon}{4}-1}\exp\left(\beta^{\frac{3\varepsilon}{4}}\right)\right)\qquad&&\text{if }\alpha>\beta,\\ &\exp\left(-c\beta^{-\frac{3}{4}}\exp\left(\beta\right)\right)\qquad&&\text{if }\alpha\leq\beta.\end{aligned}\right.

Let us present here the detail of estimating σ(N)\sigma^{(N)} for N1N\geq 1 (for N=0N=0, it is analogous). If α>β\alpha>\beta, we estimate straightforwardly as follows

σ(N)(β)\displaystyle\sigma^{(N)}(\beta) k=03N1j=1k+N+1exp((k+N+1)β)βjαk+N34+jk=03N1βexp((k+N+1)β)βk+N+14\displaystyle\lesssim\sum_{k=0}^{3N-1}\sum_{j=1}^{k+N+1}\frac{\exp\left(-(k+N+1)\beta\right)\beta^{j}}{\alpha^{\frac{k+N-3}{4}+j}}\lesssim\sum_{k=0}^{3N-1}\frac{\beta\exp\left(-(k+N+1)\beta\right)}{\beta^{\frac{k+N+1}{4}}}
β3N4exp((N+1)β).\displaystyle\lesssim\beta^{\frac{3-N}{4}}\exp\left(-(N+1)\beta\right).

When αβ\alpha\leq\beta, we employ the first inequality in (2.23), we obtain

σ(N)(β)\displaystyle\sigma^{(N)}(\beta) k=03N1exp((k+N+1)β)βk+N+1αk+N34+k+N+1k=03N1βk+N+1exp((k+N+1)β)[β45exp(45β)]54k+54N+14\displaystyle\lesssim\sum_{k=0}^{3N-1}\frac{\exp\left(-(k+N+1)\beta\right)\beta^{k+N+1}}{\alpha^{\frac{k+N-3}{4}+k+N+1}}\lesssim\sum_{k=0}^{3N-1}\frac{\beta^{k+N+1}\exp\left(-(k+N+1)\beta\right)}{\left[\beta^{\frac{4}{5}}\exp\left(-\frac{4}{5}\beta\right)\right]^{\frac{5}{4}k+\frac{5}{4}N+\frac{1}{4}}}
β45exp(45β).\displaystyle\lesssim\beta^{\frac{4}{5}}\exp\left(-\frac{4}{5}\beta\right).

In summary, Theorem 2.2 provides the pseudomodes such that

(HVλ)Ψλ,NΨλ,N={𝒪(β3N4exp((N+1)β)) if α>β,𝒪(β45exp(45β)) if αβ,\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}=\left\{\begin{aligned} &\mathcal{O}\left(\beta^{\frac{3-N}{4}}\exp(-(N+1)\beta)\right)\qquad&&\text{ if }\alpha>\beta,\\ &\mathcal{O}\left(\beta^{\frac{4}{5}}\exp\left(-\frac{4}{5}\beta\right)\right)\qquad&&\text{ if }\alpha\leq\beta,\end{aligned}\right.

where λ\lambda\to\infty in

Ω:={α+iβ:β1 and β45exp(45β)αβ43εexp((43ε)β)}.\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\beta\gtrsim 1\text{ and }\beta^{\frac{4}{5}}\exp\left(-\frac{4}{5}\beta\right)\lesssim\alpha\lesssim\beta^{\frac{4}{3}-\varepsilon}\exp\left(\left(\frac{4}{3}-\varepsilon\right)\beta\right)\right\}. (2.24)

From the definition of Ω\Omega, we see that the pseudospectral region contains even points which stay very close to the line α=0\alpha=0 (see Figure 3 3(a)).

Refer to caption
(a) V(x)=iln(x)V(x)=i\ln(x).
Refer to caption
(b) V(x)=ix12V(x)=ix^{\frac{1}{2}}.
Refer to caption
(c) V(x)=ix2V(x)=ix^{2}.
Figure 3. Illustration of the shapes of Ω\Omega (in cyan color) with the logarithmic function VV given in (2.22) and the polynomial VV given in (2.25): (a) V(x)=iln(x)V(x)=i\ln(x): The curves are the graphs of α=β45exp(45β)\alpha=\beta^{\frac{4}{5}}\exp\left(-\frac{4}{5}\beta\right) and α=β431100exp((431100)β)\alpha=\beta^{\frac{4}{3}-\frac{1}{100}}\exp\left(\left(\frac{4}{3}-\frac{1}{100}\right)\beta\right), (b) V(x)=ix12V(x)=ix^{\frac{1}{2}}: The curves are the graphs of α=β45\alpha=\beta^{-\frac{4}{5}} and α=β41100\alpha=\beta^{4-\frac{1}{100}}, (c) V(x)=ix2V(x)=ix^{2}: The curves are the graphs of α=β25+1100\alpha=\beta^{\frac{2}{5}+\frac{1}{100}} and α=β21100\alpha=\beta^{2-\frac{1}{100}}.
Example 7.

Next, we want to study the polynomial potential on 0+\mathbb{R}_{0}^{+}:

V(x)=ixγ,x1V(x)=ix^{\gamma},\qquad x\gtrsim 1 (2.25)

where γ>0\gamma>0. All the conditions of Assumption II are satisfied with τ\tau chosen as in Example 6 and we can choose t1=t2=0t_{1}=t_{2}=0. Given β>0\beta>0, then xβ>0x_{\beta}>0 is determined by xβ=β1γx_{\beta}=\beta^{\frac{1}{\gamma}}. It is straightforward to check that the conditions (2.16) and (2.17) holds if we choose α=α(β)\alpha=\alpha(\beta) in the following way

β45(11γ)αβ43(1+1γ)ε,\beta^{\frac{4}{5}\left(1-\frac{1}{\gamma}\right)}\lesssim\alpha\lesssim\beta^{\frac{4}{3}\left(1+\frac{1}{\gamma}\right)-\varepsilon}, (2.26)

where ε>0\varepsilon>0 can be chosen arbitrarily small by an appropriate choice of small ε1\varepsilon_{1}. By the second inequality in (2.26), the term κ(β)\kappa(\beta) can be estimated as

κ(β){exp(cβ3ε4)if α>β,exp(cβ1γ+14)if αβ.\displaystyle\kappa(\beta)\lesssim\left\{\begin{aligned} &\exp\left(-c\beta^{\frac{3\varepsilon}{4}}\right)\qquad&&\text{if }\alpha>\beta,\\ &\exp\left(-c\beta^{\frac{1}{\gamma}+\frac{1}{4}}\right)\qquad&&\text{if }\alpha\leq\beta.\end{aligned}\right.

By performing the estimate analogously in Example 6 to control σ(N)(β)\sigma^{(N)}(\beta) and consider NN large enough, we obtain the decay in two following cases.

  1. a)

    If 0<γ<10<\gamma<1, we get

    (HVλ)Ψλ,NΨλ,N={𝒪(β(N+1)(1γ+14)+1) if α>β,𝒪(β45(11γ)) if αβ,\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}=\left\{\begin{aligned} &\mathcal{O}\left(\beta^{-(N+1)\left(\frac{1}{\gamma}+\frac{1}{4}\right)+1}\right)\qquad&&\text{ if }\alpha>\beta,\\ &\mathcal{O}\left(\beta^{\frac{4}{5}\left(1-\frac{1}{\gamma}\right)}\right)\qquad&&\text{ if }\alpha\leq\beta,\end{aligned}\right.

    when λ\lambda\to\infty in the region

    Ω:={α+iβ:β1 and β45(11γ)αβ43(1+1γ)ε}.\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\beta\gtrsim 1\text{ and }\beta^{\frac{4}{5}\left(1-\frac{1}{\gamma}\right)}\lesssim\alpha\lesssim\beta^{\frac{4}{3}\left(1+\frac{1}{\gamma}\right)-\varepsilon}\right\}.

    The shape of this region as γ=12\gamma=\frac{1}{2} is sketched in Figure 3 3(b).

  2. b)

    If γ>1\gamma>1, we need to strengthen the first inequality in (2.26) to obtain the decay of σ(β)\sigma(\beta) and thus

    (HVλ)Ψλ,NΨλ,N={𝒪(β(N+1)(1γ+14)+1) if α>β,𝒪(β45(11γ)ε5N+14) if αβ,\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}=\left\{\begin{aligned} &\mathcal{O}\left(\beta^{-(N+1)\left(\frac{1}{\gamma}+\frac{1}{4}\right)+1}\right)\qquad&&\text{ if }\alpha>\beta,\\ &\mathcal{O}\left(\beta^{\frac{4}{5}\left(1-\frac{1}{\gamma}\right)-\varepsilon\frac{5N+1}{4}}\right)\qquad&&\text{ if }\alpha\leq\beta,\end{aligned}\right.

    as λ\lambda\to\infty in

    Ω:={α+iβ:β1 and β45(11γ)+εαβ43(1+1γ)ε}.\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\beta\gtrsim 1\text{ and }\beta^{\frac{4}{5}\left(1-\frac{1}{\gamma}\right)+\varepsilon}\lesssim\alpha\lesssim\beta^{\frac{4}{3}\left(1+\frac{1}{\gamma}\right)-\varepsilon}\right\}.

    The shape of this region as γ=2\gamma=2 is given in Figure 3 3(c).

Example 8.

The final example that we want to study is the superexponential potential

V(x)=ieex,x1.V(x)=ie^{e^{x}},\qquad x\gtrsim 1. (2.27)

All conditions of Assumption II are satisfied with τ(x)=ex\tau(x)=e^{x}, ν=0\nu=0 and t1=t2=0t_{1}=t_{2}=0. Given β>0\beta>0, the turning point xβx_{\beta} is determined by the relation

xβ=ln(ln(β)),β1.x_{\beta}=\ln\left(\ln(\beta)\right),\qquad\beta\gtrsim 1.

Then τ(xβ)=ln(β)\tau(x_{\beta})=\ln(\beta) and the conditions (2.17) are equivalent to

[βln(β)]45α[βln(β)]43ε,\left[\beta\ln(\beta)\right]^{\frac{4}{5}}\lesssim\alpha\lesssim\left[\frac{\beta}{\ln(\beta)}\right]^{\frac{4}{3}-\varepsilon}, (2.28)

where ε>0\varepsilon>0 can be chosen arbitrarily small by an appropriate choice of small ε1\varepsilon_{1} as above examples. By the second inequality in (2.28), we obtain the estimate for κ(β)\kappa(\beta):

κ(β){exp(c(βln(β)3ε4)if α>β,exp(cβ14ln(β))if αβ.\displaystyle\kappa(\beta)\lesssim\left\{\begin{aligned} &\exp\left(-c\left(\frac{\beta}{\ln(\beta}\right)^{\frac{3\varepsilon}{4}}\right)&&\text{if }\alpha>\beta,\\ &\exp\left(-c\frac{\beta^{\frac{1}{4}}}{\ln(\beta)}\right)&&\text{if }\alpha\leq\beta.\end{aligned}\right.

In order to get the decay of σ(N)(β)\sigma^{(N)}(\beta), we need to strengthen the first inequality in (2.28) as follows

β45+εln(β)45α[βln(β)]43ε.\beta^{\frac{4}{5}+\varepsilon}\ln(\beta)^{\frac{4}{5}}\lesssim\alpha\lesssim\left[\frac{\beta}{\ln(\beta)}\right]^{\frac{4}{3}-\varepsilon}. (2.29)

then it yields that

(HVλ)Ψλ,NΨλ,N={𝒪(ln(β)N+1β3N4) if α>β,𝒪(ln(β)45β45ε5N+14) if αβ,\frac{\|(H_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}=\left\{\begin{aligned} &\mathcal{O}\left(\ln(\beta)^{N+1}\beta^{\frac{3-N}{4}}\right)\qquad&&\text{ if }\alpha>\beta,\\ &\mathcal{O}\left(\ln(\beta)^{\frac{4}{5}}\beta^{\frac{4}{5}-\varepsilon\frac{5N+1}{4}}\right)\qquad&&\text{ if }\alpha\leq\beta,\end{aligned}\right.

as λ\lambda\to\infty in

Ω:={α+iβ:β1 and β45+εln(β)45α[βln(β)]43ε}.\Omega:=\left\{\alpha+i\beta\in\mathbb{C}:\beta\gtrsim 1\text{ and }\beta^{\frac{4}{5}+\varepsilon}\ln(\beta)^{\frac{4}{5}}\lesssim\alpha\lesssim\left[\frac{\beta}{\ln(\beta)}\right]^{\frac{4}{3}-\varepsilon}\right\}.
Refer to caption
Figure 4. Illustration of the shape of Ω\Omega (in cyan color) with the superexponential VV given in (2.27) in which the curves are the graphs of α=β45+1100ln(β)45\alpha=\beta^{\frac{4}{5}+\frac{1}{100}}\ln(\beta)^{\frac{4}{5}} and α=[βln(β]431100\alpha=\left[\frac{\beta}{\ln(\beta}\right]^{\frac{4}{3}-\frac{1}{100}}.

From Figure 3 and Figure 4, we see that the pseudospectrum of the operator comes close to the imaginary axis as the potential grows slowly at infinity as logarithmic functions and root-type ones xγx^{\gamma} with γ(0,1)\gamma\in(0,1). When the potential grows faster as polynomial and exponential functions, the pseudospectrum stay away the axes accordingly.

2.2.3. Application of Theorem 2.3

We recall some standard notions that is introduced in [22, Sec. IV]. The symbol associated with HhH_{h} is

f(x,ξ)=ξ4+W(x),f(x,\xi)=\xi^{4}+W(x),

and its semi-classical pseudospectrum of HhH_{h} is given by the closure of the set

Λ={ξ4+W(x):ξ3ImW(x)<0}.\Lambda=\left\{\xi^{4}+W(x):\xi^{3}\mathrm{Im}\,W^{\prime}(x)<0\right\}.

Since the sign of ξ\xi can be chosen freely, we can describe Λ\Lambda as

Λ={ξ4+W(x):ξ0,ImW(x)0},\Lambda=\left\{\xi^{4}+W(x):\xi\neq 0,\ \mathrm{Im}\,W^{\prime}(x)\neq 0\right\},

this condition is also given in Davies’ work [7] for Schrödinger operator. Then zΛz\in\Lambda if and only if there exists (ξ0,x0)2(\xi_{0},x_{0})\in\mathbb{R}^{2} such that z=ξ04+W(x0)z=\xi_{0}^{4}+W(x_{0}) with ξ0{0}\xi_{0}\in\mathbb{R}\setminus\{0\} and ImW(x0)0\mathrm{Im}\,W^{\prime}(x_{0})\neq 0. Taylor’s Theorem yields

ImW(x)ImW(x0)=ImW(x0)(xx0)+𝒪(|xx0|2).\mathrm{Im}\,W(x)-\mathrm{Im}\,W(x_{0})=\mathrm{Im}\,W^{\prime}(x_{0})(x-x_{0})+\mathcal{O}\left(|x-x_{0}|^{2}\right).

Then, there exists a neighbourhood of x0x_{0} in which the sign of the function ImW(x)ImW(x0)\mathrm{Im}\,W(x)-\leavevmode\nobreak\ \mathrm{Im}\,W(x_{0}) changes at the point x0x_{0} and we can apply Theorem 2.3 to obtain the pseudomode for the operator HhH_{h} corresponding to pseudoeigenvalue zz. Furthermore, we can extend Λ\Lambda to the set

Λ~={ξ4+W(x)|ξ0, there exists p0 such that ImW(j)(x)=0 for all j0,j2p and ImW(2p+1)(x)0}\displaystyle\widetilde{\Lambda}=\left\{\xi^{4}+W(x)\left|\begin{aligned} &\xi\neq 0,\text{ there exists }p\in\mathbb{N}_{0}\text{ such that }\\ &\mathrm{Im}\,W^{(j)}(x)=0\text{ for all }j\neq 0,j\leq 2p\text{ and }\mathrm{Im}\,W^{(2p+1)}(x)\neq 0\end{aligned}\right.\right\}

which appears in the paper [26] for Schrödinger operator, and the above result remains true. In summary, we obtained the result as in the papers [7, 26] for the biharmonic operator.

3. WKB construction

3.1. The WKB expansion

Let Pλ:P_{\lambda}:\mathbb{R}\to\mathbb{C} be a sufficient regular function depending on the parameter λ\lambda which will be determined later. We consider the formal conjugated operator of Vλ\mathscr{L}_{V}-\lambda:

ePλ(Vλ)ePλ=ePλ(d4dx4+V(x)λ)ePλ=𝒟λ+λ,\displaystyle e^{P_{\lambda}}\left(\mathscr{L}_{V}-\lambda\right)e^{-P_{\lambda}}=e^{P_{\lambda}}\left(\frac{\mathrm{d}^{4}}{\mathrm{d}x^{4}}+V(x)-\lambda\right)e^{-P_{\lambda}}=\mathcal{D}_{\lambda}+\mathcal{R}_{\lambda},

where 𝒟λ\mathcal{D}_{\lambda} has a differential expression and λ\mathcal{R}_{\lambda} has a multiplication expression

𝒟λd4dx44Pλ(1)ddx3+6[(Pλ(1))2Pλ(2)]d2dx24[Pλ(3)3Pλ(1)Pλ(2)+(Pλ(1))3]ddx,\displaystyle\mathcal{D}_{\lambda}\coloneqq\frac{\mathrm{d}^{4}}{\mathrm{d}x^{4}}-4P_{\lambda}^{(1)}\frac{\mathrm{d}}{\mathrm{d}x^{3}}+6\left[\left(P_{\lambda}^{(1)}\right)^{2}-P_{\lambda}^{(2)}\right]\frac{\mathrm{d}^{2}}{\mathrm{d}x^{2}}-4\left[P_{\lambda}^{(3)}-3P_{\lambda}^{(1)}P_{\lambda}^{(2)}+\left(P_{\lambda}^{(1)}\right)^{3}\right]\frac{\mathrm{d}}{\mathrm{d}x}, (3.1)
λPλ(4)+4Pλ(1)Pλ(3)+3(Pλ(2))26(Pλ(1))2Pλ(2)+(Pλ(1))4+V(x)λ.\displaystyle\mathcal{R}_{\lambda}\coloneqq-P_{\lambda}^{(4)}+4P_{\lambda}^{(1)}P_{\lambda}^{(3)}+3\left(P_{\lambda}^{(2)}\right)^{2}-6\left(P_{\lambda}^{(1)}\right)^{2}P_{\lambda}^{(2)}+\left(P_{\lambda}^{(1)}\right)^{4}+V(x)-\lambda.

The general WKB strategy is as follows. We look for the pseudomodes in the form

Ψλ=ξλePλ,\Psi_{\lambda}=\xi_{\lambda}e^{-P_{\lambda}},

where ξλ\xi_{\lambda} is a cut-off function whose shape is determined later depending on the behaviour of VV at infinity (±\pm\infty). From the triangular inequality, it yields that

(Vλ)ΨλΨλ=ePλ𝒟λξλ+ξλePλλΨλePλ𝒟λξλΨλ+λL(Jλ).\frac{\left\|\left(\mathscr{L}_{V}-\lambda\right)\Psi_{\lambda}\right\|}{\|\Psi_{\lambda}\|}=\frac{\left\|e^{-P_{\lambda}}\mathcal{D_{\lambda}}\xi_{\lambda}+\xi_{\lambda}e^{-P_{\lambda}}\mathcal{R}_{\lambda}\right\|}{\left\|\Psi_{\lambda}\right\|}\leq\frac{\left\|e^{-P_{\lambda}}\mathcal{D_{\lambda}}\xi_{\lambda}\right\|}{\left\|\Psi_{\lambda}\right\|}+\left\|\mathcal{R}_{\lambda}\right\|_{L^{\infty}(J_{\lambda})}. (3.2)

Here JλJ_{\lambda} is the support of the cut-off ξλ\xi_{\lambda}. The WKB idea is to look for the phase PλP_{\lambda} in the following form,

Pλ,n(x)=k=1n1λkψk(x),n0,P_{\lambda,n}(x)=\sum_{k=-1}^{n-1}\lambda^{-k}\psi_{k}(x),\qquad n\in\mathbb{N}_{0}, (3.3)

where functions (ψk)k[[1,n1]]\left(\psi_{k}\right)_{k\in[[-1,n-1]]} are to be determined by solving some ordinary differential equations (ODEs); the number nn is chosen later depending on the maximal possible order derivative of VV. In principle, by letting λ\lambda\to\infty, the first term in the right hand side of (3.2) can be shown exponentially decay thanks to consideration on the support of ξλ\xi_{\lambda} and the second term often decreases with the rate power of λ1\lambda^{-1}. The more regular the potential is, the stronger the rate of decay in (3.2) is obtained.

Let us start with n=0n=0 and put Pλ,0P_{\lambda,0} into the formula of λ\mathcal{R}_{\lambda} in (3.1), we obtain

λ,0:=λψ1(4)+4λ2ψ1(1)ψ1(3)+3λ2(ψ1(2))26λ3(ψ1(1))2ψ1(2)+λ4(ψ1(1))4+V(x)λ.\mathcal{R}_{\lambda,0}:=-\lambda\psi_{-1}^{(4)}+4\lambda^{2}\psi_{-1}^{(1)}\psi_{-1}^{(3)}+3\lambda^{2}\left(\psi_{-1}^{(2)}\right)^{2}-6\lambda^{3}\left(\psi_{-1}^{(1)}\right)^{2}\psi_{-1}^{(2)}+\lambda^{4}\left(\psi_{-1}^{(1)}\right)^{4}+V(x)-\lambda.

By solving the equation λ4(ψ1(1))4+V(x)λ=0\lambda^{4}(\psi_{-1}^{(1)})^{4}+V(x)-\lambda=0, we often call it “eikonal” equation, the forth order of λ\lambda in Rλ,0R_{\lambda,0} is removed:

λ,0:=λψ1(4)+4λ2ψ1(1)ψ1(3)+3λ2(ψ1(2))26λ3(ψ1(1))2ψ1(2).\mathcal{R}_{\lambda,0}:=-\lambda\psi_{-1}^{(4)}+4\lambda^{2}\psi_{-1}^{(1)}\psi_{-1}^{(3)}+3\lambda^{2}(\psi_{-1}^{(2)})^{2}-6\lambda^{3}(\psi_{-1}^{(1)})^{2}\psi_{-1}^{(2)}. (3.4)

From this simple observation, we wish that when we increase nn in Pλ,nP_{\lambda,n}, the order of λ\lambda appearing in the remainder λ,n\mathcal{R}_{\lambda,n} has been reduced accordingly. For n1n\in\mathbb{N}_{1}, we replace Pλ,nP_{\lambda,n} into λ\mathcal{R}_{\lambda} in (3.1), we get

λ,n\displaystyle\mathcal{R}_{\lambda,n} k=1n1λkψk(4)+4k=22n2λkα1+α2=kψα1(1)ψα2(3)+3k=22n2λkα1+α2=kψα1(2)ψα2(2)\displaystyle\coloneqq-\sum_{k=-1}^{n-1}\lambda^{-k}\psi_{k}^{(4)}+4\sum_{k=-2}^{2n-2}\lambda^{-k}\sum_{\alpha_{1}+\alpha_{2}=k}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}+3\sum_{k=-2}^{2n-2}\lambda^{-k}\sum_{\alpha_{1}+\alpha_{2}=k}\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)}
6k=33n3λkα1+α2+α3=kψα1(1)ψα2(1)ψα3(2)+k=44n4λkα1+α2+α3+α4=kψα1(1)ψα2(1)ψα3(1)ψα4(1)\displaystyle\qquad-6\sum_{k=-3}^{3n-3}\lambda^{-k}\sum_{\alpha_{1}+\alpha_{2}+\alpha_{3}=k}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(2)}+\sum_{k=-4}^{4n-4}\lambda^{-k}\sum_{\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=k}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(1)}\psi_{\alpha_{4}}^{(1)}
+V(x)λ\displaystyle\qquad+V(x)-\lambda
k=44n4λkϕk+3.\displaystyle\eqqcolon\sum_{k=-4}^{4n-4}\lambda^{-k}\phi_{k+3}.

Here the function ϕk\phi_{k} with k[[4,4n4]]k\in[[-4,4n-4]] are naturally defined by grouping together the terms attached with the same order of λ\lambda, with the exception of V(x)λV(x)-\lambda which we include in the leading order term:

λ4:\displaystyle\lambda^{4}: (ψ1(1))4+V(x)λλ4ϕ1,\displaystyle\left(\psi_{-1}^{(1)}\right)^{4}+\frac{V(x)-\lambda}{\lambda^{4}}\eqqcolon\phi_{-1},
λ3:\displaystyle\lambda^{3}: 4(ψ1(1))3ψ0(1)6(ψ1(1))2ψ1(2)ϕ0,\displaystyle 4\left(\psi_{-1}^{(1)}\right)^{3}\psi_{0}^{(1)}-6\left(\psi_{-1}^{(1)}\right)^{2}\psi_{-1}^{(2)}\eqqcolon\phi_{0},
λ2:\displaystyle\lambda^{2}: 4(ψ1(1))3ψ1(1)+6(ψ1(1))2(ψ0(1))26(ψ1(1))2ψ0(2)\displaystyle 4\left(\psi_{-1}^{(1)}\right)^{3}\psi_{1}^{(1)}+6\left(\psi_{-1}^{(1)}\right)^{2}\left(\psi_{0}^{(1)}\right)^{2}-6\left(\psi_{-1}^{(1)}\right)^{2}\psi_{0}^{(2)}
12ψ1(1)ψ0(1)ψ1(2)+3(ψ1(2))2+4ψ1(1)ψ1(3)ϕ1,\displaystyle\hskip 71.13188pt-12\psi_{-1}^{(1)}\psi_{0}^{(1)}\psi_{-1}^{(2)}+3\left(\psi_{-1}^{(2)}\right)^{2}+4\psi_{-1}^{(1)}\psi_{-1}^{(3)}\eqqcolon\phi_{1},
\displaystyle\ldots

For k[[3,4n4]]k\in[[-3,4n-4]], the formulae can be written as

ψk(4)+4α1+α2=kψα1(1)ψα2(3)+3α1+α2=kψα1(2)ψα2(2)\displaystyle-\psi_{k}^{(4)}+4\sum_{\alpha_{1}+\alpha_{2}=k}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}+3\sum_{\alpha_{1}+\alpha_{2}=k}\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)} (3.5)
6α1+α2+α3=kψα1(1)ψα2(1)ψα3(2)+α1+α2+α3+α4=kψα1(1)ψα2(1)ψα3(1)ψα4(1)ϕk+3,\displaystyle-6\sum_{\alpha_{1}+\alpha_{2}+\alpha_{3}=k}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(2)}+\sum_{\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=k}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(1)}\psi_{\alpha_{4}}^{(1)}\eqqcolon\phi_{k+3},

with the convention that ψα=0\psi_{\alpha}=0 if α2\alpha\leq-2 or αn\alpha\geq n.

Given n1n\in\mathbb{N}_{1}, requiring ϕk=0\phi_{k}=0 for all k[[1,n1]]k\in[[-1,n-1]], we obtain n+1n+1 ODEs which can be solved explicitly to find all (ψk(1))k[[1,n1]]\left(\psi_{k}^{(1)}\right)_{k\in[[-1,n-1]]} by a recursion formula

ψ1(1)\displaystyle\psi_{-1}^{(1)} =±iλ1(λV)1/4,\displaystyle=\pm i\lambda^{-1}\left(\lambda-V\right)^{1/4}, (3.6)
ψk+1(1)\displaystyle\psi_{k+1}^{(1)} =14(ψ1(1))3(ψk2(4)4α1+α2=k2ψα1(1)ψα2(3)3α1+α2=k2ψα1(2)ψα2(2)\displaystyle=\frac{1}{4(\psi_{-1}^{(1)})^{3}}\left(\psi_{k-2}^{(4)}-4\sum_{\alpha_{1}+\alpha_{2}=k-2}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}-3\sum_{\alpha_{1}+\alpha_{2}=k-2}\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)}\right.
+6α1+α2+α3=k2ψα1(1)ψα2(1)ψα3(2)α1+α2+α3+α4=k2α1,α2,α3,α4k+1ψα1(1)ψα2(1)ψα3(1)ψα4(1)),\displaystyle\left.\hskip 56.9055pt+6\sum_{\alpha_{1}+\alpha_{2}+\alpha_{3}=k-2}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(2)}-\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=k-2\\ \alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4}\neq k+1\end{subarray}}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(1)}\psi_{\alpha_{4}}^{(1)}\right), (3.7)

for k[[1,n2]]k\in[[-1,n-2]], with the convention that ψα=0\psi_{\alpha}=0 if α2\alpha\leq-2 or αn\alpha\geq n. After solving these ODEs, the WKB remainder is

λ,n=k=n34n4λkϕk+3,n1.\mathcal{R}_{\lambda,n}=\sum_{k=n-3}^{4n-4}\lambda^{-k}\phi_{k+3},\qquad n\in\mathbb{N}_{1}. (3.8)

Since λV(x)\lambda-V(x) is a complex-valued function, the forth root appearing in (3.6) is considered as the principal branch of the forth root which is defined as

z14=12(|z|1/2+12(|z|+Rez)1/2)1/2+i2Imz(|z|+Rez)1/2(|z|1/2+12(|z|+Rez)1/2)1/2.z^{\frac{1}{4}}=\frac{1}{\sqrt{2}}\left(|z|^{1/2}+\frac{1}{\sqrt{2}}\left(|z|+\mathrm{Re}\,z\right)^{1/2}\right)^{1/2}+\frac{i}{2}\frac{\mathrm{Im}\,z}{\left(|z|+\mathrm{Re}\,z\right)^{1/2}\left(|z|^{1/2}+\frac{1}{\sqrt{2}}\left(|z|+\mathrm{Re}\,z\right)^{1/2}\right)^{1/2}}. (3.9)

Although there are four solutions for the eikonal equation, that is ±λ1(λV)1/4\pm\lambda^{-1}\left(\lambda-V\right)^{1/4} and ±iλ1(λV)1/4\pm i\lambda^{-1}\left(\lambda-V\right)^{1/4}, but only the latter are suitable for our pseudomodes. The choice of the sign in the definition of ψ1(1)\psi_{-1}^{(1)} will be determined by the sign of ImV\mathrm{Im}\,V at infinity (see Remark 4.6).

3.2. Structure of solutions of the transport equations and the WKB remainder

From now on, we assume that we are dealing with the plus sign in the formula of ψ1\psi_{-1} in (3.6), unless otherwise stated. Let us list some first solutions of the first transport equations, ψk(1)\psi_{k}^{(1)} for k0k\geq 0, to see which structure they are equipped with:

ψ0(1)=38V(1)Vλ,ψ1(1)=iλVλ(1/4)(516V(2)Vλ+45128(V(1))2Vλ2).\displaystyle\psi_{0}^{(1)}=-\frac{3}{8}\frac{V^{(1)}}{V_{\lambda}},\qquad\psi_{1}^{(1)}=\frac{i\lambda}{V_{\lambda}^{(1/4)}}\left(\frac{5}{16}\frac{V^{(2)}}{V_{\lambda}}+\frac{45}{128}\frac{(V^{(1)})^{2}}{V_{\lambda}^{2}}\right).

If we continue, we will see that ψ2(1)\psi_{2}^{(1)} and ψ3(1)\psi_{3}^{(1)} has the form

ψ2(1)=λ2Vλ24{V(3)Vλ,V(1)V(2)Vλ2,(V(1))3Vλ3},\displaystyle\psi_{2}^{(1)}=\frac{\lambda^{2}}{V_{\lambda}^{\frac{2}{4}}}\left\{\frac{V^{(3)}}{V_{\lambda}},\frac{V^{(1)}V^{(2)}}{V_{\lambda}^{2}},\frac{\left(V^{(1)}\right)^{3}}{V_{\lambda}^{3}}\right\},
ψ3(1)=λ3Vλ34{V(4)Vλ,V(1)V(3)Vλ2,(V(2))2Vλ2,(V(1))2V(2)Vλ3,(V(1))4Vλ4},\displaystyle\psi_{3}^{(1)}=\frac{\lambda^{3}}{V_{\lambda}^{\frac{3}{4}}}\left\{\frac{V^{(4)}}{V_{\lambda}},\frac{V^{(1)}V^{(3)}}{V_{\lambda}^{2}},\frac{\left(V^{(2)}\right)^{2}}{V_{\lambda}^{2}},\frac{\left(V^{(1)}\right)^{2}V^{(2)}}{V_{\lambda}^{3}},\frac{\left(V^{(1)}\right)^{4}}{V_{\lambda}^{4}}\right\},

where the bracket denotes a linear combination of all elements in the bracket with complex coefficients. To estimate the transport solutions later, the coefficients attached with these elements are not important, instead the structure they share together is essential. That is: for each k1k\in\mathbb{N}_{1}, each element in the bracket of ψk(1)\psi_{k}^{(1)} has the form

(V(1))α1(V(2))α2(V(s))αsVλj,\frac{(V^{(1)})^{\alpha_{1}}(V^{(2)})^{\alpha_{2}}...(V^{(s)})^{\alpha_{s}}}{V_{\lambda}^{j}},

in which s=k+1js=k+1-j and all (αi)i[[1,s]]0s\left(\alpha_{i}\right)_{i\in[[1,s]]}\in\mathbb{N}_{0}^{s} satisfy

α1+α2++αs=j,1α1+2α2++sαs=k+1.\alpha_{1}+\alpha_{2}+\ldots+\alpha_{s}=j,\qquad 1\alpha_{1}+2\alpha_{2}+\ldots+s\alpha_{s}=k+1.

This is the content of the following lemma, but first, some notations should be introduced.

Notation 3.1.

For j,r0j,r\in\mathbb{N}_{0} such that jrj\leq r, we employ the following notations

Dr,j:={αr,jcα(V(1))α1(V(2))α2(V(rj+1))αrj+1:cα},D_{r,j}:=\left\{\displaystyle\sum_{\alpha\in\mathcal{I}_{r,j}}c_{\alpha}(V^{(1)})^{\alpha_{1}}(V^{(2)})^{\alpha_{2}}...(V^{(r-j+1)})^{\alpha_{r-j+1}}:c_{\alpha}\in\mathbb{C}\right\},

where

r,j:={α0rj+1:p=1rj+1pαp=r and p=1rj+1αp=j}.\mathcal{I}_{r,j}:=\left\{\alpha\in\mathbb{N}_{0}^{r-j+1}:\sum_{p=1}^{r-j+1}p\alpha_{p}=r\text{ and }\sum_{p=1}^{r-j+1}\alpha_{p}=j\right\}\,. (3.10)

When r,j=\mathcal{I}_{r,j}=\emptyset, we make a convention that Dr,j={0}D_{r,j}=\{0\}. Thus, Dr,0={0}D_{r,0}=\{0\} if r1r\geq 1.

Lemma 3.2.

Let n0n\in\mathbb{N}_{0}, VWlocn+3,2()V\in W_{\mathrm{loc}}^{n+3,2}(\mathbb{R}) and functions (ψk(1))k[[1,n1]]\left(\psi_{k}^{(1)}\right)_{k\in[[-1,n-1]]} be determined by (3.6) and (3.7). When xx\in\mathbb{R} such that Vλ(x)0V_{\lambda}(x)\in\mathbb{C}\setminus\mathbb{R}_{0}^{-}, then we have

ψk(m)(x)=λkVλ(x)k4j=0k+mdk+m,j(x)Vλ(x)j,m[[1,n+3k]], with dr,jDr,j.\psi_{k}^{(m)}(x)=\frac{\lambda^{k}}{V_{\lambda}(x)^{\frac{k}{4}}}\sum_{j=0}^{k+m}\frac{d_{k+m,j}(x)}{V_{\lambda}(x)^{j}},\qquad m\in[[1,n+3-k]],\text{ with }d_{r,j}\in D_{r,j}. (3.11)

Moreover, dr,0=0d_{r,0}=0 if r1r\geq 1.

The condition that the range of VλV_{\lambda} needs to stay away from (,0](-\infty,0] is added to ensure that Vλ14V_{\lambda}^{\frac{1}{4}} is well-defined (i.e. non-multi-valued) and differentiable inherited from the differentiability of VV (since the principal branch of forth root is holomorphic on (,0]\mathbb{C}\setminus(-\infty,0]). From (3.4) and (3.6), the remainder when we solve up to ψ1\psi_{-1} can be calculated explicitly:

λ,0(x)=iVλ(x)1/4(14V(3)(x)Vλ(x)+916V(1)(x)V(2)(x)Vλ2(x)+2164(V(1)(x))3Vλ3(x))\displaystyle\mathcal{R}_{\lambda,0}(x)=iV_{\lambda}(x)^{1/4}\left(\frac{1}{4}\frac{V^{(3)}(x)}{V_{\lambda}(x)}+\frac{9}{16}\frac{V^{(1)}(x)V^{(2)}(x)}{V_{\lambda}^{2}(x)}+\frac{21}{64}\frac{(V^{(1)}(x))^{3}}{V_{\lambda}^{3}(x)}\right) (3.12)
+Vλ(x)2/4(V(2)(x)Vλ(x)+916(V(1)(x))2Vλ(x)2)+iVλ(x)3/4(32V(1)(x)Vλ(x)).\displaystyle\hskip 45.52458pt+V_{\lambda}(x)^{2/4}\left(\frac{V^{(2)}(x)}{V_{\lambda}(x)}+\frac{9}{16}\frac{(V^{(1)}(x))^{2}}{V_{\lambda}(x)^{2}}\right)+iV_{\lambda}(x)^{3/4}\left(-\frac{3}{2}\frac{V^{(1)}(x)}{V_{\lambda}(x)}\right).

We see that the remainder contains the elements which still share the structure mentioned above. Thanks to the recursion formula (3.7), we can show by induction that the shape of the remainder λ,n\mathcal{R}_{\lambda,n} can be written by means of Notation 3.1 as follows:

Lemma 3.3.

Let n0n\in\mathbb{N}_{0}, VWlocn+3,2()V\in W_{\mathrm{loc}}^{n+3,2}(\mathbb{R}) and functions (ψk(1))k[[1,n1]]\left(\psi_{k}^{(1)}\right)_{k\in[[-1,n-1]]} be determined by (3.6) and (3.7), (ϕk)k[[1,4n1]]\left(\phi_{k}\right)_{k\in[[-1,4n-1]]} be as in (3.5) and Rλ,nR_{\lambda,n} as in (3.8). When xx\in\mathbb{R} such that Vλ(x)0V_{\lambda}(x)\in\mathbb{C}\setminus\mathbb{R}_{0}^{-}, then the maximal order derivative of VV in λ,n\mathcal{R}_{\lambda,n} is n+3n+3 and

λ,0(x)=k=021Vλ(x)k34j=1k+1dk+1,j(x)Vλj(x),\displaystyle\mathcal{R}_{\lambda,0}(x)=\sum_{k=0}^{2}\frac{1}{V_{\lambda}(x)^{\frac{k-3}{4}}}\sum_{j=1}^{k+1}\frac{d_{k+1,j}(x)}{V_{\lambda}^{j}(x)},
λ,n(x)=k=03n11Vλ(x)k+n34j=1k+n+1dk+n+1,j(x)Vλj(x),n1, with dr,jDr,j.\displaystyle\mathcal{R}_{\lambda,n}(x)=\sum_{k=0}^{3n-1}\frac{1}{V_{\lambda}(x)^{\frac{k+n-3}{4}}}\sum_{j=1}^{k+n+1}\frac{d_{k+n+1,j}(x)}{V_{\lambda}^{j}(x)},\qquad n\geq 1,\text{ with }d_{r,j}\in D_{r,j}.

The proofs of these lemmata is postponed to Appendix A.

4. Pseudomodes for large real pseudoeigenvalues

We reserve this section for proving Theorem 2.1. In other words, we are going to construct the pseudomode for the perturbed biharmonic operator when the real part of the spectral parameter λ\lambda is considered largely.

From the assumptions (2.2) and (2.7), there exist constants a±>0a_{\pm}>0 such that, for all βB=[β,β+]\beta\in B=[-\beta_{-},\beta_{+}],

ImV(x)βImV(x)1,\displaystyle\mathrm{Im}\,V(x)-\beta\lesssim\mathrm{Im}\,V(x)\lesssim-1,\qquad xI{x0:xa},\displaystyle x\in I^{-}\coloneqq\left\{x\in\mathbb{R}_{0}^{-}:x\leq-a_{-}\right\}, (4.1)
ImV(x)βImV(x)1,\displaystyle\mathrm{Im}\,V(x)-\beta\gtrsim\mathrm{Im}\,V(x)\gtrsim 1,\qquad xI+{x0+:xa+}.\displaystyle x\in I^{+}\coloneqq\left\{x\in\mathbb{R}_{0}^{+}:x\geq a_{+}\right\}.

Notice that the constants in the above notations \lesssim and \gtrsim are uniformly on β\beta. Possibly considering larger a±a_{\pm}, we assume that all the remain assumptions of Assumption I also happen on I±I^{\pm}. Furthermore, two constants a±a_{\pm} will be used as universal constants in this section, i.e. the size of a±a_{\pm} can be changed a finite number of times but we still keep denoting them as a±a_{\pm}.

4.1. Shapes of the cut-off functions

The cut-off function is employed to complete two tasks in our construction of pseudomodes: first, by attaching with the functions created by WKB method, the cut-off make them belong to the domain of the operator; second, the cut-off makes (λV)14(\lambda-V)^{\frac{1}{4}} well-defined (i.e. non-multi-valued) and differentiable in its support. In the latter, the differentiability of (λV)14(\lambda-V)^{\frac{1}{4}} comes from the analyticity of the forth root on (,0]\mathbb{C}\setminus(-\infty,0] and the regularity of VV with the requirement that λV(x)(,0]\lambda-V(x)\in\mathbb{C}\setminus(-\infty,0] on the support of the cut-off. Let us denote by ξ:[0,1]\xi:\mathbb{R}\to[0,1] the cut-off function satisfying the following properties

{ξ𝒞0(),ξ(x)=1on (δα+Δα,δα+Δα+),ξ(x)=0on (δα,δα+),\left\{\begin{aligned} &\xi\in\mathcal{C}_{0}^{\infty}(\mathbb{R}),\\ &\xi(x)=1\qquad\text{on }\left(-\delta_{\alpha}^{-}+\Delta_{\alpha}^{-},\delta_{\alpha}^{+}-\Delta_{\alpha}^{+}\right),\\ &\xi(x)=0\qquad\text{on }\mathbb{R}\setminus\left(-\delta_{\alpha}^{-},\delta_{\alpha}^{+}\right),\end{aligned}\right. (4.2)

where δα±\delta_{\alpha}^{\pm} and Δα±<δα±\Delta_{\alpha}^{\pm}<\delta_{\alpha}^{\pm} are α\alpha-dependent positive numbers which will be determined later. Notice that the cut-off ξ\xi can be selected in such a way that

ξ(j)L(±)(Δα±)j,j[[1,4]].\|\xi^{(j)}\|_{L^{\infty}(\mathbb{R}_{\pm})}\lesssim\left(\Delta_{\alpha}^{\pm}\right)^{-j},\qquad j\in[[1,4]]. (4.3)

To simplify the notation, we introduce the following sets

Jα:=[δα,0],\displaystyle J_{\alpha}^{-}=\left[-\delta_{\alpha}^{-},0\right],\qquad Jα+:=[0,δα+],\displaystyle J_{\alpha}^{+}=\left[0,\delta_{\alpha}^{+}\right],\qquad Jα:=JαJα+.\displaystyle J_{\alpha}=J_{\alpha}^{-}\cup J_{\alpha}^{+}.

We use the following lemma (see the proof in [20, Lemma 3.1]) to define the boundary δα±\delta_{\alpha}^{\pm} of the cut-off ξ\xi.

Lemma 4.1.

Let g:0+0+g:\mathbb{R}_{0}^{+}\to\mathbb{R}_{0}^{+} be a continuous function and let α\alpha be a positive number, we define

δ(α):=inf{x0:g(x)=α}.\delta(\alpha):=\inf\left\{x\geq 0:g(x)=\alpha\right\}.

Then δ(α)\delta(\alpha) can be infinite (inf=+)(\inf\emptyset=+\infty), however, when gg is unbounded at ++\infty and for all sufficiently large α>0\alpha>0, the number δ(α)\delta(\alpha) is finite and

limα+δ(α)=+.\lim_{\alpha\to+\infty}\delta(\alpha)=+\infty.

Furthermore, if α>g(0)\alpha>g(0) then

g(x)α,x[0,δ(α)].g(x)\leq\alpha,\qquad x\in[0,\delta(\alpha)].

We consider the following cases:

  1. a)

    When VV is unbounded at ±\pm\infty: From the assumption (2.5), we have

    τ±4+ε1(x)|ImV(x)|,|ReV(x)|3+ε1τ±4(x)|ImV(x)|4,xI±.\tau_{\pm}^{4+\varepsilon_{1}}(x)\lesssim\left|\mathrm{Im}\,V(x)\right|,\qquad\left|\mathrm{Re}\,V(x)\right|^{3+\varepsilon_{1}}\tau_{\pm}^{4}(x)\lesssim\left|\mathrm{Im}\,V(x)\right|^{4},\qquad x\in I^{\pm}.

    By choosing η1(0,1)\eta_{1}\in(0,1) such that 4+ε1=41η14+\varepsilon_{1}=\frac{4}{1-\eta_{1}}, we will obtain, for all xI±x\in I^{\pm},

    {τ±(x)41η1|ImV(x)|,|ImV(x)||ImV(x)τ±(x)|43+η1,|ReV(x)|11η1|ImV(x)τ±(x)|43+η1.\left\{\begin{aligned} &\tau_{\pm}(x)^{\frac{4}{1-\eta_{1}}}\lesssim\left|\mathrm{Im}\,V(x)\right|,\\ &\left|\mathrm{Im}\,V(x)\right|\lesssim\left|\frac{\mathrm{Im}\,V(x)}{\tau_{\pm}(x)}\right|^{\frac{4}{3+\eta_{1}}},\\ &\left|\mathrm{Re}\,V(x)\right|^{\frac{1}{1-\eta_{1}}}\lesssim\left|\frac{\mathrm{Im}\,V(x)}{\tau_{\pm}(x)}\right|^{\frac{4}{3+\eta_{1}}}.\end{aligned}\right. (4.4)
  2. b)

    When VV is bounded at ±\pm\infty, from the assumption (2.6), by choosing η2(0,1)\eta_{2}\in(0,1) such that 13ε2=1η23+η2\frac{1}{3}-\varepsilon_{2}=\frac{1-\eta_{2}}{3+\eta_{2}}, we have

    τ±(x)|x|1η23+η2,xI±.\tau_{\pm}(x)\lesssim|x|^{\frac{1-\eta_{2}}{3+\eta_{2}}},\qquad x\in I^{\pm}. (4.5)

By using Lemma 4.1, we can define the boundary of the cut-off ξ\xi

δα±:=inf{x0:g±(x)=α}\delta_{\alpha}^{\pm}:=\inf\left\{x\geq 0:g_{\pm}(x)=\alpha\right\} (4.6)

through defining functions g±:[0,+)[0,+)g_{\pm}:[0,+\infty)\to[0,+\infty) as follows

g±(x):={|ImV(±x)τ±(±x)|43+η1 if V is unbounded at ±,|x|43+η2 if V is bounded at ±.\displaystyle g_{\pm}(x)=\left\{\begin{aligned} &\left|\frac{\mathrm{Im}\,V(\pm x)}{\tau_{\pm}(\pm x)}\right|^{\frac{4}{3+\eta_{1}}}\qquad&&\text{ if }V\text{ is unbounded at }\pm\infty,\\ &|x|^{\frac{4}{3+\eta_{2}}}\qquad&&\text{ if }V\text{ is bounded at }\pm\infty.\end{aligned}\right. (4.7)

In the latter case, since g±g_{\pm} is strictly increasing, we can work out precisely δα±\delta_{\alpha}^{\pm} whose formula is given at Theorem 2.1. In the first case, VV is continuous since VWlocn+3,()V\in W^{n+3,\infty}_{\textup{loc}}(\mathbb{R}), so are the functions g±g_{\pm}. Furthermore, thanks to τ±(0)>0\tau_{\pm}(0)>0, the last two conditions in (4.4), g±g_{\pm} is unbounded at ±\pm\infty and bounded at 0. In practice, it is not easy to get the exact solution δα±\delta_{\alpha}^{\pm} of the equation g±(x)=αg_{\pm}(x)=\alpha, instead we may approximate this solution by means of the symbol \approx introduced in the Notation 1.2 (see Examples 3 and 4).

Next, we define the remain ingredient of the cut-off ξ\xi in (4.2), that is Δα±\Delta_{\alpha}^{\pm}. When VV is unbounded at ±\pm\infty, with the aid of (2.3), there exist a constant κ>0\kappa>0 such that, for sufficiently large |x||x|,

ητ±(x)|x|ν±4.\frac{\eta}{\tau_{\pm}(x)}\leq\frac{|x|^{-\nu_{\pm}}}{4}. (4.8)

Let us define

Δα±={ητ±(δα±)if V is unbounded at ±,14δα±if V is bounded at ±.\Delta_{\alpha}^{\pm}=\left\{\begin{aligned} &\frac{\eta}{\tau_{\pm}(\delta_{\alpha}^{\pm})}\qquad&&\text{if }V\text{ is unbounded at }\pm\infty,\\ &\frac{1}{4}\delta_{\alpha}^{\pm}\qquad&&\text{if }V\text{ is bounded at }\pm\infty.\end{aligned}\right. (4.9)

Notice that since ν±1\nu_{\pm}\geq-1, from (4.8), it implies that Δα±δα±/4\Delta_{\alpha}^{\pm}\leq\delta_{\alpha}^{\pm}/4.

Proposition 4.2.

For all sufficiently large α>0\alpha>0, δα±\delta_{\alpha}^{\pm} are finite and limα+δα±=+\displaystyle\lim_{\alpha\to+\infty}\delta_{\alpha}^{\pm}=+\infty. Furthermore, the following hold for α1\alpha\gtrsim 1 and for all βB\beta\in B,

  1. 1)

    On JαJ_{\alpha},

    |ReV(x)|=o(α),|ImV(x)|α,|λV(x)|α,\displaystyle\left|\mathrm{Re}\,V(x)\right|=o(\alpha),\qquad\left|\mathrm{Im}\,V(x)\right|\lesssim\alpha,\qquad|\lambda-V(x)|\approx\alpha, (4.10)
    τ±(x){α1η14if V is unbounded at ±,α1η24if V is bounded at ±.\displaystyle\tau_{\pm}(x)\lesssim\left\{\begin{aligned} &\alpha^{\frac{1-\eta_{1}}{4}}\qquad&&\text{if }V\text{ is unbounded at }\pm\infty,\\ &\alpha^{\frac{1-\eta_{2}}{4}}\qquad&&\text{if }V\text{ is bounded at }\pm\infty.\end{aligned}\right. (4.11)
  2. 2)

    When VV is unbounded at ±\pm\infty, we have

    τ±(x)τ(x),ImV(x)ImV(±δα±),x:|xδα±|2Δα±.\tau_{\pm}(x)\approx\tau(x),\qquad\mathrm{Im}\,V(x)\approx\mathrm{Im}\,V(\pm\delta_{\alpha}^{\pm}),\qquad x\in\mathbb{R}:|x-\delta_{\alpha}^{\pm}|\leq 2\Delta_{\alpha}^{\pm}. (4.12)
Proof.

The statements related to the boundary δα±\delta_{\alpha}^{\pm} of the cut-off function ξ\xi is obtained from Lemma 4.1. We will give a proof for x0x\geq 0, the case x0x\leq 0 is analogous.

  1. 1)

    All the estimates in (4.10) and (4.11) will be claimed on I+Jα+I^{+}\cap J_{\alpha}^{+}. In order to have them on Jα+J_{\alpha}^{+}, the continuity of VV and τ+\tau_{+} on [0,a+][0,a_{+}] is employed.

    1. a)

      When VV is bounded at ++\infty, the claim on the real and imaginary part of VV in (4.10) is obvious. The last one in (4.10) is deduced from the triangle inequality

      |α||V(x)β||λV(x)||α|+|V(x)β|,|\alpha|-|V(x)-\beta|\leq\left|\lambda-V(x)\right|\leq|\alpha|+|V(x)-\beta|,

      and boundedness of BB on \mathbb{R} when α>0\alpha>0 is considered largely enough. In order to estimate τ+(x)\tau_{+}(x), we apply Lemma 4.1 with the definition of g+g_{+} in (4.7) and the aid of (4.5), for all xI+Jα+x\in I^{+}\cap J_{\alpha}^{+},

      τ+(x)41η2|x|43+η2α.\tau_{+}(x)^{\frac{4}{1-\eta_{2}}}\lesssim|x|^{\frac{4}{3+\eta_{2}}}\leq\alpha.
    2. b)

      When VV is unbounded at ++\infty, we apply again Lemma 4.1 with the definition of g+g_{+} in (4.7) and the help of (4.4), we have, for all xI+Jα+x\in I^{+}\cap J_{\alpha}^{+},

      |ImV(x)|g+(x)α,|ReV(x)|g+(x)1η1α1η1.\displaystyle\left|\mathrm{Im}\,V(x)\right|\lesssim g_{+}(x)\leq\alpha,\quad\left|\mathrm{Re}\,V(x)\right|\lesssim g_{+}(x)^{1-\eta_{1}}\leq\alpha^{1-\eta_{1}}.

      Then the estimate for λV\lambda-V in (4.10) is followed from

      |αReV(x)||λV(x)||αReV(x)|+|βImV(x)|.\left|\alpha-\mathrm{Re}\,V(x)\right|\leq\left|\lambda-V(x)\right|\leq\left|\alpha-\mathrm{Re}\,V(x)\right|+\left|\beta-\mathrm{Im}\,V(x)\right|.

      The above bound of τ+\tau_{+} in (4.11) is inferred from (4.4) and Lemma 4.1 that, for all xI+Jα+x\in I^{+}\cap J_{\alpha}^{+},

      τ+(x)41η1|ImV(x)|g+(x)α.\tau_{+}(x)^{\frac{4}{1-\eta_{1}}}\lesssim\left|\mathrm{Im}\,V(x)\right|\lesssim g_{+}(x)\leq\alpha.
  2. 2)

    First of all, let us show that, for sufficient large x>0x>0 and every |h|xν+2|h|\leq\frac{x^{-\nu_{+}}}{2}, we have

    τ(x+h)τ(x).\tau(x+h)\approx\tau(x).

    Indeed, from the assumption (2.3),

    |ln|τ(x+h)||τ(x)||=|xx+hτ(1)(t)τ(t)dt|\displaystyle\left|\ln\frac{|\tau(x+h)|}{|\tau(x)|}\right|=\left|\int_{x}^{x+h}\frac{\tau^{(1)}(t)}{\tau(t)}\,\mathrm{d}t\right| {xx+h|t|νdth0,x+hx|t|νdth0,\displaystyle\lesssim\left\{\begin{aligned} &\int_{x}^{x+h}|t|^{\nu}\,\mathrm{d}t\qquad&&h\geq 0,\\ &\int_{x+h}^{x}|t|^{\nu}\,\mathrm{d}t\qquad&&h\leq 0,\end{aligned}\right.
    {h(x+h)νh0,ν0,hxνh0,ν<0,(h)xνh0,ν0,(h)(x+h)νh0,ν<0,\displaystyle\leq\left\{\begin{aligned} &h(x+h)^{\nu}\qquad&&h\geq 0,\,\nu\geq 0,\\ &hx^{\nu}\qquad&&h\geq 0,\,\nu<0,\\ &(-h)x^{\nu}\qquad&&h\leq 0,\,\nu\geq 0,\\ &(-h)(x+h)^{\nu}\qquad&&h\leq 0,\,\nu<0,\\ \end{aligned}\right.
    1.\displaystyle\lesssim 1.

    In the last inequality, since ν+1\nu_{+}\geq-1, we used the observation that, for all |h|xν+2|h|\leq\frac{x^{-\nu_{+}}}{2} and for x>0x>0,

    x2x+h3x2.\frac{x}{2}\leq x+h\leq\frac{3x}{2}.

    Then, the first estimate for τ+(x)\tau_{+}(x) is deduced by replacing the above xx by δα+\delta_{\alpha}^{+} with the notice that 2Δα+(δα+)ν+2\displaystyle 2\Delta_{\alpha}^{+}\leq\frac{\left(\delta_{\alpha}^{+}\right)^{-\nu_{+}}}{2}, thanks to (4.8) and the definition of Δα+\Delta_{\alpha}^{+}. We employ this idea for the function ImV\mathrm{Im}\,V as following: for large δα+\delta_{\alpha}^{+} and for all |h|2Δα+|h|\leq 2\Delta_{\alpha}^{+}, we have

    |ln|ImV(δα++h)||ImV(δα+)||=|δα+δα++hImV(1)(t)ImV(t)dt|\displaystyle\left|\ln\frac{|\mathrm{Im}\,V(\delta_{\alpha}^{+}+h)|}{|\mathrm{Im}\,V(\delta_{\alpha}^{+})|}\right|=\left|\int_{\delta_{\alpha}^{+}}^{\delta_{\alpha}^{+}+h}\frac{\mathrm{Im}\,V^{(1)}(t)}{\mathrm{Im}\,V(t)}\,\mathrm{d}t\right| {δα+δα++hτ+(t)dtif h0,δα++hδα+τ+(t)dtif h0.\displaystyle\leq\left\{\begin{aligned} &\int_{\delta_{\alpha}^{+}}^{\delta_{\alpha}^{+}+h}\tau_{+}(t)\,\mathrm{d}t&&\text{if }h\geq 0,\\ &\int_{\delta_{\alpha}^{+}+h}^{\delta_{\alpha}^{+}}\tau_{+}(t)\,\mathrm{d}t&&\text{if }h\leq 0.\end{aligned}\right.
    τ+(δα+)Δα+=η.\displaystyle\lesssim\tau_{+}(\delta_{\alpha}^{+})\Delta_{\alpha}^{+}=\eta.

    Here, in the second inequality, we have used the estimate for τ+\tau_{+} in (4.12). Thus, the estimate for ImV\mathrm{Im}\,V in (4.12) is followed.

4.2. Pseudomode estimate

Let Assumption I hold for some N0N\in\mathbb{N}_{0}. Let ψ1(1)\psi_{-1}^{(1)} be determined by (3.6) with the plus sign and (ψk(1))k[[0,N1]]\left(\psi_{k}^{(1)}\right)_{k\in[[0,N-1]]} be determined by (3.7), in order to obtain the primitive functions (ψk)k[[1,N1]]\left(\psi_{k}\right)_{k\in[[-1,N-1]]}, we fix the initial data for them

ψk(0):=0,k[[1,N1]].\psi_{k}(0):=0,\qquad\forall k\in[[-1,N-1]].

Let us define the pseudomode for Theorem 2.1 as follows

Ψλ,N:=ξλexp(Pλ,N),\Psi_{\lambda,N}:=\xi_{\lambda}\exp(-P_{\lambda,N}),

where

  • ξλ\xi_{\lambda} is the cut-off defined in (4.2) with δα±\delta_{\alpha}^{\pm} and Δα±\Delta_{\alpha}^{\pm} as in (4.6) and (4.9),

  • Pλ,N=k=1N1λkψk(t)\displaystyle P_{\lambda,N}=\sum_{k=-1}^{N-1}\lambda^{-k}\psi_{k}(t) defined as in (3.3).

With the intention of estimating the pseudomode Ψλ,N\Psi_{\lambda,N} later, we firstly provide some estimates for the functions (ψk(1))k[[1,N1]]\left(\psi_{k}^{(1)}\right)_{k\in[[-1,N-1]]} in the following lemma.

Lemma 4.3.

For α1\alpha\gtrsim 1 and for all βB\beta\in B, we have

Re(λψ1(1)(x))ImV(x)βα34\displaystyle\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(x)\right)\approx\frac{\mathrm{Im}\,V(x)-\beta}{\alpha^{\frac{3}{4}}}\qquad on I±Jα±,\displaystyle\text{on }I^{\pm}\cap J_{\alpha}^{\pm}, (4.13)
|Re(λψ1(1)(x))|1α34\displaystyle\left|\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(x)\right)\right|\lesssim\frac{1}{\alpha^{\frac{3}{4}}}\qquad on [a,a+],\displaystyle\text{on }[-a_{-},a_{+}],

and for all k[[1,N1]]k\in[[-1,N-1]], for all m[[1,3]]m\in[[1,3]] such that k+m1k+m\geq 1, we have

|λkψk(m)(x)||V(x)|τ±(x)k+mαk4+1\displaystyle\left|\lambda^{-k}\psi_{k}^{(m)}(x)\right|\lesssim\frac{|V(x)|\tau_{\pm}(x)^{k+m}}{\alpha^{\frac{k}{4}+1}}\qquad on I±Jα±,\displaystyle\text{ on }I^{\pm}\cap J_{\alpha}^{\pm}, (4.14)
|λkψk(m)(x)|1αk4+1\displaystyle\left|\lambda^{-k}\psi_{k}^{(m)}(x)\right|\lesssim\frac{1}{\alpha^{\frac{k}{4}+1}}\qquad on [a,a+].\displaystyle\text{ on }[-a_{-},a_{+}].
Proof.

From the formula of the eikonal solution (3.6) and the principal forth root given by (3.9), we have

Re(λψ1(1))=12ImVβ(|Vλ|+αReV)1/2(|Vλ|1/2+12(|Vλ|+αReV)1/2)1/2.\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}\right)=\frac{1}{2}\frac{\mathrm{Im}\,V-\beta}{\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\left(|V_{\lambda}|^{1/2}+\frac{1}{\sqrt{2}}\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\right)^{1/2}}. (4.15)

Thanks to the application of (4.10) for the denominator of Re(λψ1(1))\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}\right), the first estimate in (4.13) is obtained directly by the fixed sign of ImVβ\mathrm{Im}\,V-\beta on each II^{-} and I+I^{+} (see (4.1)), while the second one in (4.13) is attained from the continuity of ImV\mathrm{Im}\,V on [a,a+][-a_{-},a_{+}] and the boundedness of BB.

For each k[[1,N1]]k\in[[-1,N-1]] and for each m[[1,3]]m\in[[1,3]] such that k+m1k+m\geq 1, Lemma 3.2 and the assumption (2.2) combining with the characteristic (3.10) of the set Ir,jI_{r,j} yield that, on I+Jα+I^{+}\cap J_{\alpha}^{+},

|λkψk(m)|\displaystyle\left|\lambda^{-k}\psi_{k}^{(m)}\right| 1|Vλ|k4j=1k+m|dk+m,j||Vλ|j\displaystyle\leq\frac{1}{\left|V_{\lambda}\right|^{\frac{k}{4}}}\sum_{j=1}^{k+m}\frac{|d_{k+m,j}|}{|V_{\lambda}|^{j}}
1|Vλ|k4j=1k+mαIk+m,j|V(1)|α1|V(2)|α2|V(k+m+1j)|αk+m+1j|Vλ|j\displaystyle\lesssim\frac{1}{|V_{\lambda}|^{\frac{k}{4}}}\sum_{j=1}^{k+m}\frac{\displaystyle\sum_{\alpha\in I_{k+m,j}}|V^{(1)}|^{\alpha_{1}}|V^{(2)}|^{\alpha_{2}}\ldots|V^{(k+m+1-j)}|^{\alpha_{k+m+1-j}}}{|V_{\lambda}|^{j}}
τ+k+m|Vλ|k4j=1k+m|V|j|Vλ|j.\displaystyle\lesssim\frac{\tau_{+}^{k+m}}{|V_{\lambda}|^{\frac{k}{4}}}\sum_{j=1}^{k+m}\frac{|V|^{j}}{|V_{\lambda}|^{j}}.

From (4.10), it implies that |V|α|Vλ||V|\lesssim\alpha\lesssim|V_{\lambda}| on JαJ_{\alpha}. Then the first estimate in (4.14) for x>0x>0 is obtained:

|λkψk(m)|τ+k+m|V||Vλ|k4+1j=0k+m1|V|j|Vλ|jτ+k+m|V|αk4+1.\left|\lambda^{-k}\psi_{k}^{(m)}\right|\lesssim\frac{\tau_{+}^{k+m}|V|}{|V_{\lambda}|^{\frac{k}{4}+1}}\sum_{j=0}^{k+m-1}\frac{|V|^{j}}{|V_{\lambda}|^{j}}\lesssim\frac{\tau_{+}^{k+m}|V|}{\alpha^{\frac{k}{4}+1}}.

On IJαI^{-}\cap J_{\alpha}^{-}, all the above estimates are analogous. For k[[1,N1]]k\in[[-1,N-1]] and m[[1,3]]m\in[[1,3]], we observe that the maximal derivatives of VV appearing in the expression of λkψk(m)\lambda^{-k}\psi_{k}^{(m)} is k+m+11=k+mk+m+1-1=k+m that is at most N+2N+2. Since VWlocN+3,()V\in W_{\textup{loc}}^{N+3,\infty}(\mathbb{R}), all the derivatives of VV in λkψk(m)\lambda^{-k}\psi_{k}^{(m)} are continuous. The second one in (4.14) follows immediately from the boundedness of the derivatives of VV on [a,a+][-a_{-},a_{+}]. ∎

Remark 4.4 (The assumption on ReV\mathrm{Re}\,V).

Let us explain why we set up the condition (2.5) on ReV\mathrm{Re}\,V. We consider the polynomial potential

V(x)=xρ+isign(x)|x|γ with ρ>0,γ>0.V(x)=-x^{\rho}+i\ \textup{sign}(x)|x|^{\gamma}\qquad\text{ with }\rho>0,\,\gamma>0.

It is obvious that the assumption (2.1) is fulfilled. In this case, we can choose τ±(x)=(x2+1)12\tau_{\pm}(x)=(x^{2}+1)^{-\frac{1}{2}} to satisfy assumptions (2.2) and (2.4) and we have τ±(x)|x|1\tau_{\pm}(x)\approx|x|^{-1} for |x|1|x|\gtrsim 1 then the condition (2.3) is satisfied with ν±=1\nu_{\pm}=-1. The assumption (2.5) reads: there exists ε1>0\varepsilon_{1}>0 such that

3ρ4+ρε14γ.3\rho-4+\rho\varepsilon_{1}\leq 4\gamma.

We assume on the contrary that

3ρ4>4γ.3\rho-4>4\gamma.

By change of variable t=α1ρst=\alpha^{\frac{1}{\rho}}s, it yields that, for all x>0x>0,

0x|Re(λψ1(1)(t))|dt0x|tγ|+|β|(α+tρ)34dtα4γ3ρ+44ρ0+sγ+|β|αγ(1+sρ)34dt=o(1),α+.\int_{0}^{x}\left|\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\right|\,\mathrm{d}t\lesssim\int_{0}^{x}\frac{|t^{\gamma}|+|\beta|}{\left(\alpha+t^{\rho}\right)^{\frac{3}{4}}}\,\mathrm{d}t\lesssim\alpha^{\frac{4\gamma-3\rho+4}{4\rho}}\int_{0}^{+\infty}\frac{s^{\gamma}+\frac{|\beta|}{\alpha^{\gamma}}}{\left(1+s^{\rho}\right)^{\frac{3}{4}}}\,\mathrm{d}t=o(1),\qquad\alpha\to+\infty.

It means that the dominant part 0xλRe(λψ1(1)(t))dt\int_{0}^{x}\lambda\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\,\mathrm{d}t in the expansion of Pλ,NP_{\lambda,N} (as we see later in the next proposition) is bounded (uniformly in λ\lambda) on R+R_{+} and this will spoil completely the decay of our pseudomode.

Proposition 4.5.

There exists c>0c>0 such that, for α1\alpha\gtrsim 1 and for all βB\beta\in B,

(𝒟λ,Nξ)exp(Pλ,N)ξexp(Pλ,N)exp(cαη4),\frac{\|(\mathcal{D}_{\lambda,N}\xi)\exp(-P_{\lambda,N})\|}{\|\xi\exp(-P_{\lambda,N})\|}\lesssim\exp\left(-c\alpha^{\frac{\eta}{4}}\right), (4.16)

where

  • 𝒟λ,N\mathcal{D}_{\lambda,N} is the differential representation DλD_{\lambda} after replacing PP as Pλ,NP_{\lambda,N} in (3.1),

  • η:={η1if V is unbounded at ±,η2if V is bounded at ±.\eta:=\left\{\begin{aligned} &\eta_{1}\qquad&&\text{if }V\text{ is unbounded at }\pm\infty,\\ &\eta_{2}\qquad&&\text{if }V\text{ is bounded at }\pm\infty.\end{aligned}\right.

Proof.

The idea of the proof is as follows.

  1. (1)

    We will deal with the denominator of (4.16) first by showing that it is bounded below by a constant which is independent of α\alpha.

  2. (2)

    In order to handle the numerator of (4.16), we will show that the eikonal term λψ1(1)\lambda\psi_{-1}^{(1)} dominates over the other terms λkψk(1)\lambda^{-k}\psi_{k}^{(1)} with k0k\geq 0 and thus there exists constants C±>0C_{\pm}>0 such that, for all xI±Jα±x\in I^{\pm}\cap J_{\alpha}^{\pm},

    |exp(Pλ,N(x))|exp(C±α3/4±a±xImV(t)dt).\left|\exp(-P_{\lambda,N}(x))\right|\lesssim\exp\left(-\frac{C_{\pm}}{\alpha^{3/4}}\int_{\pm a_{\pm}}^{x}\mathrm{Im}\,V(t)\,\mathrm{d}t\right). (4.17)
  3. (3)

    From (4.17), we show that there exists a constant c1>0c_{1}>0 such that, for all x[δα,δα+Δα]x\in\left[-\delta_{\alpha}^{-},-\delta_{\alpha}^{-}+\Delta_{\alpha}^{-}\right] and for all x[δα+Δα+,δα+]x\in\left[\delta_{\alpha}^{+}-\Delta_{\alpha}^{+},\delta_{\alpha}^{+}\right],

    |exp(Pλ,N(x))|exp(c1αη4).\left|\exp\left(-P_{\lambda,N}(x)\right)\right|\lesssim\exp\left(-c_{1}\alpha^{\frac{\eta}{4}}\right). (4.18)
  4. (4)

    Finally, we use (4.18) to control Dλ,NξD_{\lambda,N}\xi in (4.16).

The details are as follows.

  1. (1)

    Let us recall that

    Pλ,N(x)=k=1N10xλkψk(1)(t)dt.P_{\lambda,N}(x)=\sum_{k=-1}^{N-1}\int_{0}^{x}\lambda^{-k}\psi_{k}^{(1)}(t)\,\mathrm{d}t.

    Thanks to the estimates (4.13) and (4.14), we have the bound of (4.16),

    |RePλ,N(x)|1α34,x[0,a+].\left|\mathrm{Re}\,P_{\lambda,N}(x)\right|\lesssim\frac{1}{\alpha^{\frac{3}{4}}},\qquad x\in[0,a_{+}]. (4.19)

    Then, there is a constant C1>0C_{1}>0 such that

    |ξexp(Pλ,n(x))|2dx0a+exp(2RePλ,N(x))dxa+exp(C1α3/4).\displaystyle\int_{\mathbb{R}}|\xi\exp(-P_{\lambda,n}(x))|^{2}\,\mathrm{d}x\geq\int_{0}^{a^{+}}\exp\left(-2\mathrm{Re}\,P_{\lambda,N}(x)\right)\,\mathrm{d}x\geq a_{+}\exp\left(-\frac{C_{1}}{\alpha^{3/4}}\right).

    Therefore, by considering large α>0\alpha>0, we get

    ξexp(Pλ,N)1.\left\|\xi\exp\left(-P_{\lambda,N}\right)\right\|\gtrsim 1.
  2. (2)

    Next, we will prove (4.17). On I+Jα+I^{+}\cap J^{+}_{\alpha}, thanks to (4.14) and (4.11), we have

    |k=0N1Re(λkψk(1)(t))|\displaystyle\left|\sum_{k=0}^{N-1}\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\right| k=0N1|V(t)|τ+(t)k+1αk4+1=|V(t)|τ+(t)αk=0N1(τ+(t)α14)k\displaystyle\lesssim\sum_{k=0}^{N-1}\frac{|V(t)|\tau_{+}(t)^{k+1}}{\alpha^{\frac{k}{4}+1}}=\frac{|V(t)|\tau_{+}(t)}{\alpha}\sum_{k=0}^{N-1}\left(\frac{\tau_{+}(t)}{\alpha^{\frac{1}{4}}}\right)^{k}
    |V(t)|τ+(t)α.\displaystyle\lesssim\frac{|V(t)|\tau_{+}(t)}{\alpha}.

    Combining this with (4.13) and (4.1), it implies that

    |k=0N1Re(λkψk(1)(t))|Re(λψ1(1)(t))|V(t)|τ+(t)ImV(t)α14{αη14if V is unbounded at +,αη24if V is bounded at +.\displaystyle\frac{\displaystyle\left|\sum_{k=0}^{N-1}\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\right|}{\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)}\lesssim\frac{|V(t)|\tau_{+}(t)}{\mathrm{Im}\,V(t)\alpha^{\frac{1}{4}}}\lesssim\left\{\begin{aligned} &\alpha^{-\frac{\eta_{1}}{4}}\qquad\text{if }V\text{ is unbounded at }+\infty,\\ &\alpha^{-\frac{\eta_{2}}{4}}\qquad\text{if }V\text{ is bounded at }+\infty.\end{aligned}\right.

    Indeed, we consider that cases:

    1. i)

      If VV is bounded at ++\infty, the above estimate comes from (4.1) and (4.11).

    2. ii)

      If VV is unbounded at ++\infty, the triangular inequality leads to

      |V(t)|τ+(t)ImV(t)α14|ReV(t)|τ+(t)ImV(t)α14+τ+(t)α14,\displaystyle\frac{|V(t)|\tau_{+}(t)}{\mathrm{Im}\,V(t)\alpha^{\frac{1}{4}}}\leq\frac{|\mathrm{Re}\,V(t)|\tau_{+}(t)}{\mathrm{Im}\,V(t)\alpha^{\frac{1}{4}}}+\frac{\tau_{+}(t)}{\alpha^{\frac{1}{4}}},

      then (4.4), the definition of g+g_{+} in (4.7) and Lemma 4.1 allow us to control the term with ReV\mathrm{Re}\,V as follows

      |ReV(t)|τ+(t)ImV(t)α14g+(t)1η1τ+(t)ImV(t)α14=g+(t)15η14α14α5η14.\displaystyle\frac{|\mathrm{Re}\,V(t)|\tau_{+}(t)}{\mathrm{Im}\,V(t)\alpha^{\frac{1}{4}}}\lesssim\frac{g_{+}(t)^{1-\eta_{1}}\tau_{+}(t)}{\mathrm{Im}\,V(t)\alpha^{\frac{1}{4}}}=\frac{g_{+}(t)^{\frac{1-5\eta_{1}}{4}}}{\alpha^{\frac{1}{4}}}\leq\alpha^{-\frac{5\eta_{1}}{4}}.

      In the last inequality, we assumed that ε1\varepsilon_{1} small enough in Assumption (2.5). Hence using (4.11) to control the term τ+(t)α14\frac{\tau_{+}(t)}{\alpha^{\frac{1}{4}}}, we have a conclusion.

    Therefore, (4.17) is obtained by employing (4.19) , (4.13) and (4.1). In detail, for all xI+Jα+x\in I^{+}\cap J_{\alpha}^{+}, we have (with some constant C+C_{+})

    |exp(Pλ,N(x))|\displaystyle\left|\exp\left(-P_{\lambda,N}(x)\right)\right| =exp(RePλ,N(a+)a+xk=1N1Re(λkψk(1)(t))dt)\displaystyle=\exp\left(-\mathrm{Re}\,P_{\lambda,N}(a_{+})-\int_{a_{+}}^{x}\sum_{k=-1}^{N-1}\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\,\mathrm{d}t\right)
    exp(a+x(1o(1))Re(λψ1(1))dt)\displaystyle\lesssim\exp\left(-\int_{a_{+}}^{x}(1-o(1))\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}\right)\,\mathrm{d}t\right)
    exp(C+α3/4a+xImV(t)dt).\displaystyle\lesssim\exp\left(-\frac{C_{+}}{\alpha^{3/4}}\int_{a_{+}}^{x}\mathrm{Im}\,V(t)\,\mathrm{d}t\right).

    The proof for xIJαx\in I^{-}\cap J_{\alpha}^{-} is the same.

  3. (3)

    In order to prove (4.18), we consider two cases:

    1. i)

      If VV is unbounded at ++\infty, from the property (4.12) and the definitions of δα+\delta_{\alpha}^{+} in (4.6) and g+(δα+)g_{+}(\delta_{\alpha}^{+}) in (4.7), we have that, for x[δα+Δα+,δα+]x\in\left[\delta_{\alpha}^{+}-\Delta_{\alpha}^{+},\delta_{\alpha}^{+}\right],

      a+xImV(t)dt\displaystyle\int_{a_{+}}^{x}\mathrm{Im}\,V(t)\,\mathrm{d}t δα+2Δα+xImV(t)dt(Δα+)ImV(δα+)\displaystyle\geq\int_{\delta_{\alpha}^{+}-2\Delta_{\alpha}^{+}}^{x}\mathrm{Im}\,V(t)\,\mathrm{d}t\gtrsim\left(\Delta_{\alpha}^{+}\right)\mathrm{Im}\,V(\delta_{\alpha}^{+})
      ImV(δα+)τ(δα+)=α3+η14.\displaystyle\gtrsim\frac{\mathrm{Im}\,V(\delta_{\alpha}^{+})}{\tau(\delta_{\alpha}^{+})}=\alpha^{\frac{3+\eta_{1}}{4}}.
    2. ii)

      If VV is bounded at ++\infty, we use the assumption (4.1) to obtain,

      a+xImV(t)dtδα+2Δα+xImV(t)dtΔα+=α3+η24.\displaystyle\int_{a_{+}}^{x}\mathrm{Im}\,V(t)\,\mathrm{d}t\gtrsim\int_{\delta_{\alpha}^{+}-2\Delta_{\alpha}^{+}}^{x}\mathrm{Im}\,V(t)\,\mathrm{d}t\gtrsim\Delta_{\alpha}^{+}=\alpha^{\frac{3+\eta_{2}}{4}}.

    Then, (4.18) is followed directly from (4.17).

  4. (4)

    In order to control the terms attached with ξ()\xi^{(\ell)} for [[1,4]]\ell\in[[1,4]], we notice that, for m[[1,3]]m\in[[1,3]]

    |Pλ,N(m)(x)|k=1N1|λkψk(m)(x)|k=1N1τ±(x)k+mαk4,xI±Jα±.\displaystyle\left|P_{\lambda,N}^{(m)}(x)\right|\leq\sum_{k=-1}^{N-1}\left|\lambda^{-k}\psi_{k}^{(m)}(x)\right|\lesssim\sum_{k=-1}^{N-1}\frac{\tau_{\pm}(x)^{k+m}}{\alpha^{\frac{k}{4}}},\qquad x\in I^{\pm}\cap J_{\alpha}^{\pm}.

    Here we employed (4.14) and (3.6). Thanks to the upper bound of τ±\tau_{\pm} by some power of α\alpha in (4.11), we can bound Pλ,n(m)P_{\lambda,n}^{(m)} by a polynomial of α\alpha and thus they are also rapidly decaying when they are attached with exp(c1αη4)\exp(-c_{1}\alpha^{\frac{\eta}{4}}). For example, we give a detail on how to deal with the terms attached with ξ(4)\xi^{(4)} and ξ(3)\xi^{(3)}, the other terms are estimated similarly:

    1. a)

      The attached with ξ(4)\xi^{(4)}, by employing (4.3) and (4.18), we have

      δαδα+Δα|ξ(4)(x)|2|exp(Pλ,N(x))|2dx+δα+Δα+δα+|ξ(4)(x)|2|exp(Pλ,N(x))|2dx\displaystyle\int_{-\delta_{\alpha}^{-}}^{-\delta_{\alpha}^{-}+\Delta_{\alpha}^{-}}|\xi^{(4)}(x)|^{2}\left|\exp\left(-P_{\lambda,N}(x)\right)\right|^{2}\,\mathrm{d}x+\int_{\delta_{\alpha}^{+}-\Delta_{\alpha}^{+}}^{\delta_{\alpha}^{+}}|\xi^{(4)}(x)|^{2}\left|\exp\left(-P_{\lambda,N}(x)\right)\right|^{2}\,\mathrm{d}x
      (Δα)7exp(2c1αη4)+(Δα+)7exp(2c1αη4).\displaystyle\lesssim\left(\Delta_{\alpha}^{-}\right)^{-7}\exp\left(-2c_{1}\alpha^{\frac{\eta}{4}}\right)+\left(\Delta_{\alpha}^{+}\right)^{-7}\exp\left(-2c_{1}\alpha^{\frac{\eta}{4}}\right).

      By using (4.11) for Δα±\Delta_{\alpha}^{\pm} defined in (4.9) when VV is unbounded at ±\pm\infty and δα±=α3+η242\delta_{\alpha}^{\pm}=\frac{\alpha^{\frac{3+\eta_{2}}{4}}}{2} when VV is bounded at ±\pm\infty, it implies that

      (Δα±)1{α1η14if V is unbounded at ±,α3+η24if V is bounded at ±.\left(\Delta_{\alpha}^{\pm}\right)^{-1}\lesssim\left\{\begin{aligned} &\alpha^{\frac{1-\eta_{1}}{4}}&&\text{if }V\text{ is unbounded at }\pm\infty,\\ &\alpha^{-\frac{3+\eta_{2}}{4}}&&\text{if }V\text{ is bounded at }\pm\infty.\end{aligned}\right.

      Therefore, there exists c2>0c_{2}>0 such that

      ξ(4)exp(Pλ,N(x))exp(c2αη4).\left\|\xi^{(4)}\exp\left(-P_{\lambda,N}(x)\right)\right\|\lesssim\exp\left(-c_{2}\alpha^{\frac{\eta}{4}}\right).
    2. b)

      The attached with ξ(3)\xi^{(3)}, we have

      |ξ(3)(x)Pλ,N(1)exp(Pλ,N(x))|\displaystyle\left|\xi^{(3)}(x)P_{\lambda,N}^{(1)}\exp\left(-P_{\lambda,N}(x)\right)\right| (Δα±)3(k=1N1τ+(x)k+1αk4)exp(c1αη4)\displaystyle\lesssim\left(\Delta_{\alpha}^{\pm}\right)^{-3}\left(\sum_{k=-1}^{N-1}\frac{\tau_{+}(x)^{k+1}}{\alpha^{\frac{k}{4}}}\right)\exp\left(-c_{1}\alpha^{\frac{\eta}{4}}\right)
      (Δα±)3(k=1N1α(k+1)(1η)k4)exp(c1αη4)\displaystyle\lesssim\left(\Delta_{\alpha}^{\pm}\right)^{-3}\left(\sum_{k=-1}^{N-1}\alpha^{\frac{(k+1)(1-\eta)-k}{4}}\right)\exp(-c_{1}\alpha^{\frac{\eta}{4}})
      exp(c3αη4).\displaystyle\lesssim\exp\left(-c_{3}\alpha^{\frac{\eta}{4}}\right).

    Consequently, we put everything together, we obtain (4.16).

4.3. Remainder estimate (Proof of Theorem 2.1)

Obviously, Ψλ,N\Psi_{\lambda,N} belongs to the domain of V\mathscr{L}_{V} because of its support. By the estimate (3.2) and Proposition 4.5, we have

(Vλ)Ψλ,NΨλ,Nexp(cαη4)+λ,NL(Jα).\displaystyle\frac{\|(\mathscr{L}_{V}-\lambda)\Psi_{\lambda,N}\|}{\|\Psi_{\lambda,N}\|}\lesssim\exp(-c\alpha^{\frac{\eta}{4}})+\|\mathcal{R}_{\lambda,N}\|_{L^{\infty}(J_{\alpha})}.

Let n=Nn=N in Lemma 3.3, estimate as in the proof of Lemma 4.3 and employ (4.11), it yields that, for N1N\geq 1,

|λ,N(x)|\displaystyle\left|\mathcal{R}_{\lambda,N}(x)\right| k=03N11|Vλ|k+N34j=1k+N+1|dk+N+1,j||Vλ|jk=03N1τ±(x)k+N+1|Vλ|k+N34j=1k+N+1|V(x)|j|Vλ|j\displaystyle\lesssim\sum_{k=0}^{3N-1}\frac{1}{|V_{\lambda}|^{\frac{k+N-3}{4}}}\sum_{j=1}^{k+N+1}\frac{|d_{k+N+1,j}|}{|V_{\lambda}|^{j}}\lesssim\sum_{k=0}^{3N-1}\frac{\tau_{\pm}(x)^{k+N+1}}{|V_{\lambda}|^{\frac{k+N-3}{4}}}\sum_{j=1}^{k+N+1}\frac{|V(x)|^{j}}{|V_{\lambda}|^{j}}
k=03N1τ±(x)k+N+1|V(x)||Vλ|k+N+14|V(x)|τ±(x)N+1αN+14,xI±Jλ±.\displaystyle\lesssim\sum_{k=0}^{3N-1}\frac{\tau_{\pm}(x)^{k+N+1}|V(x)|}{|V_{\lambda}|^{\frac{k+N+1}{4}}}\lesssim\frac{|V(x)|\tau_{\pm}(x)^{N+1}}{\alpha^{\frac{N+1}{4}}},\qquad x\in I^{\pm}\cap J_{\lambda}^{\pm}.

Similarly, since VWlocN+3,()V\in W^{N+3,\infty}_{\textup{loc}}(\mathbb{R}), we have the following estimate on the compact set [a,a+][-a_{-},a_{+}]:

|λ,N(x)|αN+14,x[a,a+].\displaystyle\left|\mathcal{R}_{\lambda,N}(x)\right|\lesssim\alpha^{-\frac{N+1}{4}},\qquad x\in[-a_{-},a_{+}].

Likewise, with the same reason, we have the same estimate for N=0N=0. Thus, the estimate (2.8) is followed for all N0N\geq 0.

Remark 4.6.

From the above construction, we see that if we change the sign of ImV\mathrm{Im}\,V in condition (2.1) as follows

lim supx+ImV(x)<0<lim infxImV(x);\limsup_{x\to+\infty}\mathrm{Im}\,V(x)<0<\liminf_{x\to-\infty}\mathrm{Im}\,V(x); (4.20)

then the previous analysis still works. Indeed, what we need to do is just to choose the minus sign in the formula of ψ1(1)\psi_{-1}^{(1)} in (3.6). Then, we have

Re(λψ1(1))=12βImV(|Vλ|+αReV)1/2(|Vλ|1/2+12(|Vλ|+αReV)1/2)1/2.\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}\right)=\frac{1}{2}\frac{\beta-\mathrm{Im}\,V}{\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\left(|V_{\lambda}|^{1/2}+\frac{1}{\sqrt{2}}\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\right)^{1/2}}.

By fixing β±0+\beta_{\pm}\in\mathbb{R}_{0}^{+} such that

lim supx+ImV(x)<β and β+<lim infxImV(x),\limsup_{x\to+\infty}\mathrm{Im}\,V(x)<-\beta_{-}\text{ and }\beta_{+}<\liminf_{x\to-\infty}\mathrm{Im}\,V(x),

then there exist constants a±>0a_{\pm}>0 such that, for all βB=[β,β+]\beta\in B=[-\beta_{-},\beta_{+}],

ImV(x)βImV(x)1,\displaystyle\mathrm{Im}\,V(x)-\beta\lesssim\mathrm{Im}\,V(x)\lesssim-1,\qquad xI+{x0+:xa+},\displaystyle x\in I^{+}\coloneqq\left\{x\in\mathbb{R}_{0}^{+}:x\geq a_{+}\right\},
ImV(x)βImV(x)1,\displaystyle\mathrm{Im}\,V(x)-\beta\gtrsim\mathrm{Im}\,V(x)\gtrsim 1,\qquad xI{x0:xa}.\displaystyle x\in I^{-}\coloneqq\left\{x\in\mathbb{R}_{0}^{-}:x\leq-a_{-}\right\}.

By repeating the procedure when proving (4.13), we have

Re(λψ1(1)(x))βImV(x)α34ImV(x)α34\displaystyle\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(x)\right)\approx\frac{\beta-\mathrm{Im}\,V(x)}{\alpha^{\frac{3}{4}}}\gtrsim-\frac{\mathrm{Im}\,V(x)}{\alpha^{\frac{3}{4}}}\qquad on I+Jα+,\displaystyle\text{ on }I^{+}\cap J_{\alpha}^{+}, (4.21)
Re(λψ1(1)(x))βImV(x)α34ImV(x)α34\displaystyle\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(x)\right)\approx\frac{\beta-\mathrm{Im}\,V(x)}{\alpha^{\frac{3}{4}}}\lesssim-\frac{\mathrm{Im}\,V(x)}{\alpha^{\frac{3}{4}}}\qquad on IJα.\displaystyle\text{ on }I^{-}\cap J_{\alpha}^{-}.

Therefore, the pseudomode now possesses the right sign for the decay. Although all the other terms (ψk(1))0kN1\left(\psi_{k}^{(1)}\right)_{0\leq k\leq N-1} also change their signs accordingly, but it does not matter because they are all estimated with the absolute value.

4.4. Decaying potentials

We reserve this section for constructing the pseudomodes for the potentials in Example 5. Since the condition 2.1 is not met by the decaying of the potentials, we can not apply directly the previous constructions. However, the shape of the pseudomodes is the same as in the beginning of Subsection 4.2, just the definition of δλ±\delta_{\lambda}^{\pm} (replace for δα±\delta_{\alpha}^{\pm}) should be defined differently. In the coming paragraphs, when we say λ=α+iβ\lambda=\alpha+i\beta\to\infty, we mean α+\alpha\to+\infty and |β|0|\beta|\to 0 (β\beta can be zero). Let a+>0a_{+}>0 such that

ImV(x)=|x|γ,xa+.\displaystyle\mathrm{Im}\,V(x)=|x|^{-\gamma},\qquad x\geq a_{+}.

We seek for the boundary δλ+\delta_{\lambda}^{+} of the cut-off such that the first term in the expansion very large when λ\lambda\to\infty. Since VV is bounded, we still have |λV(x)|α|\lambda-V(x)|\approx\alpha, thus, for x>a+x>a_{+},

a+xRe(λψ1(1)(t))dt1α34a+x[ImV(t)β]dt=x1γ[11γβxγ]a+1γ1γ+βa+α34.\int_{a_{+}}^{x}\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\,\mathrm{d}t\gtrsim\frac{1}{\alpha^{\frac{3}{4}}}\int_{a_{+}}^{x}\left[\mathrm{Im}\,V(t)-\beta\right]\,\mathrm{d}t=\frac{x^{1-\gamma}\left[\frac{1}{1-\gamma}-\beta x^{\gamma}\right]-\frac{a_{+}^{1-\gamma}}{1-\gamma}+\beta a_{+}}{\alpha^{\frac{3}{4}}}.

Here, in order that the inequality happens, we fixed the sign ImVβ>0\mathrm{Im}\,V-\beta>0 on [a+,δλ+][a_{+},\delta_{\lambda}^{+}] by assuming that |β|(δλ+)γ=o(1)|\beta|\left(\delta_{\lambda}^{+}\right)^{\gamma}=o(1) as λ\lambda\to\infty. Combining this assumption for δλ+\delta_{\lambda}^{+} with the expect that the right hand side of the above estimate very large, δλ±\delta_{\lambda}^{\pm} should read

α34(δλ+)γ1+|β|(δλ+)γ=o(1),λ.\alpha^{\frac{3}{4}}\left(\delta_{\lambda}^{+}\right)^{\gamma-1}+|\beta|\left(\delta_{\lambda}^{+}\right)^{\gamma}=o(1),\qquad\lambda\to\infty. (4.22)

The existence of finite positive number δλ+\delta_{\lambda}^{+} satisfying limλδλ+=+\displaystyle\lim_{\lambda\to\infty}\delta_{\lambda}^{+}=+\infty and (4.22) is equivalent to the constraint (2.21) on α\alpha and β\beta. Indeed, this is due to the following inequality for all s>0s>0,

α34sγ1+|β|sγcγα34γ|β|1γ,cγ=1γγ(1γ)1γ,\alpha^{\frac{3}{4}}s^{\gamma-1}+|\beta|s^{\gamma}\geq c_{\gamma}\alpha^{\frac{3}{4}\gamma}|\beta|^{1-\gamma},\qquad c_{\gamma}=\frac{1}{\gamma^{\gamma}(1-\gamma)^{1-\gamma}},

and the choice of δλ+\delta_{\lambda}^{+}, for example, as follows

δλ+={α34|β|1 if β0,α11γ if β=0.\delta_{\lambda}^{+}=\left\{\begin{aligned} &\alpha^{\frac{3}{4}}|\beta|^{-1}&&\text{ if }\beta\neq 0,\\ &\alpha^{\frac{1}{1-\gamma}}&&\text{ if }\beta=0.\end{aligned}\right. (4.23)

Step (1) of Proposition 4.5 is easily to be obtained since all the estimates in (4.13) and (4.14) on [0,a+][0,a_{+}] still hold. Since the potential VV still satisfy the assumption (2.2) with τ(x)=(x2+1)12\tau(x)=\left(x^{2}+1\right)^{-\frac{1}{2}}, the estimate (4.14) keep being true for all t[a+,δλ+]t\in\left[a_{+},\delta_{\lambda}^{+}\right] and thus

|k=0N1Re(λkψk(1)(t))|k=0N1|V(t)|τ+(t)k+1αk4+1|ImV(t)|α.\left|\sum_{k=0}^{N-1}\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\right|\lesssim\sum_{k=0}^{N-1}\frac{|V(t)|\tau_{+}(t)^{k+1}}{\alpha^{\frac{k}{4}+1}}\lesssim\frac{|\mathrm{Im}\,V(t)|}{\alpha}.

With the choice of δλ+\delta_{\lambda}^{+} satisfying (4.22), we have, for all t[a+,δλ+]t\in\left[a_{+},\delta_{\lambda}^{+}\right],

ImV(t)β=1βtγtγImV(t).\mathrm{Im}\,V(t)-\beta=\frac{1-\beta t^{\gamma}}{t^{\gamma}}\gtrsim\mathrm{Im}\,V(t).

Therefore, we estimate as in Step (2) of Proposition 4.5, that is the terms λkψk(1)\lambda^{-k}\psi_{k}^{(1)} for k0k\geq 0 can be neglected in the expansion of Pλ,NP_{\lambda,N} and we also obtain (4.17) for all x[a+,δλ+]x\in[a_{+},\delta_{\lambda}^{+}]. By choosing Δα=δλ+/4\Delta_{\alpha}=\delta_{\lambda}^{+}/4, we have for all x[δλ+Δλ+,δλ+]x\in[\delta_{\lambda}^{+}-\Delta_{\lambda}^{+},\delta_{\lambda}^{+}],

a+xImV(t)dtδλ+2Δλ+xtγdt(δλ+)1γ,\displaystyle\int_{a_{+}}^{x}\mathrm{Im}\,V(t)\,\mathrm{d}t\gtrsim\int_{\delta_{\lambda}^{+}-2\Delta_{\lambda}^{+}}^{x}t^{-\gamma}\,\mathrm{d}t\gtrsim\left(\delta_{\lambda}^{+}\right)^{1-\gamma},

and thus, there exists c1>0c_{1}>0 such that, for all x[δλ+Δλ+,δλ+]x\in[\delta_{\lambda}^{+}-\Delta_{\lambda}^{+},\delta_{\lambda}^{+}],

|exp(Pλ,N(x))|exp(c1α34(δλ+)γ1).\left|\exp\left(-P_{\lambda,N}(x)\right)\right|\lesssim\exp\left(-\frac{c_{1}}{\alpha^{\frac{3}{4}}\left(\delta_{\lambda}^{+}\right)^{\gamma-1}}\right). (4.24)

Thanks to (4.22), we know that the right hand side of (4.24) has a decay as λ\lambda\to\infty. If we strengthen (2.21) to (with some ε>0\varepsilon>0)

|β|α34γ1γ+ε=𝒪(1),λ,|\beta|\alpha^{\frac{3}{4}\frac{\gamma}{1-\gamma}+\varepsilon}=\mathcal{O}(1),\qquad\lambda\to\infty,

and choose δλ+\delta_{\lambda}^{+} as in (4.23), the decay of the right hand side of (4.24) is given by (with some c>0c>0 and η>0\eta>0)

|exp(Pλ,N(x))|=𝒪(exp(cαη)).\left|\exp\left(-P_{\lambda,N}(x)\right)\right|=\mathcal{O}\left(\exp\left(-c\alpha^{\eta}\right)\right).

The claim for δλ\delta_{\lambda}^{-}, Δλ\Delta_{\lambda}^{-} and the estimates on the negative axis are as same as the above. Therefore, by the same manner as Step (4), we obtain

(𝒟λ,Nξ)exp(Pλ,N)ξexp(Pλ,N)=o(1),λ.\frac{\|(\mathcal{D}_{\lambda,N}\xi)\exp(-P_{\lambda,N})\|}{\|\xi\exp(-P_{\lambda,N})\|}=o(1),\qquad\lambda\to\infty.

Concerning the remainder λ,N\mathcal{R}_{\lambda,N}, since VV and τ\tau are bounded on \mathbb{R}, we indeed have

λ,NL(Jλ)αN+14.\|\mathcal{R}_{\lambda,N}\|_{L^{\infty}(J_{\lambda})}\lesssim\alpha^{-\frac{N+1}{4}}.

5. Pseudomodes for large imagine pseudoeigenvalues

5.1. Pseudomode construction

Let Assumption II hold for some N0N\in\mathbb{N}_{0} and let us define the pseudomode for Theorem 2.2 as follows

Ψλ,N:=ξλexp(Pλ,N),\Psi_{\lambda,N}:=\xi_{\lambda}\exp\left(-P_{\lambda,N}\right),

where

  • ξλ\xi_{\lambda} is the cut-off function chosen such that, with Δβ\Delta_{\beta} and JβJ_{\beta} defined as in (2.15),

    ξλ𝒞0(+),0ξλ1,\displaystyle\xi_{\lambda}\in\mathcal{C}_{0}^{\infty}(\mathbb{R}_{+}),\quad 0\leq\xi_{\lambda}\leq 1, (5.1)
    ξλ(x)=1, for all x(xβΔβ,xβ+Δβ)=:Jβ,\displaystyle\xi_{\lambda}(x)=1,\qquad\text{ for all }x\in\left(x_{\beta}-\Delta_{\beta},x_{\beta}+\Delta_{\beta}\right)=:J_{\beta}^{\prime},
    ξλ(x)=0, for all x(xβ2Δβ,xβ+2Δβ)=Jβ.\displaystyle\xi_{\lambda}(x)=0,\qquad\text{ for all }x\notin\left(x_{\beta}-2\Delta_{\beta},x_{\beta}+2\Delta_{\beta}\right)=J_{\beta}.
  • Pλ,N(x)=k=1N1xβxλkψk(1)(t)dt\displaystyle P_{\lambda,N}(x)=\sum_{k=-1}^{N-1}\int_{x_{\beta}}^{x}\lambda^{-k}\psi_{k}^{(1)}(t)\,\mathrm{d}t in which ψ1(1)\psi_{-1}^{(1)} determined by (3.6) with the plus sign and (ψk(1))k[[0,N1]]\left(\psi_{k}^{(1)}\right)_{k\in[[0,N-1]]} determined by (3.7).

From (2.4) and (2.13), we can deduce that

τ(x)τ(xβ),ImV(x)ImV(xβ),ImV(1)(x)ImV(1)(xβ),xJβ.\tau(x)\approx\tau(x_{\beta}),\qquad\mathrm{Im}\,V(x)\approx\mathrm{Im}\,V(x_{\beta}),\qquad\mathrm{Im}\,V^{(1)}(x)\approx\mathrm{Im}\,V^{(1)}(x_{\beta}),\qquad x\in J_{\beta}. (5.2)
Proposition 5.1.

There exists c>0c>0 such that, for all β1\beta\gtrsim 1, we have

(𝒟λ,Nξ)exp(Pλ,N)L2(0+)ξexp(Pλ,N)L2(0+)exp(cImV(1)(xβ)τ(xβ)2α34+β34).\frac{\left\|(\mathcal{D}_{\lambda,N}\xi)\exp(-P_{\lambda,N})\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)}}{\left\|\xi\exp(-P_{\lambda,N})\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)}}\lesssim\exp\left(-c\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}\right). (5.3)
Proof.

Following are the lines of steps to prove the Theorem 2.2.

  1. (1)

    We begin the proof by showing that for sufficiently large β>0\beta>0, the set of admissible α\alpha in (2.16) and (5.6) is non-empty.

  2. (2)

    Under these conditions for α\alpha, we can show that, there exists a constant C>0C>0 such that

    (𝒟λ,Nξ)exp(Pλ,N)L2(0+)exp(CImV(1)(xβ)τ(xβ)2α34+β34).\left\|(\mathcal{D}_{\lambda,N}\xi)\exp(-P_{\lambda,N})\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)}\lesssim\exp\left(-C\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}\right). (5.4)
  3. (3)

    Since the support JβJ_{\beta} of pseudomode makes a move as β+\beta\to+\infty, we can not bound below the denominator ξexp(Pλ,N)L2(0+)2\left\|\xi\exp\left(-P_{\lambda,N}\right)\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)}^{2} by a constant as in [Prop. 4.5,Step (1)]. However, we can bound it below by a small term whose inverse can be controlled by the right hand side of (5.4). Then it yields (5.3).

Before entering the details of the proof, for simplifying the later computation, let us write

4+ε=41η4+\varepsilon=\frac{4}{1-\eta}

as in (4.4) where η1=η(0,1)\eta_{1}=\eta\in(0,1). Then, the condition (2.11) is rewritten in terms of η\eta as follows

(t1,t2)[0,η5(3+η)]2,t11η4t2<η20.\left(t_{1},t_{2}\right)\in\left[0,\frac{\eta}{5(3+\eta)}\right]^{2},\qquad t_{1}-\frac{1-\eta}{4}t_{2}<\frac{\eta}{20}. (5.5)

Similarly, the condition (2.17) for α\alpha is revised to

[βτ(xβ)]45|α|[βτ(xβ)1]43+η.\left[\beta\tau(x_{\beta})\right]^{\frac{4}{5}}\lesssim|\alpha|\lesssim\left[\beta\tau(x_{\beta})^{-1}\right]^{\frac{4}{3+\eta}}. (5.6)

Following are the details of the proof:

  1. (1)

    In order to show that the set of α\alpha satisfying (2.16) and (5.6) is non-empty, we can choose, for instance,

    α=α(β):=[βτ(xβ)1]43+η.\alpha=\alpha(\beta):=\left[\beta\tau(x_{\beta})^{-1}\right]^{\frac{4}{3+\eta}}.

    Then, from (4.4) which is a direct consequence of the assumption (2.5), we have

    τ(xβ)4β1η,β[βτ(xβ)1]43+η,|ReV(x)|11η[βτ(xβ)1]43+η,\tau(x_{\beta})^{4}\lesssim\beta^{1-\eta},\qquad\beta\lesssim\left[\beta\tau(x_{\beta})^{-1}\right]^{\frac{4}{3+\eta}},\qquad\left|\mathrm{Re}\,V(x)\right|^{\frac{1}{1-\eta}}\lesssim\left[\beta\tau(x_{\beta})^{-1}\right]^{\frac{4}{3+\eta}}, (5.7)

    and thus, we deduce (5.6) from

    [βτ(xβ)]45β[βτ(xβ)1]43+η.\left[\beta\tau(x_{\beta})\right]^{\frac{4}{5}}\lesssim\beta\lesssim\left[\beta\tau(x_{\beta})^{-1}\right]^{\frac{4}{3+\eta}}.

    Furthermore, (2.16) is also followed from (5.7):

    |ReV(x)|α1η,xJβ.|\mathrm{Re}\,V(x)|\lesssim\alpha^{1-\eta},\qquad x\in J_{\beta}.

    Therefore, this choice of α\alpha satisfies (2.16) and (5.6).

  2. (2)

    Following the work of Lemma 4.3 and Proposition 4.5, we firstly perform the estimate for the real part of the eikonal term xβxλψ1(1)(t)dt\displaystyle\int_{x_{\beta}}^{x}\lambda\psi_{-1}^{(1)}(t)\,\mathrm{d}t and then prove that the other transport terms xβxλkψk(1)(t)dt\displaystyle\int_{x_{\beta}}^{x}\lambda^{k}\psi_{k}^{(1)}(t)\,\mathrm{d}t, for k[[0,N1]]k\in[[0,N-1]] play less important roles than the eikonal one. Let us recall that

    Re(λψ1(1))=12ImVβ(|Vλ|+αReV)1/2(|Vλ|1/2+12(|Vλ|+αReV)1/2)1/2.\displaystyle\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}\right)=\frac{1}{2}\frac{\mathrm{Im}\,V-\beta}{\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\left(|V_{\lambda}|^{1/2}+\frac{1}{\sqrt{2}}\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\right)^{1/2}}.

    Thanks to (2.16) and (5.2), we have

    (|Vλ|+αReV)1/2(|Vλ|1/2+12(|Vλ|+αReV)1/2)1/2|α|34+β34.\displaystyle\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\left(|V_{\lambda}|^{1/2}+\frac{1}{\sqrt{2}}\left(|V_{\lambda}|+\alpha-\mathrm{Re}\,V\right)^{1/2}\right)^{1/2}\approx|\alpha|^{\frac{3}{4}}+\beta^{\frac{3}{4}}.

    By observing the sign of the term ImV(x)β\mathrm{Im}\,V(x)-\beta on the left and on the right of xβx_{\beta} on JβJ_{\beta}, it implies that, for all xJβx\in J_{\beta},

    xβxRe(λψ1(1)(t))dtxβx[ImV(t)β]dt|α|34+β44ImV(1)(xβ)(xxβ)2|α|34+β44.\int_{x_{\beta}}^{x}\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\,\mathrm{d}t\approx\frac{\displaystyle\int_{x_{\beta}}^{x}\left[\mathrm{Im}\,V(t)-\beta\right]\,\mathrm{d}t}{|\alpha|^{\frac{3}{4}}+\beta^{\frac{4}{4}}}\approx\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})(x-x_{\beta})^{2}}{|\alpha|^{\frac{3}{4}}+\beta^{\frac{4}{4}}}. (5.8)

    Here in the last estimate, we changed variable twice in integrals and employing (5.2), in detail, that is, for all xJβx\in J_{\beta},

    xβxImV(t)βdt\displaystyle\int_{x_{\beta}}^{x}\mathrm{Im}\,V(t)-\beta\,\mathrm{d}t =(xxβ)20101ξImV(1)(xβ+τξ(xxβ))dτdξ\displaystyle=(x-x_{\beta})^{2}\int_{0}^{1}\int_{0}^{1}\xi\mathrm{Im}\,V^{(1)}\left(x_{\beta}+\tau\xi(x-x_{\beta})\right)\,\mathrm{d}\tau\mathrm{d}\xi
    ImV(1)(xβ)(xxβ)2.\displaystyle\approx\mathrm{Im}\,V^{(1)}(x_{\beta})(x-x_{\beta})^{2}.

    On JβJβJ_{\beta}\setminus J_{\beta}^{\prime}, all xx stays away from the turning point xβx_{\beta} a distance Δβ=ητ(xβ)\Delta_{\beta}=\frac{\eta}{\tau(x_{\beta})}, hence, for every xJβJβx\in J_{\beta}\setminus J_{\beta}^{\prime},

    xβxRe(λψ1(1)(t))dt\displaystyle\int_{x_{\beta}}^{x}\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\,\mathrm{d}t ImV(1)(xβ)τ(xβ)2|α|34+β34\displaystyle\gtrsim\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{|\alpha|^{\frac{3}{4}}+\beta^{\frac{3}{4}}}
    {|α|η5βη5(3+η)t1τ(xβ)t2η5(3+η) if |α|>β,β14t1τ(xβ)t21 if |α|β.\displaystyle\gtrsim\left\{\begin{aligned} &|\alpha|^{\frac{\eta}{5}}\beta^{\frac{\eta}{5(3+\eta)}-t_{1}}\tau(x_{\beta})^{t_{2}-\frac{\eta}{5(3+\eta)}}&&\text{ if }|\alpha|>\beta,\\ &\beta^{\frac{1}{4}-t_{1}}\tau(x_{\beta})^{t_{2}-1}&&\text{ if }|\alpha|\leq\beta.\\ \end{aligned}\right. (5.9)

    In the second inequality, we used (2.4) and (5.6). Notice that, from the assumption (5.5), the powers of β\beta and τ\tau are related by the following inequalities

    η5(3+η)t1>1η4(η5(3+η)t2),14t1\displaystyle\frac{\eta}{5(3+\eta)}-t_{1}>\frac{1-\eta}{4}\left(\frac{\eta}{5(3+\eta)}-t_{2}\right),\qquad\frac{1}{4}-t_{1} >1η4(1t2)+η5.\displaystyle>\frac{1-\eta}{4}\left(1-t_{2}\right)+\frac{\eta}{5}.

    Combine this with the first inequality in (5.7) and the fact 0η5(3+η)t2<1t20\leq\frac{\eta}{5(3+\eta)}-t_{2}<1-t_{2}, we obtain, for all xJβJβx\in J_{\beta}\setminus J_{\beta}^{\prime},

    xβxRe(λψ1(1)(t))dtImV(1)(xβ)τ(xβ)2|α|34+β34{|α|η5 if |α|>β,βη5 if |α|β.\int_{x_{\beta}}^{x}\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\,\mathrm{d}t\gtrsim\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{|\alpha|^{\frac{3}{4}}+\beta^{\frac{3}{4}}}\gtrsim\left\{\begin{aligned} &|\alpha|^{\frac{\eta}{5}}&&\text{ if }|\alpha|>\beta,\\ &\beta^{\frac{\eta}{5}}&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right. (5.10)

    Next, in the same manner of proving (4.14), by the choice of α\alpha in (2.16) and (5.6) collaborating with the first inequality in (5.7), we can show that, for k[[1,N1]]k\in[[-1,N-1]] and m[[1,3]]m\in[[1,3]] such that k+m1k+m\geq 1, and for all tJβt\in J_{\beta},

    |λkψk(m)(t)|\displaystyle\left|\lambda^{-k}\psi_{k}^{(m)}(t)\right| j=1k+mτk+m(t)|V(t)|j|Vλ|j+k4τ(xβ)k+m|α|k4j=1k+m(1+β|α|)j\displaystyle\lesssim\sum_{j=1}^{k+m}\frac{\tau^{k+m}(t)|V(t)|^{j}}{|V_{\lambda}|^{j+\frac{k}{4}}}\lesssim\frac{\tau(x_{\beta})^{k+m}}{|\alpha|^{\frac{k}{4}}}\sum_{j=1}^{k+m}\left(1+\frac{\beta}{|\alpha|}\right)^{j}
    τ(xβ)k+m|α|k4(1+β|α|)k+m\displaystyle\lesssim\frac{\tau(x_{\beta})^{k+m}}{|\alpha|^{\frac{k}{4}}}\left(1+\frac{\beta}{|\alpha|}\right)^{k+m}
    {[τ(xβ)β14]kτ(xβ)m if |α|>β,[βτ(xβ)|α|54]k+4m5[βτ(xβ)]m5 if |α|β,\displaystyle\lesssim\left\{\begin{aligned} &\left[\frac{\tau(x_{\beta})}{\beta^{\frac{1}{4}}}\right]^{k}\tau(x_{\beta})^{m}&&\text{ if }|\alpha|>\beta,\\ &\left[\frac{\beta\tau(x_{\beta})}{|\alpha|^{\frac{5}{4}}}\right]^{k+\frac{4m}{5}}[\beta\tau(x_{\beta})]^{\frac{m}{5}}&&\text{ if }|\alpha|\leq\beta,\end{aligned}\right.
    {βkη4τ(xβ)m if |α|>β,[βτ(xβ)]m5 if |α|β.\displaystyle\lesssim\left\{\begin{aligned} &\beta^{-\frac{k\eta}{4}}\tau(x_{\beta})^{m}&&\text{ if }|\alpha|>\beta,\\ &[\beta\tau(x_{\beta})]^{\frac{m}{5}}&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right. (5.11)

    In particular, for k0k\geq 0 and at m=1m=1, by employing (5.10) for the case |α|>β|\alpha|>\beta and (5.9) for the other case, we obtain, for all xJβJβx\in J_{\beta}\setminus J_{\beta}^{\prime},

    |xβxλkψk(1)(t)dt|ImV(1)(xβ)τ(xβ)2α34+β34\displaystyle\frac{\displaystyle\left|\int_{x_{\beta}}^{x}\lambda^{-k}\psi_{k}^{(1)}(t)\,\mathrm{d}t\right|}{\displaystyle\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}} {|α|η5 if |α|>β,[β120t1τ(xβ)t215]1 if |α|β.\displaystyle\lesssim\left\{\begin{aligned} &|\alpha|^{-\frac{\eta}{5}}\qquad&&\text{ if }|\alpha|>\beta,\\ &\left[\beta^{\frac{1}{20}-t_{1}}\tau(x_{\beta})^{t_{2}-\frac{1}{5}}\right]^{-1}\qquad&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right.

    Furthermore, in the case |α|β|\alpha|\leq\beta, we notice that

    120t1=1η4(15t2)+η20(t11η4t2).\frac{1}{20}-t_{1}=\frac{1-\eta}{4}\left(\frac{1}{5}-t_{2}\right)+\frac{\eta}{20}-\left(t_{1}-\frac{1-\eta}{4}t_{2}\right). (5.12)

    Accordingly, for all k[[0,N1]]k\in[[0,N-1]] and for all xJβJβx\in J_{\beta}\setminus J_{\beta}^{\prime}, we get

    |xβxλkψk(1)(t)dt|xβxRe(λψ1(1)(t))dt\displaystyle\frac{\displaystyle\left|\int_{x_{\beta}}^{x}\lambda^{-k}\psi_{k}^{(1)}(t)\,\mathrm{d}t\right|}{\displaystyle\int_{x_{\beta}}^{x}\mathrm{Re}\,(\lambda\psi_{-1}^{(1)}(t))\,\mathrm{d}t} |xβxλkψk(1)(t)dt|ImV(1)(xβ)τ(xβ)2α34+β34\displaystyle\lesssim\frac{\displaystyle\left|\int_{x_{\beta}}^{x}\lambda^{-k}\psi_{k}^{(1)}(t)\,\mathrm{d}t\right|}{\displaystyle\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}}
    {|α|η5 if |α|>β,β[η20(t11η4t2)] if |α|β.\displaystyle\lesssim\left\{\begin{aligned} &|\alpha|^{-\frac{\eta}{5}}&&\text{ if }|\alpha|>\beta,\\ &\beta^{-\left[\frac{\eta}{20}-\left(t_{1}-\frac{1-\eta}{4}t_{2}\right)\right]}&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right. (5.13)

    Therefore, for all xJβJβx\in J_{\beta}\setminus J_{\beta}^{\prime}, we obtain (with some constant C1>0C_{1}>0)

    |exp(Pλ,N(x))|exp(C1ImV(1)(xβ)τ(xβ)2α34+β34).\left|\exp\left(-P_{\lambda,N}(x)\right)\right|\lesssim\exp\left(-C_{1}\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}\right).

    By using ((2)) and the fact |λψ1(1)|=|Vλ|14|α|14+β14\left|\lambda\psi_{-1}^{(1)}\right|=\left|V_{\lambda}\right|^{\frac{1}{4}}\lesssim|\alpha|^{\frac{1}{4}}+\beta^{\frac{1}{4}}, we have, for all xJβx\in J_{\beta},

    |Pλ,N(m)(x)|{|α|m4 if |α|>β,βm4 if |α|β.\left|P_{\lambda,N}^{(m)}(x)\right|\lesssim\left\{\begin{aligned} &|\alpha|^{\frac{m}{4}}&&\text{ if }|\alpha|>\beta,\\ &\beta^{\frac{m}{4}}&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right.

    With the help of (5.10), we can control all appearing polynomial terms in Dλ,NξD_{\lambda,N}\xi to get the estimate (5.4).

  3. (3)

    In this step, we will check that ξexp(Pλ,N)L2(0+)\left\|\xi\exp\left(-P_{\lambda,N}\right)\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)} is not too small. To do that, we set Δ~β=τ(xβ)45β120\widetilde{\Delta}_{\beta}=\tau(x_{\beta})^{-\frac{4}{5}}\beta^{-\frac{1}{20}}. Then, by (5.7), we have Δ~β<Δβ\widetilde{\Delta}_{\beta}<\Delta_{\beta} and thus

    ξexp(Pλ,N)L2(0+)2\displaystyle\left\|\xi\exp\left(-P_{\lambda,N}\right)\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)}^{2} xβxβ+Δ~βexp(2k=1N1|xβxRe(λkψk(1)(t))dt|)dx\displaystyle\geq\int_{x_{\beta}}^{x_{\beta}+\widetilde{\Delta}_{\beta}}\exp\left(-2\sum_{k=-1}^{N-1}\left|\int_{x_{\beta}}^{x}\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\,\mathrm{d}t\right|\right)\,\mathrm{d}x
    Δ~βexp(2k=1N1xβxβ+Δ~β|Re(λkψk(1)(t))|dt).\displaystyle\geq\widetilde{\Delta}_{\beta}\exp\left(-2\sum_{k=-1}^{N-1}\int_{x_{\beta}}^{x_{\beta}+\widetilde{\Delta}_{\beta}}\left|\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\right|\,\mathrm{d}t\right).

    For the integral of Re(λψ1)\mathrm{Re}\,\left(\lambda\psi_{-1}\right), we make use of (5.8) for |α|>β|\alpha|>\beta and the fact |Re(λψ1(1))||Vλ|34β34\left|\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}\right)\right|\leq\left|V_{\lambda}\right|^{\frac{3}{4}}\lesssim\beta^{\frac{3}{4}} for |α|β|\alpha|\leq\beta, then we have

    xβxβ+Δ~β|Re(λψ1(1)(t))|dt{|α|34ImV(1)(xβ)(Δ~β)2 if |α|>β,β14Δ~β if |α|β.\int_{x_{\beta}}^{x_{\beta}+\widetilde{\Delta}_{\beta}}\left|\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\right|\,\mathrm{d}t\lesssim\left\{\begin{aligned} &|\alpha|^{-\frac{3}{4}}\mathrm{Im}\,V^{(1)}(x_{\beta})\left(\widetilde{\Delta}_{\beta}\right)^{2}&&\text{ if }|\alpha|>\beta,\\ &\beta^{\frac{1}{4}}\widetilde{\Delta}_{\beta}&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right.

    Thus, by the definition of Δ~β\widetilde{\Delta}_{\beta} and (5.12) combining with the first inequality in (5.7), we obtain

    xβxβ+Δ~β|Re(λψ1(1)(t))|dtImV(1)(xβ)τ(xβ)2α34+β34\displaystyle\frac{\displaystyle\int_{x_{\beta}}^{x_{\beta}+\widetilde{\Delta}_{\beta}}\left|\mathrm{Re}\,\left(\lambda\psi_{-1}^{(1)}(t)\right)\right|\,\mathrm{d}t}{\displaystyle\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}} {[τ(xβ)4β1]110 if |α|>β,[β120t1τ(xβ)t215]1 if |α|β,\displaystyle\lesssim\left\{\begin{aligned} &\left[\tau(x_{\beta})^{4}\beta^{-1}\right]^{\frac{1}{10}}&&\text{ if }|\alpha|>\beta,\\ &\left[\beta^{\frac{1}{20}-t_{1}}\tau(x_{\beta})^{t_{2}-\frac{1}{5}}\right]^{-1}&&\text{ if }|\alpha|\leq\beta,\end{aligned}\right.
    {βη10 if |α|>β,β[η20(t11η4t2)] if |α|β.\displaystyle\lesssim\left\{\begin{aligned} &\beta^{-\frac{\eta}{10}}&&\text{ if }|\alpha|>\beta,\\ &\beta^{-\left[\frac{\eta}{20}-\left(t_{1}-\frac{1-\eta}{4}t_{2}\right)\right]}&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right.

    For k[[0,N1]]k\in[[0,N-1]], we use (5.13) to get

    xβxβ+Δ~β|Re(λkψk(1)(t))|dtImV(1)(xβ)τ(xβ)2α34+β34{|α|η5 if |α|>β,β[η20(t11η4t2)] if |α|β.\displaystyle\frac{\displaystyle\int_{x_{\beta}}^{x_{\beta}+\widetilde{\Delta}_{\beta}}\left|\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\right|\,\mathrm{d}t}{\displaystyle\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}}\lesssim\left\{\begin{aligned} &|\alpha|^{-\frac{\eta}{5}}&&\text{ if }|\alpha|>\beta,\\ &\beta^{-\left[\frac{\eta}{20}-\left(t_{1}-\frac{1-\eta}{4}t_{2}\right)\right]}&&\text{ if }|\alpha|\leq\beta.\end{aligned}\right.

    From the above estimates, we have shown that

    k=1N1xβxβ+Δ~β|Re(λkψk(1)(t))|dt=o(ImV(1)(xβ)τ(xβ)2α34+β34),β+.\sum_{k=-1}^{N-1}\int_{x_{\beta}}^{x_{\beta}+\widetilde{\Delta}_{\beta}}\left|\mathrm{Re}\,\left(\lambda^{-k}\psi_{k}^{(1)}(t)\right)\right|\,\mathrm{d}t=o\left(\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}\right),\qquad\beta\to+\infty.

    Therefore, we have

    (𝒟λ,Nξ)exp(Pλ,N)L2(0+)ξexp(Pλ,N)L2(0+)\displaystyle\frac{\left\|(\mathcal{D}_{\lambda,N}\xi)\exp(-P_{\lambda,N})\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)}}{\left\|\xi\exp(-P_{\lambda,N})\right\|_{L^{2}\left(\mathbb{R}_{0}^{+}\right)}} Δ~β12exp((Co(1))ImV(1)(xβ)τ(xβ)2α34+β34).\displaystyle\lesssim\widetilde{\Delta}_{\beta}^{-\frac{1}{2}}\exp\left(-\left(C-o(1)\right)\frac{\mathrm{Im}\,V^{(1)}(x_{\beta})\tau(x_{\beta})^{-2}}{\alpha^{\frac{3}{4}}+\beta^{\frac{3}{4}}}\right).

    Thus, (5.3) is followed directly by using (5.7) and (5.10) to control the term Δ~β12\widetilde{\Delta}_{\beta}^{-\frac{1}{2}}.

5.2. Remainder estimate (Proof of Theorem 2.2)

By using the same trick as proving (4.12), we can show that

|V(x)||V(xβ)|,xJβ.|V(x)|\approx|V(x_{\beta})|,\qquad x\in J_{\beta}. (5.14)

Indeed, since VV is a complex-valued function, we should be careful with some steps when we perform estimation, more precisely, for all hh\in\mathbb{R} such that |h|2Δβ|h|\leq 2\Delta_{\beta}, we have

|ln|V(xβ+h)||V(xβ)|||Log V(xβ+h)Log V(xβ)|=|V(xβ)V(xβ+h)dzz|=|xβxβ+hV(t)V(t)dt|1,\displaystyle\left|\ln\frac{|V(x_{\beta}+h)|}{|V(x_{\beta})|}\right|\leq\left|\textup{Log }V(x_{\beta}+h)-\textup{Log }V(x_{\beta})\right|=\left|\int_{V(x_{\beta})}^{V(x_{\beta}+h)}\frac{\mathrm{d}z}{z}\right|=\left|\int_{x_{\beta}}^{x_{\beta}+h}\frac{V^{\prime}(t)}{V(t)}\,\mathrm{d}t\right|\lesssim 1,

in which Log z\textup{Log }z is the principal branch of the logarithmic function with its antiderivative 1z\frac{1}{z} and keep in mind that the range V(Jβ)V(J_{\beta}) stays away the negative semi-axis 0\mathbb{R}_{0}^{-}, and the last inequality is estimated as same as proving (4.12) by using (2.2).

We finish the proof by estimating the remainder whose shape is given in Lemma 3.3, for N1N\geq 1 and for all xJβx\in J_{\beta},

|λ,N(x)|\displaystyle\left|\mathcal{R}_{\lambda,N}(x)\right| k=03N11|Vλ(x)|k+N34j=1k+N+1|dk+N+1,j(x)||Vλ(x)|j\displaystyle\lesssim\sum_{k=0}^{3N-1}\frac{1}{\left|V_{\lambda}(x)\right|^{\frac{k+N-3}{4}}}\sum_{j=1}^{k+N+1}\frac{\left|d_{k+N+1,j}(x)\right|}{\left|V_{\lambda}(x)\right|^{j}}
k=03N1j=1k+N+1τ(x)k+N+1|V(x)|j|Vλ(x)|k+N34+j\displaystyle\lesssim\sum_{k=0}^{3N-1}\sum_{j=1}^{k+N+1}\frac{\tau(x)^{k+N+1}\left|V(x)\right|^{j}}{\left|V_{\lambda}(x)\right|^{\frac{k+N-3}{4}+j}}
k=03N1j=1k+N+1τ(xβ)k+N+1(|ReV(xβ)|j+βj)|α|k+N34+j.\displaystyle\lesssim\sum_{k=0}^{3N-1}\sum_{j=1}^{k+N+1}\frac{\tau(x_{\beta})^{k+N+1}\left(\left|\mathrm{Re}\,V(x_{\beta})\right|^{j}+\beta^{j}\right)}{\left|\alpha\right|^{\frac{k+N-3}{4}+j}}.

Here, we employed (2.2) to control dk+N+1,jd_{k+N+1,j}, (2.16) for VλV_{\lambda} in the denominator, (5.7) for function τ\tau and (5.14) for VV. The remainder λ,0\mathcal{R}_{\lambda,0} is estimated analogously.

5.3. Pseudomode for semi-classical biharmonic operator (Proof of Theorem 2.3)

Since RezReW(x0)=μ>0\mathrm{Re}\,z-\mathrm{Re}\,W(x_{0})=\mu>0, there exists an interval centered at x0x_{0} included in II, denoted as J(x02Δ,x0+2Δ)IJ\coloneqq(x_{0}-2\Delta,x_{0}+2\Delta)\subset I, such that the function RezReW(t)\mathrm{Re}\,z-\mathrm{Re}\,W(t) is positive in JJ. Without loss of generality, we assume further that ImW(x)ImW(x0)<0\mathrm{Im}\,W(x)-\mathrm{Im}\,W(x_{0})<0 for all x(x02Δ,x0)x\in(x_{0}-2\Delta,x_{0}) and ImW(x)ImW(x0)>0\mathrm{Im}\,W(x)-\mathrm{Im}\,W(x_{0})>0 for all x(x0,x0+2Δ)x\in(x_{0},x_{0}+2\Delta). Let us write our semi-classical problem in our previous setting by factoring the parameter h4h^{4} out

Hhz=h4(d4dx4+Vh(x)λh),H_{h}-z=h^{4}\left(\frac{\mathrm{d}^{4}}{\mathrm{d}x^{4}}+V_{h}(x)-\lambda_{h}\right),

where Vh(x):=h4W(x)V_{h}(x):=h^{-4}W(x) and λh:=h4z\lambda_{h}:=h^{-4}z. Being inspired by the above analysis, the pseudomode is achieved around the point x0x_{0} satisfying the equation ImVh(x0)=Imλh\mathrm{Im}\,V_{h}(x_{0})=\mathrm{Im}\,\lambda_{h}, i.e. ImW(x0)=Imz\mathrm{Im}\,W(x_{0})=\mathrm{Im}\,z. We should keep in mind that the point x0x_{0} here is fixed. We set up the pseudomode as follows

Ψh,N=ξexp(Ph,N),\Psi_{h,N}=\xi\exp\left(-P_{h,N}\right),

in which

  • ξ𝒞0()\xi\in\mathcal{C}_{0}^{\infty}(\mathbb{R}) is a cut-off function which is equal 11 on J(x0Δ,x0+Δ)J^{\prime}\coloneqq\left(x_{0}-\Delta,x_{0}+\Delta\right) and equal 0 in the complement of JJ in \mathbb{R},

  • Ph,N(x)k=1N1x0xλhkψk(1)(t)dt\displaystyle P_{h,N}(x)\coloneqq\sum_{k=-1}^{N-1}\int_{x_{0}}^{x}\lambda_{h}^{-k}\psi_{k}^{(1)}(t)\,\mathrm{d}t where ψ1(1)\psi_{-1}^{(1)} is determined by (3.6) with the plus sign and (ψk(1))k[[0,N1]]\left(\psi_{k}^{(1)}\right)_{k\in[[0,N-1]]} is determined by (3.7), in which VV is replaced by VhV_{h} and λ\lambda is replaced by λh\lambda_{h}. If the above sign of ImW(x)ImW(x0)\mathrm{Im}\,W(x)-\mathrm{Im}\,W(x_{0}) changes inversely on the left and on the right of x0x_{0}, we merely choose the minus sign in (3.6).

Notice that

λhVh(t)=h4[(RezReW(t))+i(ImW(x0)ImW(t))],\lambda_{h}-V_{h}(t)=h^{-4}\left[\left(\mathrm{Re}\,z-\mathrm{Re}\,W(t)\right)+i\left(\mathrm{Im}\,W(x_{0})-\mathrm{Im}\,W(t)\right)\right],

the function (λhVh)14\left(\lambda_{h}-V_{h}\right)^{\frac{1}{4}} is well-defined and so are all functions ψk(1)\psi_{k}^{(1)}, for k[[1,N1]]k\in[[-1,N-1]] on JJ. By the formula of Re(λhψ1(1)(t))\mathrm{Re}\,\left(\lambda_{h}\psi_{-1}^{(1)}(t)\right) in (4.15), we obtain, for all tx0t\geq x_{0},

Re(λhψ1(1)(t))h1ImW(t)ImW(x0)(RezReW(t))34+|ImW(t)ImW(x0)|34,t(x0,x0+2Δ)\displaystyle\mathrm{Re}\,\left(\lambda_{h}\psi_{-1}^{(1)}(t)\right)\gtrsim h^{-1}\frac{\mathrm{Im}\,W(t)-\mathrm{Im}\,W(x_{0})}{\left(\mathrm{Re}\,z-\mathrm{Re}\,W(t)\right)^{\frac{3}{4}}+|\mathrm{Im}\,W(t)-\mathrm{Im}\,W(x_{0})|^{\frac{3}{4}}},\qquad t\in(x_{0},x_{0}+2\Delta)

and

Re(λhψ1(1)(t))h1ImW(t)ImW(x0)(RezReW(t))34+|ImW(t)ImW(x0)|34,t(x02Δ,x0).\displaystyle\mathrm{Re}\,\left(\lambda_{h}\psi_{-1}^{(1)}(t)\right)\lesssim h^{-1}\frac{\mathrm{Im}\,W(t)-\mathrm{Im}\,W(x_{0})}{\left(\mathrm{Re}\,z-\mathrm{Re}\,W(t)\right)^{\frac{3}{4}}+|\mathrm{Im}\,W(t)-\mathrm{Im}\,W(x_{0})|^{\frac{3}{4}}},\qquad t\in(x_{0}-2\Delta,x_{0}).

On JJJ\setminus J^{\prime}, the function x0x(ImW(t)ImW(x0))dt\int_{x_{0}}^{x}\left(\mathrm{Im}\,W(t)-\mathrm{Im}\,W(x_{0})\right)\,\mathrm{d}t is bounded below by a positive constant, we have

x0xRe(λhψ1(1)(t))dth1.\int_{x_{0}}^{x}\mathrm{Re}\,\left(\lambda_{h}\psi_{-1}^{(1)}(t)\right)\,\mathrm{d}t\gtrsim h^{-1}.

Thanks to the shape of the WKB solutions in Lemma 3.2, it can be seen that, for all k1k\geq-1 and for all tJt\in J,

|λhkψk(1)(t)|hk.\left|\lambda_{h}^{-k}\psi_{k}^{(1)}(t)\right|\lesssim h^{k}.

It yields that the transport terms (ψk(1))k0\left(\psi_{k}^{(1)}\right)_{k\geq 0} are harmless in the expansion of pseudomode and thus there exists c1>0c_{1}>0 such that, for all xJJx\in J\setminus J^{\prime},

|exp(Ph,N(x))|exp(c1h1).\left|\exp\left(-P_{h,N}(x)\right)\right|\lesssim\exp(-c_{1}h^{-1}).

By considering on the fixed support JJ, we also obtain (with some c2>0c_{2}>0)

(𝒟λh,Nξ)exp(Ph,N)L2()exp(c2h1).\left\|(\mathcal{D}_{\lambda_{h},N}\xi)\exp(-P_{h,N})\right\|_{L^{2}\left(\mathbb{R}\right)}\lesssim\exp(-c_{2}h^{-1}).

Let Δ~=hΔ\widetilde{\Delta}=h\Delta with h<1h<1, we have, for all x[x0,x0+Δ~]Jx\in[x_{0},x_{0}+\widetilde{\Delta}]\subset J^{\prime},

|Ph,N(x)|k=1N1Δ~hk1.\left|P_{h,N}(x)\right|\leq\sum_{k=-1}^{N-1}\widetilde{\Delta}h^{k}\lesssim 1.

Thus, as in [Prop. 4.5, Step (1)], it implies that

Ψh1.\|\Psi_{h}\|\gtrsim 1.

Concerning the remainder, we have, for all N0N\geq 0 and xJx\in J,

h4|λ,N(x)|hN+1.\displaystyle h^{4}\left|\mathcal{R}_{\lambda,N}(x)\right|\lesssim h^{N+1}.

The conclusion of Theorem 2.3 is followed.

Appendix A The solutions of the transport equations and the WKB remainder

This appendix is devoted to the proofs of Lemma 3.2 and Lemma 3.3. These lemmata describe the structure of the transport solutions (ψk(1))k[[1,n1]]\left(\psi_{k}^{(1)}\right)_{k\in[[-1,n-1]]} and the WKB remainder λ,n\mathcal{R}_{\lambda,n} which are made from the elements of Dr,jD_{r,j} defined in Notation 3.1. First, let us recall here some simple observations about Dr,jD_{r,j} which are mentioned in [21, Appendix A].

  1. (1)

    If r1r\geq 1, then Dr,0={0}D_{r,0}=\{0\}. If r<jr<j, then Dr,j={0}D_{r,j}=\{0\}.

  2. (2)

    D0,0=D_{0,0}=\mathbb{C}.

  3. (3)

    Dr,j+Dr,j=Dr,jD_{r,j}+D_{r,j}=D_{r,j}.

  4. (4)

    cDr,j=Djr,scD_{r,j}=D_{j}^{r,s} for any constant cc\in\mathbb{C}.

  5. (5)

    Dr1,j1Dr2,j2Dr1+r2,j1+j2D_{r_{1},j_{1}}D_{r_{2},j_{2}}\subset D_{r_{1}+r_{2},j_{1}+j_{2}}.

Proof of Lemma 3.2.

The induction method is employed to prove the following statement with respect to the index k[[1,n1]]k\in[[-1,n-1]]:

ψk(m)λkVλk4j=0k+mDk+m,jVλj,m[[1,n+3k]].\psi_{k}^{(m)}\in\frac{\lambda^{k}}{V_{\lambda}^{\frac{k}{4}}}\sum_{j=0}^{k+m}\frac{D_{k+m,j}}{V_{\lambda}^{j}},\qquad m\in[[1,n+3-k]]. (A.1)

Base step: We check that the statement is true for k=1k=-1. It is a direct application of Faà di Bruno’s formula for the high derivative of the composition of the functions f(x):=x14f(x):=x^{\frac{1}{4}} and g(x):=Vλ(x)g(x):=V_{\lambda}(x), for 1\ell\in\mathbb{N}_{1},

ddxf(g(x))=j=1f(j)(g(x))B,j(g(1)(x),g(2)(x),,g(j+1)(x)).\frac{\mathrm{d}^{\ell}}{\mathrm{d}x^{\ell}}f(g(x))=\sum_{j=1}^{\ell}f^{(j)}(g(x))B_{\ell,j}\left(g^{(1)}(x),g^{(2)}(x),\ldots,g^{(\ell-j+1)}(x)\right).

Here B,jB_{\ell,j} are Bell polynomials which have formulae

B,j(x1,x2,,xj+1)=αI.j!α1!α2!αj+1!(x11!)α1(x22!)α2(xj+1(j+1)!)αj+1,B_{\ell,j}\left(x_{1},x_{2},\ldots,x_{\ell-j+1}\right)=\sum_{\alpha\in I_{\ell.j}}\frac{\ell!}{\alpha_{1}!\alpha_{2}!\ldots\alpha_{\ell-j+1}!}\left(\frac{x_{1}}{1!}\right)^{\alpha_{1}}\left(\frac{x_{2}}{2!}\right)^{\alpha_{2}}\ldots\left(\frac{x_{\ell-j+1}}{(\ell-j+1)!}\right)^{\alpha_{\ell-j+1}},

where α=(α1,α2,,αj+1)\alpha=(\alpha_{1},\alpha_{2},\ldots,\alpha_{\ell-j+1}) and I,jI_{\ell,j} is the set defined in (3.10). It is not difficult to see that f(j)(x)=cjx14jf^{(j)}(x)=c_{j}x^{\frac{1}{4}-j} for some cjc_{j}\in\mathbb{R} and

B,j(g(1)(x),g(2)(x),,g(j+1)(x))D,j.B_{\ell,j}\left(g^{(1)}(x),g^{(2)}(x),\ldots,g^{(\ell-j+1)}(x)\right)\in D_{\ell,j}.

By plugging these objects in the formula of ddxψ1(1)=iλ1ddxVλ14\frac{\mathrm{d}^{\ell}}{\mathrm{d}x^{\ell}}\psi_{-1}^{(1)}=i\lambda^{-1}\frac{\mathrm{d}^{\ell}}{\mathrm{d}x^{\ell}}V_{\lambda}^{\frac{1}{4}}, regarding the rule (1), it implies that, for all 1\ell\in\mathbb{N}_{1},

ψ1(+1)λ1Vλ14j=0D,jVλj.\psi_{-1}^{(\ell+1)}\in\frac{\lambda^{-1}}{V_{\lambda}^{-\frac{1}{4}}}\sum_{j=0}^{\ell}\frac{D_{\ell,j}}{V_{\lambda}^{j}}.

Thanks to the rule (2), the claim in the case k=1k=-1 is confirmed, for all m1m\in\mathbb{N}_{1},

ψ1(m)λ1Vλ14j=0m1Dm1,jVλj.\psi_{-1}^{(m)}\in\frac{\lambda^{-1}}{V_{\lambda}^{-\frac{1}{4}}}\sum_{j=0}^{m-1}\frac{D_{m-1,j}}{V_{\lambda}^{j}}.

Since VWlocn+3,2()V\in W^{n+3,2}_{\mathrm{loc}}(\mathbb{R}), the maximal order of the derivative that ψ1\psi_{-1} can be taken is n+4n+4, in which V(n+3)V^{(n+3)} appears in Dn+3,1D_{n+3,1} (see Notation 3.1).
Inductive step: Let q[[1,n2]]q\in[[-1,n-2]], we assume that (A.1) holds for all k[[1,q]]k\in[[-1,q]], we need to show that

ψq+1(m)λq+1Vλq+14j=0q+m+1Dq+1+m,jVλj,m[[1,n+2q]].\psi_{q+1}^{(m)}\in\frac{\lambda^{q+1}}{V_{\lambda}^{\frac{q+1}{4}}}\sum_{j=0}^{q+m+1}\frac{D_{q+1+m,j}}{V_{\lambda}^{j}},\qquad m\in[[1,n+2-q]]. (A.2)

From the formula (3.7), by using the Leibniz product rule, we obtain

ψq+1(m)=\displaystyle\psi_{q+1}^{(m)}= (ψq+1(1))(m1)=i=0m1(m1i)[14(ψ1(1))3](m1i)×\displaystyle\left(\psi_{q+1}^{(1)}\right)^{(m-1)}=\sum_{i=0}^{m-1}\begin{pmatrix}m-1\\ i\end{pmatrix}\left[\frac{1}{4\left(\psi_{-1}^{(1)}\right)^{3}}\right]^{(m-1-i)}\times
[ψq2(4+i)4α1+α2=q2(ψα1(1)ψα2(3))(i)3α1+α2=q2(ψα1(2)ψα2(2))(i)\displaystyle\hskip 76.82234pt\left[\psi_{q-2}^{(4+i)}-4\sum_{\alpha_{1}+\alpha_{2}=q-2}\left(\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}\right)^{(i)}-3\sum_{\alpha_{1}+\alpha_{2}=q-2}\left(\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)}\right)^{(i)}\right.
+6α1+α2+α3=q2(ψα1(1)ψα2(1)ψα3(2))(i)α1+α2+α3+α4=q2α1,α2,α3,α4q+1(ψα1(1)ψα2(1)ψα3(1)ψα4(1))(i)].\displaystyle\left.\hskip 56.9055pt+6\sum_{\alpha_{1}+\alpha_{2}+\alpha_{3}=q-2}\left(\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(2)}\right)^{(i)}-\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=q-2\\ \alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4}\neq q+1\end{subarray}}\left(\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(1)}\psi_{\alpha_{4}}^{(1)}\right)^{(i)}\right].

We compute each term appearing in the above formula:

  1. a)

    For the derivatives of 1(ψ1(1))3\frac{1}{\left(\psi_{-1}^{(1)}\right)^{3}}: we do the same manner as computing the derivatives of ψ1(1)\psi_{-1}^{(1)} in the base step by considering the functions f(x)=x34f(x)=x^{-\frac{3}{4}} and g(x)=Vλg(x)=V_{\lambda}. Since f(j)(x)=cjx34jf^{(j)}(x)=c_{j}x^{-\frac{3}{4}-j} for some cjc_{j}\in\mathbb{R}, we obtain

    (1(ψ1(1))3)()λ3Vλ34j=0D,jVλj,[[0,n+3]].\left(\frac{1}{\left(\psi_{-1}^{(1)}\right)^{3}}\right)^{(\ell)}\in\frac{\lambda^{3}}{V_{\lambda}^{\frac{3}{4}}}\sum_{j=0}^{\ell}\frac{D_{\ell,j}}{V_{\lambda}^{j}},\qquad\ell\in[[0,n+3]]. (A.3)
  2. b)

    For the derivatives of ψq2\psi_{q-2}: for each m[[1,n+2q]]m\in[[1,n+2-q]], i[[0,m1]]i\in[[0,m-1]], we have 4+i[[1,n+3(q2)]]4+i\in[[1,n+3-(q-2)]]. It means that we can use the induction assumption for the derivatives of ψq2\psi_{q-2}:

    ψq2(4+i)λq2Vλq24j=0q2+4+iDq2+4+i,jVλj=λq2Vλq24j=0q+2+iDq+2+i,jVλj.\psi_{q-2}^{(4+i)}\in\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q-2+4+i}\frac{D_{q-2+4+i,j}}{V_{\lambda}^{j}}=\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q+2+i}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}}. (A.4)
  3. c)

    The derivatives of ψα1(1)ψα2(3)\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)} and ψα1(2)ψα2(2)\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)} where α1+α2=q2\alpha_{1}+\alpha_{2}=q-2. By the Leibniz product rule again, we have, for each i[[0,m1]]i\in[[0,m-1]],

    (ψα1(1)ψα2(3))(i)==0i(i)ψα1(+1)ψα2(i+3).\displaystyle\left(\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}\right)^{(i)}=\sum_{\ell=0}^{i}\begin{pmatrix}i\\ \ell\end{pmatrix}\psi_{\alpha_{1}}^{(\ell+1)}\psi_{\alpha_{2}}^{(i-\ell+3)}.

    For each m[[1,n+2q]]m\in[[1,n+2-q]], i[[0,m1]]i\in[[0,m-1]] and [[0,i]]\ell\in[[0,i]], we notice that for α11\alpha_{1}\geq-1, α21\alpha_{2}\geq-1 and α1+α2=q2\alpha_{1}+\alpha_{2}=q-2, we have +1[[1,n+3α1]]\ell+1\in[[1,n+3-\alpha_{1}]] and i+3[[1,n+3α2]]i-\ell+3\in[[1,n+3-\alpha_{2}]]. This enables us to use the induction assumption to rewrite both ψα1(+1)\psi_{\alpha_{1}}^{(\ell+1)} and ψα2(i+3)\psi_{\alpha_{2}}^{(i-\ell+3)}. Then, we arrive at

    ψα1(+1)ψα2(i+3)\displaystyle\psi_{\alpha_{1}}^{(\ell+1)}\psi_{\alpha_{2}}^{(i-\ell+3)} (λα1Vλα14j1=0α1++1Dα1++1,j1Vλj1)(λα2Vλα24j2=0α2+i+3Dα2+i+3,j2Vλj2)\displaystyle\in\left(\frac{\lambda^{\alpha_{1}}}{V_{\lambda}^{\frac{\alpha_{1}}{4}}}\sum_{j_{1}=0}^{\alpha_{1}+\ell+1}\frac{D_{\alpha_{1}+\ell+1,j_{1}}}{V_{\lambda}^{j_{1}}}\right)\left(\frac{\lambda^{\alpha_{2}}}{V_{\lambda}^{\frac{\alpha_{2}}{4}}}\sum_{j_{2}=0}^{\alpha_{2}+i-\ell+3}\frac{D_{\alpha_{2}+i-\ell+3,j_{2}}}{V_{\lambda}^{j_{2}}}\right)
    =λα1+α2Vλα1+α24j=0α1+α2+i+4q1+q2=jDα1++1,q1Vλq1Dα2+i+3,q2Vλq2\displaystyle=\frac{\lambda^{\alpha_{1}+\alpha_{2}}}{V_{\lambda}^{\frac{\alpha_{1}+\alpha_{2}}{4}}}\sum_{j=0}^{\alpha_{1}+\alpha_{2}+i+4}\sum_{q_{1}+q_{2}=j}\frac{D_{\alpha_{1}+\ell+1,q_{1}}}{V_{\lambda}^{q_{1}}}\frac{D_{\alpha_{2}+i-\ell+3,q_{2}}}{V_{\lambda}^{q_{2}}}
    λq2Vλq24j=0q+2+iq1+q2=jDq+2+i,jVλj\displaystyle\subset\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q+2+i}\sum_{q_{1}+q_{2}=j}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}}
    =λq2Vλq24j=0q+2+iDq+2+i,jVλj,\displaystyle=\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q+2+i}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}},

    in which we used

    • -

      the rule (1) for the first equality, that is Dα1++1,q1={0}D_{\alpha_{1}+\ell+1,q_{1}}=\{0\} and Dα2+i+3,q2={0}D_{\alpha_{2}+i-\ell+3,q_{2}}=\{0\} when q1>α1++1q_{1}>\alpha_{1}+\ell+1 and if q2>α2+i+3q_{2}>\alpha_{2}+i-\ell+3,

    • -

      the rule (5) and α1+α2=q2\alpha_{1}+\alpha_{2}=q-2 for the inclusion,

    • -

      the rule (3) for the second equality.

    Since the resulting terms do not depend on \ell any more, by the rule (3) and (4), it implies that

    (ψα1(1)ψα2(3))(i)=0i(i)λq2Vλq24j=0q+i+2q1+q2=jDq+2+i,jVλj=λq2Vλq24j=0q+2+iDq+2+i,jVλj.\left(\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}\right)^{(i)}\in\sum_{\ell=0}^{i}\begin{pmatrix}i\\ \ell\end{pmatrix}\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q+i+2}\sum_{q_{1}+q_{2}=j}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}}=\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q+2+i}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}}. (A.5)
  4. d)

    In the same manner as above, we have

    (ψα1(2)ψα2(2))(i)\displaystyle\left(\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)}\right)^{(i)} λk2Vλk24j=0q+2+iDq+2+i,jVλj,\displaystyle\in\frac{\lambda^{k-2}}{V_{\lambda}^{\frac{k-2}{4}}}\sum_{j=0}^{q+2+i}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}}, (A.6)
    (ψα1(1)ψα2(1)ψα3(2))(i)\displaystyle\left(\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(2)}\right)^{(i)} λq2Vλq24j=0q+2+iDq+2+i,jVλj,\displaystyle\in\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q+2+i}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}},
    (ψα1(1)ψα2(1)ψα3(1)ψα4(1))(i)\displaystyle\left(\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(1)}\psi_{\alpha_{4}}^{(1)}\right)^{(i)} λq2Vλq24j=0q+2+iDq+2+i,jVλj.\displaystyle\in\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j=0}^{q+2+i}\frac{D_{q+2+i,j}}{V_{\lambda}^{j}}.

To finish the proof, we put (A.3), (A.4), (A.5) and (A.6) together into the formula of ψq+1(m)\psi_{q+1}^{(m)}, we obtain

ψq+1(m)\displaystyle\psi_{q+1}^{(m)}\in i=0m1(λ3Vλ34j1=0m1iDm1i,j1Vλj1)(λq2Vλq24j2=0q+2+iDq+2+i,j2Vλj2)\displaystyle\sum_{i=0}^{m-1}\left(\frac{\lambda^{3}}{V_{\lambda}^{\frac{3}{4}}}\sum_{j_{1}=0}^{m-1-i}\frac{D_{m-1-i,j_{1}}}{V_{\lambda}^{j_{1}}}\right)\left(\frac{\lambda^{q-2}}{V_{\lambda}^{\frac{q-2}{4}}}\sum_{j_{2}=0}^{q+2+i}\frac{D_{q+2+i,j_{2}}}{V_{\lambda}^{j_{2}}}\right)
=\displaystyle= i=0m1λq+1Vλq+14j=0q+m+1q1+q2=jDm1i,q1Vλq1Dq+2+i,q2Vλq2\displaystyle\sum_{i=0}^{m-1}\frac{\lambda^{q+1}}{V_{\lambda}^{\frac{q+1}{4}}}\sum_{j=0}^{q+m+1}\sum_{q_{1}+q_{2}=j}\frac{D_{m-1-i,q_{1}}}{V_{\lambda}^{q_{1}}}\frac{D_{q+2+i,q_{2}}}{V_{\lambda}^{q_{2}}}
\displaystyle\subset i=0m1λq+1Vλq+14j=0q+m+1q1+q2=jDq+m+1,q1+q2Vλq1+q2\displaystyle\sum_{i=0}^{m-1}\frac{\lambda^{q+1}}{V_{\lambda}^{\frac{q+1}{4}}}\sum_{j=0}^{q+m+1}\sum_{q_{1}+q_{2}=j}\frac{D_{q+m+1,q_{1}+q_{2}}}{V_{\lambda}^{q_{1}+q_{2}}}
=\displaystyle= λq+1Vλq+14j=0q+m+1Dq+m+1,jVλj,\displaystyle\frac{\lambda^{q+1}}{V_{\lambda}^{\frac{q+1}{4}}}\sum_{j=0}^{q+m+1}\frac{D_{q+m+1,j}}{V_{\lambda}^{j}},

where we employed the rule (3) for the first membership, the rule (1) for the second equality as proving in (A.5), the rule (5) for the inclusion and the rule (3) for the final equality. Therefore, the inductive claim in (A.2) is proved. ∎

Proof of Lemma (3.3).

When n=0n=0, the conclusion of the Lemma follows from (3.4) and the fact that ψ1(1)=λ1iVλ14\psi_{-1}^{(1)}=\lambda^{-1}iV_{\lambda}^{\frac{1}{4}}. When n>0n>0, thanks to (3.5) and (3.8), the reminder λ,n\mathcal{R}_{\lambda,n} can be written in the following way

λ,n=\displaystyle\mathcal{R}_{\lambda,n}= k=n34n4λkϕk+3\displaystyle\sum_{k=n-3}^{4n-4}\lambda^{-k}\phi_{k+3}
=\displaystyle= k=n34n4λk(ψk(4)+4α1+α2=k1α1,α2n1ψα1(1)ψα2(3)+3α1+α2=k1α1,α2n1ψα1(2)ψα2(2)\displaystyle\sum_{k=n-3}^{4n-4}\lambda^{-k}\left(-\psi_{k}^{(4)}+4\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=k\\ -1\leq\alpha_{1},\alpha_{2}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}+3\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=k\\ -1\leq\alpha_{1},\alpha_{2}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)}\right.
6α1+α2+α3=k1α1,α2,α3n1ψα1(1)ψα2(1)ψα3(2)+α1+α2+α3+α4=k1α1,α2,α3,α4n1ψα1(1)ψα2(1)ψα3(1)ψα4(1)).\displaystyle\left.\hskip 56.9055pt-6\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}+\alpha_{3}=k\\ -1\leq\alpha_{1},\alpha_{2},\alpha_{3}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(2)}+\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=k\\ -1\leq\alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(1)}\psi_{\alpha_{4}}^{(1)}\right).

By using the fact that ψk=0\psi_{k}=0 for all k>n1k>n-1, we can reduce the indices of the sums as follows

λ,n=\displaystyle\mathcal{R}_{\lambda,n}= k=n3n1λkψk(4)S1+4k=n32n2λkα1+α2=k1α1,α2n1ψα1(1)ψα2(3)S2+3k=n32n2λkα1+α2=k1α1,α2n1ψα1(2)ψα2(2)S3\displaystyle-\underbrace{\sum_{k=n-3}^{n-1}\lambda^{-k}\psi_{k}^{(4)}}_{S_{1}}+4\underbrace{\sum_{k=n-3}^{2n-2}\lambda^{-k}\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=k\\ -1\leq\alpha_{1},\alpha_{2}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(3)}}_{S_{2}}+3\underbrace{\sum_{k=n-3}^{2n-2}\lambda^{-k}\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=k\\ -1\leq\alpha_{1},\alpha_{2}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(2)}\psi_{\alpha_{2}}^{(2)}}_{S_{3}}
6k=n33n3λkα1+α2+α3=k1α1,α2,α3n1ψα1(1)ψα2(1)ψα3(2)S4+k=n34n4λkα1+α2+α3+α4=k1α1,α2,α3,α4n1ψα1(1)ψα2(1)ψα3(1)ψα4(1)S5.\displaystyle-6\underbrace{\sum_{k=n-3}^{3n-3}\lambda^{-k}\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}+\alpha_{3}=k\\ -1\leq\alpha_{1},\alpha_{2},\alpha_{3}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(2)}}_{S_{4}}+\underbrace{\sum_{k=n-3}^{4n-4}\lambda^{-k}\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}=k\\ -1\leq\alpha_{1},\alpha_{2},\alpha_{3},\alpha_{4}\leq n-1\end{subarray}}\psi_{\alpha_{1}}^{(1)}\psi_{\alpha_{2}}^{(1)}\psi_{\alpha_{3}}^{(1)}\psi_{\alpha_{4}}^{(1)}}_{S_{5}}.

For each k[[1,n1]]k\in[[-1,n-1]], Lemma 3.2 shows us that the maximal possible derivative of VV in ψk(1)\psi_{k}^{(1)} is k+1k+1. Therefore, the maximal possible derivative of VV in λ,n\mathcal{R}_{\lambda,n} that may appear in ψn1(4)\psi_{n-1}^{(4)} is n+3n+3.

Using Lemma 3.2, we obtain

S1k=n3n11Vλk4j=0k+4Dk+4,jVλj=k=021Vλk+n34j=0k+n+1Dk+n+1,jVλj.\displaystyle S_{1}\in\sum_{k=n-3}^{n-1}\frac{1}{V_{\lambda}^{\frac{k}{4}}}\sum_{j=0}^{k+4}\frac{D_{k+4,j}}{V_{\lambda}^{j}}=\sum_{k=0}^{2}\frac{1}{V_{\lambda}^{\frac{k+n-3}{4}}}\sum_{j=0}^{k+n+1}\frac{D_{k+n+1,j}}{V_{\lambda}^{j}}.

In order to estimate SmS_{m} for m[[2,5]]m\in[[2,5]], we do the same trick as proving (A.6). For example, to estimate S2S_{2}, we do as follows

S2\displaystyle S_{2}\in k=n32n2λkα1+α2=k1α1,α2n1(λα1Vλα14j1=0α1+1Dα1+1,j1Vλj1)(λα2Vλα24j2=0α2+3Dα2+3,j2Vλj2)\displaystyle\sum_{k=n-3}^{2n-2}\lambda^{-k}\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=k\\ -1\leq\alpha_{1},\alpha_{2}\leq n-1\end{subarray}}\left(\frac{\lambda^{\alpha_{1}}}{V_{\lambda}^{\frac{\alpha_{1}}{4}}}\sum_{j_{1}=0}^{\alpha_{1}+1}\frac{D_{\alpha_{1}+1,j_{1}}}{V_{\lambda}^{j_{1}}}\right)\left(\frac{\lambda^{\alpha_{2}}}{V_{\lambda}^{\frac{\alpha_{2}}{4}}}\sum_{j_{2}=0}^{\alpha_{2}+3}\frac{D_{\alpha_{2}+3,j_{2}}}{V_{\lambda}^{j_{2}}}\right)
=\displaystyle= k=n32n21Vλk4α1+α2=k1α1,α2n1j=0α1+α2+4q1+q2=jDα1+1,q1Vλq1Dα2+3,q2Vλq2\displaystyle\sum_{k=n-3}^{2n-2}\frac{1}{V_{\lambda}^{\frac{k}{4}}}\sum_{\begin{subarray}{c}\alpha_{1}+\alpha_{2}=k\\ -1\leq\alpha_{1},\alpha_{2}\leq n-1\end{subarray}}\sum_{j=0}^{\alpha_{1}+\alpha_{2}+4}\sum_{q_{1}+q_{2}=j}\frac{D_{\alpha_{1}+1,q_{1}}}{V_{\lambda}^{q_{1}}}\frac{D_{\alpha_{2}+3,q_{2}}}{V_{\lambda}^{q_{2}}}
\displaystyle\subset k=n32n21Vλk4j=0k+4Dk+4,jVλj=k=0n+11Vλk+n34j=0k+n+1Dk+n+1,jVλj.\displaystyle\sum_{k=n-3}^{2n-2}\frac{1}{V_{\lambda}^{\frac{k}{4}}}\sum_{j=0}^{k+4}\frac{D_{k+4,j}}{V_{\lambda}^{j}}=\sum_{k=0}^{n+1}\frac{1}{V_{\lambda}^{\frac{k+n-3}{4}}}\sum_{j=0}^{k+n+1}\frac{D_{k+n+1,j}}{V_{\lambda}^{j}}.

Similarly, we obtain

S3k=0n+11Vλk+n34j=0k+n+1Dk+n+1,jVλj,\displaystyle S_{3}\in\sum_{k=0}^{n+1}\frac{1}{V_{\lambda}^{\frac{k+n-3}{4}}}\sum_{j=0}^{k+n+1}\frac{D_{k+n+1,j}}{V_{\lambda}^{j}},
S4k=02n1Vλk+n34j=0k+n+1Dk+n+1,jVλj,\displaystyle S_{4}\in\sum_{k=0}^{2n}\frac{1}{V_{\lambda}^{\frac{k+n-3}{4}}}\sum_{j=0}^{k+n+1}\frac{D_{k+n+1,j}}{V_{\lambda}^{j}},
S5k=03n11Vλk+n34j=0k+n+1Dk+n+1,jVλj.\displaystyle S_{5}\in\sum_{k=0}^{3n-1}\frac{1}{V_{\lambda}^{\frac{k+n-3}{4}}}\sum_{j=0}^{k+n+1}\frac{D_{k+n+1,j}}{V_{\lambda}^{j}}.

Therefore, thanks to the rules (1) (for j=0j=0 and r1r\geq 1, Dr,j={0}D_{r,j}=\{0\}) and (3), the conclusion of Lemma 3.3 is obtained. ∎

Acknowledgement

I would like to thank Professor David Krejčiřík for introducing and encouraging me to study this great subject. Especially, I am very grateful to him for giving me many precious opportunities to continue my research career. This project was supported by the EXPRO grant number 20-17749X of the Czech Science Foundation (GAČR).

References

  • [1] A. Arifoski and P. Siegl. Pseudospectra of the damped wave equation with unbounded damping. SIAM J. Math. Anal., 52(2):1343–1362, 2020.
  • [2] J. M. Arrieta, F. Ferraresso, and P. D. Lamberti. Spectral analysis of the biharmonic operator subject to Neumann boundary conditions on dumbbell domains. Integral Equations Operator Theory, 89(3):377–408, 2017.
  • [3] Y. Bonthonneau and N. Raymond. WKB constructions in bidimensional magnetic wells. Math. Res. Lett., 27(3):647–663, 2020.
  • [4] L. S. Boulton. Non-self-adjoint harmonic oscillator, compact semigroups and pseudospectra. J. Operator Theory, 47(2):413–429, 2002.
  • [5] B. Colbois and L. Provenzano. Neumann eigenvalues of the biharmonic operator on domains: geometric bounds and related results. arXiv:1907.02252 [math.SP], 2019.
  • [6] L. Cossetti and D. Krejčiřík. Absence of eigenvalues of non-self-adjoint Robin Laplacians on the half-space. Proc. Lond. Math. Soc. (3), 121(3):584–616, 2020.
  • [7] E. B. Davies. Semi-classical states for non-self-adjoint Schrödinger operators. Comm. Math. Phys., 200(1):35–41, 1999.
  • [8] E. B. Davies. Linear operators and their spectra. Cambridge University Press, 2007.
  • [9] A. Enblom. Estimates for eigenvalues of Schrödinger operators with complex-valued potentials. Lett. Math. Phys., 106(2):197–220, 2016.
  • [10] F. Ferraresso and L. Provenzano. On the eigenvalues of the biharmonic operator with Neumann boundary conditions on a thin set. arXiv:2108.03969 [math.SP], 2021.
  • [11] Y. Guedes Bonthonneau, T. Nguyen Duc, N. Raymond, and S. Vũ Ngọc. Magnetic WKB constructions on surfaces. Rev. Math. Phys., 33(7):Paper No. 2150022, 41, 2021.
  • [12] B. Helffer. Spectral theory and its applications. Cambridge University Press, New York, 2013.
  • [13] R. Henry. Spectral instability for the complex Airy operator and even non-selfadjoint anharmonic oscillators. J. Spectr. Theory, 4:349–364, 2014.
  • [14] R. Henry. Spectral projections of the complex cubic oscillator. Ann. H. Poincaré, 15:2025–2043, 2014.
  • [15] R. Henry and D. Krejčiřík. Pseudospectra of the Schrödinger operator with a discontinuous complex potential. J. Spectr. Theory, 7(3):659–697, 2017.
  • [16] M. Hitrik, J. Sjöstrand, and J. Viola. Resolvent estimates for elliptic quadratic differential operators. Anal. PDE, 6:181–196, 2013.
  • [17] A. Hulko. On the number of eigenvalues of the biharmonic operator on 3\mathbb{R}^{3} perturbed by a complex potential. Rep. Math. Phys., 81(3):373–383, 2018.
  • [18] O. O. Ibrogimov, D. Krejčiřík, and A. Laptev. Sharp bounds for eigenvalues of biharmonic operators with complex potentials in low dimensions. Math. Nachr., 294(7):1333–1349, 2021.
  • [19] D. Krejčiřík. Complex magnetic fields: an improved Hardy-Laptev-Weidl inequality and quasi-self-adjointness. SIAM J. Math. Anal., 51(2):790–807, 2019.
  • [20] D. Krejčiřík and T. Nguyen Duc. Pseudomodes for non-self-adjoint dirac operators. arXiv:2011.09433v2 [math-ph], 2020.
  • [21] D. Krejčiřík and P. Siegl. Pseudomodes for Schrödinger operators with complex potentials. J. Funct. Anal., 276(9):2856–2900, 2019.
  • [22] D. Krejčiřík, P. Siegl, M. Tater, and J. Viola. Pseudospectra in non-Hermitian quantum mechanics. J. Math. Phys., 56(10):103513, 32, 2015.
  • [23] A. Laptev and O. Safronov. Eigenvalue estimates for Schrödinger operators with complex potentials. Comm. Math. Phys., 292(1):29–54, 2009.
  • [24] B. Mityagin, P. Siegl, and J. Viola. Concentration of eigenfunctions of schroedinger operators. arXiv:1910.10048 [math.SP], 2020.
  • [25] R. Novák. On the pseudospectrum of the harmonic oscillator with imaginary cubic potential. Int. J. Theor. Phys., 54:4142–4153, 2015.
  • [26] K. Pravda-Starov. A general result about the pseudo-spectrum of Schrödinger operators. Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci., 460(2042):471–477, 2004.
  • [27] L. N. Trefethen and M. Embree. Spectra and pseudospectra. Princeton University Press, Princeton, NJ, 2005. The behavior of nonnormal matrices and operators.