Processing math: 85%
Research article Special Issues

Bounds on eigenvalues of perturbed Lamé operators with complex potentials

  • Received: 06 August 2021 Accepted: 20 September 2021 Published: 28 September 2021
  • Several recent papers have focused their attention in proving the correct analogue to the Lieb-Thirring inequalities for non self-adjoint operators and in finding bounds on the distribution of their eigenvalues in the complex plane. This paper provides some improvement in the state of the art in this topic. Precisely, we address the question of finding quantitative bounds on the discrete spectrum of the perturbed Lamé operator of elasticity Δ+V in terms of Lp-norms of the potential. Original results within the self-adjoint framework are provided too.

    Citation: Lucrezia Cossetti. Bounds on eigenvalues of perturbed Lamé operators with complex potentials[J]. Mathematics in Engineering, 2022, 4(5): 1-29. doi: 10.3934/mine.2022037

    Related Papers:

    [1] Biagio Cassano, Lucrezia Cossetti, Luca Fanelli . Spectral enclosures for the damped elastic wave equation. Mathematics in Engineering, 2022, 4(6): 1-10. doi: 10.3934/mine.2022052
    [2] Isabeau Birindelli, Kevin R. Payne . Principal eigenvalues for k-Hessian operators by maximum principle methods. Mathematics in Engineering, 2021, 3(3): 1-37. doi: 10.3934/mine.2021021
    [3] Stefano Biagi, Serena Dipierro, Enrico Valdinoci, Eugenio Vecchi . A Hong-Krahn-Szegö inequality for mixed local and nonlocal operators. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023014
    [4] Antonio Iannizzotto, Giovanni Porru . Optimization problems in rearrangement classes for fractional $ p $-Laplacian equations. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002
    [5] Roberto Feola, Felice Iandoli, Federico Murgante . Long-time stability of the quantum hydrodynamic system on irrational tori. Mathematics in Engineering, 2022, 4(3): 1-24. doi: 10.3934/mine.2022023
    [6] G. Gaeta, G. Pucacco . Near-resonances and detuning in classical and quantum mechanics. Mathematics in Engineering, 2023, 5(1): 1-44. doi: 10.3934/mine.2023005
    [7] Marco Cirant, Kevin R. Payne . Comparison principles for viscosity solutions of elliptic branches of fully nonlinear equations independent of the gradient. Mathematics in Engineering, 2021, 3(4): 1-45. doi: 10.3934/mine.2021030
    [8] Giuseppe Procopio, Massimiliano Giona . Bitensorial formulation of the singularity method for Stokes flows. Mathematics in Engineering, 2023, 5(2): 1-34. doi: 10.3934/mine.2023046
    [9] Pier Domenico Lamberti, Michele Zaccaron . Spectral stability of the curlcurl operator via uniform Gaffney inequalities on perturbed electromagnetic cavities. Mathematics in Engineering, 2023, 5(1): 1-31. doi: 10.3934/mine.2023018
    [10] Konstantinos T. Gkikas . Nonlinear nonlocal equations involving subcritical or power nonlinearities and measure data. Mathematics in Engineering, 2024, 6(1): 45-80. doi: 10.3934/mine.2024003
  • Several recent papers have focused their attention in proving the correct analogue to the Lieb-Thirring inequalities for non self-adjoint operators and in finding bounds on the distribution of their eigenvalues in the complex plane. This paper provides some improvement in the state of the art in this topic. Precisely, we address the question of finding quantitative bounds on the discrete spectrum of the perturbed Lamé operator of elasticity Δ+V in terms of Lp-norms of the potential. Original results within the self-adjoint framework are provided too.



    This paper is devoted to providing bounds on the location of discrete eigenvalues of operators of the form

    Δ+V (1.1)

    acting on [L2(Rd)]d, the Hilbert space of vector fields with components in L2(Rd). The operator (1.1) is introduced in a standard way as an m-accreative operator obtained as a form sum of the Lamé operator Δ with domain [H1(Rd)]d and a relatively form-bounded potential V. We refer to Appendix A for details. The free Lamé operator Δ is defined as the (self-adjoint) Friedrichs extension of the minimal operator

    Δu:=μΔu(λ+μ) div u,u[C0(Rd)]d, (1.2)

    with Δ:=ΔICd×d the diagonal matrix with the Laplace operator on the diagonal and where  div  is the standard divergence operator (refer to [24,Sec. 2.1] for details). Here we slightly abused the notation using the same symbol Δ both for the symmetric operator (1.2) and its self-adjoint realisation. The two material-dependent constants λ,μR (tipically called Lamé's coefficients) are assumed to satisfy the standard conditions

    μ>0,λ+2μ>0, (1.3)

    which guarantee the strong ellipticity of Δ (refer to [24,Sec. 2.2] for the a brief summary in matter of ellipticity for systems of second order differential operators). V represents the operator of multiplication by VICd×d, with V:RdC. Being V complex-valued leads to a (possibly) non self-adjoint setting.

    The distribution of eigenvalues of self-adjoint operators has been intensively studied for several decades and nowadays the usage of powerful techniques such as spectral theorem and variational principles have become the standard approach for addressing this issue. However, as these tools are no longer available in a non self-adjoint setting, the generalization of spectral bounds to the non self-adjoint framework is not straightforward and requires a diverse strategy.

    Nevertheless, a systematic, albeit recent, approach to the study of eigenvalue estimates for perturbed operators with complex-valued potentials, in particular for Schrödinger operators Δ+V, has been successfully developed and has already a bibliography. It is very well known consequence of Sobolev inequalities that if V is real-valued then the distance from the origin of every eigenvalue z lying in the negative semi-axis (discrete eigenvalue) can be bounded in terms of Lp-norm of the potential (see [49,57] and [13] for a more recent improvement), more precisely the following bound

    |z|γCγ,d||V||γ+d2Lγ+d2(Rd) (1.4)

    holds true for every γ1/2 if d=1, γ>0 if d=2 and γ0 if d3, with Cγ,d independent of V.

    In 2001, Abramov, Aslanyan and Davies [1] generalized the previous result to complex-valued potentials showing that every discrete eigenvalue z of the one-dimensional Schrödinger operator d2/dx2+V, i.e., zC[0,), lies in the complex plane within a 1/4||V||2L1(R) radius of the origin. In order to overcome the lack of the aforementioned self-adjoint-based tools, the authors introduced a strategy built on the Birman-Schwinger principle (see, the pioneering Birman's paper [7] or the recent work [46] for a systematic exposition of abstract Birman-Schwinger principles and some interesting explicit applications), which has permeated all the subsequent works on this topic. Ten years later Frank [40] went beyond the one dimensional restriction proving the validity of (1.4) for any d2 and 0<γ1/2, covering also the endpoint case γ=0 if d3. His notable accomplishment derives from replacing a pointwise bound for the Green function of d2/dx2z, used in [1] and highly sensitive of the one dimensional framework, with the much deeper uniform resolvent estimates due to Kenig, Ruiz and Sogge [50] (see below for further details). The result in [40] partially proved the conjecture in [52] according to which (1.4) holds for any d2 and 0<γd/2, leaving opened only the case 1/2<γ<d/2. This range was covered later by Frank and Simon in [45] for radial potential. In the same work the general case was investigated too. In these respects the authors provided the construction of a sequence of real-valued potentials Vn with ||Vn||p0 for d2 and for any p>(d+1)/2 such that Δ+Vn has eigenvalue 1. Even if this does not disprove the conjecture, as it is stated for discrete eigenvalues only, the result is still relevant in view of the recent generalization in [45] of Frank's work [40] to positive (embedded) eigenvalues. Recent related results can be also found in [41].

    In passing, let us stress that the extensive bibliography devoted to spectral bounds in a non self-adjoint context that the appearance of the cited work [1] stimulated, did not remain long confined to the class of Schrödinger operators and more generally to second order differential operators. Indeed subsequent investigations have shown the robustness of the approach introduced in [1] by fruitfully testing it to several other models. Without attempting to be exhaustive we mention [48] for an adaptation to the discrete Schrödinger setting, see also [51] where matrix-valued damped wave operators are concerned (see also [15] for the elasticity counterpart). Lower order operators, such as Dirac or fractional Schrödinger models, are investigated in [18,26,28,30,31,36,58] (see also [27,39]) and in [17] respectively in the continuous and discrete scenario; as for higher order operators refer to [34,47]. Associated spectral stability results obtained with different techniques and related tools can be found in [3,10,11,16,25,35,44].

    As a final remark, we mention that since the turn of the millennium, when the interest on non self-adjoint operators started to take hold, this subject has exhibited a rapid development and nowadays the analysis of such Hamiltonians spreads out to cover several diverse branches of spectral theory. Just with the aim of giving a (non comprehensive) overview of the several questions posed in this context, focusing on Schrödinger operators, we cite [8,19,43,62] in matter of existence and estimate on the number of eigenvalues, [9,32,33,41,42,52] as regards with the non self-adjoint counterpart of Lieb-Thirring inequalities and we refer to the aformentioned quoted works for the topic related to eigenvalues bounds for non self-adjoint Schrödinger operators.

    The main aim of our paper is to investigate on spectral properties in the elasticity setting, specifically providing bounds on the distribution of discrete (possibly complex) eigenvalues of Lamé operators (1.2) in terms of Lp-norm of the potential. In other words we want to show the validity of a suitable analogous of (1.4) in this diverse framework.

    A first motivation to our purpose is the trivial observation that in d=1 our operator Δ turns into a scalar differential operator and, even more relevant, it is simply a multiple of the Laplacian, more precisely

    Δ:=μd2dx2(λ+μ)d2dx2=(λ+2μ)d2dx2. (1.5)

    Therefore, by virtue of the aforementioned analogous theorem in [1] (cfr Theorem 4) for non self-adjoint Schrödinger operators d2/dx2+V, the following result comes as no surprise and it is indeed its straightforward consequence.

    Theorem 1.1. Let d=1 and assume that VL1(R). Then any eigenvalue zC[0,) of the perturbed Lamé operator Δ+V satisfies

    |z|1/212λ+2μ||V||L1(R). (1.6)

    The validity of Theorem 1.1, compared with Theorem 4 in [1], together with the aforementioned extensions to higher dimensional Schrödinger operators, motivate the question of whether an estimate similar to (1.6) holds true in d2.

    As already said, starting from the celebrated paper of Abramov, Aslanyan and Davies [1], the usage of the Birman-Schwinger principle has been recognized as a crucial tool to get this type of bounds in a non self-adjoint setting of problem. Roughly speaking this permits to re-phrase conveniently the eigenvalues problem associated to a perturbed operator, say H0+V, in terms of the eigenvalues problem of an integral operator suitably related to the latter. Precisely, the following proposition holds true.

    Proposition 1.1 (Birman-Schwinger principle). Let zσ(H0). Then

    zσp(H0+V)1σp(V1/2(H0z)1|V|1/2),

    with V1/2:=|V|1/2 sgn (V). (If zC, then  sgn (z) represents the complex signum function defined by  sgn (z):=z/|z| if z0 and  sgn (0):=0).

    The formal statement of Proposition 1.1 can be made rigorous under suitable hypotheses on H0 and V. We refer again to the recent work [46,Thm. 6,Thm.7] by Hansmann and Krejčiřík for a rigorous exposition of abstract Birman-Schwinger principles.

    In the context of elasticity, H0 is replaced by the Lamé operator Δ, therefore the corresponding Birman-Schwinger operator has the form V1/2(Δz)1|V|1/2 and makes sense if zC[0,) as σ(Δ)=[0,). This leads to the need for an explicit expression of the resolvent operator associated with Δ. In this regards it turns out that (Δz)1 has a favorable form, i.e.,

    (Δz)1g=1μ(Δzμ)1Pg+1λ+2μ(Δzλ+2μ)1(IP)g, (1.7)

    where P is the so-called Leray projection operator, customarily used in elasticity to decompose [L2(Rd)]d vector fields into a divergence-free component plus a gradient (refer to the preliminary section for further details).

    This expression shows that, as soon as the decomposition, also known as Helmholtz decomposition, g=Pg+(IP)g is operated, the resolvent operator (Δz)1 splits into a sum of two vector-valued resolvent operators associated with the Laplacian acting, respectively, on the components Pg and (IP)g of g. This fact, together with the validity of the corresponding results for Schrödinger, strongly suggests a positive answer to our question of whether Theorem 1.1 extends to higher dimensions and indeed it is confirmed by the following result proved in this paper.

    Theorem 1.2. Let d2 and assume that VLγ+d2(Rd), with 0<γ1/2 if d=2 and 0γ1/2 if d3. Then any eigenvalue zC[0,) of the perturbed Lamé operator Δ+V satisfies

    |z|γCγ,d,λ,μ||V||γ+d2Lγ+d2(Rd), (1.8)

    with a constant Cγ,d,λ,μ independent of V.

    Incidentally, observe that if d3 and γ=0, the previous theorem provides a sufficient condition which guarantees absence of discrete eigenvalues of Δ+V with VLd2(Rd). More specifically, if

    C0,d,λ,μ||V||d2Ld2(Rd)<1, (1.9)

    then the discrete spectrum σd(Δ+V) is empty. In comparison with this result, we should mention a prior work of the present author [24] also related to the problem of establishing sufficient conditions which disprove presence of eigenvalues. In [24], with a completely different approach based on the multipliers method as previously applied to Schrödinger operators by Fanelli, Krejčiřík and Vega in [37,38], total absence of eigenvalues, i.e., absence of both discrete and embedded eigenvalues, of Δ+V was proved under the following Hardy-type subordination

    Rd|x|2|V(x)|2|u|2dxΛ2Rd|u|2dx,u[H1(Rd)]d,d3, (1.10)

    where Λ is a suitable small constant (see condition (4) in [24]).

    As (1.9), condition (1.10) is intrinsically a smallness condition, it is true, on the other hand it is satisfied by potentials with quite rough local singularities, e.g., |x|2, which, instead, are ruled out by the Lp-type condition in Theorem 1.2.

    In attempt of including potentials with local stronger singularities, such as inverse-square type, we generalize Theorem 1.2 by measuring the size of the potential in (1.8) in terms of less restrictive norms. Specifically, as first generalization, we consider potentials in the Morrey-Campanato class Lα,p(Rd) which is defined for α>0 and 1pd/α by

    ||V||Lα,p(Rd):=supx,rrα(rdBr(x)|V(x)|pdx)1p<.

    In passing, notice that 1/|x|αLd/α,(Rd)Lα,p, for α>0 (we emphasize particularly the case α=2) and 1p<d/α, however 1/|x|αLd/α=Lα,d/α.

    More precisely we shall prove the following theorem.

    Theorem 1.3. Let d2 and assume that VLα,p(Rd) with (d1)(2γ+d)/2(d2γ)<pγ+d/2 and let 0<γ1/2 if d=2 and 0γ1/2 if d3. Then any eigenvalue zC[0,) of the perturbed Lamé operator Δ+V satisfies

    |z|γCγ,p,d,λ,μ||V||γ+d2Lα,p(Rd), (1.11)

    with α=2d/(2γ+d) and a constant Cγ,p,d,λ,μ independent of V.

    As a byproduct, in higher dimensions, the previous theorem provides a sufficient condition to guarantee absence of discrete eigenvalues. More precisely, the following corollary is immediate consequence of Theorem 1.3.

    Corollary 1.1. Let d3, (d1)/2<pd/2 and assume VL2,p(Rd) and

    C0,p,d,λ,μ||V||d2L2,p(Rd)<1,

    with C0,p,d,λ,μ as in Theorem 1.3 when γ=0. Then the perturbed Lamé operator Δ+V has no eigenvalue in C[0,).

    Observe that Theorem 1.3 does extend Theorem 1.2, indeed from

    Lγ+d2(Rd)=L2d2γ+d,γ+d2(Rd)L2d2γ+d,p(Rd),

    which holds true for 1pγ+d/2, in particular it follows that

    ||V||L2d2γ+d,p(Rd)Cγ,p,d||V||Lγ+d2(Rd),

    which immediately gives (1.8) as a consequence of (1.11).

    Actually our investigation goes even further providing eigenvalues bounds of the type (1.8) in terms of potentials belonging to the Kerman-Saywer space KSα(Rd), which is defined for 0<α<d by

    ||V||KSα(Rd):=supQ(Q|V(x)|dx)1QQ|V(x)||V(y)||xy|dαdxdy<,

    where the supremum is taken over all dyadic cubes Q in Rd.

    As Lα,p(Rd)KSα(Rd) if p1 (see Section 2 in [5]) it is true that the Kerman-Sayer class is wider than the Morrey-Campanato class. On the other hand, it turns out that assuming solely VKSα is not enough to get bound (1.11) with ||V||Lα,p(Rd) replaced by ||V||KSα(Rd). Indeed, additionally, we will ask the potential V to belong to the Muckenhoupt A2(Rd) class of weights which is defined, in general, for 1<p< as the set of measurable non-negative function w such that

    Qp(w):=supQ(1|Q|Qw(x)dx)(1|Q|Qw(x)1p1dx)p1C, (1.12)

    where Q is any cube in Rd and C is a constant independent of Q.

    More precisely, we shall prove the following result.

    Theorem 1.4. Let d2 let 1/3γ<1/2 if d=2 and 0γ<1/2 if d3. Let α=2d2γ+dβ and β=(d+2γ)(d1)2(d2γ) and assume that VβKSα(Rd). If moreover VA2(Rd), then any eigenvalue zC[0,) of the perturbed Lamé operator Δ+V satisfies

    |z|γCγ,d,λ,μQ2(|V|)2γ+d||Vβ||1β(γ+d2)KSα, (1.13)

    with a constant Cγ,d,λ,μ independent of V.

    As a consequence of the previous result, one gets the following corollary on absence of discrete eigenvalues.

    Corollary 1.2. Let d3 and assume Vd12KSd1(Rd), VA2(Rd) and

    C0,d,λ,μQ2(|V|)d||Vd12||dd1KSd1<1,

    with C0,d,λ,μ as in Theorem 1.4 when γ=0. Then the perturbed Lamé operator Δ+V has no eigenvalue in C[0,).

    It is worth comparing Theorem (1.4) with the analogous result in [53] (cfr. Theorem 1.1) for Schrödinger operators. Here, the bound (1.13) was obtained without the additional assumption VA2(Rd), then showing a peculiar feature of the Lamé operator.

    Roughly speaking, the philosophy is that thanks to the Helmholtz decomposition which, as shown in (1.7), makes the resolvent operator (Δz)1 "behave" like a sum of two resolvent (Δz)1, at first we can perform our analysis at the level of the much more investigated Schrödinger operators, estimating the two components in (1.7) separately. Then, in order to get bound (1.11) and (1.13), respectively, these two pieces have to be recombined together on weighted L2-spaces and it is at this step that the A2 assumption comes into play. We refer to Section 4 for greater details.

    Marginally, observe that, even though the aforementioned strategy underpins the proof of both (1.11) and (1.13), good property of the Lα,p class allowed to drop the A2 assumption in Theorem 1.3 (see Lemma 2.8 below).

    It is very well known fact that distinguishing whether Δ+V has finite (possibly empty) or infinite discrete spectrum depends on the large x fall-off of the potential. More specifically, it is mainly consequence of the uncertainty principle, quantified by the Hardy inequality

    Δ(d2)241|x|2

    that the borderline is marked by an inverse-square type behavior at infinity.

    In scattering theory and in particular in matter of determining existence of wave operators, again the large x behavior of the potential plays a central role, in this context the threshold is given by a Coulomb-type asymptotic decay and positive results require |x|α with α>1. Observe that our previous theorems essentially restrict to |x|α decay with α>2d/d+1 and therefore are not fully satisfactory in the perspective of their possible application to stationary scattering theory. The gap is filled in the following result by using a suitable interpolation argument.

    Theorem 1.5. Let d2, γ>12 and α>γ12. Let q=2γ+(d1)/2 and assume VLq(x2αdx). Then any eigenvalue zC[0,) of the perturbed Lamé operator Δ+V satisfies

    |z|γCγ,α,d,λ,μ||V||qLq(x2αdx),

    with a constant Cγ,α,d,λ,μ independent of V.

    Here we used the notation x:=(1+|x|2)1/2, moreover, given a measurable function w, Lp(wdx) stands for the w-weighted Lp space on Rd with measure w(x)dx.

    The rest of the paper is organized as follows: In Section 2, in attempt of making the paper sufficiently self-contained, we collect some preliminary facts on the Helmholtz decomposition. Here, some properties of Lamé operator are provided too. Among them, particular emphasis will be given to uniform estimates for the resolvent operator (Δz)1 which will represent the main ingredient in the proof of our aforementioned results.

    Contrarily to the much more investigated Schrödinger operator, up to our knowledge, eigenvalue bounds of the form (1.4) for the perturbed Lamé operator are unknown even in the self-adjoint situation. Although in this case the proof of (1.4) follows almost verbatim the instead-well-known one for Schrödinger, we decided to dedicate Section 3 to prove it anyhow. The advantage of this choice comes out in the possibility of explicitly showing the deep differences and difficulties that arise passing from the self-adjoint to the non self-adjoint framework which, instead, is fully analyzed in Section 4. In particular, Section 4 is devoted to the proof of the main results stated in the introduction. We mention that in the meantime the publication process for this paper was finalized, the results presented here were generalised in [14] to cover matrix-valued potential and, more importantly, embedded eigenvalues as well.

    Notations.

    In this paper both scalar and vector-valued functions are considered. In attempt of produce no confusion, we clarify here that notation like f,g,u are reserved for vector-fields, instead ϕ,ψ are set aside for scalar functions.

    When the letter adopted for denoting a vector field contains already a subscript, the standard subscript notation j for the j-th component will be replaced by the superscript (j), e.g., the j-th component of the vector field uS is indicated by u(j)S.

    We use the following definition for the Lp- norm of a vector field u[Lp(Rd)]d:

    ||u||[Lp(Rd)]d:=(dj=1||uj||pLp(Rd))1p.

    In order to lighten the presentation, in the following we usually abbreviate both ||||Lp(Rd) and ||||[Lp(Rd)]d, the Lp-norm of scalar and vector-valued functions, respectively, with ||||p. This, in general, could create some ambiguity, on the other hand, it will be clear from the context and the notation used there if ||||p stands for one or the other norm.

    Again, the notation , will denote both ,L2(Rd) and ,[L2(Rd)]d, where the latter extends in an obvious way the usual definition of the former, namely given f,g[L2(Rd)]d, one defines

    f,g[L2(Rd)]d:=dj=1fj,gjL2(Rd).

    Let E and F be two Banach spaces and let T:EF be a bounded linear operator from E into F. The notation ||T||EF will be used to denote the operator norm of T.

    This section is concerned with recalling some properties connected with the Helmholtz decomposition together with stating and proving some related consequences on Lamé operators that will strongly enter the proof of our results later. If the first part wants to be just a remainder of very well known results on the Helmholtz decomposition and therefore can be safely skipped by any reader already familiar with this topic, the subsequent subsections, namely Subsection 2.2 and Subsection 2.3, represent an important part of the paper. More specifically, it is there that uniform resolvent estimates for the resolvent operator (Δz)1 are provided which are important in their own right.

    Theorem 2.1 (Helmholtz decomposition). Let d2 and let ΩRd be either an open, bounded, simply connected, Lipschitz domain or the entire Rd. Then any square-integrable vector field f=(f(1),f(2),,f(d))[L2(Ω)]d can be uniquely decomposed as

    f=fS+fP,

    where fS is a divergence-free vector field with null normal derivative and fP is a gradient. Moreover the two components are orthogonal in a L2- sense. Specifically, the Pythagorean identity

    ||f||2[L2(Ω)]d=||fS||2[L2(Ω)]d+||fP||2[L2(Ω)]d (2.1)

    holds true.

    Proof. Though the proof simply relies on classical techniques, for sake of completeness we provide a brief sketch of the proof.

    Let f[L2(Ω)]d and let ν denote the unit outward normal vector at the boundary Ω. We consider the trivial decomposition f=fψ+ψ. In order for fψ to be the divergence-free component fS of f with null normal derivative, ψ must satisfy the boundary value problem

    {Δψ= div finΩψν=fνonΩ. (2.2)

    In passing, observe that (2.2) is a Poisson problem with Neumann boundary conditions, therefore it admits a unique solution ψH1(Ω) modulo additive constants.

    Let us now prove the uniqueness of the decomposition. Let f1,f2[L2(Ω)]d and let ψ1,ψ2H1(Ω) be such that  div f1= div f2=0 in Ω and f1ν=f2ν=0 on Ω and with the property that f=f1ψ1=f2ψ2 which gives

    f1f2=(ψ1ψ2).

    Multiplying the latter by f1f2, integrating over Ω and integrating by parts, one has

    Ω|f1f2|2dx=Ω(f1f2)(ψ1ψ2)dx=Ω div (f1f2)(ψ1ψ2)dx=0.

    It follows that f1=f2 and therefore ψ1=ψ2 up to an additive constant. To sum up, we have proved that f can be uniquely written as a sum of a divergence free vector field fS with null normal derivative on the boundary Ω and a gradient fP, where an explicit expression for fS and fP is provided by the former construction, specifically

    fS:=fψ,fP:=ψ, (2.3)

    with ψH1(Ω) unique (up to additive constant) solution of (2.2).

    At last, observe that the L2-orthogonality of fS and fP and in particular identity (2.1) are immediate consequences of the properties of the two components. This yields the proof.

    Remark 2.1. Notice that, as soon as the vector field f is more regular, for instance, say f[H1(Ω)]d as it suits our purposes, then the components fS and fP of the Helmholtz decomposition are even H1- orthogonal and in particular

    ||f||2[L2(Ω)]d=||fS||2[L2(Ω)]d+||fP||2[L2(Ω)]d.

    Now we are in position to provide the rigorous definition of the so-called Leray projection operator already mentioned in the introduction.

    Definition 2.1. Consider the setting of Theorem 2.1. Let P be the orthogonal projection of [L2(Ω)]d into the subspace of divergence-free vector fields with null normal derivative on Ω. Then P is called Leray projection operator. More precisely, for any f[L2(Ω)]d, it holds that

    Pf=P(fS+fP)=fS,

    where f=fS+fP is the Helmholtz decomposition of f as constructed in the previous theorem.

    The following result is easily proven.

    Lemma 2.1. Consider the setting of Theorem 2.1. The Leray projection operator P:[L2(Ω)]d[L2(Ω)]d is bounded.

    Proof. By definition, for any f[L2(Ω)]d, Pf:=fψ, where ψH1(Ω) is the unique solution to the boundary value problem (2.2). In particular, ψ achieves Ω|ψf|2dx=infϕH1(Ω)J(ϕ), where

    J(ϕ):=Ω|ϕf|2dx.

    Now, from infϕH1(Ω)J(ϕ)Ω|f|2dx, we immediately get ||Pf||[L2(Ω)]d||f||[L2(Ω)]d, which is the thesis.

    As showed by the following result, in the specific case Ω=Rd, the Leray projection operator has a favorable form in terms of Riesz transform R=(R1,R2,,Rd) defined for any ϕL2(Rd), in Fourier space, by

    ^Rjϕ(ξ)=iξj|ξ|ˆϕ(ξ),j=1,2,,d. (2.4)

    Lemma 2.2. Consider the setting of Definition 2.1, fix Ω=Rd and let f[L2(Rd)]d be any square-integrable vector field on Rd. Then the j-th component of Pf can be written as

    (Pf)j=fj+dk=1RjRkfk, (2.5)

    for any j=1,2,,d and where R denote the Riesz transform defined in (2.4).

    Proof. Let f[L2(Rd)]d. It follows from Theorem 2.1 and Definition 2.1 that Pf=fS=fψ, where ψ satisfies Δψ= div f. In Fourier space this yields

    ˆψ(ξ)=idk=1ξk|ξ|2^fk(ξ).

    In particular, for any j=1,2,,d, this gives

    ^jψ(ξ):=iξjˆψ(ξ)=dk=1ξjξk|ξ|2^fk(ξ)

    and so, by using the Fourier representation of the Riesz transform (2.4),

    jψ=dk=1RjRkfk, (2.6)

    for any j=1,2,,d. From the latter, one immediately gets (2.5) and this concludes the proof.

    In passing, observe that since ψ is a solution of (2.2), then it follows easily by elliptic estimates that

    ||ψ||[L2(Ω)]d||f||[L2(Ω)]d (2.7)

    (the same would follow from the boundedness of the Leray operator P (see Lemma 2.1) but in this case we would get a worse bound).

    Now, we want to show that as a consequence of the representation (2.6) of ψ in terms of the Riesz transform, we can get boundedness of type (2.7) replacing the L2-norm with suitable Lp-weighted norms.

    Notice that, in general, proving boundedness in weighted Lp-space does not come as a mere consequence of elliptic estimates as for (2.7) and, in fact, does require a more involved analysis. In our case, it will follow from special well known properties of the Riesz transform that we summarize in the following lemma.

    Lemma 2.3 (Boundedness Riesz transform). Let 1<p< and p such that 1/p+1/p=1 and let w be a weight in the Ap(Rd)-class (see definition (1.12)). Then, for any j=1,2,,d, the following bounds on the operator norms of the Riesz transform Rj hold true:

    ||Rj||LpLp=cot(π2p)=:cp,p:=max{p,p}, (2.8)
    ||Rj||Lp(w)Lp(w)cp,dQp(w)r,r:=max{1,p/p}. (2.9)

    Moreover, both the bound are sharp, i.e., the best possible bound is established.

    Proof. The proof of the sharp bound (2.8) can be found in [4] (see also [12]), inequality (2.9) can be found in [59] (see also [23]).

    Now we are in position to state and prove the aforementioned boundedness of the operator ψ. More precisely, we are interested in proving the following result.

    Lemma 2.4. Let uS(Rd) and consider ψ the unique solution to

    Δψ= div u. (2.10)

    Then, for any 1<p<, the following estimates hold true:

    ||ψ||Lpc2pd||u||Lp, (2.11)
    ||ψ||Lp(w)c2p,dQp(w)2rd||u||Lp(w), (2.12)

    where w,cp,cp,d and r are as in Lemma 2.3.

    Proof. We will prove only (2.11), the proof of (2.12) is analogous.

    Using the representation (2.6), the bound for the Riesz operator (2.8) and the Hölder inequality for discrete measures, we get

    ||ψ||Lp=(dj=1||jψ||pLp)1p(dj=1(dk=1||RjRkuk||Lp)p)1p=c2pd1pdk=1||uk||Lpc2pd||u||Lp.

    This gives (2.11). Bound (2.12) follows in the same way using (2.9).

    As a consequence of the previous lemma, we are able to prove the following (almost-) orthogonality result that we will strongly use in the future.

    Lemma 2.5. Let uS(Rd) and consider u=uS+uP the Helmholtz decomposition of u. Then, for 1<p<, the following estimates hold true:

    ||uS||Lp+||uP||Lp(1+2c2pd)||u||Lp, (2.13)
    ||uS||Lp(w)+||uS||Lp(w)(1+2c2p,dQp(w)2rd)||u||Lp(w), (2.14)

    where w,cp,cp,d and r are as in Lemma 2.3.

    Proof. The proof is a direct consequence of Lemma 2.4. We know from Theorem 2.1 (see (2.3)) that

    uS=uψ,uP=ψ,

    with ψ the unique solution to the Poisson problem (2.10). Then, it is easy to see that estimate (2.13) holds true, indeed

    ||uS||Lp+||uP||Lp||u||Lp+2||ψ||Lp(1+2c2pd)||u||Lp,

    where in the last inequality we used (2.11). As it is analogue, we skip the proof of (2.14).

    Notice that both Lemma 2.4 and Lemma 2.5 are stated for function in the Schwarz class S(Rd). By density, it is straightforward to see that they can be extended to a (weighted) Lp framework.

    As already mentioned in the introduction, the usage, as a starting point in our proofs, of an adaptation of the Birman-Schwinger principle to our elasticity context, requires a better understanding of the action of the resolvent operator (Δz)1, well-defined for any zC[0,), associated with the Lamé operator.

    The following easy consequence of Helmholtz decomposition will be useful to this end.

    Lemma 2.6. Let d2 and let f be a suitably smooth vector field sufficiently rapidly decaying at infinity. Then Δ acts on f=fS+fP as

    Δf=μΔfS(λ+2μ)ΔfP, (2.15)

    where fS is a divergence free vector field and fP a gradient.

    Now we are in position to show the validity of identity (1.7), stated in the introduction, which follows as a consequence of (2.15) together with the H1-orthogonality of the components of the Helmholtz decomposition. This is object of the following lemma.

    Lemma 2.7. Let zC[0,) and g[L2(Rd)]d. Then the identity

    (Δz)1g=1μ(Δzμ)1gS+1λ+2μ(Δzλ+2μ)1gP (2.16)

    holds true, where g=gS+gP is the Helmholtz decomposition of g.

    Proof. Given g[L2(Rd)]d, we want to obtain an explicit expression of the vector field f, defined by

    f:=(Δz)1g. (2.17)

    Observe that, since zσ(Δ)=[0,), the previous is equivalent to

    (Δz)f=g.

    Now, writing the Helmholtz decomposition of f and g, namely f=fS+fP and g=gS+gP, respectively, and using (2.15), the previous can be re-written as

    μΔfS(λ+2μ)ΔfPzfSzfP=gS+gP.

    The H1- orthogonality of the two components of the decomposition enables us to split this intertwining equation for both the two components into a system of two decoupled equations, i.e.,

    {μΔfSzfS=gS,(λ+2μ)ΔfPzfP=gP,

    or, equivalently,

    {μ(Δzμ)fS=gS,(λ+2μ)(Δzλ+2μ)fP=gP.

    Hypothesis (1.3) ensures that zμ,zλ+2μσ(Δ), hence

    fS=1μ(Δzμ)1gS,fP=1λ+2μ(Δzλ+2μ)1gP.

    Plugging these explicit expressions in f=fS+fP, from (2.17) we obtain (2.16).

    We underline that Lemma 2.7 was already proved in [6], its statement and proof were provided also here only for reader's convenience.

    As already mentioned, in a non self-adjoint framework, the unavailability of a variational characterization of the spectrum causes that Sobolev inequalities no longer suffice to prove spectral bounds. To overcome this lack, uniform resolvent estimates have been recognized as a crucial tool to fruitfully address this problem. For this reason in this subsection we shall prove uniform estimate for the operator (Δz)1 that will be the fundamental ingredient in the proof of our main results later on.

    Observe that from the representation (2.16), it is reasonable to expect that estimates for (Δz)1 should follow as a consequence of the corresponding estimates (if available) for the resolvent operator (Δz)1 associated to the Laplacian. This is true, indeed, as we will see repeatedly in this paper, the underlying strategy to treat issues concerning the Lamé operator is, actually, to operate at first at the level of the Laplacian taking advantage of the representation (2.16) once the Helmholtz decomposition is operated. On the other hand, this procedure has the main drawback of producing separated outcomes on the single components of the Helmholtz decomposition, which later must be recombined together in order to get meaningful results in application. This need of recombination requires providing suitable (almost-) orthogonality results for the Helmholtz components, roughly speaking some inequality like ||uS||+||uP||||u|| (see Lemma 2.5), which turn out to be highly non-trivial to get and require deep result from harmonic analysis and in particular from singular integrals theory (see Lemma 2.3).

    The following theorem collects the main estimates for the resolvent operator (Δz)1 that we will use for our purpose.

    Theorem 2.2 (Uniform estimates for (Δz)1). Let zC[0,). Then the following estimates for (Δz)1 hold true.

    i) Let 1<p6/5 if d=2, 2d/(d+2)p2(d+1)/(d+3) if d3 and let p such that 1/p+1/p=1. Then

    ||(Δz)1||LpLpCp,d|z|d+22+dp. (2.18)

    ii) Let α>1/2. Then

    ||(Δz)1||L2(x2α)L2(x2α)Cα,d|z|12. (2.19)

    iii) Let 3/2<α<2 if d=2, 2d/(d+1)<α2 if d3 and let (d1)/2(α1)<pd/α. Then for any non-negative function V in Lα,p(Rd)

    ||(Δz)1||L2(V1)L2(V)Cα,p,d||V||Lα,p(Rd)|z|1+α2. (2.20)

    iv) Let 3/2α<2 if d=2, d1α<d if d3 and let β=(2αd+1)/2. Then for any non-negative function V such that |V|βKSα(Rd)

    ||(Δz)1||L2(V1)L2(V)Cα,d|||V|β||1βKSα|z|αd+12αd+1. (2.21)

    Proof. Proof of (2.18) can be found in the work [50] by Kenig, Ruiz and Sogge. Estimate (2.19) is proved in the pioneering work by Agmon [2] (Lemma 4.1 there), see also [63]. Proof of (2.20) was provided by Frank in [40] (see also [20,22]). Finally, estimate (2.21) is proved by Lee and Seo in [53].

    Now we are in position to state the corresponding estimate for the resolvent operator (Δz)1.

    Theorem 2.3. Let zC[0,). Then, under the same hypotheses of Theorem 2.2, the following estimates for (Δz)1 hold true.

    ||(Δz)1||LpLpCp,d,λ,μ|z|d+22+dp. (2.22)
    ||(Δz)1||L2(x2α)L2(x2α)Cα,d,λ,μ|z|12. (2.23)
    ||(Δz)1||L2(V1)L2(V)Cα,p,d,λ,μ||V||Lα,p(Rd)|z|1+α2. (2.24)

    If, in addition, VA2(Rd), then

    ||(Δz)1||L2(V1)L2(V)Cα,d,λ,μ|||V|β||1βKSα|z|αd+12αd+1. (2.25)

    Remark 2.2. Notice that in order to prove (2.25), that is when the bound on the resolvent operator norm is measured in term of potentials in the Kerman-Saywer space, the additional assumption of V belonging to the A2 class of weights is required, on the contrary this auxiliary hypothesis is not needed when the estimate involves potentials in the Morrey-Campanato class (see (2.24)). This fact is mainly due to a good behavior of functions in the Morrey-Campanato space in relation with the Ap class of weights. This property is clarified in the following result (refer to [21], Lemma 1).

    Lemma 2.8. Let V be a non-negative function in Lα,p(Rd) with 0<α<d and 1<pd/α. If r is such that 1<r<p, then W:=(MVr)1/rA1(Rd)Lα,p(Rd), where M denotes the Hardy-Littlewood maximal operator, defined for any ψL1loc(Rd) by Mψ(x)=supr>01|Br(0)|Br(0)|ψ(xy)|dy, with Br(0) the Euclidean ball centered at the origin with radius r. Furthermore there exists a constant c independent of V such that

    ||W||Lα,pc||V||Lα,p.

    In passing, notice that V(x)W(x) for almost every xRd.

    Now we are in position to prove Theorem 2.3.

    Proof of Theorem 2.3. As already mentioned, the proof will basically rely on the interplay between the favorable representation (2.16) of (Δz)1 in terms of the resolvent of the Laplace operator together with the uniform estimates for (Δz)1 summarized in Theorem 2.2 and the orthogonality result stated in Lemma 2.5.

    Let us first consider the proof of (2.22).

    Let G be any vector-valued function in [Lp(Rd)]d. From (2.16) one easily has

    ||(Δz)1G||Lp1μ||(Δzμ)1GS||Lp+1λ+2μ||(Δzλ+2μ)1GP||Lp, (2.26)

    where G=GS+GP is the Helmholtz decomposition of G.

    We shall explicitly estimate the term involving the S-component. The analogous term for the P- component can be treated similarly. It follows from (2.18) that, for any j=1,2,,d,

    ||(Δzμ)1G(j)S||LpCp,dμd+22+dp|z|d+22+dp||G(j)S||Lp.

    This, along with the sub-additivity of the concave function |x|p, for 0<p1, and the Hölder inequality for discrete measures, gives

    ||(Δzμ)1GS||Lp=(dj=1||(Δzμ)1G(j)S||pLp)1pCp,dμd+22+dp|z|d+22+dpdj=1||G(j)S||LpCp,dμd+22+dp|z|d+22+dpd1p(dj=1||G(j)S||pLp)1p=Cp,dμd+22+dp|z|d+22+dpd1p||GS||Lp.

    The same computation for the term involving the P- component provides

    ||(Δzλ+2μ)GP||LpCp,d(λ+2μ)d+22+dp|z|d+22+dpd1p||GP||Lp.

    Plugging the previous two bounds together in (2.26), one has

    ||(Δz)1G||LpCp,d,λ,μ|z|d+22+dp(||GS||Lp+||GP||Lp),

    where Cp,d,λ,μ:=Cp,dd1pmax{μd2dp,(λ+2μ)d2dp}.

    Hence, estimate (2.22) follows immediately from the latter as a consequence of (2.13).

    Bound (2.23) can be proved with a few modifications to the argument above, namely using (2.19) and (2.14) with p=2 (notice that x2αA2(Rd)) instead of (2.18) and (2.13), respectively. Similarly bound (2.25) follows as a consequence of (2.21) and (2.14) with p=2.

    Now, we turn to the proof of (2.24).

    Let WA1(Rd)Lα,p(Rd) be as in Lemma 2.8. Observe that, since WLα,p(Rd), then (2.20) is available. Moreover, as WA1(Rd), in particular WA2(Rd) and (2.14) with p=2 is valid too. Therefore, it comes as a slight modification of the argument above proving that the following analogue of (2.25) for W holds true, namely one has

    ||(Δz)1G||L2(W)Cα,p,d,λ,μ||W||Lα,p(Rd)|z|1+α2||G||L2(W1). (2.27)

    Since V(x)W(x) almost everywhere and as ||W||Lα,p(Rd)c||V||Lα,p(Rd), using estimate (2.27) one gets

    ||(Δz)1G||L2(V)||(Δz)1G||L2(W)Cα,p,d,λ,μ||W||Lα,p(Rd)|z|1+α2||G||L2(W1)cCα,p,d,λ,μ||V||Lα,p(Rd)|z|1+α2||G||L2(V1),

    which is (2.24). This concludes the proof of the theorem.

    This section is concerned with the proof of eigenvalue bounds of the form (1.4) for the self-adjoint perturbed Lamé operator. More precisely, we shall prove the following result.

    Theorem 3.1. Let V be real-valued and and let γ1/2 if d=1, γ>0 if d=2 and γ0 if d=3. Then any negative eigenvalue z of the perturbed Lamé operator Δ+V satisfies

    |z|γCγ,d,λ,μ||V||γ+d2Lγ+d2(Rd), (3.1)

    with a constant Cγ,d,λ,μ independent of V.

    Here V denotes the negative part of V, i.e., V(x):=max{V(x),0}.

    Before providing the proof of this theorem, let us comment on the corresponding inequalities of type (3.1) for self-adjoint Schrödinger operators. As the remarks provided below then naturally carry over to Lamé operators, the choice of discussing the case of the Laplacian only, finds its reasons solely in the intent of lightening the discussion.

    In the case of Δ+V with real-valued potential V, estimate (3.1) was first found by Keller [49] in d=1 and, later, generalized to an inequality for the negative eigenvalues power sum known as Lieb-Thirring inequality:

    zσd(Δ+V)|z|γLγ,d||V||γ+d2Lγ+d2(Rd), (3.2)

    where γ1/2 if d=1, γ>0 if d=2 and γ0 if d3 (same conditions as in Theorem 3.1) (see [57] and [29,54,60,61,64] for the endpoint cases). In passing, observe that bounds on single eigenvalues, like (3.1), represent a much weaker version of the Lieb-Thirring type inequalities (3.2).

    Now, some comment on inequalities (3.1) for Δ+V (in fact on the stronger bound (3.2)) are listed below (we refer to [56], Chapter 4, for further details).

    Remark 3.1. Contrarily to the case of complex-valued potentials, here, as a consequence of the variational principles (no more available in the non self-adjoint context), only the negative part of V, namely V, plays a role. Of course, since Δ is a non-negative operator, if V is also non-negative then so is Δ+V and therefore the variational characterization of the spectrum guarantees that no negative eigenvalues can occur. If V changes its sign, that is if both the positive and negative part of V=V+V are non-trivial, it is true that both parts influence the negative eigenvalues, but as Δ+VΔV, it is a consequence of the minimax principle that an upper estimate for the absolute value of negative eigenvalues of ΔV provides automatically the same upper estimate for the negative eigenvalues of the complete Hamiltonian Δ+V (actually, the same reasoning applies to the eigenvalue power sum). Indeed the effect of V+ on negative eigenvalues is only to increase their size.

    Remark 3.2. It is not difficult to see that if z is an eigenvalue of Δ+V with eigenfunction ψ, then ϕα():=ψ(α) is an eigenfunction of Δ+Vα(x) where Vα()=α2V(α) with eigenvalue α2z. By a simple scaling argument, this gives that p=γ+d/2 is the only possible exponent for which an inequality of the following type

    zσd(Δ+V)|z|γLγ,d||V||pLp(Rd)

    can hold. Notice that here our underlining domain is the entire Euclidean space Rd.

    Remark 3.3. Let us underline that there are "natural" constraints on the validity of inequalities of type (3.2) that can be easily justified. We emphasize here the pathological behavior of dimensions d=1,2. It is well known that, due to the lack of a Hardy-type inequality, the free Hamiltonian Δ is critical in low dimensions, which means that the addition of any arbitrarily small non-trivial negative potential V makes the bottom of the spectrum of the corresponding perturbed operator Δ+V negative, thus ensuring existence of negative eigenvalues. On the other hand, if an inequality of the form (3.2) with γ=0 holds, then the left-hand side would turn into the counting function of negative eigenvalues and so, as a consequence of the aforementioned criticality, it is an integer greater or equal to one for any such potential. On the contrary, since the right-hand side can be made arbitrarily small, for instance, assuming L0,d||V||d/2Ld/2(Rd)<1, would give an evident contradiction.

    We can now turn to proof of Theorem 3.1. As we will show later, it will come as a consequence of the following lemma which provides dimension-dependent estimates for the expectation value ψ,Vψ:=RdV|ψ|2dx of the potential energy V in the state ψH1(Rd).

    Lemma 3.1. Let d1 and let VLγ+d/2(Rd) be a real-valued function. The following estimates for ψ,Vψ hold true.

    i) If d=1 and γ12, then

    ψ,Vψ||V||γ+12||ψ||2(2γ1)2γ+12||ψ||42γ+1. (3.3)

    ii) If d=2 and γ>0, then

    ψ,Vψ||V||γ+1||ψ||22(γ+1)γ. (3.4)

    iii) If d3 and γ0, then

    ψ,Vψ||V||γ+d2||ψ||4γ2γ+d2||ψ||2d2γ+d2dd2. (3.5)

    Proof. Let us start with the proof of (3.5).

    It is an easy consequence of Hölder inequality that

    ψ,Vψ:=RdV|ψ|2dx||V||γ+d2||ψ||22(2γ+d)2γ+d2.

    Being 22(2γ+d)/(2γ+d2)2d/(d2), we can use the interpolation inequality to get

    ||ψ||22(2γ+d)2γ+d2||ψ||4γ2γ+d2||ψ||2d2γ+d2dd2.

    Plugging the latter in the former gives (3.5).

    Now let us consider d=1,2. Estimate (3.4) is immediate consequence of Hölder inequality and the same holds for the case γ=1/2 in d=1. Finally, the remaining case γ>1/2 follows, as in the three dimensional framework, from Hölder and interpolation inequality, using that 22(2γ+1)/(2γ1)<. This proves (3.3) and concludes the proof of the lemma.

    In passing, observe that if ψH1(Rd), then the norms on the right hand side of (3.3), (3.4) and (3.5) are finite. This is a consequence of the Sobolev embeddings

    H1(Rd)Lq(Rd)where{q=ifd=1,2q<ifd=2,q=2d/(d2)ifd3,

    which are quantified by the inequalities contained in the following lemma (see [55,Ch. 8] or [56,Sec. 2.2.1]).

    Lemma 3.2 (Sobolev inequalities). Let d1 and let ψH1(Rd).

    i) If d=1, then

    R|dψdx|2dx||ψ||22||ψ||4. (3.6)

    ii) If d=2, then

    R2|ψ|2dxS2,q||ψ||4q22||ψ||2qq2q,2<q<. (3.7)

    iii) If d3, then

    Rd|ψ|2dxSd||ψ||22dd2. (3.8)

    Here, S2,q and Sd denote the optimal Sobolev constants ([55,56]).

    Now we are in position to prove Theorem 3.1.

    In this setting, the variational characterization of the spectrum states that for any u[H1(Rd)]d,

    infσ(Δ+V)=inf||u||[L2(Rd)]d=1u,(ΔV)u.

    Therefore, in order to get (3.1) it is sufficient to prove the following lower bound

    u,(ΔV)uC1γγ,d,λ,μ(RdVγ+d2dx)1γ (3.9)

    for any u[H1(Rd)]d with ||u||[L2(Rd)]d=1. In order to estimate u,(ΔV)u, we consider the Helmholtz decomposition u=uS+uP of u. It follows from the explicit expression (2.15) of Δ and from the H1-orthogonality of uS and uP (see Subsection 2.1) that

    u,(ΔV)u=uS,μΔuS+uP,(λ+2μ)ΔuPu,Vu=μdj=1Rd|u(j)S|2dx+(λ+2μ)dj=1Rd|u(j)P|2dxdj=1u(j),Vu(j). (3.10)

    Let us start considering the case d3.

    Using Sobolev inequality (3.8) on the j-th component of the vector-field uS, one gets

    dj=1Rd|u(j)S|2dxSddj=1||u(j)S||22dd2Sd(dj=1||u(j)S||2dd22dd2)d2d=:Sd||uS||22dd2, (3.11)

    where in the last inequality we used the sub-additivity of the concave function |x|d2d.

    The same computation performed for uP gives

    dj=1Rd|u(j)P|2dxSd||uP||22dd2. (3.12)

    Now we are in position to estimate u,Vu. Using bound (3.5) in Lemma 3.1, two times the Hölder inequality for discrete measures and, finally, the Young inequality, abap/p+bq/q which holds for all positive a,b and 1/p+1/q=1, we get for some ε>0

    dj=1u(j),Vu(j)||V||γ+d2dj=1||u(j)||4γ2γ+d2||u(j)||2d2γ+d2dd2d22γ+d||V||γ+d2||u||4γ2γ+d2||u||2d2γ+d2dd2d22γ+dε1+d2γ2γ2γ+d||V||1+d2γγ+d2+ε1+2γdd1+22γ+d2γ+d||u||22dd2d22γ+dε1+d2γ2γ2γ+d||V||1+d2γγ+d2+2ε1+2γdd1+22γ+d2γ+d[||uS||22dd2+||uP||22dd2], (3.13)

    where in the last inequality we simply used the inequality ||u||22dd22[||uS||22dd2+||uP||22dd2].

    Now, plugging (3.11), (3.12) and (3.13) in (3.10), one has

    u,(ΔV)u(min{μ,λ+2μ}Sd2ε1+2γdd1+22γ+d2γ+d)(||uS||22dd2+||uP||22dd2)d22γ+dε1+d2γ2γ2γ+d||V||1+d2γγ+d2.

    Choosing a suitable small ε=:εγ,d,λ,μ, one gets (3.9) with

    C1γγ,d,λ,μ:=d22γ+dε1+d2γγ,d,λ,μ2γ2γ+d.

    Hence, bound (3.1) is proved if d3. We skip the proof of the analogous bounds in the lower dimensional cases, namely d=1,2. Indeed these follow from the corresponding estimates in Lemma 3.1 and Sobolev inequalities (Lemma 3.2) with minor modifications from the reasoning above.

    This section is concerned with the proof of the eigenvalue bounds contained in Theorems 1.2–1.5. As we will see, with the estimates of Theorem 2.3 in hand, the proofs will follow smoothly.

    As already mentioned in the introduction, the starting point in our proofs is the Birman-Schwinger principle. In our context it states that if zC[0,) is an eigenvalue of Δ+V, then 1 is an eigenvalue of the Birman-Schwinger operator V1/2(Δz)1|V|1/2 on [L2(Rd)]d. This implies that the operator norm of the latter is at least 1. Therefore, in order to get the bound (1.8), we are reduced to prove

    ||V12(Δz)1|V|12||γ+d2L2L2Cγ,d,λ,μ|z|γ||V||γ+d2Lγ+d2. (4.1)

    The same strategy, with the needed modifications, then will be also applied to prove the corresponding bounds in Theorem 1.3, Theorem 1.4 and 1.5.

    Providing bounds for ||V12(Δz)1|V|12||L2L2 is estimating the quantity |f,V1/2(Δz)1|V|12g|, for f,g[L2(Rd)]d. To simplify the notation we introduce the function G:=|V|12g. Using the Hölder inequality and estimate (2.22) for the resolvent (Δz)1, we have

    |f,V12(Δz)1G|||f|V|12||p||(Δz)1G||pCp,d,λ,μ|z|d+22+dp||f|V|12||p||G||p.

    Thus, recalling that G=|V|1/2g, one has

    |f,V12(Δz)1|V|12g|Cp,d,λ,μ|z|d+22+dp||f|V|12||p|||V|12g||p. (4.2)

    Using the Hölder inequality and its version for discrete measures and the sub-additivity property of the concave function |x|p, for 0<p1, one gets

    ||f|V|12||p=(dj=1||fj|V|12||pp)1p(dj=1||fj||p2||V||p2p2p)1p||V||12p2pdj=1||fj||2d12||V||12p2p||f||2.

    The same estimate for the term involving g gives

    |||V|12g||pd12||V||12p2p||g||2.

    Plugging these two bounds in (4.2) we end up with the following inequality

    |f,V12(Δz)1|V|12g|Cp,d,λ,μd|z|d+22+dp||V||p2p||f||2||g||2.

    Now, choosing p=2(2γ+d)2γ+d+2 (observe that the restriction on γ in Theorem 1.2 guarantees that p satisfies the hypotheses in Theorem 2.3) and taking the supremum over all f,g[L2(Rd)]d with norm less than or equal to one, we get (4.1). This concludes the the proof of Theorem 1.2.

    As in the proof of the previous result, we are reduced to prove the following bound

    ||V12(Δz)1|V|12||γ+d2L2L2Cγ,d,λ,μ|z|γ||V||γ+d2Lα,p(Rd).

    We pick a strictly positive function ϕLα,p and we define a strictly positive approximation of our potential, that is Vε(x):=sup{|V(x)|,εϕ(x)}. Using Cauchy-Schwarz inequality and estimate (2.24) for the resolvent (Δz)1, we have

    |f,V12(Δz)1|V|12g|||f|V|/Vε||2||(Δz)1|V|12g||L2(Vε)Cα,d,λ,μ|z|1+α2||Vε||Lα,p||f|V|/Vε||2||g|V|/Vε||2Cα,d,λ,μ|z|1+α2||Vε||Lα,p||f||2||g||2.

    Thus, the theorem is proved once ε goes to zero, taking the supremum over all f,g[L2(Rd)]d with norm less than or equal to one and by choosing α=2d2γ+d.

    Again we are reduced to prove the following bound

    ||V12(Δz)1|V|12||γ+d2L2L2Cγ,d,λ,μ|z|γ||Vβ||1β(γ+d2)KSα.

    The same strategy used above, with the usage of (2.25) instead of (2.24), gives

    |f,V12(Δz)1|V|12g|||f|V|/Vε||2||(Δz)1|V|12g||L2(Vε)Cα,p,d,λ,μ|z|αd+12αd+1||Vβε||1βKSα||f|V|/Vε||2||g|V|/Vε||2Cα,p,d,λ,μ|z|αd+12αd+1||Vβε||1βKSα||f||2||g||2.

    Thus, the theorem is proved once ε goes to zero, taking the supremum over all f,g[L2(Rd)]d with norm less than or equal to one and by choosing α=d(d1)d2γ.

    As before, our problem is reduced to proving the following bound

    ||V12(Δz)1|V|12||qL2L2Cγ,α,d,λ,μ|z|γ||V||qLq(x2αdx), (4.3)

    with q=2γ+(d1)/2. First, observe that from (2.22), with the choice p=2(d+1)/(d+3) one has

    ||(Δz)1||LpLpCd,λ,μ|z|1d+1. (4.4)

    In passing, notice that (1.8) was obtained with the usage of (2.18) with the choice p=2(2γ+d)/(2γ+d+2) (see last part of Proof of Theorem 1.2). Then p=2(d+1)/(d+3) corresponds to the case \gamma = 1/2 in (1.8) which gave the aforementioned decay threshold 2d/(d+1) .

    We know from (2.23) that

    \begin{equation} ||(-\Delta^\ast-z)^{-1}||_{L^2(\langle x \rangle^{2\alpha}) \to L^2(\langle x \rangle^{-2\alpha})}\leq C_{\alpha, d, \lambda, \mu} |z|^{-\frac{1}{2}}, \qquad \alpha > \frac{1}{2}. \end{equation} (4.5)

    Using Riesz-Thorin interpolation between estimate (4.4) and (4.5), we get

    \begin{equation*} ||(-\Delta^\ast-z)^{-1}||_{L^{p_\theta}(\langle x \rangle^{\alpha \theta p_\theta}) \to L^{p_\theta'}(\langle x \rangle^{-\alpha \theta p_\theta'})} \leq C_{\alpha, d, \lambda, \mu}|z|^{-\frac{1-\theta}{d+1} - \frac{\theta}{2}}, \qquad \frac{1}{p_\theta} = \frac{1-\theta}{p} + \frac{\theta}{2}, \end{equation*}

    with p = 2(d+1)/(d+3), \alpha > 1/2 and 1/p_\theta + 1/p_\theta' = 1.

    From this fact it is easy to get

    \begin{equation*} \begin{split} |\langle f, V_\frac{1}{2} (-\Delta^\ast - z)^{-1} |V|^\frac{1}{2} g\rangle| &\leq ||f |V|^\frac{1}{2} \langle x \rangle^{\alpha \theta}||_{{p_\theta}} ||(-\Delta^\ast-z)^{-1} |V|^\frac{1}{2} g||_{L^{p_\theta'}(\langle x \rangle^{-\alpha \theta p_\theta'})}\\ &\leq C_{\alpha, d, \lambda, \mu} |z|^{-\frac{1-\theta}{d+1} - \frac{\theta}{2}} ||f |V|^\frac{1}{2} \langle x \rangle^{\alpha \theta}||_{{p_\theta}} || \langle x \rangle^{\alpha \theta}|V|^\frac{1}{2} g||_{{p_\theta}}\\ & \leq C_{\alpha, d, \lambda, \mu}\, d |z|^{-\frac{1-\theta}{d+1} - \frac{\theta}{2}} ||f||_2 ||g||_2 |||V| \langle x \rangle^{2\alpha \theta}||_{\frac{p_\theta}{2-p_\theta}}. \end{split} \end{equation*}

    Taking the supremum over all f, g\in [L^2(\mathbb{R}^d)]^d with norm less than or equal to one and raising the resulting inequality to the power p_\theta/(2-p_\theta) gives

    \begin{equation*} ||V_\frac{1}{2} (-\Delta^\ast -z)^{-1} |V|^\frac{1}{2}||_{L^2 \to L^2}^\frac{p_\theta}{2-p_\theta} \leq C_{\alpha, d, \lambda, \mu} |z|^{-\big(\frac{1-\theta}{d+1} + \frac{\theta}{2}\big) \tfrac{p_\theta}{2-p_\theta}} |||V| \langle x \rangle^{2\alpha \theta}||_{\frac{p_\theta}{2-p_\theta}}^\frac{p_\theta}{2-p_\theta}. \end{equation*}

    Here we abuse the notation by using the same symbol C_{\alpha, d, \lambda, \mu} for different constants.

    Calling

    \begin{equation*} \gamma: = \left(\frac{1-\theta}{d+1}+\frac{\theta}{2}\right) \frac{p_\theta}{2-p_\theta}, \end{equation*}

    this clearly gives \frac{p_\theta}{2-p_\theta} = 2\gamma\, \frac{d+1}{2-\theta + d\theta}, since we also have \frac{1}{p_\theta} = \frac{1-\theta}{p} + \frac{\theta}{2}, with p = \frac{2(d+1)}{d+3}, this leads to the constraint \theta = 1-\frac{d+1}{4\gamma + d-1}. With these choices one has

    \begin{equation*} ||V_\frac{1}{2} (-\Delta^\ast -z)^{-1} |V|^\frac{1}{2}||_{L^2\to L^2}^q \leq C_{\gamma, \alpha, d, \lambda, \mu} |z|^{-\gamma} ||V||_{L^q(\langle x \rangle^{2\alpha(\gamma-1)}\, dx)}^q, \end{equation*}

    with q = 2\gamma + (d-1)/2, which is the bound (4.3) once renaming \alpha(2\gamma -1) = \alpha.

    The author is deeply grateful to Prof. L. Fanelli for bringing the attention to the setting of the problem and for valuable comments on the draft of the manuscript. The author also thanks Prof. D. Krejčiřík for useful discussions which greatly enrich the paper. The author is greatful to Prof. M. Correggi for helpful references' suggestions. The author thanks the hospitality of Sapienza University of Rome where this work was initiated and Czech Technical University in Prague where it was developed in part. The author gratefully acknowledges financial support by the Deutsche Forschungsgemeinschaft (DFG) through CRC 1173.

    The author declares no conflict of interest.

    In this appendix we give the rigorous definition of the operator -\Delta^\ast + V. This is introduced as an m -accretive operator obtained as a form sum of the free Lamé operator -\Delta^\ast with domain [H^1(\mathbb{R}^d)]^d and a relatively form-bounded potential VI_{ \mathbb{C}^d\times \mathbb{C}^d} with V\colon \mathbb{R}^d \to \mathbb{C}. More precisely, our standing assumption is that there exist numbers a\in (0, 1) and b\in \mathbb{R} such that, for all u \in [H^1(\mathbb{R}^d)]^d,

    \begin{equation} \int_{ \mathbb{R}^d} |V||u|^2\, dx\leq a\, h_0^\ast[u] + b \int_{ \mathbb{R}^d} |u|^2\, dx, \end{equation} (A.1)

    where

    \begin{equation*} h_0^\ast[u]: = \lambda \int_{ \mathbb{R}^d}| \text{ div }(u)|^2\, dx + \frac{\mu}{2} \int_{ \mathbb{R}^d} |\nabla u + \nabla^T u|^2\, dx. \end{equation*}

    Here \nabla u denotes the matrix with entries the first derivatives of u, namely (\nabla u)_{jk} = \frac{\partial u_j}{\partial x_k}, whereas \nabla^T u is the transpose matrix of \nabla u, namely (\nabla^T u)_{jk} = \frac{\partial u_k}{\partial x_j}. We refer the reader to [24,Sec. 2.1] for more details. In the following we will show that under the hypotheses of Thm. 1.1–Thm. 1.5, V satisfies (A.1).

    As a warm up we consider the easiest case d = 1 first. If u\in H^1(\mathbb{R}), then by Sobolev embedding u\in L^\infty(\mathbb{R}). In particular (3.6) holds. Using (3.6), the 1D expression of the Lamé operator (1.5) and the Cauchy Schwarz inequality one easily has that for any \varepsilon > 0

    \begin{equation*} \begin{split} \int_{ \mathbb{R}} |V||u|^2\, dx &\leq \|u\|_{L^\infty( \mathbb{R})}^2 \|V\|_{L^1( \mathbb{R})}\\ &\leq \|u\|_{L^2( \mathbb{R})} \Bigg(\int_{ \mathbb{R}} \Big|\frac{du}{dx}\Big|^2\, dx\Bigg)^{\!1/2} \|V\|_{L^1( \mathbb{R})}\\ & = \frac{1}{\sqrt{\varepsilon}} \|u\|_{L^2( \mathbb{R})} \sqrt{\varepsilon} \Bigg(\frac{1}{\lambda + 2\mu} (\lambda+ 2\mu)\int_{ \mathbb{R}} \Big|\frac{du}{dx}\Big|^2\, dx\Bigg)^{\!1/2} \|V\|_{L^1( \mathbb{R})}\\ &\leq \frac{\|V\|_{L^1( \mathbb{R})}}{\sqrt{\lambda + 2\mu}} \Big[\varepsilon h_0^\ast[u] + \frac{1}{\varepsilon} \|u\|_{L^2( \mathbb{R})}^2 \Big]. \end{split} \end{equation*}

    Since \varepsilon is arbitrary one gets that if V\in L^1(\mathbb{R}), then (A.1) holds and -\Delta^\ast + V is well defined as an m -accretive operator on H^1(\mathbb{R}).

    In order to tackle the higher dimensional framework d\geq 2 we need to state a preliminary lemma. For simplicity, in the following we shorten the notation for the L^2-L^2 operator norm \|\cdot\|_{L^2(\mathbb{R}^d)\to L^2(\mathbb{R}^d)} to the more compact \|\cdot\|.

    Lemma A.1. Let d\geq 2 and z\in \mathbb{C}\setminus [0, \infty). For \gamma as in Theorem 1.2, if V\in L^{\gamma+ \frac{d}{2}}(\mathbb{R}^d), then

    \begin{equation} \||V|^{1/2}(-\Delta^\ast -z)^{-1}|V|^{1/2}\|\leq C|z|^{-\frac{2\gamma}{2\gamma +d}} \|V\|_{L^{\gamma + \frac{d}{2}}( \mathbb{R}^d)}. \end{equation} (A.2)

    For \gamma, \alpha and p as in Theorem 1.3, if V\in \mathcal{L}^{\alpha, p}(\mathbb{R}^d), then

    \begin{equation} \||V|^{1/2}(-\Delta^\ast -z)^{-1}|V|^{1/2}\|\leq C|z|^{-\frac{2\gamma}{2\gamma +d}} \|V\|_{\mathcal{L}^{\alpha, p}( \mathbb{R}^d)}. \end{equation} (A.3)

    For \gamma, \alpha and \beta as in Theorem 1.4, if V^\beta\in \mathcal{KS}_{\alpha}(\mathbb{R}^d) and if V\in A_2(\mathbb{R}^d), then

    \begin{equation} \||V|^{1/2}(-\Delta^\ast -z)^{-1}|V|^{1/2}\|\leq C|z|^{-\frac{2\gamma}{2\gamma +d}} Q_2(|V|)^2 \||V|^\beta\|_{\mathcal{KS}_{\alpha}( \mathbb{R}^d)}^\frac{1}{\beta}. \end{equation} (A.4)

    For \gamma, \alpha and q as in Theorem 1.5, if V\in L^q(\langle x \rangle^{2\alpha} dx), then

    \begin{equation} \||V|^{1/2}(-\Delta^\ast -z)^{-1}|V|^{1/2}\|\leq C|z|^{-\frac{\gamma}{q}} \|V\|_{L^{q}(\langle x \rangle^{\alpha}dx)}. \end{equation} (A.5)

    The constant C may change from (A.2) to (A.5) but in each estimate it is independent of V.

    Proof. The proof of estimates (A.2)–(A.5) follows with almost no modification of the proofs of Theorem 1.2–Theorem 1.5, respectively (see Sec. 4).

    With Lemma A.1 at hand we see that V satisfies (A.1) whenever the hypotheses of Thm. 1.2–Thm. 1.5 are satisfied. We consider first the case \gamma\neq 0. The special case \gamma = 0 will be treated separately.

    Let z\in (-\infty, 0). We write estimates (A.2)–(A.5) in a unified way as follows:

    \begin{equation*} \||V|^{1/2}(-\Delta^\ast-z)^{-1} |V|^{1/2}\|\leq C|z|^{-\frac{2\gamma}{2\gamma + d}}\|V\|, \end{equation*}

    where C might depend on V and where \|V\| denotes one of the norms of the right hand side of estimates (A.2)–(A.5). Defining A: = |V|^{1/2}(-\Delta^\ast -z)^{-1/2} one has that A^\ast = (-\Delta^\ast- z)^{-1/2}|V|^{1/2} and consequently AA^\ast = |V|^{1/2}(-\Delta^\ast-z)^{-1} |V|^{1/2}. Using that \|AA^\ast\| = \|A^\ast\|^2 = \|A\|^2 we get

    \begin{equation*} \begin{split} \int_{ \mathbb{R}^d} |V||u|^2& = \||V|^{1/2}(-\Delta^\ast-z)^{-1/2}(-\Delta^\ast-z)^{1/2}u\|_{L^2( \mathbb{R}^d)}^2\\ &\leq \||V|^{1/2}(-\Delta^\ast-z)^{-1/2}\|^2 \|(-\Delta^\ast-z)^{1/2}u\|_{L^2( \mathbb{R}^d)}^2\\ & = \||V|^{1/2}(-\Delta^\ast-z)^{-1} |V|^{1/2}\| \|(-\Delta^\ast-z)^{1/2}u\|_{L^2( \mathbb{R}^d)}^2\\ &\leq C |z|^{-\frac{2\gamma}{2\gamma + d}}\|V\|\langle u, (-\Delta^\ast-z)u \rangle\\ & = C |z|^{-\frac{2\gamma}{2\gamma + d}}\|V\| (h_0^\ast[u] +|z| \|u\|_{L^2( \mathbb{R}^d)}^2)\\ & = C\|V\|\big(|z|^{-\frac{2\gamma}{2\gamma + d}}h_0^\ast[u] + |z|^{\frac{d}{2\gamma+d}}\|u\|_{L^2( \mathbb{R}^d)}^2\big). \end{split} \end{equation*}

    Thus, bound (A.1) is obtained in the limit |z|\to \infty.

    Now it is left the case \gamma = 0. We shall see the explicit computations only in the Lebesgue setting, namely when V\in L^\frac{d}{2}(\mathbb{R}^d), when V belongs to the Morrey-Campanato, Keller-Sawyer and weighted L^q class one can argue similarly.

    Let R > 0, we split the left hand side of (A.1) as follows

    \begin{equation} \begin{split} \int_{ \mathbb{R}^d} |V||u|^2\, dx& = \int_{\{x\in \mathbb{R}^d\colon |V(x)| > R\}} |V||u|^2\, dx + \int_{\{x\in \mathbb{R}^d\colon |V(x)|\leq R\}}|V||u|^2, dx\\ & = I_{ > R} + I_{\leq R}. \end{split} \end{equation} (A.6)

    For I_{\leq R} one easily has

    \begin{equation*} I_{\leq R}\leq R \| u\|_{L^2( \mathbb{R}^d)}^2. \end{equation*}

    As for I_{ > R}, using Hölder and Sobolev inequality (3.8) we have

    \begin{equation*} \begin{split} I_{ > R}&\leq \Big(\int_{\{x\in \mathbb{R}^d\colon |V(x)| > R\}} |V|^{\frac{d}{2}}\, dx\Big)^{\frac{2}{d}}\|u\|_{L^{\frac{2d}{d-2}}}^2\\ &\leq S_d^{-1} \Big(\int_{\{x\in \mathbb{R}^d\colon |V(x)| > R\}} |V|^{\frac{d}{2}}\, dx\Big)^{\frac{2}{d}} \|\nabla u\|_{L^2( \mathbb{R}^d)}^2\\ &\leq C\Big(\int_{\{x\in \mathbb{R}^d\colon |V(x)| > R\}} |V|^{\frac{d}{2}}\, dx\Big)^{\frac{2}{d}} h_0^\ast[u], \end{split} \end{equation*}

    where in the last step we have used that \|\nabla u\|_{L^2(\mathbb{R}^d)}^2\leq Ch_0^\ast[u] (see [24,Sec. 2.1]).

    Plugging the previous estimates for I_{ > R} and I_{\leq R} in (A.6) one eventually has

    \begin{equation} \int_{ \mathbb{R}^d} |V||u|^2\, dx\leq C\Big(\int_{\{x\in \mathbb{R}^d\colon |V(x)| > R\}} |V|^{\frac{d}{2}}\, dx\Big)^{\frac{2}{d}} h_0^\ast[u] + R \| u\|_{L^2( \mathbb{R}^d)}^2. \end{equation} (A.7)

    Since the measure of the set \{x\in \mathbb{R}^d\colon |V(x)| > R\} tends to zero as R goes to infinity, one has that (A.1) follows from (A.7).



    [1] A. A. Abramov, A. Aslanyan, E. B. Davies, Bounds on complex eigenvalues and resonances, J. Phys. A, 34 (2001), 57–72. doi: 10.1088/0305-4470/34/1/304
    [2] S. Agmon, Spectral properties of Schrödinger operators and scattering theory, Ann. Scuola Norm. Sup. Pisa Cl. Sci., 2 (1975), 151–218.
    [3] S. Avramska-Lukarska, D. Hundertmark, H. Kovařík, Absence of positive eigenvalues for magnetic Schrödinger operators, 2020, arXiv: 2003.07294.
    [4] R. Bañuelos, G. Wang, Sharp inequalities for martingales with applications to the Beurling-Ahlfors and Riesz transforms, Duke Math. J., 80 (1995), 575–600.
    [5] J. A. Barceló, J. M. Bennet, A. Ruiz, M. C. Vilela, Local smoothing for Kato potentials in three dimensions, Math. Nachr., 282 (2009), 1391–1405. doi: 10.1002/mana.200610808
    [6] J. A. Barceló, M. Folch-Gabayet, S. Pérez-Esteva, A. Ruiz, M. C. Vilela, Limiting absorption principles for the Navier equation in elasticity, Ann. Sc. Norm. Super. Pisa Cl. Sci., 11 (2012), 817–842.
    [7] M. Sh. Birman, Discrete spectrum in the gaps of a continuous one for perturbations with large coupling constant, In: Estimates and asymptotics for discrete spectra of integral and differential equations, Providence: Amer. Math. Soc., 1991, 57–73.
    [8] S. Bögli, Schrödinger operator with non-zero accumulation points of complex eigenvalues, Commun. Math. Phys., 352 (2017), 629–639. doi: 10.1007/s00220-016-2806-5
    [9] V. Bruneau, E. M. Ouhabaz, Lieb-Thirring estimates for non self-adjoint Schrödinger operators, J. Math. Phys., 49 (2008), 093504. doi: 10.1063/1.2969028
    [10] N. Boussaid, P. D'Ancona, L. Fanelli, Virial identity and weak dispersion for the magnetic Dirac equation, J. Math. Pure. Appl., 95 (2011), 137–150. doi: 10.1016/j.matpur.2010.10.004
    [11] F. Cacciafesta, Virial identity and dispersive estimates for the n-dimensional Dirac equation, J. Math. Sci. Univ. Tokyo, 18 (2011), 441–463.
    [12] A. P. Calderón, A. Zygmund, On singular integrals, Am. J. Math., 78 (1956), 289–309.
    [13] E. A. Carlen, R. L. Frank, E. H. Lieb, Stability estimates for the lowest eigenvalue of a Schrödinger operator, Geom. Funct. Anal., 24 (2014), 63–84. doi: 10.1007/s00039-014-0253-z
    [14] B. Cassano, L. Cossetti, L. Fanelli, Eigenvalue bounds and spectral stability of Lamé operators with complex potentials, J. Differ. Equations, 298 (2021), 528–559. doi: 10.1016/j.jde.2021.07.017
    [15] B. Cassano, L. Cossetti, L. Fanelli, Spectral enclosures for the damped elastic wave equation, 2021, arXiv: 2108.07676.
    [16] B. Cassano, P. D'Ancona, Scattering in the energy space for the NLS with variable coefficients, Math. Ann., 366 (2016), 479–543. doi: 10.1007/s00208-015-1335-4
    [17] B. Cassano, O. O. Ibrogimov, D. Krejčiřík, F. Štampach, Location of eigenvalues of non-self-adjoint discrete Dirac operators, Ann. Henri Poincaré, 21 (2020), 2193–2217. doi: 10.1007/s00023-020-00916-2
    [18] B. Cassano, F. Pizzichillo, L. Vega, A Hardy-type inequality and some spectral characterizations for the Dirac–Coulomb operator, Rev. Mat. Complut., 33 (2020), 1–18. doi: 10.1007/s13163-019-00311-4
    [19] S. Chanillo, B. Helffer, A. Laptev, Nonlinear eigenvalues and analytic hypoellipticity, J. Funct. Anal., 209 (2004), 425–433. doi: 10.1016/S0022-1236(03)00105-8
    [20] S. Chanillo, E. Sawyer, Unique continuation for \Delta + v and the C. Fefferman-Phong class, Trans. Amer. Math. Soc., 318 (1990), 275–300.
    [21] F. Chiarenza, M. Frasca, A remark on a paper by C. Fefferman, Proc. Amer. Math. Soc., 108 (1990), 407–409.
    [22] F. Chiarenza, A. Ruiz, Uniform L^2- weighted Sobolev inequalities, Proc. Amer. Math. Soc., 112 (1991), 53–64.
    [23] R. R. Coifman, C. Fefferman, Weighted norm inequalities for maximal functions and singular integrals, Stud. Math., 51 (1974), 241–250. doi: 10.4064/sm-51-3-241-250
    [24] L. Cossetti, Uniform resolvent estimates and absence of eigenvalues for Lamé operators with subordinated complex potentials, J. Math. Anal. Appl., 1 (2017), 336–360.
    [25] L. Cossetti, L. Fanelli, D. Krejčiřík, Absence of eigenvalues of Dirac and Pauli Hamiltonians via the method of multipliers, Commun. Math. Phys., 379 (2020), 633–691. doi: 10.1007/s00220-020-03853-7
    [26] J.-C. Cuenin, Eigenvalue bounds for Dirac and Fractional Schrödinger operators with complex potentials, J. Funct. Anal., 272 (2017), 2987–3018. doi: 10.1016/j.jfa.2016.12.008
    [27] J.-C. Cuenin, Eigenvalue bounds for bilayer graphene, Ann. Henri Poincaré, 20 (2019), 1501–1516. doi: 10.1007/s00023-019-00770-x
    [28] J.-C. Cuenin, A. Laptev, C. Tretter, Eigenvalue estimates for non-selfadjoint Dirac operators on the real line, Ann. Henri Poincaré, 15 (2014), 707–736. doi: 10.1007/s00023-013-0259-3
    [29] M. Cwikel, Weak type estimates for singular values and the number of bound states of Schrödinger operators, Ann. Math., 106 (1977), 93–102. doi: 10.2307/1971160
    [30] P. D'Ancona, L. Fanelli, D. Krejčiřík, N. M. Schiavone, Localization of eigenvalues for non-self-adjoint Dirac and Klein-Gordon operators, Nonlinear Anal., 214, (2022), 112565.
    [31] P. D'Ancona, L. Fanelli, N. M. Schiavone, Eigenvalue bounds for non-selfadjoint Dirac operators, Math. Ann., 2021, https://doi.org/10.1007/s00208-021-02158-x.
    [32] M. Demuth, M. Hansmann, G. Katriel, On the discrete spectrum of non-selfadjoint operators, J. Funct. Anal., 257 (2009), 2742–2759. doi: 10.1016/j.jfa.2009.07.018
    [33] M. Demuth, M. Hansmann, G. Katriel, Lieb-Thirring type inequalities for Schrödinger operators with a complex-valued potential, Integr. Equ. Oper. Theory, 75 (2013), 1–5. doi: 10.1007/s00020-012-2021-5
    [34] A. Enblom, Estimates for eigenvalues of Schrödinger operators with complex-valued potentials, Lett. Math. Phys., 106 (2016), 197–220. doi: 10.1007/s11005-015-0810-x
    [35] L. Fanelli, Non-trapping magnetic fields and Morrey-Campanato estimates for Schrödinger operators, J. Math. Anal. Appl., 357 (2017), 1–14.
    [36] L. Fanelli, D. Krejčiřík, Location of eigenvalues of three-dimensional non-self-adjoint Dirac operators, Lett. Math. Phys., 109 (2019), 1473–1485. doi: 10.1007/s11005-018-01155-7
    [37] L. Fanelli, D. Krejčiřík, L. Vega, Spectral stability of Schrödinger operators with subordinated complex potentials, J. Spectr. Theory, 8 (2018), 575–604. doi: 10.4171/JST/208
    [38] L. Fanelli, D. Krejčiřík, L. Vega, Absence of eigenvalues of two-dimensional magnetic Schrödinger operators, J. Funct. Anal., 275 (2018), 2453–2472. doi: 10.1016/j.jfa.2018.08.007
    [39] F. Ferrulli, A. Laptev, O. Safronov, Eigenvalues of the bilayer graphene operator with a complex valued potential, Anal. Math. Phys., 9 (2019), 1535–1546. doi: 10.1007/s13324-018-0262-4
    [40] R. L. Frank, Eigenvalue bounds for Schrödinger operators with complex potentials, Bull. Lond. Math. Soc., 43 (2011), 745–750. doi: 10.1112/blms/bdr008
    [41] R. L. Frank, Eigenvalue bounds for Schrödinger operators with complex potentials. III, Trans. Amer. Math. Soc., 370 (2018), 219–240.
    [42] R. L. Frank, A. Laptev, E. H. Lieb, R. Seiringer, Lieb-Thirring inequalities for Schrödinger operators with complex-valued potentials, Lett. Math. Phys., 77 (2006), 309–316. doi: 10.1007/s11005-006-0095-1
    [43] R. L. Frank, A. Laptev, O. Safronov, On the number of eigenavlues of Schrödinger operators with complex potentials, J. London Math. Soc., 94 (2016), 377–390. doi: 10.1112/jlms/jdw039
    [44] R. L. Frank, M. Loss, Which magnetic fields support a zero mode?, 2020, arXiv: 2012.13646.
    [45] R. L. Frank, B. Simon, Eigenvalue bounds for Schrödinger operators with complex potentials. II, J. Spectr. Theory, 7 (2017), 633–658. doi: 10.4171/JST/173
    [46] M. Hansmann, D. Krejčiřík, The abstract Birman-Schwinger principle and spectral stability, 2020, arXiv: 2010.15102.
    [47] O. O. Ibrogimov, D. Krejčiřík, A. Laptev, Sharp bounds for eigenvalues of biharmonic operators with complex potentials in low dimensions, Math. Nachr., 294 (2021), 1333–1349. doi: 10.1002/mana.202000196
    [48] O. O. Ibrogimov, F. Štampach, Spectral enclosures for non-self-adjoint discrete Schrödinger operators, Integr. Equ. Oper. Theory, 91 (2019), 53. doi: 10.1007/s00020-019-2553-z
    [49] J. B. Keller, Lower bounds and isoperimetric inequalities for eigenvalues of the Schrödinger equation, J. Math. Phys., 2 (1961), 262–266. doi: 10.1063/1.1703708
    [50] C. E. Kenig, A. Ruiz, C. D. Sogge, Uniform Sobolev inequalities and unique continuation for second order constant coefficient differential operators, Duke Math. J., 55 (1987), 329–347.
    [51] D. Krejčiřík, T. Kurimaiová, From Lieb–Thirring inequalities to spectral enclosures for the damped wave equation, Integr. Equ. Oper. Theory, 92 (2020), 47. doi: 10.1007/s00020-020-02607-3
    [52] A. Laptev, O. Safronov, Eigenvalue estimates for Schrödinger operators with complex potentials, Commun. Math. Phys., 292 (2009), 29–54. doi: 10.1007/s00220-009-0883-4
    [53] Y. Lee, I. Seo, A note on eigenvalue bounds for Schrödinger operators, J. Math. Anal. Appl., 470 (2019), 340–347. doi: 10.1016/j.jmaa.2018.10.006
    [54] E. H. Lieb, The number of bound states of one-body Schrödinger operators and the Weyl problem, In: The stability of matter: from atoms to stars, Berlin, Heidelberg: Springer, 1997,241–252.
    [55] E. H. Lieb, M. Loss, Analysis, 2 Eds., Providence, Rhode Island: American Mathematical Society, 2001.
    [56] E. H. Lieb, R. Seiringer, The stability of matter in quantum mechanics, Cambridge University Press, 2010.
    [57] E. H. Lieb, W. Thirring, Inequalities for the moments of the eigenvalues of the Schrödinger Hamiltonian and their relation to Sobolev inequalities, In: Studies in mathematical physics, Princeton: Princeton University Press, 1976,269–303.
    [58] H. Mizutani, N. M. Schiavone, Keller-type bounds for Dirac operators perturbed by rigid potentials, 2021, arXiv: 2108.12854.
    [59] S. Petermichl, The sharp weighted bound for the Riesz transforms, Proc. Amer. Math. Soc., 136 (2008), 1237–1249.
    [60] G. V. Rozenblyum, Distribution of the discrete spectrum of singular differential operators, Sov. Math. Dokl., 13 (1972), 245–249.
    [61] G. V. Rozenblyum, Distribution of the discrete spectrum of singular differential operators, Soviet Math. (Iz. VUZ), 20 (1976), 63–71.
    [62] S. A. Stepin, An estimate for the number of eigenvalues of the Schrödinger operator with complex potential, Sb. Math., 208 (2017), 269–284. doi: 10.1070/SM8686
    [63] J. Sylvester, An estimate for the free Helmholtz equation that scales, Inverse Probl. Imaging, 3 (2009), 333–351. doi: 10.3934/ipi.2009.3.333
    [64] T. Weidl, On the Lieb-Thirring constants L_{\gamma, 1} for \gamma \geq \frac{1}{2}, Commun. Math. Phys., 178 (1996), 135–146. doi: 10.1007/BF02104912
  • This article has been cited by:

    1. Seongyeon Kim, Yehyun Kwon, Sanghyuk Lee, Ihyeok Seo, Strichartz and uniform Sobolev inequalities for the elastic wave equation, 2022, 151, 0002-9939, 239, 10.1090/proc/16101
    2. Borbala Gerhat, Orif O. Ibrogimov, Petr Siegl, On the Point Spectrum in the Ekman Boundary Layer Problem, 2022, 392, 0010-3616, 377, 10.1007/s00220-022-04321-0
    3. Biagio Cassano, Lucrezia Cossetti, Luca Fanelli, Spectral enclosures for the damped elastic wave equation, 2022, 4, 2640-3501, 1, 10.3934/mine.2022052
    4. Dario Mazzoleni, Benedetta Pellacci, Calculus of variations and nonlinear analysis: advances and applications, 2023, 5, 2640-3501, 1, 10.3934/mine.2023059
    5. Lucrezia Cossetti, Luca Fanelli, Nico M. Schiavone, 2024, Chapter 8, 978-981-97-0363-0, 225, 10.1007/978-981-97-0364-7_8
    6. Lucrezia Cossetti, Luca Fanelli, David Krejčiřík, Uniform resolvent estimates and absence of eigenvalues of biharmonic operators with complex potentials, 2024, 287, 00221236, 110646, 10.1016/j.jfa.2024.110646
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2130) PDF downloads(105) Cited by(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog