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

On Critical Dipoles in Dimensions n3n\geqslant 3

S. Blake Allan Department of Mathematics, Baylor University, Sid Richardson Bldg., 1410 S. 4th Street, Waco, TX 76706, USA [email protected]  and  Fritz Gesztesy Department of Mathematics, Baylor University, Sid Richardson Bldg., 1410 S. 4th Street, Waco, TX 76706, USA [email protected] http://www.baylor.edu/math/index.php?id=935340
Abstract.

We reconsider generalizations of Hardy’s inequality corresponding to the case of (point) dipole potentials Vγ(x)=γ(u,x)|x|3V_{\gamma}(x)=\gamma(u,x)|x|^{-3}, xn\{0}x\in{\mathbb{R}}^{n}\backslash\{0\}, γ[0,)\gamma\in[0,\infty), unu\in{\mathbb{R}}^{n}, |u|=1|u|=1, nn\in{\mathbb{N}}, n3n\geqslant 3. More precisely, for n3n\geqslant 3, we provide an alternative proof of the existence of a critical dipole coupling constant γc,n>0\gamma_{c,n}>0, such that

for all γ[0,γc,n]\gamma\in[0,\gamma_{c,n}], and all unu\in{\mathbb{R}}^{n}, |u|=1|u|=1,
ndnx|(f)(x)|2±γndnx(u,x)|x|3|f(x)|2,fD1(n).\displaystyle\quad\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}\geqslant\pm\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2},\quad f\in D^{1}({\mathbb{R}}^{n}).

with D1(n)D^{1}({\mathbb{R}}^{n}) denoting the completion of C0(n)C_{0}^{\infty}({\mathbb{R}}^{n}) with respect to the norm induced by the gradient. Here γc,n\gamma_{c,n} is sharp, that is, the largest possible such constant. Moreover, we discuss upper and lower bounds for γc,n>0\gamma_{c,n}>0 and develop a numerical scheme for approximating γc,n\gamma_{c,n}.

This quadratic form inequality will be a consequence of the fact

[Δ+γ(u,x)|x|3]|C0(n\{0})¯0 if and only if  0γγc,n\overline{\big{[}-\Delta+\gamma(u,x)|x|^{-3}\big{]}\big{|}_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}}\geqslant 0\,\text{ if and only if }\,0\leqslant\gamma\leqslant\gamma_{c,n}

in L2(n)L^{2}({\mathbb{R}}^{n}) (with T¯\overline{T} the operator closure of the linear operator TT).

We also consider the case of multicenter dipole interactions with dipoles centered on an infinite discrete set.

Key words and phrases:
Hardy-type inequalities, Schrödinger operators, dipole potentials.
2010 Mathematics Subject Classification:
Primary: 35A23, 35J30; Secondary: 47A63, 47F05.
J. Diff. Eq. (to appear).

1. Introduction

The celebrated (multi-dimensional) Hardy inequality,

ndnx|(f)(x)|2[(n2)/2]2ndnx|x|2|f(x)|2,fD1(n),n,n3,\displaystyle\begin{split}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}\geqslant[(n-2)/2]^{2}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f(x)|^{2},&\\ f\in D^{1}({\mathbb{R}}^{n}),\;n\in{\mathbb{N}},\;n\geqslant 3,&\end{split} (1.1)

the first in an infinite sequence of higher-order Birman–Hardy–Rellich-type inequalities, received enormous attention in the literature due to its ubiquity in self-adjointness and spectral theory problems associated with second-order differential operators with strongly singular coefficients, see, for instance, [4], [6, Ch. 1], [20, Sect. 1.5], [21, Ch. 5], [34], [41], [43], [44], [45, Part 1], [50][54], [67, Ch. 2, Sect. 21], [74, Ch. 2], [77] and the extensive literature cited therein. We also note that inequality (LABEL:1.1) is closely related to Heisenberg’s uncertainty relation as discussed in [27].

The basics behind the (point) dipole Hamiltonian Δ+Vγ(x)-\Delta+V_{\gamma}(x), with potential

Vγ(x)=γ(u,x)|x|3,xn\{0},γ[0,),un,|u|=1,n3V_{\gamma}(x)=\gamma\frac{(u,x)}{|x|^{3}},\quad x\in{\mathbb{R}}^{n}\backslash\{0\},\;\gamma\in[0,\infty),\;u\in{\mathbb{R}}^{n},\;|u|=1,\;n\geqslant 3 (1.2)

(with (a,b)(a,b) denoting the Euclidean scalar product of a,bna,b\in{\mathbb{R}}^{n}), in the physically relevant case n=3n=3, have been discussed in great detail in the 1980 paper by Hunziker and Günther [48]. In particular, these authors point out some of the existing fallacies to be found in the physics literature in connection with dipole potentials and their ability to bind electrons. The primary goal in this paper has been the attempt to extend the three-dimensional results on dipole potentials in [48] to the general case n4n\geqslant 4 and thereby rederiving and complementing some of the results obtained by Felli, Marchini, and Terracini [29], [30] (see also [28], [31], [80]). While Felli, Marchini, and Terracini primarily rely on variational techniques, we will focus more on an operator and spectral theoretic approach. To facilitate a comparison between the existing literature on this topic and the results presented in the present paper, we next summarize some of the principal achievements in [28], [29], [30], [48], [80].

However, we first emphasize that these sources also discuss a number of facts that go beyond the scope of our paper: For instance, Hunziker and Günther [48] also consider non-binding criteria for Hamiltonians with MM point charges and applications to electronic spectra of an NN-electron Hamiltonian in the presence of MM point charges (nuclei). In addition, Felli, Marchini and Terracini [29], [30] discuss more general operators where the point dipole potential VγV_{\gamma} in (1.2) is replaced by111For simplicity of notation, we will omit the standard surface measure dn1ωd^{n-1}\omega in L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1}), and similarly, the Lebesgue measure dnxd^{n}x in L2(n)L^{2}({\mathbb{R}}^{n}).

a(x/|x|)|x|2,xn\{0},aL(𝕊n1),a(x/|x|)|x|^{-2},\;x\in{\mathbb{R}}^{n}\backslash\{0\},\quad a\in L^{\infty}({\mathbb{S}}^{n-1}), (1.3)

and hence (1.2) represents the special case

aγ(x/|x|)=γ(u,x/|x|),γ[0,),un,|u|=1,xn\{0}.a_{\gamma}(x/|x|)=\gamma(u,x/|x|),\quad\gamma\in[0,\infty),\;u\in{\mathbb{R}}^{n},\,|u|=1,\;x\in{\mathbb{R}}^{n}\backslash\{0\}. (1.4)

These authors also provide a discussion of strict positivity of the underlying quadratic form Q{γj}1jM(,)Q_{\{\gamma_{j}\}_{1\leqslant j\leqslant M}}(\,\cdot\,,\,\cdot\,), MM\in{\mathbb{N}}, in the multi-center case,

Q{γj}1jM(f,f)=ndnx[|(f)(x)|2+j=1Mγj(u,(xxj))|xxj|3|f(x)|2],γj[0,),xjn,xjxk for jk, 1j,kM,fD1(n),\displaystyle\begin{split}Q_{\{\gamma_{j}\}_{1\leqslant j\leqslant M}}(f,f)=\int_{{\mathbb{R}}^{n}}d^{n}x\,\bigg{[}|(\nabla f)(x)|^{2}+\sum_{j=1}^{M}\gamma_{j}\frac{(u,(x-x_{j}))}{|x-x_{j}|^{3}}|f(x)|^{2}\bigg{]},&\\ \gamma_{j}\in[0,\infty),\;x_{j}\in{\mathbb{R}}^{n},\,x_{j}\neq x_{k}\text{ for }j\neq k,\,1\leqslant j,k\leqslant M,\;f\in D^{1}({\mathbb{R}}^{n}),&\end{split} (1.5)

and its analog with γj(u,(xxj)/|xxj|)\gamma_{j}(u,(x-x_{j})/|x-x_{j}|) replaced by aγj()a_{\gamma_{j}}(\,\cdot\,) restricted to suitable neighborhoods of xjx_{j}, 1jM1\leqslant j\leqslant M. In this context also the problem of “localization of binding”, a notion going back to Ovchinnikov and Sigal [68], is discussed in [30]. In addition, applications to a class of nonlinear PDEs are discussed in [80].

Turning to the topics directly treated in this paper and their relation to results in [28], [29], [30], [48], [80], we start by noting that the dipole-modified Hardy-type inequality reads as follows: For each n3n\geqslant 3, there exists a critical dipole coupling constant γc,n>0\gamma_{c,n}>0, such that

for all γ[0,γc,n], and all un|u|=1,ndnx|(f)(x)|2±γndnx(u,x)|x|3|f(x)|2,fD1(n).\displaystyle\begin{split}&\text{for all $\gamma\in[0,\gamma_{c,n}]$, and all $u\in{\mathbb{R}}^{n}$, $|u|=1$,}\\ &\quad\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}\geqslant\pm\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2},\quad f\in D^{1}({\mathbb{R}}^{n}).\end{split} (1.6)

Here γc,n>0\gamma_{c,n}>0 is optimal, that is, the largest possible such constant, and we recall that D1(n)D^{1}({\mathbb{R}}^{n}) denotes the completion of C0(n)C_{0}^{\infty}({\mathbb{R}}^{n}) with respect to the norm (ndnx|(g)(x)|2)1/2\big{(}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla g)(x)|^{2}\big{)}^{1/2}, gC0(n)g\in C_{0}^{\infty}({\mathbb{R}}^{n}).

The critical constant γc,n\gamma_{c,n} can be characterized by the Rayleigh quotient

γc,n1\displaystyle\gamma_{c,n}^{-1} =supfD1(n){0}{ndnx(u,x)|x|3|f(x)|2ndnx|f(x)|2}\displaystyle=-\underset{f\in D^{1}({\mathbb{R}}^{n})\setminus\{0\}}{\sup}\,\left\{\frac{\int_{{\mathbb{R}}^{n}}\,d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2}}{\int_{{\mathbb{R}}^{n}}d^{n}x\,|\nabla f(x)|^{2}}\right\} (1.7)
=supφH01((0,π))\{0}{0πdθn1[cos(θn1)]|φ(θn1)|2\displaystyle=\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}\Bigg{\{}\int_{0}^{\pi}d\theta_{n-1}\,[-\cos(\theta_{n-1})]|\varphi(\theta_{n-1})|^{2}
×[0πdθn1|φ(θn1)|2+[(n2)(n4)/4][sin(θn1)]2|φ(θn1)|2]1},\displaystyle\quad\times\bigg{[}\int_{0}^{\pi}d\theta_{n-1}\,|\varphi^{\prime}(\theta_{n-1})|^{2}+[(n-2)(n-4)/4][\sin(\theta_{n-1})]^{-2}|\varphi(\theta_{n-1})|^{2}\bigg{]}^{-1}\Bigg{\}},
n3,\displaystyle\hskip 273.14662ptn\geqslant 3, (1.8)

see [29]. To obtain (1.8) one introduces polar coordinates,

x=rω,r=|x|(0,),ω=ω(θ1,,θn1)=x/|x|𝕊n1,x=r\omega,\quad r=|x|\in(0,\infty),\quad\omega=\omega(\theta_{1},\dots,\theta_{n-1})=x/|x|\in{\mathbb{S}}^{n-1}, (1.9)

as in (A.1)–(A.4). By (LABEL:1.1), clearly,

(n2)2/4γc,n,n3.(n-2)^{2}/4\leqslant\gamma_{c,n},\quad n\geqslant 3. (1.10)

The existence of γc,n\gamma_{c,n} as a finite positive number is shown for n=3n=3 in [48] and for nn\in{\mathbb{N}}, n3n\geqslant 3, in [29].

Next, one decomposes

Lγ\displaystyle L_{\gamma} =Δ+Vγ(x),xn\{0},\displaystyle=-\Delta+V_{\gamma}(x),\quad x\in{\mathbb{R}}^{n}\backslash\{0\}, (1.11)
=[d2dr2n1rddr]IL2(𝕊n1)+1r2Λγ,n,r(0,),γ0,n3,\displaystyle=\bigg{[}-\frac{d^{2}}{dr^{2}}-\frac{n-1}{r}\frac{d}{dr}\bigg{]}\otimes I_{L^{2}({\mathbb{S}}^{n-1})}+\frac{1}{r^{2}}\otimes\Lambda_{\gamma,n},\quad r\in(0,\infty),\;\gamma\geqslant 0,\;n\geqslant 3,

acting in L2((0,);rn1dr)L2(𝕊n1)L^{2}((0,\infty);r^{n-1}dr)\otimes L^{2}({\mathbb{S}}^{n-1}), where

Λγ,n=Δ𝕊n1+γcos(θn1),dom(Λγ,n)=dom(Δ𝕊n1),γ0,n3,\displaystyle\Lambda_{\gamma,n}=-\Delta_{{\mathbb{S}}^{n-1}}+\gamma\cos(\theta_{n-1}),\quad\operatorname{dom}(\Lambda_{\gamma,n})=\operatorname{dom}(-\Delta_{{\mathbb{S}}^{n-1}}),\quad\gamma\geqslant 0,\;n\geqslant 3, (1.12)

with Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}} the Laplace–Beltrami operator in L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1}) (see Appendix A). It is shown in [29] that γc,n\gamma_{c,n} is also characterized by

λγc,n,n,0=(n2)2/4,n3,\lambda_{\gamma_{c,n},n,0}=-(n-2)^{2}/4,\quad n\geqslant 3, (1.13)

where λγ,n,0\lambda_{\gamma,n,0} denotes the lowest eigenvalue of Λγ,n\Lambda_{\gamma,n}. We rederive (1.13) via ODE methods and also prove in Theorem 3.1, that

dλγ,n,0dγλγ,n,0γ<0,γ>0,n3\frac{d\lambda_{\gamma,n,0}}{d\gamma}\leqslant\frac{\lambda_{\gamma,n,0}}{\gamma}<0,\quad\gamma>0,\;n\geqslant 3 (1.14)

(this extends the n=3n=3 result in [48] to nn\in{\mathbb{N}}, n3n\geqslant 3) and also provide the two-sided bounds

γ2(n1)2λγ,n,0γ2In/2(2γ/(n1))I(n2)/2(2γ/(n1))<0,γ>0,n3,-\frac{\gamma^{2}}{(n-1)^{2}}\leqslant\lambda_{\gamma,n,0}\leqslant-\frac{\gamma}{2}\frac{I_{n/2}(2\gamma/(n-1))}{I_{(n-2)/2}(2\gamma/(n-1))}<0,\quad\gamma>0,\;n\geqslant 3, (1.15)

with Iν()I_{\nu}(\,\cdot\,) the regular modified Bessel function of order ν\nu\in{\mathbb{C}}. (The additional lower bound γλγ,n,0-\gamma\leqslant\lambda_{\gamma,n,0} is of course evident since 1cos(θn1)-1\leqslant\cos(\theta_{n-1}).) Moreover, employing the fact that

y′′(x)+{[s2(1/4)]/sin2(x)}y(x)=zy(x),z,s[0,),x(0,π),-y^{\prime\prime}(x)+\big{\{}\big{[}s^{2}-(1/4)\big{]}\big{/}\sin^{2}(x)\big{\}}y(x)=zy(x),\quad z\in{\mathbb{C}},\;s\in[0,\infty),\;x\in(0,\pi), (1.16)

is exactly solvable in terms of hypergeometric functions leads to inequality (3.49), and combining the latter with (1.8) enables us to prove the existence of C0(0,)C_{0}\in(0,\infty) such that

γc,n=nC0(n2)(n4)[1+o(1)],\gamma_{c,n}\underset{n\to\infty}{=}C_{0}(n-2)(n-4)[1+o(1)], (1.17)

and the two-sided bounds

15π[(n2)(n4)+4]/32γc,n{1/4,n=3,1,n=4,33/2[(n2)(n4)+1]/8,n5,15\pi[(n-2)(n-4)+4]/32\geqslant\gamma_{c,n}\geqslant\begin{cases}1/4,&n=3,\\ 1,&n=4,\\ 3^{3/2}[(n-2)(n-4)+1]/8,&n\geqslant 5,\end{cases} (1.18)

in Theorem 3.3.

Briefly turning to the content of each section, Section 2 offers a detailed treatment of the angular momentum decomposition of HγH_{\gamma}, the self-adjoint realization of Lγ=Δ+Vγ()L_{\gamma}=-\Delta+V_{\gamma}(\,\cdot\,) (i.e., the Friedrichs extension of Lγ|C0(n\{0})L_{\gamma}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}) in L2(n)L^{2}({\mathbb{R}}^{n}), and hence together with Appendix A on spherical harmonics and the Laplace–Beltrami operator in L2(Sn1)L^{2}(S^{n-1}), n2n\geqslant 2, provides the background information for the bulk of this paper. Equations (1.14), (1.15), (1.17), and (1.18) represent our principal new results in Section 3. Section 4 develops a numerical approach to γc,n\gamma_{c,n} that exhibits γc,n1-\gamma_{c,n}^{-1} as the smallest (negative) eigenvalue of a particular triangular operator K(γc,n)K(\gamma_{c,n}) in 2(0)\ell^{2}({\mathbb{N}}_{0}) with vanishing diagonal elements (cf. (4.20), (4.25)). In addition, we prove that finite truncations of K(γc,n)K(\gamma_{c,n}) yield a convergent and efficient approximation scheme for γc,n\gamma_{c,n}. Finally, Section 5 considers the extension to multicenter dipole Hamiltonians of the form

L{γj}jJ=Δ+jJγj(u,(xxj))|xxj|3χBn(xj;ε/4)(x)+W0(x),L_{\{\gamma_{j}\}_{j\in J}}=-\Delta+\sum_{j\in J}\gamma_{j}\frac{(u,(x-x_{j}))}{|x-x_{j}|^{3}}\chi_{B_{n}(x_{j};\varepsilon/4)}(x)+W_{0}(x), (1.19)

where JJ\subseteq{\mathbb{N}} is an index set, ε>0\varepsilon>0, χBn(x0;η)\chi_{B_{n}(x_{0};\eta)} denotes the characteristic function of the open ball Bn(x0;η)nB_{n}(x_{0};\eta)\subset{\mathbb{R}}^{n} with center x0nx_{0}\in{\mathbb{R}}^{n} and radius η>0\eta>0, {xj}jJn\{x_{j}\}_{j\in J}\subset{\mathbb{R}}^{n}, nn\in{\mathbb{N}}, n3n\geqslant 3, and

infj,jJ|xjxj|ε,0γjγ0<γc,n,jJ,W0L(n), W0 real-valued a.e. on n.\displaystyle\begin{split}&\inf_{j,j^{\prime}\in J}|x_{j}-x_{j^{\prime}}|\geqslant\varepsilon,\quad 0\leqslant\gamma_{j}\leqslant\gamma_{0}<\gamma_{c,n},\;j\in J,\\ &\,W_{0}\in L^{\infty}({\mathbb{R}}^{n}),\,\text{ $W_{0}$ real-valued a.e.~{}on ${\mathbb{R}}^{n}$.}\end{split} (1.20)

In particular, {xj}jJ\{x_{j}\}_{j\in J} is permitted to be an infinite, discrete set, for instance, a lattice. Relying on results proven in [38], we derive the optimal result that L{γj}jJ|C0(n\{xj}jJ)L_{\{\gamma_{j}\}_{j\in J}}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{x_{j}\}_{j\in J})} is bounded from below (resp., essentially self-adjoint) if each individual Lγj|C0(n\{0})L_{\gamma_{j}}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}, jJj\in J, is bounded from below (resp., essentially self-adjoint). This extends results in [30], where JJ is assumed to be finite.

2. The Dipole Hamiltonian

In this section we provide a discussion of the angular momentum decomposition of the nn-dimensional Laplacian Δ-\Delta, introduce the dipole Hamiltonian HγH_{\gamma}, the principal object of this paper, and discuss an analogous decomposition of the latter.

In spherical coordinates (A.1), the Laplace differential expression in nn dimensions takes the form

Δ=2r2n1rr1r2Δ𝕊n1\displaystyle-\Delta=-\frac{{\partial}^{2}}{{\partial}r^{2}}-\frac{n-1}{r}\frac{{\partial}}{{\partial}r}-\frac{1}{r^{2}}\Delta_{{\mathbb{S}}^{n-1}} (2.1)

where Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}} denotes the Laplace–Beltrami operator222We will call Δ-\Delta the Laplacian to guarantee nonnegativity of the underlying L2(n)L^{2}({\mathbb{R}}^{n})-realization (and analogously for the L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1})-realization of the Laplace–Beltrami operator Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}}). associated with the (n1)(n-1)-dimensional unit sphere 𝕊n1{\mathbb{S}}^{n-1} in n{\mathbb{R}}^{n}, see (A.16). When acting in L2(n)L^{2}({\mathbb{R}}^{n}), which in spherical coordinates can be written as L2(n)L2((0,);rn1dr)L2(𝕊n1)L^{2}({\mathbb{R}}^{n})\simeq L^{2}((0,\infty);r^{n-1}dr)\otimes L^{2}({\mathbb{S}}^{n-1}), (2.1) becomes

Δ=[d2dr2n1rddr]IL2(𝕊n1)1r2Δ𝕊n1\displaystyle-\Delta=\bigg{[}-\frac{d^{2}}{dr^{2}}-\frac{n-1}{r}\frac{d}{dr}\bigg{]}\otimes I_{L^{2}({\mathbb{S}}^{n-1})}-\frac{1}{r^{2}}\otimes\Delta_{{\mathbb{S}}^{n-1}} (2.2)

(with I𝒳I_{{\mathcal{X}}} denoting the identity operator in 𝒳{\mathcal{X}}). The Laplace–Beltrami operator Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}} in L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1}), with domain dom(Δ𝕊n1)=H2(𝕊n1)\operatorname{dom}(-\Delta_{{\mathbb{S}}^{n-1}})=H^{2}\big{(}{\mathbb{S}}^{n-1}\big{)} (cf., e.g., [7]), is known to be essentially self-adjoint and nonnegative on C0(𝕊n1)C_{0}^{\infty}({\mathbb{S}}^{n-1}) (cf. [20, Theorem 5.2.3]). Recalling the treatment in [72, p. 160–161], one decomposes the space L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1}) into an infinite orthogonal sum, yielding

L2(n)L2((0,);rn1dr)L2(𝕊n1)==0L2((0,);rn1dr)𝒴n,\displaystyle\begin{split}L^{2}({\mathbb{R}}^{n})&\simeq L^{2}((0,\infty);r^{n-1}dr)\otimes L^{2}({\mathbb{S}}^{n-1})\\ &=\bigoplus\limits_{\ell=0}^{\infty}L^{2}((0,\infty);r^{n-1}dr)\otimes{\mathcal{Y}}_{\ell}^{n},\end{split} (2.3)

where 𝒴n{\mathcal{Y}}_{\ell}^{n} is the eigenspace of Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}} corresponding to the eigenvalue (+n2)\ell(\ell+n-2), 0\ell\in{\mathbb{N}}_{0}, as

σ(Δ𝕊n1)={(+n2)}0.\sigma(-\Delta_{{\mathbb{S}}^{n-1}})=\{\ell(\ell+n-2)\}_{\ell\in{\mathbb{N}}_{0}}. (2.4)

In particular, this results in

Δ==0[d2dr2n1rddr+(+n2)r2]I𝒴n,\displaystyle-\Delta=\bigoplus\limits_{\ell=0}^{\infty}\left[-\frac{d^{2}}{dr^{2}}-\frac{n-1}{r}\frac{d}{dr}+\frac{\ell(\ell+n-2)}{r^{2}}\right]\otimes I_{{\mathcal{Y}}_{\ell}^{n}}, (2.5)

in the space (2.3).

To simplify matters, replacing the measure rn1drr^{n-1}dr by drdr and simultaneously removing the term (n1)r1(d/dr)(n-1)r^{-1}(d/dr), one introduces the unitary operator

Un={L2((0,);rn1dr)L2((0,);dr),f(r)r(n1)/2f(r),\displaystyle U_{n}=\begin{cases}L^{2}((0,\infty);r^{n-1}dr)\rightarrow L^{2}((0,\infty);dr),\\[2.84526pt] f(r)\mapsto r^{(n-1)/2}f(r),\end{cases} (2.6)

under which (2.5) becomes

Δ==0Un1[d2dr2+[(n1)(n3)/4]+(+n2)r2]UnI𝒴n\displaystyle-\Delta=\bigoplus\limits_{\ell=0}^{\infty}U_{n}^{-1}\left[-\frac{d^{2}}{dr^{2}}+\frac{[(n-1)(n-3)/4]+\ell(\ell+n-2)}{r^{2}}\right]U_{n}\otimes I_{{\mathcal{Y}}_{\ell}^{n}} (2.7)

acting in the space (2.3). The precise self-adjoint L2L^{2}-realization of Δ-\Delta in the space (2.3) then is of the form

H0==0Un1hn,UnI𝒴n,H_{0}=\bigoplus\limits_{\ell=0}^{\infty}U_{n}^{-1}h_{n,\ell}\,U_{n}\otimes I_{{\mathcal{Y}}_{\ell}^{n}}, (2.8)

where hn,h_{n,\ell}, 0\ell\in{\mathbb{N}}_{0}, represents the Friedrichs extension of

[d2dr2+[(n1)(n3)/4]+(+n2)r2]|C0((0,)),0,\displaystyle\left[-\frac{d^{2}}{dr^{2}}+\frac{[(n-1)(n-3)/4]+\ell(\ell+n-2)}{r^{2}}\right]\bigg{|}_{C_{0}^{\infty}((0,\infty))},\quad\ell\in{\mathbb{N}}_{0}, (2.9)

in L2((0,);dr)L^{2}((0,\infty);dr). For explicit operator domains and boundary conditions (the latter for n=2,3n=2,3 only) we refer to (2.34)–(2.37). It is well-known (cf. [72, Sect. IX.7, Appendix to X.1]) that

H0=Δ,dom(H0)=H2(n),\displaystyle H_{0}=-\Delta,\quad\operatorname{dom}(H_{0})=H^{2}({\mathbb{R}}^{n}), (2.10)
H0|C0(n) is essentially self-adjoint,\displaystyle H_{0}|_{C_{0}^{\infty}({\mathbb{R}}^{n})}\,\text{ is essentially self-adjoint,} (2.11)
H0|C0(n\{0}) is essentially self-adjoint if and only if n4.\displaystyle H_{0}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}\,\text{ is essentially self-adjoint if and only if $n\geqslant 4$.} (2.12)

Next, we turn to the dipole potential

Vγ(x)=γ(u,x)|x|3,xn\{0},γ0,n2,\displaystyle V_{\gamma}(x)=\gamma\frac{(u,x)}{|x|^{3}},\quad x\in{\mathbb{R}}^{n}\backslash\{0\},\;\gamma\geqslant 0,\;n\geqslant 2, (2.13)

where unu\in{\mathbb{R}}^{n} is a unit vector in the direction of the dipole, the strength of the dipole equals γ0\gamma\geqslant 0, and (,)(\,\cdot\,,\,\cdot\,) represents the Euclidean scalar product in n{\mathbb{R}}^{n}. Upon an appropriate rotation, one can always choose the coordinate system in such a manner that (u,x)=|x|cos(θn1)(u,x)=|x|\cos(\theta_{n-1}), implying

Vγ(x)=γcos(θn1)|x|2,xn\{0},γ0,n2.\displaystyle V_{\gamma}(x)=\gamma\frac{\cos(\theta_{n-1})}{|x|^{2}},\quad x\in{\mathbb{R}}^{n}\backslash\{0\},\;\gamma\geqslant 0,\;n\geqslant 2. (2.14)

In the following we primarily restrict ourselves to the case n3n\geqslant 3 and comment on the exceptional case n=2n=2 at the end of Section 3. The differential expression associated with Hamiltonian for this system then becomes

Lγ=Δ+Vγ(x),xn\{0},γ0,n3,\displaystyle L_{\gamma}=-\Delta+V_{\gamma}(x),\quad x\in{\mathbb{R}}^{n}\backslash\{0\},\;\gamma\geqslant 0,\;n\geqslant 3, (2.15)

acting in L2(n)L^{2}({\mathbb{R}}^{n}). In analogy to (2.2), (2.15) can be represented as

Lγ=[d2dr2n1rddr]IL2(𝕊n1)+1r2Λγ,n,γ0,n3,\displaystyle L_{\gamma}=\bigg{[}-\frac{d^{2}}{dr^{2}}-\frac{n-1}{r}\frac{d}{dr}\bigg{]}\otimes I_{L^{2}({\mathbb{S}}^{n-1})}+\frac{1}{r^{2}}\otimes\Lambda_{\gamma,n},\quad\gamma\geqslant 0,\;n\geqslant 3, (2.16)

acting in L2((0,);rn1dr)L2(𝕊n1)L^{2}((0,\infty);r^{n-1}dr)\otimes L^{2}({\mathbb{S}}^{n-1}), where

Λγ,n=Δ𝕊n1+γcos(θn1),dom(Λγ,n)=dom(Δ𝕊n1),γ0,n3,\displaystyle\Lambda_{\gamma,n}=-\Delta_{{\mathbb{S}}^{n-1}}+\gamma\cos(\theta_{n-1}),\quad\operatorname{dom}(\Lambda_{\gamma,n})=\operatorname{dom}(-\Delta_{{\mathbb{S}}^{n-1}}),\quad\gamma\geqslant 0,\;n\geqslant 3, (2.17)

is self-adjoint in L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1}) (since γcos(θn1)\gamma\cos(\theta_{n-1}) is a bounded self-adjoint operator in L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1})). Applying the angular momentum decomposition to LγL_{\gamma}, but this time with respect to the eigenspaces of Λγ,n\Lambda_{\gamma,n}, then results in

L2(n)=L2((0,);rn1dr)L2(𝕊n1)==0L2((0,);rn1dr)𝒴γ,n,n3,\displaystyle\begin{split}L^{2}({\mathbb{R}}^{n})&=L^{2}((0,\infty);r^{n-1}\,dr)\otimes L^{2}({\mathbb{S}}^{n-1})\\ &=\bigoplus\limits_{\ell=0}^{\infty}L^{2}((0,\infty);r^{n-1}\,dr)\otimes{\mathcal{Y}}_{\gamma,\ell}^{n},\quad n\geqslant 3,\end{split} (2.18)

where 𝒴γ,n{\mathcal{Y}}_{\gamma,\ell}^{n} represents the eigenspace of Λγ,n\Lambda_{\gamma,n} corresponding to the eigenvalue λγ,n,\lambda_{\gamma,n,\ell}, as

σ(Λγ,n)={λγ,n,}0.\sigma(\Lambda_{\gamma,n})=\{\lambda_{\gamma,n,\ell}\}_{\ell\in{\mathbb{N}}_{0}}. (2.19)

We will order the eigenvalues of Λγ,n\Lambda_{\gamma,n} according to magnitude, that is,

λγ,n,λγ,n,+1,γ0,0,n3,\lambda_{\gamma,n,\ell}\leqslant\lambda_{\gamma,n,\ell+1},\quad\gamma\geqslant 0,\;\ell\in{\mathbb{N}}_{0},\;n\geqslant 3, (2.20)

repeating them according to their multiplicity. The analog of (2.7) in the space (2.18) then becomes

Lγ==0Un1[d2dr2+[(n1)(n3)/4]+λγ,n,r2]UnI𝒴γ,n,n3.\displaystyle L_{\gamma}=\bigoplus\limits_{\ell=0}^{\infty}U_{n}^{-1}\left[-\frac{d^{2}}{dr^{2}}+\frac{[(n-1)(n-3)/4]+\lambda_{\gamma,n,\ell}}{r^{2}}\right]U_{n}\otimes I_{{\mathcal{Y}}_{\gamma,\ell}^{n}},\quad n\geqslant 3. (2.21)
Remark 2.1.

Since et(Δ𝕊n1)e^{-t(-\Delta_{{\mathbb{S}}^{n-1}})}, t0t\geqslant 0, has a continuous and nonnegative integral kernel (see, e,g., [20, Theorem 5.2.1]), it is positivity improving in L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1}). Hence, so is etΛγ,ne^{-t\Lambda_{\gamma,n}}, t0t\geqslant 0, by (a special case of) [73, Theorem XIII.45]. Thus, by [73, Theorem XIII.44] one concludes that

the lowest eigenvalue λγ,n,0\lambda_{\gamma,n,0} of Λγ,n\Lambda_{\gamma,n} is simple for all γ0\gamma\geqslant 0. (2.22)

\diamond

In order to deal exclusively with operators which are bounded from below we now make the the following assumption.

Hypothesis 2.2.

Suppose that nn\in{\mathbb{N}}, n3n\geqslant 3, and γ0\gamma\geqslant 0 are such that

λγ,n,0(n2)2/4.\lambda_{\gamma,n,0}\geqslant-(n-2)^{2}/4. (2.23)

Inequality (2.23) is inspired by Hardy’s inequality (LABEL:1.1) (cf. [6, Sect. 1.2], [56, p. 345], [58, Ch. 3], [59, Ch. 1], [67, Ch. 1]), which in turn implies

[d2dr2+cr2]|C0((0,))0 if and only if c1/4.\bigg{[}-\frac{d^{2}}{dr^{2}}+\frac{c}{r^{2}}\bigg{]}\bigg{|}_{C_{0}^{\infty}((0,\infty))}\geqslant 0\,\text{ if and only if $c\geqslant-1/4$.} (2.24)

In fact, “0\geqslant 0” in (2.24) can be replaced by “bounded from below”. Assumption (2.23) is equivalent to

[(n1)(n3)/4]+λγ,n,01/4.[(n-1)(n-3)/4]+\lambda_{\gamma,n,0}\geqslant-1/4. (2.25)
Remark 2.3.

Since the perturbation γcos(θn1)\gamma\cos(\theta_{n-1}), γ[0,)\gamma\in[0,\infty), of Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}} in (2.17) is bounded from below and from above,

γIL2(𝕊n1)γcos(θn1)γIL2(𝕊n1),-\gamma I_{L^{2}({\mathbb{S}}^{n-1})}\leqslant\gamma\cos(\theta_{n-1})\leqslant\gamma I_{L^{2}({\mathbb{S}}^{n-1})}, (2.26)

and Δ𝕊n10-\Delta_{{\mathbb{S}}^{n-1}}\geqslant 0, it is clear that

λγ,n,0γ, that is, Λγ,nγIL2(𝕊n1),\lambda_{\gamma,n,0}\geqslant-\gamma,\,\text{ that is, }\,\Lambda_{\gamma,n}\geqslant-\gamma I_{L^{2}({\mathbb{S}}^{n-1})}, (2.27)

and λ0,n,0=0\lambda_{0,n,0}=0. In particular, for n3n\geqslant 3 and 0γ0\leqslant\gamma sufficiently small, Hypothesis 2.2 will be satisfied. We are particularly interested in the existence of a critical γc,n>0\gamma_{c,n}>0 such that

λγc,n,n,0=(n2)2/4,\lambda_{\gamma_{c,n},n,0}=-(n-2)^{2}/4, (2.28)

and whether or not

λγ,n,0<(n2)2/4,γ(γc,n,γ2),\lambda_{\gamma,n,0}<-(n-2)^{2}/4,\quad\gamma\in(\gamma_{c,n},\gamma_{2}), (2.29)

for a γ2(γc,n,)\gamma_{2}\in(\gamma_{c,n},\infty), with

λγ,n,0(n2)2/4,γ(γ2,γ3),\lambda_{\gamma,n,0}\geqslant-(n-2)^{2}/4,\quad\gamma\in(\gamma_{2},\gamma_{3}), (2.30)

for a γ3(γ2,)\gamma_{3}\in(\gamma_{2},\infty), etc. This will be clarified in the next section (demonstrating that γ2=\gamma_{2}=\infty). \diamond

Given Hypothesis 2.2, the precise self-adjoint L2(n)L^{2}({\mathbb{R}}^{n})-realization of LγL_{\gamma} in the space (2.18) is then of the form

Hγ==0Un1hγ,n,UnI𝒴γ,n,γ0,n3,H_{\gamma}=\bigoplus\limits_{\ell=0}^{\infty}U_{n}^{-1}h_{\gamma,n,\ell}\,U_{n}\otimes I_{{\mathcal{Y}}_{\gamma,\ell}^{n}},\quad\gamma\geqslant 0,\;n\geqslant 3, (2.31)

where hγ,n,h_{\gamma,n,\ell}, 0\ell\in{\mathbb{N}}_{0}, represents the Friedrichs extension of

[d2dr2+[(n1)(n3)/4]+λγ,n,r2]|C0((0,)),r>0,γ0,n3,0,\displaystyle\left[-\frac{d^{2}}{dr^{2}}+\frac{[(n-1)(n-3)/4]+\lambda_{\gamma,n,\ell}}{r^{2}}\right]\bigg{|}_{C_{0}^{\infty}((0,\infty))},\quad r>0,\;\gamma\geqslant 0,\;n\geqslant 3,\;\ell\in{\mathbb{N}}_{0}, (2.32)

in L2((0,);dr)L^{2}((0,\infty);dr). Explicitly, as discussed, for instance, in [37], [40], the Friedrichs extension of hγ,n,h_{\gamma,n,\ell}, 0\ell\in{\mathbb{N}}_{0}, can be determined from the fact that the Friedrichs extension hα,Fh_{\alpha,F} in L2((0,);dr)L^{2}((0,\infty);dr) of

hα=[d2dr2+α2(1/4)r2]|C0((0,)),r>0,α[0,),h_{\alpha}=\bigg{[}-\frac{d^{2}}{dr^{2}}+\frac{\alpha^{2}-(1/4)}{r^{2}}\bigg{]}\bigg{|}_{C_{0}^{\infty}((0,\infty))},\quad r>0,\;\alpha\in[0,\infty), (2.33)

is given by

hα,F=d2dr2+α2(1/4)r2,r>0,α[0,),\displaystyle h_{\alpha,F}=-\frac{d^{2}}{dr^{2}}+\frac{\alpha^{2}-(1/4)}{r^{2}},\quad r>0,\;\alpha\in[0,\infty), (2.34)
dom(hα,F)={fL2((0,);dr)|f,fACloc((0,));f~α(0)=0;\displaystyle\operatorname{dom}\big{(}h_{\alpha,F}\big{)}=\big{\{}f\in L^{2}((0,\infty);dr)\,\big{|}\,f,f^{\prime}\in AC_{loc}((0,\infty));\,\widetilde{f}_{\alpha}(0)=0; (2.35)
(f′′+[α2(1/4)]r2f)L2((0,);dr)},α[0,1),\displaystyle\hskip 71.13188pt(-f^{\prime\prime}+\big{[}\alpha^{2}-(1/4)\big{]}r^{-2}f)\in L^{2}((0,\infty);dr)\big{\}},\quad\alpha\in[0,1),
dom(hα,F)={fL2((0,);dr)|f,fACloc((0,));\displaystyle\operatorname{dom}\big{(}h_{\alpha,F}\big{)}=\big{\{}f\in L^{2}((0,\infty);dr)\,\big{|}\,f,f^{\prime}\in AC_{loc}((0,\infty)); (2.36)
(f′′+[α2(1/4)]r2f)L2((0,);dr)},α[1,),\displaystyle\hskip 71.13188pt(-f^{\prime\prime}+\big{[}\alpha^{2}-(1/4)\big{]}r^{-2}f)\in L^{2}((0,\infty);dr)\big{\}},\quad\alpha\in[1,\infty),

where

f~α(0)={limr0[r1/2ln(1/r)]1f(r),α=0,limr02αrα(1/2)f(r),α(0,1).\widetilde{f}_{\alpha}(0)=\begin{cases}\lim_{r\downarrow 0}\big{[}r^{1/2}\text{\rm ln}(1/r)\big{]}^{-1}f(r),&\alpha=0,\\[2.84526pt] \lim_{r\downarrow 0}2\alpha r^{\alpha-(1/2)}f(r),&\alpha\in(0,1).\end{cases} (2.37)

Next we note the following fact.

Lemma 2.4.

Given the operator Λγ,n\Lambda_{\gamma,n}, γ0\gamma\geqslant 0, in L2(𝕊n1)L^{2}({\mathbb{S}}^{n-1}) as introduced in (2.17), one infers that

limγ0λγ,n,=(+n2),0,\lim_{\gamma\downarrow 0}\lambda_{\gamma,n,\ell}=\ell(\ell+n-2),\quad\ell\in{\mathbb{N}}_{0}, (2.38)

recalling that {(+n2)}0\{\ell(\ell+n-2)\}_{\ell\in{\mathbb{N}}_{0}} are the corresponding eigenvalues of the unperturbed operator, Λ0,n=Δ𝕊n1\Lambda_{0,n}=-\Delta_{{\mathbb{S}}^{n-1}}, the Laplace–Beltrami operator ((cf. (2.4))).

Proof.

This is a special case of Rellich’s theorem in the form recorded, for instance, in [73, Theorems XII.3 and XII.13]. ∎

Lemma 2.5.

Assume Hypothesis 2.2, that is, suppose that

λγ,n,0(n2)2/4,n3.\lambda_{\gamma,n,0}\geqslant-(n-2)^{2}/4,\quad n\geqslant 3. (2.39)

Then HγH_{\gamma} has purely absolutely continuous spectrum,

σ(Hγ)=σac(Hγ)=[0,).\sigma(H_{\gamma})=\sigma_{ac}(H_{\gamma})=[0,\infty). (2.40)
Proof.

First, one notes that HγH_{\gamma} is bounded from below if and only if each hγ,n,h_{\gamma,n,\ell}, 0\ell\in{\mathbb{N}}_{0}, is bounded from below. The ordinary differential operators hγ,n,h_{\gamma,n,\ell}, 0\ell\in{\mathbb{N}}_{0}, are well-known to have purely absolutely continuous spectrum equal to [0,)[0,\infty), as proven, for instance in [25] and [42]. Thus the result follows from the special case of direct sums (instead of direct integrals) in [73, Theorem XIII.85 (f)]. ∎

3. Criticality

We now turn to one of the principal questions – a discussion of which γ0\gamma\geqslant 0 cause HγH_{\gamma} to be bounded from below.

The natural space to which Hardy’s inequality and its analog in connection with a dipole potential extends is the space D1(n)D^{1}({\mathbb{R}}^{n}) (sometimes also denoted D01(n)D_{0}^{1}({\mathbb{R}}^{n}), or D1,2(n)D^{1,2}({\mathbb{R}}^{n})) obtained as the closure of C0(n)C_{0}^{\infty}({\mathbb{R}}^{n}) with respect to the gradient norm,

D1(n)=C0(n)¯,f=(ndnx|(f)(x)|2)1/2,fC0(n),\displaystyle D^{1}({\mathbb{R}}^{n})=\overline{C_{0}^{\infty}({\mathbb{R}}^{n})}^{\|\,\cdot\,\|_{\nabla}},\quad\|f\|_{\nabla}=\bigg{(}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}\bigg{)}^{1/2},\quad f\in C_{0}^{\infty}({\mathbb{R}}^{n}), (3.1)

see also [61, pp. 201–204].

Theorem 3.1.

Assume Hypothesis 2.2. Then for all n3n\geqslant 3, there exists a unique critical dipole moment γc,n>0\gamma_{c,n}>0 characterized by

λγc,n,n,0=(n2)2/4\lambda_{\gamma_{c,n},n,0}=-(n-2)^{2}/4 (3.2)

((cf. (2.28) in Remark 2.3)). Moreover, λγ,n,0\lambda_{\gamma,n,0} is strictly monotonically decreasing with respect to γ(0,)\gamma\in(0,\infty), λ0,n,0=0\lambda_{0,n,0}=0, and

dλγ,n,0dγλγ,n,0γ<0 as well as λγ,n,0γ,γ(0,).\frac{d\lambda_{\gamma,n,0}}{d\gamma}\leqslant\frac{\lambda_{\gamma,n,0}}{\gamma}<0\,\text{ as well as }\,\lambda_{\gamma,n,0}\geqslant-\gamma,\quad\gamma\in(0,\infty). (3.3)

Moreover,

γ2(n1)2λγ,n,0γ2In/2(2γ/(n1))I(n2)/2(2γ/(n1))<0,γ(0,),-\frac{\gamma^{2}}{(n-1)^{2}}\leqslant\lambda_{\gamma,n,0}\leqslant-\frac{\gamma}{2}\frac{I_{n/2}(2\gamma/(n-1))}{I_{(n-2)/2}(2\gamma/(n-1))}<0,\quad\gamma\in(0,\infty), (3.4)

hold. In particular, HγH_{\gamma} is bounded from below, and then Hγ0H_{\gamma}\geqslant 0, if and only if γ[0,γc,n]\gamma\in[0,\gamma_{c,n}]. Consequently,

for all γ[0,γc,n], and all un|u|=1,ndnx|(f)(x)|2±γndnx(u,x)|x|3|f(x)|2,fD1(n).\displaystyle\begin{split}&\text{for all $\gamma\in[0,\gamma_{c,n}]$, and all $u\in{\mathbb{R}}^{n}$, $|u|=1$,}\\ &\quad\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}\geqslant\pm\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2},\quad f\in D^{1}({\mathbb{R}}^{n}).\end{split} (3.5)

The constant γc,n>0\gamma_{c,n}>0 in (3.5) is optimal ((i.e., the largest possible )), in addition,

γc,n(n2)2/4.\gamma_{c,n}\geqslant(n-2)^{2}/4. (3.6)

Finally,

σ(Hγ)=σac(Hγ)=[0,),γ[0,γc,n].\sigma(H_{\gamma})=\sigma_{ac}(H_{\gamma})=[0,\infty),\quad\gamma\in[0,\gamma_{c,n}]. (3.7)
Proof.

Existence of some critical dipole moment γc,n>0\gamma_{c,n}>0 is clear from the discussion in Remark 2.3. To prove the remaining claims regarding λγ,n,0\lambda_{\gamma,n,0} in Theorem 3.1, we seek spherical harmonics dependent only on the final angle θn1\theta_{n-1}, as this is the only angular variable dependence of Vγ()V_{\gamma}(\,\cdot\,). From (A.10)–(A.13), one infers these are precisely the ones indexed by the particular multi-indices (,0,,0)0n(\ell,0,\ldots,0)\in{\mathbb{N}}_{0}^{n}, that is (cf. (A.10)–(A.14)),

Y(,0,,0)(θn1)=[[(n2)/2](n2)!(+[(n2)/2])]1/2C(n2)/2(cos(θn1)),0,θn1[0,π).\displaystyle\begin{split}Y_{(\ell,0,\ldots,0)}(\theta_{n-1})=\bigg{[}\frac{[(n-2)/2](n-2)_{\ell}}{\ell!(\ell+[(n-2)/2])}\bigg{]}^{1/2}C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1})),&\\[2.84526pt] \ell\in{\mathbb{N}}_{0},\;\theta_{n-1}\in[0,\pi).&\end{split} (3.8)

Introducing the subspace

n=lin.span{Y(,0,,0)}0.\displaystyle{\mathcal{L}}^{n}={\rm lin.span}\{Y_{(\ell,0,\ldots,0)}\}_{\ell\in{\mathbb{N}}_{0}}. (3.9)

and restricting the Laplace–Beltrami differential expression (A.16) to n{\mathcal{L}}^{n}, one finds for (2.17),

Λγ,n\displaystyle\Lambda_{\gamma,{\mathcal{L}}^{n}} =d2dθn12(n2)cot(θn1)ddθn1+γcos(θn1),θn1(0,π),\displaystyle=-\frac{d^{2}}{d\theta_{n-1}^{2}}-(n-2)\cot(\theta_{n-1})\frac{d}{d\theta_{n-1}}+\gamma\cos(\theta_{n-1}),\quad\theta_{n-1}\in(0,\pi), (3.10)

acting on functions in L2((0,π);[sin(θn1)]n2dθn1)L^{2}\big{(}(0,\pi);[\sin(\theta_{n-1})]^{n-2}d\theta_{n-1}\big{)}. Reverting from the weighted measure [sin(θn1)]n2dθn1[\sin(\theta_{n-1})]^{n-2}d\theta_{n-1} to Lebesgue measure dθn1d\theta_{n-1} on (0,π)(0,\pi) in a unitary fashion then yields the differential expression Λ~γ,n\widetilde{\Lambda}_{\gamma,{\mathcal{L}}^{n}} given by

Λ~γ,n=d2dθn12+(n2)(n4)4sin2(θn1)(n2)24+γcos(θn1),θn1(0,π),\widetilde{\Lambda}_{\gamma,{\mathcal{L}}^{n}}=-\frac{d^{2}}{d\theta_{n-1}^{2}}+\frac{(n-2)(n-4)}{4\sin^{2}(\theta_{n-1})}-\frac{(n-2)^{2}}{4}+\gamma\cos(\theta_{n-1}),\quad\theta_{n-1}\in(0,\pi), (3.11)

now acting on functions in L2((0,π);dθn1)L^{2}((0,\pi);d\theta_{n-1}).

Next, introducing the change of variable ξ=cos(θn1)(1,1)\xi=\cos(\theta_{n-1})\in(-1,1), Λγ,n\Lambda_{\gamma,{\mathcal{L}}^{n}} in (3.10) turns into

Λ¯γ,n=(1ξ2)(n3)/2[ddξ(1ξ2)(n1)/2ddξ+γ(1ξ2)(n3)/2ξ],ξ(1,1),\displaystyle\begin{split}{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}&=\big{(}1-\xi^{2}\big{)}^{-(n-3)/2}\bigg{[}-\frac{d}{d\xi}\big{(}1-\xi^{2}\big{)}^{(n-1)/2}\frac{d}{d\xi}+\gamma\big{(}1-\xi^{2}\big{)}^{(n-3)/2}\xi\bigg{]},\\ &\hskip 239.00298pt\xi\in(-1,1),\end{split} (3.12)

acting on functions in L2((1,1);(1ξ2)(n3)/2dξ)L^{2}\Big{(}(-1,1);\big{(}1-\xi^{2}\big{)}^{(n-3)/2}d\xi\Big{)}. We also note that reverting from the weighted measure (1ξ2)(n3)/2dξ\big{(}1-\xi^{2}\big{)}^{(n-3)/2}d\xi to Lebesgue measure dξd\xi on (1,1)(-1,1) in a unitary fashion then finally yields the differential expression Λ¯~γ,n\widetilde{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}} given by

Λ¯~γ,n=ddξ(1ξ2)(n1)/2ddξ+(n3)24(1ξ2)(n1)(n3)4+γξ,ξ(1,1),\displaystyle\begin{split}\widetilde{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}&=-\frac{d}{d\xi}\big{(}1-\xi^{2}\big{)}^{(n-1)/2}\frac{d}{d\xi}+\frac{(n-3)^{2}}{4\big{(}1-\xi^{2}\big{)}}-\frac{(n-1)(n-3)}{4}+\gamma\,\xi,\\ &\hskip 210.55022pt\quad\xi\in(-1,1),\end{split} (3.13)

acting on functions in L2((1,1);dξ)L^{2}((-1,1);d\xi). One observes that the first two terms on the right-hand side of (3.13) represent the Legendre operator LμL_{\mu} in L2((1,1);dξ)L^{2}((-1,1);d\xi) associated with the differential expression

Lμ=ddξ(1ξ2)(n1)/2ddξ+μ2(1ξ2),μ[0,),ξ(1,1),L_{\mu}=-\frac{d}{d\xi}\big{(}1-\xi^{2}\big{)}^{(n-1)/2}\frac{d}{d\xi}+\frac{\mu^{2}}{\big{(}1-\xi^{2}\big{)}},\quad\mu\in[0,\infty),\;\xi\in(-1,1), (3.14)

which is in the limit circle case at ±1\pm 1 if μ[0,1)\mu\in[0,1) and in the limit point case at ±1\pm 1 if μ[1,)\mu\in[1,\infty), as discussed in detail in [24]. In particular, applying this fact to Λγ,n\Lambda_{\gamma,{\mathcal{L}}^{n}}, Λ~γ,n\widetilde{\Lambda}_{\gamma,{\mathcal{L}}^{n}}, Λ¯γ,n{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}, and Λ¯~γ,n\widetilde{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}} yields the necessity of the Friedrichs boundary condition for n=3,4n=3,4, whereas for nn\in{\mathbb{N}}, n5n\geqslant 5, Λγ,n\Lambda_{\gamma,{\mathcal{L}}^{n}} and Λ~γ,n\widetilde{\Lambda}_{\gamma,{\mathcal{L}}^{n}} (resp., Λ¯γ,n{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}} and Λ¯~γ,n\widetilde{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}) are essentially self-adjoint on C0((0,π))C_{0}^{\infty}((0,\pi)) (resp., C0((1,1))C_{0}^{\infty}((-1,1))) and hence the associated maximally defined operators are self-adjoint. For the explicit form of the Friedrichs boundary condition corresponding to (3.14) and hence (3.13) we also refer to [24]. Due to the θn12\theta_{n-1}^{-2} (resp., (πθn1)2(\pi-\theta_{n-1})^{-2}) singularity at θn1=0\theta_{n-1}=0 (resp., θn1=π\theta_{n-1}=\pi), the Friedrichs extension corresponding to Λ~γ,n\widetilde{\Lambda}_{\gamma,{\mathcal{L}}^{n}} in (3.11) is clear from (2.33)–(2.37).

Following [48] in the special case n=3n=3, choosing ψdom(Λ¯γ,n)\psi\in\operatorname{dom}\big{(}{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}\big{)} normalized,

ψL2((1,1);(1ξ2)(n3)/2dξ)=1,\|\psi\|_{L^{2}((-1,1);(1-\xi^{2})^{(n-3)/2}d\xi)}=1, (3.15)

an appropriate integration by parts yields

(ψ,Λ¯γ,nψ)L2((1,1);(1ξ2)(n3)/2dξ)\displaystyle(\psi,{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}\psi)_{L^{2}((-1,1);(1-\xi^{2})^{(n-3)/2}d\xi)}
=11𝑑ξ[(1ξ2)(n1)/2|ψ(ξ)|2+γ(1ξ2)(n3)/2ξ|ψ(ξ)|2]\displaystyle\quad=\int_{-1}^{1}d\xi\Big{[}\big{(}1-\xi^{2}\big{)}^{(n-1)/2}|\psi^{\prime}(\xi)|^{2}+\gamma\big{(}1-\xi^{2}\big{)}^{(n-3)/2}\xi|\psi(\xi)|^{2}\Big{]} (3.16)
=11𝑑ξ(1ξ2)(n1)/2[|ψ(ξ)+γ(n1)1ψ(ξ)|2γ2(n1)2|ψ(ξ)|2]\displaystyle\quad=\int_{-1}^{1}d\xi\big{(}1-\xi^{2}\big{)}^{(n-1)/2}\Big{[}\big{|}\psi^{\prime}(\xi)+\gamma(n-1)^{-1}\psi(\xi)\big{|}^{2}-\gamma^{2}(n-1)^{-2}|\psi(\xi)|^{2}\Big{]} (3.17)
γ2(n1)211𝑑ξ(1ξ2)(n1)/2|ψ(ξ)|2\displaystyle\quad\geqslant-\frac{\gamma^{2}}{(n-1)^{2}}\int_{-1}^{1}d\xi\big{(}1-\xi^{2}\big{)}^{(n-1)/2}|\psi(\xi)|^{2}
γ2(n1)211𝑑ξ(1ξ2)(n3)/2|ψ(ξ)|2=γ2(n1)2.\displaystyle\quad\geqslant-\frac{\gamma^{2}}{(n-1)^{2}}\int_{-1}^{1}d\xi\big{(}1-\xi^{2}\big{)}^{(n-3)/2}|\psi(\xi)|^{2}=-\frac{\gamma^{2}}{(n-1)^{2}}. (3.18)

In particular, choosing for ψ\psi a normalized eigenfunction of Λ¯γ,n{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}} corresponding to the eigenvalue λγ,n,0\lambda_{\gamma,n,0} in (3.16) implies the lower bound

λγ,n,0γ2/(n1)2.\lambda_{\gamma,n,0}\geqslant-\gamma^{2}\big{/}(n-1)^{2}. (3.19)

On the other hand (following once more [48] in the special case n=3n=3), employing the normalized trial function (cf. [46, no. 3.387])

ϕγ(ξ)=Cγeγ(n1)1ξ,ξ(1,1),\displaystyle\phi_{\gamma}(\xi)=C_{\gamma}\,e^{-\gamma(n-1)^{-1}\xi},\quad\xi\in(-1,1),
Cγ=π1/4[γ/(n1)](n2)/4[Γ((n1)/2)]1/2[I(n2)/2(2γ/(n1))]1/2,\displaystyle C_{\gamma}=\pi^{-1/4}[\gamma/(n-1)]^{(n-2)/4}[\Gamma((n-1)/2)]^{-1/2}[I_{(n-2)/2}(2\gamma/(n-1))]^{-1/2}, (3.20)
ϕγL2((1,1);(1ξ2)(n3)/2dξ)=1,\displaystyle\|\phi_{\gamma}\|_{L^{2}((-1,1);(1-\xi^{2})^{(n-3)/2}d\xi)}=1,

with Iν()I_{\nu}(\,\cdot\,) the regular modified Bessel function of order ν\nu\in{\mathbb{C}} (cf. [1, Sect. 9.6]), an application of the min/max principle and (3.17) yield the upper bound

λγ,n,0\displaystyle\lambda_{\gamma,n,0} (ϕγ,Λ¯γ,nϕγ)L2((1,1);(1ξ2)(n3)/2dξ)\displaystyle\leqslant(\phi_{\gamma},{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}\phi_{\gamma})_{L^{2}((-1,1);(1-\xi^{2})^{(n-3)/2}d\xi)}
=γ2(n1)211𝑑ξ(1ξ2)(n1)/2ϕγ(ξ)2\displaystyle=-\frac{\gamma^{2}}{(n-1)^{2}}\int_{-1}^{1}d\xi\,\big{(}1-\xi^{2}\big{)}^{(n-1)/2}\phi_{\gamma}(\xi)^{2}
=γ2In/2(2γ/(n1))I(n2)/2(2γ/(n1))<0,γ(0,),\displaystyle=-\frac{\gamma}{2}\frac{I_{n/2}(2\gamma/(n-1))}{I_{(n-2)/2}(2\gamma/(n-1))}<0,\quad\gamma\in(0,\infty), (3.21)

employing [46, no. 3.387] once again. Thus, (3.21) implies that

λγ,n,0<0,γ(0,),\lambda_{\gamma,n,0}<0,\quad\gamma\in(0,\infty), (3.22)

and one infers a quadratic upper bound as γ0\gamma\downarrow 0

γ2In/2(2γ/(n1))I(n2)/2(2γ/(n1))=γ0γ2n(n1)[1+O(γ2)],-\frac{\gamma}{2}\frac{I_{n/2}(2\gamma/(n-1))}{I_{(n-2)/2}(2\gamma/(n-1))}\underset{\gamma\downarrow 0}{=}-\frac{\gamma^{2}}{n(n-1)}\big{[}1+O\big{(}\gamma^{2}\big{)}\big{]}, (3.23)

in addition to the quadratic lower bound in (3.19).

Next, recalling that the lowest eigenvalue λγ,n,0\lambda_{\gamma,n,0} of Λγ,n\Lambda_{\gamma,n} is simple for all γ0\gamma\geqslant 0 (and is also the lowest eigenvalue of Λγ,n\Lambda_{\gamma,{\mathcal{L}}^{n}}, Λ~γ,n\widetilde{\Lambda}_{\gamma,{\mathcal{L}}^{n}}, and Λ¯γ,n{\underline{\Lambda}}_{\gamma,{\mathcal{L}}^{n}}), we denote by ψγ,0dom(Λγ,n)\psi_{\gamma,0}\in\operatorname{dom}(\Lambda_{\gamma,n}) the corresponding normalized eigenfunction, that is,

Λγ,nψγ,0=λγ,n,0ψγ,0,ψγ,0L2(𝕊n1)=1,γ[0,).\Lambda_{\gamma,n}\psi_{\gamma,0}=\lambda_{\gamma,n,0}\psi_{\gamma,0},\quad\|\psi_{\gamma,0}\|_{L^{2}({\mathbb{S}}^{n-1})}=1,\quad\gamma\in[0,\infty). (3.24)

Thus, one gets

λγ,n,0=(ψγ,0,Λγ,nψγ,0)L2(𝕊n1)=(ψγ,0,[Δ𝕊n1+γcos(θn1)]ψγ,0)L2(𝕊n1).\displaystyle\begin{split}\lambda_{\gamma,n,0}&=(\psi_{\gamma,0},\Lambda_{\gamma,n}\psi_{\gamma,0})_{L^{2}({\mathbb{S}}^{n-1})}\\ &=(\psi_{\gamma,0},[-\Delta_{{\mathbb{S}}^{n-1}}+\gamma\cos(\theta_{n-1})]\psi_{\gamma,0})_{L^{2}({\mathbb{S}}^{n-1})}.\end{split} (3.25)

Moreover, one observes that {Λγ,n}γ[0,)\{\Lambda_{\gamma,n}\}_{\gamma\in[0,\infty)} is a self-adjoint analytic (in fact, entire) family of type (A)(A) in the sense of Kato (cf. [56, Sect. VII.2, p. 375–379], [73, p. 16]), implying analyticity of λγ,n,0\lambda_{\gamma,n,0} and ψγ,0\psi_{\gamma,0} with respect to γ\gamma in a complex neighborhood of [0,)[0,\infty). In particular, λγ,n,0\lambda_{\gamma,n,0} is differentiable with respect to γ\gamma, and the Feynman–Hellmann Theorem [82, p. 151] (see also [76, Theorem 1.4.7]) yields that

dλγ,n,0dγ=(ψγ,0,cos(θn1)ψγ,0)L2(𝕊n1),γ(0,).\displaystyle\frac{d\lambda_{\gamma,n,0}}{d\gamma}=(\psi_{\gamma,0},\cos(\theta_{n-1})\psi_{\gamma,0})_{L^{2}({\mathbb{S}}^{n-1})},\quad\gamma\in(0,\infty). (3.26)

Returning to the discussion of (2.27) in Remark 2.3, employing Δ𝕊n10-\Delta_{{\mathbb{S}}^{n-1}}\geqslant 0, one obtains

λγ,n,0=(ψγ,0,Λγ,nψγ,0)L2(𝕊n1)(ψγ,0,γcos(θn1)ψγ,0)L2(𝕊n1)γ,\lambda_{\gamma,n,0}=(\psi_{\gamma,0},\Lambda_{\gamma,n}\psi_{\gamma,0})_{L^{2}({\mathbb{S}}^{n-1})}\geqslant(\psi_{\gamma,0},\gamma\cos(\theta_{n-1})\psi_{\gamma,0})_{L^{2}({\mathbb{S}}^{n-1})}\geqslant-\gamma, (3.27)

implying,

dλγ,n,0dγ=(ψγ,0,cos(θn1)ψγ,0)L2(𝕊n1)λγ,n,0γ<0,γ(0,),\displaystyle\frac{d\lambda_{\gamma,n,0}}{d\gamma}=(\psi_{\gamma,0},\cos(\theta_{n-1})\psi_{\gamma,0})_{L^{2}({\mathbb{S}}^{n-1})}\leqslant\frac{\lambda_{\gamma,n,0}}{\gamma}<0,\quad\gamma\in(0,\infty), (3.28)

by the strict negativity of λγ,n,0\lambda_{\gamma,n,0} for γ>0\gamma>0 derived in (3.21).

Given the existence of a unique critical dipole moment γc,n>0\gamma_{c,n}>0 one concludes from (2.24), (2.31), and (2.32) the following fact:

Hγ|C0(n\{0}) is bounded from below, in fact, nonnegative,if and only if γ[0,γc,n],\displaystyle\begin{split}&H_{\gamma}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}\,\text{ is bounded from below, in fact, nonnegative,}\\ &\quad\text{if and only if }\,\gamma\in[0,\gamma_{c,n}],\end{split} (3.29)

and an integration by parts thus yields

±γndnx(u,x)|x|3|g(x)|2ndnx|(g)(x)|2,gC0(n),γ[0,γc,n].\pm\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|g(x)|^{2}\leqslant\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla g)(x)|^{2},\quad g\in C_{0}^{\infty}({\mathbb{R}}^{n}),\quad\gamma\in[0,\gamma_{c,n}]. (3.30)

It remains to extend (3.30) to elements fD1()f\in D^{1}({\mathbb{R}}). As in the case of the Hardy inequality (LABEL:1.1), this follows from invoking a Fatou-type argument to be outlined next.

Since C0(n)C_{0}^{\infty}({\mathbb{R}}^{n}) is dense in D1(n)D^{1}({\mathbb{R}}^{n}), given fD1(n)f\in D^{1}({\mathbb{R}}^{n}) we pick a sequence {fj}jC0(n)\{f_{j}\}_{j\in{\mathbb{N}}}\subset C_{0}^{\infty}({\mathbb{R}}^{n}) such that limjfjfD1(n)=0\lim_{j\to\infty}\|f_{j}-f\|_{D^{1}({\mathbb{R}}^{n})}=0, and, by passing to a subsequence, we may assume without loss of generality (see (3.31) below) that fjjff_{j}\underset{j\to\infty}{\longrightarrow}f a.e. on n{\mathbb{R}}^{n}. (For the remainder of this proof f,fjf,f_{j}, jj\in{\mathbb{N}}, will always be assumed to have the properties just discussed.) Indeed, the Sobolev inequality (see, e.g., [61, Theorem 8.3], [79]),

fL2(n)2SnfL2(n)2,fD1(n),2=2n/(n2),Sn=[n(n2)/4]22/nπ(n+1)/nΓ((n+1)/2)2/n,n3,\displaystyle\begin{split}&\|\nabla f\|_{L^{2}({\mathbb{R}}^{n})}^{2}\geqslant S_{n}\|f\|_{L^{2^{*}}({\mathbb{R}}^{n})}^{2},\quad f\in D^{1}({\mathbb{R}}^{n}),\quad 2^{*}=2n/(n-2),\\ &\,S_{n}=[n(n-2)/4]2^{2/n}\pi^{(n+1)/n}\Gamma((n+1)/2)^{-2/n},\quad n\geqslant 3,\end{split} (3.31)

(Γ()\Gamma(\,\cdot\,) the Gamma function, cf. [1, Sect. 6.1]), yields convergence of fjf_{j} to ff in L2(n)L^{2^{*}}({\mathbb{R}}^{n}) and hence permits the selection of a subsequence that converges pointwise a.e. Thus, given Hardy’s inequality for functions in C0(n)C_{0}^{\infty}({\mathbb{R}}^{n}), a well-known fact (see, e.g., [6, Corollary 1.2.6]),

[(n2)2/4]ndnx|x|2|g(x)|2ndnx|(g)(x)|2,gC0(n),\big{[}(n-2)^{2}/4\big{]}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|g(x)|^{2}\leqslant\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla g)(x)|^{2},\quad g\in C_{0}^{\infty}({\mathbb{R}}^{n}), (3.32)

one obtains,

[(n2)2/4]ndnx|x|2|fj(x)|2ndnx|[(fjf+f)](x)|2\displaystyle\big{[}(n-2)^{2}/4\big{]}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f_{j}(x)|^{2}\leqslant\int_{{\mathbb{R}}^{n}}d^{n}x\,|[\nabla(f_{j}-f+f)](x)|^{2}
2ndnx|[(fjf)](x)|2+2ndnx|(f)(x)|2C,\displaystyle\quad\leqslant 2\int_{{\mathbb{R}}^{n}}d^{n}x\,|[\nabla(f_{j}-f)](x)|^{2}+2\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}\leqslant C, (3.33)

for some C(0,)C\in(0,\infty) independent of jj\in{\mathbb{N}}. Thus,

[(n2)2/4]ndnx|x|2|f(x)|2C,fD1(n),\big{[}(n-2)^{2}/4\big{]}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f(x)|^{2}\leqslant C,\quad f\in D^{1}({\mathbb{R}}^{n}), (3.34)

by a consequence of Fatou’s Lemma (see, e.g., [61, p. 21]. Hence,

[(n2)2/4]ndnx|x|2|f(x)|2=[(n2)2/4]ndnxlimj|x|2|fj(x)|2\displaystyle\big{[}(n-2)^{2}/4\big{]}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f(x)|^{2}=\big{[}(n-2)^{2}/4\big{]}\int_{{\mathbb{R}}^{n}}d^{n}x\,\lim_{j\to\infty}|x|^{-2}|f_{j}(x)|^{2}
=[(n2)2/4]ndnxlim infj|x|2|fj(x)|2\displaystyle\quad=\big{[}(n-2)^{2}/4\big{]}\int_{{\mathbb{R}}^{n}}d^{n}x\,\liminf_{j\to\infty}|x|^{-2}|f_{j}(x)|^{2}
[(n2)2/4]lim infjndnx|x|2|fj(x)|2(by Fatou’s Lemma)\displaystyle\quad\leqslant\big{[}(n-2)^{2}/4\big{]}\liminf_{j\to\infty}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f_{j}(x)|^{2}\quad\text{(by Fatou's Lemma)}
lim infjndnx|(fj)(x)|2(by (3.32))\displaystyle\quad\leqslant\liminf_{j\to\infty}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f_{j})(x)|^{2}\quad\text{(by \eqref{3.22})}
=limjndnx|(fj)(x)|2=ndnx|(f)(x)|2,\displaystyle\quad=\lim_{j\to\infty}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f_{j})(x)|^{2}=\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}, (3.35)

extends Hardy’s inequality (3.32) from C0(n)C_{0}^{\infty}({\mathbb{R}}^{n}) to D1(n)D^{1}({\mathbb{R}}^{n}). Hardy’s inequality on D1(n)D^{1}({\mathbb{R}}^{n}) also implies that

limjndnx|x|2|f(x)fj(x)|2=0,\lim_{j\to\infty}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f(x)-f_{j}(x)|^{2}=0, (3.36)

in particular,

limjndnx|x|2|fj(x)|2=ndnx|x|2|f(x)|2.\lim_{j\to\infty}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f_{j}(x)|^{2}=\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f(x)|^{2}. (3.37)

Since

|(u,x)||x|11,xn\{0},un,|u|=1,|(u,x)||x|^{-1}\leqslant 1,\quad x\in{\mathbb{R}}^{n}\backslash\{0\},\quad u\in{\mathbb{R}}^{n},\;|u|=1, (3.38)

(3.34) also implies

[(n2)2/4]ndnx|(u,x)||x|3|f(x)|2C,fD1(n),\big{[}(n-2)^{2}/4\big{]}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(u,x)||x|^{-3}|f(x)|^{2}\leqslant C,\quad f\in D^{1}({\mathbb{R}}^{n}), (3.39)

similarly, (3.36), (3.37), and Hölder’s inequality imply

limjndnx(u,x)|x|3|fj(x)|2=ndnx(u,x)|x|3|f(x)|2.\lim_{j\to\infty}\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f_{j}(x)|^{2}=\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2}. (3.40)

Thus, for γ[0,γc,n]\gamma\in[0,\gamma_{c,n}],

±γndnx(u,x)|x|3|f(x)|2=±limjγndnx(u,x)|x|3|fj(x)|2(by (3.40))\displaystyle\pm\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2}=\pm\lim_{j\to\infty}\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f_{j}(x)|^{2}\quad\text{(by \eqref{3.24c})}
limjndnx|(fj)(x)|2(by (3.30))\displaystyle\quad\leqslant\lim_{j\to\infty}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f_{j})(x)|^{2}\quad\text{(by \eqref{3.24AA})}
=ndnx|(f)(x)|2,\displaystyle\quad=\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}, (3.41)

finally implying (3.5). Moreover, (3.38) also yields

ndnx|(f)(x)|2\displaystyle\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2} [(n2)/2]2ndnx|x|2|f(x)|2\displaystyle\geqslant[(n-2)/2]^{2}\int_{{\mathbb{R}}^{n}}d^{n}x\,|x|^{-2}|f(x)|^{2} (3.42)
[(n2)/2]2ndnx|(u,x)||x|3|f(x)|2,fD1(n),\displaystyle\geqslant[(n-2)/2]^{2}\int_{{\mathbb{R}}^{n}}d^{n}x\,|(u,x)||x|^{-3}|f(x)|^{2},\quad f\in D^{1}({\mathbb{R}}^{n}),

and hence (3.6).

Finally, (3.7) is clear from Lemma 2.5 and the strict monotonicity of λγ,n,0\lambda_{\gamma,n,0} with respect to γ0\gamma\geqslant 0. ∎

Remark 3.2.

(i)(i) Theorem 3.1 demonstrates that γ2=\gamma_{2}=\infty in Remark 2.3.
(ii)(ii) Inequality (3.6), that is, γc,n(n2)2/4\gamma_{c,n}\geqslant(n-2)^{2}/4, shows that γc,n\gamma_{c,n} grows at least like cn2cn^{2} for appropriate c>0c>0 as nn\to\infty. \diamond

Next, we improve upon Remark 3.2(ii)(ii) for n5n\geqslant 5 as follows:

Theorem 3.3.

Assume Hypothesis 2.2. Then there exists C0(0,)C_{0}\in(0,\infty) such that

γc,n=nC0(n2)(n4)[1+o(1)],\gamma_{c,n}\underset{n\to\infty}{=}C_{0}(n-2)(n-4)[1+o(1)], (3.43)

in addition,

15π[(n2)(n4)+4]/32γc,n{1/4,n=3,1,n=4,33/2[(n2)(n4)+1]/8,n5,15\pi[(n-2)(n-4)+4]/32\geqslant\gamma_{c,n}\geqslant\begin{cases}1/4,&n=3,\\ 1,&n=4,\\ 3^{3/2}[(n-2)(n-4)+1]/8,&n\geqslant 5,\end{cases} (3.44)
Proof.

Employing [29, eq. (1) and Remark 1], one considers the Rayleigh quotient

Γn(γ(u,x)|x|1)=γsupfD1(n){0}{ndnx(u,x)|x|3|f(x)|2ndnx|f(x)|2},γ(0,),n3,\displaystyle\begin{split}\Gamma_{n}\big{(}\gamma(u,x)|x|^{-1}\big{)}=-\gamma\underset{f\in D^{1}({\mathbb{R}}^{n})\setminus\{0\}}{\sup}\,\left\{\frac{\int_{{\mathbb{R}}^{n}}\,d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2}}{\int_{{\mathbb{R}}^{n}}d^{n}x\,|\nabla f(x)|^{2}}\right\},&\\ \gamma\in(0,\infty),\;n\geqslant 3,&\end{split} (3.45)

and notes that Γn(γ(u,x)|x|1)1\Gamma_{n}\big{(}\gamma(u,x)|x|^{-1}\big{)}\uparrow 1 as γγc,n\gamma\uparrow\gamma_{c,n}, implying (cf. (3.11))

γc,n1\displaystyle\gamma_{c,n}^{-1} =supφH01((0,π))\{0}{0πdθn1[cos(θn1)]|φ(θn1)|2\displaystyle=\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}\Bigg{\{}\int_{0}^{\pi}d\theta_{n-1}\,[-\cos(\theta_{n-1})]|\varphi(\theta_{n-1})|^{2}
×[0πdθn1|φ(θn1)|2+[(n2)(n4)/4][sin(θn1)]2|φ(θn1)|2]1},\displaystyle\quad\times\bigg{[}\int_{0}^{\pi}d\theta_{n-1}\,|\varphi^{\prime}(\theta_{n-1})|^{2}+[(n-2)(n-4)/4][\sin(\theta_{n-1})]^{-2}|\varphi(\theta_{n-1})|^{2}\bigg{]}^{-1}\Bigg{\}},
n3.\displaystyle\hskip 270.30118ptn\geqslant 3. (3.46)

Employing the fact that

(d2dx2+s2(1/4)sin2(x))|C0((0,π))[(1/2)+s]2IL2((0,π);dx),s0,\bigg{(}-\frac{d^{2}}{dx^{2}}+\frac{s^{2}-(1/4)}{\sin^{2}(x)}\bigg{)}\bigg{|}_{C_{0}^{\infty}((0,\pi))}\geqslant[(1/2)+s]^{2}I_{L^{2}((0,\pi);dx)},\quad s\geqslant 0, (3.47)

(this follows from [36, Sect. 4] for s>0s>0 and extends to s=0s=0 utilizing [37, Subsect. 6.1]) one concludes the following variant of Hardy’s inequality (upon taking s=0s=0) with optimal constants 1/41/4,

0π𝑑x|φ(x)|2140π𝑑x|φ(x)|2sin2(x)+140π𝑑x|φ(x)|2,φC0((0,π)),\int_{0}^{\pi}dx\,|\varphi^{\prime}(x)|^{2}\geqslant\frac{1}{4}\int_{0}^{\pi}dx\,\frac{|\varphi(x)|^{2}}{\sin^{2}(x)}+\frac{1}{4}\int_{0}^{\pi}dx\,|\varphi(x)|^{2},\quad\varphi\in C_{0}^{\infty}((0,\pi)), (3.48)

which, by a density argument, extends to

0π𝑑x|φ(x)|2140π𝑑x|φ(x)|2sin2(x)+140π𝑑x|φ(x)|2,φH01((0,π))\int_{0}^{\pi}dx\,|\varphi^{\prime}(x)|^{2}\geqslant\frac{1}{4}\int_{0}^{\pi}dx\,\frac{|\varphi(x)|^{2}}{\sin^{2}(x)}+\frac{1}{4}\int_{0}^{\pi}dx\,|\varphi(x)|^{2},\quad\varphi\in H_{0}^{1}((0,\pi)) (3.49)

(see [39]). Thus, employing (3.49) in (3.46) yields

γc,n1\displaystyle\gamma_{c,n}^{-1} supφH01((0,π))\{0}{0πdθn1[cos(θn1)]|φ(θn1)|2\displaystyle\leqslant\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}\Bigg{\{}\int_{0}^{\pi}d\theta_{n-1}\,[-\cos(\theta_{n-1})]|\varphi(\theta_{n-1})|^{2}
×[0πdθn1{(1/4)+[(n2)(n4)/4]}[sin(θn1)]2|φ(θn1)|2]1}\displaystyle\quad\times\bigg{[}\int_{0}^{\pi}d\theta_{n-1}\,\big{\{}(1/4)+[(n-2)(n-4)/4]\big{\}}[\sin(\theta_{n-1})]^{-2}|\varphi(\theta_{n-1})|^{2}\bigg{]}^{-1}\Bigg{\}}
4(n2)(n4)+1supφH01((0,π))\{0}{0π[sin(θn1)]2dθn1\displaystyle\leqslant\frac{4}{(n-2)(n-4)+1}\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}\Bigg{\{}\int_{0}^{\pi}[\sin(\theta_{n-1})]^{-2}d\theta_{n-1}\,
×[cos(θn1)][sin(θn1)]2|φ(θn1)|2\displaystyle\hskip 163.60333pt\times[-\cos(\theta_{n-1})][\sin(\theta_{n-1})]^{2}|\varphi(\theta_{n-1})|^{2}
×[0π[sin(θn1)]2dθn1|φ(θn1)|2]1}\displaystyle\hskip 113.81102pt\times\bigg{[}\int_{0}^{\pi}[\sin(\theta_{n-1})]^{-2}d\theta_{n-1}\,|\varphi(\theta_{n-1})|^{2}\bigg{]}^{-1}\Bigg{\}} (3.50)
=4(n2)(n4)+1supφH01((0,π))\{0}{π/2π[sin(θn1)]2dθn1\displaystyle=\frac{4}{(n-2)(n-4)+1}\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}\Bigg{\{}\int_{\pi/2}^{\pi}[\sin(\theta_{n-1})]^{-2}d\theta_{n-1}
×[cos(θn1)][sin(θn1)]2|φ(θn1)|2\displaystyle\hskip 163.60333pt\times[-\cos(\theta_{n-1})][\sin(\theta_{n-1})]^{2}|\varphi(\theta_{n-1})|^{2}
×[π/2π[sin(θn1)]2dθn1|φ(θn1)|2]1}\displaystyle\hskip 139.41832pt\times\bigg{[}\int_{\pi/2}^{\pi}[\sin(\theta_{n-1})]^{-2}d\theta_{n-1}\,|\varphi(\theta_{n-1})|^{2}\bigg{]}^{-1}\Bigg{\}}
[8/33/2][(n2)(n4)+1]1,n4.\displaystyle\leqslant\big{[}8\big{/}3^{3/2}\big{]}[(n-2)(n-4)+1]^{-1},\quad n\geqslant 4. (3.51)

Here we used the estimate,

cos(θ)sin2(θ)2/33/2,θ[π/2,π],-\cos(\theta)\sin^{2}(\theta)\leqslant 2\big{/}3^{3/2},\quad\theta\in[\pi/2,\pi], (3.52)

and the fact that due to the sign change of cos(θ)\cos(\theta) as θ\theta crosses π/2\pi/2, the numerator in (3.50) diminishes and the denominator in (3.50) increases, altogether diminishing the ratio in (3.50) if φ()\varphi(\,\cdot\,) has support in [0,π/2][0,\pi/2]. Thus, one is justified assuming that φ()\varphi(\,\cdot\,) has support in [π/2,π][\pi/2,\pi] only.

In the case n=3n=3, the factor [(n2)(n4)+1]/4[(n-2)(n-4)+1]/4 in ](3.50) vanishes, and hence we now employ the additional term φL2((0,π);dx)2/4\|\varphi\|^{2}_{L^{2}((0,\pi);dx)}/4 in (3.49) to arrive at

γc,31\displaystyle\gamma_{c,3}^{-1} supφH01((0,π))\{0}{0π𝑑θ2[cos(θ2)]|φ(θ2)|2[140π𝑑θ2|φ(θ2)|2]1}\displaystyle\leqslant\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}\Bigg{\{}\int_{0}^{\pi}d\theta_{2}\,[-\cos(\theta_{2})]|\varphi(\theta_{2})|^{2}\bigg{[}\frac{1}{4}\int_{0}^{\pi}d\theta_{2}\,|\varphi(\theta_{2})|^{2}\bigg{]}^{-1}\Bigg{\}}
4supφH01((0,π))\{0}{π/2π𝑑θ2[cos(θ2)]|φ(θ2)|2/π/2π𝑑θ2|φ(θ2)|2}\displaystyle\leqslant 4\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}\Bigg{\{}\int_{\pi/2}^{\pi}d\theta_{2}\,[-\cos(\theta_{2})]|\varphi(\theta_{2})|^{2}\bigg{/}\int_{\pi/2}^{\pi}d\theta_{2}\,|\varphi(\theta_{2})|^{2}\Bigg{\}}
4.\displaystyle\leqslant 4. (3.53)

Altogether, this implies the lower bound in (3.44) and hence improves on Remark 3.2(ii)(ii) for n5n\geqslant 5. (For n=4n=4 one can include the term φL2((0,π));dx)2/4\|\varphi\|^{2}_{L^{2}((0,\pi));dx)}/4 to improve the lower bound, but the actual details become so unwieldy that we refrain from doing so.) For n=3,4n=3,4 we just recalled (3.6).

Next, introducing the functionals

Fn(φ)=0π𝑑θn1[cos(θn1)]|φ(θn1)|2\displaystyle F_{n}(\varphi)=\int_{0}^{\pi}d\theta_{n-1}\,[-\cos(\theta_{n-1})]|\varphi(\theta_{n-1})|^{2}
×[0π𝑑θn1{|φ(θn1)|2+[(n2)(n4)/4][sin(θn1)]2|φ(θn1)|2}]1,\displaystyle\quad\times\bigg{[}\int_{0}^{\pi}d\theta_{n-1}\,\big{\{}|\varphi^{\prime}(\theta_{n-1})|^{2}+[(n-2)(n-4)/4][\sin(\theta_{n-1})]^{-2}|\varphi(\theta_{n-1})|^{2}\big{\}}\bigg{]}^{-1},
φH01((0,π))\{0},n,n3,\displaystyle\hskip 156.49014pt\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\},\;n\in{\mathbb{N}},\;n\geqslant 3, (3.54)

one concludes as in (3.51) that

Fn(φ)π/2π[sin(θn1)]2𝑑θn1[cos(θn1)]sin2(θn1)|φ(θn1)|2π/2π[sin(θn1)]2𝑑θn1|φ(θn1)|22/33/2,n,n3,\displaystyle\begin{split}F_{n}(\varphi)&\leqslant\frac{\int_{\pi/2}^{\pi}[\sin(\theta_{n-1})]^{-2}d\theta_{n-1}\,[-\cos(\theta_{n-1})]\sin^{2}(\theta_{n-1})|\varphi(\theta_{n-1})|^{2}}{\int_{\pi/2}^{\pi}[\sin(\theta_{n-1})]^{-2}d\theta_{n-1}\,|\varphi(\theta_{n-1})|^{2}}\\ &\leqslant 2\big{/}3^{3/2},\quad n\in{\mathbb{N}},\;n\geqslant 3,\end{split} (3.55)

is uniformly bounded with respect to nn and strictly monotonically decreasing with respect to nn. Consequently, also

γc,n1(n2)(n4)/4=supφH01((0,π))\{0}Fn(φ)\gamma_{c,n}^{-1}(n-2)(n-4)/4=\underset{\varphi\in H_{0}^{1}((0,\pi))\backslash\{0\}}{\sup}F_{n}(\varphi) (3.56)

is bounded and monotonically decreasing with respect to nn and hence has a limit as nn\to\infty, proving (3.43).

Finally, to prove the upper bound in (3.44) one can argue as follows. Introducing φ0H01((0,π))\{0}\varphi_{0}\in H^{1}_{0}((0,\pi))\backslash\{0\} via

φ0(θ)={0,θ[0,π/2],sin(2θ),θ[π/2,π],\varphi_{0}(\theta)=\begin{cases}0,&\theta\in[0,\pi/2],\\ \sin(2\theta),&\theta\in[\pi/2,\pi],\end{cases} (3.57)

then,

π/2π𝑑θ[cos(θ)]sin2(2θ)=8/15,\displaystyle\int_{\pi/2}^{\pi}d\theta\,[-\cos(\theta)]\sin^{2}(2\theta)=8/15,
π/2π𝑑θ{4cos2(2θ)+[(n2)(n4)/4][sin(θ)]24sin2(θ)cos2(θ)}\displaystyle\int_{\pi/2}^{\pi}d\theta\,\big{\{}4\cos^{2}(2\theta)+[(n-2)(n-4)/4][\sin(\theta)]^{-2}4\sin^{2}(\theta)\cos^{2}(\theta)\big{\}} (3.58)
=π/2π𝑑θ[4cos2(2θ)+(n2)(n4)cos2(θ)]=π[4+(n2)(n4)]/4,\displaystyle\quad=\int_{\pi/2}^{\pi}d\theta\,\big{[}4\cos^{2}(2\theta)+(n-2)(n-4)\cos^{2}(\theta)\big{]}=\pi[4+(n-2)(n-4)]/4,

and hence (cf. (3.46))

γc,n13215π[(n2)(n4)+4],n3,\gamma_{c,n}^{-1}\geqslant\frac{32}{15\pi[(n-2)(n-4)+4]},\quad n\geqslant 3, (3.59)

completes the proof of (3.44). ∎

Remark 3.4.

(i)(i) Since H01((0,π))H^{1}_{0}((0,\pi)) embeds compactly into L2((0,π);dθ)L^{2}((0,\pi);d\theta), the supremum in (3.46) (unlike that in (LABEL:3.40a)) is actually attained, that is, for a particular φnH01((0,π))\{0}\varphi_{n}\in H^{1}_{0}((0,\pi))\backslash\{0\},

γc,n1\displaystyle\gamma_{c,n}^{-1} =0π𝑑θn1[cos(θn1)]|φn(θn1)|2\displaystyle=\int_{0}^{\pi}d\theta_{n-1}\,[-\cos(\theta_{n-1})]|\varphi_{n}(\theta_{n-1})|^{2}
×[0πdθn1|φ(θn1)|2+[(n2)(n4)/4][sin(θn1)]2|φn(θn1)|2]1},\displaystyle\quad\times\bigg{[}\int_{0}^{\pi}d\theta_{n-1}\,|\varphi^{\prime}(\theta_{n-1})|^{2}+[(n-2)(n-4)/4][\sin(\theta_{n-1})]^{-2}|\varphi_{n}(\theta_{n-1})|^{2}\bigg{]}^{-1}\Bigg{\}},
n3.\displaystyle\hskip 256.0748pt\quad n\geqslant 3. (3.60)

However, since the nn-dependence of φn\varphi_{n} appears to be beyond our control, computing the exact value of C0C_{0} in (3.43) remains elusive.
(ii)(ii) The differential equation underlying (3.46) is of the type

y′′(θ)+[(n2)(n4)/4][sin(θ)]2y(θ)=γc,ncos(θ)y(θ),θ(0,π),-y^{\prime\prime}(\theta)+[(n-2)(n-4)/4][\sin(\theta)]^{-2}y(\theta)=-\gamma_{c,n}\cos(\theta)y(\theta),\quad\theta\in(0,\pi), (3.61)

which naturally leads to the Birman–Schwinger-type eigenvalue problem

(hn1/2[cos(θ)]hn1/2v)(θ)=λnv(θ),v=hn1/2y,\displaystyle\Big{(}h_{n}^{-1/2}[-\cos(\theta)]h_{n}^{-1/2}v\Big{)}(\theta)=\lambda_{n}v(\theta),\quad v=h_{n}^{1/2}y, (3.62)

where hnh_{n} denotes the Friedrichs extension of the preminimal operator h.n,min\overset{\textbf{\Large.}}{h}_{n,min} in L2((0,π);dθ)L^{2}((0,\pi);d\theta) defined by

(h.n,ming)(θ)=g′′(θ)+[(n2)(n4)/4][sin(θ)]2g(θ),gC0((0,π)).\big{(}\overset{\textbf{\Large.}}{h}_{n,min}g\big{)}(\theta)=-g^{\prime\prime}(\theta)+[(n-2)(n-4)/4][\sin(\theta)]^{-2}g(\theta),\quad g\in C_{0}^{\infty}((0,\pi)). (3.63)

One observes that h.n,min\overset{\textbf{\Large.}}{h}_{n,min} is essentially self-adjoint for n5n\geqslant 5 and hence boundary conditions at θ=0,π\theta=0,\pi, familiar for singular second-order differential operators of Bessel-type (see [37, Subsection 6.1]), are only required for n=3,4n=3,4. The Birman–Schwinger operator

Tn=hn1/2[cos(θ)]hn1/2T_{n}=h_{n}^{-1/2}[-\cos(\theta)]h_{n}^{-1/2} (3.64)

in L2((0,π);dθ)L^{2}((0,\pi);d\theta) is compact (in fact, Hilbert–Schmidt) upon inspecting its integral kernel and hence by the Raleigh–Ritz quotient in (3.46), γc,n1\gamma_{c,n}^{-1} is the largest eigenvalue for TnT_{n}. Finally, introducing the unitary operator

(Uf)(θ)=f(πθ),θ(0,π),fL2((0,π);dθ),(Uf)(\theta)=f(\pi-\theta),\quad\theta\in(0,\pi),\;f\in L^{2}((0,\pi);d\theta), (3.65)

in L2((0,π);dθ)L^{2}((0,\pi);d\theta), one verifies that

UTnU1=Tn,UT_{n}U^{-1}=-T_{n}, (3.66)

and hence the spectrum of TnT_{n} is symmetric with respect to the origin. \diamond

It is well-known that Hardy’s inequality (LABEL:1.1) is strict, that is, equality holds in (LABEL:1.1) for some fD1(n)f\in D^{1}({\mathbb{R}}^{n}) if and only if f=0f=0. More general results regarding strictness for weighted Hardy–Sobolev or Caffarelli–Kohn–Nirenberg inequalities based on variational techniques can be found, for instance, in [13], [15]. Strictness in the case of the Hardy inequality was discussed in [87]. Thus, we next turn to strictness of inequality (3.5) on H1(n)H^{1}({\mathbb{R}}^{n}) employing a quadratic form approach.

To set the stage we briefly recall a few facts on quadratic forms generated by symmetric operators AA bounded from below and the associated Friedrichs extension (to be denoted by AFA_{F}) of AA.

Let AA be a densely defined symmetric operator in the Hilbert space {\mathcal{H}} bounded from below, that is, AAA\subseteq A^{*} and for some cc\in{\mathbb{R}}, AcIA\geqslant cI_{{\mathcal{H}}}. Without loss of generality we put c=0c=0 in the following. We denote by A¯\overline{A} the closure of AA in {\mathcal{H}}, and introduce the associated forms in {\mathcal{H}},

qA(f,g)=(f,Ag),f,gdom(qA)=dom(A),\displaystyle q_{A}(f,g)=(f,Ag)_{{\mathcal{H}}},\quad f,g\in\operatorname{dom}(q_{A})=\operatorname{dom}(A), (3.67)
qA¯(f,g)=(f,A¯g),f,gdom(qA¯)=dom(A¯),\displaystyle q_{\overline{A}}(f,g)=(f,{\overline{A}}g)_{{\mathcal{H}}},\quad f,g\in\operatorname{dom}(q_{\overline{A}})=\operatorname{dom}({\overline{A}}), (3.68)

then the closures of qAq_{A} and qA¯q_{\overline{A}} coincide in {\mathcal{H}} (cf., e.g., [8, Lemma 5.1.12])

qA¯=qA¯¯\overline{q_{A}}=\overline{q_{\overline{A}}} (3.69)

and the first representation theorem for forms (see, e.g., [23, Theorem 4.2.4], [56, Theorem VI.2.1, Sect. VI.2.3]) yields

qA¯(f,g)=(f,AFg),fdom(qA¯),gdom(AF),\overline{q_{A}}(f,g)=(f,A_{F}g)_{{\mathcal{H}}},\quad f\in\operatorname{dom}(\overline{q_{A}}),\;g\in\operatorname{dom}(A_{F}), (3.70)

where AF0A_{F}\geqslant 0 represents the self-adjoint Friedrichs extension of AA. Due to the fact (3.69), one infers (cf., e.g., [8, Lemma 5.3.1])

AF=(A¯)F.A_{F}=(\overline{A})_{F}. (3.71)

The second representation theorem for forms (see, e.g., [23, Theorem 4.2.8], [56, Theorem VI.2.123]) then yields the additional result

qA¯(f,g)=(AF1/2f,AF1/2g),f,gdom(qA¯)=dom(AF1/2).\overline{q_{A}}(f,g)=\big{(}A_{F}^{1/2}f,A_{F}^{1/2}g\big{)}_{{\mathcal{H}}},\quad f,g\in\operatorname{dom}(\overline{q_{A}})=\operatorname{dom}\big{(}A_{F}^{1/2}\big{)}. (3.72)

Moreover, one has the fact (see, e.g., [8, Theorem 5.3.3], [23, Corollary 4.2.7], [75, Theorem 10.17])

dom(AF)=dom(qA¯)dom(A)=dom(AF1/2)dom(A).\operatorname{dom}(A_{F})=\operatorname{dom}(\overline{q_{A}})\cap\operatorname{dom}(A^{*})=\operatorname{dom}\big{(}A_{F}^{1/2}\big{)}\cap\operatorname{dom}(A^{*}). (3.73)
Theorem 3.5.

Assume Hypothesis 2.2. Then inequality (3.5) is strict on H1(n)H^{1}({\mathbb{R}}^{n}), n3n\geqslant 3, that is, equality holds in (3.5) for some fH1(n)f\in H^{1}({\mathbb{R}}^{n}) if and only if f=0f=0.

Proof.

We first discuss the simpler case γ[0,γc,n)\gamma\in[0,\gamma_{c,n}). In this case the inequality (3.5) implies that the sesquilinear form γqu\gamma\,q_{u}, unu\in{\mathbb{R}}^{n}, |u|=1|u|=1, where

qu(f,g)=ndnx(u,x)|x|3f(x)¯g(x),f,gdom(qu)=H1(n),q_{u}(f,g)=\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}\overline{f(x)}g(x),\quad f,g\in\operatorname{dom}(q_{u})=H^{1}({\mathbb{R}}^{n}), (3.74)

is bounded relative to the form QH0Q_{H_{0}} of the Laplacian H0=ΔH_{0}=-\Delta on dom(H0)=H2(n)\operatorname{dom}(H_{0})=H^{2}({\mathbb{R}}^{n}),

QH0(f,g)=(f,g)[L2(n)]n=(H01/2f,H01/2g)L2(n),f,gdom(QH0)=H1(n),\displaystyle\begin{split}Q_{H_{0}}(f,g)=(\nabla f,\nabla g)_{[L^{2}({\mathbb{R}}^{n})]^{n}}=\big{(}H_{0}^{1/2}f,H_{0}^{1/2}g\big{)}_{L^{2}({\mathbb{R}}^{n})},&\\ f,g\in\operatorname{dom}(Q_{H_{0}})=H^{1}({\mathbb{R}}^{n}),&\end{split} (3.75)

with relative bound strictly less than one. Hence the form

Qγ(f,g)=QH0(f,g)+γqu(f,g),f,gdom(Qγ)=H1(n),γ[0,γc,n),Q_{\gamma}(f,g)=Q_{H_{0}}(f,g)+\gamma\,q_{u}(f,g),\quad f,g\in\operatorname{dom}(Q_{\gamma})=H^{1}({\mathbb{R}}^{n}),\;\gamma\in[0,\gamma_{c,n}), (3.76)

is densely defined, nonnegative, and closed. Moreover, since C0(n\{0})C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\}) is dense in H1(n)H^{1}({\mathbb{R}}^{n}) for n2n\geqslant 2 (cf., e.g., [26, p. 33–35]), that is, C0(n\{0})C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\}) is a core for QH0Q_{H_{0}} (equivalently, a form core for H0H_{0}), and hence also a core for QγQ_{\gamma},

Qγ|C0(n\{0})¯=Qγ,γ[0,γc,n).\overline{Q_{\gamma}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}}=Q_{\gamma},\quad\gamma\in[0,\gamma_{c,n}). (3.77)

Thus, the self-adjoint, nonnegative operator HQγH_{Q_{\gamma}}, uniquely associated with QγQ_{\gamma} by the first representation theorem for forms coincides with the Friedrichs extension of the minimal operator associated with the differential expression LγL_{\gamma} in (2.15),

Hγ,min=(Δ+Vγ)|C0(n\{0})0,γ[0,γc,n),H_{\gamma,min}=(-\Delta+V_{\gamma})|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}\geqslant 0,\quad\gamma\in[0,\gamma_{c,n}), (3.78)

that is,

HQγ=(Hγ,min)F0,γ[0,γc,n).H_{Q_{\gamma}}=(H_{\gamma,min})_{F}\geqslant 0,\quad\gamma\in[0,\gamma_{c,n}). (3.79)

In turn, since HγH_{\gamma} coincides with the direct sum of Friedrichs extensions in (2.31), one concludes that HQγH_{Q_{\gamma}} coincides with HγH_{\gamma}, and hence,

HQγ=Hγ=((Δ+Vγ)|C0(n\{0}))F,γ[0,γc,n),H_{Q_{\gamma}}=H_{\gamma}=\big{(}(-\Delta+V_{\gamma})|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}\big{)}_{F},\quad\gamma\in[0,\gamma_{c,n}), (3.80)

in particular,

Qγ(f,g)=(Hγ1/2f,Hγ1/2g)L2(n),f,gdom(Qγ)=H1(n),γ[0,γc,n).Q_{\gamma}(f,g)=\big{(}H_{\gamma}^{1/2}f,H_{\gamma}^{1/2}g\big{)}_{L^{2}({\mathbb{R}}^{n})},\quad f,g\in\operatorname{dom}(Q_{\gamma})=H^{1}({\mathbb{R}}^{n}),\;\gamma\in[0,\gamma_{c,n}). (3.81)

Thus, equality in the inequality (3.5) for some f0H1(n)f_{0}\in H^{1}({\mathbb{R}}^{n}) implies

0\displaystyle 0 =ndnx|(f0)(x)|2+γndnx(u,x)|x|3|f0(x)|2\displaystyle=\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f_{0})(x)|^{2}+\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f_{0}(x)|^{2} (3.82)
=QH0(f0,f0)+γqu(f0,f0)=Qγ(f0,f0)=Hγ1/2f0L2(n)2,γ[0,γc,n),\displaystyle=Q_{H_{0}}(f_{0},f_{0})+\gamma\,q_{u}(f_{0},f_{0})=Q_{\gamma}(f_{0},f_{0})=\big{\|}H_{\gamma}^{1/2}f_{0}\big{\|}_{L^{2}({\mathbb{R}}^{n})}^{2},\quad\gamma\in[0,\gamma_{c,n}),

and hence,

f0ker(Hγ1/2)=ker(Hγ)={0},γ[0,γc,n),f_{0}\in\ker\big{(}H_{\gamma}^{1/2}\big{)}=\ker(H_{\gamma})=\{0\},\quad\gamma\in[0,\gamma_{c,n}), (3.83)

by (3.7).

The case γ=γc,n\gamma=\gamma_{c,n} follows analogous lines but is a bit more involved as now the form γc,nqu\gamma_{c,n}\,q_{u} is bounded relative to the form QH0Q_{H_{0}} with relative bound equal to one.

Since by inequality (3.5)

Hγc,n,min=(Δ+Vγc,n)|C0(n\{0})0,H_{\gamma_{c,n},min}=(-\Delta+V_{\gamma_{c,n}})|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}\geqslant 0, (3.84)

the form

Q.γc,n(f,g)\displaystyle\overset{\textbf{\Large.}}{Q}_{\gamma_{c,n}}(f,g) =(f,Hγc,n,ming)L2(n)\displaystyle=(f,H_{\gamma_{c,n},min}g)_{L^{2}({\mathbb{R}}^{n})} (3.85)
=QH0(f,g)+γc,nqu(f,g),f,gdom(Q.γc,n)=C0(n\{0}),\displaystyle=Q_{H_{0}}(f,g)+\gamma_{c,n}q_{u}(f,g),\quad f,g\in\operatorname{dom}\big{(}\overset{\textbf{\Large.}}{Q}_{\gamma_{c,n}}\big{)}=C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\}),

is closable and we denote its closure in L2(n)L^{2}({\mathbb{R}}^{n}) by Qγc,nQ_{\gamma_{c,n}}. Thus, Qγc,n0Q_{\gamma_{c,n}}\geqslant 0 and hence the self-adjoint, nonnegative operator HQγc,nH_{Q_{\gamma_{c,n}}} uniquely associated with Qγc,nQ_{\gamma_{c,n}} in L2(n)L^{2}({\mathbb{R}}^{n}) is the Friedrichs extension of Hγc,n,minH_{\gamma_{c,n},min},

HQγc,n=(Hγc,n,min)F0.H_{Q_{\gamma_{c,n}}}=(H_{\gamma_{c,n},min})_{F}\geqslant 0. (3.86)

Again, by the second representation theorem

Qγc,n(f,g)=(HQγc,n1/2f,HQγc,n1/2g)L2(n),f,gdom(Qγc,n)=dom(HQγc,n1/2).Q_{\gamma_{c,n}}(f,g)=\big{(}H_{Q_{\gamma_{c,n}}}^{1/2}f,H_{Q_{\gamma_{c,n}}}^{1/2}g\big{)}_{L^{2}({\mathbb{R}}^{n})},\quad f,g\in\operatorname{dom}(Q_{\gamma_{c,n}})=\operatorname{dom}\big{(}H_{Q_{\gamma_{c,n}}}^{1/2}\big{)}. (3.87)

However, unlike in the case γ[0,γc,n)\gamma\in[0,\gamma_{c,n}), since the form γc,nqu\gamma_{c,n}\,q_{u} is bounded relative to the form QH0Q_{H_{0}} with relative bound equal to one, one now has possible cancellations between the forms QH0Q_{H_{0}} and γc,nqu\gamma_{c,n}\,q_{u} and hence concludes that

dom(Qγc,n)=dom(HQγc,n1/2)H1(n)\operatorname{dom}(Q_{\gamma_{c,n}})=\operatorname{dom}\big{(}H_{Q_{\gamma_{c,n}}}^{1/2}\big{)}\supseteq H^{1}({\mathbb{R}}^{n}) (3.88)

(see also Remark 3.6). The rest of the proof now follows the case γ[0,γc,n)\gamma\in[0,\gamma_{c,n}) line by line, in particular,

HQγc,n=Hγc,n=((Δ+Vγc,n)|C0(n\{0}))F,H_{Q_{\gamma_{c,n}}}=H_{\gamma_{c,n}}=\big{(}(-\Delta+V_{\gamma_{c,n}})|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}\big{)}_{F}, (3.89)

and

ker(Hγc,n1/2)=ker(Hγc,n)={0},\ker\big{(}H_{\gamma_{c,n}}^{1/2}\big{)}=\ker(H_{\gamma_{c,n}})=\{0\}, (3.90)

again by (3.7), then proves strictness of (3.5) also for γ=γc,n\gamma=\gamma_{c,n}. ∎

Remark 3.6.

We briefly illustrate the possibility of cancellations in Hγc,nH_{\gamma_{c,n}}. Let ψγ,0=ψγ,0(θn1)\psi_{\gamma,0}=\psi_{\gamma,0}(\theta_{n-1}), γ0\gamma\geqslant 0, be the unique (up to constant multiples) eigenfunction of Λγ,n\Lambda_{\gamma,n} corresponding to its lowest eigenvalue λγ,n,0\lambda_{\gamma,n,0}, see (3.24), and introduce

Ψγ,0(x)=|x|(n2)/2ψγ,0(θn1),xn\{0},γ0.\Psi_{\gamma,0}(x)=|x|^{-(n-2)/2}\psi_{\gamma,0}(\theta_{n-1}),\quad x\in{\mathbb{R}}^{n}\backslash\{0\},\;\gamma\geqslant 0. (3.91)

Then an elementary computation reveals that

LγΨγ,0={[(n2)2/4]+λγ,n,0}|x|(n+2)/2ψγ,0(θn1),L_{\gamma}\Psi_{\gamma,0}=\big{\{}\big{[}(n-2)^{2}/4\big{]}+\lambda_{\gamma,n,0}\big{\}}|x|^{-(n+2)/2}\psi_{\gamma,0}(\theta_{n-1}), (3.92)

in the sense of distributions. In particular, if γ=γc,n\gamma=\gamma_{c,n}, and hence, λγc,n,n,0=(n2)2/4\lambda_{\gamma_{c,n},n,0}=-(n-2)^{2}/4, one obtains

Lγc,nΨγc,n,0=0,L_{\gamma_{c,n}}\Psi_{\gamma_{c,n},0}=0, (3.93)

in the distributional sense. Thus, introducing

f0(x)=|x|(n2)/2ψγc,n,0(θn1)ϕ(|x|),xn\{0},f_{0}(x)=|x|^{-(n-2)/2}\psi_{\gamma_{c,n},0}(\theta_{n-1})\phi(|x|),\quad x\in{\mathbb{R}}^{n}\backslash\{0\}, (3.94)

one concludes that

f0dom(Hγc,n)dom(Hγc,n1/2).f_{0}\in\operatorname{dom}(H_{\gamma_{c,n}})\subset\operatorname{dom}\big{(}H_{\gamma_{c,n}}^{1/2}\big{)}. (3.95)

However, since (f0/r)L2(n)(\partial f_{0}/\partial r)\notin L^{2}({\mathbb{R}}^{n}), the elementary fact

|(f0r)(x)|=|x|x|(f0)(x)||(f0)(x)|\bigg{|}\bigg{(}\frac{\partial f_{0}}{\partial r}\bigg{)}(x)\bigg{|}=\bigg{|}\frac{x}{|x|}\cdot(\nabla f_{0})(x)\bigg{|}\leqslant|(\nabla f_{0})(x)| (3.96)

implies that

f0H1(n),(u,x)|x|3|f0|2L1(n),f_{0}\notin H^{1}({\mathbb{R}}^{n}),\;(u,x)|x|^{-3}|f_{0}|^{2}\notin L^{1}({\mathbb{R}}^{n}), (3.97)

illustrating possible cancellations between H0H_{0} and γc,n(u,x)|x|3\gamma_{c,n}(u,x)|x|^{-3}. \diamond

Remark 3.7.

Next, we briefly discuss the remaining case n=2n=2. In this situation, the Laplace–Beltrami operator Δ𝕊1-\Delta_{{\mathbb{S}}^{1}} in L2(𝕊1)L^{2}\big{(}{\mathbb{S}}^{1}\big{)} can be characterized by

(Δ𝕊1f)(θ1)=f′′(θ1),θ1(0,2π),\displaystyle(-\Delta_{{\mathbb{S}}^{1}}f)(\theta_{1})=-f^{\prime\prime}(\theta_{1}),\quad\theta_{1}\in(0,2\pi),
fdom(Δ𝕊1)={gL2((0,2π);dθ1)|g,gAC([0,2π]);\displaystyle f\in\operatorname{dom}(-\Delta_{{\mathbb{S}}^{1}})=\big{\{}g\in L^{2}((0,2\pi);d\theta_{1})\,\big{|}\,g,g^{\prime}\in AC([0,2\pi]); (3.98)
g(0)=g(2π),g(0)=g(2π);g′′L2((0,2π);dθ1)},\displaystyle\hskip 88.2037ptg(0)=g(2\pi),\,g^{\prime}(0)=g^{\prime}(2\pi);\,g^{\prime\prime}\in L^{2}((0,2\pi);d\theta_{1})\big{\}},

with

σ(Δ𝕊1)={2}0,\displaystyle\;\sigma(-\Delta_{{\mathbb{S}}^{1}})=\big{\{}\ell^{2}\big{\}}_{\ell\in{\mathbb{N}}_{0}}, (3.99)
Δ𝕊1e±iθ1=2e±iθ1,θ1(0,2π),0.\displaystyle-\Delta_{{\mathbb{S}}^{1}}e^{\pm i\ell\theta_{1}}=\ell^{2}e^{\pm i\ell\theta_{1}},\quad\theta_{1}\in(0,2\pi),\;\ell\in{\mathbb{N}}_{0}. (3.100)

The resulting Mathieu operator Λγ,2\Lambda_{\gamma,2} in L2((0,2π);dθ1)L^{2}((0,2\pi);d\theta_{1}) (cf. (2.17)), of the form

Λγ,2=d2dθ12+γcos(θ1),dom(Λγ,2)=dom(Δ𝕊1),\Lambda_{\gamma,2}=-\frac{d^{2}}{d\theta_{1}^{2}}+\gamma\cos(\theta_{1}),\quad\operatorname{dom}(\Lambda_{\gamma,2})=\operatorname{dom}(-\Delta_{{\mathbb{S}}^{1}}), (3.101)

has extensively been studied in the literature, see, for instance [63, Ch. 2]. More generally, the least periodic eigenvalue of Hill operators (i.e., situations where cos(θ1)\cos(\theta_{1}) is replaced by a 2π2\pi-periodic, locally integrable potential q(θ1)q(\theta_{1})) has received enormous attention, see for instance, [10], [35], [55], [64], [69], [70], [78], [85], and [90]. Applied to the Mathieu operator Λγ,2\Lambda_{\gamma,2} at hand, the results obtained (cf. the discussion in [35]) imply,

γ2/[8π2]<λγ,2,0<0,γ(0,),there exists c0(0,) such that λγ,2,0c0γ2,γ[0,1].\displaystyle\begin{split}&-\gamma^{2}\big{/}\big{[}8\pi^{2}\big{]}<\lambda_{\gamma,2,0}<0,\quad\gamma\in(0,\infty),\\ &\quad\text{there exists $c_{0}\in(0,\infty)$ such that }\,\lambda_{\gamma,2,0}\leqslant-c_{0}\gamma^{2},\quad\gamma\in[0,1].\end{split} (3.102)

In particular, this proves the absence of a critical coupling constant 0<γc,20<\gamma_{c,2} for n=2n=2 (equivalently, the critical constant in two dimensions equals zero, γc,2=0\gamma_{c,2}=0), explaining why we had to limit ourselves to n3n\geqslant 3 in the bulk of this paper. \diamond

Remark 3.8.

While thus far we focused primarily on lower semiboundedness of HγH_{\gamma}, the direct sum considerations in Section 2 equally apply to essential self-adjointness of Hγ|C0(n\{0})H_{\gamma}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}. Indeed, returning to the operator in (2.24), one notes that

[d2dr2+cr2]|C0((0,)) is essentially self-adjoint if and only if c3/4.\bigg{[}-\frac{d^{2}}{dr^{2}}+\frac{c}{r^{2}}\bigg{]}\bigg{|}_{C_{0}^{\infty}((0,\infty))}\,\text{ is essentially self-adjoint if and only if $c\geqslant 3/4$.} (3.103)

The criterion (3.103) combined with (2.31), (2.32) thus implies that

Hγ|C0(n\{0}) is essentially self-adjointif and only if λγ,n,0n(n4)/4.\displaystyle\begin{split}&H_{\gamma}|_{C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\})}\,\text{ is essentially self-adjoint}\\ &\quad\text{if and only if }\,\lambda_{\gamma,n,0}\geqslant-n(n-4)/4.\end{split} (3.104)

\diamond

4. A Numerical Approach

Having verified the existence and uniqueness of critical dipole moments γc,n\gamma_{c,n} for all dimensions n3n\geqslant 3, and having shown some of the properties of λγ,n,0\lambda_{\gamma,n,0}, this section is devoted to a description of a numerical method for computing γc,n\gamma_{c,n}, in analogy to the Legendre expansion in [16].

To set up the numerical algorithm one can argue as follows: Given (3.24), we are interested in solving this eigenvalue problem in the particular scenario where γ\gamma ranges from 0 to γc,n\gamma_{c,n}, observing that λγc,n,n,0=(n2)2/4\lambda_{\gamma_{c,n},n,0}=-(n-2)^{2}/4 (cf. (2.28)). Restricting Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}} in (A.16) to n{\mathcal{L}}^{n} as in (LABEL:3.3)–(3.10), (3.24) reduces to solving the eigenvalue problem associated with Λγ,n\Lambda_{\gamma,{\mathcal{L}}^{n}} in L2((0,π);[sin(θn1)]n2dθn1)L^{2}\big{(}(0,\pi);[\sin(\theta_{n-1})]^{n-2}d\theta_{n-1}\big{)} of the type,

d2Ψγ(θn1)dθn12(n2)cot(θn1)dΨγ(θn1)dθn1+γcos(θn1)Ψγ(θn1)=λγ,n,0Ψγ(θn1),γ(0,γc,n],θn1(0,π).\displaystyle\begin{split}&-\frac{d^{2}\Psi_{\gamma}(\theta_{n-1})}{d\theta_{n-1}^{2}}-(n-2)\cot(\theta_{n-1})\frac{d\Psi_{\gamma}(\theta_{n-1})}{d\theta_{n-1}}+\gamma\cos(\theta_{n-1})\Psi_{\gamma}(\theta_{n-1})\\ &\quad=\lambda_{\gamma,n,0}\Psi_{\gamma}(\theta_{n-1}),\quad\gamma\in(0,\gamma_{c,n}],\;\theta_{n-1}\in(0,\pi).\end{split} (4.1)

Expanding Ψγ()\Psi_{\gamma}(\,\cdot\,) in normalized Gegenbauer polynomials, one obtains

Ψγ(θn1)==0d(γ)[!(2+n2)24nπΓ(+n2)]1/2Γ((n2)/2)C(n2)/2(cos(θn1)),\displaystyle\Psi_{\gamma}(\theta_{n-1})=\sum\limits_{\ell=0}^{\infty}d_{\ell}(\gamma)\bigg{[}\frac{\ell!(2\ell+n-2)}{2^{4-n}\pi\Gamma(\ell+n-2)}\bigg{]}^{1/2}\Gamma((n-2)/2)C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1})),
θn1(0,π),\displaystyle\hskip 256.0748pt\theta_{n-1}\in(0,\pi), (4.2)

where d(γ)d_{\ell}(\gamma) are appropriate expansion coefficients. Since the Gegenbauer polynomial C(n2)/2(cos())C_{\ell}^{(n-2)/2}(\cos(\,\cdot\,)) is an eigenfunction of Δ𝕊n1-\Delta_{{\mathbb{S}}^{n-1}} corresponding to the eigenvalue (+n2)\ell(\ell+n-2), (4.1) becomes

=0[(+n2)+γcos(θn1)λγ,n,0]d(γ)[!(2+n2)24nπΓ(+n2)]1/2×Γ((n2)/2)C(n2)/2(cos(θn1))=0.\displaystyle\begin{split}&\sum\limits_{\ell=0}^{\infty}[\ell(\ell+n-2)+\gamma\cos(\theta_{n-1})-\lambda_{\gamma,n,0}]\,d_{\ell}(\gamma)\bigg{[}\frac{\ell!(2\ell+n-2)}{2^{4-n}\pi\Gamma(\ell+n-2)}\bigg{]}^{1/2}\\ &\hskip 15.649pt\times\Gamma((n-2)/2)C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1}))=0.\end{split} (4.3)

Next, we will exploit the following recurrence relation of Gegenbauer polynomials,

cos(θn1)C(n2)/2(cos(θn1))\displaystyle\cos(\theta_{n-1})C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1}))
=+12+n2(C+1(n2)/2(cos(θn1))++n3+1C1(n2)/2(cos(θn1))),\displaystyle\quad=\frac{\ell+1}{2\ell+n-2}\left(C_{\ell+1}^{(n-2)/2}(\cos(\theta_{n-1}))+\frac{\ell+n-3}{\ell+1}C_{\ell-1}^{(n-2)/2}(\cos(\theta_{n-1}))\right),
0\displaystyle\hskip 273.14662pt\ell\in{\mathbb{N}}_{0} (4.4)

(with C1(n2)/2()0C_{-1}^{(n-2)/2}(\,\cdot\,)\equiv 0) to expand the term γcos(θn1)\gamma\cos(\theta_{n-1}). For the (1)(\ell-1)-term, one infers

γcos(θn1)d1(γ)[(1)!(2+n4))224nπΓ(+n3)]1/2Γ((n2)/2)C1(n2)/2(cos(θn1))\displaystyle\gamma\cos(\theta_{n-1})d_{\ell-1}(\gamma)\bigg{[}\frac{(\ell-1)!(2\ell+n-4))^{2}}{2^{4-n}\pi\Gamma(\ell+n-3)}\bigg{]}^{1/2}\Gamma((n-2)/2)C_{\ell-1}^{(n-2)/2}(\cos(\theta_{n-1}))
=γd1(γ)[(1)!(2+n4)24nπΓ(+n3)]1/2Γ((n2)/2)2+n4C(n2)/2(cos(θn1)),\displaystyle\quad=\gamma d_{\ell-1}(\gamma)\bigg{[}\frac{(\ell-1)!(2\ell+n-4)}{2^{4-n}\pi\Gamma(\ell+n-3)}\bigg{]}^{1/2}\frac{\ell\,\Gamma((n-2)/2)}{2\ell+n-4}C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1})),
0\displaystyle\hskip 281.6821pt\ell\in{\mathbb{N}}_{0} (4.5)

(where d1(γ)=0d_{-1}(\gamma)=0) and for the (+1)(\ell+1)-term, one obtains

γcos(θn1)d+1(γ)[(+1)!(2+n)24nπΓ(+n1)]1/2Γ((n2)/2)C+1(n2)/2(cos(θn1))\displaystyle\gamma\cos(\theta_{n-1})d_{\ell+1}(\gamma)\bigg{[}\frac{(\ell+1)!(2\ell+n)}{2^{4-n}\pi\Gamma(\ell+n-1)}\bigg{]}^{1/2}\Gamma((n-2)/2)C_{\ell+1}^{(n-2)/2}(\cos(\theta_{n-1}))
=γd+1(γ)[(+1)!(2+n)24nπΓ(+n1)]1/2Γ((n2)/2)+n22+nC(n2)/2(cos(θn1)),\displaystyle\quad=\gamma d_{\ell+1}(\gamma)\bigg{[}\frac{(\ell+1)!(2\ell+n)}{2^{4-n}\pi\Gamma(\ell+n-1)}\bigg{]}^{1/2}\Gamma((n-2)/2)\frac{\ell+n-2}{2\ell+n}C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1})),
0.\displaystyle\hskip 284.52756pt\ell\in{\mathbb{N}}_{0}. (4.6)

The \ell-term maintains its form

[(+n2)λγ,n,0]d(γ)[!(2+n2)24nπΓ(+n2)]1/2Γ((n2)/2)\displaystyle[\ell(\ell+n-2)-\lambda_{\gamma,n,0}]\,d_{\ell}(\gamma)\bigg{[}\frac{\ell!(2\ell+n-2)}{2^{4-n}\pi\Gamma(\ell+n-2)}\bigg{]}^{1/2}\Gamma((n-2)/2)
×C(n2)/2(cos(θn1)),0,\displaystyle\quad\times C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1})),\quad\ell\in{\mathbb{N}}_{0}, (4.7)

so one can divide all terms by the normalizing factor from (4.2) (since the orthogonality of the Gegenbauer polynomials mandates that every term under the sum in (4.3) individually vanishes), obtaining

=0{[(+n2)λγ,n,0]d(γ)\displaystyle\sum\limits_{\ell=0}^{\infty}\bigg{\{}[\ell(\ell+n-2)-\lambda_{\gamma,n,0}]d_{\ell}(\gamma)
+γ([(2+n4)(+n3)(2+n2)]1/22+n4d1(γ)\displaystyle\hskip 19.91692pt+\gamma\bigg{(}\bigg{[}\frac{(2\ell+n-4)(\ell+n-3)}{\ell(2\ell+n-2)}\bigg{]}^{1/2}\,\frac{\ell}{2\ell+n-4}d_{\ell-1}(\gamma) (4.8)
+[(+1)(2+n)(2+n2)(+n2)]1/2+n22+nd+1(γ))}C(n2)/2(cos(θn1))=0.\displaystyle\hskip 19.91692pt+\bigg{[}\frac{(\ell+1)(2\ell+n)}{(2\ell+n-2)(\ell+n-2)}\bigg{]}^{1/2}\frac{\ell+n-2}{2\ell+n}\,d_{\ell+1}(\gamma)\bigg{)}\bigg{\}}C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1}))=0.

Setting each coefficient equal to zero results in

[(+n2)λγ,n,0]d(γ)+γ([(+n3)(2+n4)(2+n2)]1/2d1(γ)+[(+1)(+n2)(2+n2)(2+n)]1/2d+1(γ))=0,0,\displaystyle\begin{split}&[\ell(\ell+n-2)-\lambda_{\gamma,n,0}]d_{\ell}(\gamma)+\gamma\bigg{(}\bigg{[}\frac{\ell(\ell+n-3)}{(2\ell+n-4)(2\ell+n-2)}\bigg{]}^{1/2}d_{\ell-1}(\gamma)\\ &\quad+\bigg{[}\frac{(\ell+1)(\ell+n-2)}{(2\ell+n-2)(2\ell+n)}\bigg{]}^{1/2}d_{\ell+1}(\gamma)\bigg{)}=0,\quad\ell\in{\mathbb{N}}_{0},\end{split} (4.9)

which one can rewrite as

[(+1)(+n2)(2+n2)(2+n)]1/2d+1(γ)+[(+n3)(2+n4)(2+n2)]1/2d1(γ)\displaystyle\bigg{[}\frac{(\ell+1)(\ell+n-2)}{(2\ell+n-2)(2\ell+n)}\bigg{]}^{1/2}d_{\ell+1}(\gamma)+\bigg{[}\frac{\ell(\ell+n-3)}{(2\ell+n-4)(2\ell+n-2)}\bigg{]}^{1/2}d_{\ell-1}(\gamma)
=1γ[(+n2)λγ,n,0]d(γ),γ(0,γc,n],0.\displaystyle\quad=-\frac{1}{\gamma}[\ell(\ell+n-2)-\lambda_{\gamma,n,0}]\,d_{\ell}(\gamma),\quad\gamma\in(0,\gamma_{c,n}],\;\ell\in{\mathbb{N}}_{0}. (4.10)

Equation (4.10) can be expressed as the generalized Jacobi operator eigenvalue problem in 2(0;w)\ell^{2}({\mathbb{N}}_{0};w),

Jd(γ)=1γw(γ)d(γ),γ(0,γc,n],\displaystyle Jd(\gamma)=-\frac{1}{\gamma}w(\gamma)d(\gamma),\quad\gamma\in(0,\gamma_{c,n}], (4.11)

where

Jd(γ)={a+1d+1(γ)+ad1(γ),,a1d1(γ),=0,w(γ)d(γ)=(w(γ)d(γ))0,\displaystyle Jd(\gamma)=\begin{cases}a_{\ell+1}d_{\ell+1}(\gamma)+a_{\ell}d_{\ell-1}(\gamma),&\ell\in{\mathbb{N}},\\ a_{1}d_{1}(\gamma),&\ell=0,\end{cases}\quad w(\gamma)d(\gamma)=\big{(}w_{\ell}(\gamma)d_{\ell}(\gamma)\big{)}_{\ell\in{\mathbb{N}}_{0}}, (4.12)

and

a=[(+n3)(2+n4)(2+n2)]1/2,w(γ)=(+n2)λγ,n,0,0.\displaystyle\begin{split}&a_{\ell}=\bigg{[}\frac{\ell(\ell+n-3)}{(2\ell+n-4)(2\ell+n-2)}\bigg{]}^{1/2},\quad w_{\ell}(\gamma)=\ell(\ell+n-2)-\lambda_{\gamma,n,0},\quad\ell\in{\mathbb{N}}_{0}.\end{split} (4.13)

(One observes that w0(γ)γ00w_{0}(\gamma)\underset{\gamma\downarrow 0}{\longrightarrow}0 by (2.38).)

Explicitly, (4.12) yields the self-adjoint Jacobi operator JJ in 2(0;w)\ell^{2}({\mathbb{N}}_{0};w) represented as a semi-infinite matrix with respect to the standard Kronecker-δ\delta basis

Jd(γ)\displaystyle Jd(\gamma) =(0a10a10a200a20a300a30a400)(d0(γ)d1(γ)d2(γ))\displaystyle=\begin{pmatrix}0&a_{1}&0&\ldots&&&\\ a_{1}&0&a_{2}&0&\ldots&&\\ 0&a_{2}&0&a_{3}&0&\ldots&\\ \vdots&0&a_{3}&0&a_{4}&0&\ldots\\ &\vdots&0&\ddots&\ddots&\ddots&\\ \end{pmatrix}\begin{pmatrix}d_{0}(\gamma)\\ d_{1}(\gamma)\\ d_{2}(\gamma)\\ \vdots\\ \phantom{\vdots}\\ \end{pmatrix}
=1γ(w0(γ)00w1(γ)00w2(γ)00w3(γ)0)(d0(γ)d1(γ)d2(γ))\displaystyle=-\frac{1}{\gamma}\begin{pmatrix}w_{0}(\gamma)&0&\ldots&&&\\ 0&w_{1}(\gamma)&0&\ldots&&\\ \vdots&0&w_{2}(\gamma)&0&\ldots&\\ &\vdots&0&w_{3}(\gamma)&0&\ldots\\ &&\vdots&\ddots&\ddots&\ddots\\ \end{pmatrix}\begin{pmatrix}d_{0}(\gamma)\\ d_{1}(\gamma)\\ d_{2}(\gamma)\\ \vdots\\ \phantom{\vdots}\vspace{1mm}\\ \end{pmatrix}
=1γw(γ)d(γ),γ(0,γc,n].\displaystyle=-\frac{1}{\gamma}w(\gamma)d(\gamma),\quad\gamma\in(0,\gamma_{c,n}]. (4.14)

One would like to calculate γc,n\gamma_{c,n} approximately using finite truncations of the matrix representation of JJ in the first line of (4.14) - the feasibility of truncations will be made precise below. In order for these approximants to converge, a transformation to a compact Jacobi operator becomes necessary. For this purpose one introduces the operator

V(γ)={2(0;w(γ))2(0),bw1/2(γ)b.\displaystyle V(\gamma)=\begin{cases}\ell^{2}({\mathbb{N}}_{0};w(\gamma))\to\ell^{2}({\mathbb{N}}_{0}),\\ b\mapsto w^{1/2}(\gamma)b.\end{cases} (4.15)

That V(γ)V(\gamma) is unitary may be seen from

V(γ)b2(0)=w1/2(γ)b2(0)=b2(0;w(γ)),b2(0;w(γ)),\displaystyle||V(\gamma)b||_{\ell^{2}({\mathbb{N}}_{0})}=||w^{1/2}(\gamma)b||_{\ell^{2}({\mathbb{N}}_{0})}=||b||_{\ell^{2}({\mathbb{N}}_{0};w(\gamma))},\quad b\in\ell^{2}({\mathbb{N}}_{0};w(\gamma)), (4.16)

and the fact that V(γ)V(\gamma) is surjective and defined on all of 2(0;w)\ell^{2}({\mathbb{N}}_{0};w). Next, one transforms (4.14) into

V(γ)1JV(γ)1V(γ)d=1γV(γ)d(γ),d(γ)2(0;w(γ)),V(\gamma)^{-1}JV(\gamma)^{-1}V(\gamma)d=-\frac{1}{\gamma}V(\gamma)d(\gamma),\quad d(\gamma)\in\ell^{2}({\mathbb{N}}_{0};w(\gamma)), (4.17)

which can equivalently be expressed as

w(γ)1/2Jw(γ)1/2d~(γ)=1γd~(γ),d~(γ)=V(γ)d(γ)2(0).w(\gamma)^{-1/2}Jw(\gamma)^{-1/2}\widetilde{d}(\gamma)=-\frac{1}{\gamma}\widetilde{d}(\gamma),\quad\widetilde{d}(\gamma)=V(\gamma)d(\gamma)\in\ell^{2}({\mathbb{N}}_{0}). (4.18)

Introducing

K(γ)=w(γ)1/2Jw(γ)1/2(2(0)),γ(0,γc,n],K(\gamma)=w(\gamma)^{-1/2}Jw(\gamma)^{-1/2}\in{\mathcal{B}}\big{(}\ell^{2}({\mathbb{N}}_{0})\big{)},\quad\gamma\in(0,\gamma_{c,n}], (4.19)

one can write (4.18) in the form

K(γ)d~(γ)=(0a~1(γ)0a~1(γ)0a~2(γ)00a~2(γ)0a~3(γ)00a~3(γ)0a~4(γ)00)(d~0(γ)d~1(γ)d~2(γ))=1γ(d~0(γ)d~1(γ)d~2(γ))=1γd~(γ),γ(0,γc,n],\displaystyle\begin{split}K(\gamma)\widetilde{d}(\gamma)&=\begin{pmatrix}0&\widetilde{a}_{1}(\gamma)&0&\ldots&&&\\ \widetilde{a}_{1}(\gamma)&0&\widetilde{a}_{2}(\gamma)&0&\ldots&&\\ 0&\widetilde{a}_{2}(\gamma)&0&\widetilde{a}_{3}(\gamma)&0&\ldots&\\ \vdots&0&\widetilde{a}_{3}(\gamma)&0&\widetilde{a}_{4}(\gamma)&0&\ldots\\ &\vdots&0&\ddots&\ddots&\ddots&\\ \end{pmatrix}\begin{pmatrix}\widetilde{d}_{0}(\gamma)\\ \widetilde{d}_{1}(\gamma)\\ \widetilde{d}_{2}(\gamma)\\ \vdots\\ \phantom{\vdots}\\ \end{pmatrix}\\ &=-\frac{1}{\gamma}\begin{pmatrix}\widetilde{d}_{0}(\gamma)\\ \widetilde{d}_{1}(\gamma)\\ \widetilde{d}_{2}(\gamma)\\ \vdots\\ \phantom{\vdots}\vspace{1mm}\\ \end{pmatrix}=-\frac{1}{\gamma}\widetilde{d}(\gamma),\quad\gamma\in(0,\gamma_{c,n}],\end{split} (4.20)

in analogy with (4.14), with

a~(γ)=w(γ)1/2aw1(γ)1/2\displaystyle\,\,\widetilde{a}_{\ell}(\gamma)=w_{\ell}(\gamma)^{-1/2}a_{\ell}\,w_{\ell-1}(\gamma)^{-1/2}
=[(+n2)λγ,n,0]1/2[(1)(+n3)λγ,n,0]1/2\displaystyle\,\,\,\qquad=[\ell(\ell+n-2)-\lambda_{\gamma,n,0}]^{-1/2}[(\ell-1)(\ell+n-3)-\lambda_{\gamma,n,0}]^{-1/2}
×[(+n3)(2+n4)(2+n2)]1/2,,\displaystyle\,\,\,\quad\qquad\times\bigg{[}\frac{\ell(\ell+n-3)}{(2\ell+n-4)(2\ell+n-2)}\bigg{]}^{1/2},\quad\ell\in{\mathbb{N}}, (4.21)
a~(γ)=O(2).\displaystyle\widetilde{a}_{\ell}(\gamma)\underset{\ell\to\infty}{=}O\big{(}\ell^{-2}\big{)}. (4.22)

Recalling once more that λγc,n,n,0=(n2)2/4\lambda_{\gamma_{c,n},n,0}=-(n-2)^{2}/4 (cf. (2.28)), and that λγ,n,0\lambda_{\gamma,n,0} as well as Ψγ()\Psi_{\gamma}(\,\cdot\,) in (4.1) are analytic with respect to γ\gamma as γ\gamma varies in a complex neighborhood of [0,)[0,\infty), taking the limit γγc,n\gamma\uparrow\gamma_{c,n} on either side of (4.20) yields

K(γc,n)d~(γc,n)=γc,n1d~(γc,n),K(\gamma_{c,n})\widetilde{d}(\gamma_{c,n})=-\gamma_{c,n}^{-1}\,\widetilde{d}(\gamma_{c,n}), (4.23)

implying

γc,n1σp(K(γc,n)).-\gamma_{c,n}^{-1}\in\sigma_{p}(K(\gamma_{c,n})). (4.24)

Next, we prove that γc,n1-\gamma_{c,n}^{-1} is the smallest (negative) eigenvalue of the operator K(γc,n)K(\gamma_{c,n}) in 2(0)\ell^{2}({\mathbb{N}}_{0}).

Proposition 4.1.

One has K(γ)(2(0))K(\gamma)\in{\mathcal{B}}_{\infty}\big{(}\ell^{2}({\mathbb{N}}_{0})\big{)}, γ(0,γc,n]\gamma\in(0,\gamma_{c,n}], and

γK(γ)I2(0),γ(0,γc,n].\gamma K(\gamma)\geqslant-I_{\ell^{2}({\mathbb{N}}_{0})},\quad\gamma\in(0,\gamma_{c,n}]. (4.25)

Moreover,

K(γc,n)(2(0)){8/[33/2],n=3,2max{a~r,n(γc,n),a~r,n1(γc,n)},n4,\displaystyle||K(\gamma_{c,n})||_{{\mathcal{B}}(\ell^{2}({\mathbb{N}}_{0}))}\leqslant\begin{cases}8\big{/}\big{[}3^{3/2}\big{]},\quad&n=3,\\ 2\max\{\widetilde{a}_{\ell_{r,n}}(\gamma_{c,n}),\widetilde{a}_{\ell_{r,n}-1}(\gamma_{c,n})\},\quad&n\geqslant 4,\end{cases} (4.26)

where

r(n)=14(62n+{2[2618n+3n2]}1/2),r,n=r(n),r(n)=\frac{1}{4}\left(6-2n+\big{\{}2\big{[}26-18n+3n^{2}\big{]}\big{\}}^{1/2}\right),\quad\ell_{r,n}=\lceil r(n)\rceil, (4.27)

with x=inf{m0|mx}\lceil x\rceil=\inf\{m\in{\mathbb{N}}_{0}\,|\,m\geqslant x\}, the ceiling function.

Proof.

The compactness assertion for K(γ)K(\gamma), γ(0,γc,n]\gamma\in(0,\gamma_{c,n}], follows from the limiting behavior a~(γ)0\widetilde{a}_{\ell}(\gamma)\underset{\ell\to\infty}{\longrightarrow}0, see, for instance, [86, p. 201].

To prove the uniform lower bound (4.25) one can argue by contradiction as follows: Fix γ(0,γc,n]\gamma\in(0,\gamma_{c,n}] and suppose there exists ε>0\varepsilon>0 such that (1+ε)σ(K(γ))-(1+\varepsilon)\in\sigma(K(\gamma)), that is, (1+ε)σp(K(γ))-(1+\varepsilon)\in\sigma_{p}(K(\gamma)). Then working backwards from (4.20) to (4.1) yields the existence of {d(ε,γ)}02(0;w)\{d_{\ell}(\varepsilon,\gamma)\}_{\ell\in{\mathbb{N}}_{0}}\in\ell^{2}({\mathbb{N}}_{0};w) such that

d2Ψε,γ(θn1)dθn12(n2)cot(θn1)dΨε,γ(θn1)dθn1+γ1+εcos(θn1)Ψε,γ(θn1)=λγ,n,0Ψε,γ(θn1),γ(0,γc,n].\displaystyle\begin{split}&-\frac{d^{2}\Psi_{\varepsilon,\gamma}(\theta_{n-1})}{d\theta_{n-1}^{2}}-(n-2)\cot(\theta_{n-1})\frac{d\Psi_{\varepsilon,\gamma}(\theta_{n-1})}{d\theta_{n-1}}+\frac{\gamma}{1+\varepsilon}\cos(\theta_{n-1})\Psi_{\varepsilon,\gamma}(\theta_{n-1})\\ &\quad=\lambda_{\gamma,n,0}\Psi_{\varepsilon,\gamma}(\theta_{n-1}),\quad\gamma\in(0,\gamma_{c,n}].\end{split} (4.28)

where

Ψε,γ(θn1)\displaystyle\Psi_{\varepsilon,\gamma}(\theta_{n-1}) ==0d(ε,γ)[!(2+n2)24nπΓ(+n2)]1/2Γ((n2)/2)C(n2)/2(cos(θn1)).\displaystyle=\sum\limits_{\ell=0}^{\infty}d_{\ell}(\varepsilon,\gamma)\bigg{[}\frac{\ell!(2\ell+n-2)}{2^{4-n}\pi\Gamma(\ell+n-2)}\bigg{]}^{1/2}\Gamma((n-2)/2)C_{\ell}^{(n-2)/2}(\cos(\theta_{n-1})). (4.29)

By the strict monotonicity of λγ,n,0\lambda_{\gamma,n,0} with respect to γ(0,γc,n]\gamma\in(0,\gamma_{c,n}] one infers that

λγ,n,0<λγ/(1+ε),n,0,\lambda_{\gamma,n,0}<\lambda_{\gamma/(1+\varepsilon),n,0}, (4.30)

and hence (4.28) contradicts the fact that by definition, λγ/(1+ε),n,0\lambda_{\gamma/(1+\varepsilon),n,0} is the lowest eigenvalue of Λγ/(1+ε),n\Lambda_{\gamma/(1+\varepsilon),{\mathcal{L}}^{n}}.

To obtain the bound (4.26), (4.27), one applies [81, Theorem 1.5] after calculating a~(γc,n)(0)||\widetilde{a}(\gamma_{c,n})||_{\ell^{\infty}({\mathbb{N}}_{0})}. As a~,n(γc,n)\widetilde{a}_{\ell,n}(\gamma_{c,n}) is bounded and tends to 0 as \ell\to\infty for all n3n\geqslant 3, it attains its supremum. To find the index where this occurs, one considers \ell as a continuous variable, and solves da~,n(γc,n)d=0\frac{d\widetilde{a}_{\ell,n}(\gamma_{c,n})}{d\ell}=0. The value r(n)r(n) emerges as the only nonnegative, real root of this expression, but as 0<r(n)\0<r(n)\in{\mathbb{R}}\backslash{\mathbb{N}} for n4n\geqslant 4, the maximum in (4.26) and the ceiling function in (4.27) are required. Since r(3)\r(3)\in{\mathbb{C}}\backslash{\mathbb{R}}, the norm a~,3(γc,n)(0)||\widetilde{a}_{\,\cdot\,,3}(\gamma_{c,n})||_{\ell^{\infty}({\mathbb{N}}_{0})} must be computed separately as a~,3(γc,n)(0)=|a~1,3(γc,n)|=8/[33/2]||\widetilde{a}_{\,\cdot\,,3}(\gamma_{c,n})||_{\ell^{\infty}({\mathbb{N}}_{0})}=|\widetilde{a}_{1,3}(\gamma_{c,n})|=8\big{/}\big{[}3^{3/2}\big{]}. ∎

To introduce the notion of finite truncations, one considers the operators

Pm=(Im000)=Im0,Km(γc,n)=PmK(γc,n)Pm,m,\displaystyle P_{m}=\begin{pmatrix}I_{{\mathbb{C}}^{m}}&0\\ 0&0\end{pmatrix}=I_{{\mathbb{C}}^{m}}\oplus 0,\quad K_{m}(\gamma_{c,n})=P_{m}K(\gamma_{c,n})P_{m},\quad m\in{\mathbb{N}}, (4.31)

on 2(0)\ell^{2}({\mathbb{N}}_{0}) (with ImI_{{\mathbb{C}}^{m}} denoting the identity matrix in m{\mathbb{C}}^{m}, mm\in{\mathbb{N}}).

We also introduce the finite N×NN\times N tri-diagonal Jacobi matrices JN(a~0,,a~N1)J_{N}(\widetilde{a}_{0},\dots,\widetilde{a}_{N-1}) in N{\mathbb{C}}^{N}, NN\in{\mathbb{N}}, N2N\geqslant 2, denoted by

JN(a~1,,a~N1)=(0a~1a~10a~2𝟎a~20a~3a~30𝟎a~N1a~N10),N,N2,J_{N}(\widetilde{a}_{1},\dots,\widetilde{a}_{N-1})=\begin{pmatrix}0&\widetilde{a}_{1}&\phantom{0}&\phantom{0}&\phantom{0}&\phantom{0}\\ \widetilde{a}_{1}&0&\widetilde{a}_{2}&\phantom{0}&\bf{0}&\phantom{0}\\ \phantom{0}&\widetilde{a}_{2}&0&\widetilde{a}_{3}&\phantom{0}&\phantom{0}\\ \phantom{0}&\phantom{0}&\widetilde{a}_{3}&0&\ddots&\phantom{0}\\ \phantom{0}&\bf{0}&\phantom{0}&\ddots&\ddots&\widetilde{a}_{N-1}\\ \phantom{0}&\phantom{0}&\phantom{0}&\phantom{0}&\widetilde{a}_{N-1}&0\\ \end{pmatrix},\quad N\in{\mathbb{N}},N\geqslant 2, (4.32)

in particular,

detN(zINJN(a~1,,a~N1))=zdetN1(zIN1JN1(a~2,,a~N1))\displaystyle{\det}_{{\mathbb{C}}^{N}}(zI_{N}-J_{N}(\widetilde{a}_{1},\dots,\widetilde{a}_{N-1}))=z{\det}_{{\mathbb{C}}^{N-1}}(zI_{N-1}-J_{N-1}(\widetilde{a}_{2},\dots,\widetilde{a}_{N-1}))
[a~1]2detN2(zIN2JN2(a~3,,a~N1)),z,\displaystyle\qquad-[\widetilde{a}_{1}]^{2}{\det}_{{\mathbb{C}}^{N-2}}(zI_{N-2}-J_{N-2}(\widetilde{a}_{3},\dots,\widetilde{a}_{N-1})),\quad z\in{\mathbb{C}},
={zP(N1)/2(z2),P(N1)/2(0)0,N odd,QN/2(z2),QN/2(0)0,N even,\displaystyle\quad=\begin{cases}zP_{(N-1)/2}\big{(}z^{2}\big{)},\quad P_{(N-1)/2}(0)\neq 0,&N\text{ odd,}\\ Q_{N/2}\big{(}z^{2}\big{)},\quad Q_{N/2}(0)\neq 0,&N\text{ even,}\end{cases} (4.33)

where P(N1)/2()P_{(N-1)/2}(\,\cdot\,) and QN/2()Q_{N/2}(\,\cdot\,) are monic polynomials of degree (N1)/2(N-1)/2 and N/2N/2, respectively.

Thus, the spectrum of each JN(a~0,,a~N1)J_{N}(\widetilde{a}_{0},\dots,\widetilde{a}_{N-1}) consists of NN real eigenvalues, symmetric with respect to the origin, the eigenvalues being simple as long as a~j>0\widetilde{a}_{j}>0, 1jN11\leqslant j\leqslant N-1 (see, e.g., [33, Theorem II.1.1], [81, Remark 1.10 and p. 120]). Explicitly,

det2((zI2J2(a~1))=z2[a~1]2,\displaystyle{\det}_{{\mathbb{C}}^{2}}((zI_{{\mathbb{C}}^{2}}-J_{2}(\widetilde{a}_{1}))=z^{2}-[\widetilde{a}_{1}]^{2},
det3((zI3J3(a~1,a~2))=z{z2[a~1]2[a~2]2},\displaystyle{\det}_{{\mathbb{C}}^{3}}((zI_{{\mathbb{C}}^{3}}-J_{3}(\widetilde{a}_{1},\widetilde{a}_{2}))=z\big{\{}z^{2}-[\widetilde{a}_{1}]^{2}-[\widetilde{a}_{2}]^{2}\big{\}},
det4((zI4J4(a~1,a~2,a~3))=z4{[a~1]2+[a~2]2+[a~3]2}z2+[a~1]2[a~3]2,\displaystyle{\det}_{{\mathbb{C}}^{4}}((zI_{{\mathbb{C}}^{4}}-J_{4}(\widetilde{a}_{1},\widetilde{a}_{2},\widetilde{a}_{3}))=z^{4}-\big{\{}[\widetilde{a}_{1}]^{2}+[\widetilde{a}_{2}]^{2}+[\widetilde{a}_{3}]^{2}\big{\}}z^{2}+[\widetilde{a}_{1}]^{2}[\widetilde{a}_{3}]^{2}, (4.34)
det5((zI5J5(a~1,a~2,a~3,a~4))=z{z4{[a~1]2+[a~2]2+[a~3]2+[a~4]2}z2\displaystyle{\det}_{{\mathbb{C}}^{5}}((zI_{{\mathbb{C}}^{5}}-J_{5}(\widetilde{a}_{1},\widetilde{a}_{2},\widetilde{a}_{3},\widetilde{a}_{4}))=z\big{\{}z^{4}-\big{\{}[\widetilde{a}_{1}]^{2}+[\widetilde{a}_{2}]^{2}+[\widetilde{a}_{3}]^{2}+[\widetilde{a}_{4}]^{2}\big{\}}z^{2}
+[a~1]2[a~3]2+[a~1]2[a~4]2+[a~2]2[a~4]2},\displaystyle\hskip 157.91287pt+[\widetilde{a}_{1}]^{2}[\widetilde{a}_{3}]^{2}+[\widetilde{a}_{1}]^{2}[\widetilde{a}_{4}]^{2}+[\widetilde{a}_{2}]^{2}[\widetilde{a}_{4}]^{2}\big{\}},
etc.

In addition, we introduce the unitary, self-adjoint, diagonal operator WW in 2(0)\ell^{2}({\mathbb{N}}_{0}) as

W=((1)pδp,q)(p,q)02,W1=W=W.W=\big{(}(-1)^{p}\delta_{p,q}\big{)}_{(p,q)\in{\mathbb{N}}_{0}^{2}},\quad W^{-1}=W=W^{*}. (4.35)
Theorem 4.2.

Given the operators K(γc,n)K(\gamma_{c,n}), Km(γc,n)K_{m}(\gamma_{c,n}), mm\in{\mathbb{N}}, and WW as in (4.20), (4.31)–(4.35), one concludes that K(γc,n)K(\gamma_{c,n}) and K(γc,n)-K(\gamma_{c,n}) as well as Km(γc,n)K_{m}(\gamma_{c,n}) and Km(γc,n)-K_{m}(\gamma_{c,n}) are unitarily equivalent,

K(γc,n)=WK(γc,n)W1,Km(γc,n)=WKm(γc,n)W1,m,-K(\gamma_{c,n})=WK(\gamma_{c,n})W^{-1},\quad-K_{m}(\gamma_{c,n})=WK_{m}(\gamma_{c,n})W^{-1},\;m\in{\mathbb{N}}, (4.36)

and hence the spectra of K(γc,n)K(\gamma_{c,n}) and Km(γc,n)K_{m}(\gamma_{c,n}), mm\in{\mathbb{N}}, are symmetric with respect to zero. Moreover, all nonzero eigenvalues of K(γc,n)K(\gamma_{c,n}) and Km(γc,n)K_{m}(\gamma_{c,n}), mm\in{\mathbb{N}}, are simple. In addition,

limmKm(γc,n)K(γc,n)(2(0))=0,\lim_{m\to\infty}\|K_{m}(\gamma_{c,n})-K(\gamma_{c,n})\|_{{\mathcal{B}}(\ell^{2}({\mathbb{N}}_{0}))}=0, (4.37)

and333Here σess()\sigma_{ess}(\,\cdot\,) denotes the essential spectrum.

σ(K(γc,n))=limmσ(Km(γc,n)),σess(K(γc,n))=σess(Km(γc,n))={0},m.\displaystyle\begin{split}&\sigma(K(\gamma_{c,n}))=\underset{m\to\infty}{\lim}\sigma(K_{m}(\gamma_{c,n})),\\ &\sigma_{ess}(K(\gamma_{c,n}))=\sigma_{ess}(K_{m}(\gamma_{c,n}))=\{0\},\;m\in{\mathbb{N}}.\end{split} (4.38)

In particular, λσ(K(γc,n))\lambda\in\sigma(K(\gamma_{c,n})) if and only if there is a sequence (λm)m(\lambda_{m})_{m\in{\mathbb{N}}} with λmσ(Km(γc,n))\lambda_{m}\in\sigma(K_{m}(\gamma_{c,n})) such that λmmλ\lambda_{m}\underset{m\to\infty}{\longrightarrow}\lambda.

Proof.

The symmetry fact (4.36) follows from an elementary computation. That all eigenvalues of K(γc,n)K(\gamma_{c,n}) are simple follows from the fact that K(γc,n)K(\gamma_{c,n}) is a half-lattice operator with a~(γc,n)>0\widetilde{a}_{\ell}(\gamma_{c,n})>0, \ell\in{\mathbb{N}}, and hence the half-lattice does not decouple into a disjoint union of subsets (resp., K(γc,n)K(\gamma_{c,n}) does not reduce to a direct sum of operators in 2(0)\ell^{2}({\mathbb{N}}_{0})). The same argument applies to the finite-lattice operators Km(γc,n)K_{m}(\gamma_{c,n}), mm\in{\mathbb{N}}.

One notices that slimmPm=I2(0)\operatorname*{s-lim}_{m\to\infty}P_{m}=I_{\ell^{2}({\mathbb{N}}_{0})}, where strong operator convergence is abbreviated by slim\operatorname*{s-lim}. Together with the compactness of K(γc,n)K(\gamma_{c,n}) given in Proposition 4.1, one obtains

limmPmK(γc,n)K(γc,n)(2(0))=0,\lim_{m\to\infty}\|P_{m}K(\gamma_{c,n})-K(\gamma_{c,n})\|_{{\mathcal{B}}(\ell^{2}({\mathbb{N}}_{0}))}=0, (4.39)

applying [3, Proposition 3.11]. The norm convergence in (4.39), together with the uniform bound Pm(2(0))=1\|P_{m}\|_{{\mathcal{B}}(\ell^{2}({\mathbb{N}}_{0}))}=1, mm\in{\mathbb{N}}, yields (4.37). The latter implies (4.38) as a consequence of [71, Theorem VIII.23 (a) and Theorem VIII.24 (a)] (see also [89, Satz 9.24 a)a)]), taking into account that norm resolvent convergence of a sequence of self-adjoint operators is equivalent to norm convergence of a uniformly bounded sequence of self-adjoint operators in a complex Hilbert space (see [71, Theorem VIII.18], [89, Satz 9.22 a) (ii)(ii)]). ∎

Returning to the dipole context, one may now compute approximants of γc,n\gamma_{c,n} by approximating the smallest negative eigenvalues of K(γc,n)K(\gamma_{c,n}) in terms of the smallest negative eigenvalue of Km(γc,n)K_{m}(\gamma_{c,n}) with increasing mm\in{\mathbb{N}}. Using K7K_{7}, (which produced 16 stable digits in the case n=3n=3), one obtains the following values and approximants for 3n103\leqslant n\leqslant 10:

nn lower bound for γc,n\gamma_{c,n} γc,n\gamma_{c,n} upper bound γc,n\gamma_{c,n}
3 0.250 1.279 4.418
4 1.000 3.790 5.890
5 2.598 7.584 10.308
6 5.846 12.672 17.672
7 10.392 19.058 27.980
8 16.238 26.742 41.233
9 23.383 35.725 57.432
10 31.826 46.006 76.576
Refer to caption
Figure 1. Dimension nn vs. Critical Dipole Moment γc,n\gamma_{c,n} (•) with its upper (\blacktriangle) and lower (\blacksquare) bounds.

Here the lower and upper bounds for γc,n\gamma_{c,n} correspond to the values displayed in (3.44) (see Fig. 1).

The result for γc,3\gamma_{c,3} is in excellent agreement with the ones found in the literature (see, e.g., [2], [12], [16], [17], [32], [60], [83], and [84]). The approximate values of γc,n\gamma_{c,n} for n4n\geqslant 4 (but surprisingly, not for n=3n=3) are in good agreement with those obtained in [29, p. 98].

5. Multicenter Extensions

Combining the results of this manuscript with those in [38] one can extend the scope of this investigation to include multicenter dipole interactions, that is, sums of point dipoles supported on an infinite discrete set (a set of distinct points spaced apart by a minimal distance ε>0\varepsilon>0). For various related studies on multicenter singular interactions, see, for instance, [11], [14], [22], [28], [29], [30], [31], [38], [48], [62], [80].

To set the stage, we recall the notion of relative form boundedness in the special context of self-adjoint operators.

Definition 5.1.

Suppose that AA is self-adjoint in a complex Hilbert space {\mathcal{H}} and bounded from below, that is, AcIA\geqslant cI_{{\mathcal{H}}} for some cc\in{{\mathbb{R}}}. Then the sesquilinear form QAQ_{A} associated with AA is denoted by

QA(f,g)=((AcI)1/2f,(AcI)1/2g)+c(f,g),fdom(|A|1/2).Q_{A}(f,g)=\big{(}(A-cI_{{\mathcal{H}}})^{1/2}f,(A-cI_{{\mathcal{H}}})^{1/2}g\big{)}_{{\mathcal{H}}}+c(f,g)_{{\mathcal{H}}},\quad f\in\operatorname{dom}\big{(}|A|^{1/2}\big{)}. (5.1)

A sesquilinear form qq in {\mathcal{H}} satisfying dom(q)dom(QA)\operatorname{dom}(q)\supseteq\operatorname{dom}(Q_{A}) is called bounded with respect to the form QAQ_{A} if for some a,b0a,b\geqslant 0,

|q(f,f)|aQA(f,f)+bf2,fdom(QA).|q(f,f)|\leqslant a\,Q_{A}(f,f)+b\,\|f\|_{{\mathcal{H}}}^{2},\quad f\in\operatorname{dom}(Q_{A}). (5.2)

The infimum of all numbers aa for which there exists b[0,)b\in[0,\infty) such that (5.1) holds is called the bound of qq with respect to QAQ_{A}.

The following result is a variant of [38, Theorem 3.2], which in turn is an abstract version of Morgan [65, Theorem 2.1] (see also [18, Proposition 3.3], [49], [57, Sect. 4]). Throughout this section, infinite sums are understood in the weak operator topology and JJ\subseteq{\mathbb{N}} denotes an index set.

Lemma 5.2.

Suppose that TT is a self-adjoint operator in {\mathcal{H}} bounded from below, TcIT\geqslant cI_{{\mathcal{H}}} for some cc\in{\mathbb{R}}, and WW is a self-adjoint operator in {\mathcal{H}} such that

dom(|T|1/2)dom(|W|1/2).\operatorname{dom}\big{(}|T|^{1/2}\big{)}\subseteq\operatorname{dom}\big{(}|W|^{1/2}\big{)}. (5.3)

We abbreviate

qW(f,g)=(|W|1/2f,sgn(W)|W|1/2g),fdom(|W|1/2).q_{W}(f,g)=\big{(}|W|^{1/2}f,\operatorname*{sgn}(W)|W|^{1/2}g\big{)}_{{\mathcal{H}}},\quad f\in\operatorname{dom}\big{(}|W|^{1/2}\big{)}. (5.4)

Let d,D(0,)d,D\in(0,\infty), e[0,)e\in[0,\infty), assume that Φj()\Phi_{j}\in{\mathcal{B}}({\mathcal{H}}), jJj\in J, leave dom(|T|1/2)\operatorname{dom}\big{(}|T|^{1/2}\big{)} invariant, that is,

Φjdom(|T|1/2)dom(|T|1/2),jJ,\Phi_{j}\operatorname{dom}\big{(}|T|^{1/2}\big{)}\subseteq\operatorname{dom}\big{(}|T|^{1/2}\big{)},\quad j\in J, (5.5)

and suppose that the following conditions (i)(i)(iii)(iii) hold:
(i)(i) jJΦjΦjI\sum_{j\in J}\Phi_{j}^{*}\Phi_{j}\leqslant I_{{\mathcal{H}}}.
(ii)(ii) jJ|qW(Φjf,Φjf)|D1|qW(f,f)|\sum_{j\in J}|q_{W}(\Phi_{j}f,\Phi_{j}f)|\geqslant D^{-1}|q_{W}(f,f)|,   fdom(|T|1/2)f\in\operatorname{dom}\big{(}|T|^{1/2}\big{)}.
(iii)(iii) jJ|T|1/2Φjf2d|T|1/2f2+ef2\sum_{j\in J}\||T|^{1/2}\Phi_{j}f\|_{{\mathcal{H}}}^{2}\leqslant d\||T|^{1/2}f\|_{{\mathcal{H}}}^{2}+e\|f\|_{{\mathcal{H}}}^{2},   fdom(|T|1/2)f\in\operatorname{dom}\big{(}|T|^{1/2}\big{)}.
Then,

|qW(Φjf,Φjf)|aqT(Φjf,Φjf)+bΦjf2,fdom(|T|1/2),jJ,|q_{W}(\Phi_{j}f,\Phi_{j}f)|\leqslant a\,q_{T}(\Phi_{j}f,\Phi_{j}f)+b\|\Phi_{j}f\|_{{\mathcal{H}}}^{2},\quad f\in\operatorname{dom}(|T|^{1/2}),\;j\in J, (5.6)

implies

|qW(f,f)|adDqT(f,f)+[ae+b]Df2,fdom(|T|1/2).|q_{W}(f,f)|\leqslant a\,d\,D\,q_{T}(f,f)+[a\,e+b]D\,\|f\|_{{\mathcal{H}}}^{2},\quad f\in\operatorname{dom}(|T|^{1/2}). (5.7)
Proof.

For fdom(|T|1/2)f\in\operatorname{dom}\big{(}|T|^{1/2}\big{)} one computes

|qW(f,f)|\displaystyle|q_{W}(f,f)| DjJ|qW(Φjf,Φjf)|(by (ii))\displaystyle\leqslant D\sum_{j\in J}|q_{W}(\Phi_{j}f,\Phi_{j}f)|\quad\text{(by $(ii)$)}
DjJ[a|T|1/2Φjf2+bΦjf2](by (5.6))\displaystyle\leqslant D\sum_{j\in J}\left[a\big{\|}|T|^{1/2}\Phi_{j}f\big{\|}_{{\mathcal{H}}}^{2}+b\|\Phi_{j}f\|_{{\mathcal{H}}}^{2}\right]\quad\text{(by \eqref{5.6})}
adD|T|1/2f2+(aeD+bD)f2(by (i) and (iii))\displaystyle\leqslant a\,d\,D\big{\|}|T|^{1/2}f\big{\|}_{{\mathcal{H}}}^{2}+(a\,e\,D+b\,D)\|f\|_{{\mathcal{H}}}^{2}\quad\text{(by $(i)$ and $(iii)$)}
=adDqT(f,f)+[ae+b]Df2,\displaystyle=a\,d\,D\,q_{T}(f,f)+[a\,e+b]D\,\|f\|_{{\mathcal{H}}}^{2}, (5.8)

completing the proof. ∎

Remark 5.3.

Considering the concrete case of

H0=Δ,dom(H0)=H2(n),\displaystyle H_{0}=-\Delta,\quad\operatorname{dom}(H_{0})=H^{2}({{\mathbb{R}}}^{n}),
QH0(f,g)=((Δ)1/2f,(Δ)1/2g)L2(n)=(f,g)[L2(n)]n,\displaystyle\,Q_{H_{0}}(f,g)=\big{(}(-\Delta)^{1/2}f,(-\Delta)^{1/2}g\big{)}_{L^{2}({\mathbb{R}}^{n})}=(\nabla f,\nabla g)_{[L^{2}({\mathbb{R}}^{n})]^{n}}, (5.9)
f,gdom(QH0)=H1(n),\displaystyle\hskip 147.95424ptf,g\in\operatorname{dom}(Q_{H_{0}})=H^{1}({\mathbb{R}}^{n}),

in L2(n)L^{2}({{\mathbb{R}}}^{n}), and assuming that WW, the operator of multiplication with a measurable and a.e. real-valued function W()W(\,\cdot\,), employing a slight abuse of notation, satisfies (5.3) (for sufficient conditions on WW, see, e.g., [88, Theorems 10.17 (b), 10.18] with r=1r=1). Let {ϕj}jJ\{\phi_{j}\}_{j\in J}, JJ\subseteq{\mathbb{N}}, be a family of smooth, real-valued functions defined on n{{\mathbb{R}}}^{n} in such a manner that for each xnx\in{{\mathbb{R}}}^{n}, there exists an open neighborhood UxnU_{x}\subset{{\mathbb{R}}}^{n} of xx such that there exist only finitely many indices kJk\in J with supp(ϕk)Ux\operatorname{supp}\,(\phi_{k})\cap U_{x}\neq\emptyset and ϕk|Ux0\phi_{k}|_{U_{x}}\neq 0, as well as

jJϕj(x)2=1,xn\sum_{j\in J}\phi_{j}(x)^{2}=1,\quad x\in{{\mathbb{R}}}^{n} (5.10)

(the sum over jJj\in J in (5.10) being finite). Finally, let Φj\Phi_{j} be the operator of multiplication by the function ϕj\phi_{j}, jJj\in J. Then one notes that for these choices, hypothesis (i)(i) holds with equality, and hypothesis (ii)(ii) with D=1D=1 follows from (i)(i). Moreover, item (iii)(iii) holds with d=1d=1 as long as

e=jJ|ϕj()|2L(n)<.e=\bigg{\|}\sum_{j\in J}|\nabla\phi_{j}(\cdot)|^{2}\bigg{\|}_{L^{\infty}({{\mathbb{R}}}^{n})}<\infty. (5.11)

To verify this, one observes that |H0|1/2ϕfL2(n)2=ndnx|(ϕ(x)f(x))|2\||H_{0}|^{1/2}\phi f\|_{L^{2}({{\mathbb{R}}}^{n})}^{2}=\int_{{{\mathbb{R}}}^{n}}d^{n}x\,|\nabla(\phi(x)f(x))|^{2} and that the cross terms vanish since jJϕj(x)(ϕj)(x)=0\sum_{j\in J}\phi_{j}(x)(\nabla\phi_{j})(x)=0, xnx\in{{\mathbb{R}}}^{n}, by condition (5.10). (We note again that the latter sum over jJj\in J contains only finitely many terms in every bounded neighborhood of xnx\in{{\mathbb{R}}}^{n}.) \diamond

Strongly singular potentials that are covered by Lemma 5.2 are, for instance, of the following form: Let JJ\subseteq{\mathbb{N}} be an index set, ε>0\varepsilon>0, and {xj}jJn\{x_{j}\}_{j\in J}\subset{{\mathbb{R}}}^{n}, nn\in{\mathbb{N}}, n3n\geqslant 3, be a set of points such that

infj,jJjj|xjxj|ε.\inf_{\begin{subarray}{c}j,j^{\prime}\in J\\ j\neq j^{\prime}\end{subarray}}|x_{j}-x_{j^{\prime}}|\geqslant\varepsilon. (5.12)

In addition, let γj\gamma_{j}\in{{\mathbb{R}}}, jJj\in J, γ0[0,)\gamma_{0}\in[0,\infty), with

|γj|γ0<(n2)2/4,jJ,|\gamma_{j}|\leqslant\gamma_{0}<(n-2)^{2}/4,\quad j\in J, (5.13)

and

W{γj}jJ(x)=jJγj|xxj|2χBn(xj;ε/4)(x)+W0(x),xn\{xj}jJ,W_{\{\gamma_{j}\}_{j\in J}}(x)=\sum_{j\in J}\gamma_{j}|x-x_{j}|^{-2}\chi_{B_{n}(x_{j};\varepsilon/4)}(x)+W_{0}(x),\quad x\in{{\mathbb{R}}}^{n}\backslash\{x_{j}\}_{j\in J}, (5.14)

with

W0L(n), W0 real-valued a.e. on n,W_{0}\in L^{\infty}({\mathbb{R}}^{n}),\,\text{ $W_{0}$ real-valued~{}a.e.~{}on ${\mathbb{R}}^{n}$,} (5.15)

and Bn(x0;r)B_{n}(x_{0};r) the open ball in n{\mathbb{R}}^{n} of radius r>0r>0, centered at x0nx_{0}\in{\mathbb{R}}^{n}.

Then an application of Hardy’s inequality in n{{\mathbb{R}}}^{n}, n3n\geqslant 3 (cf. (LABEL:1.1)), shows that W{γj}jJW_{\{\gamma_{j}\}_{j\in J}} is form bounded with respect to T0T_{0} in (5.9) with form bound strictly less than one.

At this point one can extend existing results of [29], [30] regarding quadratic form estimates for multicenter dipole interactions as follows.

Theorem 5.4.

Given (5.9)–(5.12) and W0W_{0} in (5.15), we introduce γ0,γj[0,)\gamma_{0},\gamma_{j}\in[0,\infty), jJj\in J, satisfying

0γjγ0<γc,n,jJ,0\leqslant\gamma_{j}\leqslant\gamma_{0}<\gamma_{c,n},\quad j\in J, (5.16)

and

q{γj}jJ(f,g)=jJγjndnx(u,(xxj))|xxj|3χBn(xj;ε/4)(x)f(x)¯g(x),+ndnxW0(x)f(x)¯g(x),f,gH1(n).\displaystyle\begin{split}q_{\{\gamma_{j}\}_{j\in J}}(f,g)&=\sum_{j\in J}\gamma_{j}\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,(x-x_{j}))|x-x_{j}|^{-3}\chi_{B_{n}(x_{j};\varepsilon/4)}(x)\overline{f(x)}g(x),\\ &\quad+\int_{{\mathbb{R}}^{n}}d^{n}x\,W_{0}(x)\overline{f(x)}g(x),\quad f,g\in H^{1}({\mathbb{R}}^{n}).\end{split} (5.17)

Then q{γj}jJq_{\{\gamma_{j}\}_{j\in J}} is bounded with respect to QT0Q_{T_{0}} in (5.9) with form bound strictly less than one.

Proof.

Without loss of generality we put W0=0W_{0}=0. Inequality (3.5) and the analogous inequality with unu\in{\mathbb{R}}^{n}, |u|=1|u|=1, replaced by u-u yields

for all γ[0,γc,n],ndnx|(f)(x)|2±γndnx(u,x)|x|3|f(x)|2,fH1(n),\displaystyle\begin{split}&\text{for all $\gamma\in[0,\gamma_{c,n}]$,}\\ &\quad\int_{{\mathbb{R}}^{n}}d^{n}x\,|(\nabla f)(x)|^{2}\geqslant\pm\gamma\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2},\quad f\in H^{1}({\mathbb{R}}^{n}),\end{split} (5.18)

and hence for some εγ0(0,1)\varepsilon_{\gamma_{0}}\in(0,1) and c(εγ0)(0,)c(\varepsilon_{\gamma_{0}})\in(0,\infty),

±γ0ndnx(u,x)|x|3|f(x)|2(1εγ0)f[L2(n)]n2+c(εγ0)fL2(n)2,γ0[0,γc,n),fH1(n).\displaystyle\begin{split}&\pm\gamma_{0}\int_{{\mathbb{R}}^{n}}d^{n}x\,(u,x)|x|^{-3}|f(x)|^{2}\leqslant(1-\varepsilon_{\gamma_{0}})\big{\|}\nabla f\big{\|}_{[L^{2}({\mathbb{R}}^{n})]^{n}}^{2}\\ &\quad+c(\varepsilon_{\gamma_{0}})\|f\|_{L^{2}({\mathbb{R}}^{n})}^{2},\quad\gamma_{0}\in[0,\gamma_{c,n}),\;f\in H^{1}({\mathbb{R}}^{n}).\end{split} (5.19)

Thus, an application of Lemma 5.2, taking into account that d=D=1d=D=1 as described in Remark 5.3, implies for some C(εγ0)(0,)C(\varepsilon_{\gamma_{0}})\in(0,\infty),

|q{γj}jJ(f,f)|(1εγ0)f[L2(n)]n2+C(εγ0)fL2(n)2,fH1(n),\displaystyle\begin{split}|q_{\{\gamma_{j}\}_{j\in J}}(f,f)|\leqslant(1-\varepsilon_{\gamma_{0}})\big{\|}\nabla f\big{\|}_{[L^{2}({\mathbb{R}}^{n})]^{n}}^{2}+C(\varepsilon_{\gamma_{0}})\|f\|_{L^{2}({\mathbb{R}}^{n})}^{2},&\\ f\in H^{1}({\mathbb{R}}^{n}),&\end{split} (5.20)

as was to be proven. ∎

Thus, Theorem 5.4 proves semiboundedness of the self-adjoint multicenter dipole Hamiltonian H{γj}jJH_{\{\gamma_{j}\}_{j\in J}} in L2(n)L^{2}({\mathbb{R}}^{n}), uniquely associated with the quadratic form sum

QH{γj}jJ=QT0+q{γj}jJ,dom(QH{γj}jJ)=H1(n),Q_{H_{\{\gamma_{j}\}_{j\in J}}}=Q_{T_{0}}+q_{\{\gamma_{j}\}_{j\in J}},\quad\operatorname{dom}(Q_{H_{\{\gamma_{j}\}_{j\in J}}})=H^{1}({\mathbb{R}}^{n}), (5.21)

under very general hypotheses on {xj}jJn\{x_{j}\}_{j\in J}\subset{{\mathbb{R}}}^{n} and {γj}jJ\{\gamma_{j}\}_{j\in J}. We note that [29], [30] derive sufficient conditions {xj}jJn\{x_{j}\}_{j\in J}\subset{{\mathbb{R}}}^{n} and {γj}jJ\{\gamma_{j}\}_{j\in J} to guarantee nonnegativity of H{γj}jJH_{\{\gamma_{j}\}_{j\in J}} and also discuss situations characterized by the lack of nonnegativity of H{γj}jJH_{\{\gamma_{j}\}_{j\in J}}.

Finally, we sketch how Remark 3.8 extends to the multicenter situation.

Remark 5.5.

Assume (5.9)–(5.12), let W0W_{0} as in (5.15), suppose γj[0,)\gamma_{j}\in[0,\infty), and introduce

V(x)=jJ[γj(u,(xxj))|xxj|3+V~j(|xxj|)]χBn(xj;ε/4)(x)+W0(x),\displaystyle V(x)=\sum_{j\in J}\big{[}\gamma_{j}(u,(x-x_{j}))|x-x_{j}|^{-3}+\widetilde{V}_{j}(|x-x_{j}|)\big{]}\chi_{B_{n}(x_{j};\varepsilon/4)}(x)+W_{0}(x),
xn\{xj}jJ,\displaystyle\hskip 233.3125ptx\in{{\mathbb{R}}}^{n}\backslash\{x_{j}\}_{j\in J}, (5.22)

with

rV~j(r)L1((0,ε);dr)Lloc((0,ε];dr),jJ.r\widetilde{V}_{j}(r)\in L^{1}((0,\varepsilon);dr)\cap L^{\infty}_{loc}((0,\varepsilon];dr),\quad j\in J. (5.23)

Consider the minimally defined Schrödinger operator H.{γj}jJ\overset{\textbf{\Large.}}{H}_{\{\gamma_{j}\}_{j\in J}} in L2(n)L^{2}({\mathbb{R}}^{n}) given by

H.{γj}jJ=Δ+V(),dom(H.{γj}jJ)=C0(n\{xj}jJ).\overset{\textbf{\Large.}}{H}_{\{\gamma_{j}\}_{j\in J}}=-\Delta+V(\,\cdot\,),\quad\operatorname{dom}\big{(}\overset{\textbf{\Large.}}{H}_{\{\gamma_{j}\}_{j\in J}}\big{)}=C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{x_{j}\}_{j\in J}). (5.24)

Then the criterion (3.103) and (2.31), (2.32) combined with [38, Theorems 1.1 and 5.8] imply that

H.{γj}jJ is essentially self-adjointif and only if λγj,n,0n(n4)/4 for each jJ.\displaystyle\begin{split}&\text{$\overset{\textbf{\Large.}}{H}_{\{\gamma_{j}\}_{j\in J}}$ is essentially self-adjoint}\\ &\quad\text{if and only if $\lambda_{\gamma_{j},n,0}\geqslant-n(n-4)/4$ for each $j\in J$.}\end{split} (5.25)

\diamond

Appendix A Spherical Harmonics and the Laplace–Beltrami Operator in L2(𝕊n1)L^{2}\big{(}{\mathbb{S}}^{n-1}\big{)}, n2n\geqslant 2.

In this appendix we summarize some of the results on spherical harmonics and the Laplace–Beltrami operator on the unit sphere 𝕊n1{\mathbb{S}}^{n-1} in dimensions nn\in{\mathbb{N}}, n2n\geqslant 2, following [5, Chs. 2,3], [19, Ch. 1], and [47, Ch. 2].

Assuming nn\in{\mathbb{N}}, n2n\geqslant 2, cartesian and polar coordinates (cf.  e.g., [9]) on 𝕊n1{\mathbb{S}}^{n-1} are given by

x=(x1,,xn)n,\displaystyle x=(x_{1},\dots,x_{n})\in{\mathbb{R}}^{n},
x=rω,ω=ω(θ)=ω(θ1,θ2,,θn1)=x/|x|𝕊n1,\displaystyle x=r\omega,\;\omega=\omega(\theta)=\omega(\theta_{1},\theta_{2},\dots,\theta_{n-1})=x/|x|\in{\mathbb{S}}^{n-1}, (A.1)
xk, 1kn,r=|x|[0,),θ1[0,2π),θj[0,π), 2jn1,\displaystyle x_{k}\in{\mathbb{R}},\,1\leqslant k\leqslant n,\;r=|x|\in[0,\infty),\;\theta_{1}\in[0,2\pi),\;\theta_{j}\in[0,\pi),\,2\leqslant j\leqslant n-1,

where (cf., e.g., [9], [19, Sect. 1.5])

{x1=rcos(θ1)j=2n1sin(θj),x2=rsin(θ1)j=2n1sin(θj),xn1=rcos(θn2)sin(θn1),xn=rcos(θn1).\begin{cases}x_{1}=r\cos(\theta_{1})\prod\limits_{j=2}^{n-1}\sin(\theta_{j}),\\[2.84526pt] x_{2}=r\sin(\theta_{1})\prod\limits_{j=2}^{n-1}\sin(\theta_{j}),\\ \;\vdots\\ x_{n-1}=r\cos(\theta_{n-2})\sin(\theta_{n-1}),\\ x_{n}=r\cos(\theta_{n-1}).\end{cases} (A.2)

The surface measure dn1ωd^{n-1}\omega on 𝕊n1{\mathbb{S}}^{n-1} and the volume element in n{\mathbb{R}}^{n} then read

dn1ω(θ)=dθ1j=2n1[sin(θj)]j1dθj,dnx=rn1drdn1ω(θ),d^{n-1}\omega(\theta)=d\theta_{1}\prod_{j=2}^{n-1}[\sin(\theta_{j})]^{j-1}d\theta_{j},\quad d^{n}x=r^{n-1}dr\,d^{n-1}\omega(\theta), (A.3)

in particular, the area ωn\omega_{n} of the unit sphere 𝕊n1{\mathbb{S}}^{n-1} in n{\mathbb{R}}^{n} is given by (cf.  [66, p. 2])

ωn=𝕊n1dn1ω(θ)=2πn/2/Γ(n/2).\displaystyle\omega_{n}=\int_{{\mathbb{S}}^{n-1}}d^{n-1}\omega(\theta)=2\pi^{n/2}/\Gamma(n/2). (A.4)

Turning to spherical harmonics next, we recall that a homogeneous polynomial P(x1,,xn)P(x_{1},\ldots,x_{n}) of degree 0\ell\in{\mathbb{N}}_{0} (in nn variables) satisfies P(tx1,,txn)=tnP(x1,,xn)P(tx_{1},\ldots,tx_{n})=t^{n}P(x_{1},\ldots,x_{n}) and is a linear combination of terms of degree \ell. The space of such polynomials with real coefficients is denoted 𝒫n\mathscr{P}_{\ell}^{n}. We define the harmonic homogeneous polynomials of degree \ell in nn variables by

n={P𝒫n|ΔP=0},\displaystyle\mathscr{H}_{\ell}^{n}=\big{\{}P\in\mathscr{P}_{\ell}^{n}\,\big{|}\,\Delta P=0\big{\}}, (A.5)

where Δ\Delta represents the Laplace differential expression on n{\mathbb{R}}^{n}. Restricting the elements of n\mathscr{H}_{\ell}^{n} to the sphere 𝕊n1{\mathbb{S}}^{n-1}, one obtains 𝒴n{\mathcal{Y}}_{\ell}^{n}, the space of spherical harmonics of degree \ell in nn dimensions. Spaces of different degrees are orthogonal with respect to the real inner product on the sphere,

(Y,Z)L2(𝕊n1)=𝕊n1dn1ω(θ)Y(θ)Z(θ)=0,Y𝒴n,Z𝒴n,,0,.\displaystyle\begin{split}(Y,Z)_{L^{2}({\mathbb{S}}^{n-1})}=\int_{{\mathbb{S}}^{n-1}}d^{n-1}\omega(\theta)\,Y(\theta)Z(\theta)=0,&\\ Y\in{\mathcal{Y}}_{\ell}^{n},\,Z\in{\mathcal{Y}}_{\ell^{\prime}}^{n},\;\ell,\ell^{\prime}\in{\mathbb{N}}_{0},\,\ell\neq\ell^{\prime}.&\end{split} (A.6)

The dimension of 𝒴n{\mathcal{Y}}_{\ell}^{n} equals that of n\mathscr{H}_{\ell}^{n} and is given by ([19, Corollary 1.1.4])

dim(n)=(+n1)(+n32)=2+n2+n2(+n2n2),\displaystyle\dim(\mathscr{H}_{\ell}^{n})=\binom{\ell+n-1}{\ell}-\binom{\ell+n-3}{\ell-2}=\frac{2\ell+n-2}{\ell+n-2}\binom{\ell+n-2}{n-2}, (A.7)

where we use the convention that the second binomial coefficient equals 0 when =0,1\ell=0,1, and replace the final fraction by 1 in the case where n=2n=2 and =0\ell=0. This is equivalently formulated in [66, Lemma 3, p. 4] as the generating series

1+x(1x)n1==0dim(𝒴n)x.\displaystyle\frac{1+x}{(1-x)^{n-1}}=\sum\limits_{\ell=0}^{\infty}\dim({\mathcal{Y}}_{\ell}^{n})\,x^{\ell}. (A.8)

Most importantly, the spherical harmonics are the eigenfunctions of the Laplace–Beltrami operator Δ𝕊n1\Delta_{{\mathbb{S}}^{n-1}} in L2(𝕊n1)L^{2}\big{(}{\mathbb{S}}^{n-1}\big{)}, satisfying the eigenvalue equation

(Δ𝕊n1Y)(θ)=(+n2)Y(θ),Y𝒴n,0.\displaystyle(-\Delta_{{\mathbb{S}}^{n-1}}Y)(\theta)=\ell(\ell+n-2)Y(\theta),\quad Y\in{\mathcal{Y}}_{\ell}^{n},\;\ell\in{\mathbb{N}}_{0}. (A.9)

Following [19, Sect. 1.5] an explicit characterization for the spherical harmonics reads as follows: Introducing the multi-index α=(α1,,αn)0n\alpha=(\alpha_{1},\ldots,\alpha_{n})\in{\mathbb{N}}_{0}^{n}, with |α|=j=1nαj|\alpha|=\sum\limits_{j=1}^{n}\alpha_{j}, and θ=(θ1,,θn1)\theta=(\theta_{1},\ldots,\theta_{n-1}), the spherical harmonics are of the form

Yα(θ)=[Nα]1gα(θ1)j=1n2[sin(θnj)]|αj+1|Cαjνj(cos(θnj)),\displaystyle Y_{\alpha}(\theta)=[N_{\alpha}]^{-1}g_{\alpha}(\theta_{1})\prod\limits_{j=1}^{n-2}[\sin(\theta_{n-j})]^{|\alpha^{j+1}|}C_{\alpha_{j}}^{\nu_{j}}(\cos(\theta_{n-j})), (A.10)

where

|αj|=k=jn1αk,νj=|αj+1|+[(nj1)/2],\displaystyle|\alpha^{j}|=\sum\limits_{k=j}^{n-1}\alpha_{k},\quad\,\nu_{j}=|\alpha^{j+1}|+[(n-j-1)/2], (A.11)
gα(θ1)={cos(αn1θ1),αn=0,sin(αn1θ1),αn=1,\displaystyle\,g_{\alpha}(\theta_{1})=\begin{cases}\cos(\alpha_{n-1}\theta_{1}),&\alpha_{n}=0,\\ \sin(\alpha_{n-1}\theta_{1}),&\alpha_{n}=1,\end{cases} (A.12)
[Nα]2=bαj=1n2[αj!]([(nj+1)/2])|αj+1|(αj+νj)(2νj)αj([(nj)/2])|αj+1|νj,bα={2,αn1+αn>0,1,otherwise.\displaystyle[N_{\alpha}]^{2}=b_{\alpha}\prod_{j=1}^{n-2}\frac{[\alpha_{j}!]([(n-j+1)/2])_{|\alpha^{j+1}|}(\alpha_{j}+\nu_{j})}{(2\nu_{j})_{\alpha_{j}}([(n-j)/2])_{|\alpha^{j+1}|}\nu_{j}},\quad b_{\alpha}=\begin{cases}2,&\alpha_{n-1}+\alpha_{n}>0,\\ 1,&\text{otherwise.}\end{cases} (A.13)

Here the Pochhammer symbol (x)a(x)_{a} is defined by

(x)0=1,(x)n=Γ(x+n)/Γ(n)=x(x+1)(x+n1),n,(x)_{0}=1,\quad(x)_{n}=\Gamma(x+n)/\Gamma(n)=x(x+1)\cdots(x+n-1),\quad n\in{\mathbb{N}}, (A.14)

and Cnλ()C^{\lambda}_{n}(\,\cdot\,) represent the Gegenbauer (or ultrasperical) polynomials, see, for instance, [1, Ch. 22], [19, Appendix B].

The set {Yα||α|=,αn=0,1}\{Y_{\alpha}\,|\,|\alpha|=\ell,\alpha_{n}=0,1\} represents an orthonormal basis of 𝒴n{\mathcal{Y}}_{\ell}^{n}.

Finally, we recall the expression of the Laplace–Beltrami differential expression on 𝕊n1{\mathbb{S}}^{n-1} in spherical coordinates. From [5, p. 94], [19, Lemma 1.4.2], one obtains the recursion444For clarity we indicate the space dimension nn\in{\mathbb{N}} as a subscript in the Laplacian Δn-\Delta_{n} for the remainder of this appendix.

Δ𝕊1\displaystyle-\Delta_{{\mathbb{S}}^{1}} =2θ12,\displaystyle=-\dfrac{{\partial}^{2}}{{\partial}\theta_{1}^{2}},
Δ𝕊2\displaystyle-\Delta_{{\mathbb{S}}^{2}} =1sin(θ2)θ2(sin(θ2)θ2)1sin2(θ2)2θ12,\displaystyle=-\frac{1}{\sin(\theta_{2})}\frac{\partial}{\partial\theta_{2}}\bigg{(}\sin(\theta_{2})\frac{\partial}{\partial\theta_{2}}\bigg{)}-\frac{1}{\sin^{2}(\theta_{2})}\frac{\partial^{2}}{\partial\theta_{1}^{2}}, (A.15)
Δ𝕊n1\displaystyle-\Delta_{{\mathbb{S}}^{n-1}} =2θn12(n2)cot(θn1)θn1[sin(θn1)]2Δ𝕊n2,n3.\displaystyle=-\dfrac{{\partial}^{2}}{{\partial}\theta_{n-1}^{2}}-(n-2)\cot(\theta_{n-1})\dfrac{{\partial}}{{\partial}\theta_{n-1}}-[\sin(\theta_{n-1})]^{-2}\Delta_{{\mathbb{S}}^{n-2}},\quad n\geqslant 3.

Explicitly (cf. [19, p. 19]),

Δ𝕊n1=[sin(θn1)]2nθn1[[sin(θn1)]n2θn1]j=1n2(k=j+1n1[sin(θk)]2)[sin(θj)]1jθj[[sin(θj)]j1θj]=j=1n1(k=1j1[sin(θnk)]2)[sin(θnj)]1jn×θnj[[sin(θnj)]nj1θnj].\displaystyle\begin{split}-\Delta_{{\mathbb{S}}^{n-1}}&=-[\sin(\theta_{n-1})]^{2-n}\frac{\partial}{\partial\theta_{n-1}}\bigg{[}[\sin(\theta_{n-1})]^{n-2}\frac{\partial}{\partial\theta_{n-1}}\bigg{]}\\ &\quad-\sum_{j=1}^{n-2}\bigg{(}\prod_{k=j+1}^{n-1}[\sin(\theta_{k})]^{-2}\bigg{)}[\sin(\theta_{j})]^{1-j}\frac{\partial}{\partial\theta_{j}}\bigg{[}[\sin(\theta_{j})]^{j-1}\frac{\partial}{\partial\theta_{j}}\bigg{]}\\ &=-\sum_{j=1}^{n-1}\bigg{(}\prod_{k=1}^{j-1}[\sin(\theta_{n-k})]^{-2}\bigg{)}[\sin(\theta_{n-j})]^{1-j-n}\\ &\hskip 39.83368pt\times\frac{\partial}{\partial\theta_{n-j}}\bigg{[}[\sin(\theta_{n-j})]^{n-j-1}\frac{\partial}{\partial\theta_{n-j}}\bigg{]}.\end{split} (A.16)

Acknowledgments. We are indebted to Mark Ashbaugh, Andrei Martínez-Finkel- shtein, and Gerald Teschl for very helpful comments. We are also very grateful to the anonymous referee for constructive criticism.

References

  • [1] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions, Dover, New York, 1972.
  • [2] R. F. Alvarez-Estrada and A. Galindo, Bound states in some Coulomb systems, Nuovo Cim. 44 B, 47–66 (1978).
  • [3] W. O. Amrein, Non-Relativistic Quantum Dynamics, Volume 2, Reidel, Dordrecht, 1980.
  • [4] W. Arendt, G. R. Goldstein, J. A. Goldstein, Outgrowths of Hardy’s inequality, in Recent Advances in Differential Equations and Mathematical Physics, N. Chernov, Y. Karpeshina, I. W. Knowles, R. T. Lewis, and R. Weikard (eds.), Contemp. Math. 412, 51–68, 2006.
  • [5] K. Atkinson and W. Han, Spherical Harmonics and Approximations on the Unit Sphere: An Introduction, Lecture Notes in Math., Vol. 2044, Springer, 2012.
  • [6] A. A. Balinsky, W. D. Evans, and R. T. Lewis, The Analysis and Geometry of Hardy’s Inequality, Universitext, Springer, 2015.
  • [7] J. A. Barceló, T. Luque, and S. Pérez-Esteva, Characterization of Sobolev spaces on the sphere, J. Math. Anal. Appl. 491, 124240 (2020).
  • [8] J. Behrndt, S. Hassi, and H. De Snoo, Boundary Value Problems, Weyl Functions, and Differential Operators, Monographs in Math., Vol. 108, Birkhäuser, Springer, 2020.
  • [9] L. E. Blumenson, A derivation of nn-dimensional spherical coordinates, Amer. Math. Monthly 67, 63–66 (1960).
  • [10] L. E. Blumenson, On the eigenvalues of Hill’s equation, Commun. Pure Appl. Math. 16, 261–266 (1963).
  • [11] R. Bosi, J. Dolbeault, and M. J. Esteban, Estimates for the optimal constants in multipolar Hardy inequalities for Schrödinger and Dirac operators, Commun. Pure Appl. Anal. 7, 533–562 (2008).
  • [12] W. B. Brown and R. E. Roberts, On the critical binding of an electron by an electric dipole, J. Chem. Physics 46, 2006–2007 (1967).
  • [13] F. Catrina and Z.-Q. Wang, On the Caffarelli–Kohn–Nirenberg inequalities: Sharp constants, existence (and nonexistence), and symmetry of extremal functions, Commun. Pure Appl. Math. 54, 229–258 (2001).
  • [14] C. Cazacu, New estimates for the Hardy constants of multipolar Schrödinger operators, Commun. Contemp. Math. 18, no. 5, 1550093 (2016), 28pp.
  • [15] K. S. Chou and C. W. Chu, On the best constant for a weighted Sobolev–Hardy inequality, J. London Math. Soc. 48, 137–151 (1993).
  • [16] K. Connolly and D. J. Griffiths, Critical dipoles in one, two, and three dimensions, Am. J. Phys. 75, 524–631 (2007).
  • [17] O. H. Crawford, Bound states of a charged particle in a dipole field, Proc. Phys. Soc. 91, 279–284 (1967).
  • [18] H. L. Cycon, R. G. Froese, W. Kirsch, and B. Simon, Schrödinger Operators with Applications to Quantum Mechanics and Global Geometry, Texts and Monographs in Physics, Springer, Berlin, 1987.
  • [19] F. Dai and Y. Xu, Approximation Theory and Harmonic Analysis on Spheres and Balls, Springer, New York, 2013.
  • [20] E. B. Davies, Heat Kernels and Spectral Theory, Cambridge Tracts in Math., Vol. 92, Cambridge Univ. Press, Cambridge, 1989.
  • [21] E. B. Davies, Spectral Theory and Differential Operators, Cambridge University Press, Cambridge, 1995.
  • [22] T. Duyckaerts, Inégalités de résolvante pour l’opérateur de Schrödinger avec potentiel multipolaire critique, Bull. Soc. Math. France 134, 201–239 (2006).
  • [23] D. E. Edmunds and W. D. Evans, Spectral Theory and Differential Operators, 2nd ed., Oxford Math. Monographs, Oxford University Press, Oxford, 2018.
  • [24] W. D. Evans and R. T. Lewis, On the Rellich inequality with magnetic potentials, Math. Z. 251, 267–284 (2005).
  • [25] W. N. Everitt and H. Kalf, The Bessel differential equation and the Hankel transform, J. Comp. Appl. Math. 208, 3–19 (2007).
  • [26] W. G. Faris, Self-Adjoint Operators, Lecture Notes in Math., Vol. 433, Springer, Berlin, 1975.
  • [27] W. G. Faris, Inequalities and uncertainty principles, J. Math. Phys. 19, 461–466 (1978).
  • [28] V. Felli, E. M. Marchini, and S. Terracini, On Schrödinger opewrators with multipolar inverse square potentials, J. Funct. Anal. 250, 265–316 (2007).
  • [29] V. Felli, E. M. Marchini, and S. Terracini, On the behavior of solutions to Schrödinger equations with dipole type potentials near the singularity, Discrete Cont. Dyn. Syst. 21, 91–119 (2008).
  • [30] V. Felli, E. M. Marchini, and S. Terracini, On Schrödinger operators with multisingular inverse-square anisotropic potentials, Indiana Univ. Math. J. 58, 617–676 (2009).
  • [31] V. Felli, D. Mukherjee, and R. Ognibene, On fractional multi-singular Schrödinger operators: Positivity and localization of binding, J. Funct. Anal. 278, 108389 (2020).
  • [32] E. Fermi and E. Teller, The capture of negative mesotrons in matter, Phys. Rev. 72, 406 (1947).
  • [33] F. R. Gantmacher and M. G. Krein, Oscillation Matrices and Kernels and Small Vibrations of Mechanical Systems, rev. ed., AMS Chelsea Publ., Amer. Math. Soc., Providence, RI, 2002.
  • [34] F. Gesztesy, On non-degenerate ground states for Schrödinger operators, Rep. Math. Phys. 20, 93–109 (1984).
  • [35] F. Gesztesy, G. M. Graf, and B. Simon, The ground state energy of Schrödinger operators, Commun. Math. Phys. 150, 375–384 (1992).
  • [36] F. Gesztesy and W. Kirsch, One-dimensional Schrödinger operators with interactions singular on a discrete set, J. reine angew. Math. 362, 28–50 (1985).
  • [37] F. Gesztesy, L. L. Littlejohn, and R. Nichols, On self-adjoint boundary conditions for singular Sturm–Liouville operators bounded from below, J. Diff. Eq. 269, 6448–6491 (2020).
  • [38] F. Gesztesy, M. Mitrea, I. Nenciu, and G. Teschl, Decoupling of deficiency indices and applications to Schödinger-type operators with possibly strongly singular potentials, Adv. Math. 301, 1022–1061 (2016).
  • [39] F. Gesztesy, M. M. H. Pang, and J. Stanfill, Bessel-type operators and a refinement of Hardy’s inequality, in From Operator Theory to Orthogonal Polynomials, Combinatorics, and Number Theory. A Festschrift in honor of Lance L. Littlejohn’s 70th birthday, F. Gesztesy and A. Martinez-Finkelshtein (eds.), Operator Theory: Advances and Applications, Birkhäuser, Springer, to appear, arXiv:2102.00106.
  • [40] F. Gesztesy and L. Pittner, On the Friedrichs extension of ordinary differential operators with strongly singular potentials, Acta Phys. Austriaca 51, 259–268 (1979).
  • [41] F. Gesztesy and L. Pittner, A generalization of the virial theorem for strongly singular potentials, Rep. Math. Phys. 18, 149–162 (1980).
  • [42] F. Gesztesy and M. Zinchenko, On spectral theory for Schrödinger operators with strongly singular potentials, Math. Nachr. 279, 1041–1082 (2006).
  • [43] N. Ghoussoub and A. Moradifam, On the best possible remaining term in the Hardy inequality, Proc. Nat. Acad. Sci. 105, no. 37, 13746–13751 (2008).
  • [44] N. Ghoussoub and A. Moradifam, Bessel pairs and optimal Hardy and Hardy–Rellich inequalities, Math. Ann. 349, 1–57 (2011).
  • [45] N. Ghoussoub and A. Moradifam, Functional Inequalities: New Perspectives and New Applications, Amer. Math. Soc., Providence, RI, 2013.
  • [46] I. S. Gradshteyn and I. M. Rhyzhik, Table of Integrals, Series, and Products, Academic Press, San Diego, 1980.
  • [47] L. Hermi, On the Spectrum of the Dirichlet Laplacian and Other Elliptic Operators, Ph.D. Thesis, University of Missouri, Columbia, 1999.
  • [48] W. Hunziker and C. Günther, Bound states in dipole fields and continuity properties of electronic spectra, Helv. Phys. Acta 53, 201–208 (1980).
  • [49] R. S. Ismagilov, Conditions for the semiboundedness and discreteness of the spectrum for one-dimensional differential equations, Sov. Math. Dokl. 2, 1137–1140 (1961).
  • [50] H. Kalf, On the characterization of the Friedrichs extension of ordinary or elliptic differential operators with a strongly singular potential, J. Funct. Anal. 10, 230–250 (1972).
  • [51] H. Kalf, A characterization of the Friedrichs extension of Sturm–Liouville operators, J. London Math. Soc. (2) 17, 511–521 (1978).
  • [52] H. Kalf, Gauss’ theorem and the self-adjointness of Schrödinger operators, Arkiv Mat. 18, 19–47 (1980).
  • [53] H. Kalf, A note on the domain characterization of certain Schrödinger operators with strongly singular potentials, Proc. Roy. Soc. Edinburgh 97A, 125–130 (1984).
  • [54] H. Kalf and J. Walter, Strongly singular potentials and essential self-adjointness of singular elliptic operators in C0(n\{0})C_{0}^{\infty}({\mathbb{R}}^{n}\backslash\{0\}), J. Funct. Anal. 10, 114–130 (1972).
  • [55] T. Kato, Note on the least eigenvalue of the Hill equation, Quart. Appl. Math. 10, 292–294 (1952).
  • [56] T. Kato, Perturbation Theory for Linear Operators, Reprint of the 1980 edition, Classics in Mathematics, Springer, Berlin, 1995.
  • [57] W. Kirsch, Über Spektren stochastischer Schrödingeroperatoren, Ph.D. thesis, Ruhr-Universität Bochum, 1981.
  • [58] A. Kufner, L. Maligranda, and L.-E. Persson, The Hardy Inequality. About its History and Some Related Results, Vydavatelský Servis, Pilsen, 2007.
  • [59] A. Kufner, L.-E. Persson, and N. Samko, Weighted Inequalities of Hardy Type, 2nd ed., World Scientific, Singapore, 2017.
  • [60] J. Lévy-Leblond, Electron capture by polar molecules, Phy. Rev. 153, 1–4 (1967).
  • [61] E. H. Lieb and M. Loss, Analysis, 2nd ed., Graduate Studies in Math., Vol. 14, Ameri. Math. Soc., Providence, RI, 2001.
  • [62] M. Lucia and S. Prashanth, Criticality theory for Schrödinger operators with sungular potential, J. Diff. Eq. 265, 3400–3440 (2018); Addendum, 269, 7211–7213 (2020).
  • [63] J. Meixner and F. W. Schäfke, Mathieusche Funktionen und Sphäroidfunktionen. Mit Anwendungen auf physikalische und technische Probleme, Springer, Berlin, 1954.
  • [64] R. A. Moore, The least eigenvalue of Hill’s equation, J. d’Analyse Math. 5, 183–196 (1956/57).
  • [65] J. D. Morgan, Schrödinger operators whose potentials have separated singularities, J. Operator Th. 1, 109–115 (1979).
  • [66] C. Müller, Spherical Harmonics, Lecture Notes in Math., Vol. 17, Springer, Berlin, 1966.
  • [67] B. Opic and A. Kufner, Hardy-Type Inequalities, Pitman Research Notes in Mathematics Series, Vol. 219. Longman Scientific & Technical, Harlow, 1990.
  • [68] Yu. N. Ovchinnikov and I. M. Sigal, Number of bound states of three-body systems and Efimov’s effect, Ann. Phys. 123, 274–295 (1979).
  • [69] C. R. Putnam, On the least eigenvalue of Hill’s equation, Quart. Appl. Math. 9, 310–314 (1951).
  • [70] S. Rademacher and H. Siedentop, Accumulation rate of bound states of dipoles in graphene, J. Math. Phys. 57, 042105 (2016).
  • [71] M. Reed and B. Simon, Methods of Mathematical Physics. I: Functional Analysis. Revised and Enlarged Edition, Academic Press, New York, 1980.
  • [72] M. Reed and B. Simon, Methods of Mathematical Physics. II: Fourier Analysis, Self-Adjointness, Academic Press, New York, 1975.
  • [73] M. Reed and B. Simon, Methods of Mathematical Physics. IV: Analysis of Operators, Academic Press, New York, 1978.
  • [74] M. Ruzhansky and D. Suragan, Hardy Inequalities on Homogeneous Groups. 100 Years of Hardy Inequalities, Progress in Math., Vol. 327, Birkhäuser, Springer, Cham, 2019.
  • [75] K. Schmüdgen, Unbounded Self-adjoint Operators on Hilbert Space, Graduate Texts in Math., Vol. 265, Springer, Dordrecht, 2012
  • [76] B. Simon, Operator Theory, A Comprehensive Course in Analysis, Part 4, Amer. Math. Soc., Providence, R.I., 2015.
  • [77] B. Simon, Hardy and Rellich inequalities in non-integral dimension, J. Operator Th. 9, 143–146 (1983). Addendum, J. Operator Th. 12, 197 (1984).
  • [78] S. Stanek, A note on the oscillation of solutions of the differential equation y′′=λq(t)yy^{\prime\prime}=\lambda q(t)y with a periodic coefficient, Czech. math. J. 29, 318–323 (1979).
  • [79] G. Talenti, Best constant in Sobolev inequality, Ann. Mat. Pura Appl. (4) 110, 353–372 (1976).
  • [80] S. Terracini, On positive entire solutions to a class of equations with a singular coefficient and critical exponent, Adv. Diff. Eq. 1, 241–264 (1996).
  • [81] G. Teschl, Jacobi Operators and Completely Integrable Nonlinear Lattices, AMS, Providence, 1999.
  • [82] W. Thirring, A Course in Mathematical Physics, 3. Quantum Mechanics of Atoms and Molecules, transl. by E. M. Harrell, Springer, New York, 1981.
  • [83] J. E. Turner, Minimum dipole moment required to bind an electron, Amer. J. Phy. 45, 758–766 (1977).
  • [84] J. E. Turner and K. Fox, Minimum dipole moment required to bind an electron to a finite dipole, Phy. Letters 23, 547–549 (1966).
  • [85] P. Ungar, Stable Hill equation, Commun. Pure Appl. Math. 14, 707–710 (1961).
  • [86] W. Van Assche, Compact Jacobi matrices: from Stieltjes to Krein and M(a,b)M(a,b), Ann. Fac. Sci. Toulouse S5, 195–215 (1996).
  • [87] J. L. Vazquez and E. Zuazua, The Hardy inequality and the asymptotic behavior of the heat equation with an inverse-square potential, J. Funct. Anal. 173, 103–153 (2000).
  • [88] J. Weidmann, Linear Operators in Hilbert Spaces, Graduate Texts in Mathematics, Vol. 68, Springer, New York, 1980.
  • [89] J. Weidmann, Lineare Operatoren in Hilberträumen, Teubner, Stuttgart, 2000.
  • [90] A. Wintner, On the non-existence of conjugate points, Amer. J. Math. 73, 368–380 (1951).