On Critical Dipoles in Dimensions
Abstract.
We reconsider generalizations of Hardy’s inequality corresponding to the case of (point) dipole potentials , , , , , , . More precisely, for , we provide an alternative proof of the existence of a critical dipole coupling constant , such that
for all , and all , , | ||
with denoting the completion of with respect to the norm induced by the gradient. Here is sharp, that is, the largest possible such constant. Moreover, we discuss upper and lower bounds for and develop a numerical scheme for approximating .
This quadratic form inequality will be a consequence of the fact
in (with the operator closure of the linear operator ).
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, 47F05.1. Introduction
The celebrated (multi-dimensional) Hardy inequality,
(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 , with potential
(1.2) |
(with denoting the Euclidean scalar product of ), in the physically relevant case , 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 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 point charges and applications to electronic spectra of an -electron Hamiltonian in the presence of point charges (nuclei). In addition, Felli, Marchini and Terracini [29], [30] discuss more general operators where the point dipole potential in (1.2) is replaced by111For simplicity of notation, we will omit the standard surface measure in , and similarly, the Lebesgue measure in .
(1.3) |
and hence (1.2) represents the special case
(1.4) |
These authors also provide a discussion of strict positivity of the underlying quadratic form , , in the multi-center case,
(1.5) |
and its analog with replaced by restricted to suitable neighborhoods of , . 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 , there exists a critical dipole coupling constant , such that
(1.6) |
Here is optimal, that is, the largest possible such constant, and we recall that denotes the completion of with respect to the norm , .
The critical constant can be characterized by the Rayleigh quotient
(1.7) | ||||
(1.8) |
see [29]. To obtain (1.8) one introduces polar coordinates,
(1.9) |
as in (A.1)–(A.4). By (LABEL:1.1), clearly,
(1.10) |
The existence of as a finite positive number is shown for in [48] and for , , in [29].
Next, one decomposes
(1.11) | ||||
acting in , where
(1.12) |
with the Laplace–Beltrami operator in (see Appendix A). It is shown in [29] that is also characterized by
(1.13) |
where denotes the lowest eigenvalue of . We rederive (1.13) via ODE methods and also prove in Theorem 3.1, that
(1.14) |
(this extends the result in [48] to , ) and also provide the two-sided bounds
(1.15) |
with the regular modified Bessel function of order . (The additional lower bound is of course evident since .) Moreover, employing the fact that
(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 such that
(1.17) |
and the two-sided bounds
(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 , the self-adjoint realization of (i.e., the Friedrichs extension of ) in , and hence together with Appendix A on spherical harmonics and the Laplace–Beltrami operator in , , 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 that exhibits as the smallest (negative) eigenvalue of a particular triangular operator in with vanishing diagonal elements (cf. (4.20), (4.25)). In addition, we prove that finite truncations of yield a convergent and efficient approximation scheme for . Finally, Section 5 considers the extension to multicenter dipole Hamiltonians of the form
(1.19) |
where is an index set, , denotes the characteristic function of the open ball with center and radius , , , , and
(1.20) |
In particular, is permitted to be an infinite, discrete set, for instance, a lattice. Relying on results proven in [38], we derive the optimal result that is bounded from below (resp., essentially self-adjoint) if each individual , , is bounded from below (resp., essentially self-adjoint). This extends results in [30], where is assumed to be finite.
2. The Dipole Hamiltonian
In this section we provide a discussion of the angular momentum decomposition of the -dimensional Laplacian , introduce the dipole Hamiltonian , the principal object of this paper, and discuss an analogous decomposition of the latter.
In spherical coordinates (A.1), the Laplace differential expression in dimensions takes the form
(2.1) |
where denotes the Laplace–Beltrami operator222We will call the Laplacian to guarantee nonnegativity of the underlying -realization (and analogously for the -realization of the Laplace–Beltrami operator ). associated with the -dimensional unit sphere in , see (A.16). When acting in , which in spherical coordinates can be written as , (2.1) becomes
(2.2) |
(with denoting the identity operator in ). The Laplace–Beltrami operator in , with domain (cf., e.g., [7]), is known to be essentially self-adjoint and nonnegative on (cf. [20, Theorem 5.2.3]). Recalling the treatment in [72, p. 160–161], one decomposes the space into an infinite orthogonal sum, yielding
(2.3) |
where is the eigenspace of corresponding to the eigenvalue , , as
(2.4) |
In particular, this results in
(2.5) |
in the space (2.3).
To simplify matters, replacing the measure by and simultaneously removing the term , one introduces the unitary operator
(2.6) |
under which (2.5) becomes
(2.7) |
acting in the space (2.3). The precise self-adjoint -realization of in the space (2.3) then is of the form
(2.8) |
where , , represents the Friedrichs extension of
(2.9) |
in . For explicit operator domains and boundary conditions (the latter for only) we refer to (2.34)–(2.37). It is well-known (cf. [72, Sect. IX.7, Appendix to X.1]) that
(2.10) | |||
(2.11) | |||
(2.12) |
Next, we turn to the dipole potential
(2.13) |
where is a unit vector in the direction of the dipole, the strength of the dipole equals , and represents the Euclidean scalar product in . Upon an appropriate rotation, one can always choose the coordinate system in such a manner that , implying
(2.14) |
In the following we primarily restrict ourselves to the case and comment on the exceptional case at the end of Section 3. The differential expression associated with Hamiltonian for this system then becomes
(2.15) |
acting in . In analogy to (2.2), (2.15) can be represented as
(2.16) |
acting in , where
(2.17) |
is self-adjoint in (since is a bounded self-adjoint operator in ). Applying the angular momentum decomposition to , but this time with respect to the eigenspaces of , then results in
(2.18) |
where represents the eigenspace of corresponding to the eigenvalue , as
(2.19) |
We will order the eigenvalues of according to magnitude, that is,
(2.20) |
repeating them according to their multiplicity. The analog of (2.7) in the space (2.18) then becomes
(2.21) |
Remark 2.1.
In order to deal exclusively with operators which are bounded from below we now make the the following assumption.
Hypothesis 2.2.
Suppose that , , and are such that
(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
(2.24) |
In fact, “” in (2.24) can be replaced by “bounded from below”. Assumption (2.23) is equivalent to
(2.25) |
Remark 2.3.
Since the perturbation , , of in (2.17) is bounded from below and from above,
(2.26) |
and , it is clear that
(2.27) |
and . In particular, for and sufficiently small, Hypothesis 2.2 will be satisfied. We are particularly interested in the existence of a critical such that
(2.28) |
and whether or not
(2.29) |
for a , with
(2.30) |
for a , etc. This will be clarified in the next section (demonstrating that ).
Given Hypothesis 2.2, the precise self-adjoint -realization of in the space (2.18) is then of the form
(2.31) |
where , , represents the Friedrichs extension of
(2.32) |
in . Explicitly, as discussed, for instance, in [37], [40], the Friedrichs extension of , , can be determined from the fact that the Friedrichs extension in of
(2.33) |
is given by
(2.34) | |||
(2.35) | |||
(2.36) | |||
where
(2.37) |
Next we note the following fact.
Lemma 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
(2.39) |
Then has purely absolutely continuous spectrum,
(2.40) |
Proof.
First, one notes that is bounded from below if and only if each , , is bounded from below. The ordinary differential operators , , are well-known to have purely absolutely continuous spectrum equal to , 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 cause 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 (sometimes also denoted , or ) obtained as the closure of with respect to the gradient norm,
(3.1) |
see also [61, pp. 201–204].
Theorem 3.1.
Assume Hypothesis 2.2. Then for all , there exists a unique critical dipole moment characterized by
(3.2) |
cf. (2.28) in Remark 2.3. Moreover, is strictly monotonically decreasing with respect to , , and
(3.3) |
Moreover,
(3.4) |
hold. In particular, is bounded from below, and then , if and only if . Consequently,
(3.5) |
The constant in (3.5) is optimal i.e., the largest possible , in addition,
(3.6) |
Finally,
(3.7) |
Proof.
Existence of some critical dipole moment is clear from the discussion in Remark 2.3. To prove the remaining claims regarding in Theorem 3.1, we seek spherical harmonics dependent only on the final angle , as this is the only angular variable dependence of . From (A.10)–(A.13), one infers these are precisely the ones indexed by the particular multi-indices , that is (cf. (A.10)–(A.14)),
(3.8) |
Introducing the subspace
(3.9) |
and restricting the Laplace–Beltrami differential expression (A.16) to , one finds for (2.17),
(3.10) |
acting on functions in . Reverting from the weighted measure to Lebesgue measure on in a unitary fashion then yields the differential expression given by
(3.11) |
now acting on functions in .
Next, introducing the change of variable , in (3.10) turns into
(3.12) |
acting on functions in . We also note that reverting from the weighted measure to Lebesgue measure on in a unitary fashion then finally yields the differential expression given by
(3.13) |
acting on functions in . One observes that the first two terms on the right-hand side of (3.13) represent the Legendre operator in associated with the differential expression
(3.14) |
which is in the limit circle case at if and in the limit point case at if , as discussed in detail in [24]. In particular, applying this fact to , , , and yields the necessity of the Friedrichs boundary condition for , whereas for , , and (resp., and ) are essentially self-adjoint on (resp., ) 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 (resp., ) singularity at (resp., ), the Friedrichs extension corresponding to in (3.11) is clear from (2.33)–(2.37).
Following [48] in the special case , choosing normalized,
(3.15) |
an appropriate integration by parts yields
(3.16) | |||
(3.17) | |||
(3.18) |
In particular, choosing for a normalized eigenfunction of corresponding to the eigenvalue in (3.16) implies the lower bound
(3.19) |
On the other hand (following once more [48] in the special case ), employing the normalized trial function (cf. [46, no. 3.387])
(3.20) | |||
with the regular modified Bessel function of order (cf. [1, Sect. 9.6]), an application of the min/max principle and (3.17) yield the upper bound
(3.21) |
employing [46, no. 3.387] once again. Thus, (3.21) implies that
(3.22) |
and one infers a quadratic upper bound as
(3.23) |
in addition to the quadratic lower bound in (3.19).
Next, recalling that the lowest eigenvalue of is simple for all (and is also the lowest eigenvalue of , , and ), we denote by the corresponding normalized eigenfunction, that is,
(3.24) |
Thus, one gets
(3.25) |
Moreover, one observes that is a self-adjoint analytic (in fact, entire) family of type in the sense of Kato (cf. [56, Sect. VII.2, p. 375–379], [73, p. 16]), implying analyticity of and with respect to in a complex neighborhood of . In particular, is differentiable with respect to , and the Feynman–Hellmann Theorem [82, p. 151] (see also [76, Theorem 1.4.7]) yields that
(3.26) |
Returning to the discussion of (2.27) in Remark 2.3, employing , one obtains
(3.27) |
implying,
(3.28) |
by the strict negativity of for derived in (3.21).
Given the existence of a unique critical dipole moment one concludes from (2.24), (2.31), and (2.32) the following fact:
(3.29) |
and an integration by parts thus yields
(3.30) |
It remains to extend (3.30) to elements . As in the case of the Hardy inequality (LABEL:1.1), this follows from invoking a Fatou-type argument to be outlined next.
Since is dense in , given we pick a sequence such that , and, by passing to a subsequence, we may assume without loss of generality (see (3.31) below) that a.e. on . (For the remainder of this proof , , will always be assumed to have the properties just discussed.) Indeed, the Sobolev inequality (see, e.g., [61, Theorem 8.3], [79]),
(3.31) |
( the Gamma function, cf. [1, Sect. 6.1]), yields convergence of to in and hence permits the selection of a subsequence that converges pointwise a.e. Thus, given Hardy’s inequality for functions in , a well-known fact (see, e.g., [6, Corollary 1.2.6]),
(3.32) |
one obtains,
(3.33) |
for some independent of . Thus,
(3.34) |
by a consequence of Fatou’s Lemma (see, e.g., [61, p. 21]. Hence,
(3.35) |
extends Hardy’s inequality (3.32) from to . Hardy’s inequality on also implies that
(3.36) |
in particular,
(3.37) |
Since
(3.38) |
(3.34) also implies
(3.39) |
similarly, (3.36), (3.37), and Hölder’s inequality imply
(3.40) |
Thus, for ,
(3.41) |
finally implying (3.5). Moreover, (3.38) also yields
(3.42) | ||||
and hence (3.6).
Remark 3.2.
Next, we improve upon Remark 3.2 for as follows:
Theorem 3.3.
Proof.
Employing [29, eq. (1) and Remark 1], one considers the Rayleigh quotient
(3.45) |
and notes that as , implying (cf. (3.11))
(3.46) |
Employing the fact that
(3.47) |
(this follows from [36, Sect. 4] for and extends to utilizing [37, Subsect. 6.1]) one concludes the following variant of Hardy’s inequality (upon taking ) with optimal constants ,
(3.48) |
which, by a density argument, extends to
(3.49) |
(see [39]). Thus, employing (3.49) in (3.46) yields
(3.50) | ||||
(3.51) |
Here we used the estimate,
(3.52) |
and the fact that due to the sign change of as crosses , the numerator in (3.50) diminishes and the denominator in (3.50) increases, altogether diminishing the ratio in (3.50) if has support in . Thus, one is justified assuming that has support in only.
In the case , the factor in ](3.50) vanishes, and hence we now employ the additional term in (3.49) to arrive at
(3.53) |
Altogether, this implies the lower bound in (3.44) and hence improves on Remark 3.2 for . (For one can include the term to improve the lower bound, but the actual details become so unwieldy that we refrain from doing so.) For we just recalled (3.6).
Remark 3.4.
Since embeds compactly into , the supremum in (3.46) (unlike that in (LABEL:3.40a)) is actually attained, that is, for a particular ,
(3.60) |
However, since the -dependence of appears to be beyond our control, computing the exact value of in (3.43) remains elusive.
The differential equation underlying (3.46) is of the type
(3.61) |
which naturally leads to the Birman–Schwinger-type eigenvalue problem
(3.62) |
where denotes the Friedrichs extension of the preminimal operator in defined by
(3.63) |
One observes that is essentially self-adjoint for and hence boundary conditions at , familiar for singular second-order differential operators of Bessel-type (see [37, Subsection 6.1]), are only required for . The Birman–Schwinger operator
(3.64) |
in is compact (in fact, Hilbert–Schmidt) upon inspecting its integral kernel and hence by the Raleigh–Ritz quotient in (3.46), is the largest eigenvalue for . Finally, introducing the unitary operator
(3.65) |
in , one verifies that
(3.66) |
and hence the spectrum of is symmetric with respect to the origin.
It is well-known that Hardy’s inequality (LABEL:1.1) is strict, that is, equality holds in (LABEL:1.1) for some if and only if . 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 employing a quadratic form approach.
To set the stage we briefly recall a few facts on quadratic forms generated by symmetric operators bounded from below and the associated Friedrichs extension (to be denoted by ) of .
Let be a densely defined symmetric operator in the Hilbert space bounded from below, that is, and for some , . Without loss of generality we put in the following. We denote by the closure of in , and introduce the associated forms in ,
(3.67) | |||
(3.68) |
then the closures of and coincide in (cf., e.g., [8, Lemma 5.1.12])
(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
(3.70) |
where represents the self-adjoint Friedrichs extension of . Due to the fact (3.69), one infers (cf., e.g., [8, Lemma 5.3.1])
(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
(3.72) |
Moreover, one has the fact (see, e.g., [8, Theorem 5.3.3], [23, Corollary 4.2.7], [75, Theorem 10.17])
(3.73) |
Theorem 3.5.
Proof.
We first discuss the simpler case . In this case the inequality (3.5) implies that the sesquilinear form , , , where
(3.74) |
is bounded relative to the form of the Laplacian on ,
(3.75) |
with relative bound strictly less than one. Hence the form
(3.76) |
is densely defined, nonnegative, and closed. Moreover, since is dense in for (cf., e.g., [26, p. 33–35]), that is, is a core for (equivalently, a form core for ), and hence also a core for ,
(3.77) |
Thus, the self-adjoint, nonnegative operator , uniquely associated with by the first representation theorem for forms coincides with the Friedrichs extension of the minimal operator associated with the differential expression in (2.15),
(3.78) |
that is,
(3.79) |
In turn, since coincides with the direct sum of Friedrichs extensions in (2.31), one concludes that coincides with , and hence,
(3.80) |
in particular,
(3.81) |
Thus, equality in the inequality (3.5) for some implies
(3.82) | ||||
and hence,
(3.83) |
by (3.7).
The case follows analogous lines but is a bit more involved as now the form is bounded relative to the form with relative bound equal to one.
Since by inequality (3.5)
(3.84) |
the form
(3.85) | ||||
is closable and we denote its closure in by . Thus, and hence the self-adjoint, nonnegative operator uniquely associated with in is the Friedrichs extension of ,
(3.86) |
Again, by the second representation theorem
(3.87) |
However, unlike in the case , since the form is bounded relative to the form with relative bound equal to one, one now has possible cancellations between the forms and and hence concludes that
(3.88) |
(see also Remark 3.6). The rest of the proof now follows the case line by line, in particular,
(3.89) |
and
(3.90) |
again by (3.7), then proves strictness of (3.5) also for . ∎
Remark 3.6.
We briefly illustrate the possibility of cancellations in . Let , , be the unique (up to constant multiples) eigenfunction of corresponding to its lowest eigenvalue , see (3.24), and introduce
(3.91) |
Then an elementary computation reveals that
(3.92) |
in the sense of distributions. In particular, if , and hence, , one obtains
(3.93) |
in the distributional sense. Thus, introducing
(3.94) |
one concludes that
(3.95) |
However, since , the elementary fact
(3.96) |
implies that
(3.97) |
illustrating possible cancellations between and .
Remark 3.7.
Next, we briefly discuss the remaining case . In this situation, the Laplace–Beltrami operator in can be characterized by
(3.98) | |||
with
(3.99) | |||
(3.100) |
The resulting Mathieu operator in (cf. (2.17)), of the form
(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 is replaced by a -periodic, locally integrable potential ) has received enormous attention, see for instance, [10], [35], [55], [64], [69], [70], [78], [85], and [90]. Applied to the Mathieu operator at hand, the results obtained (cf. the discussion in [35]) imply,
(3.102) |
In particular, this proves the absence of a critical coupling constant for (equivalently, the critical constant in two dimensions equals zero, ), explaining why we had to limit ourselves to in the bulk of this paper.
Remark 3.8.
While thus far we focused primarily on lower semiboundedness of , the direct sum considerations in Section 2 equally apply to essential self-adjointness of . Indeed, returning to the operator in (2.24), one notes that
(3.103) |
The criterion (3.103) combined with (2.31), (2.32) thus implies that
(3.104) |
4. A Numerical Approach
Having verified the existence and uniqueness of critical dipole moments for all dimensions , and having shown some of the properties of , this section is devoted to a description of a numerical method for computing , 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 ranges from to , observing that (cf. (2.28)). Restricting in (A.16) to as in (LABEL:3.3)–(3.10), (3.24) reduces to solving the eigenvalue problem associated with in of the type,
(4.1) |
Expanding in normalized Gegenbauer polynomials, one obtains
(4.2) |
where are appropriate expansion coefficients. Since the Gegenbauer polynomial is an eigenfunction of corresponding to the eigenvalue , (4.1) becomes
(4.3) |
Next, we will exploit the following recurrence relation of Gegenbauer polynomials,
(4.4) |
(with ) to expand the term . For the -term, one infers
(4.5) |
(where ) and for the -term, one obtains
(4.6) |
The -term maintains its form
(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
(4.8) | |||
Setting each coefficient equal to zero results in
(4.9) |
which one can rewrite as
(4.10) |
Equation (4.10) can be expressed as the generalized Jacobi operator eigenvalue problem in ,
(4.11) |
where
(4.12) |
and
(4.13) |
(One observes that by (2.38).)
Explicitly, (4.12) yields the self-adjoint Jacobi operator in represented as a semi-infinite matrix with respect to the standard Kronecker- basis
(4.14) |
One would like to calculate approximately using finite truncations of the matrix representation of 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
(4.15) |
That is unitary may be seen from
(4.16) |
and the fact that is surjective and defined on all of . Next, one transforms (4.14) into
(4.17) |
which can equivalently be expressed as
(4.18) |
Introducing
(4.19) |
one can write (4.18) in the form
(4.20) |
in analogy with (4.14), with
(4.21) | |||
(4.22) |
Recalling once more that (cf. (2.28)), and that as well as in (4.1) are analytic with respect to as varies in a complex neighborhood of , taking the limit on either side of (4.20) yields
(4.23) |
implying
(4.24) |
Next, we prove that is the smallest (negative) eigenvalue of the operator in .
Proposition 4.1.
One has , , and
(4.25) |
Moreover,
(4.26) |
where
(4.27) |
with , the ceiling function.
Proof.
The compactness assertion for , , follows from the limiting behavior , see, for instance, [86, p. 201].
To prove the uniform lower bound (4.25) one can argue by contradiction as follows: Fix and suppose there exists such that , that is, . Then working backwards from (4.20) to (4.1) yields the existence of such that
(4.28) |
where
(4.29) |
By the strict monotonicity of with respect to one infers that
(4.30) |
and hence (4.28) contradicts the fact that by definition, is the lowest eigenvalue of .
To obtain the bound (4.26), (4.27), one applies [81, Theorem 1.5] after calculating . As is bounded and tends to 0 as for all , it attains its supremum. To find the index where this occurs, one considers as a continuous variable, and solves . The value emerges as the only nonnegative, real root of this expression, but as for , the maximum in (4.26) and the ceiling function in (4.27) are required. Since , the norm must be computed separately as . ∎
To introduce the notion of finite truncations, one considers the operators
(4.31) |
on (with denoting the identity matrix in , ).
We also introduce the finite tri-diagonal Jacobi matrices in , , , denoted by
(4.32) |
in particular,
(4.33) |
where and are monic polynomials of degree and , respectively.
Thus, the spectrum of each consists of real eigenvalues, symmetric with respect to the origin, the eigenvalues being simple as long as , (see, e.g., [33, Theorem II.1.1], [81, Remark 1.10 and p. 120]). Explicitly,
(4.34) | |||
etc. |
In addition, we introduce the unitary, self-adjoint, diagonal operator in as
(4.35) |
Theorem 4.2.
Given the operators , , , and as in (4.20), (4.31)–(4.35), one concludes that and as well as and are unitarily equivalent,
(4.36) |
and hence the spectra of and , , are symmetric with respect to zero. Moreover, all nonzero eigenvalues of and , , are simple. In addition,
(4.37) |
and333Here denotes the essential spectrum.
(4.38) |
In particular, if and only if there is a sequence with such that .
Proof.
The symmetry fact (4.36) follows from an elementary computation. That all eigenvalues of are simple follows from the fact that is a half-lattice operator with , , and hence the half-lattice does not decouple into a disjoint union of subsets (resp., does not reduce to a direct sum of operators in ). The same argument applies to the finite-lattice operators , .
One notices that , where strong operator convergence is abbreviated by . Together with the compactness of given in Proposition 4.1, one obtains
(4.39) |
applying [3, Proposition 3.11]. The norm convergence in (4.39), together with the uniform bound , , 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 ]), 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) ]). ∎
Returning to the dipole context, one may now compute approximants of by approximating the smallest negative eigenvalues of in terms of the smallest negative eigenvalue of with increasing . Using , (which produced 16 stable digits in the case ), one obtains the following values and approximants for :
lower bound for | upper bound | ||
---|---|---|---|
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 |

Here the lower and upper bounds for correspond to the values displayed in (3.44) (see Fig. 1).
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 ). 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 is self-adjoint in a complex Hilbert space and bounded from below, that is, for some . Then the sesquilinear form associated with is denoted by
(5.1) |
A sesquilinear form in satisfying is called bounded with respect to the form if for some ,
(5.2) |
The infimum of all numbers for which there exists such that (5.1) holds is called the bound of with respect to .
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 denotes an index set.
Lemma 5.2.
Suppose that is a self-adjoint operator in bounded from below, for some , and is a self-adjoint operator in such that
(5.3) |
We abbreviate
(5.4) |
Let , , assume that , , leave invariant, that is,
(5.5) |
and suppose that the following conditions – hold:
.
, .
, .
Then,
(5.6) |
implies
(5.7) |
Proof.
For one computes
(5.8) |
completing the proof. ∎
Remark 5.3.
Considering the concrete case of
(5.9) | |||
in , and assuming that , the operator of multiplication with a measurable and a.e. real-valued function , employing a slight abuse of notation, satisfies (5.3) (for sufficient conditions on , see, e.g., [88, Theorems 10.17 (b), 10.18] with ). Let , , be a family of smooth, real-valued functions defined on in such a manner that for each , there exists an open neighborhood of such that there exist only finitely many indices with and , as well as
(5.10) |
(the sum over in (5.10) being finite). Finally, let be the operator of multiplication by the function , . Then one notes that for these choices, hypothesis holds with equality, and hypothesis with follows from . Moreover, item holds with as long as
(5.11) |
To verify this, one observes that and that the cross terms vanish since , , by condition (5.10). (We note again that the latter sum over contains only finitely many terms in every bounded neighborhood of .)
Strongly singular potentials that are covered by Lemma 5.2 are, for instance, of the following form: Let be an index set, , and , , , be a set of points such that
(5.12) |
In addition, let , , , with
(5.13) |
and
(5.14) |
with
(5.15) |
and the open ball in of radius , centered at .
Then an application of Hardy’s inequality in , (cf. (LABEL:1.1)), shows that is form bounded with respect to 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.
Proof.
Without loss of generality we put . Inequality (3.5) and the analogous inequality with , , replaced by yields
(5.18) |
and hence for some and ,
(5.19) |
Thus, Theorem 5.4 proves semiboundedness of the self-adjoint multicenter dipole Hamiltonian in , uniquely associated with the quadratic form sum
(5.21) |
under very general hypotheses on and . We note that [29], [30] derive sufficient conditions and to guarantee nonnegativity of and also discuss situations characterized by the lack of nonnegativity of .
Finally, we sketch how Remark 3.8 extends to the multicenter situation.
Appendix A Spherical Harmonics and the Laplace–Beltrami Operator in , .
In this appendix we summarize some of the results on spherical harmonics and the Laplace–Beltrami operator on the unit sphere in dimensions , , following [5, Chs. 2,3], [19, Ch. 1], and [47, Ch. 2].
Assuming , , cartesian and polar coordinates (cf. e.g., [9]) on are given by
(A.1) | |||
where (cf., e.g., [9], [19, Sect. 1.5])
(A.2) |
The surface measure on and the volume element in then read
(A.3) |
in particular, the area of the unit sphere in is given by (cf. [66, p. 2])
(A.4) |
Turning to spherical harmonics next, we recall that a homogeneous polynomial of degree (in variables) satisfies and is a linear combination of terms of degree . The space of such polynomials with real coefficients is denoted . We define the harmonic homogeneous polynomials of degree in variables by
(A.5) |
where represents the Laplace differential expression on . Restricting the elements of to the sphere , one obtains , the space of spherical harmonics of degree in dimensions. Spaces of different degrees are orthogonal with respect to the real inner product on the sphere,
(A.6) |
The dimension of equals that of and is given by ([19, Corollary 1.1.4])
(A.7) |
where we use the convention that the second binomial coefficient equals 0 when , and replace the final fraction by 1 in the case where and . This is equivalently formulated in [66, Lemma 3, p. 4] as the generating series
(A.8) |
Most importantly, the spherical harmonics are the eigenfunctions of the Laplace–Beltrami operator in , satisfying the eigenvalue equation
(A.9) |
Following [19, Sect. 1.5] an explicit characterization for the spherical harmonics reads as follows: Introducing the multi-index , with , and , the spherical harmonics are of the form
(A.10) |
where
(A.11) | |||
(A.12) | |||
(A.13) |
Here the Pochhammer symbol is defined by
(A.14) |
and represent the Gegenbauer (or ultrasperical) polynomials, see, for instance, [1, Ch. 22], [19, Appendix B].
The set represents an orthonormal basis of .
Finally, we recall the expression of the Laplace–Beltrami differential expression on in spherical coordinates. From [5, p. 94], [19, Lemma 1.4.2], one obtains the recursion444For clarity we indicate the space dimension as a subscript in the Laplacian for the remainder of this appendix.
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 -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 , 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 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 , 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).