Skip to main content
Log in

Limiting Absorption Principle for Discrete Schrödinger Operators with a Wigner–von Neumann Potential and a Slowly Decaying Potential

  • Original Paper
  • Published:
Annales Henri Poincaré Aims and scope Submit manuscript

Abstract

We consider discrete Schrödinger operators on \(\mathbb {Z}^d\) for which the perturbation consists of the sum of a long-range-type potential and a Wigner–von Neumann-type potential. Still working in a framework of weighted Mourre theory, we improve the limiting absorption principle (LAP) that was obtained in Mandich (J Funct Anal 272(6):2235–2272, 2017). To our knowledge, this is a new result even in the one-dimensional case. The improvement is twofold. It weakens the assumptions on the long-range potential and provides better LAP weights. Both upgrades include logarithmic terms. The proof relies on the fact that some particular functions that contain logarithmic terms, are operator monotone. This fact is proved using Loewner’s theorem and Nevanlinna functions.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  1. Agmon, S.: Spectral properties of Schrödinger operators and scattering theory. Ann. Scuola Norm. Sup. Pisa 2, 151–218 (1975)

    MathSciNet  MATH  Google Scholar 

  2. Amrein, W.O., de Monvel, A.Boutet, Georgescu, V.: \(C_0\)-groups, Commutator Methods and Spectral Theory of \(N\)-body hamiltonians. Birkhäuser, Basel (1996)

    MATH  Google Scholar 

  3. Perry, P., Sigal, I.M., Simon, B.: Spectral analysis of \(N\)-body Schrödinger operators. Ann. of Math. 114, 519–567 (1981)

    MathSciNet  MATH  Google Scholar 

  4. Devinatz, A., Moeckel, R., Rejto, P.: A limiting absorption principle for Schrödinger operators with Von-Neumann–Wigner potentials. Int. Eq. Op. Theory 14(1), 13–68 (1991)

    MATH  Google Scholar 

  5. Rejto, P., Taboada, M.: A limiting absorption principle for Schrödinger operators with generalized Von Neumann–Wigner potentials I. Construction of approximate phase. J. Math. Anal. Appl 208, 85–108 (1997)

    MathSciNet  MATH  Google Scholar 

  6. Rejto, P., Taboada, M.: A limiting absorption principle for Schrödinger operators with generalized Von Neumann–Wigner potentials II. The proof. J. Math. Anal. Appl 208, 311–336 (1997)

    MathSciNet  MATH  Google Scholar 

  7. von Neumann, J., Wigner, E.P.: Über merkwürdige diskrete Eigenwerte. Z. Phys. 30, 465–567 (1929)

    MATH  Google Scholar 

  8. Frank, R.L., Simon, B.: Eigenvalue bounds for Schrödinger operators with complex potentials II. J. Spectr. Theory 7(3), 633–658 (2017)

    MathSciNet  MATH  Google Scholar 

  9. Golénia, S., Jecko, T.: Weighted Mourre’s commutator theory, application to Schrödinger operators with oscillating potential. J. Oper. Theory 1, 109–144 (2013)

    MATH  Google Scholar 

  10. Jecko, T.: On Schrödinger and Dirac operators with an oscillating potential. With contributions by Aiman Mbarek. Rev. Roumaine Math. Pures Appl. 64(2–3), 283–314 (2019)

  11. Jecko, T., Mbarek, A.: Limiting absorption priniciple for Schrödinger operators with oscillating potentials. Documenta Math. 22, 727–776 (2017)

    MathSciNet  MATH  Google Scholar 

  12. Mandich, M.: The limiting absorption principle for the discrete Wigner-von Neumann operator. J. Funct. Anal. 272(6), 2235–2272 (2017)

    MathSciNet  MATH  Google Scholar 

  13. Martin, A.: On the limiting absorption principle for a new class of Schrödinger Hamiltonians. Conflu Math 10(1), 63–94 (2018)

    Google Scholar 

  14. Martin, A.: A new class of Schrödinger operators without positive eigenvalues. Integr. Equ. Oper. Theory 91, 24 (2019)

    MATH  Google Scholar 

  15. Martin, A.: On the limiting absorption principle at zero energy for a class of possibly non self-adjoint Schrödinger operators (2018). arxiv:1808.07738

  16. Janas, J., Simonov, S.: Weyl-Titchmarsh type formula for discrete Schrödinger operator with Wigner-von Neumann potential. Studia Math. 201(2), 167–189 (2010)

    MathSciNet  MATH  Google Scholar 

  17. Kurasov, P., Naboko, S.: Wigner-von Neumann perturbations of a periodic potential: spectral singularities in bands. Math. Proc. Cambridge Philos. Soc. 142(1), 161–183 (2007)

    ADS  MathSciNet  MATH  Google Scholar 

  18. Kurasov, P., Simonov, S.: Weyl-Titchmarsh type formula for periodic Schrödinger operator with Wigner-von Neumann potential. Proc. Roy. Soc. Edinburgh Sect. A 143(2), 401–425 (2013)

    MathSciNet  MATH  Google Scholar 

  19. Liu, W.: Criteria for embedded eigenvalues for discrete Schrödinger Operators, International Mathematics Research Notices, rnz262, (2019). https://doi.org/10.1093/imrn/rnz262

  20. Lukic, M.: Orthogonal polynomials with recursion coefficients of generalized bounded variation. Comm. Math. Phys. 306, 485–509 (2011)

    ADS  MathSciNet  MATH  Google Scholar 

  21. Lukic, M.: Schrödinger operators with slowly decaying Wigner-von Neumann type potentials. J. Spectral Theory 3, 147–169 (2013)

    MathSciNet  MATH  Google Scholar 

  22. Lukic, M.: A class of Schrödinger operators with decaying oscillatory potentials. Comm. Math. Phys. 326, 441–458 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  23. Naboko, S., Simonov, S.: Zeroes of the spectral density of the periodic Schrödinger operator with Wigner–von Neumann potential. Math. Proc. Cambridge Philos. Soc. 153(1), 33–58 (2012)

    ADS  MathSciNet  MATH  Google Scholar 

  24. Simonov, S.: Zeroes of the spectral density of discrete Schrödinger operator with Wigner–von Neumann potential. Integral Eq. Oper. Theory 73(3), 351–364 (2012)

    MathSciNet  MATH  Google Scholar 

  25. Mourre, E.: Absence of singular continuous spectrum for certain self-adjoint operators. Comm. Math. Phys. 78, 391–408 (1981)

    ADS  MathSciNet  MATH  Google Scholar 

  26. Mourre, E.: Opérateurs conjugués et propriétés de propagation. Comm. Math. Phys. 91, 279–300 (1983)

    ADS  MathSciNet  MATH  Google Scholar 

  27. Gérard, C.: A proof of the abstract limiting absorption principle by energy estimates. J. Funct. Anal. 254(11), 2707–2724 (2008)

    MathSciNet  MATH  Google Scholar 

  28. Golénia, S., Jecko, T.: A new look at Mourre’s commutator theory. Compl. Anal. Oper. Theory 1(3), 399–422 (2007)

    MathSciNet  MATH  Google Scholar 

  29. Georgescu, V., Golénia, S.: Isometries, Fock spaces and spectral analysis of Schrödinger operators on trees. J. Funct. Anal. 227, 389–429 (2005)

    MathSciNet  MATH  Google Scholar 

  30. Remling, C.: Discrete and embedded eigenvalues for one-dimensional Schrödinger operators. Comm. Math. Phys. 271, 275–287 (2007)

    ADS  MathSciNet  MATH  Google Scholar 

  31. Mandich, M.: Sub-exponential decay of eigenfunctions for some discrete Schrödinger operators. J. Spectr. Theory 9, 21–77 (2019)

    MathSciNet  MATH  Google Scholar 

  32. Mbarek, A.: Etudes théorème d’absorption limite pour des opérateurs de Schrödinger et Dirac avec un potentiel oscillant, Phd thesis (2017). http://www.theses.fr/2017CERG0839

  33. Liu, W.: Absence of singular continuous spectrum for perturbed discrete Schrödinger operators. J. Math. Anal. Appl. 472(2), 1420–1429 (2019)

    MathSciNet  MATH  Google Scholar 

  34. Simon, B.: Bounded eigenfunctions and absolutely continuous spectra for one-dimensional Schrödinger operators. Proc. Am. Math. Soc 124(11), 3361–3369 (1996)

    MATH  Google Scholar 

  35. Boutet de Monvel, A., Sahbani, J.: On the spectral properties of discrete Schrödinger operators: the multi-dimensional case. Rev. Math. Phys. 11(9), 1061–1078 (1999)

    MathSciNet  MATH  Google Scholar 

  36. Jensen, A., Perry, P.: Commutator methods and Besov space estimates for Schrödinger operators. J. Oper. Theory 14(1), 181–188 (1985)

    MATH  Google Scholar 

  37. Agmon, S., Hörmander, L.: Asymptotic properties of solutions to differential equations with simple characteristics. J. Anal. Math. 30, 1–38 (1976)

    MathSciNet  MATH  Google Scholar 

  38. Dereziński, J., Gérard, C.: Scattering Theory of Classical and Quantum N-particle Systems. Springer-Verlag, Berlin (1997)

    MATH  Google Scholar 

  39. Heinz, E.: Beiträge zur Störungstheorie der Spektralzerlegung. Math. Ann. 123, 415–438 (1951)

    MathSciNet  MATH  Google Scholar 

  40. Kato, T.: A generalization of the Heinz inequality. Proc. Japan Acad. 37, 305–308 (1961)

    MathSciNet  MATH  Google Scholar 

  41. Reed, M., Simon, B.: Methods of Modern Mathematical Physics, Tome I: Functional Analysis. Academic Press, New York (1980). ISBN 9780125850506

    MATH  Google Scholar 

  42. Donoghue, W.: Monotone matrix functions and analytic continuation. Springer, Berlin, Heidelberg, New York (1974)

    MATH  Google Scholar 

  43. Loewner, K.: Über monotone Matrixfunktionen. Math. Z. 38, 177–216 (1934)

    MathSciNet  MATH  Google Scholar 

  44. Simon, B.: Loewner’s theorem on monotone matrix functions, Springer International Publishing, print ISBN : 978–3–030–22421–9 (2019)

  45. Hansen, F.: The fast track to Löwner’s theorem. Linear Algebra Appl. 438(11), 4557–4571 (2013)

    MathSciNet  MATH  Google Scholar 

  46. Bendat, J., Sherman, S.: Monotone and convex operator functions. Trans. Am. Math. Soc. 79(1), 58–71 (1955)

    MathSciNet  MATH  Google Scholar 

  47. Kato, T.: Perturbation theory for linear operators, Reprint of the 1980 edition. Classics in Mathematics. Springer-Verlag, Berlin, (1995). xxii+619 pp. ISBN: 3-540-58661-X

  48. Olson, M.P.: The selfadjoint operators of a von Neumann algebra form a conditionally complete lattice. Proc. Amer. Math. Soc. 28(2), 537–544 (1971)

    MathSciNet  MATH  Google Scholar 

  49. Uchiyama, M.: Commutativity of selfadjoint operators. Pacific J. Math. 161(2), 385–392 (1993)

    MathSciNet  MATH  Google Scholar 

  50. Fujii, M., Kasahara, I.: A remark on the spectral order of operators. Proc. Japan Acad. 47, 986–988 (1971)

    MathSciNet  MATH  Google Scholar 

  51. Lewin, L.: Polylogarithms and Associated Functions. Elsevier North Holland, New York (1981)

    MATH  Google Scholar 

  52. Prudnikov, A.P., Brychkov, Yu A., Marichev, O.I.: Integrals and Series, Volume I : Elementary Functions. Gordon and Breach, New York, NY (1986)

    MATH  Google Scholar 

  53. Davies, E.B.: Spectral theory and differential operators. Cambridge Studies in Adv. Math, Cambridge (1995)

    MATH  Google Scholar 

  54. Møller, J.S., Skibsted, E.: Spectral theory of time-periodic many-body systems. Adv. Math. 188(1), 137–221 (2004)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

It is a pleasure to thank Thierry Jecko for fruitful conversations on the topic, generous advice to improve the results in various ways, and especially giving us the permission to publish his Proposition 4.4 which conveniently generalizes self-adjoint operator norm estimates to submultiplicative functions. We are grateful to the two anonymous referees for a careful reading of the manuscript leading to many improvements, and especially Proposition 4.4.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sylvain Golénia.

Additional information

Communicated by Jan Derezinski.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A. Nevanlinna Functions, Operator Monotone Functions and Loewner’s Theorem

We revisit Loewner’s theorem on matrix operator monotone functions, see, e.g., [42,43,44,45]. This wonderful theorem makes a striking connection between the operator monotone functions and the Nevanlinna functions. Let \(\mathbb {C}_+\) (resp. \(\mathbb {C}_-\)) denote the complex numbers with strictly positive (resp. negative) imaginary part. A Nevanlinna function (also known as Herglotz, Pick or R function) is an analytic function on \(\mathbb {C}_+\) that maps \(\mathbb {C}_+\) to \(\overline{\mathbb {C}_+}\). A function \(f : \mathbb {C}_+ \mapsto \mathbb {C}\) is Nevanlinna if and only if it admits a representation

$$\begin{aligned} f(z) = \alpha + \beta z + \int _{\mathbb {R}} \left( \frac{1}{\lambda - z} - \frac{\lambda }{\lambda ^2+1} \right) d \mu (\lambda ), \quad z \in \mathbb {C}^+ \end{aligned}$$
(7.1)

where \(\alpha \in \mathbb {R}\), \(\beta \geqslant 0\), and \(\mu \) is a positive Borel measure on \(\mathbb {R}\) satisfying \(\int _{\mathbb {R}} (\lambda ^2+1) ^{-1} d\mu (\lambda ) < \infty \). We refer to [42, Theorem 1 of Chapter II] for a proof of this wonderful result. The integral representation is unique. The measure \(\mu \) is recovered from f by the Stieltjes inversion formula

$$\begin{aligned} \mu \left( (\lambda _1, \lambda _2] \right) = \lim \limits _{\delta \downarrow 0} \lim \limits _{\varepsilon \downarrow 0} \frac{1}{\pi } \int _{\lambda _1 + \delta } ^{\lambda _2 + \delta } \mathrm {Im} \left( f(\lambda + \mathrm {i}\varepsilon ) \right) d\lambda .\end{aligned}$$

Standard examples of Nevanlinna functions given in the literature include \(z^{p}\) for \(0 \leqslant p \leqslant 1\), \(-z^{p}\) for \(-1 \leqslant p \leqslant 0\), and the logarithm \(\ln (z)\) with the branch cut \((-\infty ,0]\).

Notation. Let (ab) be an open interval (finite or infinite). Denote P(ab) the set of Nevanlinna functions that continue analytically across (ab) into \(\mathbb {C}_-\) and where the continuation is by reflection.

Functions in P(ab) are real-valued on (ab) and their measure satisfies \(\mu \left( (a,b) \right) = 0\); see [42, Lemma 2, Chapter II]. Functions in P(ab) are strictly increasing on (ab), unless they are constant. Indeed, if f is not a constant function \(\beta \) and \(\mu \) cannot be simulaneously zero and so

$$\begin{aligned} f'(x) = \beta + \int _{\mathbb {R}} \frac{d\mu (\lambda )}{(\lambda -x)^2} > 0, \quad x \in (a,b).\end{aligned}$$

Let \(\mathcal {M}_n(\mathbb {C})\) be the set of \(n \times n\) matrices with entries in \(\mathbb {C}\), and consider a function \(f : (a,b) \mapsto \mathbb {R}\).

Definition. f is matrix monotone of order n in (ab) if \(f(T) \leqslant f(S)\) holds whenever TS in \(\mathcal {M}_n(\mathbb {C})\) are hermitian matrices with spectrum in (ab) and \(T \leqslant S\).

In 1934, Karl Loewner proved the following remarkable theorem that characterizes the matrix monotone functions:

Theorem 7.1

[43] Let \(f : (a,b) \mapsto \mathbb {R}\), where (ab) is a finite or infinite open interval. Then f is matrix operator monotone of order n in (ab) for all \(n \in \mathbb {N}\) if and only if f admits an analytic continuation that belongs to P(ab).

Loewner’s theorem is a truly wonderful result and has been reproved in several different ways, see, e.g., [42, 44, Theorem I of Chapter VII and Chapter IX for the converse] or [45] and references therein for a concise historical exposition. For the purpose of this article, we need a version of Loewner’s theorem that applies to unbounded self-adjoint operators. In Simon’s book [44, Chapter 2], it is explicitly discussed how Loewner’s theorem extends to unbounded operators. We propose below yet another proof of the extension to the semi-bounded operators (which is the case for \(\langle A \rangle \) and \(\langle N \rangle \)). For self-adjoint operators T, S which are bounded from below, the inequality \(T \leqslant S\) means \(\mathrm {Dom}[|S|^{1/2}]\subset \mathrm {Dom}[|T|^{1/2}]\) and \(\langle \psi , T \psi \rangle \leqslant \langle \psi , S \psi \rangle \) for all \(\psi \in \mathrm {Dom}[|S|^{1/2}]\).

Definition. f is operator monotone in \((a,b) \subset \mathbb {R}\) if \(f(T) \leqslant f(S)\) holds whenever TS (possibly unbounded) are self-adjoint operators in \(\mathcal {H}\) with spectrum contained in (ab) and \(T \leqslant S\).

Assuming \(f \in P(a,b)\), let \(\mu \) be the measure associated with f, see (7.1), and denote the supremum of the support of \(\mu \) by \(\varSigma _{\mu }\). In what follows, the discussion is for general self-adjoint TS, but one should keep in mind that the results are applied to \(T =\langle A \rangle \) and \(S = \sqrt{c_d} \langle N \rangle \), cf. Lemma 4.13. We start with a lemma:

Lemma 7.2

Let f belong to \(P(a,+\infty )\). Then

  1. (1)

    \(\varSigma _{\mu } \leqslant a\),

  2. (2)

    given T a self-adjoint operator with \(\inf (\sigma (T))> a >0\), we have f(T) bounded from below and \(\mathrm {Dom}[T] \subseteq \mathrm {Dom}[|f(T)|]\) with equality if and only if \(\beta >0\).

Proof

For (1), we start by noting that f belongs to \(P(a,+\infty )\) so f admits an integral representation as in (7.1). Thus, \(\varSigma _{\mu } \leqslant a\) holds thanks to the Stieltjes inversion formula and the fact that f is real-valued on \((a,+\infty )\). For (2), adding a constant to f does not alter the assumptions of the lemma and leads to the same conclusion, so we may assume \(f(\inf (\sigma (T))) >0\), where \(\inf (\sigma (T)) > a\) (recall f is non-decreasing on \((a,+\infty )\)). We start by showing that \(\lim f(x)/x = \beta \), as \(x\rightarrow +\infty \), or equivalently

$$\begin{aligned} \lim \limits _{x \rightarrow +\infty } \int _{-\infty } ^{\varSigma _{\mu }} \frac{1+\lambda x}{x(\lambda -x)(\lambda ^2+1)} d\mu (\lambda ) = 0.\end{aligned}$$

We wish to exchange the order of the limit and integration. We have

$$\begin{aligned} \big | (1+\lambda x)x^{-1}(\lambda -x)^{-1} \big | \leqslant 1, \quad \forall (\lambda ,x) \in (-\infty ,0] \times [1,+\infty ) \cup [0,a] \times [a+\sqrt{a^2+1}, + \infty ). \end{aligned}$$
(7.2)

We may apply the dominated convergence theorem, and the above limit follows. This limit implies that f(x)/x is a bounded function on \((a,+\infty )\) and hence \(\mathrm {Dom}[T] \subset \mathrm {Dom}[|f(T)|]\). For the reverse inclusion, x/f(x) is a well-defined bounded function on \((\inf (\sigma (T)),+\infty )\) iff \(\beta >0\). \(\square \)

We are now ready to prove the extension of Theorem 7.1 to semi-bounded operators:

Theorem 7.3

Let \(f : (a,b) \mapsto \mathbb {R}\), where \(0< a < b \leqslant +\infty \). Then f is operator monotone in (ab) if and only if f admits an analytic continuation that belongs to P(ab).

Proof

If f is operator monotone in (ab), then in particular it is matrix operator monotone of order n in (ab) for all \(n \in \mathbb {N}\), and so f admits an analytic continuation that belongs to P(ab) by Loewner’s theorem. This direction is in fact the hard direction in Loewner’s theroem, but the extension is trivial!

For the converse, we suppose that f admits an analytic continuation that belongs to P(ab). Consider the two separate cases \(b < +\infty \) and \(b = +\infty \). If \(b < +\infty \), then f is matrix operator monotone of order n in (ab) for all \(n \in \mathbb {N}\) by Loewner’s theorem. But, since the interval (ab) is finite, this is equivalent to being operator monotone in (ab); see, e.g., [46, Lemma 2.2] or [44, Chapter 2]. Now the case \(b=+\infty \). We have \(\varSigma _{\mu } \leqslant a\) by Lemma 7.2, and

$$\begin{aligned} f(x) = \alpha + \beta x + \int _{-\infty } ^{\varSigma _{\mu }} \left( \frac{1}{\lambda - x} - \frac{\lambda }{\lambda ^2+1} \right) d \mu (\lambda ), \quad x \in (a,+\infty ) .\end{aligned}$$

Let

$$\begin{aligned} g_r (x) := \alpha + \int _{-r} ^{\varSigma _{\mu }} \left( \frac{1}{\lambda - x} - \frac{\lambda }{\lambda ^2+1} \right) d \mu (\lambda ), \quad g(x) := f(x) - \beta x, \quad x \in (a,+\infty ). \end{aligned}$$
(7.3)

Note that \(\{g_r \}_{r \in \mathbb {R}^+}\) is of a sequence of functions of the real variable x that converges pointwise to g(x) as \(r \rightarrow + \infty \) for all \(x \in (a,+\infty )\). Moreover, since for every fixed r, \(g'_r(x) \geqslant 0\), all functions in the sequence are increasing in the variable x. We need a lemma.

Lemma 7.4

For every fixed \(\varepsilon > 0\), the subsequence \(\{ g_r(x) \}_{r > 1/ \varepsilon }\) has the property that \(g_r(x) \nearrow g(x)\) for every \(x >\max (\varSigma _{\mu } + \varepsilon , a)\).

Proof

Clearly \(g_r(x) \rightarrow g(x)\) as \(r\rightarrow +\infty \) for all \(x \in (a,+\infty )\). What needs to be shown is that the sequence \(g_r(x)\) is increasing pointwise. Let \(\varepsilon > 0\) be given and fix \(x > \max (\varSigma _{\mu } + \varepsilon , a)\). Recall a is assumed to be strictly positive. The integrand in (7.3) is equal to

$$\begin{aligned} \frac{1+\lambda x}{(\lambda -x) (\lambda ^2+1)}.\end{aligned}$$

On the one hand, \(x > \varSigma _{\mu }+ \varepsilon \) implies that \((\lambda -x)\) is negative for all \(\lambda \in \mathrm {supp} \ \mu \). On the other hand, the assumption \(x> a > 0\) implies that \((1+\lambda x)\) is negative for all \(\lambda < -1/x\). Thus, for all \(x >\max (\varSigma _{\mu } + \varepsilon , a)\) and for all \(\lambda < -1/a\), the integrand in (7.3) is strictly positive. Thus, as r increases above 1/a, the value of the integral in (7.3) strictly increases. This completes the proof of the lemma. \(\square \)

Continuation of the Proof of Theorem7.3. Now fix TS self-adjoint operators with spectrum contained in \((a,+\infty )\) and \(T \leqslant S\). Let \(\psi \in \mathcal {H}\). \(T \leqslant S\) implies \(\langle \psi , (\lambda - T)^{-1} \psi \rangle \leqslant \langle \psi , (\lambda - S)^{-1} \psi \rangle \) for all \(\lambda \leqslant \varSigma _{\mu }\), e.g., [47, Theorem VI.2.21]. We integrate over the compact \([-r,\varSigma _{\mu }]\) :

$$\begin{aligned} \int _{-r}^{\varSigma _{\mu }} \langle \psi , (\lambda - T)^{-1} \psi \rangle \ d\mu (\lambda ) \leqslant \int _{-r}^{\varSigma _{\mu }} \langle \psi , (\lambda - S)^{-1} \psi \rangle \ d\mu (\lambda ), \quad \text {for every finite} \ r > |\varSigma _{\mu }|.\end{aligned}$$

Moreover, since the integrand is norm continuous, we infer

$$\begin{aligned} \langle \psi , g_r(T) \psi \rangle \leqslant \langle \psi , g_r(S) \psi \rangle , \quad \text {for every finite} \ r > |\varSigma _{\mu }|. \end{aligned}$$
(7.4)

Next we choose in Lemma 7.4\(\varepsilon \) small enough so that \(\max (\varSigma _\mu +\varepsilon , a)<\inf (\sigma (T))\leqslant \inf (\sigma (S))\). We obtain \(g_r(x) \nearrow g(x)\) (for \(r>1/\varepsilon \)) as r goes to infinity. In turn, the monotone convergence theorem for forms, e.g., [47, Theorem VIII.3.11], ensures that \(\langle \psi , g_r(T) \psi \rangle \) converges to \(\langle \psi , g(T)\psi \rangle \) as \(r \rightarrow +\infty \) for all \(\psi \in \mathrm {Dom}[|g(T)|^{1/2}]\), and similarly for S. Taking limits in (7.4) gives \(\mathrm {Dom}[|g(S)|^{1/2}]\subset \mathrm {Dom}[|g(T)|^{1/2}]\) and

$$\begin{aligned} \langle \psi , g(T) \psi \rangle \leqslant \langle \psi , g(S) \psi \rangle , \quad \psi \in \mathrm {Dom}[|g(S)|^{1/2}].\end{aligned}$$

Thus, if \(\beta = 0\) we have

$$\begin{aligned} \langle \psi , f(T) \psi \rangle \leqslant \langle \psi , f(S) \psi \rangle , \quad \psi \in \mathrm {Dom}[|g(S)|^{1/2}]= \mathrm {Dom}[|f(S)|^{1/2}],\end{aligned}$$

whereas if \(\beta > 0\), we use the fact that \(0 < T\leqslant S\), and \(\mathrm {Dom}[S^{1/2}] = \mathrm {Dom}[|f(S)|^{1/2}]\), \(\mathrm {Dom}[T^{1/2}]= \mathrm {Dom}[|f(T)|^{1/2}]\), by Lemma 7.2, which yields

$$\begin{aligned} \langle \psi , f(T) \psi \rangle \leqslant \langle \psi , f(S) \psi \rangle , \quad \psi \in \mathrm {Dom}[S^{1/2}] = \mathrm {Dom}[|f(S)|^{1/2}].\end{aligned}$$

This gives the result. \(\square \)

To close this section, we have a remark about results in the literature on order relations for general self-adjoint operators. In [48], Olson introduced the spectral order for self-adjoint operators: \(T \preceq S\) if and only if \(E_{(-\infty ,t]}(T) \geqslant E_{(-\infty ,t]}(S)\) for all \(t \in \mathbb {R}\). Here, \(E_{(-\infty ,t]}(T)\) and \(E_{(-\infty ,t]}(S)\) are the spectral resolutions of the identity for T and S. He showed that \(T^n \leqslant S^n\) for every \(n \in \mathbb {N}\) is equivalent to \(T \preceq S\), see also [49, Proposition 5]. Furthermore, it is shown in [50] that this order relation is equivalent to \(f(T) \leqslant f(S)\) for any continuous monotone non-decreasing function f defined on an interval which contains \(\sigma (T) \cup \sigma (S)\). For the purpose of this article, we do not know if \(\langle A \rangle ^n \leqslant (\sqrt{c_d} \langle N \rangle )^n \) holds \(\forall n \in \mathbb {N}\) but if it does it would considerably simplify this article. To check this inequality by brute force seems to be unbearable.

Appendix B. Polylogarithms of Positive Order are Nevanlinna Functions

That logarithm with the standard branch cut is a Nevanlinna function follows from the identity \(\ln (z) = \ln (r) + \mathrm {i}\theta \), where \(z = r e^{\mathrm {i}\theta }\), \(r>0\), and \(\theta \in (-\pi ,\pi )\). The integral representation of the logarithm is

$$\begin{aligned} \ln (z) = \int _{-\infty } ^0 \left( \frac{1}{\lambda -z} - \frac{\lambda }{\lambda ^2+1} \right) d\lambda . \end{aligned}$$

The composition of Nevanlinna functions produces another Nevanlinna function. So, for example, \(\ln ^p(z)\) is Nevanlinna for \(0 \leqslant p \leqslant 1\). What about higher powers of the logarithm? Certainly the square and cube of the logarithm are not Nevannlina functions. Indeed writing

$$\begin{aligned} \ln ^2(z) = \ln ^2(r) - \theta ^2 + \mathrm {i}2 \theta \ln (r), \quad \text {and} \quad \ln ^3(z) = \ln ^3(r) - 3\theta ^2 \ln (r) + \mathrm {i}\theta ( 3\ln ^2(r) -\theta ^2)\end{aligned}$$

reveals that these functions do not map \(\mathbb {C}_+\) to \(\overline{\mathbb {C}_+}\). In spite of this, there are functions that are Nevanlinna and are “almost” equal to the logarithms. To motivate the idea, we note that the Stieltjes inversion formula gives

$$\begin{aligned} \lim \limits _{\delta \downarrow 0} \lim \limits _{\varepsilon \downarrow 0} \frac{1}{\pi } \int _{\lambda _1 + \delta } ^{\lambda _2 + \delta } \mathrm {Im} \left( \ln ^2(\lambda + \mathrm {i}\varepsilon ) \right) d\lambda = {\left\{ \begin{array}{ll} 0 &{} \lambda _1 \geqslant 0, \\ \int _{\lambda _1} ^{\lambda _2} 2\ln (|\lambda |) d\lambda &{} \lambda _2 \leqslant 0 \\ \end{array}\right. } \end{aligned}$$

when applied to \(\ln ^2(z)\) and

$$\begin{aligned} \lim \limits _{\delta \downarrow 0} \lim \limits _{\varepsilon \downarrow 0} \frac{1}{\pi } \int _{\lambda _1 + \delta } ^{\lambda _2 + \delta } \mathrm {Im} \left( \ln ^3(\lambda + \mathrm {i}\varepsilon ) \right) d\lambda = {\left\{ \begin{array}{ll} 0 &{} \lambda _1 \geqslant 0, \\ \int _{\lambda _1} ^{\lambda _2} \left( 3\ln ^2(| \lambda |) - \pi ^2 \right) d\lambda &{} \lambda _2 \leqslant 0 \\ \end{array}\right. } \end{aligned}$$

when applied to \(\ln ^3(z)\). This suggests to calculate the Nevanlinna functions corresponding to the measures \(\mathrm {d}\mu (\lambda ) = \varvec{1}_{\{\lambda < -1\}} \ln (-\lambda ) d\lambda \) and \(\mathrm {d}\mu (\lambda ) = \varvec{1}_{\{\lambda < -1\}} \ln ^2(-\lambda ) d\lambda \).

We introduce polylogarithms. We refer to [51, 52] for formulas and a detailed exposition. The polylogarithm of order \(\sigma \in \mathbb {C}\) is defined by the power series

$$\begin{aligned} \mathrm {Li}_{\sigma }(z) := \sum _{k=1} ^{\infty } \frac{z^k}{k^{\sigma }}.\end{aligned}$$

The definition is valid for complex \(|z| <1\) and is extended to the complex plane by analytic continuation. For the purpose of this article, we are interested in the polylogarithms with \(\sigma > 2\), or \(\sigma = 3\) if we want to simplify by taking the smallest integer above 2. The standard branch cut is \([1,+\infty )\) for \(\mathrm {Li}_1(z)\) and \((1,+\infty )\) for \(\mathrm {Li}_2(z)\) and \(\mathrm {Li}_3(z)\). The polylogarithm of order 1 can be written in terms of a logarithm as \(\mathrm {Li}_1(z) = - \ln (1-z)\). The polylogarithm of order 2 is called the dilogarithm or Spence function, while the polylogarithm of order 3 is called the trilogarithm. On p. 494 of [52], the following integral representation is given without proof:

$$\begin{aligned} \mathrm {Li}_{\sigma +1} (z) = \frac{z}{\varGamma (\sigma +1)} \int _{1} ^{\infty } \frac{\ln ^{\sigma } (\lambda ) }{\lambda (\lambda -z)} d\lambda , \end{aligned}$$
(8.1)

for \(| \mathrm {arg} (1-z) | < \pi , \mathrm {Re}(\sigma ) > -1\), or for \( z=1, \mathrm {Re}(\sigma ) > 0\). Here, \(\varGamma \) is the gamma function. Obviously (8.1) is equivalent to

$$\begin{aligned} \mathrm {Li}_{\sigma +1} (z) = - \frac{1}{\varGamma (\sigma +1)} \int _{1} ^{\infty } \frac{\ln ^{\sigma } (\lambda ) }{\lambda (\lambda ^2+1)} d\lambda + \frac{1}{\varGamma (\sigma +1)} \int _{1} ^{\infty } \left( \frac{1}{\lambda -z} - \frac{\lambda }{ \lambda ^2+1} \right) \ln ^{\sigma } (\lambda ) d\lambda . \end{aligned}$$
(8.2)

This means that \(\mathrm {Li}_{\sigma +1}(z)\) is a Nevanlinna function for \(\mathrm {Re}(\sigma ) > -1\). Although not difficult to prove, it is not clear where a Proof of (8.1) can be found in the literature. Thus, we prove it:

Proposition 8.1

(8.1) is true.

Proof

Let \(\lambda \geqslant 1\). Writing \(1/(\lambda (\lambda -z))\) as a power series in z we have \(1/(\lambda (\lambda -z)) = \sum _{k=0} ^{\infty } \lambda ^{-k-2} z^k\) for \(|z| < 1\). Then the rhs of (8.1) is equal to

$$\begin{aligned} \frac{z}{\varGamma (\sigma +1)} \int _{1} ^{\infty } \sum _{k=0} ^{\infty } \frac{\ln ^{\sigma } (\lambda ) }{\lambda ^{k+2}} z^k d\lambda = \sum _{k=1} ^{\infty } \frac{z^k}{\varGamma (\sigma +1)} \int _{1} ^{\infty } \frac{\ln ^{\sigma } (\lambda ) }{\lambda ^{k+1}} d\lambda = \sum _{k=1} ^{\infty } \frac{z^k}{k^{\sigma +1}}, \quad \text {Re}(\sigma ) >-1.\end{aligned}$$

To evaluate the last integral the change of variable \(k \ln (\lambda ) = t\) was performed, followed by the definition of the gamma function. Thus, the rhs of (8.1) is equal to \(\mathrm {Li}_{\sigma +1}(z)\) for \(|z|<1\) and \(\text {Re}(\sigma ) >-1\). The result follows by the uniqueness of the analytic continuation. \(\square \)

While we are at it, we note that \(\mathrm {Li}_{0} (z) = z/(1-z)\) is also a Nevanlinna function.

Definition

For \(\sigma \in \mathbb {C}\),

$$\begin{aligned} \varPhi _{\sigma } (z) := - \mathrm {Li}_{\sigma }(-z), \quad z \in \mathbb {C}{\setminus } (-\infty ,-1]. \end{aligned}$$
(8.3)

Clearly (8.2) implies that the \(\varPhi _{\sigma }\) are Nevanlinna for \(\text {Re}(\sigma ) >-1\) with integral representations given by:

$$\begin{aligned} \varPhi _{\sigma +1} (z) = \frac{1}{\varGamma (\sigma +1)} \int _{1} ^{\infty } \frac{\ln ^{\sigma } (\lambda ) }{\lambda (\lambda ^2+1)} d\lambda + \frac{1}{\varGamma (\sigma +1)} \int _{-\infty } ^{-1} \left( \frac{1}{\lambda -z} - \frac{\lambda }{ \lambda ^2+1} \right) \ln ^{\sigma } (-\lambda ) \mathrm{d}\lambda , \end{aligned}$$

for \(|\text {arg}(1+z) | < \pi \). Finally, the other reason we resort to polylogarithms is because they decay at the same rate as the logarithms, at least for positive integer order (this follows directly from the inversion/reflection formula [51, (6) of Appendix A.2.7] together with \(\mathrm {Li}_n(0)=0\)), namely

$$\begin{aligned} \lim \limits _{x \rightarrow +\infty } \frac{\varPhi _n(x)}{\ln ^ n (x)} = \frac{1}{n!}, \quad n \in \mathbb {N}. \end{aligned}$$
(8.4)

Appendix C. Almost Analytic Extensions and Helffer–Sjöstrand Calculus

We refer to [9, 28, 38, 53, 54] for more details. Let \(\rho \in \mathbb {R}\) and denote by \(\mathcal {S}^{\rho }(\mathbb {R})\) the class of functions \(\varphi \) in \(C^{\infty }(\mathbb {R})\) such that

$$\begin{aligned} |\varphi ^{(k)} (x)| \leqslant C_k \langle x \rangle ^{\rho -k}, \quad \text {for all} \ k \in \mathbb {N}. \end{aligned}$$
(9.1)

Lemma 9.1

[38, 53] Let \(\varphi \in \mathcal {S}^{\rho }(\mathbb {R})\), \(\rho \in \mathbb {R}\). Then for every \(N \in \mathbb {Z}^+\) and every \(c>0\), there exists a smooth function \(\tilde{\varphi }_N : \mathbb {C}\rightarrow \mathbb {C}\), called an almost analytic extension of \(\varphi \), satisfying:

$$\begin{aligned} \tilde{\varphi }_N(x+\mathrm {i}0)= & {} \varphi (x), \quad \forall x \in \mathbb {R};\nonumber \\ \mathrm {supp} \ (\tilde{\varphi }_N) \subset \varOmega:= & {} \{x+\mathrm {i}y : |y| \leqslant c \langle x \rangle \};\nonumber \\ \tilde{\varphi }_N(x+\mathrm {i}y)= & {} 0, \quad \forall y \in \mathbb {R}\ \mathrm {whenever} \ \varphi (x) = 0;\nonumber \\ \forall \ell \in \mathbb {N}\cap [0,N], \Bigg | \frac{\partial \tilde{\varphi }_N}{\partial \overline{z}}(x+\mathrm {i}y) \Bigg |\leqslant & {} c_{\ell } \langle x \rangle ^{\rho -1-\ell } |y|^{\ell } \ \mathrm {for \ some \ constants} \ c_{\ell } >0.\nonumber \\ \end{aligned}$$
(9.2)

Now let Q be a self-adjoint operator.

Lemma 9.2

Let \(\rho < 0\) and \(\varphi \in \mathcal {S}^{\rho }(\mathbb {R})\). Then for all \(k \in \mathbb {N}\) and \(N \in \mathbb {N}\):

$$\begin{aligned} \varphi ^{(k)}(Q) = \frac{\mathrm {i}(k!)}{2\pi } \int _{\mathbb {C}} \frac{\partial \tilde{\varphi }_N}{\partial \overline{z}} (z) (z-Q)^{-1-k} \mathrm{d}z \wedge \mathrm{d}\overline{z} \end{aligned}$$
(9.3)

where the integral exists in the norm topology. For \(\rho \geqslant 0\), the following limit exists:

$$\begin{aligned} \varphi ^{(k)}(Q)f = \lim \limits _{R \rightarrow \infty } \frac{\mathrm {i}(k!)}{2\pi } \int _{\mathbb {C}} \frac{\partial (\tilde{\varphi \theta _R})_N}{\partial \overline{z}} (z) (z-Q)^{-1-k} f \mathrm{d}z \wedge \mathrm{d}\overline{z}, \quad \text {for all} \ f \in \mathrm {Dom}[\langle Q \rangle ^{\rho }]. \end{aligned}$$
(9.4)

In particular, if \(\varphi \in \mathcal {S}^{\rho }(\mathbb {R})\) with \(0 \leqslant \rho < k\) and \(\varphi ^{(k)}\) is a bounded function, then \(\varphi ^{(k)}(Q)\) is a bounded operator and (9.3) holds (with the integral converging in norm).

Proposition 9.3

[28]. Let T be a bounded self-adjoint operator satisfying \(T \in \mathcal {C}^1(Q)\). Then:

$$\begin{aligned}{}[T,(z-Q)^{-1}]_{\circ } = (z-Q)^{-1}[T,Q]_{\circ }(z-Q)^{-1}, \end{aligned}$$
(9.5)

and for any \(\varphi \in \mathcal {S}^{\rho }(\mathbb {R})\) with \(\rho < 1\), \(T \in \mathcal {C}^1(\varphi (Q))\) and

$$\begin{aligned}{}[T,\varphi (Q)]_{\circ } = \frac{\mathrm {i}}{2\pi }\int _{\mathbb {C}} \frac{\partial \tilde{\varphi }_N}{\partial \overline{z}} (z-Q)^{-1}[T,Q]_{\circ }(z-Q)^{-1} \mathrm{d}z\wedge \mathrm{d}\overline{z}. \end{aligned}$$
(9.6)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Golénia, S., Mandich, MA. Limiting Absorption Principle for Discrete Schrödinger Operators with a Wigner–von Neumann Potential and a Slowly Decaying Potential. Ann. Henri Poincaré 22, 83–120 (2021). https://doi.org/10.1007/s00023-020-00971-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00023-020-00971-9

Mathematics Subject Classification

Navigation