A new spectral analysis of stationary random Schrödinger operators
Abstract.
Motivated by the long-time transport properties of quantum waves in weakly disordered media, the present work puts random Schrödinger operators into a new spectral perspective. Based on a stationary random version of a Floquet type fibration, we reduce the description of the quantum dynamics to a fibered family of abstract spectral perturbation problems on the underlying probability space. We state a natural resonance conjecture for these fibered operators: in contrast with periodic and quasiperiodic settings, this would entail that Bloch waves do not exist as extended states, but rather as resonant modes, and this would justify the expected exponential decay of time correlations. Although this resonance conjecture remains open, we develop new tools for spectral analysis on the probability space, and in particular we show how ideas from Malliavin calculus lead to rigorous Mourre type results: we obtain an approximate dynamical resonance result and the first spectral proof of the decay of time correlations on the kinetic timescale. This spectral approach suggests a whole new way of circumventing perturbative expansions and renormalization techniques.
MITIA DUERINCKX1,2***[email protected] AND CHRISTOPHER SHIRLEY1†††[email protected]
1Université Paris-Saclay, CNRS, Laboratoire de Mathématiques d’Orsay, 91405 Orsay, France
2Université Libre de Bruxelles, Département de Mathématique, 1050 Brussels, Belgium
1. Introduction
1.1. General overview
We consider random Schrödinger operators of the form
on , where is a realization of a “stationary” (that is, statistically translation-invariant) random potential , constructed on a given probability space , and we study the properties of the corresponding Schrödinger flow on ,
with initial data . This well-travelled equation models the motion of an electron in a disordered medium described by the potential , where the coupling constant stands for the strength of the disorder.
For comparison, let us first recall transport properties in the simpler case of periodic or quasiperiodic disorder. If is periodic, the energy transport is well-known to remain purely ballistic as for the free flow , cf. [3]. The proof relies on the absolute continuity of the spectrum of the periodic Schrödinger operator on , and more precisely on the existence of so-called Bloch waves, which are extended states constructed by means of standard perturbation theory as deformations of Fourier modes, with periodic. In case of a quasiperiodic potential , the problem is more involved and depends on the strength of the disorder: energy transport is expected to remain ballistic only at weak coupling or at high energies. This is rigorously established in dimension [54]. In higher dimensions , for a specific class of quasiperiodic potentials, it was recently shown that there exist initial data at high energies that indeed display ballistic transport [25, 26] (the analysis of the discrete setting is more complete [6]). Inspired by the periodic case, the proof relies on the existence of corresponding Bloch waves as extended states of the form with quasiperiodic, but their construction is more intricate since standard perturbation theory no longer applies. In [13], we provide a simple method to construct “approximate” Bloch waves and deduce ballistic transport for all data at least up to “very long” timescales both at weak coupling and at high energies. These results in the periodic and quasiperiodic settings show how Bloch waves are crucial tools to infer transport properties of the Schrödinger flow.
The present work is concerned with the more general stationary random setting in the weak coupling regime . In case of a random potential with short-range correlations, in stark contrast with the periodic and quasiperiodic cases, a celebrated conjecture by Anderson [2] states that in dimension every initial condition can be almost surely decomposed into two parts: a low-energy part that remains dynamically localized and a bulk-energy part that propagates diffusively. Despite the great recent achievements of rigorous perturbation theory in some asymptotic time regimes, e.g. [51, 18, 17, 16, 9], successfully describing the emergence of irreversible diffusion from the reversible Schrödinger dynamics, the full justification of this quantum diffusion phenomenon remains a major open problem in mathematical physics [49, 15]. More precisely, the ensemble-averaged Wigner transform of the quantum wave is known to converge to the solution of a linear Boltzmann equation on the kinetic timescale , and of a heat equation on longer times, but the justification is limited to a perturbative time regime for some small . A simplified question concerns the behavior of time correlations in form of the averaged wavefunction , which is expected to display exponential time decay: more precisely, in Fourier space, on the kinetic timescale , as ,
(1.1) |
where the decay rate would coincide with the total scattering cross section in the corresponding Boltzmann equation, and where corrections are added on longer times. A proof of this exponential decay on the kinetic timescale is given in [9] based on a perturbative expansion of a Feynman-Kac type formula. The perturbative analysis of [18, 17, 16] would further yield an improved result, but still restricted to limited timescales.
Motivated by these open questions, rather than trying to improve on perturbative expansions and renormalization techniques, we aim at developping an alternative spectral approach to describe the long-time behavior of the system beyond perturbative timescales. More precisely, we take inspiration from the periodic and quasiperiodic cases, although the behavior radically differs from the present random setting, and we investigate the role of a corresponding notion of Bloch waves. It appears that these Bloch waves are no longer extended states associated with absolutely continuous spectrum: they are expected to be only defined in a weak distributional sense in probability and to play the role of resonant modes associated with some kind of “continuous resonant spectrum”. Exploiting ideas from Malliavin calculus, we manage to appeal to perturbative Mourre’s theory, cf. Theorem 4, which leads to the construction of approximate dynamical resonances and constitutes the first spectral proof of (1.1), cf. Corollary 5. Non-perturbative refinements to reach longer times are postponed to future works, as well as the investigation of other possible dynamical consequences in closer connection with quantum diffusion.
1.2. Summary of our approach and results
We briefly describe the framework of our new approach to Schrödinger operators . First, we change the point of view and rather consider the operator as acting on the augmented Hilbert space , then studying the corresponding Schrödinger flow on ,
(1.2) |
with deterministic initial data . This can be viewed as including stochastic averaging conveniently into the functional setup; see also [45, 24]. (Note that on has absolutely continuous spectrum as a consequence of Wegner estimates when on has almost sure pure point spectrum.)
As we have shown and already used in [13], see Section 3 below for details (also [4, 24]), the operator on can be decomposed via a Fourier-type transformation as a direct integral of fibered operators acting on the elementary space , which is viewed as the space of stationary random fields,
(1.3) |
The (centered) fibered operators take the form
where and denote the stationary gradient and Laplacian on ; see Section 3.2 for proper definitions. In particular, the Schrödinger flow is decomposed as
(1.4) |
in terms of the fibered evolutions on . This partial diagonalisation via Fourier is henceforth referred to as the stationary Floquet–Bloch fibration, in analogy with the well-known corresponding construction in the periodic setting, e.g. [32, 33].
At vanishing disorder , as the constant function is an eigenfunction with eigenvalue for the unperturbed fibered operators , the associated spectral measure coincides with the Dirac mass at , and the decomposition (1.4) then reduces to the usual Fourier diagonalisation of the free flow,
(1.5) |
When the disorder is turned on but small, , the description of the Schrödinger flow is reduced to a family of (hopefully simpler) fibered perturbation problems for the spectral measures. In case of a periodic potential (that is, , cf. Remark 3.4(a)), the fibered operators have compact resolvent in view of the Rellich theorem, and thus discrete spectrum. The eigenvalue at is then typically simple and isolated, which allows to apply standard perturbation methods, e.g. [32, 3], showing that it is perturbed into isolated eigenvalues . In other words, Fourier modes are perturbed into so-called periodic Bloch waves that diagonalize the Schrödinger flow and are associated with perturbed generalized eigenvalues of the form . This entails that the flow is approximately conjugated to the free flow (1.5) in the sense of
and in particular the energy transport remains ballistic forever. The same conclusion holds in fact for all , cf. [3]. In case of a quasiperiodic potential (cf. Remark 3.4(a)), the situation is expected to be similar at weak coupling, but the existence of corresponding Bloch waves is a more subtle question: quasiperiodic fibered operators are degenerate elliptic operators, for which compactness fails, and the simple eigenvalue at is no longer isolated but embeds in dense pure point spectrum, so that no standard perturbation theory applies; see [25, 26, 13].
In case of a random potential with short-range correlations, the situation differs drastically in link with the expected diffusive behavior. We show that is typically the only eigenvalue of the fibered operators , is simple, and embeds in absolutely continuous spectrum, cf. Proposition 1 below. According to Fermi’s Golden Rule, whenever the disorder is turned on, this embedded eigenvalue is then expected to dissolve in the continuous spectrum, cf. Proposition 3, and to turn into a complex resonance at
in the lower complex half-plane. In particular, this provides a spectral explanation why approximate Bloch wave analysis leading to ballistic transport as in [13] breaks down on the kinetic timescale . For the averaged wavefunction, this leads to expect
which would indeed agree with the exponential decay (1.1) on the kinetic timescale, with , and a finer resonance analysis would yield a more accurate expansion. From a spectral perspective, fibered resonances are transferred via the fibration (1.3) to kind of a “continuous resonant spectrum” for the full operator on , cf. Remark 2.1 below.
General spectral tools are however dramatically missing to rigorously study these fibered perturbation problems on , in particular due to the lack of any relative compactness of the perturbation. We start by performing a detailed study of rudimentary spectral properties of fibered operators, cf. Propositions 1–3, emphasizing the strong dependence on the structure of the underlying probability space . Next, we appeal to Mourre’s theory [37, 1] as a rigorous approach to fibered perturbation problems. More precisely, we start by constructing Mourre conjugates for the unperturbed operators . The construction requires a surprisingly nontrivial work and relies on a deep use of Malliavin calculus, which constitutes the core of our contribution, cf. Section 5. This construction is however not compatible with the perturbation in the sense that cannot be considered as a small perturbation of in the sense of Mourre’s theory, in link with the infinite dimensionality of the probability space. For this reason, we only manage to apply perturbative Mourre’s theory under a suitable (weak) truncation, cf. Theorem 4. As a direct consequence, the decay law (1.1) is recovered at least on the kinetic timescale, cf. Corollary 5. Finally, we give a relevant formulation of resonance conjectures for fibered operators, cf. Conjectures Conjecture (LRC) — Local resonance conjecture and Conjecture (GRC) — Global resonance conjecture, which are motivated by our partial results and are shown to imply the expected decay law (1.1) to finer accuracy on all timescales, cf. Corollary 6. These conjectures are further illustrated in Section 7, where we display a toy model that shares various properties of Schrödinger operators and allows for a rigorous resonance analysis. Although these conjectures are left open, the present work sheds a new light on the study of random Schrödinger operators, in particular providing the first spectral proof of (1.1); our results will be strengthened in future works and hopefully serve as a starting point for a new line of research in the field.
Notation
-
We denote by any constant that only depends on the space dimension and on the law of the random potential . We use the notation (resp. ) for (resp. ) up to such a multiplicative constant . We write when both and hold. We add subscripts to to indicate dependence on other parameters. We denote by any quantity that is bounded by .
-
We denote by the usual Fourier transform of a smooth function on . The inverse Fourier transform is then given by in terms of the rescaled Lebesgue measure .
-
The ball centered at and of radius in is denoted by , and we write for abbreviation , , and . Without ambiguity, we occasionally also denote by the unit ball at the origin in the complex plane .
-
For a set we denote by its convex envelope, by its interior, and by its closure.
-
We denote by the set of Borel subsets of , and for we let denote the set of Borel probability measures on .
-
For a vector space , we write for its -fold tensor product, and for its -fold symmetric tensor product.
-
For , we write and .
2. Main results
This section is devoted to a brief description of our main results, while proofs and detailed statements are postponed to the next sections.
2.1. Framework
We refer to Section 3 for a suitable definition of stationarity as statistical invariance under spatial translations. Throughout, the stationary random potential is assumed real-valued and centered, . As we show, fine spectral properties crucially depend on the structure of the underlying probability space. We therefore mainly focus on Gaussian or Poisson settings, where Malliavin calculus is available and provides a useful Fock space decomposition of . More precisely, we consider the following:
-
Gaussian setting: for some Borel function and some stationary centered Gaussian random field with bounded covariance function
Equivalently, the field can be represented as
(2.1) where is a standard Gaussian white noise on and where the kernel is the convolution square root of the covariance function, .
-
Poisson setting: for some Borel function and some of the form
(2.2) where is a standard Poisson point process on and where is the single-site potential.
We say that the random potential is short-range if it has integrable decay of correlations: in the above settings, this amounts to choosing .
For shortness, in the sequel, we shall mainly restrict to the Gaussian setting, although the same results can be transferred mutatis mutandis in the Poisson case (using the corresponding version of Malliavin calculus, e.g. [44]). For simplicity, we occasionally further restrict to a random potential that is itself Gaussian: although unbounded, such potentials have a simpler action on the Fock space decomposition of .
2.2. Basic spectral theory of fibered operators
We refer to Section 3 for the construction of the stationary Floquet–Bloch fibration (1.3). Next, we start with a detailed spectral analysis of the unperturbed operators
Although the stationary Laplacian is a natural operator on and has been introduced in various settings (e.g. in the context of stochastic homogenization [42, 23]), its spectral properties have never been elucidated before, and we close this gap here. Note that some preliminary remarks on its spectrum have been made in [7, Section 3.1], see also [34, Section 2.C], namely that it is discrete if is a finite set, that there is in general no spectral gap above in contrast with the periodic setting, and that it coincides with in case of an i.i.d. structure. Interestingly, the spectrum depends crucially on the structure of the underlying probability space , as precisely formulated in Section 4.1 below in terms of a notion of “spectrum” of the probability space. In the model Gaussian setting, our result takes on the following simple guise.
Proposition 1 (Spectral decomposition of ).
Given a stationary Gaussian field on with covariance function , denote by the (nonnegative measure) Fourier transform of , and assume that the probability space is endowed with the -algebra generated by .
-
(i)
If is not periodic in any direction, then .
-
(ii)
If the measure is absolutely continuous (in particular, if is integrable), then the eigenvalue at is simple (with eigenspace ) and
We turn to the perturbed fibered operators and start with a characterization of their spectrum. While we focus here on the Gaussian setting, a more general statement is given in Section 4.2. In case of an unbounded potential , the essential self-adjointness of the perturbed operators is already a delicate issue, for which an (almost optimal) criterion is included in Appendix A, requiring for some , in line with the corresponding celebrated self-adjointness problem for Schrödinger operators on with singular potentials, cf. [27, 19, 28]; see also Proposition 5.16 for the simpler case when is itself Gaussian.
Proposition 2 (Spectrum of ).
Consider the Gaussian setting , where is a stationary Gaussian field, and assume that is nondegenerate and that holds for some . Then is essentially self-adjoint on and
The nature of the spectrum of the perturbed operators is a more involved question and is a main concern in the sequel. In view of the fibration (1.4), the perturbation of the eigenvalue at for the fibered operators is of particular interest. According to Fermi’s Golden Rule, e.g. [48, Section XII.6], this eigenvalue embedded in continuous spectrum is expected to dissolve when the perturbation is turned on. The simplest rigorous version of this key conjecture is as follows. It is based on observing that the formula for the second derivative of a hypothetic branch of eigenvalues at would be a complex number, cf. Section 4.3.
Proposition 3 (Instability of the bound state).
Let , let be a stationary random field, denote by the (nonnegative measure) Fourier transform of its covariance function, and assume that does not vanish identically on the sphere , in the sense that . Then there exists no branch
with
This basic instability result is however quite weak: for , the operator is in fact expected to have purely absolutely continuous spectrum in a neighborhood of for . In addition, in view of the resonance interpretation of Fermi’s Golden Rule, which originates in the work of Weisskopf and Wigner [53], the perturbed eigenvalue is expected to turn into a complex resonance. Relevant conjectures are formulated in Section 2.4 below.
2.3. Perturbative Mourre’s commutator approach
The perturbation problem for an eigenvalue embedded in continuous spectrum, in link with Fermi’s Golden Rule and resonances, is an active topic of research in spectral theory. Various general approaches have been successfully developed, see e.g. [14] and references therein, but none seems to be available in our probabilistic setting: a key difficulty is that the random perturbation is never compact with respect to the unperturbed operators on (unless it is degenerate). This calls for the development of robust techniques for the spectral analysis of stationary operators on the probability space. In the present contribution, we appeal to Mourre’s commutator theory [37, 1], cf. Section 5.1, which is reputedly flexible and requires no compactness. Although not allowing to deduce the existence of resonances in any strong form, Mourre’s theory would ensure similar dynamical consequences, e.g. [41, 21, 8].
More precisely, we start with the construction of a natural group of dilations on in the model Gaussian setting, cf. Section 5.3: heuristically, it amounts to dilating the underlying white noise in the representation (2.1), which constitutes a unitary group since dilations preserve the law of the white noise. The generator of this group is then checked to be a conjugate operator for the stationary Laplacian in the sense of Mourre’s theory, cf. Proposition 5.7(i). In Section 5.6, by means of suitable deformations, we further construct corresponding conjugates for the whole family of fibered operators . This appears to be surprisingly more involved than for , in link with the infinite dimensionality of the probability space: our proof makes a deep use of the Fock space structure of as provided by Malliavin calculus, thus emphasizing the interplay between spectral theory and the functional structure of the probabilistic setup. Next, we turn to perturbed operators . It appears that the perturbation is not compatible in the sense of Mourre’s theory, cf. Proposition 5.7(iii), again in link with the infinite dimensionality of the probability space. Perturbative Mourre’s theory can therefore not be applied unless we introduce a suitable (weak) Wiener truncation.
Theorem 4 (Perturbative Mourre’s theory up to truncation).
Let be a stationary Gaussian field with covariance function , let denote the Ornstein–Uhlenbeck operator for the associated Malliavin calculus, cf. Section 5.2, and for a given constant consider the truncation onto Wiener chaoses of order . Then, for any , there exists a self-adjoint operator on and an explicit core , cf. (2.4), such that the following properties hold:
-
(i)
For all , the truncated operator satisfies a Mourre relation on the interval with respect to . More precisely, its domain is invariant under the unitary group generated by , and its commutator with is well-defined and essentially self-adjoint on , is -bounded, and satisfies the following lower bound,
-
(ii)
The truncated perturbation is compatible with respect to in the sense that its iterated commutators with are well-defined on and bounded by .
In particular, the truncated perturbed operator satisfies a corresponding Mourre relation on with respect to . ∎
Based on this perturbative Mourre result, an approximate dynamical resonance analysis can be developed for truncated fibered operators in the spirit of [41, 21, 8] and leads to the exponential time decay of the corresponding averaged wavefunction. Further noting that the truncation error is easily estimated on the kinetic timescale, we can get rid of the truncation and rigorously deduce the validity of the exponential decay law (1.1) as stated below; the proof is postponed to Section 5.7.
Corollary 5 (Exponential decay law on kinetic timescale).
Let have compactly supported Fourier transform, let be a stationary Gaussian field with covariance function , and define by
that is, more explicitly,
(2.3) | |||||
Then there exists such that the Schrödinger flow satisfies for all ,
Although our truncation argument could be compared with the truncation of the Dyson series in the perturbative analysis of [51, 18, 17, 16], it only requires to estimate a truncation error, which is often a simpler matter, while the truncated evolution is intrinsically analyzed by means of Mourre’s theory, avoiding any Feynman diagram analysis or any renormalization to handle the truncated Dyson series. In addition, formal computations indicate that the truncation of the evolution at time on Wiener chaoses of order should be accurate provided . Since our truncation in Theorem 4 amounts to projecting onto chaoses of order , which is a particularly high order compared to truncations of Dyson series in [51, 18, 17, 16], the accuracy in Corollary 5 should thus follow in fact up to times . Non-perturbative approaches to fibered resonances and accuracy on even longer timescales are postponed to future works.
2.4. Exact resonance conjectures
The above results provide partial indications that the eigenvalue at of the fibered operators should dissolve in the continuous spectrum upon perturbation and turn into complex resonances. In particular, in agreement with Fermi’s Golden Rule, Corollary 5 is consistent with resonances at
As resonance theories have never been constructed for operators on the probability space, we formulate relevant conjectures that will be investigated rigorously in future works. To emphasize the relevance of our formulation, we further consider in Section 7 an illustrative toy model that shares many spectral features of Schrödinger operators and for which resonances are explicitly shown to exist in a similar sense, cf. Theorem 7.1(iii).
According to the usual definition, the operator has a resonance at in the lower complex half-plane if the resolvent on the upper half-plane , when viewed in a suitably weakened topology, extends to a meromorphic family of operators indexed by all (or at least in a complex neighborhood), and if this family admits a simple pole at . In the usual case of operators on , the suitable weakening of the topology typically consists of viewing the resolvent as a family of linear operators rather than . In the present setting on the probability space, the role of can be played for instance by the dense linear subspace of -polynomials,
(2.4) |
and the dual is then replaced by the dual space of continuous linear functionals on . In these terms, we formulate the following resonance conjecture. The linear functionals and below are referred to as the resonant and co-resonant states, respectively. Since the imaginary part of the expected branch of resonances vanishes to leading order both as (in dimension ) and as , cf. formula (2.3), we henceforth restrict to in a compact set away from .
Conjecture (LRC) — Local resonance conjecture.
Given a compact set , there are such that for all and the resolvent defined on as a family of operators can be extended meromorphically to the whole complex ball with a unique simple pole. In other words, there exist continuous collections and such that for all we can write for ,
(2.5) |
where the remainder is holomorphic on the set and has continuous dependence on . ∎
Next, we state a global version of this resonance conjecture in the case of an unbounded potential with (see e.g. Proposition 2). A direct computation shows that the spectral measure of associated with is typically supported on the whole half-axis and is only -times differentiable at in dimension , cf. proof of Lemma 4.2; this suggests that the band on which the meromorphic extension of the resolvent exists must shrink as close to .
Conjecture (GRC) — Global resonance conjecture.
The same decomposition (2.5) holds with a remainder that is holomorphic on the set for some exponent , has continuous dependence on , and satisfies a uniform bound of the form
Remark 2.1 (Continuous resonant spectrum).
When integrated along the Floquet–Bloch fibration (1.3), in dimension , the conjectured fibered resonances would yield a hammock-shaped set in the lower complex half-plane, connecting some point on the real axis to ,
This set is increasingly thinner at infinity and can reach a thickness in the middle, but it reduces to a curve for instance when the covariance is radial. This set can be viewed as a kind of “continuous resonant spectrum” for the Schrödinger operator on . While to the best of our knowledge such a notion has never been introduced in the literature, it is made rigorous for the illustrative toy model that we introduce in Section 7, cf. Theorem 7.1(iv). ∎
We show that the above conjectures imply the expected exponential decay law (1.1) for the averaged wavefunction to finer accuracy, thus providing a strong improvement and a very first workaround for the available perturbative methods [51, 18, 17, 16, 9]. Under Conjecture Conjecture (LRC) — Local resonance conjecture an accurate description of the decay law is deduced only for times , but accuracy is reached on all timescales under Conjecture Conjecture (GRC) — Global resonance conjecture. The result is expressed as a resonant-mode expansion of the Schrödinger flow in the weak sense of , and the description of the averaged wavefunction follows as a particular case; the proof is quite standard and is postponed to Section 6.1.
Corollary 6 (Consequences of resonance conjectures).
Let have Fourier transform supported in the compact set , with .
-
(i)
Under Conjecture Conjecture (LRC) — Local resonance conjecture, there holds in , for all ,
(2.6) where we have set for . More precisely, this means for all ,
In addition, the averaged wavefunction satisfies the following improved estimate,
-
(ii)
Under Conjecture Conjecture (GRC) — Global resonance conjecture, the same holds as in (i) with the errors and improved into and , respectively.
Moreover, for and , the restriction of the spectrum of to is absolutely continuous under Conjecture Conjecture (LRC) — Local resonance conjecture, and the whole spectrum is absolutely continuous under Conjecture Conjecture (GRC) — Global resonance conjecture. ∎
In order to make the above resonant-mode expansion (2.6) more striking, we note that resonances and resonant states can be computed explicitly in form of a perturbative Rayleigh–Schrödinger series. In particular, in agreement with Corollary 5, the resonance is checked to coincide to leading order with ; the proof is included in Section 6.2.
Proposition 7 (Approximate computation of resonances).
If Conjecture Conjecture (LRC) — Local resonance conjecture holds and if for all and the map
(2.7) |
is of class , then up to a gauge transformation there hold as , for all ,
where are defined in (2.3) and where and are given by
in terms of the projection onto . More precisely, the latter formulas are understood as follows, for all ,
Remark 2.2 (Full Rayleigh–Schrödinger series for resonances).
The proof of the above is easily pursued to any order. For and , if the map (2.7) is of class , then there hold as , for all ,
(2.8) |
where the coefficients are explicitly defined and can be checked to coincide with those of the formal Rayleigh–Schrödinger series for the perturbation of a bound state. This asymptotic series makes no sense in (in link with the dissolution of the bound state), but can be constructed in the weak sense of ; this partially answers in our setting a question raised in [21, p.179]. More precisely, for all , we can write
where the limits indeed exist and where for all the sequence is defined iteratively as follows: we set and for all we define as the unique solution of the regularized Rayleigh–Schrödinger recurrence equation,
The Rayleigh–Schrödinger series (2.8) is not known to be summable, hence cannot be used to actually construct resonances, which constitutes a reputed difficulty in this problem; see also [51, 18]. ∎
3. Stationary random setting and Floquet–Bloch fibration
In this section, we give a suitable definition of stationarity (or statistical translation-invariance) and we define the associated stationary differential calculus on the probability space, which was first introduced in [42] and plays a key role in the context of stochastic homogenization theory, e.g. [23, Section 7]. Next, we generalize the periodic Floquet–Bloch theory to this stationary setting, establishing in particular (1.3) and (1.4).
3.1. Stationary setting
Given a reference (complete) probability space , we start by recalling the classical notion of stationarity. In particular, a Gaussian field , that is, a family of Gaussian random variables, is an example of a stationary measurable random field if the variables have the same expectation and have covariance of the form with continuous at the origin.
Definition 3.1.
A random field on is a map such that for all the random variable is measurable. It is said to be stationary if its finite-dimensional law is shift-invariant, that is, if for any finite set the law of does not depend on the shift . In addition, it is said to be measurable if the map is jointly measurable. (In view of a result due to von Neumann [52], which can be viewed as a stochastic version of Lusin’s theorem, joint measurability is equivalent to requiring that for almost all and for all there holds as .) ∎
This basic notion of stationarity is usefully reformulated in terms of a measure-preserving action on the probability space, which draws the link with the theory of dynamical systems and ergodic theory.
Definition 3.2.
A measurable action of the group on is a collection of measurable maps that satisfy
-
—
for all ;
-
—
for all and measurable ;
-
—
the map is jointly measurable.
A random field is said to be -stationary if there exists a measurable map such that for all . ∎
This second definition yields a bijection between random variables and -stationary random fields . The random field is referred to as the -stationary extension of . In addition, given with , since there holds for any compact , we deduce that the realization belongs to for almost all . The Banach space can thus be identified with the subspace of -stationary random fields in .
While the notion of -stationarity in the sense of Definition 3.2 obviously implies measurability and stationarity in the sense of Definition 3.1, the following asserts that both are in fact essentially equivalent.
Lemma 3.3.
Let be a stationary measurable random field defined on in the sense of Definition 3.1. Then there exist a probability space , endowed with a measurable action , and a -stationary random field defined on in the sense of Definition 3.2 such that and have the same finite-dimensional law. This extends to a correspondence between -measurable random variables on and random variables on . ∎
Proof.
The proof is a variant of e.g. [31, Section 16.1]. Let denote the set of measurable functions , endowed with the cylindrical -algebra , and consider the map . This map is measurable and induces a probability measure on the measurable space . Next, define by . As is jointly measurable and stationary, we find that is a measurable action. Finally, we set , with -stationary extension , and the claim follows. We omit the details. ∎
Henceforth, we focus on the more convenient notion of -stationarity in the sense of Definition 3.2: we implicitly assume that the reference probability space is endowed with a given measurable action and we assume that the random potential is -stationary. In the sequel, for abbreviation, -stationarity is simply referred to as stationarity, and we abusively use the same notation for and (in particular, for and ).
Remarks 3.4.
-
(a)
A standard construction [42] allows to view periodic and quasiperiodic functions (as well as almost periodic functions) as instances of stationary random fields (with correlations that do not decay at infinity). In the periodic setting, the probability space is chosen as the torus endowed with the Lebesgue measure, the action is given by on , and we set . In the quasiperiodic setting, the probability space is chosen as a higher-dimensional torus with , endowed with the Lebesgue measure, the action is given by on in terms of the winding matrix , and we set . In both cases, the construction is viewed as introducing a uniform random shift.
-
(b)
Any -stationary random potential (that is, satisfying the stationarity assumption for an action of on ) can also be seen as a stationary random potential in the above sense up to considering the random ensemble of shifts. Indeed, assume that is a measurable action of on a probability space , and that is -stationary, that is, for all , , and . Endow with the product measure , where denotes the Lebesgue measure on , and define the action of on by
where for and where denotes the largest integer for . The map then defines a -stationary random field on . ∎
3.2. Stationary differential calculus
A differential calculus is naturally developed on via the measurable action on . Indeed, while the subspace of stationary random fields in is identified with the Hilbert space , the spatial weak gradient on locally square integrable functions turns into a densely defined linear operator on , which is referred to as the stationary gradient. Equivalently, can be viewed as the infinitesimal generator of the group of isometries on . The adjoint is and we denote by the corresponding stationary Laplacian. For all , we define the (Hilbert) space as the space of all elements for which the stationary extension belongs to , and we denote by the dual of . Note that coincides with the domain of , and that the stationary Laplacian is self-adjoint on . We refer e.g. to [23, Section 7] for details.
As opposed to the case of the periodic Laplacian on the torus, the stationary Laplacian on typically has absolutely continuous spectrum and no spectral gap above , cf. Section 4.1. This entails that Poincaré’s inequality does not hold on and that compact embeddings such as Rellich’s theorem also fail. This lack of compactness is related to the fact that the gradient only contains information on a finite set of directions while is typically an infinite product space.
3.3. Stationary Floquet transform
The usual periodic Floquet transform, e.g. [33], is a reformulation of Fourier series: given a function , its Floquet transform is (formally) defined by
which is periodic in , so that the Fourier inversion formula takes the form
thus leading to the following direct integral decomposition, e.g. [48, p.280],
This decomposition allows for a simple adaptation to the augmented space : given , its Floquet transform is defined by
which is periodic in , so that the Fourier inversion formula takes the form
and leads to the direct integral decomposition
We may now mimick this construction in the general stationary random setting: given , its stationary Floquet transform is defined by
so that the Fourier inversion formula takes the form
and leads to the direct integral decomposition
(3.1) |
This stationary Floquet transform was first introduced in [24, Section 3.2]; see also [4, 13]. Some key properties are collected in the following.
Lemma 3.5.
The stationary Floquet transform is a unitary operator on , and satisfies
-
(i)
for all ;
-
(ii)
for all and with .∎
3.4. Stationary Floquet–Bloch fibration
In view of (3.1), the stationary Floquet transform decomposes differential operators with stationary random coefficients (such as the Schrödinger operator ) into a direct integral of “elementary” fibered operators on the stationary space . First, the Laplacian on is self-adjoint on and is mapped by on
(3.2) |
in terms of the (centered) fibered Laplacian
As the stationary Laplacian is self-adjoint on and as is an infinitesimal perturbation, the Kato-Rellich theorem ensures that this fibered Laplacian is also self-adjoint on , and the centering ensures that constant functions belong to its kernel. Next, if the stationary random potential is uniformly bounded (the unbounded case is postponed to Appendix A), it defines a bounded multiplication operator on and the corresponding Schrödinger operator is thus self-adjoint on . Combining (3.2) with Lemma 3.5(ii), we find
(3.3) |
in terms of the (centered) fibered Schrödinger operator
which is self-adjoint on . Using direct integral representation, e.g. [48, p.280], we may reformulate the above as
(3.4) |
This decomposition of the Schrödinger operator yields a stationary version of the so-called Bloch wave decomposition of the Schrödinger flow: given a deterministic initial condition , appealing to (3.3) and to Lemma 3.5(i),
that is, (1.4). Alternatively, in terms of the -valued spectral measure of associated with the constant function ,
For vanishing disorder the spectral measures take the form and we recover the Fourier diagonalization of the free Schrödinger flow, cf. (1.5), while for each Fourier mode is deformed into a “Bloch measure” . In the periodic setting the measure is known to be discrete, leading to the deformation of the plane wave into a superposition of so-called Bloch waves, cf. [32, 3]. The picture is very different in the random setting as is rather expected to be absolutely continuous.
4. Basic spectral theory of fibered operators
This section is devoted to the proof of Propositions 1, 2, and 3. We consider general (non-Gaussian) stationary random potentials and discuss the fine dependence on the probabilistic structure. Note that our results could also be adapted to the random perturbation of a periodic Schrödinger operator, in which case fibered operators take the form , where the periodic potential models the underlying crystalline structure.
4.1. Unperturbed fibered operators
We give a full account of the spectral properties of the unperturbed operators on . We start with some general definitions. For , we denote its covariance function by , which belongs to and is positive definite. By Bochner’s theorem, the distributional Fourier transform is then a nonnegative finite measure on with total mass , and is called the spectral measure of . The set of all such spectral measures will play an important role in this section, so that we give it a name and notation.
Definition 4.1.
The spectrum of the probability space endowed with a given stationarity structure is defined as the subset
We show that the spectrum of the unperturbed operators can be completely characterized in terms of properties of .
Lemma 4.2.
Let be a stationary random field and assume that the underlying probability space is endowed with the -algebra generated by . For and we denote by the probability measure on defined by
(4.1) |
and we consider its Lebesgue decomposition
into pure point, singularly continuous, and absolutely continuous parts. Then,
-
(i)
The spectrum of the operator on is included in and there is an eigenvalue at .
-
(ii)
For , there holds
-
(iii)
The density of the absolutely continuous part of the spectral measure of associated with takes the form
Proof.
First note that the Fourier symbol of is given by , which easily implies that the operator has bounded inverse on for all . The spectrum of is therefore included in , which already proves item (i). We now wish to determine the different types of spectrum. For that purpose it suffices to proceed to the Lebesgue decomposition of the spectral measure of associated with any . We claim that this spectral measure is explicitly given by the following formula, for all ,
(4.2) |
where is defined in the statement. The conclusion directly follows from this claim since it yields for ,
where we denote by the Lebesgue decomposition of , and similarly for .
In particular, the above result implies that the spectrum can be of any type: for any measure with nontrivial pure point, singularly continuous, and absolutely continuous parts, we can construct a stationary Gaussian field such that the spectral measure coincides with , which entails that the corresponding spectrum of admits nontrivial pure point, singularly continuous, and absolutely continuous parts. Moreover, the eigenvalue at does not need to be simple in general.
In most cases of interest, the picture is however much neater: the spectrum of the fibered operator coincides with the whole interval and is made of a simple eigenvalue at embedded in absolutely continuous spectrum. This is proven to hold below either under strong structural assumptions (e.g. Gaussian structure) or under strong mixing assumptions (e.g. exponential decay of correlations, or integrable -mixing). We first recall some terminology: For any diameter and distance , we set
(4.3) |
where Rosenblatt’s -mixing coefficient is defined for any two sub--algebras as
The random field is said to be -mixing if for any there holds as . We may now state the following criterion, which in particular implies Proposition 1 when restricted to the Gaussian setting.
Proposition 4.3.
Let be a (nonzero) stationary random field with covariance , and let the probability space be endowed with the -algebra generated by .
-
(i)
Assume that one of the following two conditions holds,
-
(C1)
is Gaussian and is not periodic in any direction;
-
(C2)
has exponential decay, that is, for all .
Then the spectrum coincides with .
-
(C1)
-
(ii)
Assume that one of the following two conditions holds,
-
(C3)
is Gaussian and the (nonnegative measure) Fourier transform is absolutely continuous (this is the case for instance if is integrable);
-
(C4)
is -mixing and satisfies for all .
Then the eigenvalue at is simple (with eigenspace ) and
-
(C3)
Proof.
We split the proof into four steps, separately proving (i) and (ii) under conditions (C1), (C2), (C3), and (C4).
Step 1. Proof of (i) under condition (C1).
Since is Gaussian and centered, a repeated use of Wick’s formula yields for ,
hence, taking Fourier transform,
where denotes the th convolution power. As all terms in the sum are nonnegative, the support of therefore contains the support of , which coincides with the sum . We conclude
(4.4) |
As is nonzero and even, we note that is an additive subgroup of , so that its closure must be of the form for some linear subspace and some discrete additive subgroup . Since is supported in , if is not the whole of , we would deduce that is periodic in some direction, which is excluded by assumption. We conclude , hence . The definition (4.1) of then implies
and follows from Lemma 4.2.
Step 2. Proof of (i) under condition (C2).
The exponential decay condition entails that the Fourier transform extends holomorphically to the complex strip , and hence its support coincides with the whole of .
It then follows from (4.1) that the support of coincides with the whole interval , and therefore by Lemma 4.2.
Step 3. Proof of (ii) under condition (C3).
Recall the definition (2.4) of the set of -polynomials,
and let denote the subset of elements of with vanishing expectation. For , since is Gaussian, Wick’s formula allows to express explicitly as a linear combination of products of translated copies of the covariance function , without constant term. As the Fourier transform is assumed absolutely continuous and integrable, we conclude that is absolutely continuous and integrable as well for all . Lemma 4.2 then implies that for the spectral measure is absolutely continuous. In other words, the absolutely continuous subspace for contains .
It remains to check that is dense in . Given , by -measurability, we may approximate by a sequence in of the form for some Borel function on and some . Truncating and smoothening the Borel functions ’s, we find in of the form for some and . For each , Weierstrass’ approximation theorem then allows to replace the smooth function by a polynomial in variables. This proves that is dense in , hence is dense in .
Step 4. Proof of (ii) under condition (C4).
Arguing as in Step 3, it is enough to prove that the spectral measure is absolutely continuous for all of the form with , , and . Let .
Since is -measurable, the -mixing condition for yields,
cf. [11, Theorem 1.2.3],
The assumed integrability of the -mixing coefficient then yields , hence the nonnegative Fourier transform is absolutely continuous and belongs to . ∎
4.2. Perturbed fibered operators
We turn to the perturbed fibered operators and show that the spectrum of typically coincides with . The precise statement below is however quite intricate and depends on the structure of level sets of . This is to be compared with [43, Theorem 5.33] for the almost sure spectrum of on . Combined with Theorem A.1, this implies Proposition 2 in the Gaussian setting.
Proposition 4.4 (Spectrum of ).
Let be a stationary random field. Define the following two closed subsets of ,
Assume that satisfies the following weak mixing type condition: for all and the level set admits almost surely a bounded connected component containing a ball of radius . Then for all there holds
(4.5) | |||||
In particular, in the Gaussian setting with a nondegenerate stationary Gaussian field and with , we find , cf. [43, Theorem 5.34], hence . ∎
Remark 4.5.
The set is known as the essential range of and coincides with the spectrum of as a multiplication operator on . The set is a closed subset of and can be much smaller: in the periodic case , for instance, there holds unless is a constant. ∎
Proof of Proposition 4.4.
We split the proof into two steps, separately establishing the first and second inclusions in (4.5).
Step 1. Second inclusion in (4.5).
We only prove that , while the other inclusion is similar. If , the inclusion is trivial. It remains to consider the cases when has the form , , or , with . We focus on the case , while the other cases are easier. Without loss of generality we can assume , so that coincides with the (finite) operator norm of . Let . Since , we deduce that is invertible and we compute
Writing
and using Neumann series, we conclude that is invertible, which entails that .
Step 2. First inclusion in (4.5).
Given and , we show that there exists a sequence with such that and in , which entails . For , consider the open set and decompose it into its (at most countable) collection of connected components. Denote by the subcollection of bounded connected components. By assumption, this collection is almost surely nonempty. For all , we consider the balls included in with maximal radius. The maximum radius may be attained by different balls and we denote by the collection of their centers. As this collection is a closed bounded set in , we may choose as the first element in lexicographic order. The set defines a (nonempty) stationary point process on . Now choose a smooth cut-off function with for and for , and choose with . For , we define the random variable
(4.6) |
By assumption, the decimated stationary point process is also nonempty and we denote by its intensity. Since the remaining points in this process are all separated by a distance at least , the sum (4.6) defining contains at most one non-zero term, and we find
Next, we estimate
hence,
Finally, we compute , and the conclusion follows. ∎
4.3. Instability of the bound state
While the spectrum of the perturbed fibered operators was easily characterized in the previous section, determining its nature is substantially more involved. We recall the heuristic prediction from Fermi’s Golden Rule, e.g. [48, Section XII.6]. Given a perturbation of a self-adjoint operator on , if admits a simple eigenvalue at with normalized eigenvector , and if satisfies
(4.7) |
where denotes the orthogonal projection onto , then the eigenvalue at is expected to dissolve whenever the perturbation is turned on. The simplest rigorous version of this statement is as follows.
Lemma 4.6.
Let be two self-adjoint operators on a Hilbert space and let be a simple eigenvalue of with normalized eigenvector . If for some there exists a branch of class with
then there holds
(4.8) |
where is the orthogonal projection onto . In particular, if the right-hand side of (4.8) is not real, then there exists no such branch . This is in particular the case whenever the spectral measure of associated with is absolutely continuous in a neighborhood of and has non-vanishing density at . ∎
Proof.
Assume that there exists a branch as in the statement and denote by and the first and second derivatives with respect to at . Differentiating the eigenvalue relation yields
Taking the scalar product with , we find
(4.9) |
hence
This can be inverted in the form
(4.10) |
Now differentiating the eigenvalue equation twice, we find
hence, injecting (4.9) and taking the scalar product with ,
Injecting (4.10) then yields
Since is simple, we find , hence
and the conclusion follows. ∎
5. Perturbative Mourre’s commutator approach
This section is devoted to the use of Mourre’s theory [37, 1] to study fibered perturbation problems, in particular proving Theorem 4 and Corollary 5. We focus on the short-range Gaussian setting, that is, for some and some stationary centered Gaussian field with covariance function , and without loss of generality we assume that the probability space is endowed with the -algebra generated by .
5.1. Reminder on Mourre’s theory
We briefly recall the general purpose of Mourre’s theory and its classical application to Schrödinger operators on ; we refer to [37, 1] for details. A self-adjoint operator with domain on a Hilbert space is said to satisfy a Mourre relation on an interval with respect to a (self-adjoint) conjugate operator with domain if there exists and a compact operator such that there holds in the sense of forms,
(5.1) |
where the commutator is defined as a sesquilinear form on . The Mourre relation (5.1) is said to be strict if . For technical reasons, one typically requires that the domain of be invariant under the unitary group generated by , that is,
(5.2) |
which in particular ensures that is dense in , and one further requires to be -bounded. In that case, the sesquilinear form on automatically extends to the form of a unique -bounded self-adjoint operator.
In a semiclassical perspective, conjugate operators can be viewed as a quantum analogue of escape functions for Hamiltonian dynamical systems. The main result of Mourre’s theory [37, 1, 20] is that the relation (5.1) (together with additional regularity assumptions) entails that the eigenvalues of in have finite multiplicity and that has no singular continuous spectrum in . In addition, a strict Mourre relation implies that the spectral measure is absolutely continuous on . This is actually a simple consequence of the virial theorem: if was an eigenvalue in with normalized eigenvector , then a strict Mourre inequality would formally yield
a contradiction. Alternatively, this short formal proof can be rewritten by noting that a strict Mourre relation implies ballistic transport for the flow with respect to the conjugate operator : for there holds
hence , thus prohibiting from being an eigenvector. In addition to such spectral information, the Mourre relation (5.1) is further known to yield useful a priori estimates on boundary values of the resolvent in form of limiting absorption principles [37, 22].
We recall the standard construction of a Mourre conjugate operator for Schrödinger operators on , e.g. [37]. Considering the unitary group of dilations on , and noting that , we deduce by differentiation,
where denotes the generator of dilations, that is, on . This implies that satisfies a strict Mourre inequality on for all with conjugate operator ,
In a semiclassical perspective, the conjugate operator corresponds to the escape function for the free Hamiltonian . Next, given a -bounded potential , we compute
so that the commutator is bounded whenever the function is bounded. For small enough, this easily entails that the Schrödinger operator on also satisfies a strict Mourre inequality on . This follows from the first general property below and illustrates the flexibility of Mourre’s theory.
Lemma 5.1.
Let be a self-adjoint operator on a Hilbert space , assume that satisfies a Mourre relation (5.1) on a bounded interval with respect to a conjugate operator , that the domain of is invariant under , cf. (5.2), and that is -bounded. Let also be symmetric and -bounded.
-
(i)
Mourre relation under perturbation:
If the commutator is -bounded, then for all and small enough the perturbed operator satisfies a Mourre relation on with respect to . In addition, if satisfies a strict Mourre relation, then does too. -
(ii)
Strict relation on orthogonal complement of an eigenspace:
If has an eigenvalue with eigenprojector , then there exists a neighborhood of such that the restriction of to the range of satisfies a strict Mourre relation on with conjugate operator . ∎
Proof.
We start with the proof of (i). As the perturbation is -bounded, the operator has the same domain as for small enough in view of the Kato-Rellich theorem, hence by assumption its domain is invariant under . Furthermore, the commutator is -bounded, hence -bounded. Now, choose with . Multiplying by both sides of the Mourre relation for yields
As is -bounded, we deduce
Noting that the -boundedness of implies , and further using the -boundedness of , we deduce
Multiplying both sides by , the conclusion (i) follows for .
We turn to the proof of (ii). As commutes with , multiplying by both sides of the Mourre relation for yields, on the range of ,
in terms of , , . Multiplying both sides with , the compact operator is replaced by . Since converges strongly to on the range of as , the conclusion (ii) follows. ∎
Next, we state the following general result by Cattaneo, Graf, and Hunziker [8], showing how Mourre’s theory can be exploited to analyze the instability of embedded bound states in form of an approximate resonance theory; see also [41, 21, 50, 36, 10]. Although Mourre’s theory does not allow to deduce the existence of resonances in any strong sense, it is shown to have essentially the same dynamical consequences. The proof further allows for asymptotic expansions to finer accuracy in , as well as for a description of the flow projected on a whole class of “smooth” states rather than on only, but such improvements are not pursued here.
Theorem 5.2 (Dynamical resonances from Mourre’s theory; [8]).
Let be a self-adjoint operator on a Hilbert space , let be symmetric and -bounded, and consider the perturbation . Let be a simple eigenvalue of with normalized eigenvector , and assume that the following properties hold:
-
The iterated commutators and are -bounded for , where iterated commutators are defined by and recursively for .
-
Fermi’s condition (4.7) is satisfied, that is,
where denotes the orthogonal projection onto and where we have set for abbreviation .
Then there exists with such that for all neighborhoods of there holds for all ,
where the dynamical resonance satisfies
Idea of the proof.
We include a brief summary of the proof for the reader’s convenience, and refer to [8] for full details. Starting point is the following Feshbach-Schur complement formula for the resolvent, for , in terms of the restriction ,
Next, recall that Lemma 5.1 ensures that the restriction on the range of satisfies a strict Mourre relation close to . In view of [22], together with the -boundedness of iterated commutators and for , this strict Mourre relation implies the -smoothness of boundary values on of the resolvent
Inserting a Taylor expansion for the latter in the above Feshbach-Schur complement formula, we construct an approximate meromorphic extension for . The conclusion then follows from complex deformation techniques similarly as for true resonances as in Section 6.1. ∎
Remark 5.3.
As it is clear from the proof, cf. [8], we mention for later reference that a similar result holds if and further depend on . More precisely, assume for all that is a simple eigenvalue of with normalized eigenvector (independent of , say), that the restriction of on the range of satisfies a strict Mourre relation on a neighborhood of with conjugate operator and constant (independent of ), that the domain of is invariant under , and that iterated commutators are -bounded by for . Next, assume that the perturbation is bounded in the sense of , and that iterated commutators are -bounded and small enough in the sense that for and ,
for some large enough constant only depending on . Then the same result holds as in Theorem 5.2 above for the perturbed operator . ∎
5.2. Reminder on Malliavin calculus
We recall some notation and tools from Malliavin calculus for the fine analysis of nonlinear functionals of the underlying Gaussian field with covariance function ; we refer to [35, 40, 39] for details. We start by underlining the Hilbert structure associated with the Gaussian field . The random variables with are centered Gaussians with covariance
We consider the completion of endowed with the (semi)norm
and we denote by the quotient of this completed space with respect to the kernel of . The normed space is a separable Hilbert space and the random field satisfies the isometry relation
The map then extends as a linear isometric embedding and constitutes a so-called isonormal Gaussian process over . The structure of is conveniently characterized as follows: as , the (nonnegative measure) Fourier transform is absolutely continuous, hence the square root belongs to and the linear map
(5.3) |
extends into a unitary transformation . Note that for all the Dirac mass is (a representative of) an element of with . By definition, the linear isometric embedding is a white noise.
As a model dense subspace of , instead of considering the linear subspace of -polynomials, cf. (2.4), we define the following slightly more convenient subspace,
(5.4) |
Recall that we implicitly assume that the underlying probability space is endowed with the minimal -algebra generated by , thus ensuring that is indeed dense in . This allows to define operators and prove properties on the simpler subspace in a concrete way before extending them to by density.
For a random variable , say , we define its Malliavin derivative as
and similarly, for all , its th Malliavin derivative is given by
Note that by definition this belongs to the symmetric tensor product, . These operators on are closable on . We then set
we define the Malliavin-Sobolev space as the closure of for this norm, and we extend the th Malliavin derivative by density to this space.
Next, we define the corresponding divergence operator as the adjoint of the Malliavin derivative , and similarly, for all , the th-order divergence operator as the adjoint of . In other words, this is defined by the following integration by parts formula, for all and ,
The so-called Meyer inequalities ensure that the th divergence operator extends as a bounded operator for all , hence in particular its domain contains . For and , a direct computation yields
Due to this relation, with in particular , the divergence operator is sometimes referred to as the Skorokhod integral; see also the notion of multiple integrals below.
With the Malliavin derivative and the divergence operator at hand, we may construct the corresponding Ornstein–Uhlenbeck operator (or infinite-dimensional Laplacian)
(5.5) |
as a self-adjoint operator acting on with domain . The spectrum of is and its kernel coincides with constants. In particular, the following Poincaré inequality holds: for all with ,
This ensures the invertibility of the restriction of to , and allows to define a pseudo-inverse on in terms of the projection .
We turn to a spectral decomposition of . For that purpose, for , we first define the th multiple integral as the bounded linear operator given by the restriction of the th divergence operator, that is, for all . Alternatively, can be characterized as follows: for all with there holds
where denotes the th Hermite polynomial, that is, . The image of is known as the th Wiener chaos . Properties of Hermite polynomials easily imply the following orthogonality property: for all and ,
(5.6) |
which in particular entails that is a unitary transformation , where the symmetric tensor product is endowed with the norm
In view of (5.3), recall that is further isometric to , endowed with the norm
so that we are led to the following unitary transformations,
(5.7) |
and we write for abbreviation
As a consequence of the orthogonality property (5.6), the following Wiener chaos expansion holds in form of a (bosonic) Fock space decomposition,
(5.8) |
More precisely, for all , we can expand
for some unique collection of kernels or , where the expansion is converging in . The Stroock formula asserts
It can be checked that the th Wiener chaos coincides with the eigenspace of the Ornstein–Uhlenbeck operator associated with the eigenvalue , so that the Wiener chaos expansion (5.8) coincides with the spectral decomposition of .
5.3. A new class of operators on
This section is devoted to a general construction allowing to transfer operators on into operators on , which will be a key tool in the sequel and is analogous to second quantization in quantum field theory. Given a bounded operator on , for all , we denote by the bounded operator on given by
Via the isomorphism (5.7), we can then construct a bounded operator on the th Wiener chaos via
where is the adjoint of on . In particular, on the first chaos, this definition formally yields . Via the Wiener chaos decomposition (5.8), we then let denote the densely defined operator on given by the direct sum
As is obviously -bounded for bounded , the map provides a linear embedding , but this is however not a group homomorphism as in particular .
If is bounded and self-adjoint, then is self-adjoint on . More generally, if is unbounded on and essentially self-adjoint on some subset , then defines an essentially self-adjoint operator on , hence is essentially self-adjoint on , cf. [47, Theorem VIII.33], and in turn defines an essentially self-adjoint operator on
In particular, noting that the definition (5.4) of can be reformulated as
we deduce that if is essentially self-adjoint on , then is essentially self-adjoint on . Similarly, if leaves invariant, then leaves invariant. Also note that the operators and strongly commute since acts as on and since preserves the chaos decomposition.
Next, the following shows that the stationary gradient corresponds to the spatial gradient via this embedding .
Lemma 5.4.
There holds on . In particular, preserves the chaos decomposition and commutes strongly with . ∎
Proof.
Given , we compute
Noting that and recalling that Hermite polynomials satisfy , we deduce
(5.9) |
Next, we appeal to the following useful product formula for multiple integrals (see e.g. [39, Section 2.7.3] for a more general statement): for all and ,
(5.10) |
where we have set
Inserting this formula into (5.9), we find
Since for real-valued , the second right-hand side term vanishes and we are led to
In addition, this formula ensures that preserves the chaos decomposition. ∎
Given a self-adjoint operator on , the operator on is also self-adjoint and we may consider the corresponding unitary -groups. If preserves the real part, then the group on is shown to admit an explicit description.
Lemma 5.5.
Let be essentially self-adjoint on , and assume that the subset of real-valued functions is invariant under . Then the operator generates a unitary -group on , which has the following explicit action: for all , say ,
In particular, this entails that is multiplicative, that is, for all ,
which implies that is a derivation, that is, for all ,
Proof.
Denote by the group of operators defined on as in the statement: for all , say with and ,
hence in particular, for all ,
(5.11) |
This is well-defined since is assumed to preserve . (Note that may of course no longer have any representative in in its equivalence class in .) Noting that
and using again the assumption that preserves , we deduce that the (Gaussian) law of is invariant with respect to , hence for all and ,
This allows to extend by density as a unitary group on . In addition, as is strongly continuous on , it is easily deduced that is strongly continuous on . We denote by its skew-adjoint generator on . Differentiating (5.11) with respect to shows that the domain of contains and that on . Since is essentially self-adjoint on , is essentially self-adjoint on , and we conclude , hence . ∎
In view of the application to Mourre’s theory for Schrödinger operators, cf. Section 5.1, we recall the definition of the unitary -group of dilations on , and its generator . We then define the self-adjoint operator
(5.12) |
and the associated unitary -group given by . Due to Lemma 5.5, this satisfies, for all , say ,
(5.13) |
which entails in particular that the spaces are invariant under for all .
5.4. Chaos decomposition of fibered operators
While the unperturbed operators preserve the chaos decomposition, cf. Lemma 5.4, we show that the random potential amounts to shifting the chaoses, thus playing the role of annihilation and creation operators on the Fock space decomposition. This structure of the perturbed operators can be viewed as drawing some surprising link between random Schrödinger operators and multi-particle quantum systems.
Lemma 5.6.
Assume for simplicity that is itself a Gaussian field. Via the Wiener chaos decomposition (5.8), for all , the perturbed fibered operator on is unitarily equivalent to in terms of
in terms of
and in terms of the annihilation and creation operators
which are defined as follows for all and ,
For all , the operators and are bounded and adjoint, with
and thus the operators and are -bounded and are adjoint in particular on . In addition, they satisfy the commutator relation on . ∎
5.5. Some Mourre relations on
Drawing inspiration from the construction of Mourre conjugates for Schrödinger operators on , cf. Section 5.1, we show in item (i) below that the generator of dilations on as constructed in Section 5.3 is a conjugate for the stationary Laplacian . Nevertheless, item (iii) indicates that the perturbation by the random potential is never compatible in the sense of Mourre’s theory for this conjugate operator, which prohibits to deduce any Mourre relation for perturbed operators of the form . In spite of this, the incompatibility is shown to be comparable to the lack of boundedness of the underlying Gaussian field in the sense that it is bounded on any fixed chaos and -bounded on . Finally, in item (iv), we show that the action of the random potential as described in Lemma 5.6 on the Fock space allows to associate a natural conjugate. In other words, the stationary Laplacian describes diffusion on each chaos and the random potential describes shifts between chaoses: the transport properties of both parts are well understood and natural conjugates can be constructed for both, cf. items (i) and (iv), but the construction of a conjugate for appears particularly difficult and is left as an open problem. In a semiclassical perspective, this is related to the construction of escape functions for the random acceleration model [29, 30].
Proposition 5.7 (Some Mourre relations).
-
(i)
Conjugate operator for :
The generator of dilations on , cf. (5.12), satisfiesand the domain of is invariant under .
-
(ii)
Conjugate operators for :
-
(iii)
Incompatibility of the perturbation:
The commutator is well-defined and essentially self-adjoint on , but only -bounded provided . If is itself Gaussian, then similarly is well-defined and essentially self-adjoint on , but only -bounded provided . -
(iv)
Conjugate operator for the perturbation:
If is itself Gaussian, then there holds
Proof.
Since the operators , , , , are essentially self-adjoint on , and since is invariant under these operators, the commutators , , , , are clearly well-defined on and are explicitly computed below on that linear subspace. We split the proof into five steps.
Step 1. Proof of (i).
Given and , recalling the notation , Lemmas 5.4 and 5.5 lead to
and hence, by definition of and of ,
so that differentiating in yields
and similarly
Step 2. Proof of (ii).
The computation of the commutator follows from Step 1, and it remains to compute the other one.
For , Lemma 5.4 yields
while the definition of leads to , hence
Step 3. Proof of (iv).
Given , for and , Lemma 5.6 yields
(5.14) |
so that the commutator with takes the form,
A direct computation of the commutator of these two operators then yields after straightforward simplifications,
hence, in view of Lemma 5.6,
Step 4. Proof of (iii) for .
In view of (5.13), with , we find
and differentiating in yields
Noting that preserves the real part and that
we deduce that and are independent Gaussian random variables. Further note that cannot be degenerate: indeed, by definition of as generator of dilations, can only vanish if , which is not compatible with . We may then deduce
proving that is unbounded.
5.6. Mourre relations for fibered operators
This section is devoted to the construction of conjugates for the unperturbed fibered operators . This appears to be surprisingly more involved than for , as the group of dilations is no longer adapted and must be suitably deformed. Noting that bounds on iterated commutators are obtained similarly and that the dense subspaces and are exchangeable, the conclusion of Theorem 4 directly follows upon truncation.
Theorem 5.8 (Mourre relations for fibered operators).
-
(i)
Conjugate operator for :
For all , there exists a self-adjoint operator on , essentially self-adjoint on , such that the commutator is well-defined and essentially self-adjoint on , is -bounded, and satisfiesHence, the fibered operator satisfies a Mourre relation on with respect to , for all ,
In addition, the domain of is invariant under .
-
(ii)
Incompatibility of the perturbation:
If is itself Gaussian, then the commutator is well-defined and essentially self-adjoint on , but only -bounded provided that .
In addition, the full commutator is well-defined and essentially self-adjoint on for all , but it only satisfies the lower bound
which does not yield any Mourre relation. ∎
Before turning to the proof, we briefly underline the difficulty and explain the idea behind the construction. Proposition 5.7 (i)–(ii) leads to
which shows that the generator of dilations should be properly modified. Again drawing inspiration from the situation on the physical space, noting that on there holds
with , we consider
(5.15) |
and a similar computation as in Proposition 5.7 (i)–(ii) yields on ,
(Recall that and commute, cf. Lemma 5.4.) In order to counter the apparition of factors in this relation and obtain a proper Mourre relation, a further modification of is thus needed. More precisely, in the definition (5.15) of , the generator of dilations is a suitable conjugate for the stationary Laplacian , while is supposed to take into account the additional first-order contribution in the fibered operator . The core of the problem then lies in the factor that appears in the commutator
(5.16) |
which is related to the infinite dimensionality of the probability space. The simplest way to solve this problem would be defining
(5.17) |
where we recall that denotes the pseudo-inverse of . This indeed leads to the desired Mourre relation,
(5.18) |
However, the perturbation behaves particularly badly with respect to this conjugate operator in the sense that the commutator is not even bounded when restricted to any fixed Wiener chaos, thus excluding any meaningful use of such a relation. While on the th Wiener chaos the operator amounts to the sum , the choice (5.17) consists of rather considering the algebraic mean . Another possible choice to avoid the factor in the commutator (5.16) is to use an -norm of . We show in the following paragraphs that the latter choice has all the desired properties claimed in Theorem 5.8: it still yields a similar Mourre relation as in (5.18) and its commutator with the perturbation is -bounded.
We construct the desired conjugate operator via its action on the Fock space decomposition (5.8). As we are concerned with the suitable treatment of the first-order operator , that is, the stationary derivative in the direction , we set and first focus on the case of dimension . For all , define the function ,
with the convention . This function is clearly symmetric with respect to the variables and has the following main properties.
Lemma 5.9.
For all , the function is well-defined and is continuous on , where denotes the hypersurface
In addition, there exists a continuous function such that
(5.19) |
Proof.
The continuity of is clear outside the zero locus of , and we turn to the second part of the statement. On there holds with , hence . It remains to examine the jump of on . We claim that every line directed by the vector in intersects the hypersurface at a single point, and this would yield the conclusion. Indeed, given a point , say and , we can write with and , and belongs to only if . ∎
Next, in order to get a proper Mourre relation on the Fock space, we regularize the functions so as to replace the Dirac part in (5.19) by a positive bump function that is -uniformly bounded on . For that purpose, it is not enough to regularize the sign function in the definition of in a fixed neighborhood of , as the derivative would still produce an unbounded term due to the multiplication by . A suitable choice of the regularization is rather defined as follows. First rewrite
where only the last sign function needs to be regularized. Choose a smooth odd function such that for , for , pointwise, for , and for . We then set
which is globally well-defined and continuous. Note that
(5.20) | |||
(5.21) |
We also establish the following important property.
Lemma 5.10.
For all , there holds for all ,
Proof.
The conclusion follows from (5.21) if , so that we can henceforth assume . By symmetry we can assume and . In the case , we find
and the conclusion follows. It remains to treat the case , while the symmetric case is similar. Given , the assumption implies . As , we compute
Noting that and that there holds whenever , the above becomes, in view of the properties of ,
as claimed. ∎
We now turn to the construction of the suitable conjugate operator for . For all , we define an operator on as the multiplication by the function , and we denote by the corresponding operator on the Fock space. Next, we define the operator on the th Wiener chaos by
and via the Wiener chaos decomposition (5.8) we set on . We then consider the following operator on ,
(5.22) |
which is clearly essentially self-adjoint on given its action on Wiener chaoses; see also Lemma 5.12 below.
Remark 5.11.
The reader may wonder why this definition of is chosen instead of , which would seem more natural in view of (5.15). The computation of the relevant commutators involves , hence the derivative , which in view of the regularization of is not uniformly equal to but can vary in the whole interval (or at best in for some smaller if the cut-off function is chosen with closer to ). Due to this modification, symbols are deformed in the commutator computation, and the choice would fail at providing a Mourre relation close to . This is precisely corrected by the above choice (5.22). ∎
We first show that the operator generates an explicit unitary -group, which preserves .
Lemma 5.12.
The operator is essentially self-adjoint on and its closure generates a unitary -group on , which has the following explicit action on chaoses: for all and ,
where is defined by
In particular, in view of (5.20), for all , the subspace is invariant under this group action . ∎
Proof.
In view of the chaos decomposition (5.8), it suffices to check that for all the family defines a unitary -group on and that its self-adjoint generator is given by on . First note that the family clearly defines a unitary group on . Next, for all , we decompose
As , for , the first right-hand side term converges to in , while the second one converges to , and the claim easily follows. ∎
Next, we show that is a conjugate for the fibered operator away from the bottom of the spectrum. This completes the proof of Theorem 5.8(i). Choosing the cut-off function with closer to , and suitably increasing the factor in definition (5.22), the term in the lower bound (5.23) below could be improved into almost , but the present construction does not allow to reach a value any closer to .
Lemma 5.13.
The commutator is well-defined and essentially self-adjoint on , is -bounded, and satisfies the lower bound
(5.23) |
which entails the following Mourre relation on , for all ,
Proof.
For all , we define the operator on as the multiplication by the function , we denote by the corresponding operator defined on the th Wiener chaos by , and we set on . A direct computation on Wiener chaoses yields
Combining this with Proposition 5.7 (i)–(ii), we easily find on ,
(5.24) | |||||
where the right-hand side is well-defined and symmetric on . We split the rest of the proof into two steps.
Step 1. Proof of the lower bound (5.23) on .
Note that constants belong to the kernel of the commutator .
The above expression (5.24) for the latter yields for all with ,
The bound (5.20) implies on , hence
and we are led to
that is, the lower bound (5.23).
Step 2. Proof that the commutator is essentially self-adjoint on and that its closure is -bounded and self-adjoint on its domain .
For all , we define the operator on as the multiplication by the function , we denote by the corresponding operator defined on the th Wiener chaos by , and we set on . A direct computation on Wiener chaoses yields
hence the expression (5.24) can be rewritten as
(5.25) |
Note that the bound (5.20) ensures that is bounded by and that is bounded on , hence
Together with (5.25), this shows that the commutator is an infinitesimal perturbation of , and the conclusion follows from the Kato-Rellich theorem. ∎
We turn to the proof of Theorem 5.8(ii), that is, the incompatibility of the perturbation with respect to the above-constructed conjugate operator . In view of Proposition 5.7(iii), it remains to establish the following.
Lemma 5.14.
If is itself Gaussian, the commutator is well-defined and essentially self-adjoint on , but is only -bounded provided that . ∎
Proof.
Finally, we argue that, while well-defined and symmetric on , the full commutator is essentially self-adjoint.
Lemma 5.15.
If is itself Gaussian, the commutator is well-defined and essentially self-adjoint on . ∎
This is obtained as a particular case of the following abstract result, which is a convenient reformulation of Nelson’s theorem [38]. Note that this result also ensures that is essentially self-adjoint on when is Gaussian; a more general criterion for essential self-adjointness of in case of an unbounded potential without particular Gaussian structure is included in Appendix A.
Proposition 5.16.
Let and be self-adjoint operators on their respective domains and on a Hilbert space , and let be a symmetric operator defined on some dense linear subspace , such that
-
and are nonnegative and commute strongly, hence is self-adjoint on ;
-
is a Kato perturbation of , that is, is a core for and there is and such that there holds, for all ,
-
, that is, for all ,
Then the operator is essentially self-adjoint on . ∎
Proof.
We split the proof into two steps.
Step 1. Proof that for large enough the operator is essentially self-adjoint on and satisfies .
As and are self-adjoint on their respective domains and nonnegative and as they commute strongly, their sum is self-adjoint on , cf. [46, Lemma 4.15.1]. Next, as is a Kato perturbation of , it is also a Kato perturbation of for all , hence the Kato-Rellich theorem entails that is essentially self-adjoint on . In addition, as is a Kato perturbation of and as is nonnegative, it is easily deduced that is bounded from below, hence for large enough.
Step 2. Conclusion.
In view of Nelson’s theorem [38] in form of [19, Corollary 1.1], together with the result of Step 1, the conclusion follows provided that we can check the following two additional properties, for large enough,
-
(i)
is -bounded on ;
-
(ii)
, that is, for all ,
We start with the proof of condition (i). On the one hand, since and are nonnegative and commute strongly, we can deduce for all . On the other hand, since is a Kato perturbation of , we find for all ,
hence
which leads to
Combined with the above, this yields
hence in particular,
that is, (i). It remains to establish condition (ii). As and commute, we can write for all ,
hence, by assumption,
Again noting as in Step 1 that the nonnegativity of and the fact that is a Kato perturbation of imply that is nonnegative for large enough, the claim (ii) follows. ∎
With the above abstract result at hand, we quickly indicate how Lemma 5.15 is a simple consequence.
Proof of Lemma 5.15.
In view of (5.24), we can decompose on , in terms of
We shall appeal to Proposition 5.16 with , and it remains to check the different assumptions. First, and are both essentially self-adjoint on , as discussed in Sections 3.4 and 5.2, respectively, they are clearly nonnegative, and Lemma 5.4 ensures that they commute strongly. Also note that and leave the linear subspace invariant and that spectral projections of also leave invariant. Using projections onto a finite number of chaoses, one then easily sees that is essentially self-adjoint on . Next, we may rewrite as in (5.25),
(5.26) |
hence the boundedness of and leads to
showing that is -bounded. Proposition 5.7(iii) and Lemma 5.14 also ensure that and are -bounded. This proves that is -bounded, hence -infinitesimal. Finally, it remains to consider the commutator . Since by definition and preserve chaoses, identity (5.26) implies . In view of the explicit description of and in Proposition 5.7(iii) and in Lemma 5.14, respectively, as these commutators have a similar structure as itself on the Wiener chaos decomposition, a similar computation as in the proof of Proposition 5.7(iv) easily shows that is -bounded, and the conclusion follows. ∎
5.7. Consequences of Mourre’s relations
This section is devoted to the proof of Corollary 5. Let be a stationary Gaussian field. Given some , consider the projection onto Wiener chaoses of order . We split the proof into three steps.
Step 1. Preliminary on spectral trunctations: For all , , and supported in , there holds
(5.27) |
where the factor can be replaced by if .
Choose with for and for such that and pointwise. By definition, we find , hence
Using spectral calculus with and , this leads to
If is nonconstant, the conclusion directly follows from the Cauchy–Schwarz inequality. If , decomposing into its positive and negative parts and , we can rather estimate
Step 2. Proof that for all the flow for the truncated fibered Schrödinger operator satisfies for all ,
(5.28) |
We wish to apply Theorem 5.2 in form of Remark 5.3 with , , , , and it suffices to check the different assumptions. First, as the conjugate operator commutes with the Wiener chaos decomposition, Theorem 5.8(i) ensures that the truncated operator restricted to satisfies a strict Mourre relation on for all : we find
hence, on ,
Similar computations as in the proof of Theorem 5.8(i) show the -boundedness of iterated commutators , hence the -uniform -boundedness of
In addition, as the domain of is , as Theorem 5.8(i) ensures that is invariant under the unitary group generated by , and as the latter further preserves the chaos decomposition, we deduce that the domain of is also invariant under the unitary group generated by .
It remains to check assumptions on the perturbation . Note that , which clearly satisfies . Further, iterating the proof of Theorem 5.8(ii), we find that for all iterated commutators are -bounded, hence for all ,
which entails, by definition of ,
Choosing large enough, we may then apply Theorem 5.2 in form of Remark 5.3, to the effect of
In view of Step 1, the spectral truncation can be removed up to a further error, and the claim (5.28) follows.
Step 3. Conclusion.
In view of the result of Step 1, it remains to prove for all ,
(5.29) |
while the conclusion of Corollary 5 then follows from the fibration (1.4). Set for abbreviation and . Since the flow satisfies the equation , an iterative use of Duhamel’s formula allows to decompose, for all ,
in terms of
Noting that for all , we deduce that the truncation error can be represented as follows, for any ,
(5.30) |
where is defined similarly as with and replaced by and , respectively. A direct estimate yields
hence, noting that is bounded by on , cf. Lemma 5.6,
Similarly estimating and inserting this into (5.30), we find for all ,
Setting and choosing , we easily deduce for ,
and the claim (5.29) follows. ∎
6. Exact resonance conjectures and consequences
This section is devoted to the proof of Corollary 6 and Proposition 7 as consequences of the resonance conjectures Conjecture (LRC) — Local resonance conjecture and Conjecture (GRC) — Global resonance conjecture.
6.1. Resonant-mode expansion
We start with the proof of Corollary 6. For , the Floquet–Bloch fibration (1.4) takes the form
hence it suffices analyze for fixed and . We split the proof into three steps, separately establishing items (i) and (ii).
Step 1. Meromorphic extension of the spectral measure: Under Conjecture (LRC) — Local resonance conjecture, for all , the spectral measure is analytic on and admits a local meromorphic extension on the complex neighborhood ,
(6.1) |
which can alternatively be expressed as
and moreover, in case ,
where we write for abbreviation
(6.4) | |||||
Indeed, Stone’s formula together with Conjecture Conjecture (LRC) — Local resonance conjecture yields for ,
and (6.1) follows. Identities (6.1) and (6.1) are obvious consequences as .
Step 2. Proof of (i): Under Conjecture (LRC) — Local resonance conjecture, for all , , , , and supported in with in , there holds
where the factor in the right-hand side can be replaced by in case . Combined with (5.27) to remove the spectral truncation, this indeed yields (i).
Starting from
and using formula (6.1) for the meromorphic extension of the spectral measure , we obtain by contour deformation,
where the smooth path is a deformation of the real interval in the lower half-plane such that remains on the real axis on while the part on is deformed into a path in that stays pointwise at a distance from the origin. Using the identity and integrating by parts, we find iteratively for all ,
As the remainder is holomorphic on and has continuous dependence on for , cf. Conjecture (LRC) — Local resonance conjecture, we deduce that all its derivatives are bounded on uniformly with respect to . Hence, it follows from (6.1) that all derivatives of are uniformly bounded by on the path , and thus
(6.5) |
where in view of (6.1) the factor can be replaced by in case .
Step 3. Proof of (ii): Under Conjecture (GRC) — Global resonance conjecture, for all , , and ,
where the factor in the right-hand side can be replaced by in case .
Similarly as in Step 2, applying formula (6.1) and contour deformation, we find
where the smooth path is given by with
-
connects and , does not exit the ball of radius , and always stays at a distance from the origin;
-
and ;
-
connects and without crossing the real axis;
-
and ;
where is to be fixed later. Inserting formula (6.1) for , using the uniform bound assumed to hold on , cf. Conjecture (GRC) — Global resonance conjecture, and setting , the above turns into
Hence, by definition of ,
Optimization in yields the bound . Interpolating this with the result (6.5) of Step 3 under Conjecture (LRC) — Local resonance conjecture, the conclusion follows. ∎
6.2. Computing resonances
We turn to the proof of Proposition 7. Assuming that for all the map
is of class , we iteratively compute the first two derivatives,
Note that the resonant and co-resonant states are only defined up to multiplication by for any complex-valued function , cf. (2.5). When differentiating, this gauge invariance implies that is only defined up to multiplication by for any , next is defined up to addition of for any , and next is defined up to addition of for any .
We first compute . The resonance conjecture Conjecture (LRC) — Local resonance conjecture yields for ,
(6.6) |
with holomorphic on . Setting and , we find for ,
and similarly, exchanging the roles of and ,
For , we deduce
By gauge symmetry, as explained above, we can e.g. choose ,
(6.7) |
Next, we compute . Differentiating identity (6.6) at , using (6.7), and choosing , we find for ,
and similarly, exchanging the roles of and ,
Choosing with , we easily deduce and
Noting that the left-hand side vanishes for , we are led to
for some . By gauge symmetry, as explained above, we can e.g. choose ,
(6.8) |
Finally, we turn to the second derivatives . Differentiating identity (6.6) twice at , using (6.7) and (6.8), and choosing , we find for ,
and similarly, exchanging the roles of and ,
Choosing with , and using again the gauge invariance, we easily deduce
in terms of the projection onto . This completes the proof. ∎
7. An illustrative toy model
In this last section, we display a toy model that shares many spectral features of Schrödinger operators, but is explicitly solvable and allows for a rigorous study of its spectrum and resonances, illustrating the relevance of the resonance conjectures Conjecture (LRC) — Local resonance conjecture and Conjecture (GRC) — Global resonance conjecture. More precisely, we replace the free Schrödinger operator by
and we consider the corresponding perturbed operator
on . Via the Floquet–Bloch fibration (1.3), this operator is decomposed as
(7.1) |
in terms of the following (centered) fibered operator on the stationary space ,
For this toy model, we establish the following detailed spectral properties, which are in perfect analogy with the expected situation in the Schrödinger case. Note however that the energy transport remains ballistic, cf. item (v) below, in stark contrast with the quantum diffusion in the Schrödinger case: this could be related to the fact that the centered fibered operator in this toy model is not deformed under the fibration parameter .
Theorem 7.1 (Toy model).
Assume for simplicity that is a stationary Gaussian field on with covariance .
-
(i)
Spectral decomposition of :
The eigenvalue at is simple (with eigenspace ) and -
(ii)
Spectral decomposition of :
For , the eigenvalue at is fully absorbed in the absolutely continuous spectrum, -
(iii)
Fibered resonances:
For all , the resolvent defined on as a family of operators can be extended meromorphically to the whole complex plane with a unique simple pole atIn other words, for all , we can write for ,
where the remainder is entire, has a continuous dependence on , and satisfies the uniform bound
and where the so-called resonant and co-resonant states are explicitly defined, cf. Remark 7.2(a) below.
-
(iv)
Continuous resonant spectrum:
For small enough, the resolvent defined on as a family of operators can be extended holomorphically to , and we denote the extension by . For , this extension has the following discontinuity, as ,Next, viewed as a family of operators , the extended resolvent can be further extended to all as an entire function.
-
(v)
Ballistic transport:
For all with , there holds
Remarks 7.2.
-
(a)
Explicit formula for resonant state:
Up to a gauge transformation, the resonant and co-resonant states in item (iii) take the formwhere is formally defined as
(7.2) More precisely, the action of on is defined inductively on monomials of increasing degree: we set , and for all and ,
(7.3) while the formal representation (7.2) is understood in view of Wick’s theorem. In particular, there holds .
-
(b)
Mourre’s approach:
In view of Proposition 5.7(ii)–(iii), with the notation of Section 5.3, the commutator is -bounded and satisfies the lower boundIn other words, satisfies a Mourre relation with conjugate “up to ”. Much spectral information can be inferred from such a property, and in particular another proof of Theorem 7.1 above could essentially be deduced. This Mourre approach would be particularly useful in the discrete setting, that is, for the discrete operator on with an i.i.d. Gaussian field on : the proof below can indeed not be adapted to that case as the flow is not explicit. ∎
Proof of Theorem 7.1.
Item (i) is obtained similarly as in the Schrödinger case, cf. Section 4.1, and the proof is omitted. Item (ii) is a direct consequence of (iii). It remains to establish items (iii), (iv), and (v). Without loss of generality, we assume that is supported in .
Step 1. Proof of (iii).
It suffices to show for all and ,
(7.4) |
where we gain a factor in the right-hand side if or is constant, and a factor if both are constants. Indeed, for , we can write
so that the conclusion (iii) would follow from (7.4) after integration. By linearity, it suffices to establish (7.4) for monomials
Noting that the fibered evolution can be explicitly computed as
we find
By Wick’s formula, for , we compute
Since by assumption
and similarly , we deduce
We recognize here the inductive definition (7.3) of the resonant state , so that the above becomes
and a similar computation yields
Finally, since is Gaussian, we compute
and the conclusion (7.4) follows.
Step 2. Proof of (iv).
For and , the Floquet–Bloch fibration (7.1) yields
and thus, inserting the result of item (iii), for ,
where the first right-hand side term is entire and the second is holomorphic on . Next, for , the Fourier transforms are entire functions, which allows to extend the second right-hand side term holomorphically to the whole complex plane. Indeed, for , the Sokhotski–Plemelj formula yields
so that the function defined for by
and defined for by
is entire. This proves (iv).
Step 3. Proof of (v).
The flow can be explicitly integrated,
and is seen to satisfy ballistic transport,
Appendix A Self-adjointness with unbounded potentials
For a bounded random potential , the Schrödinger operator on is clearly self-adjoint on and the fibered operators are self-adjoint on just as for . The present appendix is concerned with the corresponding self-adjointness statement in the unbounded setting. More precisely, we show that essential self-adjointness still holds provided that has negative part for some . This condition is essentially optimal and applies in particular to the case when is a stationary Gaussian field. (Note however that this Gaussian case is much simpler in view of Malliavin calculus and can be obtained as a consequence of Nelson’s theorem in form of Proposition 5.16 with .)
A random potential defines (densely) a multiplicative operator on . If is not uniformly bounded, this operator is unbounded, so that the self-adjointness of is a subtle question and may fail, cf. [27]. Whenever realizations are quadratically controlled from below, in the sense of for some random variable , the Faris-Lavine argument [19] ensures that the Schrödinger operator on is essentially self-adjoint on . By a Borel-Cantelli argument, the quadratic lower bound holds whenever the negative part belongs to for some and satisfies more precisely . In this setting, since is essentially self-adjoint on , we may repeat the direct integral decomposition (3.4) and the fibered operators on are necessarily essentially self-adjoint on for almost all , e.g. [48, p.280]. In order to conclude for all , some continuity would be needed, which typically requires smoothness of . In order to avoid such spurious assumptions, we provide another argument below. While the usual Faris-Lavine argument is of no use in the stationary space , we draw inspiration from an earlier work by Kato [27].
Theorem A.1 (Essential self-adjointness).
Assume that the potential satisfies for some . Then for all and the operator is essentially self-adjoint on . ∎
Proof.
Let and be fixed. For , we note that the operator is well-defined on the whole of with values in the space (cf. Section 3.2 for notation), and it is obviously continuous . Let denote the restriction of with domain . Since , the operator can be viewed as a densely defined operator on and it is clearly symmetric. Its adjoint is easily seen as the restriction of to , that is, defined whenever and . In this context, the following conditions are equivalent:
-
(E1)
is essentially self-adjoint.
-
(E2)
There exist two complex numbers with such that is an injection of into .
-
(E3)
The restriction of to is the strong closure of , that is, for all with there exists a sequence such that and in .
We proceed by truncation: we define the truncated operator for , and we split the proof into four steps.
Step 1.
Proof that is essentially self-adjoint.
Assume that satisfies in .
Applying the differential inequality of [27, Lemma A] in the form
we deduce
For large enough, we have and
Since the operator is nonnegative, this implies . Hence, the operator is an injection of into , and the claim follows from the equivalence between (E1) and (E2).
Step 2. For all and , there exists a cut-off function with the following properties:
-
(i)
on ;
-
(ii)
outside ;
-
(iii)
, ;
-
(iv)
there exists such that almost surely as whenever .
Define . Choose with , , outside , , and , and choose an even function with , for , , and . We then construct the stationary function
Properties (i)–(iii) easily follow for this choice. We turn to (iv). The definition of , a union bound, and Markov’s inequality yield
Since by assumption for for some , we deduce in measure whenever . In order to establish almost sure convergence, we similarly compute, noting that is decreasing in ,
and almost sure convergence follows under the same condition on .
Step 3. Gårding inequalities:
-
(G1)
For all with , there holds and
-
(G2)
For all with , there holds
where as in Step 2 we have set .
By density, it suffices to argue for . We start with the Gårding inequality (G1) for the truncated operator . For , we compute
hence,
and the claim (G1) follows.
We turn to (G2). Similarly as in Step 2, we may construct a cut-off function with the following properties
-
(i’)
on ;
-
(ii’)
outside ;
-
(iii’)
, .
Noting that , the result (G1) yields
Computing
and noting that , we deduce
Since for integration by parts yields
we find
hence,
Since outside , the claim (G2) follows.
Step 4. Conclusion.
Let with .
In view of the equivalence between properties (E1) and (E3), it suffices to construct a sequence such that and in as .
We argue by truncation.
Let be the cut-off function defined in Step 2 and choose such that property (iv) is satisfied. We show that for all there exists a sequence such that
-
and in as ;
-
and in as .
We split the proof into three further substeps.
Substep 4.1. Proof that for all there exists a sequence such that and in as .
For all , since by Step 1 the operator is essentially self-adjoint, the equivalence between properties (E1) and (E3) implies that there exists a sequence such that and in as . By definition of , there holds , and the claim follows.
Substep 4.2. Proof that for all there holds and in as .
We start from the identity
and note that the convergence of (cf. Substep 4.1) implies and
Combining this with the Gårding inequality (G1) of Step 3 and with the convergence properties of , the claim follows.
Substep 4.3. Proof that and in as .
We start from the identity
and note that the properties of and the dominated convergence theorem lead to , , and in , hence
Combining this with the Gårding inequality (G2) of Step 3 yields
and the claim follows. ∎
Acknowledgements
The authors wish to thank Antoine Gloria, Sylvain Golenia, Felipe Hernandez, Laure Saint-Raymond, Johannes Sjöstrand, and Martin Vogel for motivating discussions at different stages of this work. Financial support is acknowledged from the CNRS-Momentum program.
Data availability statement
Data sharing is not applicable to this article as no new data were created or analyzed in this study.
References
- [1] W. O. Amrein, A. Boutet de Monvel, and V. Georgescu. -groups, commutator methods and spectral theory of -body Hamiltonians, volume 135 of Progress in Mathematics. Birkhäuser Verlag, Basel, 1996.
- [2] P. W. Anderson. Absence of Diffusion in Certain Random Lattices. Phys. Rev., 109(5):1492–1505, 1958.
- [3] J. Asch and A. Knauf. Motion in periodic potentials. Nonlinearity, 11(1):175–200, 1998.
- [4] A. Benoit and A. Gloria. Long-time homogenization and asymptotic ballistic transport of classical waves. Ann. Sci. Éc. Norm. Supér., 52(3):703–759, 2019.
- [5] T. Bodineau, I. Gallagher, L. Saint-Raymond, and S. Simonella. Fluctuation Theory in the Boltzmann–Grad Limit. To appear in J. Stat. Phys., 2020.
- [6] J. Bourgain. Anderson localization for quasi-periodic lattice Schrödinger operators on , arbitrary. Geom. Funct. Anal., 17(3):682–706, 2007.
- [7] É. Cancès, S. Lahbabi, and M. Lewin. Mean-field models for disordered crystals. J. Math. Pures Appl. (9), 100(2):241–274, 2013.
- [8] L. Cattaneo, G.M. Graf, and W. Hunziker. A general resonance theory based on Mourre’s inequality. Ann. Henri Poincaré, 7(3):583–601, 2006.
- [9] T. Chen, T. Komorowski, and L. Ryzhik. The weak coupling limit for the random Schrödinger equation: The average wave function. Arch. Ration. Mech. Anal., 227(1):387–422, 2018.
- [10] O. Costin and A. Soffer. Resonance theory for Schrödinger operators. Commun. Math. Phys., 224:133–152, 2001.
- [11] P. Doukhan. Mixing, volume 85 of Lecture Notes in Statistics. Springer-Verlag, New York, 1994.
- [12] M. Duerinckx. On the size of chaos via Glauber calculus in the classical mean-field dynamics. Comm. Math. Phys., 382:613–653, 2021.
- [13] M. Duerinckx, A. Gloria, and C. Shirley. Approximate normal forms via Floquet-Bloch theory: Nehorošev stability for linear waves in quasiperiodic media. Comm. Math. Phys., 383:633–683, 2021.
- [14] S. Dyatlov and M. Zworski. Mathematical theory of scattering resonances, volume 200 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2019.
- [15] L. Erdős. Lecture notes on quantum Brownian motion. In J. Fröhlich, M. Salmhofer, V. Mastropietro, W. De Roeck, and L. F. Cugliandolo, editors, Quantum Theory from Small to Large Scales, volume 95 of Lecture Notes of the Les Houches Summer School, August 2010, pages 3–98. Oxford University Press, Oxford, 2012.
- [16] L. Erdős, M. Salmhofer, and H.-T. Yau. Quantum diffusion of the random Schrödinger evolution in the scaling limit. II. The recollision diagrams. Comm. Math. Phys., 271(1):1–53, 2007.
- [17] L. Erdős, M. Salmhofer, and H.-T. Yau. Quantum diffusion of the random Schrödinger evolution in the scaling limit. Acta Math., 200(2):211–277, 2008.
- [18] L. Erdős and H.-T. Yau. Linear Boltzmann equation as the weak coupling limit of a random Schrödinger equation. Comm. Pure Appl. Math., 53(6):667–735, 2000.
- [19] W. G. Faris and R. B. Lavine. Commutators and self-adjointness of Hamiltonian operators. Comm. Math. Phys., 35:39–48, 1974.
- [20] V. Georgescu and C. Gérard. On the virial theorem in quantum mechanics. Comm. Math. Phys., 208(2):275–281, 1999.
- [21] W. Hunziker. Resonances, metastable states and exponential decay laws in perturbation theory. Comm. Math. Phys., 132(1):177–188, 1990.
- [22] A. Jensen, E. Mourre, and P. Perry. Multiple commutator estimates and resolvent smoothness in quantum scattering theory. Ann. Inst. H. Poincaré Phys. Théor., 41(2):207–225, 1984.
- [23] V. V. Jikov, S. M. Kozlov, and O. A. Oleĭnik. Homogenization of differential operators and integral functionals. Springer-Verlag, Berlin, 1994. Traduit du russe par G. A. Iosif′yan.
- [24] Y. Kang and J. Schenker. Diffusion of wave packets in a Markov random potential. J. Stat. Phys., 134(5-6):1005–1022, 2009.
- [25] Y. Karpeshina, Y.-R. Lee, R. Shterenberg, and G. Stolz. Ballistic transport for the Schrödinger operator with limit-periodic or quasi-periodic potential in dimension two. Comm. Math. Phys., 354(1):85–113, 2017.
- [26] Y. Karpeshina, L. Parnovski, and R. Shterenberg. Bethe-Sommerfeld conjecture and absolutely continuous spectrum of multi-dimensional quasi-periodic Schrödinger operators. Preprint, arXiv:2010.05881, 2020.
- [27] T. Kato. Schrödinger operators with singular potentials. Israel J. Math., 13(1–2):135–148, 1972.
- [28] T. Kato. Perturbation theory for linear operators. Classics in Mathematics. Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition.
- [29] H. Kesten and G. C. Papanicolaou. A limit theorem for stochastic acceleration. Comm. Math. Phys., 78(1):19–63, 1980/81.
- [30] T. Komorowski and L. Ryzhik. The stochastic acceleration problem in two dimensions. Israel J. Math., 155:157–203, 2006.
- [31] L. B. Koralov and Y. G. Sinai. Theory of probability and random processes. Universitext. Springer, Berlin, second edition, 2007.
- [32] P. Kuchment. Floquet theory for partial differential equations, volume 60 of Operator Theory: Advances and Applications. Birkhäuser Verlag, Basel, 1993.
- [33] P. Kuchment. An overview of periodic elliptic operators. Bull. Amer. Math. Soc. (N.S.), 53(3):343–414, 2016.
- [34] S. Lahbabi. Étude mathématique de modèles quantiques et classiques pour les matériaux aléatoires à l’échelle atomique. PhD thesis, Université de Cergy Pontoise, 2013.
- [35] P. Malliavin. Stochastic analysis, volume 313 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, 1997.
- [36] M. Merkli and I. M. Sigal. A time-dependent theory of quantum resonances. Commun. Math. Phys., 201:549–576, 1999.
- [37] E. Mourre. Absence of singular continuous spectrum for certain selfadjoint operators. Comm. Math. Phys., 78(3):391–408, 1980/81.
- [38] E. Nelson. Time-Ordered Operators Products of Sharp-Time Quadratic Forms. J. Functional Analysis, 11:211–219, 1972.
- [39] I. Nourdin and G. Peccati. Normal approximations with Malliavin calculus, volume 192 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2012.
- [40] D. Nualart. The Malliavin calculus and related topics. Springer-Verlag, Berlin, second edition, 2006.
- [41] A. Orth. Quantum mechanical resonance and limiting absorption: the many body problem. Comm. Math. Phys., 126(3):559–573, 1990.
- [42] G. C. Papanicolaou and S. R. S. Varadhan. Boundary value problems with rapidly oscillating random coefficients. In Random fields, Vol. I, II (Esztergom, 1979), volume 27 of Colloq. Math. Soc. János Bolyai, pages 835–873. North-Holland, Amsterdam, 1981.
- [43] L. Pastur and A. Figotin. Spectra of random and almost-periodic operators, volume 297 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1992.
- [44] G. Peccati and M. Reitzner, editors. Stochastic analysis for Poisson point processes, volume 7 of Bocconi & Springer Series. Bocconi University Press, Springer, 2016.
- [45] C.-A. Pillet. Some results on the quantum dynamics of a particle in a Markovian potential. Comm. Math. Phys., 102(2):237–254, 1985.
- [46] C. R. Putnam. Commutations Properties of Hilbert Space Operators and Related Topics, volume 36 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 2. Folge. Springer-Verlag, Berlin Heidelberg, 1967.
- [47] M. Reed and B. Simon. Methods of modern mathematical physics. I. Functional analysis. Academic Press, New York-London, 1972.
- [48] M. Reed and B. Simon. Methods of modern mathematical physics. IV. Analysis of operators. Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1978.
- [49] B. Simon. Schrödinger operators in the twenty-first century. In Mathematical physics 2000, pages 283–288. Imp. Coll. Press, London, 2000.
- [50] A. Sofer and M. I. Weinstein. Time-dependent resonance theory. Geom. Funct. Anal., 8:1086–1128, 1998.
- [51] H. Spohn. Derivation of the transport equation for electrons moving through random impurities. J. Statist. Phys., 17(6):385–412, 1977.
- [52] J. von Neumann. Über Einen Satz Von Herrn M. H. Stone. Ann. Math., 33(3):567–573, 1932.
- [53] W. Weisskopf and E. Wigner. Berechnung der natürlichen Linienbreite auf Grund der Diracschen Lichttheorie. Z. Phys., 63(1–2):54–73, 1930.
- [54] Zh. Zhao. Ballistic transport in one-dimensional quasi-periodic continuous Schrödinger equation. J. Differential Equations, 262(9):4523–4566, 2017.