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

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

Hλ,ω:=+λVωH_{\lambda,\omega}:=-\triangle+\lambda V_{\omega}

on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), where Vω:dV_{\omega}:\mathbb{R}^{d}\to\mathbb{R} is a realization of a “stationary” (that is, statistically translation-invariant) random potential VV, constructed on a given probability space (Ω,)(\Omega,\mathbb{P}), and we study the properties of the corresponding Schrödinger flow on d\mathbb{R}^{d},

ituλ,ω=Hλ,ωuλ,ω,uλ,ω|t=0=u,i\partial_{t}u_{\lambda,\omega}=H_{\lambda,\omega}u_{\lambda,\omega},\qquad u_{\lambda,\omega}|_{t=0}=u^{\circ},

with initial data uL2(d)u^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}). This well-travelled equation models the motion of an electron in a disordered medium described by the potential VωV_{\omega}, where the coupling constant λ>0\lambda>0 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 VV is periodic, the energy transport is well-known to remain purely ballistic as for the free flow uλ=0u_{\lambda=0}, cf. [3]. The proof relies on the absolute continuity of the spectrum of the periodic Schrödinger operator on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), 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, xeikxφk,λ(x)x\mapsto e^{ik\cdot x}\varphi_{k,\lambda}(x) with φk,λ\varphi_{k,\lambda} periodic. In case of a quasiperiodic potential VV, the problem is more involved and depends on the strength of the disorder: energy transport is expected to remain ballistic only at weak coupling 0<λ10<\lambda\ll 1 or at high energies. This is rigorously established in dimension d=1d=1 [54]. In higher dimensions d>1d>1, 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 xeikxφk,λ(x)x\mapsto e^{ik\cdot x}\varphi_{k,\lambda}(x) with φk,λ\varphi_{k,\lambda} 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 0<λ10<\lambda\ll 1. In case of a random potential VV with short-range correlations, in stark contrast with the periodic and quasiperiodic cases, a celebrated conjecture by Anderson [2] states that in dimension d>2d>2 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 uλu_{\lambda} is known to converge to the solution of a linear Boltzmann equation on the kinetic timescale tλ2t\sim\lambda^{-2}, and of a heat equation on longer times, but the justification is limited to a perturbative time regime tλ2ηt\ll\lambda^{-2-\eta} for some small η>0\eta>0. A simplified question concerns the behavior of time correlations in form of the averaged wavefunction 𝔼[uλt]\mathbb{E}\left[{u_{\lambda}^{t}}\right], which is expected to display exponential time decay: more precisely, in Fourier space, on the kinetic timescale tλ2t\sim\lambda^{-2}, as λ0\lambda\downarrow 0,

|𝔼[u^λt=λ2s(k)]|esαk|u^(k)|,\big{|}\mathbb{E}\big{[}{\widehat{u}_{\lambda}^{t=\lambda^{-2}s}(k)}\big{]}\big{|}\,\sim\,e^{-s\alpha_{k}}\,|\widehat{u}^{\circ}(k)|, (1.1)

where the decay rate αk>0\alpha_{k}>0 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 Hλ,ω=+λVωH_{\lambda,\omega}=-\triangle+\lambda V_{\omega}. First, we change the point of view and rather consider the operator Hλ:=+λVH_{\lambda}:=-\triangle+\lambda V as acting on the augmented Hilbert space L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega), then studying the corresponding Schrödinger flow on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega),

ituλ=Hλuλ,uλ|t=0=u,i\partial_{t}u_{\lambda}=H_{\lambda}u_{\lambda},\qquad u_{\lambda}|_{t=0}=u^{\circ}, (1.2)

with deterministic initial data uL2(d)L2(d×Ω)u^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d})\subset\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega). This can be viewed as including stochastic averaging conveniently into the functional setup; see also [45, 24]. (Note that HλH_{\lambda} on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega) has absolutely continuous spectrum as a consequence of Wegner estimates when Hλ,ωH_{\lambda,\omega} on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}) 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 HλH_{\lambda} on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega) can be decomposed via a Fourier-type transformation as a direct integral of fibered operators {Hk,λst}k\{H_{k,\lambda}^{\operatorname{st}}\}_{k} acting on the elementary space L2(Ω)\operatorname{L}^{2}(\Omega), which is viewed as the space of stationary random fields,

(Hλ,L2(d×Ω))=d(Hk,λst+|k|2,L2(Ω))𝔢k¯𝑑k,𝔢k(x):=eikx.\big{(}H_{\lambda},\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega)\big{)}=\int_{\mathbb{R}^{d}}^{\oplus}\big{(}H_{k,\lambda}^{\operatorname{st}}+|k|^{2},\operatorname{L}^{2}(\Omega)\big{)}\,\mathfrak{e}_{k}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,\qquad\mathfrak{e}_{k}(x):=e^{ik\cdot x}. (1.3)

The (centered) fibered operators take the form

Hk,λst\displaystyle H_{k,\lambda}^{\operatorname{st}} :=\displaystyle:= Hk,0st+λV,\displaystyle H_{k,0}^{\operatorname{st}}+\lambda V,
Hk,0st\displaystyle H_{k,0}^{\operatorname{st}} :=\displaystyle:= (st+ik)(st+ik)|k|2=st2ikst,\displaystyle-(\nabla^{\operatorname{st}}+ik)\cdot(\nabla^{\operatorname{st}}+ik)-|k|^{2}\,=\,-\triangle^{\operatorname{st}}-2ik\cdot\nabla^{\operatorname{st}},

where st\nabla^{\operatorname{st}} and st\triangle^{\operatorname{st}} denote the stationary gradient and Laplacian on L2(Ω)\operatorname{L}^{2}(\Omega); see Section 3.2 for proper definitions. In particular, the Schrödinger flow uλu_{\lambda} is decomposed as

uλt(x,ω)=du^(k)eikxit|k|2(eitHk,λst1)(x,ω)¯𝑑k,u_{\lambda}^{t}(x,\omega)\,=\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it|k|^{2}}\big{(}e^{-itH^{\operatorname{st}}_{k,\lambda}}1\big{)}(x,\omega)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k, (1.4)

in terms of the fibered evolutions {eitHk,λst1}k\{e^{-itH^{\operatorname{st}}_{k,\lambda}}1\}_{k} on L2(Ω)\operatorname{L}^{2}(\Omega). 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 λ=0\lambda=0, as the constant function 1L2(Ω)1\in\operatorname{L}^{2}(\Omega) is an eigenfunction with eigenvalue 0 for the unperturbed fibered operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k}, the associated spectral measure coincides with the Dirac mass at 0, and the decomposition (1.4) then reduces to the usual Fourier diagonalisation of the free flow,

uλ=0t(x)=du^(k)eikxit|k|2¯𝑑k.\displaystyle u_{\lambda=0}^{t}(x)\,=\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it|k|^{2}}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k. (1.5)

When the disorder is turned on but small, 0<λ10<\lambda\ll 1, the description of the Schrödinger flow uλu_{\lambda} is reduced to a family of (hopefully simpler) fibered perturbation problems for the spectral measures. In case of a periodic potential VV (that is, Ω=𝕋d\Omega=\mathbb{T}^{d}, cf. Remark 3.4(a)), the fibered operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} have compact resolvent in view of the Rellich theorem, and thus discrete spectrum. The eigenvalue at 0 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 {zk,λ=λ2zλ(k)}k\{z_{k,\lambda}=\lambda^{2}z_{\lambda}(k)\}_{k}. In other words, Fourier modes {𝔢k}k\{\mathfrak{e}_{k}\}_{k} are perturbed into so-called periodic Bloch waves that diagonalize the Schrödinger flow and are associated with perturbed generalized eigenvalues of the form {|k|2+λ2zλ(k)}k\{|k|^{2}+\lambda^{2}z_{\lambda}(k)\}_{k}. This entails that the flow is approximately conjugated to the free flow (1.5) in the sense of

uλt(x)=du^(k)eikxit(|k|2+λ2zλ(k))¯𝑑k+O(λ),zλC(d;),\displaystyle u_{\lambda}^{t}(x)\,=\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it(|k|^{2}+\lambda^{2}z_{\lambda}(k))}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k\,+O(\lambda),\qquad z_{\lambda}\in C^{\infty}(\mathbb{R}^{d};\mathbb{R}),

and in particular the energy transport remains ballistic forever. The same conclusion holds in fact for all λ\lambda, cf. [3]. In case of a quasiperiodic potential VV (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 {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} are degenerate elliptic operators, for which compactness fails, and the simple eigenvalue at 0 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 VV with short-range correlations, the situation differs drastically in link with the expected diffusive behavior. We show that 0 is typically the only eigenvalue of the fibered operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k}, 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

zk,λ=λ2zλ(k)=λ2𝔼[V(i0Hk,0st)1V]+Ok(λ3),z_{k,\lambda}=\lambda^{2}z_{\lambda}(k)=\lambda^{2}\mathbb{E}\left[{V(i0-H_{k,0}^{\operatorname{st}})^{-1}V}\right]+O_{k}(\lambda^{3}),

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 tλ2t\sim\lambda^{-2}. For the averaged wavefunction, this leads to expect

𝔼[uλt(x)]=du^(k)eikxit(|k|2+λ2zλ(k))¯𝑑k+O(λ),zλC(d;),\displaystyle\mathbb{E}\left[{u_{\lambda}^{t}(x)}\right]\,=\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it(|k|^{2}+\lambda^{2}z_{\lambda}(k))}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k\,+O(\lambda),\qquad z_{\lambda}\in C^{\infty}(\mathbb{R}^{d};\mathbb{C}\setminus\mathbb{R}),

which would indeed agree with the exponential decay (1.1) on the kinetic timescale, with zλ(k)αk<0\Im z_{\lambda}(k)\sim-\alpha_{k}<0, 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 HλH_{\lambda} on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega), cf. Remark 2.1 below.

General spectral tools are however dramatically missing to rigorously study these fibered perturbation problems on L2(Ω)\operatorname{L}^{2}(\Omega), 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 13, emphasizing the strong dependence on the structure of the underlying probability space (Ω,)(\Omega,\mathbb{P}). 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 {Hk,0st}k\{H^{\operatorname{st}}_{k,0}\}_{k}. 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 VV in the sense that λV\lambda V cannot be considered as a small perturbation of {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} 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

  1. \bullet

    We denote by C1C\geq 1 any constant that only depends on the space dimension dd and on the law of the random potential VV. We use the notation \lesssim (resp. \gtrsim) for C×\leq C\times (resp. 1C×\geq\frac{1}{C}\times) up to such a multiplicative constant CC. We write \simeq when both \lesssim and \gtrsim hold. We add subscripts to C,,,C,\lesssim,\gtrsim,\simeq to indicate dependence on other parameters. We denote by O(K)O(K) any quantity that is bounded by CKCK.

  2. \bullet

    We denote by f^(k):=f(k):=deikxf(x)𝑑x\widehat{f}(k):=\mathcal{F}f(k):=\int_{\mathbb{R}^{d}}e^{-ik\cdot x}f(x)\,dx the usual Fourier transform of a smooth function ff on d\mathbb{R}^{d}. The inverse Fourier transform is then given by f(x)=deikxf^(k)¯𝑑kf(x)=\int_{\mathbb{R}^{d}}e^{ik\cdot x}\widehat{f}(k)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k in terms of the rescaled Lebesgue measure ¯dk:=(2π)ddk\,{{\mathchar 22\relax\mkern-12.0mud}}k:=(2\pi)^{-d}dk.

  3. \bullet

    The ball centered at xx and of radius rr in d\mathbb{R}^{d} is denoted by Br(x)B_{r}(x), and we write for abbreviation B(x):=B1(x)B(x):=B_{1}(x), Br:=Br(0)B_{r}:=B_{r}(0), and B:=B1(0)B:=B_{1}(0). Without ambiguity, we occasionally also denote by BB the unit ball at the origin in the complex plane \mathbb{C}.

  4. \bullet

    For a set EdE\subset\mathbb{R}^{d} we denote by conv(E){\operatorname{conv}}(E) its convex envelope, by int(E)\operatorname{int}(E) its interior, and by adh(E){\operatorname{adh\,}}(E) its closure.

  5. \bullet

    We denote by (k)\mathcal{B}(\mathbb{R}^{k}) the set of Borel subsets of k\mathbb{R}^{k}, and for EkE\subset\mathbb{R}^{k} we let 𝒫(E)\mathcal{P}(E) denote the set of Borel probability measures on EE.

  6. \bullet

    For a vector space XX, we write XpX^{\otimes p} for its pp-fold tensor product, and XpX^{\odot p} for its pp-fold symmetric tensor product.

  7. \bullet

    For a,ba,b\in\mathbb{R}, we write ab:=min{a,b}a\wedge b:=\min\{a,b\} and ab:=max{a,b}a\vee b:=\max\{a,b\}.

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 VV is assumed real-valued and centered, 𝔼[V]=0\mathbb{E}\left[{V}\right]=0. 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 L2(Ω)\operatorname{L}^{2}(\Omega). More precisely, we consider the following:

  1. \bullet

    Gaussian setting: V=b(V0)V=b(V_{0}) for some Borel function b:b:\mathbb{R}\to\mathbb{R} and some stationary centered Gaussian random field V0V_{0} with bounded covariance function

    𝒞0(x):=𝔼[V0(x+y)V0(y)].\qquad\mathcal{C}_{0}(x):=\mathbb{E}\left[{V_{0}(x+y)V_{0}(y)}\right].

    Equivalently, the field V0V_{0} can be represented as

    V0(x)=d𝒞0(x+y)𝑑Z(y),\qquad V_{0}(x)=\int_{\mathbb{R}^{d}}\mathcal{C}_{0}^{\circ}(x+y)\,dZ(y), (2.1)

    where dZdZ is a standard Gaussian white noise on d\mathbb{R}^{d} and where the kernel 𝒞0\mathcal{C}_{0}^{\circ} is the convolution square root of the covariance function, 𝒞0=𝒞0𝒞0\mathcal{C}_{0}=\mathcal{C}_{0}^{\circ}\ast\mathcal{C}_{0}^{\circ}.

  2. \bullet

    Poisson setting: V=b(V0)V=b(V_{0}) for some Borel function b:b:\mathbb{R}\to\mathbb{R} and some V0V_{0} of the form

    V0(x)=y𝒫0𝒞0(x+y),\qquad V_{0}(x)=\sum_{y\in\mathcal{P}_{0}}\mathcal{C}_{0}^{\circ}(x+y), (2.2)

    where 𝒫0\mathcal{P}_{0} is a standard Poisson point process on d\mathbb{R}^{d} and where 𝒞0\mathcal{C}_{0}^{\circ} is the single-site potential.

We say that the random potential VV is short-range if it has integrable decay of correlations: in the above settings, this amounts to choosing 𝒞0L1(d)\mathcal{C}_{0}^{\circ}\in\operatorname{L}^{1}(\mathbb{R}^{d}).

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 V=V0V=V_{0} that is itself Gaussian: although unbounded, such potentials have a simpler action on the Fock space decomposition of L2(Ω)\operatorname{L}^{2}(\Omega).

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

Hk,0st:=st2ikst,kd.H^{\operatorname{st}}_{k,0}:=-\triangle^{\operatorname{st}}-2ik\cdot\nabla^{\operatorname{st}},\qquad k\in\mathbb{R}^{d}.

Although the stationary Laplacian st-\triangle^{\operatorname{st}} is a natural operator on L2(Ω)\operatorname{L}^{2}(\Omega) 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 Ω\Omega is a finite set, that there is in general no spectral gap above 0 in contrast with the periodic setting, and that it coincides with [0,)[0,\infty) in case of an i.i.d. structure. Interestingly, the spectrum depends crucially on the structure of the underlying probability space (Ω,)(\Omega,\mathbb{P}), 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 Hk,0stH_{k,0}^{\operatorname{st}}).

Given a stationary Gaussian field V0V_{0} on d\mathbb{R}^{d} with covariance function 𝒞0\mathcal{C}_{0}, denote by 𝒞^0\widehat{\mathcal{C}}_{0} the (nonnegative measure) Fourier transform of 𝒞0\mathcal{C}_{0}, and assume that the probability space (Ω,)(\Omega,\mathbb{P}) is endowed with the σ\sigma-algebra σ({V0(x)}xd)\sigma(\{V_{0}(x)\}_{x\in\mathbb{R}^{d}}) generated by V0V_{0}.

  1. (i)

    If 𝒞0\mathcal{C}_{0} is not periodic in any direction, then σ(Hk,0st)=[|k|2,)\sigma(H_{k,0}^{\operatorname{st}})=[-|k|^{2},\infty).

  2. (ii)

    If the measure 𝒞^0\widehat{\mathcal{C}}_{0} is absolutely continuous (in particular, if 𝒞0\mathcal{C}_{0} is integrable), then the eigenvalue at 0 is simple (with eigenspace \mathbb{C}) and

    σpp(Hk,0st)={0},σsc(Hk,0st)=,σ(Hk,0st)=σac(Hk,0st)=[|k|2,).\sigma_{{\operatorname{pp}}}(H_{k,0}^{\operatorname{st}})\,=\,\{0\},\quad\sigma_{\operatorname{sc}}(H_{k,0}^{\operatorname{st}})\,=\,\varnothing,\quad\sigma(H_{k,0}^{\operatorname{st}})\,=\,\sigma_{{\operatorname{ac}}}(H_{k,0}^{\operatorname{st}})\,=\,[-|k|^{2},\infty).\qed

We turn to the perturbed fibered operators {Hk,λst}k\{H^{\operatorname{st}}_{k,\lambda}\}_{k} 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 VV, the essential self-adjointness of the perturbed operators {Hk,λst}k\{H_{k,\lambda}^{\operatorname{st}}\}_{k} is already a delicate issue, for which an (almost optimal) criterion is included in Appendix A, requiring VLp(Ω)V\in\operatorname{L}^{p}(\Omega) for some p>d2p>\frac{d}{2}, in line with the corresponding celebrated self-adjointness problem for Schrödinger operators on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}) with singular potentials, cf. [27, 19, 28]; see also Proposition 5.16 for the simpler case when V=V0V=V_{0} is itself Gaussian.

Proposition 2 (Spectrum of Hk,λstH_{k,\lambda}^{\operatorname{st}}).

Consider the Gaussian setting V=b(V0)V=b(V_{0}), where V0V_{0} is a stationary Gaussian field, and assume that V0V_{0} is nondegenerate and that 𝔼[|V|p]<\mathbb{E}\left[{|V|^{p}}\right]<\infty holds for some p>d2p>\frac{d}{2}. Then Hk,λstH_{k,\lambda}^{\operatorname{st}} is essentially self-adjoint on H2L(Ω)H^{2}\cap\operatorname{L}^{\infty}(\Omega) and

σ(Hk,λst)=[|k|2+λinfessb,).\sigma(H_{k,\lambda}^{\operatorname{st}})=[-|k|^{2}+\lambda\operatorname{inf\,ess}b,\,\infty).\qed

The nature of the spectrum of the perturbed operators {Hk,λst}k\{H_{k,\lambda}^{\operatorname{st}}\}_{k} 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 0 for the fibered operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} 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 λ=0\lambda=0 would be a complex number, cf. Section 4.3.

Proposition 3 (Instability of the bound state).

Let kd{0}k\in\mathbb{R}^{d}\setminus\{0\}, let VV be a stationary random field, denote by 𝒞^\widehat{\mathcal{C}} the (nonnegative measure) Fourier transform of its covariance function, and assume that 𝒞^\widehat{\mathcal{C}} does not vanish identically on the sphere B|k|(k)\partial B_{|k|}(-k), in the sense that limε01ε𝒞^(B|k|+ε(k)B|k|ε(k))> 0\lim_{\varepsilon\downarrow 0}\frac{1}{\varepsilon}\widehat{\mathcal{C}}\big{(}B_{|k|+\varepsilon}(-k)\setminus B_{|k|-\varepsilon}(-k)\big{)}\,>\,0. Then there exists no C2C^{2} branch

[0,δ)×L2(Ω):λ(Ek,λ,ψk,λ)[0,\delta)\to\mathbb{R}\times\operatorname{L}^{2}(\Omega):\lambda\mapsto(E_{k,\lambda},\psi_{k,\lambda})

with

Hk,λstψk,λ=Ek,λψk,λ,(Ek,λ,ψk,λ)|λ=0=(0,1).H_{k,\lambda}^{\operatorname{st}}\psi_{k,\lambda}=E_{k,\lambda}\psi_{k,\lambda},\qquad(E_{k,\lambda},\psi_{k,\lambda})|_{\lambda=0}=(0,1).\qed

This basic instability result is however quite weak: for kd{0}k\in\mathbb{R}^{d}\setminus\{0\}, the operator Hk,λstH_{k,\lambda}^{\operatorname{st}} is in fact expected to have purely absolutely continuous spectrum in a neighborhood of 0 for 0<λ10<\lambda\ll 1. 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 VV is never compact with respect to the unperturbed operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} on L2(Ω)\operatorname{L}^{2}(\Omega) (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 {Utst}t\{U_{t}^{\operatorname{st}}\}_{t\in\mathbb{R}} on L2(Ω)\operatorname{L}^{2}(\Omega) 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 AstA^{\operatorname{st}} of this group is then checked to be a conjugate operator for the stationary Laplacian st-\triangle^{\operatorname{st}} 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 {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k}. This appears to be surprisingly more involved than for k=0k=0, in link with the infinite dimensionality of the probability space: our proof makes a deep use of the Fock space structure of L2(Ω)\operatorname{L}^{2}(\Omega) 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 {Hk,λst}k\{H_{k,\lambda}^{\operatorname{st}}\}_{k}. It appears that the perturbation λV\lambda V 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 V=V0V=V_{0} be a stationary Gaussian field with covariance function 𝒞0Cc(d)\mathcal{C}_{0}\in C^{\infty}_{c}(\mathbb{R}^{d}), let \mathcal{L} denote the Ornstein–Uhlenbeck operator for the associated Malliavin calculus, cf. Section 5.2, and for a given constant L0>0L_{0}>0 consider the truncation Qλ:=𝟙[0,(L0λ)2]()Q_{\lambda}:=\mathds{1}_{[0,(L_{0}\lambda)^{-2}]}(\mathcal{L}) onto Wiener chaoses of order (L0λ)2\leq(L_{0}\lambda)^{-2}. Then, for any kdk\in\mathbb{R}^{d}, there exists a self-adjoint operator CkstC_{k}^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega) and an explicit core 𝒫(Ω)\mathcal{P}(\Omega), cf. (2.4), such that the following properties hold:

  1. (i)

    For all ε>0\varepsilon>0, the truncated operator QλHk,0stQλQ_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda} satisfies a Mourre relation on the interval Jε:=[ε34|k|2,)J_{\varepsilon}:=[\varepsilon-\frac{3}{4}|k|^{2},\infty) with respect to CkstC_{k}^{\operatorname{st}}. More precisely, its domain is invariant under the unitary group generated by CkstC_{k}^{\operatorname{st}}, and its commutator with 1iCkst\frac{1}{i}C_{k}^{\operatorname{st}} is well-defined and essentially self-adjoint on 𝒫(Ω)\mathcal{P}(\Omega), is QλHk,0stQλQ_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda}-bounded, and satisfies the following lower bound,

    𝟙Jε(QλHk,0stQλ)[QλHk,0stQλ,1iCkst] 1Jε(QλHk,0stQλ)ε𝟙Jε(QλHk,0stQλ)34|k|2𝔼.\qquad\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda})\,[Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\,\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda})\,\geq\,\varepsilon\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda})-\tfrac{3}{4}|k|^{2}\mathbb{E}.
  2. (ii)

    The truncated perturbation QλλVQλQ_{\lambda}\lambda VQ_{\lambda} is compatible with respect to CkstC_{k}^{\operatorname{st}} in the sense that its iterated commutators with 1iCkst\frac{1}{i}C_{k}^{\operatorname{st}} are well-defined on 𝒫(Ω)\mathcal{P}(\Omega) and bounded by O(L01)O(L_{0}^{-1}).

In particular, the truncated perturbed operator QλHk,λstQλQ_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda} satisfies a corresponding Mourre relation on [ε+CL0134|k|2,)[\varepsilon+CL_{0}^{-1}-\frac{3}{4}|k|^{2},\infty) with respect to CkstC_{k}^{\operatorname{st}}. ∎

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 uL2(d)u^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}) have compactly supported Fourier transform, let V=V0V=V_{0} be a stationary Gaussian field with covariance function 𝒞0Cc(d)\mathcal{C}_{0}\in C^{\infty}_{c}(\mathbb{R}^{d}), and define αk,βk\alpha_{k},\beta_{k}\in\mathbb{R} by

αk+iβk:=limε01i𝔼[V(Hk,0stiε)1V],\alpha_{k}+i\beta_{k}:=\lim_{\varepsilon\downarrow 0}\tfrac{1}{i}\mathbb{E}\big{[}{V(H_{k,0}^{\operatorname{st}}-i\varepsilon)^{-1}V}\big{]},

that is, more explicitly,

αk\displaystyle\alpha_{k} :=\displaystyle:= πd𝒞^0(yk)δ(|y|2|k|2)¯dy=π2(2π)d|k|B|k|(k)𝒞^0>0,\displaystyle\pi\int_{\mathbb{R}^{d}}\widehat{\mathcal{C}}_{0}(y-k)\,\delta(|y|^{2}-|k|^{2})\,\,{{\mathchar 22\relax\mkern-12.0mud}}y\,=\,\frac{\pi}{2(2\pi)^{d}|k|}\int_{\partial B_{|k|}(-k)}\widehat{\mathcal{C}}_{0}\leavevmode\nobreak\ \leavevmode\nobreak\ >0, (2.3)
βk\displaystyle\beta_{k} :=\displaystyle:= (2π)dp.v.|k|21r(12|k|2+rB|k|2+r(k)𝒞^0)𝑑r,\displaystyle-(2\pi)^{-d}\,\operatorname{p.v.}\int_{-|k|^{2}}^{\infty}\frac{1}{r}\bigg{(}\frac{1}{2\sqrt{|k|^{2}+r}}\int_{\partial B_{\sqrt{|k|^{2}+r}}(-k)}\widehat{\mathcal{C}}_{0}\bigg{)}\,dr,

Then there exists s01s_{0}\simeq 1 such that the Schrödinger flow satisfies for all 0ss00\leq s\leq s_{0},

limλ0supxd|𝔼[uλλ2s(x)]du^(k)eikxiλ2s|k|2es(αk+iβk)¯𝑑k|= 0.\lim_{\lambda\downarrow 0}\,\sup_{x\in\mathbb{R}^{d}}\bigg{|}\mathbb{E}\big{[}{u_{\lambda}^{\lambda^{-2}s}(x)}\big{]}\,-\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-i\lambda^{-2}s|k|^{2}}e^{-s(\alpha_{k}+i\beta_{k})}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k\bigg{|}\,=\,0.\qed

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 tt on Wiener chaoses of order K\leq K should be accurate provided Kλ2tK\gg\lambda^{2}t. Since our truncation QλQ_{\lambda} in Theorem 4 amounts to projecting onto chaoses of order (L0λ)2\leq(L_{0}\lambda)^{-2}, 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 tλ4t\ll\lambda^{-4}. 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 0 of the fibered operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} 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

zk,λ=λ2(βkiαk)+ok(λ2),kd{0}.z_{k,\lambda}=\lambda^{2}(\beta_{k}-i\alpha_{k})+o_{k}(\lambda^{2}),\qquad k\in\mathbb{R}^{d}\setminus\{0\}.

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 Hk,λH_{k,\lambda} has a resonance at zk,λz_{k,\lambda} in the lower complex half-plane if the resolvent (Hk,λstz)1(H_{k,\lambda}^{\operatorname{st}}-z)^{-1} on the upper half-plane z>0\Im z>0, when viewed in a suitably weakened topology, extends to a meromorphic family of operators indexed by all zz\in\mathbb{C} (or at least in a complex neighborhood), and if this family admits a simple pole at z=zk,λz=z_{k,\lambda}. In the usual case of operators on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), the suitable weakening of the topology typically consists of viewing the resolvent as a family of linear operators Cc(d)𝒟(d)C^{\infty}_{c}(\mathbb{R}^{d})\to\mathcal{D}^{\prime}(\mathbb{R}^{d}) rather than L2(d)L2(d)\operatorname{L}^{2}(\mathbb{R}^{d})\to\operatorname{L}^{2}(\mathbb{R}^{d}). In the present setting on the probability space, the role of Cc(d)L2(d)C^{\infty}_{c}(\mathbb{R}^{d})\subset\operatorname{L}^{2}(\mathbb{R}^{d}) can be played for instance by the dense linear subspace 𝒫(Ω)L2(Ω)\mathcal{P}(\Omega)\subset\operatorname{L}^{2}(\Omega) of VV-polynomials,

𝒫(Ω):={j=1najl=1mjV(xlj):n1,aj,mj0,xljd},\mathcal{P}(\Omega)\,:=\,\bigg{\{}\sum_{j=1}^{n}a_{j}\prod_{l=1}^{m_{j}}V(x_{lj})\,:\,n\geq 1,\,a_{j}\in\mathbb{C},\,m_{j}\geq 0,\,x_{lj}\in\mathbb{R}^{d}\bigg{\}}, (2.4)

and the dual 𝒟(d)\mathcal{D}^{\prime}(\mathbb{R}^{d}) is then replaced by the dual space 𝒫(Ω)\mathcal{P}^{\prime}(\Omega) of continuous linear functionals on 𝒫(Ω)\mathcal{P}(\Omega). In these terms, we formulate the following resonance conjecture. The linear functionals Ψk,λ+\Psi_{k,\lambda}^{+} and Ψk,λ\Psi_{k,\lambda}^{-} below are referred to as the resonant and co-resonant states, respectively. Since the imaginary part of the expected branch of resonances zk,λ=λ2αk+Ok(λ3)\Im z_{k,\lambda}=-\lambda^{2}\alpha_{k}+O_{k}(\lambda^{3}) vanishes to leading order both as k0k\to 0 (in dimension d>2d>2) and as |k||k|\uparrow\infty, cf. formula (2.3), we henceforth restrict to kk in a compact set away from 0.

Conjecture (LRC) — Local resonance conjecture.


Given a compact set Kd{0}K\subset\mathbb{R}^{d}\setminus\{0\}, there are λ0,M>0\lambda_{0},M>0 such that for all kKk\in K and 0λ<λ00\leq\lambda<\lambda_{0} the resolvent z(Hk,λstz)1z\mapsto(H_{k,\lambda}^{\operatorname{st}}-z)^{-1} defined on z>0\Im z>0 as a family of operators 𝒫(Ω)𝒫(Ω)\mathcal{P}(\Omega)\to\mathcal{P}^{\prime}(\Omega) can be extended meromorphically to the whole complex ball |z|1M|z|\leq\frac{1}{M} with a unique simple pole. In other words, there exist continuous collections {zk,λ}k,λ\{z_{k,\lambda}\}_{k,\lambda}\subset\mathbb{C} and {Ψk,λ+}k,λ,{Ψk,λ}k,λ𝒫(Ω)\{\Psi_{k,\lambda}^{+}\}_{k,\lambda},\{\Psi_{k,\lambda}^{-}\}_{k,\lambda}\subset\mathcal{P}^{\prime}(\Omega) such that for all ϕ,ϕ𝒫(Ω)\phi,\phi^{\prime}\in\mathcal{P}(\Omega) we can write for z>0\Im z>0,

ϕ,(Hk,λstz)1ϕL2(Ω)=Ψk,λ+,ϕ𝒫(Ω),𝒫(Ω)¯Ψk,λ,ϕ𝒫(Ω),𝒫(Ω)zk,λz+ζk,λϕ,ϕ(z),\big{\langle}\phi^{\prime},(H_{k,\lambda}^{\operatorname{st}}-z)^{-1}\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\,=\,\frac{\overline{\langle\Psi^{+}_{k,\lambda},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\,\langle\Psi_{k,\lambda}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}{z_{k,\lambda}-z}+\zeta_{k,\lambda}^{\phi^{\prime},\phi}(z), (2.5)

where the remainder ζk,λϕ,ϕ\zeta_{k,\lambda}^{\phi^{\prime},\phi} is holomorphic on the set {z:z>0}1MB\{z:\Im z>0\}\bigcup\frac{1}{M}B and has continuous dependence on k,λk,\lambda. ∎

Next, we state a global version of this resonance conjecture in the case of an unbounded potential VV with σ(Hk,λst)=\sigma(H_{k,\lambda}^{\operatorname{st}})=\mathbb{R} (see e.g. Proposition 2). A direct computation shows that the spectral measure of Hk,0stH_{k,0}^{\operatorname{st}} associated with VV is typically supported on the whole half-axis [|k|2,)[-|k|^{2},\infty) and is only d22\frac{d-2}{2}-times differentiable at |k|2-|k|^{2} in dimension d>2d>2, cf. proof of Lemma 4.2; this suggests that the band on which the meromorphic extension of the resolvent exists must shrink as λ0\lambda\downarrow 0 close to z=|k|2z=-|k|^{2}.

Conjecture (GRC) — Global resonance conjecture.


The same decomposition (2.5) holds with a remainder ζk,λϕ,ϕ\zeta_{k,\lambda}^{\phi^{\prime},\phi} that is holomorphic on the set {z:z>1Mλρ}1MB\{z:\Im z>-\frac{1}{M}\lambda^{\rho}\}\bigcup\frac{1}{M}B for some exponent ρ<2\rho<2, has continuous dependence on k,λk,\lambda, and satisfies a uniform bound of the form

sup|z|12Mλρ|ζk,λϕ,ϕ(z)|λM.\sup_{|\Im z|\leq\frac{1}{2M}\lambda^{\rho}}|\zeta_{k,\lambda}^{\phi^{\prime},\phi}(z)|\,\leq\,\lambda^{-M}.\qed
Remark 2.1 (Continuous resonant spectrum).

When integrated along the Floquet–Bloch fibration (1.3), in dimension d>2d>2, the conjectured fibered resonances would yield a hammock-shaped set in the lower complex half-plane, connecting some point O(λ2)O(\lambda^{2}) on the real axis to ++\infty,

Σλ:={|k|2+zk,λ:kd}{|k|2+λ2(βkiαk):kd}.\qquad\Sigma_{\lambda}:=\{|k|^{2}+z_{k,\lambda}:k\in\mathbb{R}^{d}\}\approx\{|k|^{2}+\lambda^{2}(\beta_{k}-i\alpha_{k}):k\in\mathbb{R}^{d}\}.

This set is increasingly thinner at infinity and can reach a thickness O(λ2)O(\lambda^{2}) 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 Hλ=+λVH_{\lambda}=-\triangle+\lambda V on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega). 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 tλ2|logλ|t\ll\lambda^{-2}|\!\log\lambda|, 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 𝒫(Ω)\mathcal{P}^{\prime}(\Omega), 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 uL2(d)u^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}) have Fourier transform supported in the compact set Kd{0}K\subset\mathbb{R}^{d}\setminus\{0\}, with uL2(d)=1\|u^{\circ}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}=1.

  1. (i)

    Under Conjecture Conjecture (LRC) — Local resonance conjecture, there holds in L(d;𝒫(Ω))\operatorname{L}^{\infty}(\mathbb{R}^{d};\mathcal{P}^{\prime}(\Omega)), for all 0λ<λ00\leq\lambda<\lambda_{0},

    uλt(x)=du^(k)eikxit(|k|2+zk,λ)Ψk,λ,1𝒫(Ω),𝒫(Ω)Ψk,λ+;x¯𝑑k+OK,M(λ),\qquad u_{\lambda}^{t}(x)\,=\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it(|k|^{2}+z_{k,\lambda})}\,\langle\Psi_{k,\lambda}^{-},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}\,\Psi_{k,\lambda}^{+;x}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k\,+\,O_{K,M}(\lambda), (2.6)

    where we have set Ψk,λ+;x,ϕ𝒫(Ω),𝒫(Ω):=Ψk,λ+,ϕ(x,)𝒫(Ω),𝒫(Ω)\langle\Psi_{k,\lambda}^{+;x},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}:=\langle\Psi^{+}_{k,\lambda},\phi(-x,\cdot)\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)} for ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega). More precisely, this means for all ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega),

    supx|𝔼[ϕ¯uλt(x)]du^(k)eikxit(|k|2+zk,λ)Ψk,λ+;x,ϕ𝒫(Ω),𝒫(Ω)¯Ψk,λ,1𝒫(Ω),𝒫(Ω)¯𝑑k|ϕ,K,MλuL2(d).\qquad\sup_{x}\bigg{|}\mathbb{E}\big{[}{\bar{\phi}\,u_{\lambda}^{t}(x)}\big{]}\,-\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it(|k|^{2}+z_{k,\lambda})}\overline{\langle\Psi^{+;x}_{k,\lambda},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\Psi_{k,\lambda}^{-},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k\bigg{|}\,\\ \lesssim_{\phi,K,M}\,\lambda\|u^{\circ}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}.

    In addition, the averaged wavefunction satisfies the following improved estimate,

    supx|𝔼[uλt(x)]du^(k)eikxit(|k|2+zk,λ)¯𝑑k|K,Mλ2.\qquad\sup_{x}\bigg{|}\mathbb{E}\left[{u_{\lambda}^{t}(x)}\right]\,-\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it(|k|^{2}+z_{k,\lambda})}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k\bigg{|}\,\lesssim_{K,M}\,\lambda^{2}.
  2. (ii)

    Under Conjecture Conjecture (GRC) — Global resonance conjecture, the same holds as in (i) with the errors O(λ)O(\lambda) and O(λ2)O(\lambda^{2}) improved into O(λet8Mλρ)O(\lambda e^{-\frac{t}{8M}\lambda^{\rho}}) and O(λ2et8Mλρ)O(\lambda^{2}e^{-\frac{t}{8M}\lambda^{\rho}}), respectively.

Moreover, for kKk\in K and 0<λ<λ00<\lambda<\lambda_{0}, the restriction of the spectrum of Hk,λstH_{k,\lambda}^{\operatorname{st}} to (1M,1M)(-\frac{1}{M},\frac{1}{M}) 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 λ2(βkiαk)\lambda^{2}(\beta_{k}-i\alpha_{k}); 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 kKk\in K and ϕ,ϕ𝒫(Ω)\phi,\phi^{\prime}\in\mathcal{P}(\Omega) the map

[0,λ0)×𝒫(Ω)×𝒫(Ω)×Lloc(1MB):λ(zk,λ,Ψk,λ+,Ψk,λ,ζk,λϕ,ϕ)[0,\lambda_{0})\to\mathbb{C}\times\mathcal{P}^{\prime}(\Omega)\times\mathcal{P}^{\prime}(\Omega)\times\operatorname{L}^{\infty}_{\operatorname{loc}}(\tfrac{1}{M}B):\lambda\mapsto(z_{k,\lambda},\Psi_{k,\lambda}^{+},\Psi_{k,\lambda}^{-},\zeta_{k,\lambda}^{\phi^{\prime},\phi}) (2.7)

is of class C2C^{2}, then up to a gauge transformation there hold as λ0\lambda\downarrow 0, for all kKk\in K,

zk,λ\displaystyle z_{k,\lambda} =\displaystyle= λ2(βkiαk)+ok(λ2),\displaystyle\lambda^{2}(\beta_{k}-i\alpha_{k})+o_{k}(\lambda^{2}),
Ψk,λ±\displaystyle\Psi_{k,\lambda}^{\pm} =\displaystyle= 1+λΦk1,±+λ2Φk2,±+ok(λ2),\displaystyle 1+\lambda\Phi^{1,\pm}_{k}+\lambda^{2}\Phi^{2,\pm}_{k}+o_{k}(\lambda^{2}),

where αk,βk\alpha_{k},\beta_{k} are defined in (2.3) and where Φk1,±\Phi^{1,\pm}_{k} and Φk2,±\Phi^{2,\pm}_{k} are given by

Φk1,±\displaystyle\Phi_{k}^{1,\pm} :=\displaystyle:= (Hk,0sti0)1V,\displaystyle-(H_{k,0}^{\operatorname{st}}\mp i0)^{-1}V,
Φk2,±\displaystyle\Phi_{k}^{2,\pm} :=\displaystyle:= (Hk,0sti0)1ΠV(Hk,0sti0)1V,\displaystyle(H_{k,0}^{\operatorname{st}}\mp i0)^{-1}\Pi V(H_{k,0}^{\operatorname{st}}\mp i0)^{-1}V,

in terms of the projection Π:=Id𝔼\Pi:=\operatorname{Id}-\mathbb{E} onto L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C}. More precisely, the latter formulas are understood as follows, for all ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega),

Φk1,±,ϕ𝒫(Ω),𝒫(Ω)\displaystyle\langle\Phi_{k}^{1,\pm},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)} :=limε0𝔼[V(Hk,0st±iε)1ϕ],\displaystyle\leavevmode\nobreak\ :=\leavevmode\nobreak\ -\lim_{\varepsilon\downarrow 0}\mathbb{E}\left[{V(H_{k,0}^{\operatorname{st}}\pm i\varepsilon)^{-1}\phi}\right],
Φk2,±,ϕ𝒫(Ω),𝒫(Ω)\displaystyle\langle\Phi_{k}^{2,\pm},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)} :=limε0𝔼[V(Hk,0st±iε)1V(Hk,0st±iε)1Πϕ].\displaystyle\leavevmode\nobreak\ :=\leavevmode\nobreak\ \lim_{\varepsilon\downarrow 0}\mathbb{E}\left[{V(H_{k,0}^{\operatorname{st}}\pm i\varepsilon)^{-1}V(H_{k,0}^{\operatorname{st}}\pm i\varepsilon)^{-1}\Pi\phi}\right].\qed
Remark 2.2 (Full Rayleigh–Schrödinger series for resonances).

The proof of the above is easily pursued to any order. For n1n\geq 1 and ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega), if the map (2.7) is of class CnC^{n}, then there hold as λ0\lambda\downarrow 0, for all kKk\in K,

zk,λ=m=1n1λm+1νkm+ok(λn),Ψk,λ±=m=0nλmΦkm,±+ok(λn),z_{k,\lambda}=\sum_{m=1}^{n-1}\lambda^{m+1}\nu_{k}^{m}+o_{k}(\lambda^{n}),\qquad\Psi_{k,\lambda}^{\pm}=\sum_{m=0}^{n}\lambda^{m}\Phi_{k}^{m,\pm}+o_{k}(\lambda^{n}), (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 L2(Ω)\operatorname{L}^{2}(\Omega) (in link with the dissolution of the bound state), but can be constructed in the weak sense of 𝒫(Ω)\mathcal{P}^{\prime}(\Omega); this partially answers in our setting a question raised in [21, p.179]. More precisely, for all ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega), we can write

νkm:=Φkm,,V𝒫(Ω),𝒫(Ω)=limε0𝔼[Vϕkm,ε,¯],\displaystyle\nu_{k}^{m}:=\langle\Phi_{k}^{m,-},V\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}=\lim_{\varepsilon\downarrow 0}\mathbb{E}\left[{V\overline{\phi_{k}^{m,\varepsilon,-}}}\right],
Φkm,±,ϕ𝒫(Ω),𝒫(Ω):=limε0𝔼[ϕϕkm,ε,±¯],\displaystyle\langle\Phi_{k}^{m,\pm},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}:=\lim_{\varepsilon\downarrow 0}\mathbb{E}\left[{\phi\,\overline{\phi_{k}^{m,\varepsilon,\pm}}}\right],

where the limits indeed exist and where for all ε>0\varepsilon>0 the sequence (ϕkm,ε,±)mL2(Ω)(\phi_{k}^{m,\varepsilon,\pm})_{m}\subset\operatorname{L}^{2}(\Omega) is defined iteratively as follows: we set ϕk0,ε,±=1\phi_{k}^{0,\varepsilon,\pm}=1 and for all m0m\geq 0 we define ϕkm+1,ε,±\phi_{k}^{m+1,\varepsilon,\pm} as the unique solution of the regularized Rayleigh–Schrödinger recurrence equation,

(Hk,0stiε)ϕkm+1,ε,±=Vϕkm,ε,±+l=0m𝔼[Vϕkl,ε,±]ϕkml,ε,±.\qquad\big{(}H_{k,0}^{\operatorname{st}}\mp i\varepsilon\big{)}\phi_{k}^{m+1,\varepsilon,\pm}=-V\phi_{k}^{m,\varepsilon,\pm}+\sum_{l=0}^{m}\mathbb{E}\left[{V\phi_{k}^{l,\varepsilon,\pm}}\right]\phi_{k}^{m-l,\varepsilon,\pm}.

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 (Ω,)(\Omega,\mathbb{P}), we start by recalling the classical notion of stationarity. In particular, a Gaussian field V0V_{0}, that is, a family V0={V0(x,)}xdV_{0}=\{V_{0}(x,\cdot)\}_{x\in\mathbb{R}^{d}} of Gaussian random variables, is an example of a stationary measurable random field if the variables {V0(x,)}xd\{V_{0}(x,\cdot)\}_{x\in\mathbb{R}^{d}} have the same expectation and have covariance K0(x,y):=Cov[V0(x,);V0(y,)]K_{0}(x,y):=\operatorname{Cov}\left[{V_{0}(x,\cdot)};{V_{0}(y,\cdot)}\right] of the form K0(x,y)=𝒞0(xy)K_{0}(x,y)=\mathcal{C}_{0}(x-y) with 𝒞0\mathcal{C}_{0} continuous at the origin.

Definition 3.1.

A random field on d\mathbb{R}^{d} is a map ϕ:d×Ω\phi:\mathbb{R}^{d}\times\Omega\to\mathbb{C} such that for all xdx\in\mathbb{R}^{d} the random variable ϕ(x,):Ω\phi(x,\cdot):\Omega\to\mathbb{C} is measurable. It is said to be stationary if its finite-dimensional law is shift-invariant, that is, if for any finite set EdE\subset\mathbb{R}^{d} the law of {ϕ(x+y,)}xE\{\phi(x+y,\cdot)\}_{x\in E} does not depend on the shift ydy\in\mathbb{R}^{d}. In addition, it is said to be measurable if the map ϕ:d×Ω\phi:\mathbb{R}^{d}\times\Omega\to\mathbb{R} 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 xx and for all δ>0\delta>0 there holds {ωΩ:|ϕ(x+y,ω)ϕ(x,ω)|>δ}0\mathbb{P}\{\omega\in\Omega:|\phi(x+y,\omega)-\phi(x,\omega)|>\delta\}\to 0 as y0y\to 0.) ∎

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 τ:=(τx)xd\tau:=(\tau_{x})_{x\in\mathbb{R}^{d}} of the group (d,+)(\mathbb{R}^{d},+) on (Ω,)(\Omega,\mathbb{P}) is a collection of measurable maps τx:ΩΩ\tau_{x}:\Omega\to\Omega that satisfy

  • τxτy=τx+y\tau_{x}\circ\tau_{y}=\tau_{x+y} for all x,ydx,y\in\mathbb{R}^{d};

  • [τxA]=[A]\mathbb{P}\left[{\tau_{x}A}\right]=\mathbb{P}\left[{A}\right] for all xdx\in\mathbb{R}^{d} and measurable AΩA\subset\Omega;

  • the map d×ΩΩ:(x,ω)τxω\mathbb{R}^{d}\times\Omega\to\Omega:(x,\omega)\mapsto\tau_{x}\omega is jointly measurable.

A random field ϕ:d×Ω\phi:\mathbb{R}^{d}\times\Omega\to\mathbb{C} is said to be τ\tau-stationary if there exists a measurable map ϕ:Ω\phi_{\circ}:\Omega\to\mathbb{C} such that ϕ(x,ω)=ϕ(τxω)\phi(x,\omega)=\phi_{\circ}(\tau_{-x}\omega) for all x,ωx,\omega. ∎

This second definition yields a bijection between random variables ϕ:Ω\phi_{\circ}:\Omega\to\mathbb{C} and τ\tau-stationary random fields ϕ:d×Ω\phi:\mathbb{R}^{d}\times\Omega\to\mathbb{C}. The random field ϕ\phi is referred to as the τ\tau-stationary extension of ϕ\phi_{\circ}. In addition, given ϕLp(Ω)\phi_{\circ}\in\operatorname{L}^{p}(\Omega) with p1p\geq 1, since there holds 𝔼[K|ϕ|p]=|K|𝔼[|ϕ|p]\mathbb{E}\left[{\int_{K}|\phi|^{p}}\right]=|K|\,\mathbb{E}\left[{|\phi_{\circ}|^{p}}\right] for any compact KdK\subset\mathbb{R}^{d}, we deduce that the realization ϕ(,ω)\phi(\cdot,\omega) belongs to Llocp(d)\operatorname{L}^{p}_{\operatorname{loc}}(\mathbb{R}^{d}) for almost all ω\omega. The Banach space Lp(Ω)\operatorname{L}^{p}(\Omega) can thus be identified with the subspace of τ\tau-stationary random fields in Llocp(d;Lp(Ω))\operatorname{L}^{p}_{\operatorname{loc}}(\mathbb{R}^{d};\operatorname{L}^{p}(\Omega)).

While the notion of τ\tau-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 ϕ\phi be a stationary measurable random field defined on (Ω,)(\Omega,\mathbb{P}) in the sense of Definition 3.1. Then there exist a probability space (Ω,)(\Omega^{\prime},\mathbb{P}^{\prime}), endowed with a measurable action τ\tau, and a τ\tau-stationary random field ϕ\phi^{\prime} defined on (Ω,)(\Omega^{\prime},\mathbb{P}^{\prime}) in the sense of Definition 3.2 such that ϕ\phi and ϕ\phi^{\prime} have the same finite-dimensional law. This extends to a correspondence between σ(ϕ)\sigma(\phi)-measurable random variables on Ω\Omega and random variables on Ω\Omega^{\prime}. ∎

Proof.

The proof is a variant of e.g. [31, Section 16.1]. Let Ω\Omega^{\prime} denote the set of measurable functions d\mathbb{R}^{d}\to\mathbb{C}, endowed with the cylindrical σ\sigma-algebra \mathcal{F}^{\prime}, and consider the map H:ΩΩ:ωϕ(,ω)H:\Omega\to\Omega^{\prime}:\omega\mapsto\phi(\cdot,\omega). This map is measurable and induces a probability measure :=H\mathbb{P}^{\prime}:=H_{*}\mathbb{P} on the measurable space (Ω,)(\Omega^{\prime},\mathcal{F}^{\prime}). Next, define τx:ΩΩ\tau_{x}:\Omega^{\prime}\to\Omega^{\prime} by (τxω)(y):=ω(yx)(\tau_{x}\omega^{\prime})(y):=\omega^{\prime}(y-x). As ϕ\phi is jointly measurable and stationary, we find that τ\tau is a measurable action. Finally, we set ϕ(ω):=ω(0)\phi^{\prime}_{\circ}(\omega^{\prime}):=\omega^{\prime}(0), with τ\tau-stationary extension ϕ(x,ω):=ω(x)\phi^{\prime}(x,\omega^{\prime}):=\omega^{\prime}(x), and the claim follows. We omit the details. ∎

Henceforth, we focus on the more convenient notion of τ\tau-stationarity in the sense of Definition 3.2: we implicitly assume that the reference probability space (Ω,)(\Omega,\mathbb{P}) is endowed with a given measurable action τ\tau and we assume that the random potential VV is τ\tau-stationary. In the sequel, for abbreviation, τ\tau-stationarity is simply referred to as stationarity, and we abusively use the same notation for ϕ\phi and ϕ\phi_{\circ} (in particular, for VV and VV_{\circ}).

Remarks 3.4.

  1. (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 (Ω,)(\Omega,\mathbb{P}) is chosen as the torus 𝕋d\mathbb{T}^{d} endowed with the Lebesgue measure, the action τ\tau is given by τxω=ω+x\tau_{-x}\omega=\omega+x on 𝕋d\mathbb{T}^{d}, and we set ϕ(x,ω)=ϕ(ω+x)\phi(x,\omega)=\phi_{\circ}(\omega+x). In the quasiperiodic setting, the probability space is chosen as a higher-dimensional torus 𝕋M\mathbb{T}^{M} with M>dM>d, endowed with the Lebesgue measure, the action τ\tau is given by τxω=ω+Fx\tau_{-x}\omega=\omega+Fx on 𝕋M\mathbb{T}^{M} in terms of the winding matrix FM×dF\in\mathbb{R}^{M\times d}, and we set ϕ(x,ω)=ϕ(ω+Fx)\phi(x,\omega)=\phi_{\circ}(\omega+Fx). In both cases, the construction is viewed as introducing a uniform random shift.

  2. (b)

    Any d\mathbb{Z}^{d}-stationary random potential (that is, satisfying the stationarity assumption for an action of (d,+)(\mathbb{Z}^{d},+) on Ω\Omega) can also be seen as a stationary random potential in the above sense up to considering the random ensemble of shifts. Indeed, assume that τ:=(τz)zd\tau^{\prime}:=(\tau_{z}^{\prime})_{z\in\mathbb{Z}^{d}} is a measurable action of (d,+)(\mathbb{Z}^{d},+) on a probability space (Ω,)(\Omega^{\prime},\mathbb{P}^{\prime}), and that ϕ\phi is τ\tau^{\prime}-stationary, that is, ϕ(x+z,ω)=ϕ(x,τzω)\phi(x+z,\omega)=\phi(x,\tau^{\prime}_{-z}\omega) for all xdx\in\mathbb{R}^{d}, zdz\in\mathbb{Z}^{d}, and ωΩ\omega\in\Omega^{\prime}. Endow Ω:=Ω×[0,1)d\Omega:=\Omega^{\prime}\times[0,1)^{d} with the product measure :=×eb\mathbb{P}:=\mathbb{P}^{\prime}\times\mathcal{L}eb, where eb\mathcal{L}eb denotes the Lebesgue measure on [0,1)d[0,1)^{d}, and define the action τ:=(τx)xd\tau:=(\tau_{x})_{x\in\mathbb{R}^{d}} of (d,+)(\mathbb{R}^{d},+) on Ω=Ω×[0,1)d\Omega=\Omega^{\prime}\times[0,1)^{d} by

    τx(ω,y):=(τxω,y+xy+x),\tau_{x}(\omega,y):=\big{(}\tau^{\prime}_{\lfloor x\rfloor}\omega\leavevmode\nobreak\ ,\leavevmode\nobreak\ y+x-\lfloor y+x\rfloor\big{)},

    where x=(x1,,xd)\lfloor x\rfloor=(\lfloor x_{1}\rfloor,\ldots,\lfloor x_{d}\rfloor) for xdx\in\mathbb{R}^{d} and where a\lfloor a\rfloor denotes the largest integer a\leq a for aa\in\mathbb{R}. The map ψ(x,(ω,y)):=ϕ(xy,ω)\psi(x,(\omega,y)):=\phi(x-y,\omega) then defines a τ\tau-stationary random field on d×Ω\mathbb{R}^{d}\times\Omega. ∎

3.2. Stationary differential calculus

A differential calculus is naturally developed on L2(Ω)\operatorname{L}^{2}(\Omega) via the measurable action τ\tau on (Ω,)(\Omega,\mathbb{P}). Indeed, while the subspace of stationary random fields in Lloc2(d;L2(Ω))\operatorname{L}^{2}_{\operatorname{loc}}(\mathbb{R}^{d};\operatorname{L}^{2}(\Omega)) is identified with the Hilbert space L2(Ω)\operatorname{L}^{2}(\Omega), the spatial weak gradient \nabla on locally square integrable functions turns into a densely defined linear operator st\nabla^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega), which is referred to as the stationary gradient. Equivalently, st\nabla^{\operatorname{st}} can be viewed as the infinitesimal generator of the group of isometries {Tx:ϕϕ(τx)}xd\{T_{x}:\phi_{\circ}\mapsto\phi_{\circ}(\tau_{-x}\cdot)\}_{x\in\mathbb{R}^{d}} on L2(Ω)\operatorname{L}^{2}(\Omega). The adjoint is (st)=st(\nabla^{\operatorname{st}})^{*}=-\nabla^{\operatorname{st}} and we denote by st=stst-\triangle^{\operatorname{st}}=-\nabla^{\operatorname{st}}\cdot\nabla^{\operatorname{st}} the corresponding stationary Laplacian. For all s0s\geq 0, we define the (Hilbert) space Hs(Ω)H^{s}(\Omega) as the space of all elements ϕL2(Ω)\phi_{\circ}\in\operatorname{L}^{2}(\Omega) for which the stationary extension ϕ\phi belongs to Hlocs(d;L2(Ω))H^{s}_{\operatorname{loc}}(\mathbb{R}^{d};\operatorname{L}^{2}(\Omega)), and we denote by Hs(Ω)H^{-s}(\Omega) the dual of Hs(Ω)H^{s}(\Omega). Note that H1(Ω)H^{1}(\Omega) coincides with the domain of st\nabla^{\operatorname{st}}, and that the stationary Laplacian st-\triangle^{\operatorname{st}} is self-adjoint on H2(Ω)H^{2}(\Omega). 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 st-\triangle^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega) typically has absolutely continuous spectrum and no spectral gap above 0, cf. Section 4.1. This entails that Poincaré’s inequality does not hold on H1(Ω)H^{1}(\Omega) and that compact embeddings such as Rellich’s theorem also fail. This lack of compactness is related to the fact that the gradient st\nabla^{\operatorname{st}} only contains information on a finite set of directions while Ω\Omega 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 uL2(d)u\in\operatorname{L}^{2}(\mathbb{R}^{d}), its Floquet transform is (formally) defined by

𝒱peru(k,x):=ndeik(x+n)u(x+n),\mathcal{V}^{\circ}_{\operatorname{per}}u(k,x)\,:=\,\sum_{n\in\mathbb{Z}^{d}}e^{-ik\cdot(x+n)}u(x+n),

which is periodic in xx, so that the Fourier inversion formula takes the form

u(x)=2π𝕋deikx𝒱peru(k,x)¯𝑑k,u(x)\,=\,\int_{2\pi\mathbb{T}^{d}}e^{ik\cdot x}\,\mathcal{V}^{\circ}_{\operatorname{per}}u(k,x)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,

thus leading to the following direct integral decomposition, e.g. [48, p.280],

L2(d)=2π𝕋dL2(𝕋d)𝔢k¯𝑑k,𝔢k(x):=eikx.\operatorname{L}^{2}(\mathbb{R}^{d})\,=\,\int_{2\pi\mathbb{T}^{d}}^{\oplus}\operatorname{L}^{2}(\mathbb{T}^{d})\,\mathfrak{e}_{k}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,\qquad\mathfrak{e}_{k}(x):=e^{ik\cdot x}.

This decomposition allows for a simple adaptation to the augmented space L2(d×𝕋d)\operatorname{L}^{2}(\mathbb{R}^{d}\times\mathbb{T}^{d}): given uL2(d×𝕋d)u\in\operatorname{L}^{2}(\mathbb{R}^{d}\times\mathbb{T}^{d}), its Floquet transform is defined by

𝒱peru(k,q):=deikyu(y,qy)𝑑y,\mathcal{V}_{\operatorname{per}}u(k,q)\,:=\,\int_{\mathbb{R}^{d}}e^{-ik\cdot y}u(y,q-y)\,dy,

which is periodic in qq, so that the Fourier inversion formula takes the form

u(x,q)=deikx𝒱peru(k,x+q)¯𝑑k,u(x,q)\,=\,\int_{\mathbb{R}^{d}}e^{ik\cdot x}\,\mathcal{V}_{\operatorname{per}}u(k,x+q)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,

and leads to the direct integral decomposition

L2(d×𝕋d)=dL2(𝕋d)𝔢k¯𝑑k.\operatorname{L}^{2}(\mathbb{R}^{d}\times\mathbb{T}^{d})=\int_{\mathbb{R}^{d}}^{\oplus}\operatorname{L}^{2}(\mathbb{T}^{d})\,\mathfrak{e}_{k}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k.

We may now mimick this construction in the general stationary random setting: given uL2(d×Ω)u\in\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega), its stationary Floquet transform is defined by

𝒱stu(k,ω):=deikyu(y,τyω)𝑑y,\mathcal{V}_{\operatorname{st}}u(k,\omega)\,:=\,\int_{\mathbb{R}^{d}}e^{-ik\cdot y}u(y,\tau_{y}\omega)\,dy,

so that the Fourier inversion formula takes the form

u(x,ω)=deikx𝒱stu(k,τxω)¯𝑑k,u(x,\omega)\,=\,\int_{\mathbb{R}^{d}}e^{ik\cdot x}\,\mathcal{V}_{\operatorname{st}}u(k,\tau_{-x}\omega)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,

and leads to the direct integral decomposition

L2(d×Ω)=dL2(Ω)𝔢k¯𝑑k.\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega)=\int_{\mathbb{R}^{d}}^{\oplus}\operatorname{L}^{2}(\Omega)\,\mathfrak{e}_{k}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k. (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 𝒱st\mathcal{V}_{\operatorname{st}} is a unitary operator on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega), and satisfies

  1. (i)

    𝒱stg=g^\mathcal{V}_{\operatorname{st}}g=\widehat{g} for all gL2(d)L2(d×Ω)g\in\operatorname{L}^{2}(\mathbb{R}^{d})\subset\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega);

  2. (ii)

    𝒱st(ϕu)=ϕ𝒱stu\mathcal{V}_{\operatorname{st}}(\phi u)=\phi_{\circ}\mathcal{V}_{\operatorname{st}}u for all ϕL2(Ω)\phi_{\circ}\in\operatorname{L}^{2}(\Omega) and uL2(d×Ω)u\in\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega) with ϕuL2(d×Ω)\phi u\in\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega).∎

3.4. Stationary Floquet–Bloch fibration

In view of (3.1), the stationary Floquet transform 𝒱st\mathcal{V}_{\operatorname{st}} decomposes differential operators with stationary random coefficients (such as the Schrödinger operator HλH_{\lambda}) into a direct integral of “elementary” fibered operators on the stationary space L2(Ω)\operatorname{L}^{2}(\Omega). First, the Laplacian -\triangle on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega) is self-adjoint on H2(d;L2(Ω))H^{2}(\mathbb{R}^{d};\operatorname{L}^{2}(\Omega)) and is mapped by 𝒱st\mathcal{V}_{\operatorname{st}} on

𝒱st[()u](k,ω)=(st+ik)(st+ik)𝒱stu(k,ω)=(Hk,0st+|k|2)𝒱stu(k,ω),\mathcal{V}_{\operatorname{st}}[(-\triangle)u](k,\omega)\,=\,-(\nabla^{\operatorname{st}}+ik)\cdot(\nabla^{\operatorname{st}}+ik)\mathcal{V}_{\operatorname{st}}u(k,\omega)\,=\,(H^{\operatorname{st}}_{k,0}+|k|^{2})\mathcal{V}_{\operatorname{st}}u(k,\omega), (3.2)

in terms of the (centered) fibered Laplacian

Hk,0st:=(st+ik)(st+ik)|k|2=st2ikst,H^{\operatorname{st}}_{k,0}\,:=\,-(\nabla^{\operatorname{st}}+ik)\cdot(\nabla^{\operatorname{st}}+ik)-|k|^{2}=-\triangle^{\operatorname{st}}-2ik\cdot\nabla^{\operatorname{st}},

As the stationary Laplacian st-\triangle^{\operatorname{st}} is self-adjoint on H2(Ω)H^{2}(\Omega) and as 2ikst-2ik\cdot\nabla^{\operatorname{st}} is an infinitesimal perturbation, the Kato-Rellich theorem ensures that this fibered Laplacian Hk,0stH_{k,0}^{\operatorname{st}} is also self-adjoint on H2(Ω)H^{2}(\Omega), and the centering ensures that constant functions belong to its kernel. Next, if the stationary random potential VV is uniformly bounded (the unbounded case is postponed to Appendix A), it defines a bounded multiplication operator on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega) and the corresponding Schrödinger operator Hλ=+λVH_{\lambda}=-\triangle+\lambda V is thus self-adjoint on H2(d;L2(Ω))H^{2}(\mathbb{R}^{d};\operatorname{L}^{2}(\Omega)). Combining (3.2) with Lemma 3.5(ii), we find

𝒱st[Hλf](k,ω)=(Hk,λst+|k|2)𝒱stf(k,ω),\mathcal{V}_{\operatorname{st}}[H_{\lambda}f](k,\omega)=(H^{\operatorname{st}}_{k,\lambda}+|k|^{2})\mathcal{V}_{\operatorname{st}}f(k,\omega), (3.3)

in terms of the (centered) fibered Schrödinger operator

Hk,λst:=Hk,0st+λV,H^{\operatorname{st}}_{k,\lambda}\,:=\,H_{k,0}^{\operatorname{st}}+\lambda V,

which is self-adjoint on H2(Ω)H^{2}(\Omega). Using direct integral representation, e.g. [48, p.280], we may reformulate the above as

(Hλ,L2(d×Ω))=d(Hk,λst+|k|2,L2(Ω))𝔢k¯𝑑k.\big{(}H_{\lambda},\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega)\big{)}\,=\,\int_{\mathbb{R}^{d}}^{\oplus}\big{(}H_{k,\lambda}^{\operatorname{st}}+|k|^{2},\operatorname{L}^{2}(\Omega)\big{)}\,\mathfrak{e}_{k}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k. (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 uL2(d)L2(d×Ω)u^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d})\subset\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega), appealing to (3.3) and to Lemma 3.5(i),

uλt(x,ω)=(eitHλu)(x,ω)\displaystyle u_{\lambda}^{t}(x,\omega)\,=\,\big{(}e^{-itH_{\lambda}}u^{\circ}\big{)}(x,\omega) =\displaystyle= deikx𝒱st[eitHλu](k,τxω)¯𝑑k\displaystyle\int_{\mathbb{R}^{d}}e^{ik\cdot x}\,\mathcal{V}_{\operatorname{st}}\big{[}e^{-itH_{\lambda}}u^{\circ}\big{]}(k,\tau_{-x}\omega)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k
=\displaystyle= deikxit|k|2(eitHk,λst𝒱stu(k,))(τxω)¯𝑑k\displaystyle\int_{\mathbb{R}^{d}}e^{ik\cdot x-it|k|^{2}}\,\big{(}e^{-itH_{k,\lambda}^{\operatorname{st}}}\mathcal{V}_{\operatorname{st}}u^{\circ}(k,\cdot)\big{)}(\tau_{-x}\omega)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k
=\displaystyle= du^(k)eikxit|k|2(eitHk,λst1)(τxω)¯𝑑k,\displaystyle\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it|k|^{2}}\,\big{(}e^{-itH_{k,\lambda}^{\operatorname{st}}}1\big{)}(\tau_{-x}\omega)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,

that is, (1.4). Alternatively, in terms of the L2(Ω)\operatorname{L}^{2}(\Omega)-valued spectral measure μk,λ1\mu_{k,\lambda}^{1} of Hk,λstH^{\operatorname{st}}_{k,\lambda} associated with the constant function 11,

uλt(x,ω)\displaystyle u_{\lambda}^{t}(x,\omega) =\displaystyle= du^(k)eikxit(|k|2+κ)𝑑μk,λ1(κ)(τxω)¯𝑑k.\displaystyle\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\int_{\mathbb{R}}e^{ik\cdot x-it(|k|^{2}+\kappa)}\,d\mu_{k,\lambda}^{1}(\kappa)(\tau_{-x}\omega)\,\,{{\mathchar 22\relax\mkern-12.0mud}}k.

For vanishing disorder λ=0\lambda=0 the spectral measures take the form dμk,01=dδ0d\mu_{k,0}^{1}=d\delta_{0} and we recover the Fourier diagonalization of the free Schrödinger flow, cf. (1.5), while for λ>0\lambda>0 each Fourier mode eikxe^{ik\cdot x} is deformed into a “Bloch measure” eikxdμk,λ1e^{ik\cdot x}d\mu_{k,\lambda}^{1}. In the periodic setting the measure μk,λ1\mu_{k,\lambda}^{1} is known to be discrete, leading to the deformation of the plane wave eikxe^{ik\cdot x} into a superposition of so-called Bloch waves, cf. [32, 3]. The picture is very different in the random setting as μk,λ1\mu_{k,\lambda}^{1} 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 VV 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 kst+Vper+λV-\triangle^{\operatorname{st}}_{k}+V_{\operatorname{per}}+\lambda V, where the periodic potential VperV_{\operatorname{per}} models the underlying crystalline structure.

4.1. Unperturbed fibered operators

We give a full account of the spectral properties of the unperturbed operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} on L2(Ω)\operatorname{L}^{2}(\Omega). We start with some general definitions. For ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega), we denote its covariance function by 𝒞ϕ(x):=𝔼[ϕ¯()ϕ(τx)]\mathcal{C}^{\phi}(x):=\mathbb{E}\big{[}{\bar{\phi}(\cdot)\,\phi(\tau_{-x}\cdot)}\big{]}, which belongs to L(d)\operatorname{L}^{\infty}(\mathbb{R}^{d}) and is positive definite. By Bochner’s theorem, the distributional Fourier transform 𝒞^ϕ\widehat{\mathcal{C}}^{\phi} is then a nonnegative finite measure on d\mathbb{R}^{d} with total mass 𝒞^ϕ(d)=(2π)dϕL2(Ω)2\widehat{\mathcal{C}}^{\phi}(\mathbb{R}^{d})=(2\pi)^{d}\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}, and is called the spectral measure of ϕ\phi. 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 (Ω,)(\Omega,\mathbb{P}) endowed with a given stationarity structure is defined as the subset

Ω^:={(2π)d𝒞^ϕ:ϕL2(Ω),ϕL2(Ω)=1}𝒫(d).\widehat{\Omega}\,:=\,\big{\{}(2\pi)^{-d}\widehat{\mathcal{C}}^{\phi}:\phi\in\operatorname{L}^{2}(\Omega),\,\|\phi\|_{\operatorname{L}^{2}(\Omega)}=1\big{\}}\,\subset\,\mathcal{P}(\mathbb{R}^{d}).\qed

We show that the spectrum of the unperturbed operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} can be completely characterized in terms of properties of Ω^\widehat{\Omega}.

Lemma 4.2.

Let VV be a stationary random field and assume that the underlying probability space (Ω,)(\Omega,\mathbb{P}) is endowed with the σ\sigma-algebra generated by VV. For ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega) and kdk\in\mathbb{R}^{d} we denote by νkϕ\nu^{\phi}_{k} the probability measure on +\mathbb{R}^{+} defined by

νkϕ([0,t]):=(2π)d𝒞^ϕ(Bt(k)¯),t0,\displaystyle\nu^{\phi}_{k}([0,t])\,:=\,(2\pi)^{-d}\widehat{\mathcal{C}}^{\phi}\big{(}\overline{B_{t}(-k)}\big{)},\qquad t\geq 0, (4.1)

and we consider its Lebesgue decomposition

νkϕ=νk;ppϕ+νk;scϕ+νk;acϕ\nu^{\phi}_{k}=\nu^{\phi}_{k;{\operatorname{pp}}}+\nu^{\phi}_{k;{\operatorname{sc}}}+\nu^{\phi}_{k;{\operatorname{ac}}}

into pure point, singularly continuous, and absolutely continuous parts. Then,

  1. (i)

    The spectrum σ(Hk,0st)\sigma(H_{k,0}^{\operatorname{st}}) of the operator Hk,0stH_{k,0}^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega) is included in [|k|2,)[-|k|^{2},\infty) and there is an eigenvalue at 0.

  2. (ii)

    For =pp,sc,ac\ast={\operatorname{pp}},{\operatorname{sc}},{\operatorname{ac}}, there holds

    σ(Hk,0st)=adh(𝒞^ϕΩ^hk(suppνk;ϕ)),hk(t):=t2|k|2.\displaystyle\sigma_{\ast}(H_{k,0}^{\operatorname{st}})={\operatorname{adh\,}}\bigg{(}\bigcup_{\widehat{\mathcal{C}}^{\phi}\in\widehat{\Omega}}h_{k}\big{(}\operatorname{supp}\nu^{\phi}_{k;\ast}\big{)}\bigg{)},\qquad h_{k}(t):=t^{2}-|k|^{2}.
  3. (iii)

    The density of the absolutely continuous part of the spectral measure of Hk,0stH_{k,0}^{\operatorname{st}} associated with ϕ\phi takes the form

    dμk,0;acϕ,ϕdλ(λ)=𝟙[|k|2,)(λ)2λ+|k|2dνk;acϕdt(λ+|k|2).\frac{d\mu^{\phi,\phi}_{k,0;{\operatorname{ac}}}}{d\lambda}(\lambda)\,=\,\frac{\mathds{1}_{[-|k|^{2},\infty)}(\lambda)}{2\sqrt{\lambda+|k|^{2}}}\,\frac{d\nu^{\phi}_{k;{\operatorname{ac}}}}{dt}\big{(}\sqrt{\lambda+|k|^{2}}\big{)}.\qed
Proof.

First note that the Fourier symbol of Hk,0stH_{k,0}^{\operatorname{st}} is given by y|y+k|2|k|2|k|2y\mapsto|y+k|^{2}-|k|^{2}\geq-|k|^{2}, which easily implies that the operator Hk,0stEH_{k,0}^{\operatorname{st}}-E has bounded inverse on L2(Ω)\operatorname{L}^{2}(\Omega) for all E<|k|2E<-|k|^{2}. The spectrum of Hk,0stH_{k,0}^{\operatorname{st}} is therefore included in [|k|2,)[-|k|^{2},\infty), 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 μk,0ϕ,ϕ\mu_{k,0}^{\phi,\phi} of Hk,0stH_{k,0}^{\operatorname{st}} associated with any ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega). We claim that this spectral measure is explicitly given by the following formula, for all gCb()g\in C_{b}(\mathbb{R}),

g𝑑μk,0ϕ,ϕ=+g(t2|k|2)𝑑νkϕ(t),\displaystyle\int_{\mathbb{R}}g\,d\mu_{k,0}^{\phi,\phi}\,=\,\int_{\mathbb{R}^{+}}g(t^{2}-|k|^{2})\,d\nu_{k}^{\phi}(t), (4.2)

where νkϕ\nu_{k}^{\phi} is defined in the statement. The conclusion directly follows from this claim since it yields for =pp,sc,ac\ast={\operatorname{pp}},{\operatorname{sc}},{\operatorname{ac}},

g𝑑μk,0;ϕ,ϕ=g(t2|k|2)𝑑νk;ϕ(t),\displaystyle\int_{\mathbb{R}}g\,d\mu^{\phi,\phi}_{k,0;\ast}\,=\,\int_{\mathbb{R}}g(t^{2}-|k|^{2})\,d\nu^{\phi}_{k;\ast}(t),

where we denote by μk,0ϕ,ϕ=μk,0;ppϕ,ϕ+μk,0;scϕ,ϕ+μk,0;acϕ,ϕ\mu_{k,0}^{\phi,\phi}=\mu^{\phi,\phi}_{k,0;{\operatorname{pp}}}+\mu^{\phi,\phi}_{k,0;{\operatorname{sc}}}+\mu^{\phi,\phi}_{k,0;{\operatorname{ac}}} the Lebesgue decomposition of μk,0ϕ,ϕ\mu_{k,0}^{\phi,\phi}, and similarly for νkϕ\nu_{k}^{\phi}.

It remains to argue in favor of (4.2). By density, it is enough to prove it for gCb1()g\in C_{b}^{1}(\mathbb{R}). Since the Fourier symbol of Hk,0stH_{k,0}^{\operatorname{st}} is given by y|y+k|2|k|2y\mapsto|y+k|^{2}-|k|^{2}, we compute in Fourier space,

g𝑑μk,0ϕ,ϕ=𝔼[ϕ¯g(Hk,0st)ϕ]=(2π)ddg(|y+k|2|k|2)𝑑𝒞^ϕ(y),\int_{\mathbb{R}}g\,d\mu_{k,0}^{\phi,\phi}=\mathbb{E}\left[{\bar{\phi}\,g(H_{k,0}^{\operatorname{st}})\,\phi}\right]=(2\pi)^{-d}\int_{\mathbb{R}^{d}}g(|y+k|^{2}-|k|^{2})\,d\widehat{\mathcal{C}}^{\phi}(y),

and a radial change of variables then yields

g𝑑μk,0ϕ,ϕ\displaystyle\int_{\mathbb{R}}g\,d\mu_{k,0}^{\phi,\phi} =\displaystyle= (2π)dlimε012εεd𝟙|y+k|εt<|y+k|+εg(|y+k|2|k|2)𝑑𝒞^ϕ(y)𝑑t\displaystyle(2\pi)^{-d}\lim_{\varepsilon\downarrow 0}\frac{1}{2\varepsilon}\int_{-\varepsilon}^{\infty}\int_{\mathbb{R}^{d}}\mathds{1}_{|y+k|-\varepsilon\leq t<|y+k|+\varepsilon}\,g(|y+k|^{2}-|k|^{2})\,d\widehat{\mathcal{C}}^{\phi}(y)\,dt
=\displaystyle= (2π)dlimε012ε0g(t2|k|2)d𝟙|y+k|εt<|y+k|+ε𝑑𝒞^ϕ(y)𝑑t\displaystyle(2\pi)^{-d}\lim_{\varepsilon\downarrow 0}\frac{1}{2\varepsilon}\int_{0}^{\infty}g(t^{2}-|k|^{2})\int_{\mathbb{R}^{d}}\mathds{1}_{|y+k|-\varepsilon\leq t<|y+k|+\varepsilon}\,d\widehat{\mathcal{C}}^{\phi}(y)\,dt
=\displaystyle= limε00g(t2|k|2)νkϕ((tε,t+ε])2ε𝑑t\displaystyle\lim_{\varepsilon\downarrow 0}\int_{0}^{\infty}g(t^{2}-|k|^{2})\,\frac{\nu_{k}^{\phi}((t-\varepsilon,t+\varepsilon])}{2\varepsilon}\,dt
=\displaystyle= 0g(t2|k|2)𝑑νkϕ(t),\displaystyle\int_{0}^{\infty}g(t^{2}-|k|^{2})\,d\nu_{k}^{\phi}(t),

that is, (4.2). ∎

In particular, the above result implies that the spectrum σ(Hk,0st)\sigma(H_{k,0}^{\operatorname{st}}) can be of any type: for any measure μ𝒫([|k|2,))\mu\in\mathcal{P}([-|k|^{2},\infty)) with nontrivial pure point, singularly continuous, and absolutely continuous parts, we can construct a stationary Gaussian field VV such that the spectral measure μk,0V,V\mu_{k,0}^{V,V} coincides with μ\mu, which entails that the corresponding spectrum of Hk,0stH_{k,0}^{\operatorname{st}} admits nontrivial pure point, singularly continuous, and absolutely continuous parts. Moreover, the eigenvalue at 0 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 Hk,0stH_{k,0}^{\operatorname{st}} coincides with the whole interval [|k|2,)[-|k|^{2},\infty) and is made of a simple eigenvalue at 0 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 α\alpha-mixing). We first recall some terminology: For any diameter D>0D>0 and distance R>0R>0, we set

α~(R,D;V):=sup{α(σ({V(x,)}xS1),σ({V(x,)}xS2)):S1,S2(d),dist(S1,S2)R,diam(S1),diam(S2)D},\tilde{\alpha}(R,D;V):=\sup\Big{\{}\alpha\big{(}\sigma(\{V(x,\cdot)\}_{x\in S_{1}}),\sigma(\{V(x,\cdot)\}_{x\in S_{2}})\big{)}:\\ S_{1},S_{2}\in\mathcal{B}(\mathbb{R}^{d}),\,\operatorname{dist}(S_{1},S_{2})\geq R,\,\operatorname{diam}(S_{1}),\operatorname{diam}(S_{2})\leq D\Big{\}}, (4.3)

where Rosenblatt’s α\alpha-mixing coefficient is defined for any two sub-σ\sigma-algebras 𝒜1,𝒜2\mathcal{A}_{1},\mathcal{A}_{2} as

α(𝒜1,𝒜2):=sup{|[G1G2][G1][G2]|:G1𝒜1,G2𝒜2}.\alpha(\mathcal{A}_{1},\mathcal{A}_{2})\,:=\,\sup\Big{\{}\big{|}\mathbb{P}\left[{G_{1}\cap G_{2}}\right]-\mathbb{P}\left[{G_{1}}\right]\mathbb{P}\left[{G_{2}}\right]\!\big{|}:G_{1}\in\mathcal{A}_{1},\,G_{2}\in\mathcal{A}_{2}\Big{\}}.

The random field VV is said to be α\alpha-mixing if for any D<D<\infty there holds α~(R,D;V)0\tilde{\alpha}(R,D;V)\to 0 as RR\uparrow\infty. We may now state the following criterion, which in particular implies Proposition 1 when restricted to the Gaussian setting.

Proposition 4.3.

Let VV be a (nonzero) stationary random field with covariance 𝒞\mathcal{C}, and let the probability space (Ω,)(\Omega,\mathbb{P}) be endowed with the σ\sigma-algebra generated by VV.

  1. (i)

    Assume that one of the following two conditions holds,

    1. (C1)

      VV is Gaussian and 𝒞\mathcal{C} is not periodic in any direction;

    2. (C2)

      𝒞\mathcal{C} has exponential decay, that is, |𝒞(x)|Ce1C|x||\mathcal{C}(x)|\leq Ce^{-\frac{1}{C}|x|} for all xx.

    Then the spectrum σ(Hk,0st)\sigma(H_{k,0}^{\operatorname{st}}) coincides with [|k|2,)[-|k|^{2},\infty).

  2. (ii)

    Assume that one of the following two conditions holds,

    1. (C3)

      VV is Gaussian and the (nonnegative measure) Fourier transform 𝒞^\widehat{\mathcal{C}} is absolutely continuous (this is the case for instance if 𝒞\mathcal{C} is integrable);

    2. (C4)

      VV is α\alpha-mixing and satisfies dα~(|x|,D;V)𝑑x<\int_{\mathbb{R}^{d}}\,\tilde{\alpha}(|x|,D;V)\,dx\,<\,\infty for all D<D<\infty.

    Then the eigenvalue at 0 is simple (with eigenspace \mathbb{C}) and

    σpp(Hk,0st)={0},σsc(Hk,0st)=,σ(Hk,0st)=σac(Hk,0st)=[|k|2,).\qquad\sigma_{{\operatorname{pp}}}(H_{k,0}^{\operatorname{st}})\,=\,\{0\},\quad\sigma_{{\operatorname{sc}}}(H_{k,0}^{\operatorname{st}})=\varnothing,\quad\sigma(H_{k,0}^{\operatorname{st}})\,=\,\sigma_{{\operatorname{ac}}}(H_{k,0}^{\operatorname{st}})\,=\,[-|k|^{2},\infty).\qed
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 VV is Gaussian and centered, a repeated use of Wick’s formula yields for n1n\geq 1,

𝔼[V(0)nV(x)n]=m=0nm!(nm)2𝔼[Vnm]2𝒞(x)m,\mathbb{E}\left[{V(0)^{n}V(x)^{n}}\right]\,=\,\sum_{m=0}^{n}m!\,\binom{n}{m}^{2}\,\mathbb{E}\big{[}{V^{n-m}}\big{]}^{2}\,\mathcal{C}(x)^{m},

hence, taking Fourier transform,

𝒞^(Vn)=m=0nm!(nm)2𝔼[Vnm]2(𝒞^)m,\widehat{\mathcal{C}}^{(V^{n})}\,=\,\sum_{m=0}^{n}m!\,\binom{n}{m}^{2}\,\mathbb{E}\big{[}{V^{n-m}}\big{]}^{2}\,(\widehat{\mathcal{C}})^{\ast m},

where ()m(\cdot)^{\ast m} denotes the mmth convolution power. As all terms in the sum are nonnegative, the support of 𝒞^(Vn)\widehat{\mathcal{C}}^{(V^{n})} therefore contains the support of (𝒞^)n(\widehat{\mathcal{C}})^{\ast n}, which coincides with the sum m=1nsupp𝒞^\sum_{m=1}^{n}\operatorname{supp}\widehat{\mathcal{C}}. We conclude

S:=n=1m=1nsupp𝒞^adh(𝒞^ϕΩ^supp𝒞^ϕ)=:T.\displaystyle S\,:=\,\bigcup_{n=1}^{\infty}\sum_{m=1}^{n}\operatorname{supp}\widehat{\mathcal{C}}\leavevmode\nobreak\ \subset\leavevmode\nobreak\ {\operatorname{adh\,}}\bigg{(}\bigcup_{\widehat{\mathcal{C}}^{\phi}\in\widehat{\Omega}}\operatorname{supp}\widehat{\mathcal{C}}^{\phi}\bigg{)}\,=:\,T. (4.4)

As 𝒞\mathcal{C} is nonzero and even, we note that SS is an additive subgroup of d\mathbb{R}^{d}, so that its closure must be of the form A+BA+B for some linear subspace AA and some discrete additive subgroup BB. Since 𝒞^\widehat{\mathcal{C}} is supported in SS, if AA is not the whole of d\mathbb{R}^{d}, we would deduce that 𝒞\mathcal{C} is periodic in some direction, which is excluded by assumption. We conclude A=dA=\mathbb{R}^{d}, hence T=dT=\mathbb{R}^{d}. The definition (4.1) of νkϕ\nu_{k}^{\phi} then implies

adh(𝒞^ϕΩ^suppνkϕ)=[0,),\displaystyle{\operatorname{adh\,}}\bigg{(}\bigcup_{\widehat{\mathcal{C}}^{\phi}\in\widehat{\Omega}}\operatorname{supp}\nu_{k}^{\phi}\bigg{)}\,=\,[0,\infty),

and σ(Hk,0st)=[|k|2,)\sigma(H_{k,0}^{\operatorname{st}})=[-|k|^{2},\infty) follows from Lemma 4.2.

Step 2. Proof of (i) under condition (C2).
The exponential decay condition |𝒞(x)|Ce1C|x||\mathcal{C}(x)|\leq Ce^{-\frac{1}{C}|x|} entails that the Fourier transform 𝒞^\widehat{\mathcal{C}} extends holomorphically to the complex strip |z|<1C|\Im z|<\frac{1}{C}, and hence its support coincides with the whole of d\mathbb{R}^{d}. It then follows from (4.1) that the support of νkV\nu_{k}^{V} coincides with the whole interval [0,)[0,\infty), and therefore σ(Hk,0st)=[|k|2,)\sigma(H_{k,0}^{\operatorname{st}})=[-|k|^{2},\infty) by Lemma 4.2.

Step 3. Proof of (ii) under condition (C3).
Recall the definition (2.4) of the set of VV-polynomials,

𝒫(Ω):={j=1najl=1mjV(xlj,):n1,aj,mj0,xljd},\mathcal{P}(\Omega)\,:=\,\bigg{\{}\sum_{j=1}^{n}a_{j}\prod_{l=1}^{m_{j}}V(x_{lj},\cdot)\,:\,n\geq 1,\,a_{j}\in\mathbb{C},\,m_{j}\geq 0,\,x_{lj}\in\mathbb{R}^{d}\bigg{\}},

and let 𝒫0(Ω)\mathcal{P}_{0}(\Omega) denote the subset of elements of 𝒫(Ω)\mathcal{P}(\Omega) with vanishing expectation. For ϕ𝒫0(Ω)\phi\in\mathcal{P}_{0}(\Omega), since VV is Gaussian, Wick’s formula allows to express 𝒞ϕ\mathcal{C}^{\phi} explicitly as a linear combination of products of translated copies of the covariance function 𝒞\mathcal{C}, without constant term. As the Fourier transform 𝒞^\widehat{\mathcal{C}} is assumed absolutely continuous and integrable, we conclude that 𝒞^ϕ\widehat{\mathcal{C}}^{\phi} is absolutely continuous and integrable as well for all ϕ𝒫0(Ω)\phi\in\mathcal{P}_{0}(\Omega). Lemma 4.2 then implies that for ϕ𝒫0(Ω)\phi\in\mathcal{P}_{0}(\Omega) the spectral measure μk,0ϕ,ϕ\mu_{k,0}^{\phi,\phi} is absolutely continuous. In other words, the absolutely continuous subspace for Hk,0stH_{k,0}^{\operatorname{st}} contains 𝒫0(Ω)\mathcal{P}_{0}(\Omega).

It remains to check that 𝒫0(Ω)\mathcal{P}_{0}(\Omega) is dense in L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C}. Given ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega), by σ(V)\sigma(V)-measurability, we may approximate ϕ\phi by a sequence ϕnϕ\phi_{n}\to\phi in L2(Ω)\operatorname{L}^{2}(\Omega) of the form ϕn:=hn(V(x1,),,V(xn,))\phi_{n}:=h_{n}(V(x_{1},\cdot),\ldots,V(x_{n},\cdot)) for some Borel function hnh_{n} on n\mathbb{R}^{n} and some (xj)jd(x_{j})_{j}\subset\mathbb{R}^{d}. Truncating VV and smoothening the Borel functions hnh_{n}’s, we find ϕnϕ\phi^{\prime}_{n}\to\phi in L2(Ω)\operatorname{L}^{2}(\Omega) of the form ϕn:=hn(Vn(x1,),,Vn(xn,))\phi^{\prime}_{n}:=h^{\prime}_{n}(V_{n}(x_{1},\cdot),\ldots,V_{n}(x_{n},\cdot)) for some hnCc(n)h_{n}^{\prime}\in C^{\infty}_{c}(\mathbb{R}^{n}) and Vn:=(Vn)(n)V_{n}:=(V\wedge n)\vee(-n). For each nn, Weierstrass’ approximation theorem then allows to replace the smooth function hnh_{n}^{\prime} by a polynomial pnp_{n} in nn variables. This proves that 𝒫(Ω)\mathcal{P}(\Omega) is dense in L2(Ω)\operatorname{L}^{2}(\Omega), hence 𝒫0(Ω)\mathcal{P}_{0}(\Omega) is dense in L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C}.

Step 4. Proof of (ii) under condition (C4).
Arguing as in Step 3, it is enough to prove that the spectral measure 𝒞^ϕ\widehat{\mathcal{C}}^{\phi} is absolutely continuous for all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega)\ominus\mathbb{C} of the form ϕ:=h(V(x1,),,V(xn,))\phi:=h(V(x_{1},\cdot),\ldots,V(x_{n},\cdot)) with n1n\geq 1, hCc(n)h\in C^{\infty}_{c}(\mathbb{R}^{n}), and (xj)j=1nd(x_{j})_{j=1}^{n}\subset\mathbb{R}^{d}. Let R:=maxj|xj|R:=\max_{j}|x_{j}|. Since ϕ(τx)\phi(\tau_{-x}\cdot) is σ({V(y,)}yBR(x))\sigma(\{V(y,\cdot)\}_{y\in B_{R}(x)})-measurable, the α\alpha-mixing condition for VV yields, cf. [11, Theorem 1.2.3],

|𝒞ϕ(x)|=|Cov[ϕ¯();ϕ(τx)]|8hL(n)2α~((|x|2R)0,R;V).|\mathcal{C}^{\phi}(x)|=\big{|}\operatorname{Cov}\left[{\bar{\phi}(\cdot)};{\phi(\tau_{-x}\cdot)}\right]\big{|}\leq 8\|h\|_{\operatorname{L}^{\infty}(\mathbb{R}^{n})}^{2}\,\tilde{\alpha}\big{(}(|x|-2R)\vee 0,R;V\big{)}.

The assumed integrability of the α\alpha-mixing coefficient then yields 𝒞ϕL1(d)\mathcal{C}^{\phi}\in\operatorname{L}^{1}(\mathbb{R}^{d}), hence the nonnegative Fourier transform 𝒞^ϕ\widehat{\mathcal{C}}^{\phi} is absolutely continuous and belongs to L1L(d)\operatorname{L}^{1}\cap\operatorname{L}^{\infty}(\mathbb{R}^{d}). ∎

4.2. Perturbed fibered operators

We turn to the perturbed fibered operators and show that the spectrum of Hk,λst=Hk,0st+λVH_{k,\lambda}^{\operatorname{st}}=H_{k,0}^{\operatorname{st}}+\lambda V typically coincides with [|k|2+λinfessV,)[-|k|^{2}+\lambda\operatorname{inf\,ess}V,\infty). The precise statement below is however quite intricate and depends on the structure of level sets of VV. This is to be compared with [43, Theorem 5.33] for the almost sure spectrum of Hλ,ωH_{\lambda,\omega} on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}). Combined with Theorem A.1, this implies Proposition 2 in the Gaussian setting.

Proposition 4.4 (Spectrum of Hk,λstH_{k,\lambda}^{\operatorname{st}}).

Let VV be a stationary random field. Define the following two closed subsets of \mathbb{R},

σ1(V)\displaystyle\sigma_{1}(V) :=\displaystyle:= {r:[V[rϵ,r+ϵ]]>0ϵ>0},\displaystyle\big{\{}r\in\mathbb{R}\,:\,\mathbb{P}\left[{V\in[r-\epsilon,r+\epsilon]}\right]>0\leavevmode\nobreak\ \forall\epsilon>0\big{\}},
σ2(V)\displaystyle\sigma_{2}(V) :=\displaystyle:= {r:[V(x,)[rϵ,r+ϵ]xBR]>0ϵ,R>0}.\displaystyle\big{\{}r\in\mathbb{R}\,:\,\mathbb{P}\left[{V(x,\cdot)\in[r-\epsilon,r+\epsilon]\leavevmode\nobreak\ \forall x\in B_{R}}\right]>0\leavevmode\nobreak\ \forall\epsilon,R>0\big{\}}.

Assume that VV satisfies the following weak mixing type condition: for all rσ2(V)r\in\sigma_{2}(V) and ε,R>0\varepsilon,R>0 the level set V(,ω)1([rϵ,r+ϵ])V(\cdot,\omega)^{-1}([r-\epsilon,r+\epsilon]) admits almost surely a bounded connected component containing a ball of radius RR. Then for all kdk\in\mathbb{R}^{d} there holds

σ(Hk,0st)+σ2(V)\displaystyle\sigma(H_{k,0}^{\operatorname{st}})+\sigma_{2}(V) \displaystyle\subset σ(Hk,0st+V)\displaystyle\sigma(H_{k,0}^{\operatorname{st}}+V) (4.5)
\displaystyle\subset (conv(σ(Hk,0st))+σ1(V))(σ(Hk,0st)+conv(σ1(V))).\displaystyle\Big{(}{\operatorname{conv}}\big{(}\sigma(H_{k,0}^{\operatorname{st}})\big{)}+\sigma_{1}(V)\Big{)}\bigcap\Big{(}\sigma(H_{k,0}^{\operatorname{st}})+{\operatorname{conv}}\big{(}\sigma_{1}(V)\big{)}\Big{)}.

In particular, in the Gaussian setting V=b(V0)V=b(V_{0}) with V0V_{0} a nondegenerate stationary Gaussian field and with bC(d)b\in C(\mathbb{R}^{d}), we find σ1(V)=σ2(V)=[infessb,supessb]\sigma_{1}(V)=\sigma_{2}(V)=[\operatorname{inf\,ess}b,\,\operatorname{sup\,ess}b], cf. [43, Theorem 5.34], hence σ(Hk,0st+V)=[|k|2+infessb,)\sigma(H_{k,0}^{\operatorname{st}}+V)=[-|k|^{2}+\operatorname{inf\,ess}b,\,\infty). ∎

Remark 4.5.

The set σ1(V)\sigma_{1}(V) is known as the essential range of VV and coincides with the spectrum of VV as a multiplication operator on L2(Ω)\operatorname{L}^{2}(\Omega). The set σ2(V)\sigma_{2}(V) is a closed subset of σ1(V)\sigma_{1}(V) and can be much smaller: in the periodic case Ω=𝕋d\Omega=\mathbb{T}^{d}, for instance, there holds σ2(V)=\sigma_{2}(V)=\varnothing unless VV 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 σ(Hk,0st+V)σ(Hk,0st)+conv(σ1(V))\sigma(H_{k,0}^{\operatorname{st}}+V)\subset\sigma(H_{k,0}^{\operatorname{st}})+{\operatorname{conv}}(\sigma_{1}(V)), while the other inclusion is similar. If conv(σ1(V))={\operatorname{conv}}(\sigma_{1}(V))=\mathbb{R}, the inclusion is trivial. It remains to consider the cases when conv(σ1(V)){\operatorname{conv}}(\sigma_{1}(V)) has the form [a,)[a,\infty), (,b](-\infty,b], or [a,b][a,b], with a,ba,b\in\mathbb{R}. We focus on the case conv(σ1(V))=[a,b]{\operatorname{conv}}(\sigma_{1}(V))=[a,b], while the other cases are easier. Without loss of generality we can assume a=ba=-b, so that bb coincides with the (finite) operator norm of VV. Let Eσ(Hk,0st)+[b,b]E\notin\sigma(H_{k,0}^{\operatorname{st}})+[-b,b]. Since Eσ(Hk,0st)E\notin\sigma(H_{k,0}^{\operatorname{st}}), we deduce that Hk,0stEH_{k,0}^{\operatorname{st}}-E is invertible and we compute

(Hk,0stE)1VL2(Ω)L2(Ω)<bVL2(Ω)L2(Ω)=1.\|(H_{k,0}^{\operatorname{st}}-E)^{-1}V\|_{\operatorname{L}^{2}(\Omega)\to\operatorname{L}^{2}(\Omega)}<b\|V\|_{\operatorname{L}^{2}(\Omega)\to\operatorname{L}^{2}(\Omega)}=1.

Writing

Hk,0st+VE=(Hk,0stE)(Id+(Hk,0stE)1V),H_{k,0}^{\operatorname{st}}+V-E\,=\,(H_{k,0}^{\operatorname{st}}-E)\big{(}\operatorname{Id}+(H_{k,0}^{\operatorname{st}}-E)^{-1}V\big{)},

and using Neumann series, we conclude that Hk,0st+VEH_{k,0}^{\operatorname{st}}+V-E is invertible, which entails that Eσ(Hk,0st+V)E\notin\sigma(H_{k,0}^{\operatorname{st}}+V).

Step 2. First inclusion in (4.5).

Given rσ2(V)r\in\sigma_{2}(V) and E|k|2E\geq-|k|^{2}, we show that there exists a sequence (ϕn)nL2(Ω)(\phi_{n})_{n}\subset\operatorname{L}^{2}(\Omega) with ϕnL2(Ω)=1\|\phi_{n}\|_{\operatorname{L}^{2}(\Omega)}=1 such that (Hk,0stE)ϕn0(H_{k,0}^{\operatorname{st}}-E)\phi_{n}\to 0 and (Vr)ϕn0(V-r)\phi_{n}\to 0 in L2(Ω)\operatorname{L}^{2}(\Omega), which entails E+rσ(Hk,0st+V)E+r\in\sigma(H_{k,0}^{\operatorname{st}}+V). For ε>0\varepsilon>0, consider the open set Oε(ω):=int(V(,ω)1(rε,r+ε))O_{\varepsilon}(\omega):=\operatorname{int}(V(\cdot,\omega)^{-1}(r-\varepsilon,r+\varepsilon)) and decompose it into its (at most countable) collection of connected components. Denote by (Oεn(ω))n(O_{\varepsilon}^{n}(\omega))_{n} the subcollection of bounded connected components. By assumption, this collection is almost surely nonempty. For all nn, we consider the balls included in Oεn(ω)O_{\varepsilon}^{n}(\omega) with maximal radius. The maximum radius Rεn(ω)R^{n}_{\varepsilon}(\omega) may be attained by different balls and we denote by (xεn,m(ω))m(x_{\varepsilon}^{n,m}(\omega))_{m} the collection of their centers. As this collection is a closed bounded set in d\mathbb{R}^{d}, we may choose xεn(ω)x_{\varepsilon}^{n}(\omega) as the first element in lexicographic order. The set {xεn(ω)}n\{x_{\varepsilon}^{n}(\omega)\}_{n} defines a (nonempty) stationary point process on d\mathbb{R}^{d}. Now choose a smooth cut-off function χ\chi with χ(x)=1\chi(x)=1 for |x|1|x|\leq 1 and χ(x)=0\chi(x)=0 for |x|2|x|\geq 2, and choose ξd\xi\in\mathbb{R}^{d} with |ξ+k|2=E+|k|2|\xi+k|^{2}=E+|k|^{2}. For R>0R>0, we define the random variable

ϕε,R(ω)=neiξxεn(ω)χ(2Rxεn(ω)) 1Rεn(ω)R.\phi_{\varepsilon,R}(\omega)=\sum_{n}e^{i\xi\cdot x_{\varepsilon}^{n}(\omega)}\chi\big{(}\tfrac{2}{R}x_{\varepsilon}^{n}(\omega)\big{)}\,\mathds{1}_{R_{\varepsilon}^{n}(\omega)\geq R}. (4.6)

By assumption, the decimated stationary point process {xε,Rn(ω)}n:={xεn(ω)}n:Rεn(ω)R\{x_{\varepsilon,R}^{n}(\omega)\}_{n}:=\{x_{\varepsilon}^{n}(\omega)\}_{n:R_{\varepsilon}^{n}(\omega)\geq R} is also nonempty and we denote by με,R>0\mu_{\varepsilon,R}>0 its intensity. Since the remaining points in this process are all separated by a distance at least 2R2R, the sum (4.6) defining ϕε,R\phi_{\varepsilon,R} contains at most one non-zero term, and we find

2dμε,R|BR|=[n:xε,RnB12R]ϕε,RL2(Ω)2[n:xε,RnBR]=με,R|BR|.2^{-d}\mu_{\varepsilon,R}|B_{R}|\,=\,\mathbb{P}\big{[}\exists n:x_{\varepsilon,R}^{n}\in B_{\frac{1}{2}R}\big{]}\,\leq\,\|\phi_{\varepsilon,R}\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\leq\,\mathbb{P}\big{[}\exists n:x_{\varepsilon,R}^{n}\in B_{R}\big{]}\,=\,\mu_{\varepsilon,R}|B_{R}|.

Next, we estimate

|(Hk,0stE)ϕε,R|4Rn(|ξ+k||χ(2Rxε,Rn(ω))|+1R|χ(2Rxε,Rn(ω))|),|(H_{k,0}^{\operatorname{st}}-E)\phi_{\varepsilon,R}|\,\leq\,\tfrac{4}{R}\sum_{n}\Big{(}|\xi+k|\big{|}\nabla\chi\big{(}\tfrac{2}{R}x_{\varepsilon,R}^{n}(\omega)\big{)}\big{|}+\tfrac{1}{R}\big{|}\triangle\chi\big{(}\tfrac{2}{R}x_{\varepsilon,R}^{n}(\omega)\big{)}\big{|}\Big{)},

hence,

(Hk,0stE)ϕε,RL2(Ω)2k,ER2[n:xε,RnBRB12R]=R2με,R|BRB12R|R2ϕε,RL2(Ω)2.\|(H_{k,0}^{\operatorname{st}}-E)\phi_{\varepsilon,R}\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\lesssim_{k,E}\,R^{-2}\,\mathbb{P}\big{[}{\exists n:x_{\varepsilon,R}^{n}\in B_{R}\setminus B_{\frac{1}{2}R}}\big{]}\\ \,=\,R^{-2}\mu_{\varepsilon,R}|B_{R}\setminus B_{\frac{1}{2}R}|\,\lesssim\,R^{-2}\,\|\phi_{\varepsilon,R}\|_{\operatorname{L}^{2}(\Omega)}^{2}.

Finally, we compute (Vr)ϕR,εL2(Ω)ϵϕR,εL2(Ω)\|(V-r)\phi_{R,\varepsilon}\|_{\operatorname{L}^{2}(\Omega)}\leq\epsilon\|\phi_{R,\varepsilon}\|_{\operatorname{L}^{2}(\Omega)}, and the conclusion follows. ∎

4.3. Instability of the bound state

While the spectrum of the perturbed fibered operators {Hk,λst}k\{H_{k,\lambda}^{\operatorname{st}}\}_{k} 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 H+λWH+\lambda W of a self-adjoint operator HH on \mathcal{H}, if HH admits a simple eigenvalue at E0E_{0} with normalized eigenvector ψ0\psi_{0}, and if WW satisfies

limε0P¯0(Wψ0),(HE0iε)1P¯0(Wψ0)> 0,\displaystyle\lim_{\varepsilon\downarrow 0}\,\Im\,\big{\langle}\bar{P}_{0}(W\psi_{0}),(H-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0})\big{\rangle}_{\mathcal{H}}\,>\,0, (4.7)

where P¯0\bar{P}_{0} denotes the orthogonal projection onto {ψ0}\{\psi_{0}\}^{\bot}, then the eigenvalue at E0E_{0} is expected to dissolve whenever the perturbation is turned on. The simplest rigorous version of this statement is as follows.

Lemma 4.6.

Let H,WH,W be two self-adjoint operators on a Hilbert space \mathcal{H} and let E0E_{0} be a simple eigenvalue of HH with normalized eigenvector ψ0\psi_{0}. If for some δ>0\delta>0 there exists a branch [0,δ)×:λ(Eλ,ψλ)[0,\delta)\to\mathbb{R}\times\mathcal{H}:\lambda\mapsto(E_{\lambda},\psi_{\lambda}) of class C2C^{2} with

(H+λW)ψλ=Eλψλ,(Eλ,ψλ)|λ=0=(E0,ψ0),(H+\lambda W)\psi_{\lambda}=E_{\lambda}\psi_{\lambda},\qquad(E_{\lambda},\psi_{\lambda})|_{\lambda=0}=(E_{0},\psi_{0}),

then there holds

d2dλ2Eλ|λ=0=2limε0P¯0(Wψ0),(HE0iε)1P¯0(Wψ0),\displaystyle\tfrac{d^{2}}{d\lambda^{2}}E_{\lambda}|_{\lambda=0}\,=\,-2\lim_{\varepsilon\downarrow 0}\big{\langle}\bar{P}_{0}(W\psi_{0}),(H-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0})\big{\rangle}_{\mathcal{H}}, (4.8)

where P¯0u:=uψ0,uψ0\bar{P}_{0}u:=u-\langle\psi_{0},u\rangle_{\mathcal{H}}\psi_{0} is the orthogonal projection onto {ψ0}\{\psi_{0}\}^{\bot}. In particular, if the right-hand side of (4.8) is not real, then there exists no such branch λ(Eλ,ψλ)\lambda\mapsto(E_{\lambda},\psi_{\lambda}). This is in particular the case whenever the spectral measure of HH associated with P¯0(Wψ0)\bar{P}_{0}(W\psi_{0}) is absolutely continuous in a neighborhood of E0E_{0} and has non-vanishing density at E0E_{0}. ∎

Proof.

Assume that there exists a C2C^{2} branch λ(Eλ,ψλ)\lambda\mapsto(E_{\lambda},\psi_{\lambda}) as in the statement and denote by (E0,ψ0)(E_{0}^{\prime},\psi_{0}^{\prime}) and (E0′′,ψ0′′)(E_{0}^{\prime\prime},\psi_{0}^{\prime\prime}) the first and second derivatives with respect to λ\lambda at λ=0\lambda=0. Differentiating the eigenvalue relation yields

(HE0)ψ0+Wψ0=ψ0E0.(H-E_{0})\psi^{\prime}_{0}+W\psi_{0}=\psi_{0}E^{\prime}_{0}.

Taking the scalar product with ψ0\psi_{0}, we find

E0=ψ0,Wψ0,\displaystyle E^{\prime}_{0}=\langle\psi_{0},W\psi_{0}\rangle_{\mathcal{H}}, (4.9)

hence

(HE0)ψ0=P¯0(Wψ0).\displaystyle(H-E_{0})\psi^{\prime}_{0}=-\bar{P}_{0}(W\psi_{0}).

This can be inverted in the form

ψ0=(HE0iε)1P¯0(Wψ0)iε(HE0iε)1ψ0.\displaystyle\psi^{\prime}_{0}=-(H-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0})-i\varepsilon(H-E_{0}-i\varepsilon)^{-1}\psi_{0}^{\prime}. (4.10)

Now differentiating the eigenvalue equation twice, we find

(HE0)ψ0′′+2Wψ0=E0′′ψ0+2E0ψ0,(H-E_{0})\psi^{\prime\prime}_{0}+2W\psi^{\prime}_{0}=E^{\prime\prime}_{0}\psi_{0}+2E^{\prime}_{0}\psi^{\prime}_{0},

hence, injecting (4.9) and taking the scalar product with ψ0\psi_{0},

E0′′=2P¯0(Wψ0),ψ0.E^{\prime\prime}_{0}=2\big{\langle}\bar{P}_{0}(W\psi_{0}),\psi^{\prime}_{0}\big{\rangle}_{\mathcal{H}}.

Injecting (4.10) then yields

E0′′=2P¯0(Wψ0),(HE0iε)1P¯0(Wψ0)2iε(HE0iε)1P¯0(Wψ0),ψ0.E^{\prime\prime}_{0}=-2\big{\langle}\bar{P}_{0}(W\psi_{0}),(H-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0})\big{\rangle}_{\mathcal{H}}-2\big{\langle}i\varepsilon(H-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0}),\psi_{0}^{\prime}\big{\rangle}_{\mathcal{H}}.

Since E0E_{0} is simple, we find 𝟙{E0}(H)u=ψ0,uψ0\mathds{1}_{\{E_{0}\}}(H)u=\langle\psi_{0},u\rangle_{\mathcal{H}}\psi_{0}, hence

limε0iε(HE0iε)1P¯0(Wψ0),ψ0=𝟙{E0}(H)P¯0(Wψ0),ψ0=0,\lim_{\varepsilon\downarrow 0}\big{\langle}i\varepsilon(H-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0}),\psi_{0}^{\prime}\big{\rangle}_{\mathcal{H}}=\big{\langle}\mathds{1}_{\{E_{0}\}}(H)\bar{P}_{0}(W\psi_{0}),\psi_{0}^{\prime}\big{\rangle}_{\mathcal{H}}=0,

and the conclusion follows. ∎

We apply this result to our setting Hk,λst=Hk,0st+λVH_{k,\lambda}^{\operatorname{st}}=H_{k,0}^{\operatorname{st}}+\lambda V with (E0,ψ0)=(0,1)(E_{0},\psi_{0})=(0,1). The quantity in (4.7) takes the form, for kd{0}k\in\mathbb{R}^{d}\setminus\{0\},

αk:=limε0𝔼[V(Hk,0stiε)1V]\displaystyle\alpha_{k}\leavevmode\nobreak\ :=\leavevmode\nobreak\ \lim_{\varepsilon\downarrow 0}\,\Im\,\mathbb{E}\left[{V(H_{k,0}^{\operatorname{st}}-i\varepsilon)^{-1}V}\right] =\displaystyle= limε0dε𝒞^(y)(|y+k|2|k|2)2+ε2¯𝑑y\displaystyle\lim_{\varepsilon\downarrow 0}\int_{\mathbb{R}^{d}}\frac{\varepsilon\,\widehat{\mathcal{C}}(y)}{(|y+k|^{2}-|k|^{2})^{2}+\varepsilon^{2}}\,{{\mathchar 22\relax\mkern-12.0mud}}y
=\displaystyle= (2π)dπ2|k|limε012ε𝒞^(B|k|+ε(k)B|k|ε(k)),\displaystyle(2\pi)^{-d}\frac{\pi}{2|k|}\lim_{\varepsilon\downarrow 0}\frac{1}{2\varepsilon}\widehat{\mathcal{C}}\big{(}B_{|k|+\varepsilon}(-k)\setminus B_{|k|-\varepsilon}(-k)\big{)},

and Proposition 3 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, V=b(V0)V=b(V_{0}) for some bC()b\in C^{\infty}(\mathbb{R}) and some stationary centered Gaussian field V0V_{0} with covariance function 𝒞0L1L(d)\mathcal{C}_{0}\in\operatorname{L}^{1}\cap\operatorname{L}^{\infty}(\mathbb{R}^{d}), and without loss of generality we assume that the probability space (Ω,)(\Omega,\mathbb{P}) is endowed with the σ\sigma-algebra generated by V0V_{0}.

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 L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}); we refer to [37, 1] for details. A self-adjoint operator HH with domain D(H)D(H) on a Hilbert space \mathcal{H} is said to satisfy a Mourre relation on an interval JJ\subset\mathbb{R} with respect to a (self-adjoint) conjugate operator AA with domain D(A)D(A)\subset\mathcal{H} if there exists C01C_{0}\geq 1 and a compact operator KK such that there holds in the sense of forms,

𝟙J(H)[H,iA] 1J(H)1C0𝟙J(H)+K,\mathds{1}_{J}(H)\,[H,iA]\,\mathds{1}_{J}(H)\,\geq\,\tfrac{1}{C_{0}}\mathds{1}_{J}(H)+K, (5.1)

where the commutator [H,iA][H,iA] is defined as a sesquilinear form on D(H)D(A)D(H)\cap D(A). The Mourre relation (5.1) is said to be strict if K=0K=0. For technical reasons, one typically requires that the domain of HH be invariant under the unitary group {eitA}t\{e^{itA}\}_{t\in\mathbb{R}} generated by AA, that is,

eitAD(H)D(H),t,e^{itA}D(H)\subset D(H),\qquad\forall t\in\mathbb{R}, (5.2)

which in particular ensures that D(H)D(A)D(H)\cap D(A) is dense in D(H)D(H), and one further requires [H,iA][H,iA] to be HH-bounded. In that case, the sesquilinear form [H,iA][H,iA] on D(H)D(A)D(H)\cap D(A) automatically extends to the form of a unique HH-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 HH in JJ have finite multiplicity and that HH has no singular continuous spectrum in JJ. In addition, a strict Mourre relation implies that the spectral measure is absolutely continuous on JJ. This is actually a simple consequence of the virial theorem: if λ\lambda was an eigenvalue in JJ with normalized eigenvector ψ\psi, then a strict Mourre inequality would formally yield

0=ψ,[H,iA]ψ1C0ψ2,0=\langle\psi,[H,iA]\psi\rangle_{\mathcal{H}}\geq\tfrac{1}{C_{0}}\|\psi\|^{2}_{\mathcal{H}},

a contradiction. Alternatively, this short formal proof can be rewritten by noting that a strict Mourre relation implies ballistic transport for the flow eitHe^{itH} with respect to the conjugate operator AA: for ϕ𝟙J(H)\phi\in\mathds{1}_{J}(H)\mathcal{H} there holds

teitHϕ,(A)eitHϕ=eitHϕ,[H,iA]eitHϕ1C0ϕ2,\partial_{t}\langle e^{itH}\phi,(-A)e^{itH}\phi\rangle_{\mathcal{H}}=\langle e^{itH}\phi,[H,iA]e^{itH}\phi\rangle_{\mathcal{H}}\geq\tfrac{1}{C_{0}}\|\phi\|_{\mathcal{H}}^{2},

hence eitHϕ,(A)eitHϕ1C0tϕ2+ϕ,(A)ϕ\langle e^{itH}\phi,(-A)e^{itH}\phi\rangle_{\mathcal{H}}\geq\tfrac{1}{C_{0}}t\|\phi\|^{2}_{\mathcal{H}}+\langle\phi,(-A)\phi\rangle_{\mathcal{H}}, thus prohibiting ϕ\phi 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 L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), e.g. [37]. Considering the unitary group of dilations Utg:=etd/2g(et)U_{t}g:=e^{td/2}g(e^{t}\cdot) on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), and noting that Ut()Ut=e2t()U_{-t}(-\triangle)U_{t}=e^{2t}(-\triangle), we deduce by differentiation,

[,iA]=2(),[-\triangle\,,\,iA]=2(-\triangle),

where iAiA denotes the generator of dilations, that is, A:=12i(x+x)A:=\frac{1}{2i}(x\cdot\nabla+\nabla\cdot x) on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}). This implies that -\triangle satisfies a strict Mourre inequality on [ε,)[\varepsilon,\infty) for all ε>0\varepsilon>0 with conjugate operator AA,

𝟙[ε,)()[,iA] 1[ε,)() 2ε 1[ε,)().\mathds{1}_{[\varepsilon,\infty)}(-\triangle)\,[-\triangle\,,\,iA]\,\mathds{1}_{[\varepsilon,\infty)}(-\triangle)\,\geq\,2\varepsilon\,\mathds{1}_{[\varepsilon,\infty)}(-\triangle).

In a semiclassical perspective, the conjugate operator AA corresponds to the escape function (x,p)xp(x,p)\mapsto x\cdot p for the free Hamiltonian H(x,p)=12|p|2H(x,p)=\tfrac{1}{2}|p|^{2}. Next, given a \triangle-bounded potential W:dW:\mathbb{R}^{d}\to\mathbb{R}, we compute

[W,iA]=xW,[W,iA]=-x\cdot\nabla W,

so that the commutator [W,iA][W,iA] is bounded whenever the function xxW(x)x\mapsto x\cdot\nabla W(x) is bounded. For λ\lambda small enough, this easily entails that the Schrödinger operator +λW-\triangle+\lambda W on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}) also satisfies a strict Mourre inequality on [ε,)[\varepsilon,\infty). This follows from the first general property below and illustrates the flexibility of Mourre’s theory.

Lemma 5.1.

Let HH be a self-adjoint operator on a Hilbert space \mathcal{H}, assume that HH satisfies a Mourre relation (5.1) on a bounded interval JJ\subset\mathbb{R} with respect to a conjugate operator AA, that the domain of HH is invariant under {eitA}t\{e^{itA}\}_{t\in\mathbb{R}}, cf. (5.2), and that [H,iA][H,iA] is HH-bounded. Let also WW be symmetric and HH-bounded.

  1. (i)

    Mourre relation under perturbation:
    If the commutator [W,iA][W,iA] is HH-bounded, then for all JJJ^{\prime}\Subset J and λ\lambda small enough the perturbed operator Hλ:=H+λWH_{\lambda}:=H+\lambda W satisfies a Mourre relation on JJ^{\prime} with respect to AA. In addition, if HH satisfies a strict Mourre relation, then HλH_{\lambda} does too.

  2. (ii)

    Strict relation on orthogonal complement of an eigenspace:
    If HH has an eigenvalue E0JE_{0}\in J with eigenprojector P0P_{0}, then there exists a neighborhood JJJ^{\prime}\Subset J of E0E_{0} such that the restriction H¯:=P¯0HP¯0\bar{H}:=\bar{P}_{0}H\bar{P}_{0} of HH to the range of P¯0:=IdP0\bar{P}_{0}:=\operatorname{Id}-P_{0} satisfies a strict Mourre relation on JJ^{\prime} with conjugate operator A¯:=P¯0AP¯0\bar{A}:=\bar{P}_{0}A\bar{P}_{0}. ∎

Proof.

We start with the proof of (i). As the perturbation WW is HH-bounded, the operator HλH_{\lambda} has the same domain as HH for λ\lambda small enough in view of the Kato-Rellich theorem, hence by assumption its domain D(Hλ)=D(H)D(H_{\lambda})=D(H) is invariant under {eitA}t\{e^{itA}\}_{t\in\mathbb{R}}. Furthermore, the commutator [Hλ,iA][H_{\lambda},iA] is HH-bounded, hence HλH_{\lambda}-bounded. Now, choose hCc()h\in C^{\infty}_{c}(\mathbb{R}) with 𝟙Jh𝟙J\mathds{1}_{J^{\prime}}\leq h\leq\mathds{1}_{J}. Multiplying by h(H)h(H) both sides of the Mourre relation for HH yields

h(H)[H,iA]h(H)1C0h(H)+h(H)Kh(H).h(H)[H,iA]h(H)\geq\frac{1}{C_{0}}h(H)+h(H)Kh(H).

As [W,iA][W,iA] is HH-bounded, we deduce

h(H)[Hλ,iA]h(H)1C0h(H)Cλ+h(H)Kh(H).h(H)[H_{\lambda},iA]h(H)\geq\frac{1}{C_{0}}h(H)-C\lambda+h(H)Kh(H).

Noting that the HH-boundedness of WW implies h(H)h(Hλ)λhL()\|h(H)-h(H_{\lambda})\|\lesssim\lambda\|h^{\prime}\|_{\operatorname{L}^{\infty}(\mathbb{R})}, and further using the HH-boundedness of [Hλ,iA][H_{\lambda},iA], we deduce

h(Hλ)[Hλ,iA]h(Hλ)\displaystyle h(H_{\lambda})[H_{\lambda},iA]h(H_{\lambda}) \displaystyle\geq h(H)[Hλ,iA]h(H)Cλ\displaystyle h(H)[H_{\lambda},iA]h(H)-C\lambda
\displaystyle\geq 1C0h(H)2Cλ+h(H)Kh(H)\displaystyle\frac{1}{C_{0}}h(H)-2C\lambda+h(H)Kh(H)
\displaystyle\geq 1C0h(Hλ)2Cλ+h(H)Kh(H).\displaystyle\frac{1}{C_{0}}h(H_{\lambda})-2C\lambda+h(H)Kh(H).

Multiplying both sides by 𝟙J(Hλ)\mathds{1}_{J^{\prime}}(H_{\lambda}), the conclusion (i) follows for 1C02Cλ12C0\frac{1}{C_{0}}-2C\lambda\geq\frac{1}{2C_{0}}.

We turn to the proof of (ii). As P¯0\bar{P}_{0} commutes with HH, multiplying by P¯0\bar{P}_{0} both sides of the Mourre relation for HH yields, on the range of P¯0\bar{P}_{0},

𝟙J(H¯)[H¯,iA¯]𝟙J(H¯)1C0𝟙J(H¯)+K¯,\mathds{1}_{J}(\bar{H})[\bar{H},i\bar{A}]\mathds{1}_{J}(\bar{H})\geq\frac{1}{C_{0}}\mathds{1}_{J}(\bar{H})+\bar{K},

in terms of H¯:=P¯0HP¯0\bar{H}:=\bar{P}_{0}H\bar{P}_{0}, A¯:=P¯0AP¯0\bar{A}:=\bar{P}_{0}A\bar{P}_{0}, K¯:=P¯0KP¯0\bar{K}:=\bar{P}_{0}K\bar{P}_{0}. Multiplying both sides with 𝟙J(H¯)\mathds{1}_{J}(\bar{H}), the compact operator is replaced by 𝟙J(H¯)K¯𝟙J(H¯)\mathds{1}_{J}(\bar{H})\bar{K}\mathds{1}_{J}(\bar{H}). Since 𝟙J(H¯)\mathds{1}_{J}(\bar{H}) converges strongly to 0 on the range of P¯0\bar{P}_{0} as J{E0}J\to\{E_{0}\}, 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 λ\lambda, as well as for a description of the flow eiHλt𝟙J(Hλ)ψ0e^{-iH_{\lambda}t}\mathds{1}_{J^{\prime}}(H_{\lambda})\psi_{0} projected on a whole class of “smooth” states rather than on ψ0\psi_{0} only, but such improvements are not pursued here.

Theorem 5.2 (Dynamical resonances from Mourre’s theory; [8]).

Let HH be a self-adjoint operator on a Hilbert space \mathcal{H}, let WW be symmetric and HH-bounded, and consider the perturbation Hλ:=H+λWH_{\lambda}:=H+\lambda W. Let E0E_{0} be a simple eigenvalue of HH with normalized eigenvector ψ0\psi_{0}, and assume that the following properties hold:

  1.   \bullet

    There is a self-adjoint conjugate operator AA and a neighborhood JJ\subset\mathbb{R} of E0E_{0} such that HH satisfies a Mourre relation on JJ with respect to AA in the sense of (5.1). In addition, the domain of HH is invariant under {eitA}t\{e^{itA}\}_{t\in\mathbb{R}}, cf. (5.2).

  2.   \bullet

    The iterated commutators adAk(H)\operatorname{ad}^{k}_{A}(H) and adAk(W)\operatorname{ad}^{k}_{A}(W) are HH-bounded for 0k60\leq k\leq 6, where iterated commutators are defined by adA0(H)=H\operatorname{ad}_{A}^{0}(H)=H and recursively adAk+1(H)=[adAk(H),iA]\operatorname{ad}_{A}^{k+1}(H)=[\operatorname{ad}_{A}^{k}(H),iA] for k0k\geq 0.

  3.   \bullet

    Fermi’s condition (4.7) is satisfied, that is,

    limε0P¯0(Wψ0),(H¯E0iε)1P¯0(Wψ0)>0,\lim_{\varepsilon\downarrow 0}\Im\big{\langle}\bar{P}_{0}(W\psi_{0}),(\bar{H}-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0})\big{\rangle}_{\mathcal{H}}>0,

    where P¯0\bar{P}_{0} denotes the orthogonal projection onto {ψ0}\{\psi_{0}\}^{\bot} and where we have set for abbreviation H¯:=P¯0HP¯0\bar{H}:=\bar{P}_{0}H\bar{P}_{0}.

Then there exists {zλ}λ>0\{z_{\lambda}\}_{\lambda>0}\subset\mathbb{C} with zλ<0\Im z_{\lambda}<0 such that for all neighborhoods JJJ^{\prime}\Subset J of E0E_{0} there holds for all t0t\geq 0,

|ψ0,eiHλt𝟙J(Hλ)ψ0eizλt|J,Jλ2|logλ|,\Big{|}\big{\langle}\psi_{0},e^{-iH_{\lambda}t}\mathds{1}_{J^{\prime}}(H_{\lambda})\psi_{0}\big{\rangle}_{\mathcal{H}}-e^{-iz_{\lambda}t}\Big{|}\lesssim_{J,J^{\prime}}\lambda^{2}|\!\log\lambda|,

where the dynamical resonance zλz_{\lambda} satisfies

zλ=E0+λψ0,Wψ0λ2limε0P¯0(Wψ0),(H¯E0iε)1P¯0(Wψ0)+o(λ2).z_{\lambda}=E_{0}+\lambda\langle\psi_{0},W\psi_{0}\rangle_{\mathcal{H}}-\lambda^{2}\lim_{\varepsilon\downarrow 0}\big{\langle}\bar{P}_{0}(W\psi_{0}),(\bar{H}-E_{0}-i\varepsilon)^{-1}\bar{P}_{0}(W\psi_{0})\big{\rangle}_{\mathcal{H}}+o(\lambda^{2}).\qed
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 z>0\Im z>0, in terms of the restriction H¯λ:=P¯0HλP¯0\bar{H}_{\lambda}:=\bar{P}_{0}H_{\lambda}\bar{P}_{0},

ψ0,(zHλ)1ψ0=(zE0λψ0,Wψ0λ2P¯0(Wψ0),(zH¯λ)1P¯0(Wψ0))1.\big{\langle}\psi_{0},(z-H_{\lambda})^{-1}\psi_{0}\big{\rangle}_{\mathcal{H}}\\ \,=\,\Big{(}z-E_{0}-\lambda\langle\psi_{0},W\psi_{0}\rangle_{\mathcal{H}}-\lambda^{2}\big{\langle}\bar{P}_{0}(W\psi_{0}),(z-\bar{H}_{\lambda})^{-1}\bar{P}_{0}(W\psi_{0})\big{\rangle}_{\mathcal{H}}\Big{)}^{-1}.

Next, recall that Lemma 5.1 ensures that the restriction H¯λ\bar{H}_{\lambda} on the range of P¯0\bar{P}_{0} satisfies a strict Mourre relation close to E0E_{0}. In view of [22], together with the HH-boundedness of iterated commutators adAk(H)\operatorname{ad}^{k}_{A}(H) and adAk(W)\operatorname{ad}^{k}_{A}(W) for 0k60\leq k\leq 6, this strict Mourre relation implies the C4C^{4}-smoothness of boundary values on JJ of the resolvent

zP¯0(Wψ0),(zH¯λ)1P¯0(Wψ0),z>0,zJ.z\mapsto\big{\langle}\bar{P}_{0}(W\psi_{0}),(z-\bar{H}_{\lambda})^{-1}\bar{P}_{0}(W\psi_{0})\big{\rangle}_{\mathcal{H}},\qquad\Im z>0,\,\Re z\in J.

Inserting a Taylor expansion for the latter in the above Feshbach-Schur complement formula, we construct an approximate meromorphic extension for zψ0,(zHλ)1ψ0z\mapsto\langle\psi_{0},(z-H_{\lambda})^{-1}\psi_{0}\rangle_{\mathcal{H}}. 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 H=HλH=H_{\lambda}^{\circ} and W=WλW=W_{\lambda}^{\circ} further depend on λ\lambda. More precisely, assume for all λ\lambda that E0E_{0} is a simple eigenvalue of HλH_{\lambda}^{\circ} with normalized eigenvector ψ0\psi_{0} (independent of λ\lambda, say), that the restriction of HλH_{\lambda}^{\circ} on the range of P¯0\bar{P}_{0} satisfies a strict Mourre relation on a neighborhood JJ of E0E_{0} with conjugate operator AA and constant C0C_{0} (independent of λ\lambda), that the domain of HλH_{\lambda}^{\circ} is invariant under {eitA}t\{e^{itA}\}_{t\in\mathbb{R}}, and that iterated commutators adAk(Hλ)\operatorname{ad}^{k}_{A}(H_{\lambda}^{\circ}) are HλH_{\lambda}^{\circ}-bounded by C0C_{0} for 0k60\leq k\leq 6. Next, assume that the perturbation WλW_{\lambda} is bounded in the sense of A6P¯0(Wλψ0)C0\|\langle A\rangle^{6}\bar{P}_{0}(W_{\lambda}\psi_{0})\|_{\mathcal{H}}\leq C_{0}, and that iterated commutators adAk(λWλ)\operatorname{ad}_{A}^{k}(\lambda W_{\lambda}) are HλH_{\lambda}^{\circ}-bounded and small enough in the sense that for 0k60\leq k\leq 6 and ϕ\phi\in\mathcal{H},

adAk(λWλ)ϕ1C1(Hλϕ+ϕ),\|\operatorname{ad}_{A}^{k}(\lambda W_{\lambda})\phi\|_{\mathcal{H}}\leq\frac{1}{C_{1}}(\|H_{\lambda}^{\circ}\phi\|_{\mathcal{H}}+\|\phi\|_{\mathcal{H}}),

for some large enough constant C1C_{1} only depending on C0C_{0}. Then the same result holds as in Theorem 5.2 above for the perturbed operator Hλ=Hλ+λWλH_{\lambda}=H_{\lambda}^{\circ}+\lambda W_{\lambda}^{\circ}. ∎

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 V0V_{0} with covariance function 𝒞0L1L(d)\mathcal{C}_{0}\in\operatorname{L}^{1}\cap\operatorname{L}^{\infty}(\mathbb{R}^{d}); we refer to [35, 40, 39] for details. We start by underlining the Hilbert structure associated with the Gaussian field V0V_{0}. The random variables 𝒱0(ζ):=dV0ζ\mathcal{V}_{0}(\zeta):=\int_{\mathbb{R}^{d}}V_{0}\zeta with ζCc(d)\zeta\in C^{\infty}_{c}(\mathbb{R}^{d}) are centered Gaussians with covariance

𝒱0(ζ),𝒱0(ζ)L2(Ω)=d×d𝒞0(xy)ζ(x)¯ζ(y)𝑑x𝑑y.\langle\mathcal{V}_{0}(\zeta^{\prime}),\mathcal{V}_{0}(\zeta)\rangle_{\operatorname{L}^{2}(\Omega)}\,=\,\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\mathcal{C}_{0}(x-y)\,\overline{\zeta^{\prime}(x)}\,\zeta(y)\,dxdy.

We consider the completion of Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}) endowed with the (semi)norm

ζ:=ζ,ζ,ζ,ζ:=d×d𝒞0(xy)ζ(x)¯ζ(y)𝑑x𝑑y,\|\zeta\|_{\mathfrak{H}}:=\langle\zeta,\zeta\rangle_{\mathfrak{H}},\qquad\langle\zeta^{\prime},\zeta\rangle_{\mathfrak{H}}:=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\mathcal{C}_{0}(x-y)\,\overline{\zeta^{\prime}(x)}\,\zeta(y)\,dxdy,

and we denote by \mathfrak{H} the quotient of this completed space with respect to the kernel of \|\cdot\|_{\mathfrak{H}}. The normed space \mathfrak{H} is a separable Hilbert space and the random field V0V_{0} satisfies the isometry relation

𝒱0(ζ),𝒱0(ζ)L2(Ω)=ζ,ζ.\langle\mathcal{V}_{0}(\zeta^{\prime}),\mathcal{V}_{0}(\zeta)\rangle_{\operatorname{L}^{2}(\Omega)}=\langle\zeta^{\prime},\zeta\rangle_{\mathfrak{H}}.

The map 𝒱0:ζ𝒱0(ζ)\mathcal{V}_{0}:\zeta\mapsto\mathcal{V}_{0}(\zeta) then extends as a linear isometric embedding L2(Ω)\mathfrak{H}\to\operatorname{L}^{2}(\Omega) and constitutes a so-called isonormal Gaussian process over \mathfrak{H}. The structure of \mathfrak{H} is conveniently characterized as follows: as 𝒞0L1(d)\mathcal{C}_{0}\in\operatorname{L}^{1}(\mathbb{R}^{d}), the (nonnegative measure) Fourier transform 𝒞^0\widehat{\mathcal{C}}_{0} is absolutely continuous, hence the square root 𝒞^0:=(𝒞^0)1/2\widehat{\mathcal{C}}_{0}^{\circ}:=(\widehat{\mathcal{C}}_{0})^{1/2} belongs to L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}) and the linear map

K:Cc(d)L2(d):ζ𝒞0ζK:C_{c}^{\infty}(\mathbb{R}^{d})\to\operatorname{L}^{2}(\mathbb{R}^{d}):\zeta\mapsto\mathcal{C}_{0}^{\circ}\ast\zeta (5.3)

extends into a unitary transformation K:L2(d)K:\mathfrak{H}\to\operatorname{L}^{2}(\mathbb{R}^{d}). Note that for all xx the Dirac mass δx\delta_{x} is (a representative of) an element of \mathfrak{H} with Kδx=𝒞0(x)K\delta_{x}=\mathcal{C}_{0}^{\circ}(\cdot-x). By definition, the linear isometric embedding L2(d)L2(Ω):u𝒱0(K1u)\operatorname{L}^{2}(\mathbb{R}^{d})\to\operatorname{L}^{2}(\Omega):u\mapsto\mathcal{V}_{0}(K^{-1}u) is a white noise.

As a model dense subspace of L2(Ω)\operatorname{L}^{2}(\Omega), instead of considering the linear subspace 𝒫(Ω)\mathcal{P}(\Omega) of V0V_{0}-polynomials, cf. (2.4), we define the following slightly more convenient subspace,

(Ω):={g(𝒱0(ζ1),,𝒱0(ζn)):n,g:n polynomial,ζ1,,ζnCc(d;)}.\mathcal{R}(\Omega):=\Big{\{}g\big{(}\mathcal{V}_{0}(\zeta_{1}),\ldots,\mathcal{V}_{0}(\zeta_{n})\big{)}\,:\,n\in\mathbb{N},\,g:\mathbb{R}^{n}\to\mathbb{C}\text{ polynomial},\\ \,\zeta_{1},\ldots,\zeta_{n}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R})\Big{\}}. (5.4)

Recall that we implicitly assume that the underlying probability space (Ω,)(\Omega,\mathbb{P}) is endowed with the minimal σ\sigma-algebra generated by V0V_{0}, thus ensuring that (Ω)\mathcal{R}(\Omega) is indeed dense in L2(Ω)\operatorname{L}^{2}(\Omega). This allows to define operators and prove properties on the simpler subspace (Ω)\mathcal{R}(\Omega) in a concrete way before extending them to L2(Ω)\operatorname{L}^{2}(\Omega) by density.

For a random variable ϕ(Ω)\phi\in\mathcal{R}(\Omega), say ϕ=g(𝒱0(ζ1),,𝒱0(ζn))\phi=g(\mathcal{V}_{0}(\zeta_{1}),\ldots,\mathcal{V}_{0}(\zeta_{n})), we define its Malliavin derivative DϕL2(Ω;)D\phi\in\operatorname{L}^{2}(\Omega;\mathfrak{H}) as

Dϕ:=j=1nζj(jg)(𝒱0(ζ1),,𝒱0(ζn)),\displaystyle D\phi\,:=\,\sum_{j=1}^{n}\zeta_{j}\,(\partial_{j}g)\big{(}\mathcal{V}_{0}(\zeta_{1}),\ldots,\mathcal{V}_{0}(\zeta_{n})\big{)},

and similarly, for all p1p\geq 1, its ppth Malliavin derivative DpϕL2(Ω;p)D^{p}\phi\in\operatorname{L}^{2}(\Omega;\mathfrak{H}^{\otimes p}) is given by

Dpϕ:=1j1,,jpn(ζj1ζjp)(j1jppg)(𝒱0(ζ1),,𝒱0(ζn)).\displaystyle D^{p}\phi:=\sum_{1\leq j_{1},\ldots,j_{p}\leq n}\big{(}\zeta_{j_{1}}\otimes\ldots\otimes\zeta_{j_{p}}\big{)}\big{(}\partial_{j_{1}\ldots j_{p}}^{p}g\big{)}\!\big{(}\mathcal{V}_{0}(\zeta_{1}),\ldots,\mathcal{V}_{0}(\zeta_{n})\big{)}.

Note that by definition this belongs to the symmetric tensor product, DpϕL2(Ω;p)D^{p}\phi\in\operatorname{L}^{2}(\Omega;\mathfrak{H}^{\odot p}). These operators on (Ω)\mathcal{R}(\Omega) are closable on L2(Ω)\operatorname{L}^{2}(\Omega). We then set

X𝔻p,2(Ω)2:=𝔼[|X|2]+j=1p𝔼[DjXj2],\|X\|_{\mathbb{D}^{p,2}(\Omega)}^{2}:=\mathbb{E}\left[{|X|^{2}}\right]+\sum_{j=1}^{p}\mathbb{E}\left[{\|D^{j}X\|_{\mathfrak{H}^{\otimes j}}^{2}}\right],

we define the Malliavin-Sobolev space 𝔻p,2(Ω)\mathbb{D}^{p,2}(\Omega) as the closure of (Ω)\mathcal{R}(\Omega) for this norm, and we extend the ppth Malliavin derivative DpD^{p} by density to this space.

Next, we define the corresponding divergence operator DD^{*} as the adjoint of the Malliavin derivative DD, and similarly, for all p1p\geq 1, the ppth-order divergence operator (D)p(D^{*})^{p} as the adjoint of DpD^{p}. In other words, this is defined by the following integration by parts formula, for all ϕdom(D)pL2(Ω;p)\phi^{\prime}\in{\operatorname{dom}}(D^{*})^{p}\subset\operatorname{L}^{2}(\Omega;\mathfrak{H}^{\otimes p}) and ϕ(Ω)\phi\in\mathcal{R}(\Omega),

ϕ,(D)pϕL2(Ω)=𝔼[Dpϕ,ϕp].\langle\phi,(D^{*})^{p}\phi^{\prime}\rangle_{\operatorname{L}^{2}(\Omega)}=\mathbb{E}\left[{\langle D^{p}\phi,\phi^{\prime}\rangle_{\mathfrak{H}^{\otimes p}}}\right].

The so-called Meyer inequalities ensure that the ppth divergence operator (D)p(D^{*})^{p} extends as a bounded operator 𝔻m,2(Ω;p)𝔻mp,2(Ω)\mathbb{D}^{m,2}(\Omega;\mathfrak{H}^{\otimes p})\to\mathbb{D}^{m-p,2}(\Omega) for all m,npm,n\geq p, hence in particular its domain contains 𝔻p,2(Ω;p)\mathbb{D}^{p,2}(\Omega;\mathfrak{H}^{\otimes p}). For ϕ(Ω)\phi\in\mathcal{R}(\Omega) and ζ\zeta\in\mathfrak{H}, a direct computation yields

D(ζϕ)=𝒱0(ζ)ϕζ,Dϕ.D^{*}(\zeta\phi)\,=\,\mathcal{V}_{0}(\zeta)\phi-\langle\zeta,D\phi\rangle_{\mathfrak{H}}.

Due to this relation, with in particular Dζ=𝒱0(ζ)D^{*}\zeta=\mathcal{V}_{0}(\zeta), the divergence operator DD^{*} 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)

:=DD,\mathcal{L}:=D^{*}D, (5.5)

as a self-adjoint operator acting on L2(Ω)\operatorname{L}^{2}(\Omega) with domain 𝔻2,2(Ω)\mathbb{D}^{2,2}(\Omega). The spectrum of \mathcal{L} is σ()=\sigma(\mathcal{L})=\mathbb{N} and its kernel coincides with constants. In particular, the following Poincaré inequality holds: for all ϕ𝔻1,2(Ω)\phi\in\mathbb{D}^{1,2}(\Omega) with 𝔼[ϕ]=0\mathbb{E}\left[{\phi}\right]=0,

ϕL2(Ω)2𝔼[ϕ¯ϕ]=𝔼[Dϕ2].\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\leq\,\mathbb{E}\left[{\bar{\phi}\mathcal{L}\phi}\right]\,=\,\mathbb{E}\left[{\|D\phi\|_{\mathfrak{H}}^{2}}\right].

This ensures the invertibility of the restriction of \mathcal{L} to L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C}, and allows to define a pseudo-inverse 1:=Π1Π\mathcal{L}^{-1}:=\Pi\mathcal{L}^{-1}\Pi on L2(Ω)\operatorname{L}^{2}(\Omega) in terms of the projection Π:=Id𝔼\Pi:=\operatorname{Id}-\mathbb{E}.

We turn to a spectral decomposition of \mathcal{L}. For that purpose, for p0p\geq 0, we first define the ppth multiple integral IpI_{p} as the bounded linear operator pL2(Ω)\mathfrak{H}^{\odot p}\to\operatorname{L}^{2}(\Omega) given by the restriction of the ppth divergence operator, that is, Ip(ζ):=(D)pζI_{p}(\zeta):=(D^{*})^{p}\zeta for all ζp\zeta\in\mathfrak{H}^{\odot p}. Alternatively, IpI_{p} can be characterized as follows: for all ζCc(d;)\zeta\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}) with ζ=1\|\zeta\|_{\mathfrak{H}}=1 there holds

Ip(ζp)=Hp(𝒱0(ζ)),I_{p}(\zeta^{\otimes p})=H_{p}(\mathcal{V}_{0}(\zeta)),

where HpH_{p} denotes the ppth Hermite polynomial, that is, Hp(t):=e12t2(ddt)pe12t2H_{p}(t):=e^{\frac{1}{2}t^{2}}(-\frac{d}{dt})^{p}e^{-\frac{1}{2}t^{2}}. The image of IpI_{p} is known as the ppth Wiener chaos pL2(Ω)\mathcal{H}_{p}\subset\operatorname{L}^{2}(\Omega). Properties of Hermite polynomials easily imply the following orthogonality property: for all p,q0p,q\geq 0 and ζp,ζq\zeta\in\mathfrak{H}^{\odot p},\zeta^{\prime}\in\mathfrak{H}^{\odot q},

Iq(ζ),Ip(ζ)L2(Ω)=δpqp!ζ,ζp,\langle I_{q}(\zeta^{\prime}),I_{p}(\zeta)\rangle_{\operatorname{L}^{2}(\Omega)}\,=\,\delta_{pq}\,p!\,\langle\zeta^{\prime},\zeta\rangle_{\mathfrak{H}^{\otimes p}}, (5.6)

which in particular entails that IpI_{p} is a unitary transformation pp\mathfrak{H}^{\odot p}\to\mathcal{H}_{p}, where the symmetric tensor product p\mathfrak{H}^{\odot p} is endowed with the norm

ζp:=p!ζp.\|\zeta\|_{\mathfrak{H}^{\odot p}}:=\sqrt{p!}\,\|\zeta\|_{\mathfrak{H}^{\otimes p}}.

In view of (5.3), recall that p\mathfrak{H}^{\odot p} is further isometric to L2(d)p=Lsym2((d)p)\operatorname{L}^{2}(\mathbb{R}^{d})^{\odot p}=\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}), endowed with the norm

upLsym2((d)p):=p!upL2((d)p),\|u_{p}\|_{\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p})}:=\sqrt{p!}\,\|u_{p}\|_{\operatorname{L}^{2}((\mathbb{R}^{d})^{p})},

so that we are led to the following unitary transformations,

Lsym2((d)p)ppup(K1)pupIp((K1)pup),\begin{array}[]{ccccc}\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p})&\xrightarrow{\sim}&\mathfrak{H}^{\odot p}&\xrightarrow{\sim}&\mathcal{H}^{p}\\ u_{p}&\mapsto&(K^{-1})^{\otimes p}u_{p}&\mapsto&I_{p}((K^{-1})^{\otimes p}u_{p}),\end{array} (5.7)

and we write for abbreviation

Jp(up):=Ip((K1)pup).J_{p}(u_{p}):=I_{p}((K^{-1})^{\otimes p}u_{p}).

As a consequence of the orthogonality property (5.6), the following Wiener chaos expansion holds in form of a (bosonic) Fock space decomposition,

L2(Ω)=p=0pp=0Lsym2((d)p).\operatorname{L}^{2}(\Omega)\,=\,\bigoplus_{p=0}^{\infty}\mathcal{H}_{p}\,\cong\,\bigoplus_{p=0}^{\infty}\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}). (5.8)

More precisely, for all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega), we can expand

ϕ=p=0Ip(ϕp)=p=0Jp(up),\phi=\sum_{p=0}^{\infty}I_{p}(\phi_{p})=\sum_{p=0}^{\infty}J_{p}(u_{p}),

for some unique collection of kernels ϕpp\phi_{p}\in\mathfrak{H}^{\odot p} or upLsym2((d)p)u_{p}\in\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}), where the expansion is converging in L2(Ω)\operatorname{L}^{2}(\Omega). The Stroock formula asserts

ϕp=1p!𝔼[DpF],provided ϕ𝔻p,2(Ω).\phi_{p}=\frac{1}{p!}\mathbb{E}\left[{D^{p}F}\right],\qquad\text{provided $\phi\in\mathbb{D}^{p,2}(\Omega)$}.

It can be checked that the ppth Wiener chaos p\mathcal{H}_{p} coincides with the eigenspace of the Ornstein–Uhlenbeck operator \mathcal{L} associated with the eigenvalue pp, so that the Wiener chaos expansion (5.8) coincides with the spectral decomposition of \mathcal{L}.

Intuitively, higher chaoses can be viewed as characterizing higher complexity of randomness. In our study of random Schrödinger operators, the use of Wiener chaos decomposition is reminiscent of cumulant expansions for interacting particle systems, e.g. [5, 12].

5.3. A new class of operators on L2(Ω)\operatorname{L}^{2}(\Omega)

This section is devoted to a general construction allowing to transfer operators on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}) into operators on L2(Ω)\operatorname{L}^{2}(\Omega), which will be a key tool in the sequel and is analogous to second quantization in quantum field theory. Given a bounded operator TT on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), for all p0p\geq 0, we denote by Opp(T)\operatorname{Op}^{\circ}_{p}(T) the bounded operator on L2(d)p=Lsym2((d)p)\operatorname{L}^{2}(\mathbb{R}^{d})^{\odot p}=\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}) given by

Opp(T)gp=j=0p1gjTgg(pj1),gL2(d).\operatorname{Op}^{\circ}_{p}(T)\,g^{\otimes p}=\sum_{j=0}^{p-1}g^{\otimes j}\otimes Tg\otimes g^{\otimes(p-j-1)},\qquad g\in\operatorname{L}^{2}(\mathbb{R}^{d}).

Via the isomorphism (5.7), we can then construct a bounded operator Opp(T)\operatorname{Op}_{p}(T) on the ppth Wiener chaos p\mathcal{H}_{p} via

Opp(T)Jp(up):=Jp(Opp(T)up¯¯),upLsym2((d)p),\operatorname{Op}_{p}(T)J_{p}(u_{p}):=J_{p}\big{(}\,\overline{\operatorname{Op}_{p}^{\circ}(T^{*})\,\overline{u_{p}}}\,\big{)},\qquad u_{p}\in\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}),

where TT^{*} is the adjoint of TT on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}). In particular, on the first chaos, this definition formally yields Op1(T)dV0ζ=d(TK1V0)(Kζ)\operatorname{Op}_{1}(T)\int_{\mathbb{R}^{d}}V_{0}\zeta=\int_{\mathbb{R}^{d}}(TK^{-1}V_{0})(K\zeta). Via the Wiener chaos decomposition (5.8), we then let Op(T)\operatorname{Op}(T) denote the densely defined operator on L2(Ω)\operatorname{L}^{2}(\Omega) given by the direct sum

Op(T)=p=0Opp(T).\operatorname{Op}(T)=\bigoplus_{p=0}^{\infty}\operatorname{Op}_{p}(T).

As Op(T)\operatorname{Op}(T) is obviously \mathcal{L}-bounded for bounded TT, the map Op\operatorname{Op} provides a linear embedding (L2(d);L2(d))(𝔻2,2(Ω);L2(Ω))\mathcal{B}(\operatorname{L}^{2}(\mathbb{R}^{d});\operatorname{L}^{2}(\mathbb{R}^{d}))\to\mathcal{B}(\mathbb{D}^{2,2}(\Omega);\operatorname{L}^{2}(\Omega)), but this is however not a group homomorphism as in particular Op(Id)=\operatorname{Op}(\operatorname{Id})=\mathcal{L}.

If TT is bounded and self-adjoint, then Op(T)\operatorname{Op}(T) is self-adjoint on 𝔻2,2(Ω)\mathbb{D}^{2,2}(\Omega). More generally, if TT is unbounded on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}) and essentially self-adjoint on some subset 𝒞\mathcal{C}, then Opp(T)\operatorname{Op}^{\circ}_{p}(T) defines an essentially self-adjoint operator on 𝒞p\mathcal{C}^{\odot p}, hence Opp(T)\operatorname{Op}_{p}(T) is essentially self-adjoint on Jp𝒞pJ_{p}\mathcal{C}^{\odot p}, cf. [47, Theorem VIII.33], and in turn Op(T)\operatorname{Op}(T) defines an essentially self-adjoint operator on

{ϕ=p=0Jp(up):up𝒞pp, and up=0 for p large enough}.\bigg{\{}\phi=\sum_{p=0}^{\infty}J_{p}(u_{p})\,:\,u_{p}\in\mathcal{C}^{\odot p}\leavevmode\nobreak\ \forall p,\text{ and }u_{p}=0\text{ for $p$ large enough}\bigg{\}}.

In particular, noting that the definition (5.4) of (Ω)\mathcal{R}(\Omega) can be reformulated as

(Ω)={ϕ=p=0Jp(up):upCc(d)pp, and up=0 for p large enough},\mathcal{R}(\Omega)\,=\,\bigg{\{}\phi=\sum_{p=0}^{\infty}J_{p}(u_{p})\,:\,u_{p}\in C^{\infty}_{c}(\mathbb{R}^{d})^{\odot p}\leavevmode\nobreak\ \forall p,\text{ and }u_{p}=0\text{ for $p$ large enough}\bigg{\}},

we deduce that if TT is essentially self-adjoint on Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}), then Op(T)\operatorname{Op}(T) is essentially self-adjoint on (Ω)\mathcal{R}(\Omega). Similarly, if TT leaves Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}) invariant, then Op(T)\operatorname{Op}(T) leaves (Ω)\mathcal{R}(\Omega) invariant. Also note that the operators Op(T)\operatorname{Op}(T) and \mathcal{L} strongly commute since \mathcal{L} acts as pIdp\operatorname{Id} on p\mathcal{H}_{p} and since Op(T)\operatorname{Op}(T) preserves the chaos decomposition.

Next, the following shows that the stationary gradient st\nabla^{\operatorname{st}} corresponds to the spatial gradient \nabla via this embedding Op\operatorname{Op}.

Lemma 5.4.

There holds st=Op()\nabla^{\operatorname{st}}=\operatorname{Op}(\nabla) on L2(Ω)\operatorname{L}^{2}(\Omega). In particular, st\nabla^{\operatorname{st}} preserves the chaos decomposition and commutes strongly with \mathcal{L}. ∎

Proof.

Given ζCc(d;)\zeta\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}), we compute

stIp(ζp)=stHp(𝒱0(ζ))=Hp(𝒱0(ζ))st𝒱0(ζ).\nabla^{\operatorname{st}}I_{p}(\zeta^{\otimes p})=\nabla^{\operatorname{st}}H_{p}(\mathcal{V}_{0}(\zeta))=H_{p}^{\prime}(\mathcal{V}_{0}(\zeta))\nabla^{\operatorname{st}}\mathcal{V}_{0}(\zeta).

Noting that st𝒱0(ζ)=𝒱0(ζ)\nabla^{\operatorname{st}}\mathcal{V}_{0}(\zeta)=-\mathcal{V}_{0}(\nabla\zeta) and recalling that Hermite polynomials satisfy Hp=pHp1H_{p}^{\prime}=pH_{p-1}, we deduce

stIp(ζp)=p𝒱0(ζ)Hp1(𝒱0(ζ))=pI1(ζ)Ip1(ζ(p1)).\nabla^{\operatorname{st}}I_{p}(\zeta^{\otimes p})=-p\mathcal{V}_{0}(\nabla\zeta)H_{p-1}(\mathcal{V}_{0}(\zeta))=-pI_{1}(\nabla\zeta)I_{p-1}(\zeta^{\otimes(p-1)}). (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 q1q\geq 1 and ζ0,ζ1Cc(d;)\zeta_{0},\zeta_{1}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}),

I1(ζ1)Iq(ζ0q)=Iq+1(ζ1~ζ0q)+qIq1(ζ1~1ζ0q),I_{1}(\zeta_{1})I_{q}(\zeta_{0}^{\otimes q})=I_{q+1}(\zeta_{1}\widetilde{\otimes}\zeta_{0}^{\otimes q})+qI_{q-1}(\zeta_{1}\widetilde{\otimes}_{1}\zeta_{0}^{\otimes q}), (5.10)

where we have set

ζ1~ζ0q\displaystyle\zeta_{1}\widetilde{\otimes}\zeta_{0}^{\otimes q} :=\displaystyle:= 1q+1j=0qζ0jζ1ζ0(qj),\displaystyle\frac{1}{q+1}\sum_{j=0}^{q}\zeta_{0}^{\otimes j}\otimes\zeta_{1}\otimes\zeta_{0}^{\otimes(q-j)},
ζ1~1ζ0q\displaystyle\zeta_{1}\widetilde{\otimes}_{1}\zeta_{0}^{\otimes q} :=\displaystyle:= ζ1,ζ0ζ0(q1).\displaystyle\langle\zeta_{1},\zeta_{0}\rangle_{\mathfrak{H}}\,\zeta_{0}^{\otimes(q-1)}.

Inserting this formula into (5.9), we find

stIp(ζp)=pIp(ζ~ζ(p1))p(p1)Ip2(ζ~1ζ(p1)).\nabla^{\operatorname{st}}I_{p}(\zeta^{\otimes p})=-pI_{p}(\nabla\zeta\widetilde{\otimes}\zeta^{\otimes(p-1)})-p(p-1)I_{p-2}(\nabla\zeta\widetilde{\otimes}_{1}\zeta^{\otimes(p-1)}).

Since ζ,ζ=Kζ,KζL2(d)=0\langle\zeta,\nabla\zeta\rangle_{\mathfrak{H}}=\langle K\zeta,\nabla K\zeta\rangle_{\operatorname{L}^{2}(\mathbb{R}^{d})}=0 for real-valued ζ\zeta, the second right-hand side term vanishes and we are led to

stIp(ζp)=pIp(ζ~ζ(p1))=Op()Ip(ζp).\nabla^{\operatorname{st}}I_{p}(\zeta^{\otimes p})=-pI_{p}(\nabla\zeta\widetilde{\otimes}\zeta^{\otimes(p-1)})=\operatorname{Op}(\nabla)I_{p}(\zeta^{\otimes p}).

In addition, this formula ensures that st\nabla^{\operatorname{st}} preserves the chaos decomposition. ∎

Given a self-adjoint operator TT on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), the operator Op(T)\operatorname{Op}(T) on L2(Ω)\operatorname{L}^{2}(\Omega) is also self-adjoint and we may consider the corresponding unitary C0C_{0}-groups. If iTiT preserves the real part, then the group {eitOp(T)}t\{e^{it\operatorname{Op}(T)}\}_{t\in\mathbb{R}} on L2(Ω)\operatorname{L}^{2}(\Omega) is shown to admit an explicit description.

Lemma 5.5.

Let TT be essentially self-adjoint on Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}), and assume that the subset of real-valued functions L2(d;)\operatorname{L}^{2}(\mathbb{R}^{d};\mathbb{R}) is invariant under {eitT}t\{e^{itT}\}_{t\in\mathbb{R}}. Then the operator Op(T)\operatorname{Op}(T) generates a unitary C0C_{0}-group {eitOp(T)}t\{e^{it\operatorname{Op}(T)}\}_{t\in\mathbb{R}} on L2(Ω)\operatorname{L}^{2}(\Omega), which has the following explicit action: for all ϕ(Ω)\phi\in\mathcal{R}(\Omega), say ϕ=g(𝒱0(ζ1),,𝒱0(ζn))\phi=g(\mathcal{V}_{0}(\zeta_{1}),\ldots,\mathcal{V}_{0}(\zeta_{n})),

eitOp(T)ϕ=g(𝒱0(K1eitTKζ1),,𝒱0(K1eitTKζn)).e^{it\operatorname{Op}(T)}\phi\,=\,g\big{(}\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{1}),\ldots,\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{n})\big{)}.

In particular, this entails that eitOp(T)e^{it\operatorname{Op}(T)} is multiplicative, that is, for all ϕ,ψ(Ω)\phi,\psi\in\mathcal{R}(\Omega),

eitOp(T)(ϕψ)=(eitOp(T)ϕ)(eitOp(T)ψ),e^{it\operatorname{Op}(T)}(\phi\psi)=(e^{it\operatorname{Op}(T)}\phi)(e^{it\operatorname{Op}(T)}\psi),

which implies that Op(T)\operatorname{Op}(T) is a derivation, that is, for all ϕ,ψ(Ω)\phi,\psi\in\mathcal{R}(\Omega),

Op(T)(ϕψ)=ψOp(T)ϕ+ϕOp(T)ψ.\operatorname{Op}(T)(\phi\psi)=\psi\operatorname{Op}(T)\phi+\phi\operatorname{Op}(T)\psi.\qed
Proof.

Denote by {U~t}t\{\widetilde{U}_{t}\}_{t\in\mathbb{R}} the group of operators defined on (Ω)\mathcal{R}(\Omega) as in the statement: for all ϕ(Ω)\phi\in\mathcal{R}(\Omega), say ϕ=g(𝒱0(ζ1),,𝒱0(ζn))\phi=g(\mathcal{V}_{0}(\zeta_{1}),\ldots,\mathcal{V}_{0}(\zeta_{n})) with n1n\geq 1 and ζ1,,ζnCc(d;)\zeta_{1},\ldots,\zeta_{n}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}),

U~tϕ:=g(𝒱0(K1eitTKζ1),,𝒱0(K1eitTKζn)),\widetilde{U}_{t}\phi\,:=\,g\big{(}\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{1}),\ldots,\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{n})\big{)},

hence in particular, for all ζCc(d;)\zeta\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}),

U~tIp(ζp)=U~tHp(𝒱0(ζ))=Hp(𝒱0(K1eitTKζ))=Ip((K1eitTKζ)p).\widetilde{U}_{t}I_{p}(\zeta^{\otimes p})=\widetilde{U}_{t}H_{p}(\mathcal{V}_{0}(\zeta))=H_{p}\big{(}\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta)\big{)}=I_{p}\big{(}(K^{-1}e^{-itT}K\zeta)^{\otimes p}\big{)}. (5.11)

This is well-defined since eitTe^{-itT} is assumed to preserve L2(d;)\operatorname{L}^{2}(\mathbb{R}^{d};\mathbb{R}). (Note that K1eitTKζK^{-1}e^{-itT}K\zeta may of course no longer have any representative in Cc(d;)C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}) in its equivalence class in \mathfrak{H}.) Noting that

𝒱0(K1eitTKζj),𝒱0(K1eitTKζl)L2(Ω)=K1eitTKζj,K1eitTKζl=eitTKζj,eitTKζlL2(d)=Kζj,KζlL2(d)=ζj,ζl=𝒱0(ζj),𝒱0(ζl)L2(Ω),\big{\langle}\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{j}),\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{l})\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\,=\,\big{\langle}K^{-1}e^{-itT}K\zeta_{j},K^{-1}e^{-itT}K\zeta_{l}\big{\rangle}_{\mathfrak{H}}\\ \,=\,\big{\langle}e^{-itT}K\zeta_{j},e^{-itT}K\zeta_{l}\big{\rangle}_{\operatorname{L}^{2}(\mathbb{R}^{d})}\,=\,\langle K\zeta_{j},K\zeta_{l}\rangle_{\operatorname{L}^{2}(\mathbb{R}^{d})}\,=\,\langle\zeta_{j},\zeta_{l}\rangle_{\mathfrak{H}}\,=\,\langle\mathcal{V}_{0}(\zeta_{j}),\mathcal{V}_{0}(\zeta_{l})\rangle_{\operatorname{L}^{2}(\Omega)},

and using again the assumption that eitTe^{-itT} preserves L2(d;)\operatorname{L}^{2}(\mathbb{R}^{d};\mathbb{R}), we deduce that the (Gaussian) law of (𝒱0(K1eitTKζ1),,𝒱0(K1eitTKζn))\big{(}\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{1}),\ldots,\mathcal{V}_{0}(K^{-1}e^{-itT}K\zeta_{n})\big{)} is invariant with respect to tt, hence for all ϕ(Ω)\phi\in\mathcal{R}(\Omega) and tt\in\mathbb{R},

U~tϕL2(Ω)=ϕL2(Ω).\|\widetilde{U}_{t}\phi\|_{\operatorname{L}^{2}(\Omega)}=\|\phi\|_{\operatorname{L}^{2}(\Omega)}.

This allows to extend {U~t}t\{\widetilde{U}_{t}\}_{t\in\mathbb{R}} by density as a unitary group on L2(Ω)\operatorname{L}^{2}(\Omega). In addition, as {eitT}t\{e^{-itT}\}_{t\in\mathbb{R}} is strongly continuous on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), it is easily deduced that {U~t}t\{\widetilde{U}_{t}\}_{t\in\mathbb{R}} is strongly continuous on L2(Ω)\operatorname{L}^{2}(\Omega). We denote by iT~i\widetilde{T} its skew-adjoint generator on L2(Ω)\operatorname{L}^{2}(\Omega). Differentiating (5.11) with respect to tt shows that the domain of T~\widetilde{T} contains (Ω)\mathcal{R}(\Omega) and that T~=Op(T)\widetilde{T}=\operatorname{Op}(T) on (Ω)\mathcal{R}(\Omega). Since TT is essentially self-adjoint on Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}), Op(T)\operatorname{Op}(T) is essentially self-adjoint on (Ω)\mathcal{R}(\Omega), and we conclude Op(T)=T~\operatorname{Op}(T)=\widetilde{T}, hence U~t=eitOp(T)\widetilde{U}_{t}=e^{it\operatorname{Op}(T)}. ∎

In view of the application to Mourre’s theory for Schrödinger operators, cf. Section 5.1, we recall the definition of the unitary C0C_{0}-group of dilations Utg:=etd/2g(et)U_{t}g:=e^{td/2}g(e^{t}\cdot) on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), and its generator iA:=12(x+x)iA:=\frac{1}{2}\big{(}x\cdot\nabla+\nabla\cdot x\big{)}. We then define the self-adjoint operator

Ast:=Op(A),on L2(Ω),A^{\operatorname{st}}:=\operatorname{Op}(A),\qquad\text{on $\operatorname{L}^{2}(\Omega)$,} (5.12)

and the associated unitary C0C_{0}-group {Utst}t\{U_{t}^{\operatorname{st}}\}_{t\in\mathbb{R}} given by Utst:=eitAstU_{t}^{\operatorname{st}}:=e^{itA^{\operatorname{st}}}. Due to Lemma 5.5, this satisfies, for all ϕ(Ω)\phi\in\mathcal{R}(\Omega), say ϕ=g(𝒱0(ζ1),,𝒱0(ζn))\phi=g(\mathcal{V}_{0}(\zeta_{1}),\ldots,\mathcal{V}_{0}(\zeta_{n})),

Utstϕ=g(𝒱0(K1UtKζ1),,𝒱0(K1UtKζn)),U_{t}^{\operatorname{st}}\phi\,=\,g\big{(}\mathcal{V}_{0}(K^{-1}U_{-t}K\zeta_{1}),\ldots,\mathcal{V}_{0}(K^{-1}U_{-t}K\zeta_{n})\big{)}, (5.13)

which entails in particular that the spaces Hs(Ω)H^{s}(\Omega) are invariant under {Utst}t\{U_{t}^{{\operatorname{st}}}\}_{t\in\mathbb{R}} for all s0s\geq 0.

5.4. Chaos decomposition of fibered operators

While the unperturbed operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k} 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 {Hk,λst}k\{H_{k,\lambda}^{\operatorname{st}}\}_{k} can be viewed as drawing some surprising link between random Schrödinger operators and multi-particle quantum systems.

Lemma 5.6.

Assume for simplicity that V=V0V=V_{0} is itself a Gaussian field. Via the Wiener chaos decomposition (5.8), for all kdk\in\mathbb{R}^{d}, the perturbed fibered operator Hk,λst=Hk,0st+λVH_{k,\lambda}^{\operatorname{st}}=H_{k,0}^{\operatorname{st}}+\lambda V on L2(Ω)\operatorname{L}^{2}(\Omega) is unitarily equivalent to Tk+λ(a+a)T_{k}+\lambda(a+a^{*}) in terms of

Tk:=p=0Tk,p,a:=p=0ap,a:=p=0ap,onp=0Lsym2((d)p),T_{k}:=\bigoplus_{p=0}^{\infty}T_{k,p},\qquad a:=\bigoplus_{p=0}^{\infty}a_{p},\qquad a^{*}:=\bigoplus_{p=0}^{\infty}a^{*}_{p},\qquad\text{on}\leavevmode\nobreak\ \leavevmode\nobreak\ \bigoplus_{p=0}^{\infty}\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}),

in terms of

Tk,p:=Opp(+ikp)Opp(+ikp)|k|2,onLsym2((d)p),T_{k,p}:=-\operatorname{Op}^{\circ}_{p}(\nabla+\tfrac{ik}{p})\cdot\operatorname{Op}^{\circ}_{p}(\nabla+\tfrac{ik}{p})-|k|^{2},\qquad\text{on}\leavevmode\nobreak\ \leavevmode\nobreak\ \operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}),

and in terms of the annihilation and creation operators

ap\displaystyle a_{p} :\displaystyle: Lsym2((d)p+1)Lsym2((d)p),\displaystyle\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p+1})\leavevmode\nobreak\ \to\leavevmode\nobreak\ \operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}),
ap\displaystyle a_{p}^{*} :\displaystyle: Lsym2((d)p)Lsym2((d)p+1),\displaystyle\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p})\leavevmode\nobreak\ \to\leavevmode\nobreak\ \operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p+1}),

which are defined as follows for all up+1Lsym2((d)p+1)u_{p+1}\in\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p+1}) and upLsym2((d)p)u_{p}\in\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}),

(apup+1)(x1,,xp)\displaystyle(a_{p}u_{p+1})(x_{1},\ldots,x_{p}) :=\displaystyle:= j=1p+1d𝒞0(z)up+1(x1,,xj1,z,xj,,xp)𝑑z,\displaystyle\sum_{j=1}^{p+1}\int_{\mathbb{R}^{d}}\mathcal{C}_{0}^{\circ}(z)\,u_{p+1}(x_{1},\ldots,x_{j-1},z,x_{j},\ldots,x_{p})\,dz,
(apup)(x1,,xp+1)\displaystyle(a_{p}^{*}u_{p})(x_{1},\ldots,x_{p+1}) :=\displaystyle:= 1p+1j=1p+1𝒞0(xj)up(x1,,xj1,xj+1,,xp+1).\displaystyle\frac{1}{p+1}\sum_{j=1}^{p+1}\mathcal{C}_{0}^{\circ}(x_{j})\,u_{p}(x_{1},\ldots,x_{j-1},x_{j+1},\ldots,x_{p+1}).

For all pp, the operators apa_{p} and apa_{p}^{*} are bounded and adjoint, with

ap(p+1)12𝔼[|V0|2],ap(p+1)12𝔼[|V0|2],\|a_{p}\|\,\leq\,(p+1)^{\frac{1}{2}}\,\mathbb{E}\big{[}{|V_{0}|^{2}}\big{]},\qquad\|a_{p}^{*}\|\,\leq\,(p+1)^{\frac{1}{2}}\,\mathbb{E}\big{[}{|V_{0}|^{2}}\big{]},

and thus the operators aa and aa^{*} are 12\mathcal{L}^{\frac{1}{2}}-bounded and are adjoint in particular on 𝔻1,2(Ω)\mathbb{D}^{1,2}(\Omega). In addition, they satisfy the commutator relation [a,a]=𝔼[|V0|2][a,a^{*}]=\mathbb{E}\left[{|V_{0}|^{2}}\right] on 𝔻2,2(Ω)\mathbb{D}^{2,2}(\Omega). ∎

Proof.

Given ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega), we consider its Wiener chaos expansion ϕ=p=0Jp(up)\phi=\sum_{p=0}^{\infty}J_{p}(u_{p}), with upLsym2((d)p)u_{p}\in\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}). In the case V=V0V=V_{0}, we can write V=I1(δ0)V=I_{1}(\delta_{0}), hence

Hk,λstϕ=p=0Hk,0Jp(up)+λp=0I1(δ0)Jp(up),H_{k,\lambda}^{\operatorname{st}}\phi\,=\,\sum_{p=0}^{\infty}H_{k,0}J_{p}(u_{p})+\lambda\sum_{p=0}^{\infty}I_{1}(\delta_{0})J_{p}(u_{p}),

and the conclusion follows from Lemma 5.4 and the product formula (5.10). Finally, a direct computation ensures that aa and aa^{*} satisfy the usual properties of annihilation and creation operators,

up,apup+1Lsym2((d)p)=apup,up+1Lsym2((d)p+1),\displaystyle\langle u_{p},a_{p}u_{p+1}\rangle_{\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p})}=\langle a_{p}^{*}u_{p},u_{p+1}\rangle_{\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p+1})},
(apapap1ap1)up=(d(𝒞0)2)up=𝒞0(0)up=𝔼[|V0|2]up.\displaystyle(a_{p}a_{p}^{*}-a_{p-1}^{*}a_{p-1})u_{p}=\Big{(}\int_{\mathbb{R}^{d}}(\mathcal{C}_{0}^{\circ})^{2}\Big{)}u_{p}=\mathcal{C}_{0}(0)\,u_{p}=\mathbb{E}\left[{|V_{0}|^{2}}\right]u_{p}.\qed

5.5. Some Mourre relations on L2(Ω)\operatorname{L}^{2}(\Omega)

Drawing inspiration from the construction of Mourre conjugates for Schrödinger operators on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}), cf. Section 5.1, we show in item (i) below that the generator of dilations AstA^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega) as constructed in Section 5.3 is a conjugate for the stationary Laplacian st-\triangle^{\operatorname{st}}. Nevertheless, item (iii) indicates that the perturbation by the random potential VV 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 st+λV-\triangle^{\operatorname{st}}+\lambda V. 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 1/2\mathcal{L}^{1/2}-bounded on L2(Ω)\operatorname{L}^{2}(\Omega). Finally, in item (iv), we show that the action of the random potential VV as described in Lemma 5.6 on the Fock space allows to associate a natural conjugate. In other words, the stationary Laplacian st-\triangle^{\operatorname{st}} describes diffusion on each chaos and the random potential VV 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 st+λV-\triangle^{\operatorname{st}}+\lambda V 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).

  1. (i)

    Conjugate operator for st-\triangle^{\operatorname{st}}:
    The generator of dilations AstA^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega), cf. (5.12), satisfies

    [st,1iAst]=2(st),[-\triangle^{\operatorname{st}},\tfrac{1}{i}A^{\operatorname{st}}]=2\,(-\triangle^{\operatorname{st}}),

    and the domain H2(Ω)H^{2}(\Omega) of st-\triangle^{\operatorname{st}} is invariant under {Utst=eitAst}t\{U_{t}^{\operatorname{st}}=e^{itA^{\operatorname{st}}}\}_{t\in\mathbb{R}}.

  2. (ii)

    Conjugate operators for 1i1st\frac{1}{i}\nabla^{\operatorname{st}}_{1}:

    [i1st,1iAst]=i1st,[i1st,Op(ix1)]=.[i\nabla_{1}^{\operatorname{st}},\tfrac{1}{i}A^{\operatorname{st}}]=i\nabla_{1}^{\operatorname{st}},\qquad[i\nabla_{1}^{\operatorname{st}},\operatorname{Op}(ix_{1})]=\mathcal{L}.
  3. (iii)

    Incompatibility of the perturbation:
    The commutator [V,iAst][V,iA^{\operatorname{st}}] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega), but only 1/2\mathcal{L}^{1/2}-bounded provided A𝒞0L2(d)A\mathcal{C}_{0}^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}). If V=V0V=V_{0} is itself Gaussian, then similarly [V,Op(ix1)][V,\operatorname{Op}(ix_{1})] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega), but only 1/2\mathcal{L}^{1/2}-bounded provided x1𝒞0L2(d)x_{1}\mathcal{C}_{0}^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}).

  4. (iv)

    Conjugate operator for the perturbation:
    If V=V0V=V_{0} is itself Gaussian, then there holds

    [V,[,V]]=2𝔼[|V|2].[V,[\mathcal{L},V]]=2\,\mathbb{E}\left[{|V|^{2}}\right].\qed
Proof.

Since the operators st-\triangle^{\operatorname{st}}, i1sti\nabla_{1}^{\operatorname{st}}, \mathcal{L}, AstA^{\operatorname{st}}, Op(x1)\operatorname{Op}(x_{1}) are essentially self-adjoint on (Ω)\mathcal{R}(\Omega), and since (Ω)\mathcal{R}(\Omega) is invariant under these operators, the commutators [st,1iAst][-\triangle^{\operatorname{st}},\frac{1}{i}A^{\operatorname{st}}], [i1st,1iAst][i\nabla_{1}^{\operatorname{st}},\frac{1}{i}A^{\operatorname{st}}], [i1st,Op(ix1)][i\nabla_{1}^{\operatorname{st}},\operatorname{Op}(ix_{1})], [V,iAst][V,iA^{\operatorname{st}}], [V,Op(ix1)][V,\operatorname{Op}(ix_{1})] are clearly well-defined on (Ω)\mathcal{R}(\Omega) and are explicitly computed below on that linear subspace. We split the proof into five steps.

Step 1. Proof of (i).
Given p1p\geq 1 and upCc(d)pu_{p}\in C^{\infty}_{c}(\mathbb{R}^{d})^{\odot p}, recalling the notation Utst:=eitAstU_{t}^{\operatorname{st}}:=e^{itA^{\operatorname{st}}}, Lemmas 5.4 and 5.5 lead to

UtststUtstJp(up)=Jp(UtpOpp()Utpup),U_{-t}^{\operatorname{st}}\nabla^{\operatorname{st}}U_{t}^{\operatorname{st}}J_{p}(u_{p})\,=\,J_{p}\big{(}U_{t}^{\otimes p}\operatorname{Op}_{p}^{\circ}(-\nabla)U_{-t}^{\otimes p}u_{p}\big{)},

and hence, by definition of Opp()\operatorname{Op}_{p}^{\circ}(\nabla) and of UtU_{t},

UtststUtstJp(up)=etJp(Opp()up)=etstJp(up),U_{-t}^{\operatorname{st}}\nabla^{\operatorname{st}}U_{t}^{\operatorname{st}}J_{p}(u_{p})\,=\,e^{-t}J_{p}\big{(}\!\operatorname{Op}_{p}^{\circ}(-\nabla)u_{p}\big{)}\,=\,e^{-t}\nabla^{\operatorname{st}}J_{p}(u_{p}),

so that differentiating in tt yields

[st,iAst]Jp(up)=stJp(up),[\nabla^{\operatorname{st}},iA^{\operatorname{st}}]J_{p}(u_{p})=-\nabla^{\operatorname{st}}J_{p}(u_{p}),

and similarly

[st,iAst]Jp(up)=2stJp(up).[\triangle^{\operatorname{st}},iA^{\operatorname{st}}]J_{p}(u_{p})=-2\triangle^{\operatorname{st}}J_{p}(u_{p}).

Step 2. Proof of (ii).
The computation of the commutator [i1st,1iAst][i\nabla_{1}^{\operatorname{st}},\frac{1}{i}A^{\operatorname{st}}] follows from Step 1, and it remains to compute the other one. For upCc(d)pu_{p}\in C^{\infty}_{c}(\mathbb{R}^{d})^{\odot p}, Lemma 5.4 yields

[i1st,Op(ix1)]Jp(up)=Jp([Opp(1),Opp(x1)]up),[i\nabla^{\operatorname{st}}_{1},\operatorname{Op}(ix_{1})]J_{p}(u_{p})=J_{p}\big{(}[\operatorname{Op}_{p}^{\circ}(\nabla_{1}),\operatorname{Op}_{p}^{\circ}(x_{1})]u_{p}\big{)},

while the definition of Opp\operatorname{Op}_{p}^{\circ} leads to [Opp(1),Opp(x1)]=p[\operatorname{Op}_{p}^{\circ}(\nabla_{1}),\operatorname{Op}_{p}^{\circ}(x_{1})]=p, hence

[i1st,Op(ix1)]Jp(up)=pJp(up)=Jp(up).[i\nabla^{\operatorname{st}}_{1},\operatorname{Op}(ix_{1})]J_{p}(u_{p})=pJ_{p}(u_{p})=\mathcal{L}J_{p}(u_{p}).

Step 3. Proof of (iv).
Given V=V0V=V_{0}, for p1p\geq 1 and upLsym2((d)p)u_{p}\in\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}), Lemma 5.6 yields

VJp(up)=Jp+1(apup)+Jp1(ap1up),VJ_{p}(u_{p})=J_{p+1}(a_{p}^{*}u_{p})+J_{p-1}(a_{p-1}u_{p}), (5.14)

so that the commutator with \mathcal{L} takes the form,

[,V]Jp(up)=Jp+1(apup)Jp1(ap1up).[\mathcal{L},V]J_{p}(u_{p})=J_{p+1}(a_{p}^{*}u_{p})-J_{p-1}(a_{p-1}u_{p}).

A direct computation of the commutator of these two operators then yields after straightforward simplifications,

[V,[,V]]Jp(up)=2Jp((apapap1ap1)up),[V,[\mathcal{L},V]]J_{p}(u_{p})=2\,J_{p}\big{(}(a_{p}a_{p}^{*}-a_{p-1}^{*}a_{p-1})u_{p}\big{)},

hence, in view of Lemma 5.6,

[V,[,V]]Jp(up)=2𝔼[|V0|2]Jp(up).[V,[\mathcal{L},V]]J_{p}(u_{p})=2\,\mathbb{E}\left[{|V_{0}|^{2}}\right]\!J_{p}(u_{p}).

Step 4. Proof of (iii) for [V,iAst][V,iA^{\operatorname{st}}].
In view of (5.13), with V=b(V0)=b(J1(Kδ0))V=b(V_{0})=b(J_{1}(K\delta_{0})), we find

UtstVUtstϕ=b(J1(UtKδ0))ϕ,U_{-t}^{\operatorname{st}}VU_{t}^{\operatorname{st}}\phi=b\big{(}J_{1}(U_{t}K\delta_{0})\big{)}\,\phi,

and differentiating in tt yields

[V,iAst]=b(V0)V0,V0:=J1((iA)Kδ0).[V,iA^{\operatorname{st}}]=b^{\prime}(V_{0})V_{0}^{\prime},\qquad V_{0}^{\prime}:=J_{1}\big{(}(iA)K\delta_{0}\big{)}.

Noting that iAiA preserves the real part and that

V0,V0L2(Ω)=J1(Kδ0),J1((iA)Kδ0)L2(Ω)=Kδ0,(iA)Kδ0L2(d)=0,\langle V_{0},V_{0}^{\prime}\rangle_{\operatorname{L}^{2}(\Omega)}=\big{\langle}J_{1}(K\delta_{0}),J_{1}\big{(}(iA)K\delta_{0}\big{)}\big{\rangle}_{\operatorname{L}^{2}(\Omega)}=\big{\langle}K\delta_{0},(iA)K\delta_{0}\big{\rangle}_{\operatorname{L}^{2}(\mathbb{R}^{d})}=0,

we deduce that V0V_{0} and V0V_{0}^{\prime} are independent Gaussian random variables. Further note that V0V_{0}^{\prime} cannot be degenerate: indeed, by definition of AA as generator of dilations, AKδ0AK\delta_{0} can only vanish if 𝒞0(x)=Kδ0(x)|x|d/2\mathcal{C}_{0}(x)=K\delta_{0}(x)\propto|x|^{-d/2}, which is not compatible with 𝒞0(0)=𝔼[|V0|2]<\mathcal{C}_{0}(0)=\mathbb{E}\left[{|V_{0}|^{2}}\right]<\infty. We may then deduce

[V,iAst](V0)p(V0)pL2(Ω)=b(V0)(V0)p+1L2(Ω)(V0)pL2(Ω)=b(V0)L2(Ω)V0L2(Ω)2p+1p,\dfrac{\|[V,iA^{\operatorname{st}}](V^{\prime}_{0})^{p}\|}{\|(V^{\prime}_{0})^{p}\|_{\operatorname{L}^{2}(\Omega)}}=\frac{\|b^{\prime}(V_{0})(V^{\prime}_{0})^{p+1}\|_{\operatorname{L}^{2}(\Omega)}}{\|(V^{\prime}_{0})^{p}\|_{\operatorname{L}^{2}(\Omega)}}=\|b^{\prime}(V_{0})\|_{\operatorname{L}^{2}(\Omega)}\|V_{0}^{\prime}\|_{\operatorname{L}^{2}(\Omega)}\sqrt{2p+1}\xrightarrow{p\uparrow\infty}\infty,

proving that [V,iAst][V,iA^{\operatorname{st}}] is unbounded.

Next, we show that [V,iAst][V,iA^{\operatorname{st}}] is 1/2\mathcal{L}^{1/2}-bounded. For that purpose, for ϕpCc(d)p\phi_{p}\in C^{\infty}_{c}(\mathbb{R}^{d})^{\otimes p}, we use the product formula (5.10) to compute

V0Ip(ϕp)=Ip+1((K1(iA)Kδ0)~ϕp)+pIp1((K1(iA)Kδ0)~1ϕp),V_{0}^{\prime}I_{p}(\phi_{p})=I_{p+1}\big{(}(K^{-1}(iA)K\delta_{0})\widetilde{\otimes}\phi_{p}\big{)}+pI_{p-1}\big{(}(K^{-1}(iA)K\delta_{0})\widetilde{\otimes}_{1}\phi_{p}\big{)},

where the isomorphism (5.7) allows to compute

Ip+1((K1(iA)Kδ0)~ϕp)L2(Ω)\displaystyle\big{\|}I_{p+1}\big{(}(K^{-1}(iA)K\delta_{0})\widetilde{\otimes}\phi_{p}\big{)}\big{\|}_{\operatorname{L}^{2}(\Omega)} =\displaystyle= (p+1)!(K1(iA)Kδ0)~ϕp(p+1)\displaystyle\sqrt{(p+1)!}\,\big{\|}(K^{-1}(iA)K\delta_{0})\widetilde{\otimes}\phi_{p}\big{\|}_{\mathfrak{H}^{\otimes(p+1)}}
\displaystyle\leq (p+1)!K1(iA)Kδ0ϕpp\displaystyle\sqrt{(p+1)!}\,\|K^{-1}(iA)K\delta_{0}\|_{\mathfrak{H}}\|\phi_{p}\|_{\mathfrak{H}^{\otimes p}}
=\displaystyle= p+1K1(iA)Kδ0Ip(ϕp)L2(Ω),\displaystyle\sqrt{p+1}\,\|K^{-1}(iA)K\delta_{0}\|_{\mathfrak{H}}\|I_{p}(\phi_{p})\|_{\operatorname{L}^{2}(\Omega)},

and similarly

pIp1((K1(iA)Kδ0)~1ϕp)L2(Ω)\displaystyle\big{\|}pI_{p-1}\big{(}(K^{-1}(iA)K\delta_{0})\widetilde{\otimes}_{1}\phi_{p}\big{)}\big{\|}_{\operatorname{L}^{2}(\Omega)} =\displaystyle= p(p1)!(K1(iA)Kδ0)~1ϕp(p1)\displaystyle p\sqrt{(p-1)!}\,\big{\|}(K^{-1}(iA)K\delta_{0})\widetilde{\otimes}_{1}\phi_{p}\big{\|}_{\mathfrak{H}^{\otimes(p-1)}}
\displaystyle\leq p(p1)!K1(iA)Kδ0ϕpp\displaystyle p\sqrt{(p-1)!}\,\|K^{-1}(iA)K\delta_{0}\|_{\mathfrak{H}}\|\phi_{p}\|_{\mathfrak{H}^{\otimes p}}
=\displaystyle= pK1(iA)Kδ0Ip(ϕp)L2(Ω).\displaystyle\sqrt{p}\,\|K^{-1}(iA)K\delta_{0}\|_{\mathfrak{H}}\|I_{p}(\phi_{p})\|_{\operatorname{L}^{2}(\Omega)}.

For ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega) in a finite union of chaoses, the Wiener chaos decomposition (5.8) then leads to

[V,iAst]ϕL2(Ω)V0ϕL2(Ω)\displaystyle\|[V,iA^{\operatorname{st}}]\phi\|_{\operatorname{L}^{2}(\Omega)}\,\lesssim\,\|V_{0}^{\prime}\phi\|_{\operatorname{L}^{2}(\Omega)} \displaystyle\leq K1(iA)Kδ0(p=0(2p+1)Ip(ϕp)L2(Ω)2)12\displaystyle\|K^{-1}(iA)K\delta_{0}\|_{\mathfrak{H}}\bigg{(}\sum_{p=0}^{\infty}(2p+1)\|I_{p}(\phi_{p})\|_{\operatorname{L}^{2}(\Omega)}^{2}\bigg{)}^{\frac{1}{2}}
\displaystyle\lesssim A𝒞0L2(d)(+1)12ϕL2(Ω),\displaystyle\|A\mathcal{C}_{0}^{\circ}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}\|(\mathcal{L}+1)^{\frac{1}{2}}\phi\|_{\operatorname{L}^{2}(\Omega)},

and the claim follows.

Step 5. Proof of (iii) for [V,Op(ix1)][V,\operatorname{Op}(ix_{1})].
In view of the Fock space decomposition (5.14) for multiplication by V=V0V=V_{0}, cf. Lemma 5.6, a direct computation yields for all p1p\geq 1 and ϕpCc(d)p\phi_{p}\in C^{\infty}_{c}(\mathbb{R}^{d})^{\otimes p},

[V,Op(ix1)]Ip(ϕp)=Ip+1((K1ix1Kδ0)~ϕp)+pIp1((K1ix1Kδ0)~1ϕp),[V,\operatorname{Op}(ix_{1})]I_{p}(\phi_{p})=-I_{p+1}\big{(}(K^{-1}ix_{1}K\delta_{0})\widetilde{\otimes}\phi_{p}\big{)}+pI_{p-1}\big{(}(K^{-1}ix_{1}K\delta_{0})\widetilde{\otimes}_{1}\phi_{p}\big{)},

and the conclusion follows as in Step 4. ∎

5.6. Mourre relations for fibered operators

This section is devoted to the construction of conjugates for the unperturbed fibered operators {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k}. This appears to be surprisingly more involved than for k=0k=0, as the group of dilations {Utst}t\{U_{t}^{\operatorname{st}}\}_{t\in\mathbb{R}} is no longer adapted and must be suitably deformed. Noting that bounds on iterated commutators are obtained similarly and that the dense subspaces 𝒫(Ω)\mathcal{P}(\Omega) and (Ω)\mathcal{R}(\Omega) are exchangeable, the conclusion of Theorem 4 directly follows upon truncation.

Theorem 5.8 (Mourre relations for fibered operators).

  1. (i)

    Conjugate operator for {Hk,0st}k\{H_{k,0}^{\operatorname{st}}\}_{k}:
    For all kdk\in\mathbb{R}^{d}, there exists a self-adjoint operator CkstC_{k}^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega), essentially self-adjoint on (Ω)\mathcal{R}(\Omega), such that the commutator [Hk,0st,1iCkst][H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega), is Hk,0stH_{k,0}^{\operatorname{st}}-bounded, and satisfies

    [Hk,0st,1iCkst]Π(Hk,0st+34|k|2)Π.\qquad[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\geq\Pi\big{(}H_{k,0}^{\operatorname{st}}+\tfrac{3}{4}|k|^{2}\big{)}\Pi.

    Hence, the fibered operator Hk,0stH_{k,0}^{\operatorname{st}} satisfies a Mourre relation on Jε:=[ε34|k|2,)J_{\varepsilon}:=[\varepsilon-\frac{3}{4}|k|^{2},\infty) with respect to CkstC_{k}^{\operatorname{st}}, for all ε>0\varepsilon>0,

    𝟙Jε(Hk,0st)[Hk,0st,1iCkst] 1Jε(Hk,0st)ε 1Jε(Hk,0st)34|k|2𝔼.\qquad\mathds{1}_{J_{\varepsilon}}(H_{k,0}^{\operatorname{st}})\,[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\,\mathds{1}_{J_{\varepsilon}}(H_{k,0}^{\operatorname{st}})\,\geq\,\varepsilon\,\mathds{1}_{J_{\varepsilon}}(H_{k,0}^{\operatorname{st}})-\tfrac{3}{4}|k|^{2}\mathbb{E}.

    In addition, the domain H2(Ω)H^{2}(\Omega) of Hk,0stH_{k,0}^{\operatorname{st}} is invariant under {eitCkst}t\{e^{itC_{k}^{\operatorname{st}}}\}_{t\in\mathbb{R}}.

  2. (ii)

    Incompatibility of the perturbation:
    If V=V0V=V_{0} is itself Gaussian, then the commutator [V,iCkst][V,iC_{k}^{\operatorname{st}}] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega), but only 1/2\mathcal{L}^{1/2}-bounded provided that A𝒞0,x𝒞0L2(d)A\mathcal{C}_{0}^{\circ},x\mathcal{C}_{0}^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}).

In addition, the full commutator [Hk,λst,1iCkst][H_{k,\lambda}^{\operatorname{st}},\frac{1}{i}C_{k}^{\operatorname{st}}] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega) for all λ\lambda, but it only satisfies the lower bound

[Hk,λst,1iCkst]Hk,λst+34|k|2Cλ1234|k|2𝔼,[H_{k,\lambda}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\geq H_{k,\lambda}^{\operatorname{st}}+\tfrac{3}{4}|k|^{2}-C\lambda\mathcal{L}^{\frac{1}{2}}-\tfrac{3}{4}|k|^{2}\mathbb{E},

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

[Hk,0st,1iAst]=2Hk/2,0st,[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}A^{\operatorname{st}}]=2H_{k/2,0}^{\operatorname{st}},

which shows that the generator of dilations AstA^{\operatorname{st}} should be properly modified. Again drawing inspiration from the situation on the physical space, noting that on L2(d)\operatorname{L}^{2}(\mathbb{R}^{d}) there holds

[(+ik)(+ik),iAk]=2((+ik)(+ik)),\displaystyle[-(\nabla+ik)\cdot(\nabla+ik),iA_{k}]=2\big{(}-(\nabla+ik)\cdot(\nabla+ik)\big{)},

with Ak:=12i(x(+ik)+(+ik)x)A_{k}:=\frac{1}{2i}\big{(}x\cdot(\nabla+ik)+(\nabla+ik)\cdot x\big{)}, we consider

Akst:=Op(Ak)=Ast+Op(kx),A_{k}^{\operatorname{st}}:=\operatorname{Op}(A_{k})=A^{\operatorname{st}}+\operatorname{Op}(k\cdot x), (5.15)

and a similar computation as in Proposition 5.7 (i)–(ii) yields on (Ω)\mathcal{R}(\Omega),

[Hk,0st,1iAkst]=2(st(1+)ikst+|k|2).\displaystyle[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}A_{k}^{\operatorname{st}}]=2\big{(}-\triangle^{\operatorname{st}}-(1+\mathcal{L})ik\cdot\nabla^{\operatorname{st}}+|k|^{2}\mathcal{L}\big{)}.

(Recall that \mathcal{L} and st\nabla^{\operatorname{st}} commute, cf. Lemma 5.4.) In order to counter the apparition of factors \mathcal{L} in this relation and obtain a proper Mourre relation, a further modification of AkstA_{k}^{\operatorname{st}} is thus needed. More precisely, in the definition (5.15) of AkstA_{k}^{\operatorname{st}}, the generator of dilations AstA^{\operatorname{st}} is a suitable conjugate for the stationary Laplacian st-\triangle^{\operatorname{st}}, while Op(kx)\operatorname{Op}(k\cdot x) is supposed to take into account the additional first-order contribution 2ikst-2ik\cdot\nabla^{\operatorname{st}} in the fibered operator Hk,0st:=st2ikstH_{k,0}^{\operatorname{st}}:=-\triangle^{\operatorname{st}}-2ik\cdot\nabla^{\operatorname{st}}. The core of the problem then lies in the factor \mathcal{L} that appears in the commutator

[ikst,iOp(kx)]=|k|2,[ik\cdot\nabla^{\operatorname{st}},i\operatorname{Op}(k\cdot x)]=|k|^{2}\mathcal{L}, (5.16)

which is related to the infinite dimensionality of the probability space. The simplest way to solve this problem would be defining

Bkst:=Ast+1/2Op(kx)1/2,B_{k}^{\operatorname{st}}:=A^{\operatorname{st}}+\mathcal{L}^{-1/2}\operatorname{Op}(k\cdot x)\mathcal{L}^{-1/2}, (5.17)

where we recall that 1:=Π1Π\mathcal{L}^{-1}:=\Pi\mathcal{L}^{-1}\Pi denotes the pseudo-inverse of \mathcal{L}. This indeed leads to the desired Mourre relation,

[Hk,0st,1iBkst]=2(Hk,0st+|k|2)2|k|2𝔼.[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}B_{k}^{\operatorname{st}}]=2(H_{k,0}^{\operatorname{st}}+|k|^{2})-2|k|^{2}\mathbb{E}. (5.18)

However, the perturbation VV behaves particularly badly with respect to this conjugate operator BkstB_{k}^{\operatorname{st}} in the sense that the commutator [V,iBkst][V,iB_{k}^{\operatorname{st}}] is not even bounded when restricted to any fixed Wiener chaos, thus excluding any meaningful use of such a relation. While on the ppth Wiener chaos p\mathcal{H}_{p} the operator Op(kx)\operatorname{Op}(k\cdot x) amounts to the sum j=1pkxj\sum_{j=1}^{p}k\cdot x_{j}, the choice (5.17) consists of rather considering the algebraic mean 1pj=1pkxj\frac{1}{p}\sum_{j=1}^{p}k\cdot x_{j}. Another possible choice to avoid the factor \mathcal{L} in the commutator (5.16) is to use an \ell^{\infty}-norm of {kxj}j=1p\{k\cdot x_{j}\}_{j=1}^{p}. 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 VV is 1/2\mathcal{L}^{1/2}-bounded.

We construct the desired conjugate operator CkstC_{k}^{\operatorname{st}} via its action on the Fock space decomposition (5.8). As we are concerned with the suitable treatment of the first-order operator ikstik\cdot\nabla^{\operatorname{st}}, that is, the stationary derivative in the direction kk, we set z:=kxz:=k\cdot x and first focus on the case of dimension d=1d=1. For all p0p\geq 0, define the function mp:pm_{p}:\mathbb{R}^{p}\to\mathbb{R},

mp(z1,,zp)\displaystyle m_{p}(z_{1},\ldots,z_{p}) :=\displaystyle:= (maxj|zj|)sgnrp(z1,,zp),\displaystyle\textstyle\big{(}\!\max_{j}|z_{j}|\big{)}\operatorname{sgn}r_{p}(z_{1},\ldots,z_{p}),
rp(z1,,zp)\displaystyle r_{p}(z_{1},\ldots,z_{p}) :=\displaystyle:= maxjzj+minjzj,\displaystyle\textstyle\max_{j}z_{j}+\min_{j}z_{j},

with the convention sgn(0)=0\operatorname{sgn}(0)=0. This function is clearly symmetric with respect to the variables z1,,zpz_{1},\dots,z_{p} and has the following main properties.

Lemma 5.9.

For all p0p\geq 0, the function mpm_{p} is well-defined and is continuous on pSp\mathbb{R}^{p}\setminus S_{p}, where SpS_{p} denotes the hypersurface

Sp:=rp1{0}={zp:jk such that zj=zk and |zj|=maxl|zl|}.S_{p}:=r_{p}^{-1}\{0\}=\big{\{}z\in\mathbb{R}^{p}:\exists j\neq k\text{ such that $z_{j}=-z_{k}$ and $|z_{j}|=\textstyle\max_{l}|z_{l}|$}\big{\}}.

In addition, there exists a continuous function gp:Sp[0,)g_{p}:S_{p}\to[0,\infty) such that

[Opp(),mp]= 1+gpδSp 1.[\operatorname{Op}_{p}^{\circ}(\partial),m_{p}]\,=\,1+g_{p}\,\delta_{S_{p}}\,\geq\,1.\qed (5.19)
Proof.

The continuity of mpm_{p} is clear outside the zero locus SpS_{p} of rpr_{p}, and we turn to the second part of the statement. On pSp\mathbb{R}^{p}\setminus S_{p} there holds mp(z1,,zp)=zjm_{p}(z_{1},\ldots,z_{p})=z_{j} with |zj|=maxi|zi||z_{j}|=\max_{i}|z_{i}|, hence j=1pjmp=1\sum_{j=1}^{p}\partial_{j}m_{p}=1. It remains to examine the jump of mpm_{p} on SpS_{p}. We claim that every line directed by the vector (1,,1)(1,\dots,1) in p\mathbb{R}^{p} intersects the hypersurface SpS_{p} at a single point, and this would yield the conclusion. Indeed, given a point z:=(z1,,zp)pz:=(z_{1},\dots,z_{p})\in\mathbb{R}^{p}, say z1=minjzjz_{1}=\min_{j}z_{j} and z2=maxjzjz_{2}=\max_{j}z_{j}, we can write z=z+s(1,,1)z=z^{\prime}+s(1,\ldots,1) with s=12(z1+z2)s=\frac{1}{2}(z_{1}+z_{2}) and z:=(z1s,,zps)Spz^{\prime}:=(z_{1}-s,\dots,z_{p}-s)\in S_{p}, and z+t(1,,1)z^{\prime}+t(1,\ldots,1) belongs to SpS_{p} only if t=st=s. ∎

Next, in order to get a proper Mourre relation on the Fock space, we regularize the functions {mp}p\{m_{p}\}_{p} so as to replace the Dirac part in (5.19) by a positive bump function that is pp-uniformly bounded on p\mathbb{R}^{p}. For that purpose, it is not enough to regularize the sign function in the definition of mpm_{p} in a fixed neighborhood of SpS_{p}, as the derivative would still produce an unbounded term due to the multiplication by maxj|zj|\max_{j}|z_{j}|. A suitable choice of the regularization is rather defined as follows. First rewrite

mp(z1,,zp)=12(maxjzj+minjzj)+12(maxjzjminjzj)sgn(maxzj+minjzj),m_{p}(z_{1},\ldots,z_{p})=\textstyle\frac{1}{2}(\max_{j}z_{j}+\min_{j}z_{j})+\frac{1}{2}(\max_{j}z_{j}-\min_{j}z_{j})\operatorname{sgn}(\max z_{j}+\min_{j}z_{j}),

where only the last sign function needs to be regularized. Choose a smooth odd function χ:[1,1]\chi:\mathbb{R}\to[-1,1] such that χ(s)=1\chi(s)=-1 for s1s\leq-1, χ(s)=1\chi(s)=1 for s1s\geq 1, 0χ20\leq\chi^{\prime}\leq 2 pointwise, χ(s)s\chi(s)\leq s for 1s0-1\leq s\leq 0, and χ(s)s\chi(s)\geq s for 0s10\leq s\leq 1. We then set

m~p(z1,,zp):=12(maxjzj+minjzj)+12(maxjzjminjzj)χ(maxjzj+minjzjmaxjzjminjzj),\widetilde{m}_{p}(z_{1},\ldots,z_{p}):=\textstyle\frac{1}{2}(\max_{j}z_{j}+\min_{j}z_{j})+\frac{1}{2}(\max_{j}z_{j}-\min_{j}z_{j})\,\chi\big{(}\frac{\max_{j}z_{j}+\min_{j}z_{j}}{\max_{j}z_{j}-\min_{j}z_{j}}\big{)},

which is globally well-defined and continuous. Note that

1j=1pjm~p3,|(j=1pj)rm~p|χ,r1,for all r0,\displaystyle 1\leq\sum_{j=1}^{p}\partial_{j}\widetilde{m}_{p}\leq 3,\qquad\bigg{|}\Big{(}\sum_{j=1}^{p}\partial_{j}\Big{)}^{r}\widetilde{m}_{p}\bigg{|}\lesssim_{\chi,r}1,\leavevmode\nobreak\ \leavevmode\nobreak\ \text{for all $r\geq 0$}, (5.20)
|m~p(z1,,zp)||mp(z1,,zp)|=maxj|zj|.\displaystyle|\widetilde{m}_{p}(z_{1},\ldots,z_{p})|\leq|m_{p}(z_{1},\dots,z_{p})|=\textstyle\max_{j}|z_{j}|. (5.21)

We also establish the following important property.

Lemma 5.10.

For all p0p\geq 0, there holds for all z,z1,,zpz,z_{1},\dots,z_{p}\in\mathbb{R},

|m~p+1(z,z1,,zp)m~p(z1,,zp)|2|z|.\big{|}\widetilde{m}_{p+1}(z,z_{1},\dots,z_{p})-\widetilde{m}_{p}(z_{1},\dots,z_{p})\big{|}\leq 2|z|.\qed
Proof.

The conclusion follows from (5.21) if maxj|zj||z|\max_{j}|z_{j}|\leq|z|, so that we can henceforth assume maxj|zj|>|z|\max_{j}|z_{j}|>|z|. By symmetry we can assume z1=minjzjz_{1}=\min_{j}z_{j} and z2=maxjzjz_{2}=\max_{j}z_{j}. In the case z1zz2z_{1}\leq z\leq z_{2}, we find

m~p+1(z,z1,,zp)=m~p(z1,,zp),\widetilde{m}_{p+1}(z,z_{1},\dots,z_{p})=\widetilde{m}_{p}(z_{1},\dots,z_{p}),

and the conclusion follows. It remains to treat the case z1z2zz_{1}\leq z_{2}\leq z, while the symmetric case zz1z2z\leq z_{1}\leq z_{2} is similar. Given z1z2zz_{1}\leq z_{2}\leq z, the assumption maxj|zj|>|z|\max_{j}|z_{j}|>|z| implies z1<|z|z_{1}<-|z|. As z2+z1z2z1z+z1zz1\frac{z_{2}+z_{1}}{z_{2}-z_{1}}\leq\frac{z+z_{1}}{z-z_{1}}, we compute

|m~p+1(z,z1,,zp)m~p(z1,,zp)|\displaystyle\big{|}\widetilde{m}_{p+1}(z,z_{1},\ldots,z_{p})-\widetilde{m}_{p}(z_{1},\ldots,z_{p})\big{|}
\displaystyle\leq 12(zz2)(1+χ(z2+z1z2z1))+12(zz1)(χ(z+z1zz1)χ(z2+z1z2z1))\displaystyle\textstyle\frac{1}{2}(z-z_{2})\Big{(}1+\chi\big{(}\frac{z_{2}+z_{1}}{z_{2}-z_{1}}\big{)}\Big{)}+\frac{1}{2}(z-z_{1})\Big{(}\chi\big{(}\frac{z+z_{1}}{z-z_{1}}\big{)}-\chi\big{(}\frac{z_{2}+z_{1}}{z_{2}-z_{1}}\big{)}\Big{)}
\displaystyle\leq 12(zz2)(1+χ(z2+z1z2z1))+12(zz1)(1+χ(z+z1zz1)).\displaystyle\textstyle\frac{1}{2}(z-z_{2})\Big{(}1+\chi\big{(}\frac{z_{2}+z_{1}}{z_{2}-z_{1}}\big{)}\Big{)}+\frac{1}{2}(z-z_{1})\Big{(}1+\chi\big{(}\frac{z+z_{1}}{z-z_{1}}\big{)}\Big{)}.

Noting that z2+z1z2z1z+z1zz10\frac{z_{2}+z_{1}}{z_{2}-z_{1}}\leq\frac{z+z_{1}}{z-z_{1}}\leq 0 and that there holds χ(y+z1yz1)=1\chi(\frac{y+z_{1}}{y-z_{1}})=-1 whenever y0y\leq 0, the above becomes, in view of the properties of χ\chi,

|m~p+1(z,z1,,zp)m~p(z1,,zp)|\displaystyle\big{|}\widetilde{m}_{p+1}(z,z_{1},\ldots,z_{p})-\widetilde{m}_{p}(z_{1},\ldots,z_{p})\big{|} \displaystyle\leq 12(zz2)𝟙z20+12(zz1)(1+z+z1zz1)𝟙z0\displaystyle\textstyle\frac{1}{2}(z-z_{2})\mathds{1}_{z_{2}\geq 0}+\frac{1}{2}(z-z_{1})\big{(}1+\frac{z+z_{1}}{z-z_{1}}\big{)}\mathds{1}_{z\geq 0}
\displaystyle\leq 32|z|,\displaystyle\tfrac{3}{2}|z|,

as claimed. ∎

We now turn to the construction of the suitable conjugate operator for ikstik\cdot\nabla^{{\operatorname{st}}}. For all p0p\geq 0, we define an operator Mk,pM_{k,p} on Lsym2((d)p)\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}) as the multiplication by the function (x1,,xp)m~p(kx1,,kxp)(x_{1},\ldots,x_{p})\mapsto\widetilde{m}_{p}(k\cdot x_{1},\ldots,k\cdot x_{p}), and we denote by Mk=p=0Mk,pM_{k}=\bigoplus_{p=0}^{\infty}M_{k,p} the corresponding operator on the Fock space. Next, we define the operator Mk,pstM_{k,p}^{\operatorname{st}} on the ppth Wiener chaos p\mathcal{H}_{p} by

Mk,pstJp(up):=Jp(Mk,pup)=Jp(m~p(kx1,,kxp)up),M_{k,p}^{\operatorname{st}}J_{p}(u_{p}):=J_{p}(M_{k,p}u_{p})=J_{p}\big{(}\widetilde{m}_{p}(k\cdot x_{1},\ldots,k\cdot x_{p})\,u_{p}\big{)},

and via the Wiener chaos decomposition (5.8) we set Mkst:=p=0Mk,pstM_{k}^{\operatorname{st}}:=\bigoplus_{p=0}^{\infty}M_{k,p}^{\operatorname{st}} on L2(Ω)\operatorname{L}^{2}(\Omega). We then consider the following operator on L2(Ω)\operatorname{L}^{2}(\Omega),

Ckst:=Ast+12Mkst,C_{k}^{\operatorname{st}}:=A^{\operatorname{st}}+\tfrac{1}{2}M_{k}^{\operatorname{st}}, (5.22)

which is clearly essentially self-adjoint on (Ω)\mathcal{R}(\Omega) given its action on Wiener chaoses; see also Lemma 5.12 below.

Remark 5.11.

The reader may wonder why this definition of CkstC_{k}^{\operatorname{st}} is chosen instead of Ast+MkstA^{{\operatorname{st}}}+M_{k}^{\operatorname{st}}, which would seem more natural in view of (5.15). The computation of the relevant commutators involves [st,Mkst][\nabla^{\operatorname{st}},M_{k}^{\operatorname{st}}], hence the derivative j=1pjm~p\sum_{j=1}^{p}\partial_{j}\widetilde{m}_{p}, which in view of the regularization m~p\widetilde{m}_{p} of mpm_{p} is not uniformly equal to 11 but can vary in the whole interval [1,3][1,3] (or at best in [1,2+δ][1,2+\delta] for some smaller δ>0\delta>0 if the cut-off function χ\chi is chosen with χ\chi^{\prime} closer to 𝟙[1,1]\mathds{1}_{[-1,1]}). Due to this modification, symbols are deformed in the commutator computation, and the choice Ast+MkstA^{\operatorname{st}}+M_{k}^{\operatorname{st}} would fail at providing a Mourre relation close to 0. This is precisely corrected by the above choice (5.22). ∎

We first show that the operator CkstC_{k}^{\operatorname{st}} generates an explicit unitary C0C_{0}-group, which preserves Hs(Ω)H^{s}(\Omega).

Lemma 5.12.

The operator CkstC_{k}^{\operatorname{st}} is essentially self-adjoint on (Ω)\mathcal{R}(\Omega) and its closure generates a unitary C0C_{0}-group {eitCkst}t\{e^{itC_{k}^{\operatorname{st}}}\}_{t\in\mathbb{R}} on L2(Ω)\operatorname{L}^{2}(\Omega), which has the following explicit action on chaoses: for all p1p\geq 1 and upLsym2((d)p)u_{p}\in\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}),

eitCkstJp(up)=Jp(Ut,k,pup),e^{itC_{k}^{\operatorname{st}}}J_{p}(u_{p})\,=\,J_{p}(U_{t,k,p}u_{p}),

where Ut,k,pU_{t,k,p} is defined by

(Ut,k,pup)(x1,,xp):=etdp2exp(i20tm~p(kesx1,,kesxp)𝑑s)up(etx1,,etxp).(U_{t,k,p}u_{p})(x_{1},\ldots,x_{p}):=e^{t\frac{dp}{2}}\exp\Big{(}\frac{i}{2}\int_{0}^{t}\widetilde{m}_{p}(k\cdot e^{s}x_{1},\ldots,k\cdot e^{s}x_{p})\,ds\Big{)}\,u_{p}(e^{t}x_{1},\ldots,e^{t}x_{p}).

In particular, in view of (5.20), for all s0s\geq 0, the subspace Hs(Ω)H^{s}(\Omega) is invariant under this group action {eitCkst}t\{e^{itC_{k}^{\operatorname{st}}}\}_{t\in\mathbb{R}}. ∎

Proof.

In view of the chaos decomposition (5.8), it suffices to check that for all p1p\geq 1 the family {Ut,k,p}t\{U_{t,k,p}\}_{t\in\mathbb{R}} defines a unitary C0C_{0}-group on Lsym2((d)p)\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}) and that its self-adjoint generator is given by Ck,p:=Opp(A)+12Mk,pC_{k,p}:=\operatorname{Op}_{p}(A)+\frac{1}{2}M_{k,p} on Cc(d)pC^{\infty}_{c}(\mathbb{R}^{d})^{\odot p}. First note that the family {Ut,k,p}t\{U_{t,k,p}\}_{t\in\mathbb{R}} clearly defines a unitary group on Lsym2((d)p)\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}). Next, for all ϵ>0\epsilon>0, we decompose

1iε(Uε,k,pupup)(x1,,xp)=1iε(eεdp2up(eεx1,,eεxp)up(x1,,xp))+1iε(exp(i20εm~p(kesx1,,kesxp)𝑑s)1)eεdp2up(eεx1,,eεxp).\frac{1}{i\varepsilon}\big{(}U_{\varepsilon,k,p}u_{p}-u_{p})(x_{1},\ldots,x_{p})\,=\,\frac{1}{i\varepsilon}\Big{(}e^{\varepsilon\frac{dp}{2}}u_{p}(e^{\varepsilon}x_{1},\ldots,e^{\varepsilon}x_{p})-u_{p}(x_{1},\ldots,x_{p})\Big{)}\\ +\frac{1}{i\varepsilon}\bigg{(}\exp\Big{(}\frac{i}{2}\int_{0}^{\varepsilon}\widetilde{m}_{p}(k\cdot e^{s}x_{1},\ldots,k\cdot e^{s}x_{p})\,ds\Big{)}-1\bigg{)}e^{\varepsilon\frac{dp}{2}}\,u_{p}(e^{\varepsilon}x_{1},\ldots,e^{\varepsilon}x_{p}).

As ε0\varepsilon\downarrow 0, for upCc(d)pu_{p}\in C^{\infty}_{c}(\mathbb{R}^{d})^{\odot p}, the first right-hand side term converges to Opp(A)up\operatorname{Op}_{p}^{\circ}(A)u_{p} in Lsym2((d)p)\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}), while the second one converges to 12Mk,pup\frac{1}{2}M_{k,p}u_{p}, and the claim easily follows. ∎

Next, we show that CkstC_{k}^{\operatorname{st}} is a conjugate for the fibered operator Hk,0stH_{k,0}^{\operatorname{st}} away from the bottom of the spectrum. This completes the proof of Theorem 5.8(i). Choosing the cut-off function χ\chi with χ\chi^{\prime} closer to 𝟙[1,1]\mathds{1}_{[-1,1]}, and suitably increasing the factor 12\frac{1}{2} in definition (5.22), the term 34|k|2\frac{3}{4}|k|^{2} in the lower bound (5.23) below could be improved into almost 89|k|2\frac{8}{9}|k|^{2}, but the present construction does not allow to reach a value any closer to |k|2|k|^{2}.

Lemma 5.13.

The commutator [Hk,0st,1iCkst][H_{k,0}^{\operatorname{st}},\frac{1}{i}C_{k}^{\operatorname{st}}] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega), is Hk,0stH_{k,0}^{\operatorname{st}}-bounded, and satisfies the lower bound

[Hk,0st,1iCkst]Π(Hk,0st+34|k|2)Π,[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\,\geq\,\textstyle\Pi\big{(}H_{k,0}^{\operatorname{st}}+\frac{3}{4}|k|^{2}\big{)}\Pi, (5.23)

which entails the following Mourre relation on Jε:=[ε34|k|2,)J_{\varepsilon}:=[\varepsilon-\frac{3}{4}|k|^{2},\infty), for all ε>0\varepsilon>0,

𝟙Jε(Hk,0st)[Hk,0st,1iCkst]𝟙Jε(Hk,0st)ε𝟙Jε(Hk,0st)34|k|2𝔼.\mathds{1}_{J_{\varepsilon}}(H_{k,0}^{\operatorname{st}})[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\mathds{1}_{J_{\varepsilon}}(H_{k,0}^{\operatorname{st}})\,\geq\,\varepsilon\mathds{1}_{J_{\varepsilon}}(H_{k,0}^{\operatorname{st}})-\tfrac{3}{4}|k|^{2}\mathbb{E}.\qed
Proof.

For all p0p\geq 0, we define the operator Mk,pM^{\prime}_{k,p} on Lsym2((d)p)\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}) as the multiplication by the function (x1,,xp)(j=1pjm~p)(kx1,,kxp)(x_{1},\ldots,x_{p})\mapsto(\sum_{j=1}^{p}\partial_{j}\widetilde{m}_{p})(k\cdot x_{1},\ldots,k\cdot x_{p}), we denote by Mk,pstM_{k,p}^{\prime{\operatorname{st}}} the corresponding operator defined on the ppth Wiener chaos p\mathcal{H}_{p} by Mk,pstJp(up)=Jp(Mk,pup)M_{k,p}^{\prime{\operatorname{st}}}J_{p}(u_{p})=J_{p}(M^{\prime}_{k,p}u_{p}), and we set Mkst:=p=0Mk,pstM_{k}^{\prime{\operatorname{st}}}:=\bigoplus_{p=0}^{\infty}M_{k,p}^{\prime{\operatorname{st}}} on L2(Ω)\operatorname{L}^{2}(\Omega). A direct computation on Wiener chaoses yields

[st,Mkst]=kMkst.[\nabla^{\operatorname{st}},M_{k}^{\operatorname{st}}]=-kM_{k}^{\prime{\operatorname{st}}}.

Combining this with Proposition 5.7 (i)–(ii), we easily find on (Ω)\mathcal{R}(\Omega),

[Hk,0st,1iCkst]\displaystyle[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}] =\displaystyle= 2(st)2ikstik12(stMkst+Mkstst)+|k|2Mkst\displaystyle\textstyle 2(-\triangle^{\operatorname{st}})-2ik\cdot\nabla^{\operatorname{st}}-ik\cdot\frac{1}{2}(\nabla^{\operatorname{st}}M_{k}^{\prime{\operatorname{st}}}+M_{k}^{\prime{\operatorname{st}}}\nabla^{\operatorname{st}})+|k|^{2}M_{k}^{\prime{\operatorname{st}}} (5.24)
=\displaystyle= 2(Hk,0st+|k|2)\displaystyle\textstyle 2(H_{k,0}^{\operatorname{st}}+|k|^{2})
ik12((st+ik)(Mkst2)+(Mkst2)(st+ik)),\displaystyle-ik\cdot\tfrac{1}{2}\big{(}(\nabla^{\operatorname{st}}+ik)(M_{k}^{\prime{\operatorname{st}}}-2)+(M_{k}^{\prime{\operatorname{st}}}-2)(\nabla^{\operatorname{st}}+ik)\big{)},

where the right-hand side is well-defined and symmetric on (Ω)\mathcal{R}(\Omega). We split the rest of the proof into two steps.

Step 1. Proof of the lower bound (5.23) on (Ω)\mathcal{R}(\Omega).
Note that constants belong to the kernel of the commutator [Hk,0st,1iCkst][H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]. The above expression (5.24) for the latter yields for all ϕ(Ω)\phi\in\mathcal{R}(\Omega) with 𝔼[ϕ]=0\mathbb{E}\left[{\phi}\right]=0,

ϕ,[Hk,0st,1iCkst]ϕL2(Ω) 2(st+ik)ϕL2(Ω)2|k|(st+ik)ϕL2(Ω)(Mkst2)ϕL2(Ω).\big{\langle}\phi,[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\,\geq\,2\|(\nabla^{\operatorname{st}}+ik)\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}-|k|\|(\nabla^{\operatorname{st}}+ik)\phi\|_{\operatorname{L}^{2}(\Omega)}\|(M_{k}^{\prime{\operatorname{st}}}-2)\phi\|_{\operatorname{L}^{2}(\Omega)}.

The bound (5.20) implies 1Mkst31\leq M_{k}^{\prime{\operatorname{st}}}\leq 3 on L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C}, hence

ϕ,[Hk,0st,1iCkst]ϕL2(Ω) 2(st+ik)ϕL2(Ω)2|k|(st+ik)ϕL2(Ω)ϕL2(Ω),\big{\langle}\phi,[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\,\geq\,2\|(\nabla^{\operatorname{st}}+ik)\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}-|k|\|(\nabla^{\operatorname{st}}+ik)\phi\|_{\operatorname{L}^{2}(\Omega)}\|\phi\|_{\operatorname{L}^{2}(\Omega)},

and we are led to

ϕ,[Hk,0st,1iCkst]ϕL2(Ω)\displaystyle\big{\langle}\phi,[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)} \displaystyle\geq (st+ik)ϕL2(Ω)214|k|2ϕL2(Ω)2\displaystyle\|(\nabla^{\operatorname{st}}+ik)\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}-\tfrac{1}{4}|k|^{2}\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}
=\displaystyle= ϕ,(Hk,0st+34|k|2)ϕL2(Ω),\displaystyle\big{\langle}\phi,(H_{k,0}^{\operatorname{st}}+\tfrac{3}{4}|k|^{2})\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)},

that is, the lower bound (5.23).

Step 2. Proof that the commutator [Hk,0st,1iCkst][H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}] is essentially self-adjoint on (Ω)\mathcal{R}(\Omega) and that its closure is Hk,0stH_{k,0}^{\operatorname{st}}-bounded and self-adjoint on its domain H2(Ω)H^{2}(\Omega).

For all p0p\geq 0, we define the operator Mk,p′′M^{\prime\prime}_{k,p} on Lsym2((d)p)\operatorname{L}^{2}_{\operatorname{sym}}((\mathbb{R}^{d})^{p}) as the multiplication by the function (x1,,xp)(j,l=1pjl2m~p)(kx1,,kxp)(x_{1},\ldots,x_{p})\mapsto(\sum_{j,l=1}^{p}\partial_{jl}^{2}\widetilde{m}_{p})(k\cdot x_{1},\ldots,k\cdot x_{p}), we denote by Mk,p′′stM_{k,p}^{\prime\prime{\operatorname{st}}} the corresponding operator defined on the ppth Wiener chaos p\mathcal{H}_{p} by Mk,p′′stJp(up)=Jp(Mk,p′′up)M_{k,p}^{\prime\prime{\operatorname{st}}}J_{p}(u_{p})=J_{p}(M^{\prime\prime}_{k,p}u_{p}), and we set Mk′′st:=p=0Mk,p′′stM_{k}^{\prime\prime{\operatorname{st}}}:=\bigoplus_{p=0}^{\infty}M_{k,p}^{\prime\prime{\operatorname{st}}} on L2(Ω)\operatorname{L}^{2}(\Omega). A direct computation on Wiener chaoses yields

[st,Mkst]=kMk′′st,[\nabla^{\operatorname{st}},M_{k}^{\prime{\operatorname{st}}}]=-kM_{k}^{\prime\prime{\operatorname{st}}},

hence the expression (5.24) can be rewritten as

[Hk,0st,1iCkst]= 2(Hk,0st+|k|2)ik(Mkst2)(st+ik)+i2|k|2Mk′′st.[H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]\,=\,\textstyle 2(H_{k,0}^{\operatorname{st}}+|k|^{2})-ik\cdot(M_{k}^{\prime{\operatorname{st}}}-2)(\nabla^{{\operatorname{st}}}+ik)+\frac{i}{2}|k|^{2}M_{k}^{\prime\prime{\operatorname{st}}}. (5.25)

Note that the bound (5.20) ensures that Mkst2M_{k}^{\prime{\operatorname{st}}}-2 is bounded by 22 and that Mk′′stM_{k}^{\prime\prime{\operatorname{st}}} is bounded on L2(Ω)\operatorname{L}^{2}(\Omega), hence

(ik(Mkst2)(st+ik)+i2|k|2Mk′′st)ϕL2(Ω)\displaystyle\big{\|}\big{(}-ik\cdot(M_{k}^{\prime{\operatorname{st}}}-2)(\nabla^{{\operatorname{st}}}+ik)+\tfrac{i}{2}|k|^{2}M_{k}^{\prime\prime{\operatorname{st}}}\big{)}\phi\big{\|}_{\operatorname{L}^{2}(\Omega)}
\displaystyle\lesssim |k|(st+ik)ϕL2(Ω)+|k|2ϕL2(Ω)\displaystyle|k|\|(\nabla^{\operatorname{st}}+ik)\phi\|_{\operatorname{L}^{2}(\Omega)}+|k|^{2}\|\phi\|_{\operatorname{L}^{2}(\Omega)}
=\displaystyle= |k|(Hk,0st+|k|2)12ϕL2(Ω)+|k|2ϕL2(Ω).\displaystyle|k|\|(H_{k,0}^{\operatorname{st}}+|k|^{2})^{\frac{1}{2}}\phi\|_{\operatorname{L}^{2}(\Omega)}+|k|^{2}\|\phi\|_{\operatorname{L}^{2}(\Omega)}.

Together with (5.25), this shows that the commutator [Hk,0st,1iCkst][H_{k,0}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}] is an infinitesimal perturbation of 2Hk,0st2H_{k,0}^{\operatorname{st}}, 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 VV with respect to the above-constructed conjugate operator CkstC_{k}^{\operatorname{st}}. In view of Proposition 5.7(iii), it remains to establish the following.

Lemma 5.14.

If V=V0V=V_{0} is itself Gaussian, the commutator [V,iMkst][V,iM_{k}^{{\operatorname{st}}}] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega), but is only 1/2\mathcal{L}^{1/2}-bounded provided that x𝒞0L2(d)x\mathcal{C}_{0}^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}). ∎

Proof.

In view of the Fock space decomposition (5.14) for multiplication by V=V0V=V_{0}, cf. Lemma 5.6, we find

[V,Mkst]Jp(up)=Jp+1(1p+1j=1p+1𝒞0(xj)up(x1,,xj1,xj+1,,xp+1)×(m~p(kx1,,kxj1,kxj+1,,kxp+1)m~p+1(kx1,,kxp+1)))+pJp1(d𝒞0(z)up(x1,,xp1,z)×(m~p(kx1,,kxp1,kz)m~p1(kx1,,kxp1))dz).[V,M_{k}^{\operatorname{st}}]J_{p}(u_{p})\,=\,J_{p+1}\bigg{(}\frac{1}{p+1}\sum_{j=1}^{p+1}\mathcal{C}_{0}^{\circ}(x_{j})\,u_{p}(x_{1},\ldots,x_{j-1},x_{j+1},\ldots,x_{p+1})\\ \times\big{(}\widetilde{m}_{p}(k\cdot x_{1},\ldots,k\cdot x_{j-1},k\cdot x_{j+1},\ldots,k\cdot x_{p+1})-\widetilde{m}_{p+1}(k\cdot x_{1},\ldots,k\cdot x_{p+1})\big{)}\bigg{)}\\ +pJ_{p-1}\bigg{(}\int_{\mathbb{R}^{d}}\mathcal{C}_{0}^{\circ}(z)\,u_{p}(x_{1},\ldots,x_{p-1},z)\\ \times\big{(}\widetilde{m}_{p}(k\cdot x_{1},\ldots,k\cdot x_{p-1},k\cdot z)-\widetilde{m}_{p-1}(k\cdot x_{1},\ldots,k\cdot x_{p-1})\big{)}dz\bigg{)}.

In view of Lemma 5.10 and of the Wiener chaos decomposition (5.8), arguing similarly as in the proof of Proposition 5.7(iii), we easily deduce for all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega),

[V,Mkst]ϕL2(Ω)|k|(d|x|2|𝒞0(x)|2𝑑x)12(+1)12ϕL2(Ω),\|[V,M_{k}^{\operatorname{st}}]\phi\|_{\operatorname{L}^{2}(\Omega)}\,\lesssim\,|k|\Big{(}\int_{\mathbb{R}^{d}}|x|^{2}|\mathcal{C}_{0}^{\circ}(x)|^{2}dx\Big{)}^{\frac{1}{2}}\|(\mathcal{L}+1)^{\frac{1}{2}}\phi\|_{\operatorname{L}^{2}(\Omega)},

and the conclusion follows. ∎

Finally, we argue that, while well-defined and symmetric on (Ω)\mathcal{R}(\Omega), the full commutator [Hk,λst,1iCkst][H_{k,\lambda}^{\operatorname{st}},\frac{1}{i}C_{k}^{\operatorname{st}}] is essentially self-adjoint.

Lemma 5.15.

If V=V0V=V_{0} is itself Gaussian, the commutator [Hk,λst,1iCkst][H_{k,\lambda}^{\operatorname{st}},\frac{1}{i}C_{k}^{\operatorname{st}}] is well-defined and essentially self-adjoint on (Ω)\mathcal{R}(\Omega). ∎

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 Hk,λstH_{k,\lambda}^{\operatorname{st}} is essentially self-adjoint on (Ω)\mathcal{R}(\Omega) when V=V0V=V_{0} is Gaussian; a more general criterion for essential self-adjointness of Hk,λstH_{k,\lambda}^{\operatorname{st}} in case of an unbounded potential without particular Gaussian structure is included in Appendix A.

Proposition 5.16.

Let H1H_{1} and LL be self-adjoint operators on their respective domains D(H1)D(H_{1}) and D(L)D(L) on a Hilbert space \mathcal{H}, and let H2H_{2} be a symmetric operator defined on some dense linear subspace 𝒟\mathcal{D}\subset\mathcal{H}, such that

  1.   \bullet

    H1H_{1} and LL are nonnegative and commute strongly, hence H1+LH_{1}+L is self-adjoint on D(H1)D(L)D(H_{1})\cap D(L);

  2.   \bullet

    2H22H_{2} is a Kato perturbation of H1+LH_{1}+L, that is, 𝒟\mathcal{D} is a core for H1+LH_{1}+L and there is α<1\alpha<1 and C1C\geq 1 such that there holds, for all u𝒟u\in\mathcal{D},

    2H2uα(H1+L)u+Cu;2\|H_{2}u\|_{\mathcal{H}}\leq\alpha\|(H_{1}+L)u\|_{\mathcal{H}}+C\|u\|_{\mathcal{H}};
  3.   \bullet

    ±i[H2,L]H1+L+1\pm i[H_{2},L]\lesssim H_{1}+L+1, that is, for all u𝒟u\in\mathcal{D},

    |H2u,LuLu,H2u|u,(H1+L+1)u.\qquad\big{|}\langle H_{2}u,Lu\rangle_{\mathcal{H}}-\langle Lu,H_{2}u\rangle_{\mathcal{H}}\big{|}\,\lesssim\,\langle u,(H_{1}+L+1)u\rangle_{\mathcal{H}}.

Then the operator H:=H1+H2H:=H_{1}+H_{2} is essentially self-adjoint on 𝒟\mathcal{D}. ∎

Proof.

We split the proof into two steps.

Step 1. Proof that for κ\kappa large enough the operator Hκ:=H+L+κH^{\kappa}:=H+L+\kappa is essentially self-adjoint on 𝒟\mathcal{D} and satisfies Hκ1H^{\kappa}\geq 1.

As H1H_{1} and LL are self-adjoint on their respective domains and nonnegative and as they commute strongly, their sum H1κ:=H1+L+κH_{1}^{\kappa}:=H_{1}+L+\kappa is self-adjoint on D(H1)D(L)D(H_{1})\cap D(L), cf. [46, Lemma 4.15.1]. Next, as H2H_{2} is a Kato perturbation of H1+LH_{1}+L, it is also a Kato perturbation of H1κH_{1}^{\kappa} for all κ0\kappa\geq 0, hence the Kato-Rellich theorem entails that Hκ:=H1κ+H2H^{\kappa}:=H_{1}^{\kappa}+H_{2} is essentially self-adjoint on 𝒟\mathcal{D}. In addition, as H2H_{2} is a Kato perturbation of H1+LH_{1}+L and as H1+LH_{1}+L is nonnegative, it is easily deduced that H1+H2+LH_{1}+H_{2}+L is bounded from below, hence Hκ1H^{\kappa}\geq 1 for κ\kappa 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 κ\kappa large enough,

  1. (i)

    HH is HκH^{\kappa}-bounded on 𝒟\mathcal{D};

  2. (ii)

    ±i[H,Hκ]κHκ\pm i[H,H^{\kappa}]\lesssim_{\kappa}H^{\kappa}, that is, for all u𝒟u\in\mathcal{D},

    |Hu,HκuHκu,Hu|κu,Hκu.\big{|}\langle Hu,H^{\kappa}u\rangle_{\mathcal{H}}-\langle H^{\kappa}u,Hu\rangle_{\mathcal{H}}\big{|}\,\lesssim_{\kappa}\,\langle u,H^{\kappa}u\rangle_{\mathcal{H}}.

We start with the proof of condition (i). On the one hand, since H1H_{1} and LL are nonnegative and commute strongly, we can deduce LuH1κu\|Lu\|_{\mathcal{H}}\leq\|H_{1}^{\kappa}u\|_{\mathcal{H}} for all uD(H1)D(L)u\in D(H_{1})\cap D(L). On the other hand, since H2H_{2} is a Kato perturbation of H1+LH_{1}+L, we find for all u𝒟u\in\mathcal{D},

H2uα(H1+L)u+CuαH1κu+Cu,\|H_{2}u\|_{\mathcal{H}}\leq\alpha\|(H_{1}+L)u\|_{\mathcal{H}}+C\|u\|_{\mathcal{H}}\leq\alpha\|H_{1}^{\kappa}u\|_{\mathcal{H}}+C\|u\|_{\mathcal{H}},

hence

H1κuHκu+H2uHκu+αH1κu+Cu,\|H_{1}^{\kappa}u\|_{\mathcal{H}}\leq\|H^{\kappa}u\|_{\mathcal{H}}+\|H_{2}u\|_{\mathcal{H}}\leq\|H^{\kappa}u\|_{\mathcal{H}}+\alpha\|H_{1}^{\kappa}u\|_{\mathcal{H}}+C\|u\|_{\mathcal{H}},

which leads to

H1κu11αHκu+C1αu.\|H_{1}^{\kappa}u\|_{\mathcal{H}}\leq\tfrac{1}{1-\alpha}\|H^{\kappa}u\|_{\mathcal{H}}+\tfrac{C}{1-\alpha}\|u\|_{\mathcal{H}}.

Combined with the above, this yields

LuH1κuHκu+u,\|Lu\|_{\mathcal{H}}\leq\|H_{1}^{\kappa}u\|_{\mathcal{H}}\lesssim\|H^{\kappa}u\|_{\mathcal{H}}+\|u\|_{\mathcal{H}},

hence in particular,

HuHκu+Lu+κuHκu+(1+κ)u,\displaystyle\|Hu\|_{\mathcal{H}}\,\leq\,\|H^{\kappa}u\|_{\mathcal{H}}+\|Lu\|_{\mathcal{H}}+\kappa\|u\|_{\mathcal{H}}\,\lesssim\,\|H^{\kappa}u\|_{\mathcal{H}}+(1+\kappa)\|u\|_{\mathcal{H}},

that is, (i). It remains to establish condition (ii). As H1H_{1} and LL commute, we can write for all u𝒟u\in\mathcal{D},

Hu,HκuHκu,Hu=Hu,LuLu,Hu=H2u,LuLu,H2u,\langle Hu,H^{\kappa}u\rangle_{\mathcal{H}}-\langle H^{\kappa}u,Hu\rangle_{\mathcal{H}}=\langle Hu,Lu\rangle_{\mathcal{H}}-\langle Lu,Hu\rangle_{\mathcal{H}}=\langle H_{2}u,Lu\rangle_{\mathcal{H}}-\langle Lu,H_{2}u\rangle_{\mathcal{H}},

hence, by assumption,

|Hu,HκuHκu,Hu|u,(H1+L+1)u.\big{|}\langle Hu,H^{\kappa}u\rangle_{\mathcal{H}}-\langle H^{\kappa}u,Hu\rangle_{\mathcal{H}}\big{|}\lesssim\langle u,(H_{1}+L+1)u\rangle_{\mathcal{H}}.

Again noting as in Step 1 that the nonnegativity of H1+LH_{1}+L and the fact that 2H22H_{2} is a Kato perturbation of H1+LH_{1}+L imply that (H1+L)+2H2+κ(H_{1}+L)+2H_{2}+\kappa is nonnegative for κ\kappa 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 [Hk,λst,1iCkst]=H1+H2[H_{k,\lambda}^{\operatorname{st}},\tfrac{1}{i}C_{k}^{\operatorname{st}}]=H_{1}+H_{2} on (Ω)\mathcal{R}(\Omega), in terms of

H1\displaystyle H_{1} :=\displaystyle:= 2(Hk,0st+|k|2),\displaystyle 2(H_{k,0}^{\operatorname{st}}+|k|^{2}),
H2\displaystyle H_{2} :=\displaystyle:= H2,1+H2,2,\displaystyle H_{2,1}+H_{2,2},
H2,1\displaystyle H_{2,1} :=\displaystyle:= ik12((st+ik)(Mkst2)+(Mkst2)(st+ik)),\displaystyle-ik\cdot\tfrac{1}{2}\big{(}(\nabla^{\operatorname{st}}+ik)(M_{k}^{\prime{\operatorname{st}}}-2)+(M_{k}^{\prime{\operatorname{st}}}-2)(\nabla^{\operatorname{st}}+ik)\big{)},
H2,2\displaystyle H_{2,2} :=\displaystyle:= λ([V,iAst]+12[V,iMkst]).\displaystyle-\lambda\big{(}[V,iA^{\operatorname{st}}]+\tfrac{1}{2}[V,iM_{k}^{\operatorname{st}}]\big{)}.

We shall appeal to Proposition 5.16 with L:=L:=\mathcal{L}, and it remains to check the different assumptions. First, H1=2(st+ik)(st+ik)H_{1}=-2(\nabla^{\operatorname{st}}+ik)\cdot(\nabla^{\operatorname{st}}+ik) and \mathcal{L} are both essentially self-adjoint on (Ω)\mathcal{R}(\Omega), 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 H1H_{1} and \mathcal{L} leave the linear subspace (Ω)\mathcal{R}(\Omega) invariant and that spectral projections of \mathcal{L} also leave (Ω)\mathcal{R}(\Omega) invariant. Using projections onto a finite number of chaoses, one then easily sees that H1+LH_{1}+L is essentially self-adjoint on (Ω)\mathcal{R}(\Omega). Next, we may rewrite as in (5.25),

H2,1=ik(Mkst2)(st+ik)+i2|k|2Mk′′st,H_{2,1}\,=\,\textstyle-ik\cdot(M_{k}^{\prime{\operatorname{st}}}-2)(\nabla^{{\operatorname{st}}}+ik)+\frac{i}{2}|k|^{2}M_{k}^{\prime\prime{\operatorname{st}}}, (5.26)

hence the boundedness of MkstM_{k}^{\prime{\operatorname{st}}} and Mk′′stM_{k}^{\prime\prime{\operatorname{st}}} leads to

H2,1ϕL2(Ω)|k|(st+ik)ϕL2(Ω)+|k|2ϕL2(Ω)|k|H112ϕL2(Ω)+|k|2ϕL2(Ω),\|H_{2,1}\phi\|_{\operatorname{L}^{2}(\Omega)}\,\lesssim\,\textstyle|k|\|(\nabla^{{\operatorname{st}}}+ik)\phi\|_{\operatorname{L}^{2}(\Omega)}+|k|^{2}\|\phi\|_{\operatorname{L}^{2}(\Omega)}\,\lesssim\,\textstyle|k|\|H_{1}^{\frac{1}{2}}\phi\|_{\operatorname{L}^{2}(\Omega)}+|k|^{2}\|\phi\|_{\operatorname{L}^{2}(\Omega)},

showing that H2,1H_{2,1} is H11/2H_{1}^{1/2}-bounded. Proposition 5.7(iii) and Lemma 5.14 also ensure that [V,iAst][V,iA^{\operatorname{st}}] and [V,iMkst][V,iM_{k}^{\operatorname{st}}] are 1/2\mathcal{L}^{1/2}-bounded. This proves that H2H_{2} is (H1+)1/2(H_{1}+\mathcal{L})^{1/2}-bounded, hence (H1+)(H_{1}+\mathcal{L})-infinitesimal. Finally, it remains to consider the commutator [H2,][H_{2},\mathcal{L}]. Since by definition MkstM_{k}^{\prime{\operatorname{st}}} and Mk′′stM_{k}^{\prime\prime{\operatorname{st}}} preserve chaoses, identity (5.26) implies [H2,1,]=0[H_{2,1},\mathcal{L}]=0. In view of the explicit description of [V,iAst][V,iA^{\operatorname{st}}] and [V,iMkst][V,iM_{k}^{\operatorname{st}}] in Proposition 5.7(iii) and in Lemma 5.14, respectively, as these commutators have a similar structure as VV itself on the Wiener chaos decomposition, a similar computation as in the proof of Proposition 5.7(iv) easily shows that [H2,2,][H_{2,2},\mathcal{L}] is 1/2\mathcal{L}^{1/2}-bounded, and the conclusion follows. ∎

5.7. Consequences of Mourre’s relations

This section is devoted to the proof of Corollary 5. Let V=V0V=V_{0} be a stationary Gaussian field. Given some L0>0L_{0}>0, consider the projection Qλ:=1[0,(L0λ)2]()Q_{\lambda}:=1_{[0,(L_{0}\lambda)^{-2}]}(\mathcal{L}) onto Wiener chaoses of order (L0λ)2\leq(L_{0}\lambda)^{-2}. We split the proof into three steps.

Step 1. Preliminary on spectral trunctations: For all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega), L1L\geq 1, and hL()h\in\operatorname{L}^{\infty}(\mathbb{R}) supported in 12[1L,1L]\mathbb{R}\setminus\frac{1}{2}[-\frac{1}{L},\frac{1}{L}], there holds

|𝔼[ϕ¯eitHk,λsth(Hk,λst)1]|λLhL()ϕL2(Ω),\big{|}\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}h(H_{k,\lambda}^{\operatorname{st}})1}\big{]}\!\big{|}\,\lesssim\,\lambda L\,\|h\|_{\operatorname{L}^{\infty}(\mathbb{R})}\|\phi\|_{\operatorname{L}^{2}(\Omega)}, (5.27)

where the factor λL\lambda L can be replaced by (λL)2(\lambda L)^{2} if ϕ=1\phi=1.

Choose h0C()h_{0}\in C^{\infty}(\mathbb{R}) with h0(y)=1h_{0}(y)=1 for |y|12L|y|\geq\frac{1}{2L} and h0(y)=0h_{0}(y)=0 for |y|14L|y|\leq\frac{1}{4L} such that h00h_{0}\geq 0 and |h0|8L|h^{\prime}_{0}|\leq 8L pointwise. By definition, we find |h|hL()|h0||h|\leq\|h\|_{\operatorname{L}^{\infty}(\mathbb{R})}|h_{0}|, hence

h(Hk,λst)1L2(Ω)hL()h0(Hk,λst)1L2(Ω).\|h(H_{k,\lambda}^{\operatorname{st}})1\|_{\operatorname{L}^{2}(\Omega)}\leq\|h\|_{\operatorname{L}^{\infty}(\mathbb{R})}\|h_{0}(H_{k,\lambda}^{\operatorname{st}})1\|_{\operatorname{L}^{2}(\Omega)}.

Using spectral calculus with h0(0)=0h_{0}(0)=0 and Hk,λst1=λVH_{k,\lambda}^{\operatorname{st}}1=\lambda V, this leads to

h(Hk,λst)1L2(Ω)\displaystyle\|h(H_{k,\lambda}^{\operatorname{st}})1\|_{\operatorname{L}^{2}(\Omega)} \displaystyle\leq hL()01h0(sHk,λst)Hk,λst1𝑑sL2(Ω)\displaystyle\|h\|_{\operatorname{L}^{\infty}(\mathbb{R})}\Big{\|}\int_{0}^{1}h_{0}^{\prime}(sH_{k,\lambda}^{\operatorname{st}})H_{k,\lambda}^{\operatorname{st}}1\,ds\Big{\|}_{\operatorname{L}^{2}(\Omega)}
\displaystyle\leq 8λLhL()VL2(Ω).\displaystyle 8\,\lambda L\,\|h\|_{\operatorname{L}^{\infty}(\mathbb{R})}\|V\|_{\operatorname{L}^{2}(\Omega)}.

If ϕ\phi is nonconstant, the conclusion directly follows from the Cauchy–Schwarz inequality. If ϕ=1\phi=1, decomposing hh into its positive and negative parts h+h_{+} and hh_{-}, we can rather estimate

|𝔼[eitHk,λsth(Hk,λst)1]|\displaystyle\big{|}\mathbb{E}\big{[}{e^{-itH_{k,\lambda}^{\operatorname{st}}}h(H_{k,\lambda}^{\operatorname{st}})1}\big{]}\!\big{|} \displaystyle\leq (h+)12(Hk,λst)1L2(Ω)2+(h)12(Hk,λst)1L2(Ω)2\displaystyle\big{\|}(h_{+})^{\frac{1}{2}}(H_{k,\lambda}^{\operatorname{st}})1\big{\|}_{\operatorname{L}^{2}(\Omega)}^{2}+\big{\|}(h_{-})^{\frac{1}{2}}(H_{k,\lambda}^{\operatorname{st}})1\big{\|}_{\operatorname{L}^{2}(\Omega)}^{2}
\displaystyle\leq 128(λL)2hL()VL2(Ω)2.\displaystyle 128\,(\lambda L)^{2}\|h\|_{\operatorname{L}^{\infty}(\mathbb{R})}\|V\|_{\operatorname{L}^{2}(\Omega)}^{2}.

Step 2. Proof that for all kd{0}k\in\mathbb{R}^{d}\setminus\{0\} the flow for the truncated fibered Schrödinger operator QλHk,λstQλQ_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda} satisfies for all s0s\geq 0,

limλ0|𝔼[eiλ2sQλHk,λstQλ1]es(αk+iβk)|= 0.\lim_{\lambda\downarrow 0}\Big{|}\mathbb{E}\big{[}{e^{-i\lambda^{-2}s\,Q_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda}}1}\big{]}-e^{-s(\alpha_{k}+i\beta_{k})}\Big{|}\,=\,0. (5.28)

We wish to apply Theorem 5.2 in form of Remark 5.3 with Hλ:=QλHk,0stQλH_{\lambda}^{\circ}:=Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda}, Wλ:=QλVQλW_{\lambda}^{\circ}:=Q_{\lambda}VQ_{\lambda}, E0:=0E_{0}:=0, ψ0:=1\psi_{0}:=1, and it suffices to check the different assumptions. First, as the conjugate operator CkstC_{k}^{\operatorname{st}} commutes with the Wiener chaos decomposition, Theorem 5.8(i) ensures that the truncated operator QλHk,0stQλQ_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda} restricted to L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C} satisfies a strict Mourre relation on Jε:=[ε34|k|2,)J_{\varepsilon}:=[\varepsilon-\frac{3}{4}|k|^{2},\infty) for all ε>0\varepsilon>0: we find

[QλHk,0stQλ,1iCkst]ΠQλ(Hk,0st+34|k|2)QλΠ,\big{[}Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda},\tfrac{1}{i}C_{k}^{\operatorname{st}}\big{]}\geq\Pi Q_{\lambda}\big{(}H_{k,0}^{\operatorname{st}}+\tfrac{3}{4}|k|^{2}\big{)}Q_{\lambda}\Pi,

hence, on L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C},

𝟙Jε(QλHk,0stQλ)[QλHk,0stQλ,1iCkst]𝟙Jε(QλHk,0stQλ)ε 1Jε(QλHk,0stQλ).\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda})\big{[}Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda},\tfrac{1}{i}C_{k}^{\operatorname{st}}\big{]}\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda})\geq\varepsilon\,\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda}).

Similar computations as in the proof of Theorem 5.8(i) show the Hk,0stH_{k,0}^{\operatorname{st}}-boundedness of iterated commutators adCkstk(Hk,0st)\operatorname{ad}^{k}_{C_{k}^{\operatorname{st}}}(H_{k,0}^{\operatorname{st}}), hence the λ\lambda-uniform QλHk,0stQλQ_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda}-boundedness of

adCkstk(QλHk,0stQλ)=QλadCkstk(Hk,0st)Qλ.\operatorname{ad}^{k}_{C_{k}^{\operatorname{st}}}(Q_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda})=Q_{\lambda}\operatorname{ad}^{k}_{C_{k}^{\operatorname{st}}}(H_{k,0}^{\operatorname{st}})\,Q_{\lambda}.

In addition, as the domain of Hk,0stH_{k,0}^{\operatorname{st}} is H2(Ω)H^{2}(\Omega), as Theorem 5.8(i) ensures that H2(Ω)H^{2}(\Omega) is invariant under the unitary group generated by CkstC_{k}^{\operatorname{st}}, and as the latter further preserves the chaos decomposition, we deduce that the domain of QλHk,0stQλQ_{\lambda}H_{k,0}^{\operatorname{st}}Q_{\lambda} is also invariant under the unitary group generated by CkstC_{k}^{\operatorname{st}}.

It remains to check assumptions on the perturbation QλVQλQ_{\lambda}VQ_{\lambda}. Note that Π(QλVQλ1)=V\Pi(Q_{\lambda}VQ_{\lambda}1)=V, which clearly satisfies Ckst6VL2(Ω)1\|\langle C_{k}^{\operatorname{st}}\rangle^{6}V\|_{\operatorname{L}^{2}(\Omega)}\lesssim 1. Further, iterating the proof of Theorem 5.8(ii), we find that for all k0k\geq 0 iterated commutators adCkstk(V)\operatorname{ad}^{k}_{C_{k}^{\operatorname{st}}}(V) are 1/2\mathcal{L}^{1/2}-bounded, hence for all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega),

adCkstk(λQλVQλ)ϕL2(Ω)λQλ1/2ϕL2(Ω)+λϕL2(Ω),\big{\|}\operatorname{ad}^{k}_{C_{k}^{\operatorname{st}}}(\lambda Q_{\lambda}VQ_{\lambda})\phi\big{\|}_{\operatorname{L}^{2}(\Omega)}\lesssim\lambda\big{\|}Q_{\lambda}\mathcal{L}^{1/2}\phi\big{\|}_{\operatorname{L}^{2}(\Omega)}+\lambda\|\phi\|_{\operatorname{L}^{2}(\Omega)},

which entails, by definition of QλQ_{\lambda},

adCkstk(λQλVQλ)ϕL2(Ω)(λ+1L0)ϕL2(Ω).\big{\|}\operatorname{ad}^{k}_{C_{k}^{\operatorname{st}}}(\lambda Q_{\lambda}VQ_{\lambda})\phi\big{\|}_{\operatorname{L}^{2}(\Omega)}\lesssim\big{(}\lambda+\tfrac{1}{L_{0}}\big{)}\|\phi\|_{\operatorname{L}^{2}(\Omega)}.

Choosing L01L_{0}\simeq 1 large enough, we may then apply Theorem 5.2 in form of Remark 5.3, to the effect of

𝔼[eiQλHk,λstQλt 1Jε(QλHk,λstQλ) 1]=eλ2t(αk+iβk)+o(1).\mathbb{E}\left[{e^{-iQ_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda}t}\,\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda})\,1}\right]=e^{-\lambda^{2}t(\alpha_{k}+i\beta_{k})}+o(1).

In view of Step 1, the spectral truncation 𝟙Jε(QλHk,λstQλ)\mathds{1}_{J_{\varepsilon}}(Q_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda}) can be removed up to a further O(λ2)O(\lambda^{2}) error, and the claim (5.28) follows.

Step 3. Conclusion.
In view of the result of Step 1, it remains to prove for all 0s(e2𝒞0(0)12L0)10\leq s\leq(e^{2}\mathcal{C}_{0}(0)^{\frac{1}{2}}L_{0})^{-1},

limλ0eiλ2sQλHk,λstQλ1eiλ2sHk,λst1L2(Ω)= 0,\lim_{\lambda\downarrow 0}\Big{\|}e^{-i\lambda^{-2}s\,Q_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda}}1-e^{-i\lambda^{-2}sH_{k,\lambda}^{\operatorname{st}}}1\Big{\|}_{\operatorname{L}^{2}(\Omega)}\,=\,0, (5.29)

while the conclusion of Corollary 5 then follows from the fibration (1.4). Set for abbreviation uk,λt:=eitHk,λst1u_{k,\lambda}^{t}:=e^{-itH_{k,\lambda}^{\operatorname{st}}}1 and u~k,λt:=eitQλHk,λstQλ1\tilde{u}_{k,\lambda}^{t}:=e^{-itQ_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda}}1. Since the flow uk,λu_{k,\lambda} satisfies the equation ituk,λ=(Hk,0st+λV)uk,λi\partial_{t}u_{k,\lambda}=(H_{k,0}^{\operatorname{st}}+\lambda V)u_{k,\lambda}, an iterative use of Duhamel’s formula allows to decompose, for all N1N\geq 1,

uk,λt=p=0N(iλ)pukp;t+(iλ)N+1Ek,λN;t,u_{k,\lambda}^{t}\,=\,\sum_{p=0}^{N}(-i\lambda)^{p}u_{k}^{p;t}+(-i\lambda)^{N+1}E_{k,\lambda}^{N;t},

in terms of

ukp;t\displaystyle u_{k}^{p;t} :=\displaystyle:= (+)p+1δ(tj=1p+1sj)eis1Hk,0stVeis2Hk,0stVeispHk,0stV𝑑s1𝑑sp+1,\displaystyle\int_{(\mathbb{R}^{+})^{p+1}}\delta\Big{(}t-\sum_{j=1}^{p+1}s_{j}\Big{)}\,e^{-is_{1}H_{k,0}^{\operatorname{st}}}Ve^{-is_{2}H_{k,0}^{\operatorname{st}}}\ldots Ve^{-is_{p}H_{k,0}^{\operatorname{st}}}V\,ds_{1}\ldots ds_{p+1},
Ek,λN;t\displaystyle E_{k,\lambda}^{N;t} :=\displaystyle:= (+)N+2δ(tj=1N+2sj)eis1Hk,λstVeis2Hk,0stVeisN+1Hk,0stV𝑑s1𝑑sN+2.\displaystyle\int_{(\mathbb{R}^{+})^{N+2}}\delta\Big{(}t-\sum_{j=1}^{N+2}s_{j}\Big{)}\,e^{-is_{1}H_{k,\lambda}^{\operatorname{st}}}Ve^{-is_{2}H_{k,0}^{\operatorname{st}}}\ldots Ve^{-is_{N+1}H_{k,0}^{\operatorname{st}}}V\,ds_{1}\ldots ds_{N+2}.

Noting that ukp;tpu_{k}^{p;t}\in\mathcal{H}_{p} for all p0p\geq 0, we deduce that the truncation error can be represented as follows, for any N(L0λ)2N\leq(L_{0}\lambda)^{-2},

uk,λtu~k,λt=(iλ)N+1(Ek,λN;tE~k,λN;t),u_{k,\lambda}^{t}-\tilde{u}_{k,\lambda}^{t}=(-i\lambda)^{N+1}(E_{k,\lambda}^{N;t}-\tilde{E}_{k,\lambda}^{N;t}), (5.30)

where E~k,λN;t\tilde{E}_{k,\lambda}^{N;t} is defined similarly as Ek,λN;tE_{k,\lambda}^{N;t} with VV and Hk,λstH_{k,\lambda}^{\operatorname{st}} replaced by QλVQλQ_{\lambda}VQ_{\lambda} and QλHk,λstQλQ_{\lambda}H_{k,\lambda}^{\operatorname{st}}Q_{\lambda}, respectively. A direct estimate yields

Ek,λN;tL2(Ω)(+)N+2δ(tj=1N+2sj)Veis2Hk,0stVeisN+1Hk,0stVL2(Ω)𝑑s1𝑑sN+2,\|E_{k,\lambda}^{N;t}\|_{\operatorname{L}^{2}(\Omega)}\,\leq\,\int_{(\mathbb{R}^{+})^{N+2}}\delta\Big{(}t-\sum_{j=1}^{N+2}s_{j}\Big{)}\,\big{\|}Ve^{-is_{2}H_{k,0}^{\operatorname{st}}}\ldots Ve^{-is_{N+1}H_{k,0}^{\operatorname{st}}}V\big{\|}_{\operatorname{L}^{2}(\Omega)}\,ds_{1}\ldots ds_{N+2},

hence, noting that VV is bounded by 𝒞0(0)12(2p+1)12\mathcal{C}_{0}(0)^{\frac{1}{2}}(2p+1)^{\frac{1}{2}} on npn\bigcup_{n\leq p}\mathcal{H}_{n}, cf. Lemma 5.6,

Ek,λN;tL2(Ω)\displaystyle\|E_{k,\lambda}^{N;t}\|_{\operatorname{L}^{2}(\Omega)} \displaystyle\leq 𝒞0(0)12(N+1)(2N+1)!!12(+)N+2δ(tj=1N+2sj)𝑑s1𝑑sN+2\displaystyle\mathcal{C}_{0}(0)^{\frac{1}{2}(N+1)}(2N+1)!!^{\frac{1}{2}}\int_{(\mathbb{R}^{+})^{N+2}}\delta\Big{(}t-\sum_{j=1}^{N+2}s_{j}\Big{)}\,ds_{1}\ldots ds_{N+2}
\displaystyle\leq (2N+1)!!12(N+1)!(𝒞0(0)12t)N+1(e𝒞0(0)12t)N+1(N+1)!12.\displaystyle\frac{(2N+1)!!^{\frac{1}{2}}}{(N+1)!}(\mathcal{C}_{0}(0)^{\frac{1}{2}}t)^{N+1}\,\leq\,\frac{(e\mathcal{C}_{0}(0)^{\frac{1}{2}}t)^{N+1}}{(N+1)!^{\frac{1}{2}}}.

Similarly estimating E~k,λN;t\tilde{E}_{k,\lambda}^{N;t} and inserting this into (5.30), we find for all N(L0λ)2N\leq(L_{0}\lambda)^{-2},

uk,λtu~k,λtL2(Ω)2(e𝒞0(0)12λt)N+1(N+1)!12.\|u_{k,\lambda}^{t}-\tilde{u}_{k,\lambda}^{t}\|_{\operatorname{L}^{2}(\Omega)}\leq\frac{2(e\mathcal{C}_{0}(0)^{\frac{1}{2}}\lambda t)^{N+1}}{(N+1)!^{\frac{1}{2}}}.

Setting t=λ2st=\lambda^{-2}s and choosing N=(L0λ)2N=\lfloor(L_{0}\lambda)^{-2}\rfloor, we easily deduce for s(e2𝒞0(0)12L0)1s\leq(e^{2}\mathcal{C}_{0}(0)^{\frac{1}{2}}L_{0})^{-1},

uk,λλ2su~k,λλ2sL2(Ω)λ1e12(L0λ)2,\|u_{k,\lambda}^{\lambda^{-2}s}-\tilde{u}_{k,\lambda}^{\lambda^{-2}s}\|_{\operatorname{L}^{2}(\Omega)}\lesssim\lambda^{-1}e^{-\frac{1}{2}(L_{0}\lambda)^{-2}},

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 ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega), the Floquet–Bloch fibration (1.4) takes the form

𝔼[ϕ¯uλt(x)]=du^(k)eikxit|k|2𝔼[ϕ(τx)¯eitHk,λst1]¯𝑑k,\mathbb{E}\left[{\bar{\phi}\,u_{\lambda}^{t}(x)}\right]\,=\,\int_{\mathbb{R}^{d}}\widehat{u}^{\circ}(k)\,e^{ik\cdot x-it|k|^{2}}\,\mathbb{E}\big{[}{\overline{\phi(\tau_{x}\cdot)}e^{-itH_{k,\lambda}^{\operatorname{st}}}1}\big{]}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,

hence it suffices analyze 𝔼[ϕ¯eitHk,λst1]\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}1}\big{]} for fixed kk and ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega). 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 ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega), the spectral measure μk,λϕ,1\mu_{k,\lambda}^{\phi,1} is analytic on [1M,1M][-\frac{1}{M},\frac{1}{M}] and admits a local meromorphic extension νk,λϕ,1\nu_{k,\lambda}^{\phi,1} on the complex neighborhood 1MB\frac{1}{M}B,

νk,λϕ,1(z)=12iπ(ϕ,Πk,λ1𝒫(Ω),𝒫(Ω)zk,λzϕ,Πk,λ1𝒫(Ω),𝒫(Ω)zk,λ¯z)+12iπ(ζk,λϕ,1(z)ζk,λ1,ϕ¯(z)),\nu_{k,\lambda}^{\phi,1}(z)\,=\,\frac{1}{2i\pi}\bigg{(}\frac{\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{z_{k,\lambda}-z}-\frac{\langle\phi,\Pi_{k,\lambda}^{*}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{\overline{z_{k,\lambda}}-z}\bigg{)}+\frac{1}{2i\pi}\big{(}\zeta_{k,\lambda}^{\phi,1}(z)-\overline{\zeta_{k,\lambda}^{1,\phi}}(z)\big{)}, (6.1)

which can alternatively be expressed as

νk,λϕ,1(z)=λzνk,λϕ,V(z)\displaystyle\nu_{k,\lambda}^{\phi,1}(z)\,=\,\frac{\lambda}{z}\nu_{k,\lambda}^{\phi,V}(z)
=\displaystyle= λ2iπz(ϕ,Πk,λV𝒫(Ω),𝒫(Ω)zk,λzϕ,Πk,λV𝒫(Ω),𝒫(Ω)zk,λ¯z)+λ2iπz(ζk,λϕ,V(z)ζk,λV,ϕ¯(z)),\displaystyle\frac{\lambda}{2i\pi z}\bigg{(}\frac{\langle\phi,\Pi_{k,\lambda}V\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{z_{k,\lambda}-z}-\frac{\langle\phi,\Pi_{k,\lambda}^{*}V\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{\overline{z_{k,\lambda}}-z}\bigg{)}+\frac{\lambda}{2i\pi z}\big{(}\zeta_{k,\lambda}^{\phi,V}(z)-\overline{\zeta_{k,\lambda}^{V,\phi}}(z)\big{)},

and moreover, in case ϕ=1\phi=1,

νk,λ1,1(z)\displaystyle\nu_{k,\lambda}^{1,1}(z) =\displaystyle= λ2z2νk,λV,V(z)\displaystyle\frac{\lambda^{2}}{z^{2}}\nu_{k,\lambda}^{V,V}(z)
=\displaystyle= λ22iπz2(V,Πk,λV𝒫(Ω),𝒫(Ω)zk,λzV,Πk,λV𝒫(Ω),𝒫(Ω)zk,λ¯z)+λ2πz2(ζk,λV,V)(z),\displaystyle\frac{\lambda^{2}}{2i\pi z^{2}}\bigg{(}\frac{\langle V,\Pi_{k,\lambda}V\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{z_{k,\lambda}-z}-\frac{\langle V,\Pi_{k,\lambda}^{*}V\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{\overline{z_{k,\lambda}}-z}\bigg{)}+\frac{\lambda^{2}}{\pi z^{2}}\big{(}\Im\zeta_{k,\lambda}^{V,V}\big{)}(z),

where we write for abbreviation

ϕ,Πk,λϕ𝒫(Ω),𝒫(Ω)\displaystyle\langle\phi^{\prime},\Pi_{k,\lambda}\phi\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)} =\displaystyle= Ψk,λ+,ϕ𝒫(Ω),𝒫(Ω)¯Ψk,λ,ϕ𝒫(Ω),𝒫(Ω),\displaystyle\overline{\langle\Psi_{k,\lambda}^{+},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\Psi_{k,\lambda}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}, (6.4)
ϕ,Πk,λϕ𝒫(Ω),𝒫(Ω)\displaystyle\langle\phi^{\prime},\Pi^{*}_{k,\lambda}\phi\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)} =\displaystyle= Ψk,λ,ϕ𝒫(Ω),𝒫(Ω)¯Ψk,λ+,ϕ𝒫(Ω),𝒫(Ω).\displaystyle\overline{\langle\Psi_{k,\lambda}^{-},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\Psi_{k,\lambda}^{+},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}.

Indeed, Stone’s formula together with Conjecture Conjecture (LRC) — Local resonance conjecture yields for |y|<1M|y|<\frac{1}{M},

μk,λϕ,ϕ(y)=limε012iπ(ϕ,(Hk,λstyiε)1ϕL2(Ω)ϕ,(Hk,λsty+iε)1ϕL2(Ω))\displaystyle\mu_{k,\lambda}^{\phi^{\prime},\phi}(y)\,=\,\lim_{\varepsilon\downarrow 0}\frac{1}{2i\pi}\Big{(}\big{\langle}\phi^{\prime},(H^{\operatorname{st}}_{k,\lambda}-y-i\varepsilon)^{-1}\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}-\big{\langle}\phi^{\prime},(H^{\operatorname{st}}_{k,\lambda}-y+i\varepsilon)^{-1}\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\Big{)}
=\displaystyle= limε012iπ(ϕ,(Hk,λstyiε)1ϕL2(Ω)ϕ,(Hk,λstyiε)1ϕL2(Ω)¯)\displaystyle\lim_{\varepsilon\downarrow 0}\frac{1}{2i\pi}\Big{(}\big{\langle}\phi^{\prime},(H^{\operatorname{st}}_{k,\lambda}-y-i\varepsilon)^{-1}\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}-\overline{\big{\langle}\phi,(H^{\operatorname{st}}_{k,\lambda}-y-i\varepsilon)^{-1}\phi^{\prime}\big{\rangle}_{\operatorname{L}^{2}(\Omega)}}\Big{)}
=\displaystyle= 12iπ(ϕ,Πk,λϕ𝒫(Ω),𝒫(Ω)zk,λyϕ,Πk,λϕ𝒫(Ω),𝒫(Ω)zk,λ¯y)+12iπ(ζk,λϕ,ϕ(y)ζk,λϕ,ϕ¯(y)),\displaystyle\frac{1}{2i\pi}\bigg{(}\frac{\langle\phi^{\prime},\Pi_{k,\lambda}\phi\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{z_{k,\lambda}-y}-\frac{\langle\phi^{\prime},\Pi_{k,\lambda}^{*}\phi\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}}{\overline{z_{k,\lambda}}-y}\bigg{)}+\frac{1}{2i\pi}\big{(}\zeta_{k,\lambda}^{\phi^{\prime},\phi}(y)-\overline{\zeta_{k,\lambda}^{\phi,\phi^{\prime}}}(y)\big{)},

and (6.1) follows. Identities (6.1) and (6.1) are obvious consequences as Hk,λst1=λVH_{k,\lambda}^{\operatorname{st}}1=\lambda V.

Step 2. Proof of (i): Under Conjecture (LRC) — Local resonance conjecture, for all kKk\in K, 0λ<λ00\leq\lambda<\lambda_{0}, ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega), n0n\geq 0, and gCc()g\in C^{\infty}_{c}(\mathbb{R}) supported in [1M,1M][-\frac{1}{M},\frac{1}{M}] with g=1g=1 in 12[1M,1M]\frac{1}{2}[-\frac{1}{M},\frac{1}{M}], there holds

|𝔼[ϕ¯eitHk,λstg(Hk,λst)1]eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)|n,ϕ,g,Mλ(1+t)n,\Big{|}\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}g(H_{k,\lambda}^{\operatorname{st}})1}\big{]}-e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}\Big{|}\,\lesssim_{n,\phi,g,M}\,\lambda(1+t)^{-n},

where the factor λ\lambda in the right-hand side can be replaced by λ2\lambda^{2} in case ϕ=1\phi=1. Combined with (5.27) to remove the spectral truncation, this indeed yields (i).

Starting from

𝔼[ϕ¯eitHk,λstg(Hk,λst)1]=[1M,1M]eityg(y)μk,λϕ,1(y)𝑑y,\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}g(H_{k,\lambda}^{\operatorname{st}})1}\big{]}\,=\,\int_{[-\frac{1}{M},\frac{1}{M}]}e^{-ity}g(y)\,\mu_{k,\lambda}^{\phi,1}(y)\,dy,

and using formula (6.1) for the meromorphic extension νk,λϕ,1\nu_{k,\lambda}^{\phi,1} of the spectral measure μk,λϕ,1\mu_{k,\lambda}^{\phi,1}, we obtain by contour deformation,

𝔼[ϕ¯eitHk,λstg(Hk,λst)1]=eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)+γeitzg(z)νk,λϕ,1(z)𝑑z,\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}g(H_{k,\lambda}^{\operatorname{st}})1}\big{]}\,=\,e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}+\int_{\gamma}e^{-itz}g(\Re z)\,\nu_{k,\lambda}^{\phi,1}(z)\,dz,

where the smooth path γ\gamma is a deformation of the real interval [1M,1M][-\frac{1}{M},\frac{1}{M}] in the lower half-plane such that γ\gamma remains on the real axis on [1M,1M]12[1M,1M][-\frac{1}{M},\frac{1}{M}]\setminus\frac{1}{2}[-\frac{1}{M},\frac{1}{M}] while the part on 12[1M,1M]\frac{1}{2}[-\frac{1}{M},\frac{1}{M}] is deformed into a path in {z:z0,|z|12M}1MB\{z:\Im z\leq 0,|\Re z|\leq\frac{1}{2M}\}\bigcap\frac{1}{M}B that stays pointwise at a distance 14M\frac{1}{4M} from the origin. Using the identity eitz=(1+t)1(1+iddz)eitze^{-itz}\,=\,(1+t)^{-1}(1+i\tfrac{d}{dz})e^{-itz} and integrating by parts, we find iteratively for all n0n\geq 0,

|𝔼[ϕ¯eitHk,λstg(Hk,λst)1]eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)|(1+t)nγ|(1iddz)n(g(z)νk,λϕ,1(z))|d|z|.\Big{|}\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}g(H_{k,\lambda}^{\operatorname{st}})1}\big{]}-e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}\Big{|}\\ \,\leq\,(1+t)^{-n}\int_{\gamma}\big{|}(1-i\tfrac{d}{dz})^{n}\big{(}g(\Re z)\,\nu_{k,\lambda}^{\phi,1}(z)\big{)}\big{|}\,d|z|.

As the remainder ζk,λϕ,1\zeta_{k,\lambda}^{\phi,1} is holomorphic on 1MB\frac{1}{M}B and has continuous dependence on λ\lambda for 0λ<λ00\leq\lambda<\lambda_{0}, cf. Conjecture (LRC) — Local resonance conjecture, we deduce that all its derivatives are bounded on 12MB\frac{1}{2M}B uniformly with respect to λ\lambda. Hence, it follows from (6.1) that all derivatives of νk,λϕ,1\nu_{k,\lambda}^{\phi,1} are uniformly bounded by O(λ)O(\lambda) on the path γ\gamma, and thus

|𝔼[ϕ¯eitHk,λstg(Hk,λst)1]eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)|n,ϕ,g,Mλ(1+t)n,\Big{|}\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}g(H_{k,\lambda}^{\operatorname{st}})1}\big{]}-e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}\Big{|}\,\lesssim_{n,\phi,g,M}\,\lambda(1+t)^{-n}, (6.5)

where in view of (6.1) the factor λ\lambda can be replaced by λ2\lambda^{2} in case ϕ=1\phi=1.

Step 3. Proof of (ii): Under Conjecture (GRC) — Global resonance conjecture, for all kKk\in K, 0λ<λ00\leq\lambda<\lambda_{0}, and ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega),

|𝔼[ϕ¯eitHk,λst1]eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)|ϕλet8Mλρ,\Big{|}\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}1}\big{]}-e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}\Big{|}\,\lesssim_{\phi}\,\lambda e^{-\frac{t}{8M}\lambda^{\rho}},

where the factor λ\lambda in the right-hand side can be replaced by λ2\lambda^{2} in case ϕ=1\phi=1.

Similarly as in Step 2, applying formula (6.1) and contour deformation, we find

𝔼[ϕ¯eitHk,λst1]=eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)+γeitzνk,λϕ,1(z)𝑑z,\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}1}\big{]}\,=\,e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}+\int_{\gamma}e^{-itz}\nu_{k,\lambda}^{\phi,1}(z)\,dz,

where the smooth path γ\gamma is given by γ:=γ3γ2γ1γ0γ1+γ2+γ3+\gamma:=\gamma_{3}^{-}\cup\gamma_{2}^{-}\cup\gamma_{1}^{-}\cup\gamma_{0}\cup\gamma_{1}^{+}\cup\gamma_{2}^{+}\cup\gamma_{3}^{+} with

  1. \bullet

    γ0:=γ{z:|z|12M}\gamma_{0}:=\gamma\cap\{z:|\Re z|\leq\frac{1}{2M}\} connects 12M(1iλρ)\frac{1}{2M}(-1-i\lambda^{\rho}) and 12M(1iλρ)\frac{1}{2M}(1-i\lambda^{\rho}), does not exit the ball of radius 1M\frac{1}{M}, and always stays at a distance 14M\frac{1}{4M} from the origin;

  2. \bullet

    γ1:=i2Mλρ+[L,12M]\gamma_{1}^{-}:=-\frac{i}{2M}\lambda^{\rho}+[-L,-\frac{1}{2M}] and γ1+:=i2Mλρ+[12M,L]\gamma_{1}^{+}:=-\frac{i}{2M}\lambda^{\rho}+[\frac{1}{2M},L];

  3. \bullet

    γ2±\gamma_{2}^{\pm} connects ±Li2Mλρ\pm L-\frac{i}{2M}\lambda^{\rho} and ±(L+1)\pm(L+1) without crossing the real axis;

  4. \bullet

    γ3:=(,L1]\gamma_{3}^{-}:=(-\infty,-L-1] and γ3+:=[L+1,)\gamma_{3}^{+}:=[L+1,\infty);

where L2L\geq 2 is to be fixed later. Inserting formula (6.1) for νk,λϕ,1\nu_{k,\lambda}^{\phi,1}, using the uniform bound assumed to hold on γ\gamma, cf. Conjecture (GRC) — Global resonance conjecture, and setting γ3:=γ3γ3+\gamma_{3}:=\gamma_{3}^{-}\cup\gamma_{3}^{+}, the above turns into

|𝔼[ϕ¯eitHk,λstg(Hk,λst)1]eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)|ϕ,gλ1Mγγ3|eitz||z|d|z|+λ|γ3eityyμk,λϕ,V(y)𝑑y|.\Big{|}\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}g(H_{k,\lambda}^{\operatorname{st}})1}\big{]}-e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}\Big{|}\\ \,\lesssim_{\phi,g}\,\lambda^{1-M}\int_{\gamma\setminus\gamma_{3}}\frac{|e^{-itz}|}{|z|}\,d|z|+\lambda\,\Big{|}\int_{\gamma_{3}}\frac{e^{-ity}}{y}\mu_{k,\lambda}^{\phi,V}(y)\,dy\Big{|}.

Hence, by definition of γ\gamma,

|𝔼[ϕ¯eitHk,λstg(Hk,λst)1]eitzk,λϕ,Πk,λ1𝒫(Ω),𝒫(Ω)|ϕ,gλ1ML1+λ1Met2Mλρ(1+1MLdyy)ϕ,gλ1M(L1+et2MλρlogL).\Big{|}\mathbb{E}\big{[}{\bar{\phi}\,e^{-itH_{k,\lambda}^{\operatorname{st}}}g(H_{k,\lambda}^{\operatorname{st}})1}\big{]}-e^{-itz_{k,\lambda}}\langle\phi,\Pi_{k,\lambda}1\rangle_{\mathcal{P}(\Omega),\mathcal{P}^{\prime}(\Omega)}\Big{|}\\ \,\lesssim_{\phi,g}\,\lambda^{1-M}L^{-1}+\lambda^{1-M}e^{-\frac{t}{2M}\lambda^{\rho}}\bigg{(}1+\int_{\frac{1}{M}}^{L}\frac{dy}{y}\bigg{)}\,\lesssim_{\phi,g}\,\lambda^{1-M}\Big{(}L^{-1}+e^{-\frac{t}{2M}\lambda^{\rho}}\log L\Big{)}.

Optimization in L2L\geq 2 yields the bound λ1Met4Mλρ\lambda^{1-M}e^{-\frac{t}{4M}\lambda^{\rho}}. 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 ϕ,ϕ𝒫(Ω)\phi,\phi^{\prime}\in\mathcal{P}(\Omega) the map

[0,λ0)×𝒫(Ω)×𝒫(Ω)×Lloc(1MB):λ(zk,λ,Ψk,λ+,Ψk,λ,ζk,λϕ,ϕ)[0,\lambda_{0})\to\mathbb{C}\times\mathcal{P}^{\prime}(\Omega)\times\mathcal{P}^{\prime}(\Omega)\times\operatorname{L}^{\infty}_{\operatorname{loc}}(\tfrac{1}{M}B):\lambda\mapsto\big{(}z_{k,\lambda},\Psi_{k,\lambda}^{+},\Psi_{k,\lambda}^{-},\zeta_{k,\lambda}^{\phi,\phi^{\prime}}\big{)}

is of class C2C^{2}, we iteratively compute the first two derivatives,

(zk,0,Ψk,0+,Ψk,0)\displaystyle\big{(}z_{k,0},\Psi_{k,0}^{+},\Psi_{k,0}^{-}\big{)} :=\displaystyle:= (zk,λ,Ψk,λ+,Ψk,λ)|λ=0,\displaystyle\big{(}z_{k,\lambda},\Psi_{k,\lambda}^{+},\Psi_{k,\lambda}^{-}\big{)}\big{|}_{\lambda=0},
(zk,0,Ψk,0+,Ψk,0)\displaystyle\big{(}z_{k,0}^{\prime},\Psi_{k,0}^{+\prime},\Psi_{k,0}^{-\prime}\big{)} :=\displaystyle:= ddλ(zk,λ,Ψk,λ+,Ψk,λ)|λ=0,\displaystyle\tfrac{d}{d\lambda}\big{(}z_{k,\lambda},\Psi_{k,\lambda}^{+},\Psi_{k,\lambda}^{-}\big{)}\big{|}_{\lambda=0},
(zk,0′′,Ψk,0+′′,Ψk,0′′)\displaystyle\big{(}z_{k,0}^{\prime\prime},\Psi_{k,0}^{+\prime\prime},\Psi_{k,0}^{-\prime\prime}\big{)} :=\displaystyle:= (ddλ)2(zk,λ,Ψk,λ+,Ψk,λ)|λ=0.\displaystyle\big{(}\tfrac{d}{d\lambda}\big{)}^{2}\big{(}z_{k,\lambda},\Psi_{k,\lambda}^{+},\Psi_{k,\lambda}^{-}\big{)}\big{|}_{\lambda=0}.

Note that the resonant and co-resonant states (Ψk,λ+,Ψk,λ)(\Psi^{+}_{k,\lambda},\Psi^{-}_{k,\lambda}) are only defined up to multiplication by (aλ,a¯λ1)(a_{\lambda},\bar{a}_{\lambda}^{-1}) for any complex-valued function λaλ\lambda\mapsto a_{\lambda}, cf. (2.5). When differentiating, this gauge invariance implies that (Ψk,0+,Ψk,0)(\Psi^{+}_{k,0},\Psi^{-}_{k,0}) is only defined up to multiplication by (α0,α¯01)(\alpha_{0},\bar{\alpha}_{0}^{-1}) for any α0\alpha_{0}\in\mathbb{C}, next (Ψk,0+,Ψk,0)(\Psi^{+\prime}_{k,0},\Psi^{-\prime}_{k,0}) is defined up to addition of (α1Ψk,0+,α¯1Ψk,0)(\alpha_{1}\Psi^{+}_{k,0},-\bar{\alpha}_{1}\Psi^{-}_{k,0}) for any α1\alpha_{1}\in\mathbb{C}, and next (Ψk,0+′′,Ψk,0′′)(\Psi^{+\prime\prime}_{k,0},\Psi^{-\prime\prime}_{k,0}) is defined up to addition of (α2Ψk,0+,α¯2Ψk,0)(\alpha_{2}\Psi^{+}_{k,0},-\bar{\alpha}_{2}\Psi^{-}_{k,0}) for any α2\alpha_{2}\in\mathbb{C}.

We first compute zk,0,Ψk,0+,Ψk,0z_{k,0},\Psi_{k,0}^{+},\Psi_{k,0}^{-}. The resonance conjecture Conjecture (LRC) — Local resonance conjecture yields for z>0\Im z>0,

ϕ,(Hk,λstz)1ϕL2(Ω)=Ψk,λ+,ϕ𝒫(Ω),𝒫(Ω)¯Ψk,λ,ϕ𝒫(Ω,𝒫(Ω)zk,λz+ζk,λϕ,ϕ(z),\big{\langle}{\phi^{\prime}},(H_{k,\lambda}^{\operatorname{st}}-z)^{-1}\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}=\frac{\overline{\langle\Psi_{k,\lambda}^{+},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\Psi_{k,\lambda}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega,\mathcal{P}(\Omega)}}{z_{k,\lambda}-z}+\zeta_{k,\lambda}^{\phi^{\prime},\phi}(z), (6.6)

with ζk,λϕ,ϕ\zeta_{k,\lambda}^{\phi^{\prime},\phi} holomorphic on {z:z>0}1MB\{z:\Im z>0\}\bigcup\frac{1}{M}B. Setting λ=0\lambda=0 and ϕ=1\phi^{\prime}=1, we find for z>0\Im z>0,

1z𝔼[ϕ]=𝔼[(Hk,0stz)1ϕ]=Ψk,0+,1𝒫(Ω),𝒫(Ω)¯Ψk,0,ϕ𝒫(Ω,𝒫(Ω)zk,0z+ζk,01,ϕ(z),-\frac{1}{z}\mathbb{E}\left[{\phi}\right]=\mathbb{E}\left[{(H_{k,0}^{\operatorname{st}}-z)^{-1}\phi}\right]=\frac{\overline{\langle\Psi_{k,0}^{+},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\Psi_{k,0}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega,\mathcal{P}(\Omega)}}{z_{k,0}-z}+\zeta_{k,0}^{1,\phi}(z),

and similarly, exchanging the roles of ϕ\phi and ϕ\phi^{\prime},

1z¯𝔼[ϕ]=Ψk,0,1𝒫(Ω),𝒫(Ω)¯Ψk,0+,ϕ𝒫(Ω,𝒫(Ω)z¯k,0z¯+ζk,0ϕ,1(z)¯.-\frac{1}{\bar{z}}\mathbb{E}\left[{\phi}\right]=\frac{\overline{\langle\Psi_{k,0}^{-},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\Psi_{k,0}^{+},\phi\rangle_{\mathcal{P}^{\prime}(\Omega,\mathcal{P}(\Omega)}}{\bar{z}_{k,0}-\bar{z}}+\overline{\zeta_{k,0}^{\phi,1}(z)}.

For z0z\to 0, we deduce

zk,0=0,Ψk,0+=α,Ψk,0=α¯1,for some α.z_{k,0}=0,\qquad\Psi_{k,0}^{+}=\alpha,\qquad\Psi_{k,0}^{-}=\bar{\alpha}^{-1},\qquad\text{for some $\alpha\in\mathbb{C}$}.

By gauge symmetry, as explained above, we can e.g. choose α=1\alpha=1,

zk,0=0,Ψk,0+=Ψk,0=1.z_{k,0}=0,\qquad\Psi_{k,0}^{+}=\Psi_{k,0}^{-}=1. (6.7)

Next, we compute zk,0,Ψk,0+,Ψk,0z_{k,0}^{\prime},\Psi^{+\prime}_{k,0},\Psi^{-\prime}_{k,0}. Differentiating identity (6.6) at λ=0\lambda=0, using (6.7), and choosing ϕ=1\phi^{\prime}=1, we find for z>0\Im z>0,

𝔼[V(Hk,0stz)1ϕ]=1zzk,0𝔼[ϕ]Ψk,0+,1𝒫(Ω),𝒫(Ω)¯𝔼[ϕ]Ψk,0,ϕ𝒫(Ω),𝒫(Ω)+zλζk,λ1,ϕ(z)|λ=0,\mathbb{E}\left[{V(H_{k,0}^{\operatorname{st}}-z)^{-1}\phi}\right]\\ =-\frac{1}{z}z_{k,0}^{\prime}\mathbb{E}\left[{\phi}\right]-\overline{\langle\Psi_{k,0}^{+\prime},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\mathbb{E}\left[{\phi}\right]-\langle\Psi_{k,0}^{-\prime},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}+z\partial_{\lambda}\zeta_{k,\lambda}^{1,\phi}(z)\big{|}_{\lambda=0},

and similarly, exchanging the roles of ϕ\phi and ϕ\phi^{\prime},

𝔼[V(Hk,0stz¯)1ϕ]=1z¯z¯k,0𝔼[ϕ]Ψk,0,1𝒫(Ω),𝒫(Ω)¯𝔼[ϕ]Ψk,0+,ϕ𝒫(Ω),𝒫(Ω)+zλζk,λϕ,1(z)¯|λ=0.\mathbb{E}\left[{V(H_{k,0}^{\operatorname{st}}-\bar{z})^{-1}\phi}\right]\\ =-\frac{1}{\bar{z}}\bar{z}_{k,0}^{\prime}\mathbb{E}\big{[}{\phi}\big{]}-\overline{\langle\Psi_{k,0}^{-\prime},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\mathbb{E}\big{[}{\phi}\big{]}-\langle\Psi_{k,0}^{+\prime},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}+\overline{z\partial_{\lambda}\zeta_{k,\lambda}^{\phi,1}(z)}\big{|}_{\lambda=0}.

Choosing z=iεz=i\varepsilon with ε0\varepsilon\downarrow 0, we easily deduce zk,0=0z_{k,0}^{\prime}=0 and

𝔼[V(Hk,0sti0)1ϕ]=Ψk,0±,1𝒫(Ω),𝒫(Ω)¯𝔼[ϕ]Ψk,0,ϕ𝒫(Ω),𝒫(Ω).\displaystyle\mathbb{E}\left[{V(H_{k,0}^{\operatorname{st}}\mp i0)^{-1}\phi}\right]=-\overline{\langle\Psi_{k,0}^{\pm\prime},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\,\mathbb{E}\left[{\phi}\right]-\langle\Psi_{k,0}^{\mp\prime},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}.

Noting that the left-hand side vanishes for ϕ=1\phi=1, we are led to

Ψk,0,ϕ𝒫(Ω),𝒫(Ω)\displaystyle\langle\Psi_{k,0}^{-\prime},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)} =\displaystyle= 𝔼[V(Hk,0sti0)1ϕ]+β𝔼[ϕ],\displaystyle-\mathbb{E}\big{[}{V(H_{k,0}^{\operatorname{st}}-i0)^{-1}\phi}\big{]}+\beta\mathbb{E}\left[{\phi}\right],
Ψk,0+,ϕ𝒫(Ω),𝒫(Ω)\displaystyle\langle\Psi_{k,0}^{+\prime},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)} =\displaystyle= 𝔼[V(Hk,0st+i0)1ϕ]β¯𝔼[ϕ],\displaystyle-\mathbb{E}\big{[}{V(H_{k,0}^{\operatorname{st}}+i0)^{-1}\phi}\big{]}-\bar{\beta}\mathbb{E}\left[{\phi}\right],

for some β\beta\in\mathbb{C}. By gauge symmetry, as explained above, we can e.g. choose β=0\beta=0,

zk,0=0,Ψk,0±=(Hk,0sti0)1V.z_{k,0}^{\prime}=0,\qquad\Psi_{k,0}^{\pm\prime}=-(H_{k,0}^{\operatorname{st}}\mp i0)^{-1}V. (6.8)

Finally, we turn to the second derivatives zk,0′′,Ψk,0+′′,Ψk,0′′z_{k,0}^{\prime\prime},\Psi_{k,0}^{+\prime\prime},\Psi_{k,0}^{-\prime\prime}. Differentiating identity (6.6) twice at λ=0\lambda=0, using (6.7) and (6.8), and choosing ϕ=1\phi^{\prime}=1, we find for z>0\Im z>0,

2𝔼[V(Hk,0stz)1V(Hk,0stz)1ϕ]=1zzk,0′′𝔼[ϕ]+Ψk,0+′′,1𝒫(Ω),𝒫(Ω)¯𝔼[ϕ]+Ψk,0′′,ϕ𝒫(Ω),𝒫(Ω)zλ2ζk,λ1,ϕ(z)|λ=0,2\,\mathbb{E}\big{[}{V(H_{k,0}^{\operatorname{st}}-z)^{-1}V(H_{k,0}^{\operatorname{st}}-z)^{-1}\phi}\big{]}\\ =\frac{1}{z}z_{k,0}^{\prime\prime}\,\mathbb{E}\left[{\phi}\right]+\overline{\langle\Psi_{k,0}^{+\prime\prime},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\mathbb{E}\left[{\phi}\right]+\langle\Psi_{k,0}^{-\prime\prime},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}-z\partial_{\lambda}^{2}\zeta_{k,\lambda}^{1,\phi}(z)\big{|}_{\lambda=0},

and similarly, exchanging the roles of ϕ\phi and ϕ\phi^{\prime},

2𝔼[V(Hk,0stz¯)1V(Hk,0stz¯)1ϕ]=1z¯z¯k,0′′𝔼[ϕ]+Ψk,0′′,1𝒫(Ω),𝒫(Ω)¯𝔼[ϕ]+Ψk,0+′′,ϕ𝒫(Ω),𝒫(Ω)zλ2ζk,λϕ,1(z)¯|λ=0.2\,\mathbb{E}\big{[}{V(H_{k,0}^{\operatorname{st}}-\bar{z})^{-1}V(H_{k,0}^{\operatorname{st}}-\bar{z})^{-1}\phi}\big{]}\\ =\frac{1}{\bar{z}}\bar{z}_{k,0}^{\prime\prime}\mathbb{E}\big{[}{\phi}\big{]}+\overline{\langle\Psi_{k,0}^{-\prime\prime},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\mathbb{E}\left[{\phi}\right]+\langle\Psi_{k,0}^{+\prime\prime},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}-\overline{z\partial_{\lambda}^{2}\zeta_{k,\lambda}^{\phi,1}(z)}\big{|}_{\lambda=0}.

Choosing z=iεz=i\varepsilon with ε0\varepsilon\downarrow 0, and using again the gauge invariance, we easily deduce

zk,0′′\displaystyle z_{k,0}^{\prime\prime} =\displaystyle= 2𝔼[V(Hk,0sti0)1V],\displaystyle-2\,\mathbb{E}\big{[}{V(H_{k,0}^{\operatorname{st}}-i0)^{-1}V}\big{]},
Ψk,0±′′\displaystyle\Psi_{k,0}^{\pm\prime\prime} =\displaystyle= 2(Hk,0sti0)1ΠV(Hk,0sti0)1V,\displaystyle 2(H_{k,0}^{\operatorname{st}}\mp i0)^{-1}\Pi V(H_{k,0}^{\operatorname{st}}\mp i0)^{-1}V,

in terms of the projection Πϕ:=ϕ𝔼[ϕ]\Pi\phi:=\phi-\mathbb{E}\left[{\phi}\right] onto L2(Ω)\operatorname{L}^{2}(\Omega)\ominus\mathbb{C}. 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 H0=H_{0}=-\triangle by

H~0:=1i1:=1ie1,\widetilde{H}_{0}:=\tfrac{1}{i}\nabla_{1}:=\tfrac{1}{i}e_{1}\cdot\nabla,

and we consider the corresponding perturbed operator

H~λ:=1i1+λV\widetilde{H}_{\lambda}:=\tfrac{1}{i}\nabla_{1}+\lambda V

on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega). Via the Floquet–Bloch fibration (1.3), this operator is decomposed as

(H~λ,L2(d×Ω))=d(H~λst+k1,L2(Ω))𝔢k¯𝑑k,𝔢k(x):=eikx,\big{(}\widetilde{H}_{\lambda},\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega)\big{)}\,=\,\int_{\mathbb{R}^{d}}^{\oplus}\big{(}\widetilde{H}_{\lambda}^{\operatorname{st}}+k_{1},\operatorname{L}^{2}(\Omega)\big{)}\,\mathfrak{e}_{k}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,\qquad\mathfrak{e}_{k}(x):=e^{ik\cdot x}, (7.1)

in terms of the following (centered) fibered operator on the stationary space L2(Ω)\operatorname{L}^{2}(\Omega),

H~λst:=H~0st+λV,H~0st:=1i1st:=1ie1st.\widetilde{H}_{\lambda}^{\operatorname{st}}\,:=\,\widetilde{H}_{0}^{\operatorname{st}}+\lambda V,\qquad\widetilde{H}_{0}^{\operatorname{st}}\,:=\,\tfrac{1}{i}\nabla^{\operatorname{st}}_{1}\,:=\,\tfrac{1}{i}e_{1}\cdot\nabla^{\operatorname{st}}.

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 H~λst\widetilde{H}_{\lambda}^{\operatorname{st}} in this toy model is not deformed under the fibration parameter kk.

Theorem 7.1 (Toy model).

Assume for simplicity that V=V0V=V_{0} is a stationary Gaussian field on with covariance 𝒞0Cc(d)\mathcal{C}_{0}\in C^{\infty}_{c}(\mathbb{R}^{d}).

  1. (i)

    Spectral decomposition of H~0st\widetilde{H}_{0}^{\operatorname{st}}:
    The eigenvalue at 0 is simple (with eigenspace \mathbb{C}) and

    σpp(H~0st)={0},σsc(H~0st)=,σ(H~0st)=σac(H~0st)=.\qquad\sigma_{\operatorname{pp}}(\widetilde{H}_{0}^{\operatorname{st}})\,=\,\{0\},\qquad\sigma_{\operatorname{sc}}(\widetilde{H}_{0}^{\operatorname{st}})\,=\,\varnothing,\qquad\sigma(\widetilde{H}_{0}^{\operatorname{st}})\,=\,\sigma_{\operatorname{ac}}(\widetilde{H}_{0}^{\operatorname{st}})\,=\,\mathbb{R}.
  2. (ii)

    Spectral decomposition of H~λst\widetilde{H}_{\lambda}^{\operatorname{st}}:
    For λ>0\lambda>0, the eigenvalue at 0 is fully absorbed in the absolutely continuous spectrum,

    σpp(H~λst)=σsc(H~λst)=,σ(H~λst)=σac(H~λst)=.\qquad\sigma_{\operatorname{pp}}(\widetilde{H}_{\lambda}^{\operatorname{st}})\,=\,\sigma_{\operatorname{sc}}(\widetilde{H}_{\lambda}^{\operatorname{st}})\,=\,\varnothing,\qquad\sigma(\widetilde{H}_{\lambda}^{\operatorname{st}})\,=\,\sigma_{\operatorname{ac}}(\widetilde{H}_{\lambda}^{\operatorname{st}})\,=\,\mathbb{R}.
  3. (iii)

    Fibered resonances:
    For all λ>0\lambda>0, the resolvent z(H~λstz)1z\mapsto(\widetilde{H}^{\operatorname{st}}_{\lambda}-z)^{-1} defined on z>0\Im z>0 as a family of operators 𝒫(Ω)𝒫(Ω)\mathcal{P}(\Omega)\to\mathcal{P}^{\prime}(\Omega) can be extended meromorphically to the whole complex plane with a unique simple pole at

    z=iλ2α,α:=0𝒞0(se1)𝑑s.\qquad z=-i\lambda^{2}\alpha_{\circ},\qquad\alpha_{\circ}:=\int_{0}^{\infty}\mathcal{C}_{0}(se_{1})\,ds.

    In other words, for all ϕ,ϕ𝒫(Ω)\phi,\phi^{\prime}\in\mathcal{P}(\Omega), we can write for z>0\Im z>0,

    𝔼[ϕ¯(H~λstz)1ϕ]=Ψ~λ+,ϕ𝒫(Ω),𝒫(Ω)¯Ψ~λ,ϕ𝒫(Ω),𝒫(Ω)iλ2αz+ζ~λϕ,ϕ(z),\qquad\mathbb{E}\big{[}{\bar{\phi}^{\prime}(\widetilde{H}^{\operatorname{st}}_{\lambda}-z)^{-1}\phi}\big{]}\,=\,\frac{\overline{\langle\widetilde{\Psi}_{\lambda}^{+},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\widetilde{\Psi}_{\lambda}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}{-i\lambda^{2}\alpha_{\circ}-z}+\widetilde{\zeta}_{\lambda}^{\phi^{\prime},\phi}(z),

    where the remainder ζ~λϕ,ϕ\widetilde{\zeta}_{\lambda}^{\phi^{\prime},\phi} is entire, has a continuous dependence on λ0\lambda\geq 0, and satisfies the uniform bound

    |ζ~λϕ,ϕ(z)|ϕ,ϕ{1,if neither ϕ nor ϕ is constant;λ,if ϕ or ϕ is constant;λ2,if ϕ and ϕ are constant;\qquad|\widetilde{\zeta}_{\lambda}^{\phi^{\prime},\phi}(z)|\,\lesssim_{\phi,\phi^{\prime}}\,\begin{cases}1,&\text{if neither $\phi$ nor $\phi^{\prime}$ is constant};\\ \lambda,&\text{if $\phi$ or $\phi^{\prime}$ is constant};\\ \lambda^{2},&\text{if $\phi$ and $\phi^{\prime}$ are constant};\end{cases}

    and where the so-called resonant and co-resonant states Ψ~λ+,Ψ~λ𝒫(Ω)\widetilde{\Psi}_{\lambda}^{+},\widetilde{\Psi}_{\lambda}^{-}\in\mathcal{P}^{\prime}(\Omega) are explicitly defined, cf. Remark 7.2(a) below.

  4. (iv)

    Continuous resonant spectrum:
    For λ>0\lambda>0 small enough, the resolvent z(H~λz)1z\mapsto(\widetilde{H}_{\lambda}-z)^{-1} defined on z>0\Im z>0 as a family of operators L2(d)𝒫(Ω)L2(d)𝒫(Ω)\operatorname{L}^{2}(\mathbb{R}^{d})\otimes\mathcal{P}(\Omega)\to\operatorname{L}^{2}(\mathbb{R}^{d})\otimes\mathcal{P}^{\prime}(\Omega) can be extended holomorphically to z>λ2α\Im z>-\lambda^{2}\alpha_{\circ}, and we denote the extension by Rλ(z)R_{\lambda}(z). For ϕ𝒫(Ω)\phi\in\mathcal{P}(\Omega), this extension has the following discontinuity, as zλ2α\Im z\downarrow-\lambda^{2}\alpha_{\circ},

    supgL2(d)=1ϕg,Rλ(z)ϕgL2(d×Ω)iΨ~λ+,ϕ𝒫(Ω),𝒫(Ω)¯Ψ~λ,ϕ𝒫(Ω),𝒫(Ω)λ2α+z.\displaystyle\qquad\sup_{\|g\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}=1}\big{\langle}\phi g,R_{\lambda}(z)\phi g\big{\rangle}_{\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega)}\,\sim\,i\frac{\overline{\langle\widetilde{\Psi}_{\lambda}^{+},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\widetilde{\Psi}_{\lambda}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}{\lambda^{2}\alpha_{\circ}+\Im z}.

    Next, viewed as a family of operators Lcomp2(d)𝒫(Ω)Lloc2(d)𝒫(Ω)\operatorname{L}^{2}_{\operatorname{comp}}(\mathbb{R}^{d})\otimes\mathcal{P}(\Omega)\to\operatorname{L}^{2}_{\operatorname{loc}}(\mathbb{R}^{d})\otimes\mathcal{P}^{\prime}(\Omega), the extended resolvent RλR_{\lambda} can be further extended to all \mathbb{C} as an entire function.

  5. (v)

    Ballistic transport:
    For all uL2(d)u^{\circ}\in\operatorname{L}^{2}(\mathbb{R}^{d}) with xuL2(d)<\|xu^{\circ}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}<\infty, there holds

    limt1t𝔼[xuλtL2(d)2]12=uL2(d).\qquad\lim_{t\uparrow\infty}\frac{1}{t}\,\mathbb{E}\left[{\|xu_{\lambda}^{t}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}^{2}}\right]^{\frac{1}{2}}\,=\,\|u^{\circ}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}.\qed
Remarks 7.2.

  1. (a)

    Explicit formula for resonant state:
    Up to a gauge transformation, the resonant and co-resonant states Ψ~λ+,Ψ~λ𝒫(Ω)\widetilde{\Psi}_{\lambda}^{+},\widetilde{\Psi}_{\lambda}^{-}\in\mathcal{P}^{\prime}(\Omega) in item (iii) take the form

    Ψ~λ±=e12λ20s𝒞0(se1)𝑑sΨ~λ,±,\widetilde{\Psi}_{\lambda}^{\pm}\,=\,e^{\frac{1}{2}\lambda^{2}\int_{0}^{\infty}s\,\mathcal{C}_{0}(se_{1})\,ds}\,\widetilde{\Psi}_{\lambda}^{\circ,\pm},

    where Ψ~λ,±𝒫(Ω)\widetilde{\Psi}_{\lambda}^{\circ,\pm}\in\mathcal{P}^{\prime}(\Omega) is formally defined as

    Ψ~λ,±=e±λi0V(se1,)𝑑s𝔼[e±λi0V(se1,)𝑑s].\widetilde{\Psi}_{\lambda}^{\circ,\pm}\leavevmode\nobreak\ \leavevmode\nobreak\ \text{``$=$''}\leavevmode\nobreak\ \leavevmode\nobreak\ \frac{e^{\pm\frac{\lambda}{i}\int_{0}^{\infty}V(\mp se_{1},\cdot)\,ds}}{\mathbb{E}\big{[}{e^{\pm\frac{\lambda}{i}\int_{0}^{\infty}V(\mp se_{1},\cdot)\,ds}}\big{]}}. (7.2)

    More precisely, the action of Ψ~λ,±\widetilde{\Psi}_{\lambda}^{\circ,\pm} on 𝒫(Ω)\mathcal{P}(\Omega) is defined inductively on monomials of increasing degree: we set Ψ~λ,±,1𝒫(Ω),𝒫(Ω)=1\langle\widetilde{\Psi}_{\lambda}^{\circ,\pm},1\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}=1, and for all n1n\geq 1 and x1,,xndx_{1},\ldots,x_{n}\in\mathbb{R}^{d},

    Ψ~λ,±,j=1nV(xj,)𝒫(Ω),𝒫(Ω)=l=2n𝒞0(x1xl)Ψ~λ,±,2jnjlV(xj,)𝒫(Ω),𝒫(Ω)λi(0𝒞0(x1±se1)𝑑s)Ψ~λ,±,j=2nV(xj,)𝒫(Ω),𝒫(Ω),\qquad\bigg{\langle}\widetilde{\Psi}_{\lambda}^{\circ,\pm},\prod_{j=1}^{n}V(x_{j},\cdot)\bigg{\rangle}_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}\,=\,\sum_{l=2}^{n}\mathcal{C}_{0}(x_{1}-x_{l})\,\bigg{\langle}\widetilde{\Psi}_{\lambda}^{\circ,\pm},\prod_{2\leq j\leq n\atop j\neq l}V(x_{j},\cdot)\bigg{\rangle}_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}\\ \mp\frac{\lambda}{i}\Big{(}\int_{0}^{\infty}\mathcal{C}_{0}(x_{1}\pm se_{1})\,ds\Big{)}\,\bigg{\langle}\widetilde{\Psi}_{\lambda}^{\circ,\pm},\prod_{j=2}^{n}V(x_{j},\cdot)\bigg{\rangle}_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}, (7.3)

    while the formal representation (7.2) is understood in view of Wick’s theorem. In particular, there holds Ψ~λ±,ϕ𝒫(Ω),𝒫(Ω)=𝔼[ϕ]±iλ0𝔼[V(se1,)ϕ]𝑑s+Oϕ(λ2)\langle\widetilde{\Psi}_{\lambda}^{\pm},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}=\mathbb{E}\left[{\phi}\right]\pm i\lambda\int_{0}^{\infty}\mathbb{E}\left[{V(\mp se_{1},\cdot)\,\phi}\right]ds+O_{\phi}(\lambda^{2}).

  2. (b)

    Mourre’s approach:
    In view of Proposition 5.7(ii)–(iii), with the notation of Section 5.3, the commutator [H~λ,1iOp(x1)][\widetilde{H}_{\lambda},\tfrac{1}{i}\operatorname{Op}(x_{1})] is \mathcal{L}-bounded and satisfies the lower bound

    [H~λst,1iOp(x1)]Cλ12.[\widetilde{H}_{\lambda}^{\operatorname{st}},\tfrac{1}{i}\operatorname{Op}(x_{1})]\,\geq\,\mathcal{L}-C\lambda\mathcal{L}^{\frac{1}{2}}.

    In other words, H~λst\tilde{H}_{\lambda}^{\operatorname{st}} satisfies a Mourre relation with conjugate Op(x1)\operatorname{Op}(x_{1}) “up to \mathcal{L}”. 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 H~λst=1i1+λV\widetilde{H}_{\lambda}^{\operatorname{st}}=\frac{1}{i}\nabla_{1}+\lambda V on 2(d)\ell^{2}(\mathbb{Z}^{d}) with an i.i.d. Gaussian field VV on d\mathbb{Z}^{d}: 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 𝒞0\mathcal{C}_{0} is supported in BB.

Step 1. Proof of (iii).
It suffices to show for all ϕ,ϕ𝒫(Ω)\phi,\phi^{\prime}\in\mathcal{P}(\Omega) and t0t\geq 0,

|𝔼[ϕ¯eitH~λstϕ]Ψ~λ+,ϕ𝒫(Ω),𝒫(Ω)¯Ψ~λ,ϕ𝒫(Ω),𝒫(Ω)etλ2α|ϕ,ϕ 1|t|Cϕ,ϕ,\displaystyle\Big{|}\mathbb{E}\big{[}{\bar{\phi}^{\prime}\,e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi}\big{]}-\overline{\langle\widetilde{\Psi}_{\lambda}^{+},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\widetilde{\Psi}_{\lambda}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}\,e^{-t\lambda^{2}\alpha_{\circ}}\Big{|}\,\lesssim_{\phi,\phi^{\prime}}\,\mathds{1}_{|t|\leq C_{\phi,\phi^{\prime}}}, (7.4)

where we gain a factor λ\lambda in the right-hand side if ϕ\phi or ϕ\phi^{\prime} is constant, and a factor λ2\lambda^{2} if both are constants. Indeed, for z>0\Im z>0, we can write

𝔼[ϕ¯(H~λstz)1ϕ]=i0eitz𝔼[ϕ¯eitH~λstϕ]𝑑t,\mathbb{E}\big{[}{\bar{\phi}^{\prime}(\widetilde{H}_{\lambda}^{\operatorname{st}}-z)^{-1}\phi}\big{]}\,=\,i\int_{0}^{\infty}e^{itz}\,\mathbb{E}\big{[}{\bar{\phi}^{\prime}\,e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi}\big{]}\,dt,

so that the conclusion (iii) would follow from (7.4) after integration. By linearity, it suffices to establish (7.4) for monomials

ϕ=j=1nV(xj,),ϕ=j=1mV(yj,).\phi=\prod_{j=1}^{n}V(x_{j},\cdot),\qquad\phi^{\prime}=\prod_{j=1}^{m}V(y_{j},\cdot).

Noting that the fibered evolution ϕλt:=eitH~λstϕ\phi_{\lambda}^{t}:=e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi can be explicitly computed as

ϕλt(ω)=ϕ(te1,ω)ψλt(ω),ψλt(ω):=exp(λi0tV(se1,ω)𝑑s),\phi_{\lambda}^{t}(\omega)\,=\,\phi(-te_{1},\omega)\,\psi_{\lambda}^{t}(\omega),\qquad\psi_{\lambda}^{t}(\omega)\,:=\,\exp\Big{(}\frac{\lambda}{i}\int_{0}^{t}V(-se_{1},\omega)\,ds\Big{)},

we find

𝔼[ϕeitH~λstϕ]=𝔼[(j=1mV(yj,))(j=1nV(xjte1,))ψλt].\mathbb{E}\big{[}{\phi^{\prime}\,e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi}\big{]}=\mathbb{E}\bigg{[}{\Big{(}\prod_{j=1}^{m}V(y_{j},\cdot)\Big{)}\Big{(}\prod_{j=1}^{n}V(x_{j}-te_{1},\cdot)\Big{)}\,\psi_{\lambda}^{t}}\bigg{]}.

By Wick’s formula, for m1m\geq 1, we compute

𝔼[ϕeitH~λstϕ]=l=2m𝒞0(y1yl)𝔼[(2jmjlV(yj,))(j=1nV(xjte1,))ψλt]+l=2n𝒞0(y1xl+te1)𝔼[(j=2mV(yj,))(1jnjlV(xjte1,))ψλt]+λi(0t𝒞0(y1+se1)𝑑s)𝔼[(j=2mV(yj,))(j=1nV(xjte1,))ψλt].\mathbb{E}\big{[}{\phi^{\prime}\,e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi}\big{]}=\sum_{l=2}^{m}\mathcal{C}_{0}(y_{1}-y_{l})\,\mathbb{E}\bigg{[}{\Big{(}\prod_{2\leq j\leq m\atop j\neq l}V(y_{j},\cdot)\Big{)}\Big{(}\prod_{j=1}^{n}V(x_{j}-te_{1},\cdot)\Big{)}\,\psi_{\lambda}^{t}}\bigg{]}\\ +\sum_{l=2}^{n}\mathcal{C}_{0}(y_{1}-x_{l}+te_{1})\,\mathbb{E}\bigg{[}{\Big{(}\prod_{j=2}^{m}V(y_{j},\cdot)\Big{)}\Big{(}\prod_{1\leq j\leq n\atop j\neq l}V(x_{j}-te_{1},\cdot)\Big{)}\,\psi_{\lambda}^{t}}\bigg{]}\\ +\frac{\lambda}{i}\Big{(}\int_{0}^{t}\mathcal{C}_{0}(y_{1}+se_{1})\,ds\Big{)}\,\mathbb{E}\bigg{[}{\Big{(}\prod_{j=2}^{m}V(y_{j},\cdot)\Big{)}\Big{(}\prod_{j=1}^{n}V(x_{j}-te_{1},\cdot)\Big{)}\,\psi_{\lambda}^{t}}\bigg{]}.

Since by assumption

|𝒞0(y1xl+te1)|𝟙|t|1+|y1|+|xl|,|\mathcal{C}_{0}(y_{1}-x_{l}+te_{1})|\lesssim\mathds{1}_{|t|\leq 1+|y_{1}|+|x_{l}|},

and similarly |t𝒞0(y1+se1)𝑑s|𝟙|t|1+|y1|\big{|}\int_{t}^{\infty}\mathcal{C}_{0}(y_{1}+se_{1})\,ds\big{|}\lesssim\mathds{1}_{|t|\leq 1+|y_{1}|}, we deduce

𝔼[ϕeitH~λstϕ]=l=2m𝒞0(y1yl)𝔼[(2jmjlV(yj,))(j=1nV(xjte1,))ψλt]+λi(0𝒞0(y1+se1)𝑑s)𝔼[(j=2mV(yj,))(j=1nV(xjte1,))ψλt]+Oϕ,ϕ(𝟙|t|Cϕ,ϕ).\mathbb{E}\big{[}{\phi^{\prime}\,e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi}\big{]}=\sum_{l=2}^{m}\mathcal{C}_{0}(y_{1}-y_{l})\,\mathbb{E}\bigg{[}{\Big{(}\prod_{2\leq j\leq m\atop j\neq l}V(y_{j},\cdot)\Big{)}\Big{(}\prod_{j=1}^{n}V(x_{j}-te_{1},\cdot)\Big{)}\,\psi_{\lambda}^{t}}\bigg{]}\\ +\frac{\lambda}{i}\Big{(}\int_{0}^{\infty}\mathcal{C}_{0}(y_{1}+se_{1})\,ds\Big{)}\,\mathbb{E}\bigg{[}{\Big{(}\prod_{j=2}^{m}V(y_{j},\cdot)\Big{)}\Big{(}\prod_{j=1}^{n}V(x_{j}-te_{1},\cdot)\Big{)}\,\psi_{\lambda}^{t}}\bigg{]}\,+\,O_{\phi,\phi^{\prime}}(\mathds{1}_{|t|\leq C_{\phi,\phi^{\prime}}}).

We recognize here the inductive definition (7.3) of the resonant state Ψ~λ+\widetilde{\Psi}_{\lambda}^{+}, so that the above becomes

𝔼[ϕeitH~λstϕ]=Ψ~λ,+,ϕ𝒫(Ω),𝒫(Ω)¯𝔼[(j=1nV(xjte1,))ψλt]+Oϕ,ϕ(𝟙|t|Cϕ,ϕ),\mathbb{E}\big{[}{\phi^{\prime}\,e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi}\big{]}=\overline{\langle\widetilde{\Psi}_{\lambda,\circ}^{+},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\,\mathbb{E}\bigg{[}{\Big{(}\prod_{j=1}^{n}V(x_{j}-te_{1},\cdot)\Big{)}\,\psi_{\lambda}^{t}}\bigg{]}+\,O_{\phi,\phi^{\prime}}(\mathds{1}_{|t|\leq C_{\phi,\phi^{\prime}}}),

and a similar computation yields

𝔼[ϕeitH~λstϕ]=Ψ~λ,+,ϕ𝒫(Ω),𝒫(Ω)¯Ψ~λ,,ϕ𝒫(Ω),𝒫(Ω)𝔼[ψλt]+Oϕ,ϕ(𝟙|t|Cϕ,ϕ).\mathbb{E}\big{[}{\phi^{\prime}\,e^{-it\widetilde{H}_{\lambda}^{\operatorname{st}}}\phi}\big{]}=\overline{\langle\widetilde{\Psi}_{\lambda,\circ}^{+},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\widetilde{\Psi}_{\lambda,\circ}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}\,\mathbb{E}\left[{\psi_{\lambda}^{t}}\right]+\,O_{\phi,\phi^{\prime}}(\mathds{1}_{|t|\leq C_{\phi,\phi^{\prime}}}).

Finally, since 0tV(se1,)𝑑s\int_{0}^{t}V(-se_{1},\cdot)\,ds is Gaussian, we compute

𝔼[ψλt]=e12λ2Var[0tV(se1,)𝑑s]\displaystyle\mathbb{E}\big{[}{\psi_{\lambda}^{t}}\big{]}\,=\,e^{-\frac{1}{2}\lambda^{2}\operatorname{Var}\left[{\int_{0}^{t}V(-se_{1},\cdot)ds}\right]} =\displaystyle= eλ20t(ts)𝒞0(se1)𝑑s\displaystyle e^{-\lambda^{2}\int_{0}^{t}(t-s)\mathcal{C}_{0}(se_{1})ds}
=\displaystyle= eλ2t0𝒞0(se1)𝑑seλ20s𝒞0(se1)𝑑s+OM(λ2𝟙|t|1),\displaystyle e^{-\lambda^{2}t\int_{0}^{\infty}\mathcal{C}_{0}(se_{1})ds}e^{\lambda^{2}\int_{0}^{\infty}s\,\mathcal{C}_{0}(se_{1})ds}+O_{M}(\lambda^{2}\mathds{1}_{|t|\leq 1}),

and the conclusion (7.4) follows.

Step 2. Proof of (iv).
For g,gL2(d)g,g^{\prime}\in\operatorname{L}^{2}(\mathbb{R}^{d}) and ϕ,ϕ𝒫(Ω)\phi,\phi^{\prime}\in\mathcal{P}(\Omega), the Floquet–Bloch fibration (7.1) yields

ϕg,(H~λz)1(ϕg)L2(d×Ω)=dg^(k)¯g^(k)ϕ,(H~λst+k1z)1ϕL2(Ω)¯𝑑k,\big{\langle}\phi^{\prime}g^{\prime},(\widetilde{H}_{\lambda}-z)^{-1}(\phi g)\big{\rangle}_{\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega)}\,=\,\int_{\mathbb{R}^{d}}\overline{\widehat{g}^{\prime}(k)}\widehat{g}(k)\,\big{\langle}\phi^{\prime},(\widetilde{H}_{\lambda}^{\operatorname{st}}+k_{1}-z)^{-1}\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\,\,{{\mathchar 22\relax\mkern-12.0mud}}k,

and thus, inserting the result of item (iii), for z>0\Im z>0,

ϕg,(H~λz)1(ϕg)L2(d×Ω)=dg^(k)¯g^(k)ζ~λϕ,ϕ(zk1)¯𝑑k+Ψ~λ+,ϕ𝒫(Ω),𝒫(Ω)¯Ψ~λ,ϕ𝒫(Ω),𝒫(Ω)dg^(k)¯g^(k)k1iλ2αz¯𝑑k,\big{\langle}\phi^{\prime}g^{\prime},(\widetilde{H}_{\lambda}-z)^{-1}(\phi g)\big{\rangle}_{\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega)}\,=\,\int_{\mathbb{R}^{d}}\overline{\widehat{g}^{\prime}(k)}\widehat{g}(k)\,\widetilde{\zeta}_{\lambda}^{\phi^{\prime},\phi}(z-k_{1})\,\,{{\mathchar 22\relax\mkern-12.0mud}}k\\ +\overline{\langle\widetilde{\Psi}_{\lambda}^{+},\phi^{\prime}\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}}\langle\widetilde{\Psi}_{\lambda}^{-},\phi\rangle_{\mathcal{P}^{\prime}(\Omega),\mathcal{P}(\Omega)}\int_{\mathbb{R}^{d}}\frac{\overline{\widehat{g}^{\prime}(k)}\widehat{g}(k)}{k_{1}-i\lambda^{2}\alpha_{\circ}-z}\,{{\mathchar 22\relax\mkern-12.0mud}}k,

where the first right-hand side term is entire and the second is holomorphic on z>λ2α\Im z>-\lambda^{2}\alpha_{\circ}. Next, for g,gLcomp2(d)g,g^{\prime}\in\operatorname{L}^{2}_{\operatorname{comp}}(\mathbb{R}^{d}), the Fourier transforms g^,g^\widehat{g},\widehat{g}^{\prime} are entire functions, which allows to extend the second right-hand side term holomorphically to the whole complex plane. Indeed, for yy\in\mathbb{R}, the Sokhotski–Plemelj formula yields

limε0dg^(k)¯g^(k)k1yiε¯𝑑k=p.v.dg^(k)¯g^(k)k1y¯𝑑k±iπd1g^(y,k)¯g^(y,k)¯𝑑k,\lim_{\varepsilon\downarrow 0}\int_{\mathbb{R}^{d}}\frac{\overline{\widehat{g}^{\prime}(k)}\widehat{g}(k)}{k_{1}-y\mp i\varepsilon}\,{{\mathchar 22\relax\mkern-12.0mud}}k\,=\,\operatorname{p.v.}\int_{\mathbb{R}^{d}}\frac{\overline{\widehat{g}^{\prime}(k)}\widehat{g}(k)}{k_{1}-y}\,{{\mathchar 22\relax\mkern-12.0mud}}k\pm i\pi\int_{\mathbb{R}^{d-1}}\overline{\widehat{g}^{\prime}(y,k^{\prime})}\widehat{g}(y,k^{\prime})\,\,{{\mathchar 22\relax\mkern-12.0mud}}k^{\prime},

so that the function Tλ(z)T_{\lambda}(z) defined for z>λ2α\Im z>-\lambda^{2}\alpha_{\circ} by

Tλ(z)=dg^(k)¯g^(k)k1iλ2αz¯𝑑k,T_{\lambda}(z)=\int_{\mathbb{R}^{d}}\frac{\overline{\widehat{g}^{\prime}(k)}\widehat{g}(k)}{k_{1}-i\lambda^{2}\alpha_{\circ}-z}\,{{\mathchar 22\relax\mkern-12.0mud}}k,

and defined for z<λ2α\Im z<-\lambda^{2}\alpha_{\circ} by

Tλ(z)=dg^(k)¯g^(k)k1iλ2αz¯𝑑k+2iπd1g^(z+iλ2α,k)¯g^(z+iλ2α,k)¯𝑑k,T_{\lambda}(z)=\int_{\mathbb{R}^{d}}\frac{\overline{\widehat{g}^{\prime}(k)}\widehat{g}(k)}{k_{1}-i\lambda^{2}\alpha_{\circ}-z}\,{{\mathchar 22\relax\mkern-12.0mud}}k+2i\pi\int_{\mathbb{R}^{d-1}}\overline{\widehat{g}^{\prime}(z+i\lambda^{2}\alpha_{\circ},k^{\prime})}\widehat{g}(z+i\lambda^{2}\alpha_{\circ},k^{\prime})\,\,{{\mathchar 22\relax\mkern-12.0mud}}k^{\prime},

is entire. This proves (iv).

Step 3. Proof of (v).
The flow can be explicitly integrated,

uλt(x)=u(xte1)eiλ0tV(xse1)𝑑s,u_{\lambda}^{t}(x)=u^{\circ}(x-te_{1})\,e^{-i\lambda\int_{0}^{t}V(x-se_{1})ds},

and is seen to satisfy ballistic transport,

1t𝔼[xuλtL2(d)2]12=1t𝔼[d|x|2|u(xte1)|2𝑑x]12tuL2(d).\frac{1}{t}\,\mathbb{E}\left[{\|xu_{\lambda}^{t}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}^{2}}\right]^{\frac{1}{2}}\,=\,\frac{1}{t}\,\mathbb{E}\left[{\int_{\mathbb{R}^{d}}|x|^{2}|u^{\circ}(x-te_{1})|^{2}dx}\right]^{\frac{1}{2}}\,\xrightarrow{t\uparrow\infty}\,\|u^{\circ}\|_{\operatorname{L}^{2}(\mathbb{R}^{d})}.\qed

Appendix A Self-adjointness with unbounded potentials

For a bounded random potential VV, the Schrödinger operator Hλ=+λVH_{\lambda}=-\triangle+\lambda V on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega) is clearly self-adjoint on L2(Ω;H2(d))\operatorname{L}^{2}(\Omega;H^{2}(\mathbb{R}^{d})) and the fibered operators {Hk,λst}k\{H_{k,\lambda}^{\operatorname{st}}\}_{k} are self-adjoint on H2(Ω)H^{2}(\Omega) just as for λ=0\lambda=0. 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 VL2(Ω)V\in\operatorname{L}^{2}(\Omega) has negative part VLp(Ω)V_{-}\in\operatorname{L}^{p}(\Omega) for some p>d2p>\frac{d}{2}. This condition is essentially optimal and applies in particular to the case when V=V0V=V_{0} 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 L=L=\mathcal{L}.)

A random potential VL2(Ω)V\in\operatorname{L}^{2}(\Omega) defines (densely) a multiplicative operator on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega). If VV is not uniformly bounded, this operator is unbounded, so that the self-adjointness of +λV-\triangle+\lambda V is a subtle question and may fail, cf. [27]. Whenever realizations Vω=V(,ω)V_{\omega}=V(\cdot,\omega) are quadratically controlled from below, in the sense of V(x,ω)M(ω)(1+|x|2)V(x,\omega)\geq-M(\omega)\,(1+|x|^{2}) for some random variable ML2(Ω)M\in\operatorname{L}^{2}(\Omega), the Faris-Lavine argument [19] ensures that the Schrödinger operator +λV-\triangle+\lambda V on L2(d×Ω)\operatorname{L}^{2}(\mathbb{R}^{d}\times\Omega) is essentially self-adjoint on Cc(d;L(Ω))C^{\infty}_{c}(\mathbb{R}^{d};\operatorname{L}^{\infty}(\Omega)). By a Borel-Cantelli argument, the quadratic lower bound holds whenever the negative part VV_{-} belongs to Lp(Ω)\operatorname{L}^{p}(\Omega) for some p>d2p>\frac{d}{2} and satisfies more precisely sup|x|1V(x,)Lp(Ω)\sup_{|x|\leq 1}V_{-}(x,\cdot)\in\operatorname{L}^{p}(\Omega). In this setting, since +λV-\triangle+\lambda V is essentially self-adjoint on Cc(d;L(Ω))C^{\infty}_{c}(\mathbb{R}^{d};\operatorname{L}^{\infty}(\Omega)), we may repeat the direct integral decomposition (3.4) and the fibered operators Hk,λstH^{\operatorname{st}}_{k,\lambda} on L2(Ω)\operatorname{L}^{2}(\Omega) are necessarily essentially self-adjoint on H2L(Ω)H^{2}\cap\operatorname{L}^{\infty}(\Omega) for almost all kdk\in\mathbb{R}^{d}, e.g. [48, p.280]. In order to conclude for all kk, some continuity would be needed, which typically requires smoothness of VV. 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 L2(Ω)\operatorname{L}^{2}(\Omega), we draw inspiration from an earlier work by Kato [27].

Theorem A.1 (Essential self-adjointness).

Assume that the potential VL2(Ω)V\in\operatorname{L}^{2}(\Omega) satisfies sup|x|1V(τx)Lp(Ω)\sup_{|x|\leq 1}V_{-}(\tau_{x}\cdot)\in\operatorname{L}^{p}(\Omega) for some p>d2p>\frac{d}{2}. Then for all λ0\lambda\geq 0 and kdk\in\mathbb{R}^{d} the operator Hk,λstH_{k,\lambda}^{\operatorname{st}} is essentially self-adjoint on H2L(Ω)H^{2}\cap\operatorname{L}^{\infty}(\Omega). ∎

Proof.

Let kdk\in\mathbb{R}^{d} and λ0\lambda\geq 0 be fixed. For VL2(Ω)V\in\operatorname{L}^{2}(\Omega), we note that the operator Hk,λstH_{k,\lambda}^{\operatorname{st}} is well-defined on the whole of L2(Ω)\operatorname{L}^{2}(\Omega) with values in the space 𝒟:=L1(Ω)+H2(Ω)\mathcal{D}^{\prime}:=\operatorname{L}^{1}(\Omega)+H^{-2}(\Omega) (cf. Section 3.2 for notation), and it is obviously continuous L2(Ω)𝒟\operatorname{L}^{2}(\Omega)\to\mathcal{D}^{\prime}. Let H˙k,λst\dot{H}_{k,\lambda}^{\operatorname{st}} denote the restriction of Hk,λstH_{k,\lambda}^{\operatorname{st}} with domain 𝒟:=H2L(Ω)\mathcal{D}:=H^{2}\cap\operatorname{L}^{\infty}(\Omega). Since Hk,λst𝒟L2(Ω)H_{k,\lambda}^{\operatorname{st}}\mathcal{D}\subset\operatorname{L}^{2}(\Omega), the operator H˙k,λst\dot{H}_{k,\lambda}^{\operatorname{st}} can be viewed as a densely defined operator on L2(Ω)\operatorname{L}^{2}(\Omega) and it is clearly symmetric. Its adjoint (H˙k,λst)(\dot{H}_{k,\lambda}^{\operatorname{st}})^{*} is easily seen as the restriction of Hk,λstH_{k,\lambda}^{\operatorname{st}} to L2(Ω)\operatorname{L}^{2}(\Omega), that is, defined whenever uL2(Ω)u\in\operatorname{L}^{2}(\Omega) and Hk,λstL2(Ω)H_{k,\lambda}^{\operatorname{st}}\in\operatorname{L}^{2}(\Omega). In this context, the following conditions are equivalent:

  1. (E1)

    H˙k,λst\dot{H}_{k,\lambda}^{\operatorname{st}} is essentially self-adjoint.

  2. (E2)

    There exist two complex numbers z±z_{\pm} with z±0\Im z_{\pm}\gtrless 0 such that Hk,λst+z±H_{k,\lambda}^{\operatorname{st}}+z_{\pm} is an injection of L2(Ω)\operatorname{L}^{2}(\Omega) into 𝒟\mathcal{D}^{\prime}.

  3. (E3)

    The restriction of Hk,λstH_{k,\lambda}^{\operatorname{st}} to L2(Ω)\operatorname{L}^{2}(\Omega) is the strong closure of H˙k,λst\dot{H}_{k,\lambda}^{\operatorname{st}}, that is, for all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega) with Hk,λstϕL2(Ω)H_{k,\lambda}^{\operatorname{st}}\phi\in\operatorname{L}^{2}(\Omega) there exists a sequence (ϕn)n𝒟(\phi_{n})_{n}\subset\mathcal{D} such that ϕnϕ\phi_{n}\to\phi and Hk,λstϕnHk,λstϕH_{k,\lambda}^{\operatorname{st}}\phi_{n}\to H_{k,\lambda}^{\operatorname{st}}\phi in L2(Ω)\operatorname{L}^{2}(\Omega).

We proceed by truncation: we define the truncated operator Hk,λ;Rst:=Hk,0st+λV𝟙VRH_{k,\lambda;R}^{\operatorname{st}}:=H_{k,0}^{\operatorname{st}}+\lambda V\mathds{1}_{V\geq-R} for R1R\geq 1, and we split the proof into four steps.

Step 1. Proof that Hk,λ;RstH_{k,\lambda;R}^{\operatorname{st}} is essentially self-adjoint.
Assume that ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega) satisfies (Hk,λ;Rst+z)ϕ=0(H_{k,\lambda;R}^{\operatorname{st}}+z)\phi=0 in 𝒟\mathcal{D}^{\prime}. Applying the differential inequality of [27, Lemma A] in the form

st|ϕ|(ϕ¯|ϕ|(st+ik)(st+ik)ϕ)=(ϕ¯|ϕ|(Hk,0st+|k|2)ϕ),-\triangle^{\operatorname{st}}|\phi|\,\leq\,-\Re\big{(}\tfrac{\bar{\phi}}{|\phi|}(\nabla^{\operatorname{st}}+ik)\cdot(\nabla^{\operatorname{st}}+ik)\phi\big{)}\,=\,\Re\big{(}\tfrac{\bar{\phi}}{|\phi|}(H_{k,0}^{\operatorname{st}}+|k|^{2})\phi\big{)},

we deduce

st|ϕ|(|k|2zλV𝟙VR)|ϕ|(|k|2+λRz)|ϕ|.\displaystyle-\triangle^{\operatorname{st}}|\phi|\,\leq\,\Re\big{(}|k|^{2}-z-\lambda V\mathds{1}_{V\geq-R}\big{)}|\phi|\,\leq\,(|k|^{2}+\lambda R-\Re z)|\phi|.

For z\Re z large enough, we have c:=z|k|2λR1c:=\Re z-|k|^{2}-\lambda R\geq 1 and

(cst)|ϕ|\displaystyle(c-\triangle^{\operatorname{st}})|\phi| \displaystyle\leq 0.\displaystyle 0.

Since the operator st-\triangle^{\operatorname{st}} is nonnegative, this implies ϕ=0\phi=0. Hence, the operator Hk,λ;Rst+zH_{k,\lambda;R}^{\operatorname{st}}+z is an injection of L2(Ω)\operatorname{L}^{2}(\Omega) into 𝒟\mathcal{D}^{\prime}, and the claim follows from the equivalence between (E1) and (E2).

Step 2. For all α0\alpha\geq 0 and R1R\geq 1, there exists a cut-off function χRαL2(Ω;[0,1])\chi_{R}^{\alpha}\in\operatorname{L}^{2}(\Omega;[0,1]) with the following properties:

  1. (i)

    χRα=0\chi_{R}^{\alpha}=0 on ER:={ω:y4Bsuch thatV(y,ω)R}E_{R}:=\{\omega:\exists y\in 4B\leavevmode\nobreak\ \text{such that}\leavevmode\nobreak\ V(y,\omega)\leq-R\};

  2. (ii)

    χRα=1\chi_{R}^{\alpha}=1 outside ERα:={ω:y(Rα+7)BwithV(y,ω)R}E_{R}^{\alpha}:=\{\omega:\exists\,y\in(R^{\alpha}+7)B\leavevmode\nobreak\ \text{with}\leavevmode\nobreak\ V(y,\omega)\leq-R\};

  3. (iii)

    |stχRα|Rα|\nabla^{\operatorname{st}}\chi_{R}^{\alpha}|\lesssim R^{-\alpha}, |(st)2χRα|Rα|(\nabla^{\operatorname{st}})^{2}\chi_{R}^{\alpha}|\lesssim R^{-\alpha};

  4. (iv)

    there exists ε>0\varepsilon>0 such that χRα1\chi_{R}^{\alpha}\to 1 almost surely as RR\uparrow\infty whenever α12+ε\alpha\leq\frac{1}{2}+\varepsilon.

Define ER,ω:={xd:yB4(x)such thatV(y,ω)R}E_{R,\omega}:=\{x\in\mathbb{R}^{d}:\exists y\in B_{4}(x)\leavevmode\nobreak\ \text{such that}\leavevmode\nobreak\ V(y,\omega)\leq-R\}. Choose ρCc(d)\rho\in C^{\infty}_{c}(\mathbb{R}^{d}) with dρ=1\int_{\mathbb{R}^{d}}\rho=1, ρ0\rho\geq 0, ρ=0\rho=0 outside 2B2B, |ρ|1|\nabla\rho|\lesssim 1, and |2ρ|1|\nabla^{2}\rho|\lesssim 1, and choose an even function χC()\chi\in C^{\infty}(\mathbb{R}) with χ(0)=0\chi(0)=0, χ(s)=1\chi(s)=1 for |s|1|s|\geq 1, |χ|1|\chi^{\prime}|\lesssim 1, and |χ′′|1|\chi^{\prime\prime}|\lesssim 1. We then construct the stationary function

χRα(x,ω):=χ(1Rαdρ(xy)dist(y,ER,ω+2B)𝑑y).\chi_{R}^{\alpha}(x,\omega):=\chi\Big{(}\frac{1}{R^{\alpha}}\int_{\mathbb{R}^{d}}\rho(x-y)\,\operatorname{dist}(y,E_{R,\omega}+2B)\,dy\Big{)}.

Properties (i)–(iii) easily follow for this choice. We turn to (iv). The definition of χRα\chi_{R}^{\alpha}, a union bound, and Markov’s inequality yield

[χRα<1]\displaystyle\mathbb{P}\left[{\chi_{R}^{\alpha}<1}\right] \displaystyle\leq {ω:infy2Bdist(y,ER,ω+2B)Rα}\displaystyle\mathbb{P}\Big{\{}\omega:\inf_{y\in 2B}\operatorname{dist}(y,E_{R,\omega}+2B)\leq R^{\alpha}\big{\}}
\displaystyle\leq {ω:dist(0,ER,ω)Rα+4}\displaystyle\mathbb{P}\big{\{}\omega:\operatorname{dist}(0,E_{R,\omega})\leq R^{\alpha}+4\big{\}}
\displaystyle\leq {ω:yBRα+8,V(y,ω)R}\displaystyle\mathbb{P}\big{\{}\omega:\exists y\in B_{R^{\alpha}+8},\,V(y,\omega)\leq-R\big{\}}
\displaystyle\lesssim Rαd[infBVR]\displaystyle R^{\alpha d}\,\mathbb{P}\big{[}{\textstyle\inf_{B}V\leq-R}\big{]}
\displaystyle\leq Rαdp𝔼[(infBV)p].\displaystyle R^{\alpha d-p}\,\mathbb{E}\big{[}{\textstyle(\inf_{B}V)_{-}^{p}}\big{]}.

Since by assumption 𝔼[(infBV)p]<\mathbb{E}\big{[}{(\textstyle\inf_{B}V)_{-}^{p}}\big{]}<\infty for p=d2+εp=\frac{d}{2}+\varepsilon for some ε>0\varepsilon>0, we deduce χRα1\chi_{R}^{\alpha}\to 1 in measure whenever α<12+εd\alpha<\frac{1}{2}+\frac{\varepsilon}{d}. In order to establish almost sure convergence, we similarly compute, noting that ER,ωE_{R,\omega} is decreasing in RR,

[χRα↛1]\displaystyle\mathbb{P}\left[{\chi_{R}^{\alpha}\not\to 1}\right] \displaystyle\leq limR0{ω:RR0such thatdist(0,ER,ω)Rα+4}\displaystyle\lim_{R_{0}\uparrow\infty}\mathbb{P}\Big{\{}\omega:\exists R\geq R_{0}\leavevmode\nobreak\ \text{such that}\leavevmode\nobreak\ \operatorname{dist}(0,E_{R,\omega})\leq R^{\alpha}+4\Big{\}}
\displaystyle\leq limR0n=0{ω:dist(0,E2nR0,ω)(2n+1R0)α+4}\displaystyle\lim_{R_{0}\uparrow\infty}\sum_{n=0}^{\infty}\mathbb{P}\Big{\{}\omega:\operatorname{dist}(0,E_{2^{n}R_{0},\omega})\leq(2^{n+1}R_{0})^{\alpha}+4\Big{\}}
\displaystyle\lesssim limR0n=0(2nR0)αdp𝔼[(infBV)p],\displaystyle\lim_{R_{0}\uparrow\infty}\sum_{n=0}^{\infty}(2^{n}R_{0})^{\alpha d-p}\,\mathbb{E}\big{[}{(\textstyle\inf_{B}V)_{-}^{p}}\big{]},

and almost sure convergence χRα1\chi_{R}^{\alpha}\to 1 follows under the same condition on α\alpha.

Step 3. Gårding inequalities:

  1.  (G1)

    For all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega) with Hk,λ;RstϕH1(Ω)H_{k,\lambda;R}^{\operatorname{st}}\phi\in H^{-1}(\Omega), there holds ϕH1(Ω)\phi\in H^{1}(\Omega) and

    stϕL2(Ω)2 4Hk,λ;RstϕH1(Ω)2+(1+8|k|2+4λR)ϕL2(Ω)2.\|\nabla^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\leq\,4\|H_{k,\lambda;R}^{\operatorname{st}}\phi\|_{H^{-1}(\Omega)}^{2}+(1+8|k|^{2}+4\lambda R)\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}.
  2.  (G2)

    For all ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega) with Hk,λstϕL2(Ω)H_{k,\lambda}^{\operatorname{st}}\phi\in\operatorname{L}^{2}(\Omega), there holds

    𝟙ΩERstϕL2(Ω)2Hk,λstϕL2(Ω)2+(1+|k|2+λR)ϕL2(Ω)2,\|\mathds{1}_{\Omega\setminus E_{R}}\nabla^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\lesssim\,\|H_{k,\lambda}^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}+(1+|k|^{2}+\lambda R)\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2},

    where as in Step 2 we have set ER:={ω:y4BwithV(y,ω)R}E_{R}:=\{\omega:\exists\,y\in 4B\leavevmode\nobreak\ \text{with}\leavevmode\nobreak\ V(y,\omega)\leq-R\}.

By density, it suffices to argue for ϕ𝒟\phi\in\mathcal{D}. We start with the Gårding inequality (G1) for the truncated operator Hk,λ;RstH_{k,\lambda;R}^{\operatorname{st}}. For ϕ𝒟\phi\in\mathcal{D}, we compute

ϕ,Hk,λ;RstϕL2(Ω)\displaystyle\Re\langle\phi,H_{k,\lambda;R}^{\operatorname{st}}\phi\rangle_{\operatorname{L}^{2}(\Omega)} \displaystyle\geq ϕ,Hk,0stϕL2(Ω)λRϕL2(Ω)2\displaystyle\Re\langle\phi,H_{k,0}^{\operatorname{st}}\phi\rangle_{\operatorname{L}^{2}(\Omega)}-\lambda R\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}
\displaystyle\geq 12stϕL2(Ω)2(2|k|2+λR)ϕL2(Ω)2,\displaystyle\frac{1}{2}\|\nabla^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}-(2|k|^{2}+\lambda R)\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2},

hence,

12stϕL2(Ω)2Hk,λ;RstϕH1(Ω)2+14ϕH1(Ω)2+(2|k|2+λR)ϕL2(Ω)2,\frac{1}{2}\|\nabla^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\leq\,\|H_{k,\lambda;R}^{\operatorname{st}}\phi\|_{H^{-1}(\Omega)}^{2}+\frac{1}{4}\|\phi\|_{H^{1}(\Omega)}^{2}+(2|k|^{2}+\lambda R)\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2},

and the claim (G1) follows.

We turn to (G2). Similarly as in Step 2, we may construct a cut-off function χR\chi_{R}^{\prime} with the following properties

  1. (i’)

    χR=0\chi_{R}^{\prime}=0 on ER:={ω:V(ω)R}E_{R}^{\prime}:=\{\omega:V(\omega)\leq-R\};

  2. (ii’)

    χR=1\chi_{R}^{\prime}=1 outside ERE_{R};

  3. (iii’)

    |stχR|1|\nabla^{\operatorname{st}}\chi_{R}^{\prime}|\lesssim 1, |(st)2χR|1|(\nabla^{\operatorname{st}})^{2}\chi_{R}^{\prime}|\lesssim 1.

Noting that Hk,λst(ϕχR)=Hk,λ;Rst(ϕχR)H_{k,\lambda}^{\operatorname{st}}(\phi\chi_{R}^{\prime})=H_{k,\lambda;R}^{\operatorname{st}}(\phi\chi_{R}^{\prime}), the result (G1) yields

st(ϕχR)L2(Ω)2 4Hk,λst(ϕχR)H1(Ω)2+(1+8|k|2+4λR)ϕL2(Ω)2.\|\nabla^{\operatorname{st}}(\phi\chi_{R}^{\prime})\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\leq\,4\|H_{k,\lambda}^{\operatorname{st}}(\phi\chi_{R}^{\prime})\|_{H^{-1}(\Omega)}^{2}+(1+8|k|^{2}+4\lambda R)\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}.

Computing

Hk,λst(ϕχR)=χRHk,λstϕ+ϕHk,0stχR2stχRstϕ,H_{k,\lambda}^{\operatorname{st}}(\phi\chi_{R}^{\prime})\,=\,\chi_{R}^{\prime}H_{k,\lambda}^{\operatorname{st}}\phi+\phi H_{k,0}^{\operatorname{st}}\chi_{R}^{\prime}-2\nabla^{\operatorname{st}}\chi_{R}^{\prime}\cdot\nabla^{\operatorname{st}}\phi,

and noting that |Hk,0stχR|1+|k||H_{k,0}^{\operatorname{st}}\chi_{R}^{\prime}|\lesssim 1+|k|, we deduce

st(ϕχR)L2(Ω)2Hk,λstϕL2(Ω)2+stχRstϕH1(Ω)2+(1+|k|2+λR)ϕL2(Ω)2.\|\nabla^{\operatorname{st}}(\phi\chi_{R}^{\prime})\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\lesssim\,\|H_{k,\lambda}^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}+\|\nabla^{\operatorname{st}}\chi_{R}^{\prime}\cdot\nabla^{\operatorname{st}}\phi\|_{H^{-1}(\Omega)}^{2}+(1+|k|^{2}+\lambda R)\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}.

Since for ϕH1(Ω)\phi^{\prime}\in H^{1}(\Omega) integration by parts yields

|ϕ,stχRstϕL2(Ω)|=|stϕstχR+ϕstχR,ϕL2(Ω)|ϕH1(Ω)ϕL2(Ω),\big{|}\big{\langle}\phi^{\prime},\nabla^{\operatorname{st}}\chi_{R}^{\prime}\cdot\nabla^{\operatorname{st}}\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\big{|}\,=\,\big{|}\big{\langle}\nabla^{\operatorname{st}}\phi^{\prime}\cdot\nabla^{\operatorname{st}}\chi_{R}^{\prime}+\phi^{\prime}\triangle^{\operatorname{st}}\chi_{R}^{\prime}\,,\,\phi\big{\rangle}_{\operatorname{L}^{2}(\Omega)}\big{|}\,\lesssim\,\|\phi^{\prime}\|_{H^{1}(\Omega)}\|\phi\|_{\operatorname{L}^{2}(\Omega)},

we find

stχRstϕH1(Ω)ϕL2(Ω),\|\nabla^{\operatorname{st}}\chi_{R}^{\prime}\cdot\nabla^{\operatorname{st}}\phi\|_{H^{-1}(\Omega)}\,\lesssim\,\|\phi\|_{\operatorname{L}^{2}(\Omega)},

hence,

st(ϕχR)L2(Ω)2Hk,λstϕL2(Ω)2+(1+|k|2+λR)ϕL2(Ω)2.\|\nabla^{\operatorname{st}}(\phi\chi_{R}^{\prime})\|_{\operatorname{L}^{2}(\Omega)}^{2}\,\leq\,\|H_{k,\lambda}^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}+(1+|k|^{2}+\lambda R)\|\phi\|_{\operatorname{L}^{2}(\Omega)}^{2}.

Since χR=1\chi_{R}^{\prime}=1 outside ERE_{R}, the claim (G2) follows.

Step 4. Conclusion.
Let ϕL2(Ω)\phi\in\operatorname{L}^{2}(\Omega) with Hk,λstϕL2(Ω)H_{k,\lambda}^{\operatorname{st}}\phi\in\operatorname{L}^{2}(\Omega). In view of the equivalence between properties (E1) and (E3), it suffices to construct a sequence (ϕn)n𝒟(\phi_{n})_{n}\subset\mathcal{D} such that ϕnϕ\phi_{n}\to\phi and Hk,λstϕnHk,λstϕH_{k,\lambda}^{\operatorname{st}}\phi_{n}\to H_{k,\lambda}^{\operatorname{st}}\phi in L2(Ω)\operatorname{L}^{2}(\Omega) as nn\uparrow\infty. We argue by truncation. Let χRα\chi_{R}^{\alpha} be the cut-off function defined in Step 2 and choose α>12\alpha>\frac{1}{2} such that property (iv) is satisfied. We show that for all R1R\geq 1 there exists a sequence (ϕn,R)n𝒟(\phi_{n,R})_{n}\subset\mathcal{D} such that

  1.   \bullet

    ϕn,RχRαϕ(χRα)2\phi_{n,R}\chi_{R}^{\alpha}\to\phi(\chi_{R}^{\alpha})^{2} and Hk,λst(ϕn,RχRα)Hk,λst(ϕ(χRα)2)H_{k,\lambda}^{\operatorname{st}}(\phi_{n,R}\chi_{R}^{\alpha})\to H_{k,\lambda}^{\operatorname{st}}(\phi(\chi_{R}^{\alpha})^{2}) in L2(Ω)\operatorname{L}^{2}(\Omega) as nn\uparrow\infty;

  2.   \bullet

    ϕ(χRα)2ϕ\phi(\chi_{R}^{\alpha})^{2}\to\phi and Hk,λst(ϕ(χRα)2)Hk,λstϕH_{k,\lambda}^{\operatorname{st}}(\phi(\chi_{R}^{\alpha})^{2})\to H_{k,\lambda}^{\operatorname{st}}\phi in L2(Ω)\operatorname{L}^{2}(\Omega) as RR\uparrow\infty.

We split the proof into three further substeps.

Substep 4.1. Proof that for all R1R\geq 1 there exists a sequence (ϕn,R)n𝒟(\phi_{n,R})_{n}\subset\mathcal{D} such that ϕn,RϕχRα\phi_{n,R}\to\phi\chi_{R}^{\alpha} and Hk,λ;Rstϕn,RHk,λst(ϕχRα)H_{k,\lambda;R}^{\operatorname{st}}\phi_{n,R}\to H_{k,\lambda}^{\operatorname{st}}(\phi\chi_{R}^{\alpha}) in L2(Ω)\operatorname{L}^{2}(\Omega) as nn\uparrow\infty.

For all R1R\geq 1, since by Step 1 the operator Hk,λ;RstH_{k,\lambda;R}^{\operatorname{st}} is essentially self-adjoint, the equivalence between properties (E1) and (E3) implies that there exists a sequence (ϕn,R)n(\phi_{n,R})_{n} such that ϕn,RϕχRα\phi_{n,R}\to\phi\chi_{R}^{\alpha} and Hk,λ;Rstϕn,RHk,λ;Rst(ϕχRα)H_{k,\lambda;R}^{\operatorname{st}}\phi_{n,R}\to H_{k,\lambda;R}^{\operatorname{st}}(\phi\chi_{R}^{\alpha}) in L2(Ω)\operatorname{L}^{2}(\Omega) as nn\uparrow\infty. By definition of χRα\chi_{R}^{\alpha}, there holds Hk,λst(ϕχRα)=Hk,λ;Rst(ϕχRα)H_{k,\lambda}^{\operatorname{st}}(\phi\chi_{R}^{\alpha})=H_{k,\lambda;R}^{\operatorname{st}}(\phi\chi_{R}^{\alpha}), and the claim follows.

Substep 4.2. Proof that for all R1R\geq 1 there holds ϕn,RχRαϕ(χRα)2\phi_{n,R}\chi_{R}^{\alpha}\to\phi(\chi_{R}^{\alpha})^{2} and Hk,λst(ϕn,RχRα)Hk,λst(ϕ(χRα)2)H_{k,\lambda}^{\operatorname{st}}(\phi_{n,R}\chi_{R}^{\alpha})\to H_{k,\lambda}^{\operatorname{st}}(\phi(\chi_{R}^{\alpha})^{2}) in L2(Ω)\operatorname{L}^{2}(\Omega) as nn\uparrow\infty.

We start from the identity

Hk,λst(ϕn,RχRα)Hk,λst(ϕ(χRα)2)=χRαHk,λst(ϕn,RϕχRα)+(ϕn,RϕχRα)Hk,0stχRα2stχRαst(ϕn,RϕχRα),H_{k,\lambda}^{\operatorname{st}}\big{(}\phi_{n,R}\chi_{R}^{\alpha}\big{)}-H_{k,\lambda}^{\operatorname{st}}\big{(}\phi(\chi_{R}^{\alpha})^{2}\big{)}=\chi_{R}^{\alpha}H_{k,\lambda}^{\operatorname{st}}(\phi_{n,R}-\phi\chi_{R}^{\alpha})+(\phi_{n,R}-\phi\chi_{R}^{\alpha})H_{k,0}^{\operatorname{st}}\chi_{R}^{\alpha}\\ -2\nabla^{\operatorname{st}}\chi_{R}^{\alpha}\cdot\nabla^{\operatorname{st}}(\phi_{n,R}-\phi\chi_{R}^{\alpha}),

and note that the convergence of (ϕn,R)n(\phi_{n,R})_{n} (cf. Substep 4.1) implies ϕn,RχRαϕ(χRα)2\phi_{n,R}\chi_{R}^{\alpha}\to\phi(\chi_{R}^{\alpha})^{2} and

lim supnHk,λst(ϕn,RχRα)Hk,λst(ϕ(χRα)2)L2(Ω)st(ϕn,RϕχRα)L2(Ω).\limsup_{n\uparrow\infty}\big{\|}H_{k,\lambda}^{\operatorname{st}}\big{(}\phi_{n,R}\chi_{R}^{\alpha}\big{)}-H_{k,\lambda}^{\operatorname{st}}\big{(}\phi(\chi_{R}^{\alpha})^{2}\big{)}\big{\|}_{\operatorname{L}^{2}(\Omega)}\,\lesssim\,\|\nabla^{\operatorname{st}}(\phi_{n,R}-\phi\chi_{R}^{\alpha})\|_{\operatorname{L}^{2}(\Omega)}.

Combining this with the Gårding inequality (G1) of Step 3 and with the convergence properties of (ϕn,R)n(\phi_{n,R})_{n}, the claim follows.

Substep 4.3. Proof that ϕ(χRα)2ϕ\phi(\chi_{R}^{\alpha})^{2}\to\phi and Hk,λst(ϕ(χRα)2)Hk,λstϕH_{k,\lambda}^{\operatorname{st}}(\phi(\chi_{R}^{\alpha})^{2})\to H_{k,\lambda}^{\operatorname{st}}\phi in L2(Ω)\operatorname{L}^{2}(\Omega) as RR\uparrow\infty.

We start from the identity

Hk,λst(ϕ(χRα)2)=(χRα)2Hk,λstϕ+ϕHk,0st(χRα)22st(χRα)2stϕ,H_{k,\lambda}^{\operatorname{st}}\big{(}\phi(\chi_{R}^{\alpha})^{2}\big{)}=(\chi_{R}^{\alpha})^{2}H_{k,\lambda}^{\operatorname{st}}\phi+\phi H_{k,0}^{\operatorname{st}}(\chi_{R}^{\alpha})^{2}-2\nabla^{\operatorname{st}}(\chi_{R}^{\alpha})^{2}\cdot\nabla^{\operatorname{st}}\phi,

and note that the properties of χRα\chi_{R}^{\alpha} and the dominated convergence theorem lead to ϕ(χRα)2ϕ\phi(\chi_{R}^{\alpha})^{2}\to\phi, (χRα)2Hk,λstϕHk,λstϕ(\chi_{R}^{\alpha})^{2}H_{k,\lambda}^{\operatorname{st}}\phi\to H_{k,\lambda}^{\operatorname{st}}\phi, and ϕHk,0st(χRα)20\phi H_{k,0}^{\operatorname{st}}(\chi_{R}^{\alpha})^{2}\to 0 in L2(Ω)\operatorname{L}^{2}(\Omega), hence

lim supRHk,λst(ϕ(χRα)2)Hk,λstϕL2(Ω)lim supRRα𝟙ΩERstϕL2(Ω).\limsup_{R\uparrow\infty}\big{\|}H_{k,\lambda}^{\operatorname{st}}\big{(}\phi(\chi_{R}^{\alpha})^{2}\big{)}-H_{k,\lambda}^{\operatorname{st}}\phi\big{\|}_{\operatorname{L}^{2}(\Omega)}\,\lesssim\,\limsup_{R\uparrow\infty}R^{-\alpha}\|\mathds{1}_{\Omega\setminus E_{R}}\nabla^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}.

Combining this with the Gårding inequality (G2) of Step 3 yields

lim supRHk,λst(ϕ(χRα)2)Hk,λstϕL2(Ω)k,λlim supRRα(Hk,λstϕL2(Ω)+R12ϕL2(Ω))= 0,\limsup_{R\uparrow\infty}\big{\|}H_{k,\lambda}^{\operatorname{st}}\big{(}\phi(\chi_{R}^{\alpha})^{2}\big{)}-H_{k,\lambda}^{\operatorname{st}}\phi\big{\|}_{\operatorname{L}^{2}(\Omega)}\,\lesssim_{k,\lambda}\,\limsup_{R\uparrow\infty}R^{-\alpha}\big{(}\|H_{k,\lambda}^{\operatorname{st}}\phi\|_{\operatorname{L}^{2}(\Omega)}+R^{\frac{1}{2}}\|\phi\|_{\operatorname{L}^{2}(\Omega)}\big{)}\,=\,0,

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. C0C_{0}-groups, commutator methods and spectral theory of NN-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 𝕫d\mathbb{z}^{d}, dd 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. Iosifyan.
  • [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.