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

On the localization regime of certain random operators within Hartree-Fock theory

Rodrigo Matos
Department of Mathematics, PUC-Rio, 22451-900, Rio de Janeiro, Brasil
[
Abstract

Localization results for a class of random Schrödinger operators within the Hartree-Fock approximation are proved in two regimes: large disorder and weak disorder/extreme energies. A large disorder threshold λHF\lambda_{\mathrm{HF}} analogous to the threshold λAnd\lambda_{\mathrm{And}} obtained in [J. Schenker, Lett. Math. Phys, Vol 105, 1 (2015)] is provided. We also show certain stability results for this large disorder threshold by giving examples of distributions for which λHF\lambda_{\mathrm{HF}} converges to λAnd\lambda_{\mathrm{And}}, or to a number arbitrarily close to it, as the interaction strength tends to zero.

I Introduction

In recent decades there has been intense activity regarding mathematical aspects of disordered systems. In the Anderson model in dimension two or higher, there is an extensive literature regarding localization in the regimes of large disorder or at spectral edges. In this context, proofs either follow the strategy of the multiscale analysis, see Bourgain-Kenig ; Carm-Klein-Mart ; Ding-Smart ; Fr-M-S-S85 ; Fr-Spenc ; Germ-Hislop-Klein ; GKBootstrap ; Kir-Stol-Stolz ; vonD-K and also the surveys Kirschsurv ; Klein-surv , or the method of fractional moments, dating back to A-Molc ; Aiz-weakdis and further developed in both discrete and continuous settings A-S-F-H ; Aiz-Elg-N-S-S and also in the context of non-monotone potentials Elg-Taut-Ves ; E-S-S , see also the survey Stolz-surv and the monograph A-W-B . Localization in the context of weak disorder and existence of the so-called Lisfshitz tails were also extensively studied, see Aiz-weakdis ; Elgart-LT3d ; RojasM-Geb1 ; RojasM-Geb2 ; Klopp-Lif ; Klopp-Nak ; Wang-univ and references therein. For results on complete localization in one dimension using large-deviation techniques we refer to Jit-Z ; Bucaj1dloc .

In the past years, there has been a number of developments in the context of many-body disordered systems such as systems with a finite number of particles A-W-Mult ; Chul-Suh ; Chu-Suh-ind ; the quantum XYXY AR-N-S-S ; H-S-Stolz ; Klein-Perez and XXZXXZ E-K22 ; E-K-S-2 ; E-K-S-1 spin chains; systems of hardcore particles Beaud-W ; harmonic oscillators in the presence of disorder N-S-S-harmonic ; particle-oscillator interactions S-M-Holst . Unlike in the single-particle Anderson-type models, where the notions of localization aimed at are usually spectral or dynamical localization, the challenges in the context of true many-body quantum systems start at defining the correct objects and notions of localization for each model.

One alternative to explore interactive quantum systems while remaining closer to the single-particle Schrödinger operator setting is to approximate the true many-body Hamiltonian by an effective one, as in the case of mean field theories and the Hartree/Hartree-Fock approximations which are widely studied beyond the setting of disordered systems B-Lieb-S ; Bene-Porta-Schlein ; Canc-Lab-Lew ; CHENN2022109702 ; Hainzl-Lew-Spa ; Lah-1 ; Lieb-Simon .

In the disordered setting, Anderson localization in the Hartree-Fock approximation was first studied in Duc . There, through the multiscale analysis technique, spectral localization was obtained in the presence of a spectral gap at both large disorder and at spectral edges. Recently in M-S localization properties of the disordered Hubbard model at positive temperature within the Hartree-Fock approximation have been established via the Aizenman-Molchanov fractional moment technique. There, exponential dynamical localization ( in fact, decay of eigenfunction correlators) is shown to hold at large disorder in dimension d2d\geq 2 and at any disorder in dimension d=1d=1 provided the interaction strength is sufficiently small. No assumption on the existence of a spectral gap is made but, in contrast, the interactions are modelled at positive temperature. The present manuscript is devoted to localization properties of random operators in the form

Hω=Δ+λVω+gVeff,ωH_{\omega}=-\Delta+\lambda V_{\omega}+gV_{{\mathrm{eff}},\omega} (I.1)

where {Vω(n)}nd\{V_{\omega}(n)\}_{n\in\mathbb{Z}^{d}} are independent, identically distributed random variables and Veff,ωV_{{\mathrm{eff}},\omega} is a multiplication operator implicitly defined by

Veff,ω(n)=mda(n,m)δm,F(Hω)δmfor allnd.V_{{\mathrm{eff}},\omega}(n)=\sum_{m\in\mathbb{Z}^{d}}a(n,m)\langle\delta_{m},F(H_{\omega})\delta_{m}\rangle\,\text{for all}\,n\in\mathbb{Z}^{d}. (I.2)

Here |a(m,n)|Caeγd(m,n)|a(m,n)|\leq C_{a}e^{-\gamma d(m,n)} will be assumed to decay sufficiently fast ( see 1-7 for the precise assumptions) with respect to a metric d:d×dd:\mathbb{Z}^{d}\times\mathbb{Z}^{d}\rightarrow\mathbb{R}, Ca>0C_{a}>0, γ>0\gamma>0 and FF is an analytic function on a strip {|Imz|<η}\{|\mathrm{Im}z|<\eta\} which is bounded. It is worth noting that the above setting allows for the decay of |a(m,n)||a(m,n)| to be of polynomial type. The above model is somewhat analogous (in the Hartree-Fock setting) to models in the single-particle setting with fast decaying potentials which still exhibit monotonicity properties (for instance, the ones studied in Kir-Stol-Stolz ). In the particular case where F(z)=11+eβ(zμ¯)F(z)=\frac{1}{1+e^{\beta(z-\bar{\mu})}} is the Fermi-Dirac function at temperature β1>0\beta^{-1}>0 and a chemical potential μ¯\bar{\mu}\in\mathbb{R} and a(m,n)=δmna(m,n)=\delta_{mn} ( with δmn\delta_{mn} the Kronecker delta) (I.2) simplifies to operators already studied in M-S . There for a fixed β>0\beta>0, dynamical localization is shown in any dimension provided |g|<g0|g|<g_{0} and λ>λ0\lambda>\lambda_{0} for certain constants g0g_{0} and λ0\lambda_{0} which depend on β\beta and dd but a more concrete estimate for λ0\lambda_{0} was not pursued there. In this note, we generalized the large disorder result of M-S to the operator (I.2), obtain a novel result of localization at weak disorder/extreme energies and, moreover, study the question of stability of the large disorder threshold under ‘weak’ interactions, inspired by the analysis of Schenkl . In particular, it is proven here that there is a large disorder threshold λHF\lambda_{\mathrm{HF}} such that the operators given by (I.2) exhibit dynamical localization provided λ>λHF\lambda>\lambda_{HF} as long as |g|F|g|\norm{F}_{\infty} is sufficiently small. Moreover, we show that λHFλAnd\lambda_{HF}\to\lambda_{\mathrm{And}} as |g|F0|g|\norm{F}_{\infty}\to 0 where λAnd\lambda_{\mathrm{And}} is the solution of the transcendental equation

λAnd=2ρμdeln(λAnd2ρ)\lambda_{\mathrm{And}}=2\norm{\rho}_{\infty}\mu_{d}e\ln\left(\frac{\lambda_{\mathrm{And}}}{2\norm{\rho}_{\infty}}\right) (I.3)

with μd\mu_{d} the connectivity constant of d\mathbb{Z}^{d}. For the uniform distribution in [1,1][-1,1], in which case 2ρ=12\norm{\rho}_{\infty}=1, this value of λAnd\lambda_{\mathrm{And}} was obtained for the Anderson model in Schenkl and coincides with Anderson’s original prediction in Anderson . To the best of our knowledge, in arbitrary dimension λAnd\lambda_{\mathrm{And}} in (I.3) is the best rigorous large disorder threshold proved with current methods. It is worth noting that letting |g|F0|g|\norm{F}_{\infty}\to 0 formally in (I.2) we obtain the Anderson model HAnd=Δ+VωH_{\mathrm{And}}=-\Delta+V_{\omega}.

We now comment on other technical merits of the present work, for further technical aspects we refer to section II.1 below. Our first observation is that even for the non-interacting Anderson model HAndH_{\mathrm{And}}, the fractional moment method requires the random variables VωV_{\omega} to have a density ρ\rho which is “sufficiently regular". Thus, it is to be expected that a direct application of this technique to interacting models, which is the approach adopted here and also in M-S , will require further regularity of ρ\rho. The previous paper M-S covers a large class of probability distributions with suppρ=\mathrm{supp}\rho=\mathbb{R} by making use of the symmetry F(z)=1F(z)F(z)=1-F(-z) and decay properties of the Fermi-Dirac function when Rez\mathrm{Re}z\to\infty in order to obtain certain improved Combes-Thomas bounds. Such bounds reflect decay of the effective potential at a given site Veff,ω(n)V_{{\mathrm{eff}},\omega}(n) when the local potential ω(m)\omega(m) is changed at a site mnm\neq n. However, such bounds appear not to be available in the generality studied here. In fact, they seem not to be available even when one restricts (I.2) to the case of nearest neighbor lattice fermions, i.e., when a(m,n)=1a(m,n)=1 if and only if |mn|=1|m-n|=1 where |m|=|m1|++|md||m|=|m_{1}|+\cdots+|m_{d}| and thus Veff,ω(n)=nnδn,F(Hω)δnV_{{\mathrm{eff}},\omega}(n)=\sum_{n^{\prime}\sim n}\langle\delta_{n^{\prime}},F(H_{\omega})\delta_{n^{\prime}}\rangle with nnn^{\prime}\sim n indicating that nn^{\prime} and nn are nearest neighbors on d\mathbb{Z}^{d}. The key observation surrounding the present paper is that there is a trade off between the regularity/decay properties of FF on the real line, the decay properties of the interaction kernel a(m,n)a(m,n) and the density ρ\rho. Namely, by reducing the class of probability distributions covered by our main result, we are able to include interactions of a much longer range, including the case where a(m,n)a(m,n) only decays in an algebraic fashion and where F(z)F(z) is bounded of a strip but does not necessarily decay as Rez\mathrm{Re}z\to\infty. In conclusion, even though the methods employed here to obtain the a-priori bound on fractional moments of the Green’s function follow the general scheme of M-S , in order to prove our stability result, we need to keep an explicit dependence on all parameters involved η,λ,g,F\eta,\lambda,g,\norm{F}_{\infty} and now have the inclusion the decay rate γ\gamma of a(m,n)a(m,n) as well. Once an a-priori bound on fractional moments of the Green’s function is obtained, we follow the approach of Schenker in Schenkl in order to get the best large disorder threshold which seem to be available with current methods which turns out to converge to λAnd\lambda_{\mathrm{And}} when |g|F0|g|\norm{F}_{\infty}\to 0.

This paper is dedicated to Abel Klein in occasion of his 78th birthday. Klein’s contributions to the field go well beyond the aforementioned works and can hardly be overstated. Certain aspects of the present work were also inspired by Klein’s efforts. For instance, the idea of studying distributions near a suitable chosen density (for which explicit calculations are available) used below in assumption 6 is analogous to Acosta-Klein , where analyticity of the density of states on a strip is shown for distributions sufficiently close to the Cauchy distribution. Moreover, throughout the note, Combes-Thomas type bounds for kernels of analytic functions of HωH_{\omega} are used extensively. In GK-CT such bounds are obtained in great generality which provides hope for future extensions of the results below.

II Model, Statement of the main results and proof strategies

This note concerns random operators of the form

Hω=A+λVω+gVeff,ωH_{\omega}=A+\lambda V_{\omega}+gV_{{\mathrm{eff}},\omega} (II.1)

acting on 2(d)\ell^{2}\left(\mathbb{Z}^{d}\right) as follows:

  1. 1.

    (Aψ)(n)=|nn|=1ψ(n)(A\psi)(n)=\sum_{|n^{\prime}-n|=1}\psi(n^{\prime}) for each ψ2(d)\psi\in\ell^{2}\left(\mathbb{Z}^{d}\right), i.e. A:2(d)2(d)A:\ell^{2}\left(\mathbb{Z}^{d}\right)\rightarrow\ell^{2}\left(\mathbb{Z}^{d}\right) is the adjacency operator of d\mathbb{Z}^{d}.

  2. 2.

    (Vωψ)(n)=ω(n)ψ(n)(V_{\omega}\psi)(n)=\omega(n)\psi(n) for each ψ2(d)\psi\in\ell^{2}\left(\mathbb{Z}^{d}\right) where {ω(n)}nd\{\omega(n)\}_{n\in\mathbb{Z}^{d}} are independent, identically distributed random variables with a bounded density ρ\rho.

  3. 3.

    The effective potential Veff,ω:2(d)2(d)V_{{\mathrm{eff}},\omega}:\ell^{2}\left(\mathbb{Z}^{d}\right)\rightarrow\ell^{2}\left(\mathbb{Z}^{d}\right) is a multiplication operator implicitly defined by

    Veff,ω(n)=mda(n,m)δm,F(Hω)δmfor allnd.V_{{\mathrm{eff}},\omega}(n)=\sum_{m\in\mathbb{Z}^{d}}a(n,m)\langle\delta_{m},F(H_{\omega})\delta_{m}\rangle\,\,\,\text{for all}\,\,\,n\in\mathbb{Z}^{d}. (II.2)

    We impose the following conditions on a(n,m)a(n,m) and FF.

  4. 4.

    There exists η>η0>0\eta>\eta_{0}>0 such that FF is an analytic function on the strip

    𝒮η={|Imz|<η}\mathcal{S}_{\eta}=\{|\mathrm{Im}z|<\eta\}

    and bounded on its closure 𝒮η¯\overline{\mathcal{S}_{\eta}}. Moreover, we assume that F()F(\mathbb{R})\subset\mathbb{R}.

  5. 5.

    The values a(m,n)a(m,n) are real numbers for all m,ndm,n\in\mathbb{Z}^{d} and

    |a(m,n)|Caeγad(m,n)|a(m,n)|\leq C_{a}e^{-\gamma_{a}d(m,n)} (II.3)

    for constants Ca>0C_{a}>0 and γa>0\gamma_{a}>0 and some metric d:d×dd:\mathbb{Z}^{d}\times\mathbb{Z}^{d}\rightarrow\mathbb{R} for which there exists δ(0,γa/2)\delta\in(0,\gamma_{a}/2) such that

    Sδγa,:=supndmde(δγa)d(m,n)<.\norm{S_{\delta-\gamma_{a}}}_{\infty,\infty}:=\sup_{n\in\mathbb{Z}^{d}}\sum_{m\in\mathbb{Z}^{d}}e^{(\delta-\gamma_{a})d(m,n)}<\infty. (II.4)
  6. 6.

    We also assume that suppρ=\mathrm{supp}\rho=\mathbb{R} and that for some c1>0c_{1}>0 and ε1>0\varepsilon_{1}>0

    ρ(v1)ρ(v2)ec1|v1v2|,for allv1,v2\frac{\rho(v_{1})}{\rho(v_{2})}\geq e^{-c_{1}|v_{1}-v_{2}|},\,\,\,\text{for all}\,\,v_{1},v_{2}\in\mathbb{R} (II.5)

    and

    supvρ(v)ρ(α)eε1|vα|𝑑α<\sup_{v\in\mathbb{R}}\frac{\rho(v)}{\int^{\infty}_{-\infty}\rho(\alpha)e^{-\varepsilon_{1}\absolutevalue{v-\alpha}}\,d\alpha}<\infty (II.6)
Remark 1

Assumptions 1-6 suffice for the first result of this note, namely, localization at large disorder given by Theorem 1 below, and also for the stability bounds on the large disorder threshold of Corolary 2. It is worth observing that assumption 6 holds, for instance, for the Cauchy distribution and also for the (negative) exponential distribution.

For the results of localization at weak disorder/extreme energies we will make the following additional requirement.

  1. 7.

    We further assume that ρ(v)=h(v)ecρ|v|\rho(v)=h(v)e^{-c_{\rho}|v|} for some cρ>0c_{\rho}>0 where

    h(v1)h(v2)eε2|v1v2|for allv1,v2\frac{h(v_{1})}{h(v_{2})}\geq e^{-\varepsilon_{2}|v_{1}-v_{2}|}\,\,\,\text{for all}\,\,v_{1},v_{2}\in\mathbb{R} (II.7)

    for some ε2(0,12cρ)\varepsilon_{2}\in(0,\frac{1}{2}c_{\rho}).

Remark 2

Intuitively speaking, assumption 7 means that ρ(v)\rho(v) is near the (negative and two sided) exponential distribution with density ρ1(v)=cρ2ecρ|v|\rho_{1}(v)=\frac{c_{\rho}}{2}e^{-c_{\rho}|v|} and suppρ1=\mathrm{supp}\rho_{1}=\mathbb{R}. With the help of the mean value theorem and Young’s inequality, one may check that this assumption applies to densities of the form ρ(v)=h(v)ec|v|\rho(v)=h(v)e^{-c|v|} with h(v)=Ck,ε(1+ε|v|k)eα|v|h(v)=C_{k,\varepsilon}(1+\varepsilon|v|^{k})e^{-\alpha|v|}, α>0\alpha>0, k>1k>1 and ε\varepsilon sufficiently small depending on α\alpha and kk.

Working with finite volume restrictions of both HωH_{\omega} and also Veff,ωV_{{\mathrm{eff}},\omega} will turn out convenient thus we let ΛL=[L,L]dd\Lambda_{L}=[-L,L]^{d}\cap\mathbb{Z}^{d} and

Veff,ω,L(n)=mΛLa(n,m)δm,F(Hω,L)δmfor allnΛL.V_{{\mathrm{eff}},\omega,L}(n)=\sum_{m\in\Lambda_{L}}a(n,m)\langle\delta_{m},F(H_{\omega,L})\delta_{m}\rangle\,\,\,\text{for all}\,n\in\Lambda_{L}. (II.8)

where Hω,L=𝟙L(A+Vω+Veff,ω,L)𝟙LH_{\omega,L}=\mathds{1}_{L}(A+V_{\omega}+V_{{\mathrm{eff}},\omega,L})\mathds{1}_{L} and 𝟙L:2(d)2(ΛL)\mathds{1}_{L}:\ell^{2}\left(\mathbb{Z}^{d}\right)\rightarrow\ell^{2}\left(\Lambda_{L}\right) is the projection onto span{δl:lΛL}\mathrm{span}\{\delta_{l}:\,\,\,l\in\Lambda_{L}\}. We will often write

U(n)=ω(n)+gλVeff,ω,L,nΛΛLU(n)=\omega(n)+\frac{g}{\lambda}V_{{\mathrm{eff}},\omega,L},\,\,\,n\in\Lambda^{\prime}\subset\Lambda_{L} (II.9)

to denote the “full potential" at site nn. It will be shown below in Lemma 8 that under assumptions 1-6 the conditional distribution of U(n0)=vU(n_{0})=v at specified values of {U(n)}nΛ{n0}\{U(n)\}_{n\in\Lambda^{\prime}\setminus\{n_{0}\}} has a density which is bounded with an upper bound independent of the parameters ω,Λ\omega,\Lambda^{\prime} and LL. This upper bound is denoted herein by MM_{\infty} and the conditional density by ρn0eff=ρn0,Leff,Λ\rho^{{\mathrm{eff}}}_{n_{0}}=\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}. We also recall the definition of the eigenfunction correlators for an operator HH:

QI(m,n):=sup|φ|1|δm,φ(H)δn|Q_{I}(m,n):=\sup_{|\varphi|\leq 1}\absolutevalue{\langle\delta_{m},\varphi(H)\delta_{n}\rangle} (II.10)

where the supremum is taken over Borel measurable functions φ\varphi bounded by one and supported on the interval II. In case I=I=\mathbb{R} we simply write Q(m,n)Q(m,n). In what follows we denote by QI,LΛ(m,n)Q^{\Lambda^{\prime}}_{I,L}(m,n) the eigenfunction correlators of Hω,LΛ=𝟙ΛHω,L𝟙ΛH^{\Lambda^{\prime}}_{\omega,L}={\mathds{1}}_{\Lambda^{\prime}}H_{\omega,L}{\mathds{1}}_{\Lambda^{\prime}} for ΛΛL\Lambda^{\prime}\subset\Lambda_{L} and by 𝔼(f)\mathbb{E}(f) the expected value of ff with respect to the probability space in question.

Our first result is the following.

Theorem 1

Under assumptions 1-6 there exist λHF=λHF(g,η0,F,d,ρ,γa)\lambda_{\mathrm{HF}}=\lambda_{\mathrm{HF}}(g,\eta_{0},\norm{F}_{\infty},d,\rho,\gamma_{a}) and g0=g0(Ca,d,ρ,λ,γa,η0)g_{0}=g_{0}(C_{a},d,\rho,\lambda,\gamma_{a},\eta_{0}) such that for all λ>λHF\lambda>\lambda_{\mathrm{HF}} and |g|F<g0\absolutevalue{g}\norm{F}_{\infty}<g_{0} we have that

𝔼(QLΛ(m,n))Ceν|mn|\mathbb{E}\left(Q^{\Lambda^{\prime}}_{L}(m,n)\right)\leq Ce^{-\nu^{\prime}|m-n|} (II.11)

for some ν>0\nu^{\prime}>0 and C>0C>0 independent of LL and Λ\Lambda^{\prime}. Moreover, λHF\lambda_{\mathrm{HF}} satisfies

λHF=2Mμdeln(λHF2M)\lambda_{\mathrm{HF}}=2M_{\infty}\mu_{d}e\ln\left(\frac{\lambda_{\mathrm{HF}}}{2M_{\infty}}\right) (II.12)

where μd\mu_{d} is the conective constant of d\mathbb{Z}^{d} and M=M(η0,d)M_{\infty}=M_{\infty}(\eta_{0},d) is given by

M=supωΩsupLsupΛΛLsupn0Λsupvρn0,Leff,Λ(v).M_{\infty}=\sup_{\omega\in\Omega}\sup_{L\in\mathbb{N}}\sup_{\Lambda^{\prime}\subset\Lambda_{L}}\sup_{n_{0}\in\Lambda}\sup_{v\in\mathbb{R}}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v). (II.13)
Remark 3

It readily follows that the analogue of (II.11) also holds in the infinite volume, see Lemma 15 below and (A-W-B, , Proposition 7.6).

Theorem 1 above extends to the present context a result of Schenker Schenkl , who obtained the large disorder threshold λAnd\lambda_{\mathrm{And}} which solves

λAnd=2ρμdeln(λAnd2ρ).\lambda_{\mathrm{And}}=2\norm{\rho}_{\infty}\mu_{d}e\ln\left(\frac{\lambda_{\mathrm{And}}}{2\norm{\rho}_{\infty}}\right). (II.14)

for the Anderson model with a uniformly distributed potential on [1,1][-1,1].

We also show that λHF\lambda_{\mathrm{HF}} is close to λAnd\lambda_{\mathrm{And}} in a quantified fashion.

Corollary 2

Let λHF\lambda_{\mathrm{HF}} be as in (II.12) and λAnd\lambda_{\mathrm{And}} be given by (II.14). Under assumptions 1-6 we have that |λHFλAnd|0\absolutevalue{\lambda_{\mathrm{HF}}-\lambda_{\mathrm{And}}}\to 0 as |g|F0|g|\norm{F}_{\infty}\to 0.

Before stating our second theorem we let, for each n0Λn_{0}\in\Lambda^{\prime}

ψsn0(z)=|v|sρn0,Leff,Λ(v)|vz|s𝑑v,\psi^{n_{0}}_{s}(z)=\int^{\infty}_{-\infty}\frac{|v|^{s}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)}{|v-z|^{s}}\,dv, (II.15)
ϕsn0(z)=ρn0,Leff,Λ(v)|vz|s𝑑v\phi^{n_{0}}_{s}(z)=\int^{\infty}_{-\infty}\frac{\rho^{\mathrm{eff},\Lambda^{\prime}}_{n_{0},L}(v)}{\absolutevalue{v-z}^{s}}\,dv (II.16)

and

Ds,1=supLsupΛΛLsupz,n0Λψsn0(z)ϕsn0(z).D_{s,1}=\sup_{L\in\mathbb{N}}\sup_{\Lambda^{\prime}\subset\Lambda_{L}}\sup_{z\in\mathbb{C},n_{0}\in\Lambda^{\prime}}\frac{\psi^{n_{0}}_{s}(z)}{\phi^{n_{0}}_{s}(z)}. (II.17)

As we shall see below, under assumptions 1-7 the measure ρn0,Leff,Λ(v)dv\rho^{\mathrm{eff},\Lambda^{\prime}}_{n_{0},L}(v)\,dv is 11-moment regular in the sense of (A-W-B, , Definition 8.5) meaning that Ds,1<D_{s,1}<\infty for all s(0,1)s\in(0,1). We also define the Green’s function of HωH_{\omega} at zσ(Hω)z\in\mathbb{C}\setminus\sigma(H_{\omega}) by

G(m,n;z)=δm,(Hωz)1δnG(m,n;z)=\langle\delta_{m},(H_{\omega}-z)^{-1}\delta_{n}\rangle (II.18)

and let, for ΛΛL\Lambda^{\prime}\subset\Lambda_{L}, GL(m,n;z)G_{L}(m,n;z) and GLΛ(m,n;z)G^{\Lambda^{\prime}}_{L}(m,n;z) be the Green’s function of Hω,LH_{\omega,L} and Hω,LΛ=𝟙ΛHω,L𝟙ΛH^{\Lambda^{\prime}}_{\omega,L}={\mathds{1}}_{\Lambda^{\prime}}H_{\omega,L}{\mathds{1}}_{\Lambda^{\prime}}, respectively:

GL(m,n;z)=δm,(Hω,Lz)1δnandGLΛ(m,n;z)=δm,(Hω,LΛz)1δn.G_{L}(m,n;z)=\langle\delta_{m},(H_{\omega,L}-z)^{-1}\delta_{n}\rangle\,\,\text{and}\,\,G^{\Lambda^{\prime}}_{L}(m,n;z)=\langle\delta_{m},(H^{\Lambda^{\prime}}_{\omega,L}-z)^{-1}\delta_{n}\rangle. (II.19)

We emphasize that the effective potential is Veff,ω,LV_{{\mathrm{eff}},\omega,L} for both of the above operators. Finally, we denote by G0(m,n;z)G_{0}(m,n;z) the Green’s function of the “free” operator AA, namely

G0(m,n;z)=δm,(Az)1δn.G_{0}(m,n;z)=\langle\delta_{m},(A-z)^{-1}\delta_{n}\rangle. (II.20)

We are now ready to state our second Theorem which yields localization at weak disorder/extreme energies provided the interaction strength is not too large relative to the remaining parameters.

Theorem 3

Given II\subset\mathbb{R} there exist λ0=λ0(I)\lambda_{0}=\lambda_{0}(I) and g1=g1(Ca,d,ρ,λ,γa,η0)g_{1}=g_{1}(C_{a},d,\rho,\lambda,\gamma_{a},\eta_{0}) such that whenever |g|F<g1\absolutevalue{g}\norm{F}_{\infty}<g_{1} and |λ|<λ0|\lambda|<\lambda_{0} we have that

𝔼(QI,LΛ(m,n))Ceν|mn|\mathbb{E}\left(Q^{\Lambda^{\prime}}_{I,L}(m,n)\right)\leq Ce^{-\nu^{\prime}|m-n|} (II.21)

for some ν>0\nu^{\prime}>0 and C>0C>0 independent of ΛΛL\Lambda^{\prime}\subset{\Lambda_{L}} and LL. Moreover, we have that

λ0=sups(0,1)supμ>0infEIλ^s,μ(E)\lambda_{0}=\sup_{s\in(0,1)}\sup_{\mu>0}\inf_{E\in I}\widehat{\lambda}_{s,\mu}(E) (II.22)

where

λ^s,μ(E)=(Ds,1supδ0supudvd|G0(u,v;E+iδ)|seμ|uv|)1s.\widehat{\lambda}_{s,\mu}(E)=\left(D_{s,1}\sup_{\delta\neq 0}\sup_{u\in\mathbb{Z}^{d}}\sum_{v\in\mathbb{Z}^{d}}|G_{0}(u,v;E+i\delta)|^{s}e^{\mu|u-v|}\right)^{\frac{-1}{s}}. (II.23)
Remark 4

By the Combes-Thomas bound (A-W-B, , Theorem 10.5), Theorem 3 is applicable when Iσ(A)=I\cap\sigma(A)=\emptyset. In particular, since it was assumed that suppρ=\mathrm{supp}\rho=\mathbb{R}, this yields a non-trivial result for all λ0\lambda\neq 0. We choose the above formulation for general II\subset\mathbb{R} for future reference, as in more general settings localization at weak disorder may be established away from the 1\ell^{1} spectrum of the deterministic part of HωH_{\omega}, see (A-W-B, , Theorem 10.4) and comments therein.

II.1 Proof strategy: discrete subharmonicty bounds

The proofs of Theorems 1 and 3 follow the general scheme of the Aizenman-Molchanov fractional moment method A-Molc ; Aiz-weakdis and further refinements of their technique, in particular the one in Schenkl , combined with tools from M-S (and a few technical improvements on it). Their approach requires the random potential to be sufficiently regular (even though it allows for certain singularities) which is the case in the Anderson model HAnd,ω=A+λVωH_{\mathrm{And},\omega}=A+\lambda V_{\omega} given by assumptions 1 and 2. The first difficulty in the present work is that the full random potential is of the form Uω(n)=ω(n)+gλVeff,ω(n)U_{\omega}(n)=\omega(n)+\frac{g}{\lambda}V_{{\mathrm{eff}},\omega}(n) thus U(n)U(n) and U(m)U(m) are correlated for all values of mm and nn and, a-priori, their regularity is unknown. While correlations are not necessarily a problem for the fractional moment technique, as it is well-known and already stated in the Aizenman-Molchanov original work A-Molc , in order to prove localization one needs at least some regularity on the conditional distributions of U(n0)U(n_{0}), for each n0n_{0}, when the remaining variables {U(n)}nn0\{U(n)\}_{n\neq n_{0}} are specified. Moreover, the involved bounds should be uniform in n0n_{0}. At an intuitive level, such requirement on the conditional distributions amounts to the variables U(m)U(m) and U(n)U(n) being less and less correlated as |mn|+|m-n|\to+\infty so that some of the regularity of ω(n0)\omega(n_{0}) is persists in the conditional distribution of U(n0)U(n_{0}). The technical implementation of the above reasoning essentially consists of two main parts, each of them having of a few steps. The first part is completely deterministic and aims at showing that, in terms of the metric dd in which |a(m,n)|Caeγad(m,n)\absolutevalue{a(m,n)}\leq C_{a}e^{-\gamma_{a}d(m,n)}, the effective potential Veff,ωV_{{\mathrm{eff}},\omega} is a quasilocal function of the random variables {ω(n)}nd\{\omega(n)\}_{n\in\mathbb{Z}^{d}}. The second part involves applying the fractional moment method in the spirit of Schenkl once the regularity of the {U(n)}nΛ\{U(n)\}_{n\in\Lambda^{\prime}} is determined.

Before stating the main steps of the proof, let us remark that for simplicity we do not always mention finite-volume restrictions in this sketch. Nonetheless, their introduction is technically important for the arguments, as it will be clear later in the note. Moreover, each point of the outline below is carried out in the appropriate smallness regime (cf. Theorems 1 and 3).

  1. (i)

    Step 1: Show that

    |Veff,ω(n)ω(l)|C1eδd(n,l)\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega}(n)}{\partial\omega(l)}}\leq C_{1}e^{-\delta d(n,l)}

    holds for every ωΩ\omega\in\Omega, n,lΛn,l\in\Lambda^{\prime} with C1C_{1}, δ>0\delta>0 independent of ω\omega and Λ\Lambda^{\prime}. This will allow us to make the change of variables ω(n)U(n):=ω(n)+gλVeff,ω(n)\omega(n)\mapsto U(n):=\omega(n)+\frac{g}{\lambda}V_{{\mathrm{eff}},\omega}(n) and guarantee that the map ωU\omega\mapsto U is a diffeomorphism in |Λ|\mathbb{R}^{\absolutevalue{\Lambda^{\prime}}} for each finite set Λd\Lambda^{\prime}\subset\mathbb{Z}^{d}.

  2. (ii)

    Step 2: Fix n0Λn_{0}\in\Lambda^{\prime} and α\alpha\in\mathbb{R}. Let Uα(n)=U(n)+(αU(n0))δn0U_{\alpha}(n)=U(n)+(\alpha-U(n_{0}))\delta_{n_{0}} be a rank-one perturbation of UU at n0n_{0} and define {ωα}nΛ\{\omega_{\alpha}\}_{n\in\Lambda^{\prime}} to be such that Uα(n)=ωα(n)+gλVeff,ωα(n)U_{\alpha}(n)=\omega_{\alpha}(n)+\frac{g}{\lambda}V_{{\mathrm{eff}},\omega_{\alpha}}(n) for all nΛn\in\Lambda^{\prime}. Then for some δ>0\delta>0 and C>0C>0 we have that

    |ω(n)ωα(n)|C1|αU(n0)|eδd(n,n0).\absolutevalue{\omega(n)-\omega_{\alpha}(n)}\leq C_{1}|\alpha-U(n_{0})|e^{-\delta d(n,n_{0})}. (II.24)

    This step, along with assumption 6 will allow us to control fluctuations of the density ρ\rho which naturally appear when computing the conditional density ρn0eff.\rho^{{\mathrm{eff}}}_{n_{0}}.

  3. (iii)

    Step 3: Prove that

    |Veff,ω(n)ω(l)Veff,ωα(n)ω(l)|eδ(d(n,l)+d(n,n0)).\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega}(n)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha}}(n)}{\partial\omega(l)}}\lesssim e^{-\delta\left(d(n,l)+d(n,n_{0})\right)}. (II.25)

    This step will help us control fluctuations in the Jacobian of the above change of variables which also appear in the expression for ρn0eff\rho^{{\mathrm{eff}}}_{n_{0}}.

Once the above steps are completed, the second part of the proof makes use of probabilistic techniques.

  1. (iv)

    Step 4: Use the bounds from Steps 1-3 to conclude that under assumptions 1-6 the conditional density ρn0eff\rho^{{\mathrm{eff}}}_{n_{0}} exists and is (uniformly) bounded. Moreover, under assumptions 1-7 conclude that ρn0eff\rho^{{\mathrm{eff}}}_{n_{0}} exhibits some additional regularity.

  2. (v)

    Step 5: Complete the proof using the fractional moment technique.

While the overall strategy outlined above is similar to the one in M-S there are some key technical differences. Firstly, by obtaining the cancellation directly on (iii) we are able to avoid having to bound the second derivatives 2Veff,ω(n)ω(m)ω(l)\frac{\partial^{2}V_{{\mathrm{eff}},\omega}(n)}{\partial\omega(m)\partial\omega(l)} which shortens the proof quite a bit, especially for the model studied here where a:d×da:\mathbb{Z}^{d}\times\mathbb{Z}^{d}\rightarrow\mathbb{R} may be non-local. Secondly, in Step 4 the observation that further regularity of ρn0eff\rho^{{\mathrm{eff}}}_{n_{0}} can be obtained under assumption 7, which ultimately yields the localization at weak disorder/extreme energies result, is also new. A third difference is present in Step 5. Namely, while localization at large disorder was obtained in (M-S, , Theorem 2), in the case where a(m,n)=δmna(m,n)=\delta_{mn}, the explicit dependence of the large disorder threshold on the remaining parameters is not given (although it can certainly be inferred from the proof). Here we provide a self-consistent equation for the large disorder threshold in (II.12). Moreover we show that under assumption 7 this threshold is somewhat sharp from the point of view of what is currently known for the Anderson model from Schenkl . Indeed, within the class of exponential distributions, we show that the difference between the large disorder threshold λAnd\lambda_{\mathrm{And}} of the non-interactive setting ( cf. (II.14)) and λHF\lambda_{\mathrm{HF}} given by (II.12) can be made arbitrarily small when the interaction strength tends to zero.

Turning to the question of how to show the quasilocality bounds in Steps (i)-(iii), the following Lemma will be useful since Veff,ωV_{{\mathrm{eff}},\omega} and, by extension, its partial derivatives {Veff,ω(n)ω(l)}n,lΛ\{\frac{\partial V_{{\mathrm{eff}},\omega}(n)}{\partial\omega(l)}\}_{n,l\in\Lambda^{\prime}} are only implicitly defined and hence the desired control of them can only be achieved via inequalities of self-consistent nature.

Lemma 4

(A-W-B, , Theorem 9.2) Let 𝔾\mathbb{G} be a countable set and K:(𝔾)(𝔾)K:\ell^{\infty}\left(\mathbb{G}\right)\rightarrow\ell^{\infty}\left(\mathbb{G}\right) be given by (Kφ)(n)=u𝔾K(n,u)φ(u)(K\varphi)(n)=\sum_{u\in\mathbb{G}}K(n,u)\varphi(u) with K(n,u)0K(n,u)\geq 0 and

K,:=supn𝔾u𝔾K(n,u)<1.\|K\|_{\infty,\infty}:=\sup_{n\in\mathbb{G}}\sum_{u\in\mathbb{G}}K(n,u)<1. (II.26)

Let W:(𝔾)(0,)W:\ell^{\infty}\left(\mathbb{G}\right)\rightarrow(0,\infty) and ψ(𝔾)\psi\in\ell^{\infty}\left(\mathbb{G}\right) be positive functions such that

b1:=u𝔾W(u)ψ(u)<andb2:=supm𝔾u𝔾W(u)W(m)K(u,m)<1.b_{1}:=\sum_{u\in\mathbb{G}}W(u)\psi(u)<\infty\,\,\mathrm{and}\,\,b_{2}:=\sup_{m\in\mathbb{G}}\sum_{u\in\mathbb{G}}\frac{W(u)}{W(m)}K(u,m)<1. (II.27)

Then, any φ(𝔾)\varphi\in\ell^{\infty}\left(\mathbb{G}\right) which satisfies

0φ(n)ψ(n)+(Kφ)(n)foralln𝔾0\leq\varphi(n)\leq\psi(n)+(K\varphi)(n)\,\,\,\,\mathrm{for}\,\mathrm{all}\,\,n\in\mathbb{G}

also obeys the bound

ndW(n)φ(n)b11b2foralln𝔾.\sum_{n\in\mathbb{Z}^{d}}W(n)\varphi(n)\leq\frac{b_{1}}{1-b_{2}}\,\,\,\,\mathrm{for}\,\mathrm{all}\,\,n\in\mathbb{G}. (II.28)

The first instance where Lemma 4 is applied is in Step 1 with the choice

φ1(n)=|Veff(n)ω(l)|,W(n)=eδ|nl|,δ=min{ν,γa/2},\varphi_{1}(n)=\absolutevalue{\frac{\partial V_{{\mathrm{eff}}}(n)}{\partial\omega(l)}},\,\,W(n)=e^{\delta|n-l|},\,\,\delta=\min\{\nu,\gamma_{a}/2\}, (II.29)

where ν\nu is given below in (III.11) and γa\gamma_{a} is as in 5. To accomplish Step 2, Lemma 4 is applied to

φ2(n)=|ω(n)ωα(n)|δnn0,W(n)=eδ|nl|.\varphi_{2}(n)=\absolutevalue{\omega(n)-\omega_{\alpha}(n)}\delta_{n\neq n_{0}},\,\,\,W(n)=e^{\delta|n-l|}. (II.30)

with δ\delta as above. In Step 3, Lemma 4 is applied to

φ3(n)=|Veff(n)ω(l)Veff,ωα(n)ω(l)|,W(n)=eδ|nl|.\varphi_{3}(n)=\absolutevalue{\frac{\partial V_{{\mathrm{eff}}}(n)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha}}(n)}{\partial\omega(l)}},\,\,\,W(n)=e^{\delta|n-l|}. (II.31)

Finally, in Step 5 Lemma 4 is applied to different functions depending on whether we wish to show decay of the Green’s function in the large disorder or in the weak disorder/extreme energies regime. In the large disorder regime of Theorem 1, thanks to an a-priori bound which follows from Lemma 8 below, Lemma 4 is applied to a fixed ndn\in\mathbb{Z}^{d} letting

φ(m)=supΛd𝔼(|GΛ(m,n;z)|s),W(m)=eν|mn|,\varphi(m)=\sup_{\Lambda\subset\mathbb{Z}^{d}}\mathbb{E}\left(\absolutevalue{G^{\Lambda}(m,n;z)}^{s}\right),\,\,\,W(m)=e^{\nu^{\prime}|m-n|}, (II.32)

for a suitable ν>0\nu^{\prime}>0 and choosing

K(m,u)=2sMsλsδ|mu|=1,ψ(m)=2sMsλsδm,n.K(m,u)=\frac{2^{s}M^{s}_{\infty}}{\lambda^{s}}\delta_{|m-u|=1},\,\,\,\psi(m)=\frac{2^{s}M^{s}_{\infty}}{\lambda^{s}}\delta_{m,n}. (II.33)

In the regime of weak disorder/extreme energies of Theorem 11, Lemma 4 can be applied to

K(m,u)=Ds,1|λ|s|G0(m,u;z)|s,ψ(m)=|G0(m,n;z)|sK(m,u)=D_{s,1}\absolutevalue{\lambda}^{s}\absolutevalue{G_{0}(m,u;z)}^{s},\,\,\,\psi(m)=\absolutevalue{G_{0}(m,n;z)}^{s} (II.34)

thanks to Lemma 9 below which implies a decoupling estimate for the Green’s function fractional moments cf. (A-W-B, , Theorems 8.7 and 10.4)

The remainder of this note is organized as follows: in Section III we show the quasilocality bounds of Steps 1 and 2 above, in Section IV we show the cancellation bound of Step 3, in Section V we state and prove the technical Lemmas on the conditional densities ρn0eff\rho^{{\mathrm{eff}}}_{n_{0}}.The proofs of Theorems 1 and 3 as well as Corollary 2 are given in Sections VI and VII. In the Appendix we provide some basic facts about existence of the effective potential and norm resolvent convergence of finite volume restrictions to the infinite volume operator.

III First order decay bounds on the effective potential

Let us collect some basic facts which will be repeatedly used in this note. Firstly, if Hω,LH_{\omega,L} is as above we can write F(Hω,L)=12πi(1Hω,L+tiη1Hω,L+t+iη)f(t)𝑑tF(H_{\omega,L})=\frac{1}{2\pi i}\int^{\infty}_{-\infty}\left(\frac{1}{H_{\omega,L}+t-i\eta}-\frac{1}{H_{\omega,L}+t+i\eta}\right)f(t)\,dt for f=F++F+DFf=F_{+}+F_{-}+D\ast F, where F±(u)=F(u±iηi0)F_{\pm}(u)=F(u\pm i\eta\mp i0) and D(u)=ηπ(η2+u2)D(u)=\frac{\eta}{\pi\left(\eta^{2}+u^{2}\right)} the Poisson kernel, see (A-G, , Appendix D). In particular, the inequality f3F\|f\|_{\infty}\leq 3\|F\|_{\infty} holds. The formula

Veff,ω,L(n)=12πiKL(n,ω;t)f(t)𝑑tV_{{\mathrm{eff}},\omega,L}(n)=\frac{1}{2\pi i}\int^{\infty}_{-\infty}K_{L}(n,\omega;t)f(t)\,dt (III.1)

with

KL(n,ω;t)=mda(n,m)(GL(m,m;tiη)GL(m,m;t+iη))K_{L}(n,\omega;t)=\sum_{m\in\mathbb{Z}^{d}}a(n,m)\left(G_{L}(m,m;t-i\eta)-G_{L}(m,m;t+i\eta)\right) (III.2)

readily follows and is a useful representation for the effective potential. It is shown below that it yields, for each n,lΛLn,l\in\Lambda_{L}, self-consistent equations for the derivatives Veff,ω,L(n)ω(l)\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)} which in turn imply the desired exponential decay in Step 1 of the proof strategies given earlier. We introduce ν>0\nu>0 such that

supnd|nn|=1eνd(n,n)<η/2.\sup_{n\in\mathbb{Z}^{d}}\sum_{|n^{\prime}-n|=1}e^{\nu d(n,n^{\prime})}<\eta/2. (III.3)

The decay rate in the Lemma below will be dictated by ν\nu and γa\gamma_{a}.

Lemma 5

Let ν>0\nu>0 be as in (III.3) and γa\gamma_{a} as in Assumption 5. For each LL\in\mathbb{N}, lΛL=[L,L]ddl\in\Lambda_{L}=[-L,L]^{d}\cap\mathbb{Z}^{d} and any δ<min{γa,2ν}\delta<\min\{\gamma_{a},2\nu\} the inequality

nΛeδd(n,l)|Veff,ω,L(n)ω(l)|C1\sum_{n\in\Lambda}e^{\delta d(n,l)}\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}}\leq C_{1} (III.4)

holds whenever Ca722Fη|g|SδγSδ2ν<12\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}|g|S_{\delta-\gamma}S_{\delta-2\nu}<\frac{1}{2}, with d:d×dd:\mathbb{Z}^{d}\times\mathbb{Z}^{d}\rightarrow\mathbb{R} as in assumption 5,

C1=λCa1442FηSδγSδ2ν,C_{1}=\lambda\frac{C_{a}144\sqrt{2}\norm{F}_{\infty}}{\eta}S_{\delta-\gamma}S_{\delta-2\nu}, (III.5)

and

Sβ:=supudvdeβd(u,v).S_{\beta}:=\sup_{u\in\mathbb{Z}^{d}}\sum_{v\in\mathbb{Z}^{d}}e^{\beta d(u,v)}. (III.6)

Proof.

Denote by Pl:2(d)Span{δl}P_{l}:\ell^{2}\left(\mathbb{Z}^{d}\right)\rightarrow\mathrm{Span}\{\delta_{l}\} the projection onto Span{δl}\mathrm{Span}\{\delta_{l}\}. Using difference quotients, it is immediate to check that

ω(l)1HLz=λ1HLzPl1HLzg1HLzVeff,ω,Lω(l)1HLz.\frac{\partial}{\partial\omega(l)}\frac{1}{H_{L}-z}=-\lambda\frac{1}{H_{L}-z}P_{l}\frac{1}{H_{L}-z}-g\frac{1}{H_{L}-z}\frac{\partial V_{{\mathrm{eff}},\omega,L}}{\partial\omega(l)}\frac{1}{H_{L}-z}. (III.7)

Taking matrix elements we obtain from (III.2) that

KL(n,ω;t)ω(l)=mΛLa(n,m)(λrL(m,l;t)gkΛLrL(m,k;t)Veff,ω,L(k)ω(l))\frac{\partial K_{L}(n,\omega;t)}{\partial\omega(l)}=\sum_{m\in\Lambda_{L}}a(n,m)\left(-\lambda r_{L}(m,l;t)-g\sum_{k\in\Lambda_{L}}r_{L}(m,k;t)\frac{\partial V_{{\mathrm{eff}},\omega,L}(k)}{\partial\omega(l)}\right) (III.8)

with

rL(u,v;t):=GL(u,v;tiη)GL(v,u;tiη)GL(u,v;t+iη)GL(v,u;t+iη).r_{L}(u,v;t):=G_{L}(u,v;t-i\eta)G_{L}(v,u;t-i\eta)-G_{L}(u,v;t+i\eta)G_{L}(v,u;t+i\eta). (III.9)

The above derivatives of the kernel KL(n,ω;t)K_{L}(n,\omega;t) are shown to decay exponentially in d(n,l)d(n,l) as follows. We first rewrite rL(u,v;t)r_{L}(u,v;t) as

rL(u,v;t)=\displaystyle r_{L}(u,v;t)= (GL(u,v;tiη)GL(u,v;t+iη))GL(v,u;tiη)\displaystyle\left(G_{L}(u,v;t-i\eta)-G_{L}(u,v;t+i\eta)\right)G_{L}(v,u;t-i\eta) (III.10)
+GL(u,v;t+iη)(GL(v,u;tiη)GL(v,u;t+iη)).\displaystyle+G_{L}(u,v;t+i\eta)\left(G_{L}(v,u;t-i\eta)-G_{L}(v,u;t+i\eta)\right).

For the operators studied here the Combes-Thomas bound (A-W-B, , Theorem 10.5) yields

|GL(u,v;z)|2ηeνd(u,v),z|G_{L}(u,v;z)|\leq\frac{2}{\eta}e^{-\nu d(u,v)},\,\,\,\,z\in\mathbb{C}\setminus\mathbb{R} (III.11)

for all ν>0\nu>0 satisfying (III.3). Moreover, by (A-G, , Appendix D, Lemma 3) we have the following inequality:

|GL(u,v;t+iη)GL(u,v;tiη)|12ηeνd(u,v)δu,1(HLt)2+η2/2δu1/2δv,1(HLt)2+η2/2δv1/2.|G_{L}(u,v;t+i\eta)-G_{L}(u,v;t-i\eta)|\leq 12\eta e^{-\nu d(u,v)}\langle\delta_{u},{\frac{1}{(H_{L}-t)^{2}+\eta^{2}/2}}\delta_{u}\rangle^{1/2}\langle\delta_{v},{\frac{1}{(H_{L}-t)^{2}+\eta^{2}/2}}\delta_{v}\rangle^{1/2}. (III.12)

We remark that the above result, as the usual Combes-Thomas bound, may also be applied to the metric dd instead of the usual metric of d\mathbb{Z}^{d}. One then obtains

|rL(u,v;t)|48e2νd(u,v)δu,1(HLt)2+η2/2δu1/2δv,1(HLt)2+η2/2δv1/2.\absolutevalue{r_{L}(u,v;t)}\leq 48e^{-2\nu d(u,v)}\langle\delta_{u},{\frac{1}{(H_{L}-t)^{2}+\eta^{2}/2}}\delta_{u}\rangle^{1/2}\langle\delta_{v},{\frac{1}{(H_{L}-t)^{2}+\eta^{2}/2}}\delta_{v}\rangle^{1/2}. (III.13)

By the spectral measure representation and the Cauchy-Schwarz inequality, the right-hand side of (III.13) can be controlled via

δu,1(HLt)2+η2/2δu1/2δv,1(HLt)2+η2/2δv1/2𝑑t2πη.\int^{\infty}_{-\infty}\langle\delta_{u},{\frac{1}{(H_{L}-t)^{2}+\eta^{2}/2}}\delta_{u}\rangle^{1/2}\langle\delta_{v},{\frac{1}{(H_{L}-t)^{2}+\eta^{2}/2}}\delta_{v}\rangle^{1/2}\,dt\leq\frac{\sqrt{2}\pi}{\eta}. (III.14)

Therefore,

12π|rL(u,v;t)f(t)|𝑑t722Fηe2νd(u,v).\frac{1}{2\pi}\int^{\infty}_{-\infty}\absolutevalue{r_{L}(u,v;t)f(t)}\,dt\leq\frac{72\sqrt{2}\norm{F}_{\infty}}{\eta}e^{-2\nu d(u,v)}. (III.15)

Keeping in mind assumption 5 and combining (III.1), (III.8) and (III.15) we reach the inequality

|Veff,ω,L(n)ω(l)|\displaystyle\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}} Ca722FηmΛLλeγad(n,m)2νd(m,l)\displaystyle\leq\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}\sum_{m\in\Lambda_{L}}\lambda e^{-\gamma_{a}d(n,m)-2\nu d(m,l)} (III.16)
+Ca722Fη|g|kΛLeγad(n,m)2νd(m,k)|Veff,ω,L(k)ω(l)|.\displaystyle+\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}|g|\sum_{k\in\Lambda_{L}}e^{-\gamma_{a}d(n,m)-2\nu d(m,k)}\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(k)}{\partial\omega(l)}}.

We now apply Lemma 4 with fixed lΛLl\in\Lambda_{L} and the choices φ(n)=|Veff,ω,L(n)ω(l)|\varphi(n)=\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}},

ψ(n)=Ca722FηλmΛLeγad(n,m)2νd(m,l)\psi(n)=\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}\lambda\sum_{m\in\Lambda_{L}}e^{-\gamma_{a}d(n,m)-2\nu d(m,l)} (III.17)
K(n,u)=Ca722Fη|g|mΛLeγad(n,m)2νd(m,u)K(n,u)=\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}|g|\sum_{m\in\Lambda_{L}}e^{-\gamma_{a}d(n,m)-2\nu d(m,u)} (III.18)

in the regime where K,<1\norm{K}_{\infty,\infty}<1, i.e. when

Ca722Fη|g|SγS2ν<1.\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}|g|S_{-\gamma}S_{-2\nu}<1. (III.19)

In this context, introducing the weight function W(n)=eδd(n,l)W(n)=e^{\delta d(n,l)} with δ<min{γa,2ν}\delta<\min\{\gamma_{a},2\nu\} we reach

b1:=ndW(n)ψ(n)Ca722FηλSδγSδ2νb_{1}:=\sum_{n\in\mathbb{Z}^{d}}W(n)\psi(n)\leq\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}\lambda S_{\delta-\gamma}S_{\delta-2\nu} (III.20)

and

b2=supndndW(n)W(n)K(n,n)Ca722Fη|g|SδγSδ2ν.b_{2}=\sup_{n^{\prime}\in\mathbb{Z}^{d}}\sum_{n\in\mathbb{Z}^{d}}\frac{W(n)}{W(n^{\prime})}K(n,n^{\prime})\leq\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}|g|S_{\delta-\gamma}S_{\delta-2\nu}. (III.21)

In particular, under the more restrictive assumption

Ca722Fη|g|SδγSδ2ν<12\frac{C_{a}72\sqrt{2}\norm{F}_{\infty}}{\eta}|g|S_{\delta-\gamma}S_{\delta-2\nu}<\frac{1}{2} (III.22)

we find that 11b22\frac{1}{1-b_{2}}\leq 2 and thus φ2b1\varphi\leq 2b_{1}, finishing the proof.

Given an enumeration n1,,ω(n|Λ|)n_{1},\ldots,\omega(n_{\absolutevalue{\Lambda^{\prime}}}) of the points in Λ\Lambda^{\prime}, it readily follows that within the smallness regime described in Lemma 5, the map 𝒯:|Λ||Λ|\mathcal{T}:\mathbb{R}^{\absolutevalue{\Lambda^{\prime}}}\rightarrow\mathbb{R}^{\absolutevalue{\Lambda^{\prime}}} given by

𝒯(ω(n1),,ω(n|Λ|))=(ULΛ(n1),,ULΛ(n|Λ|)),U(n):=ω(n)+gλVeff,ω,L(n).\mathcal{T}(\omega(n_{1}),\ldots,\omega(n_{\absolutevalue{\Lambda^{\prime}}}))=(U^{\Lambda^{\prime}}_{L}(n_{1}),\ldots,U^{\Lambda^{\prime}}_{L}(n_{\absolutevalue{\Lambda^{\prime}}})),\,\,\,U(n):=\omega(n)+\frac{g}{\lambda}V_{{\mathrm{eff}},\omega,L}(n). (III.23)

is a diffeomorphism.

We are now ready to quantify the change in ω\omega after resampling. Fix n0Λn_{0}\in\Lambda^{\prime} and define Uα,LΛ(n)=U(n)+(αU(n0))δn0U^{\Lambda^{\prime}}_{\alpha,L}(n)=U(n)+(\alpha-U(n_{0}))\delta_{n_{0}} for nΛn\in\Lambda^{\prime}. Then, Uα,LΛU^{\Lambda^{\prime}}_{\alpha,L} is interpreted as the “full" potential in Λ\Lambda^{\prime} with value changed to α\alpha at n0n_{0}. Denote by {ωα,LΛ(n)}nΛ\{\omega^{\Lambda^{\prime}}_{\alpha,L}(n)\}_{n\in\Lambda^{\prime}} the random variables for which Uα(n)=ωα(n)+Veff,ωα,LΛ,L(n)U_{\alpha}(n)=\omega_{\alpha}(n)+V_{{\mathrm{eff}},\omega^{\Lambda^{\prime}}_{\alpha,L},L}(n). In this setting we have the quasilocality result below.

Lemma 6

Let C1C_{1} be as in (III.5). Whenever b2=|g|λC1<1/2b_{2}=\frac{|g|}{\lambda}C_{1}<1/2 and δ<min{γa,2ν}\delta<\min\{\gamma_{a},2\nu\} we have

nΛ{n0}eδd(n,n0)|ωα,LΛ(n)ω(n)|2|g|C1λ(|αU(n0)|+2|g|Veffλ).\sum_{n\in\Lambda^{\prime}\setminus\{n_{0}\}}e^{\delta d(n,n_{0})}\absolutevalue{\omega^{\Lambda^{\prime}}_{\alpha,L}(n)-\omega(n)}\leq\frac{2|g|C_{1}}{\lambda}\left(|\alpha-U(n_{0})|+2\frac{|g|\norm{V_{{\mathrm{eff}}}}_{\infty}}{\lambda}\right). (III.24)

Proof. For simplicity we denote ωα,LΛ\omega^{\Lambda^{\prime}}_{\alpha,L} by ωα\omega_{\alpha} in this proof. Observe that there exists ω^α={ω^α}nΛ\hat{\omega}_{\alpha}=\{\hat{\omega}_{\alpha}\}_{n\in\Lambda^{\prime}} with ω^α(n)(ω(n),ωα(n))\hat{\omega}_{\alpha}(n)\in(\omega(n),\omega_{\alpha}(n)) such that for each nΛ{n0}n\in\Lambda^{\prime}\setminus\{n_{0}\}.

|ωα(n)ω(n)|\displaystyle\absolutevalue{\omega_{\alpha}(n)-\omega(n)} =|g|λ|Veff,ωα,L(n)Veff,ω,L(n)|\displaystyle=\frac{|g|}{\lambda}\absolutevalue{V_{{\mathrm{eff}},\omega_{\alpha},L}(n)-V_{{\mathrm{eff}},\omega,L}(n)}
|g|λ|Veff(n,ω^α)ω(n0)|(|αU(n0)|+|g|λ|Veff,ωα,L(n0)Veff,ω,L(n0)|)\displaystyle\leq\frac{|g|}{\lambda}\absolutevalue{\frac{\partial V_{{\mathrm{eff}}}(n,\hat{\omega}_{\alpha})}{\partial\omega(n_{0})}}\left(|\alpha-U(n_{0})|+\frac{|g|}{\lambda}\absolutevalue{V_{{\mathrm{eff}},\omega_{\alpha},L}(n_{0})-V_{{\mathrm{eff}},\omega,L}(n_{0})}\right)
+lΛ{n0}|g|λ|Veff(n,ω^α)ω(l)||ωα(l)ω(l)|.\displaystyle+\sum_{l\in\Lambda^{\prime}\setminus\{n_{0}\}}\frac{|g|}{\lambda}\absolutevalue{\frac{\partial V_{{\mathrm{eff}}}(n,\hat{\omega}_{\alpha})}{\partial\omega(l)}}\absolutevalue{\omega_{\alpha}(l)-\omega(l)}.

Thanks to Lemma 5, whenever |g|C1λ<12\frac{\absolutevalue{g}C_{1}}{\lambda}<\frac{1}{2} we can apply Lemma 4 with the choices φ(n)=|ωα(n)ω(n)|δnn0\varphi(n)=\absolutevalue{\omega_{\alpha}(n)-\omega(n)}\delta_{n\neq n_{0}}, W(n)=eδd(n,n0)W(n)=e^{\delta d(n,n_{0})},

ψ(n)=|g|λ|Veff(n,ω^α)ω(n0)|(|αU(n0)|+|g|λ|Veff,ωα,L(n0)Veff,ω,L(n0)|)\psi(n)=\frac{|g|}{\lambda}\absolutevalue{\frac{\partial V_{{\mathrm{eff}}}(n,\hat{\omega}_{\alpha})}{\partial\omega(n_{0})}}\left(|\alpha-U(n_{0})|+\frac{|g|}{\lambda}\absolutevalue{V_{{\mathrm{eff}},\omega_{\alpha},L}(n_{0})-V_{{\mathrm{eff}},\omega,L}(n_{0})}\right) (III.25)

and

K(n,u)=|g|λ|Veff(n,ω^α)ω(u)|,K(n,u)=\frac{|g|}{\lambda}\absolutevalue{\frac{\partial V_{{\mathrm{eff}}}(n,\hat{\omega}_{\alpha})}{\partial\omega(u)}}, (III.26)

finishing the proof.

IV Second order decay bounds on the effective potential

This section is devoted to the cancellation bounds of Step 3 of the proof outline. From now on throughout the paper we denote by ν\nu any positive number satisfying (III.3).

Lemma 7

Let LL\in\mathbb{N}. Whenever |g|Ca722FηSδδ02Sδ02<12\absolutevalue{g}C_{a}\frac{72\sqrt{2}\norm{F}_{\infty}}{\eta}S_{\frac{\delta-\delta_{0}}{2}}S_{-\frac{\delta_{0}}{2}}<\frac{1}{2} we have, for each δ<δ0:=min{ν,γa}\delta<\delta_{0}:=\min\{\nu,\gamma_{a}\} and lΛLl\in\Lambda_{L},

|g|λnΛLeδ2d(n,l)|Veff,ω,L(n)ω(l)Veff,ωα,L(n)ω(l)|C2|αU(n0)|eδ2d(n0,l)\frac{\absolutevalue{g}}{\lambda}\sum_{n\in\Lambda_{L}}e^{\frac{\delta}{2}d(n,l)}\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(n)}{\partial\omega(l)}}\leq C_{2}\absolutevalue{\alpha-U(n_{0})}e^{-\frac{\delta}{2}d(n_{0},l)} (IV.1)

with

C2=48FCaη2Sδδ02Sν(λ|g|+|g|2C1)C_{2}=\frac{48\norm{F}_{\infty}C_{a}}{\eta^{2}}S_{\frac{\delta-\delta_{0}}{2}}S_{-\nu}(\lambda|g|+|g|^{2}C_{1}) (IV.2)

and C1C_{1} as in (III.5).

Proof.

By (III.8) we find that if n,lΛLn,l\in\Lambda_{L}

Veff,ω,L(n)ω(l)Veff,ωα,L(n)ω(l)\displaystyle\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(n)}{\partial\omega(l)} =λmΛLa(n,m)(rL(m,l)rLα(m,l))\displaystyle=-\lambda\sum_{m\in\Lambda_{L}}a(n,m)(r_{L}(m,l)-r^{\alpha}_{L}(m,l))
gmΛLa(n,m)kΛL(rL(m,k)rLα(m,k))Veff,ω,L(k)ω(l)\displaystyle-g\sum_{m\in\Lambda_{L}}a(n,m)\sum_{k\in\Lambda_{L}}(r_{L}(m,k)-r^{\alpha}_{L}(m,k))\frac{\partial V_{{\mathrm{eff}},\omega,L}(k)}{\partial\omega(l)}
gmΛLa(n,m)kΛLrLα(m,k)(Veff,ω,L(k)ω(l)Veff,ωα,L(k)ω(l))\displaystyle-g\sum_{m\in\Lambda_{L}}a(n,m)\sum_{k\in\Lambda_{L}}r^{\alpha}_{L}(m,k)\left(\frac{\partial V_{{\mathrm{eff}},\omega,L}(k)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(k)}{{\partial\omega(l)}}\right)

where

rL(u,v)=12πirL(u,v;t)f(t)𝑑t,r_{L}(u,v)=\frac{1}{2\pi i}\int^{\infty}_{-\infty}r_{L}(u,v;t)f(t)\,dt, (IV.3)

rL(u,v;t)r_{L}(u,v;t) as in (III.10) and rLα(u,v)r^{\alpha}_{L}(u,v) similarly defined with ω\omega replaced by ωα,L\omega_{\alpha,L}.

With these definitions, letting z=tiηz=t-i\eta we reach

|rL(m,k;t)rLα(m,k;t)|\displaystyle\absolutevalue{r_{L}(m,k;t)-r^{\alpha}_{L}(m,k;t)} |GL(m,k;z)GLα(m,k;z)||GL(k,m;z)|\displaystyle\leq\absolutevalue{G_{L}(m,k;z)-G^{\alpha}_{L}(m,k;z)}\absolutevalue{G_{L}(k,m;z)}
+|GL(k,m;z)GLα(k,m;z)||GLα(m,k;z)|\displaystyle+\absolutevalue{G_{L}(k,m;z)-G^{\alpha}_{L}(k,m;z)}\absolutevalue{G^{\alpha}_{L}(m,k;z)}
+|GL(m,k;z¯)GLα(m,k;z¯)||GL(k,m;z¯)|\displaystyle+\absolutevalue{G_{L}(m,k;\bar{z})-G^{\alpha}_{L}(m,k;\bar{z})}\absolutevalue{G_{L}(k,m;\bar{z})}
+|GL(k,m;z¯)GLα(k,m;z¯)||GLα(m,k;z¯)|.\displaystyle+\absolutevalue{G_{L}(k,m;\bar{z})-G^{\alpha}_{L}(k,m;\bar{z})}\absolutevalue{G^{\alpha}_{L}(m,k;\bar{z})}.

Note that by definition of ωα\omega_{\alpha} we have that

|GL(m,k;z)GLα(m,k;z)|=λ|αU(n0)||GL(m,n0;z)||GLα(n0,k;z)|\absolutevalue{G_{L}(m,k;z)-G^{\alpha}_{L}(m,k;z)}=\lambda\absolutevalue{\alpha-U(n_{0})}\absolutevalue{G_{L}(m,n_{0};z)}{\absolutevalue{G^{\alpha}_{L}(n_{0},k;z)}} (IV.4)

for all m,kΛLm,k\in\Lambda_{L}. In particular

|rL(m,l)rLα(m,l)|λ|αU(n0)|242Fη2eν(d(m,l)+d(m,n0)+d(n0,l)).\absolutevalue{r_{L}(m,l)-r^{\alpha}_{L}(m,l)}\leq\lambda\absolutevalue{\alpha-U(n_{0})}\frac{24\sqrt{2}\norm{F}_{\infty}}{\eta^{2}}e^{-\nu\left(d(m,l)+d(m,n_{0})+d(n_{0},l)\right)}. (IV.5)

Indeed, (IV.5) follows from (IV.4) and a similar argument to the one in (III.14) with the help of the following Combes-Thomas type bound cf. (M-S, , Lemma 18)

|GL(u,v;t±iη)|2δv,1(HLt)2+η2/2δv1/2eνd(u,v)|G_{L}(u,v;t\pm i\eta)|\leq\sqrt{2}\langle\delta_{v},\frac{1}{(H_{L}-t)^{2}+\eta^{2}/2}\delta_{v}\rangle^{1/2}e^{-\nu d(u,v)} (IV.6)

applied separately to |GL(m,n0;z)|\absolutevalue{G_{L}(m,n_{0};z)} and |GLα(n0,k;z)|{\absolutevalue{G^{\alpha}_{L}(n_{0},k;z)}}.

Thus, assumption 1, (III.15) and (IV.5) imply

|Veff,ω,L(n)ω(l)Veff,ωα,L(n)ω(l)|λ2|αU(n0)|242FCaη2eνd(n0,l)mdeγd(m,n)ν(d(m,l)+d(m,n0))\displaystyle\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(n)}{\partial\omega(l)}}\leq\lambda^{2}\absolutevalue{\alpha-U(n_{0})}\frac{24\sqrt{2}\norm{F}_{\infty}C_{a}}{\eta^{2}}e^{-\nu d(n_{0},l)}\sum_{m\in\mathbb{Z}^{d}}e^{-\gamma d(m,n)-\nu(d(m,l)+d(m,n_{0}))}
+|g|λ|αU(n0)|242FCaη2mdkdeγd(m,n)ν(d(m,k)+d(m,n0)+d(n0,k))|Veff,ω,L(k)ω(l)|\displaystyle+\absolutevalue{g}\lambda\absolutevalue{\alpha-U(n_{0})}\frac{24\sqrt{2}\norm{F}_{\infty}C_{a}}{\eta^{2}}\sum_{m\in\mathbb{Z}^{d}}\sum_{k\in\mathbb{Z}^{d}}e^{-\gamma d(m,n)-\nu\left(d(m,k)+d(m,n_{0})+d(n_{0},k)\right)}\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(k)}{\partial\omega(l)}}
+|g|Ca722Fηmdkdeγd(m,n)2νd(m,k)|Veff,ω,L(k)ω(l)Veff,ωα,L(k)ω(l)|.\displaystyle+\absolutevalue{g}C_{a}\frac{72\sqrt{2}\norm{F}_{\infty}}{\eta}\sum_{m\in\mathbb{Z}^{d}}\sum_{k\in\mathbb{Z}^{d}}e^{-\gamma d(m,n)-2\nu d(m,k)}\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(k)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(k)}{{\partial\omega(l)}}}.

Thus, if δ0=min{ν,γa}\delta_{0}=\min\{\nu,\gamma_{a}\}, δ<δ0\delta<\delta_{0} and C1C_{1} is as in (III.5)

|Veff,ω,L(n)ω(l)Veff,ωα,L(n)ω(l)|λ2|αU(n0)|242FCaη2eνd(n0,l)eδ0d(n,l)Sν\displaystyle\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(n)}{\partial\omega(l)}}\leq\lambda^{2}\absolutevalue{\alpha-U(n_{0})}\frac{24\sqrt{2}\norm{F}_{\infty}C_{a}}{\eta^{2}}e^{-\nu d(n_{0},l)}e^{-\delta_{0}d(n,l)}S_{-\nu}
+|g|λ|αU(n0)|242FCaη2C1Sδ02eδ2(d(n0,l)+d(n,l))\displaystyle+\absolutevalue{g}\lambda\absolutevalue{\alpha-U(n_{0})}\frac{24\sqrt{2}\norm{F}_{\infty}C_{a}}{\eta^{2}}C_{1}S_{-\frac{\delta_{0}}{2}}e^{-\frac{\delta}{2}(d(n_{0},l)+d(n,l))}
+|g|Ca722FηSδ02kdeδ02d(n,k)|Veff,ω,L(k)ω(l)Veff,ωα,L(k)ω(l)|.\displaystyle+\absolutevalue{g}C_{a}\frac{72\sqrt{2}\norm{F}_{\infty}}{\eta}S_{-\frac{\delta_{0}}{2}}\sum_{k\in\mathbb{Z}^{d}}e^{-\frac{\delta_{0}}{2}d(n,k)}\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(k)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(k)}{{\partial\omega(l)}}}.

In particular, if b2=|g|Ca722FηSδδ02Sδ02<12b_{2}=\absolutevalue{g}C_{a}\frac{72\sqrt{2}\norm{F}_{\infty}}{\eta}S_{\frac{\delta-\delta_{0}}{2}}S_{-\frac{\delta_{0}}{2}}<\frac{1}{2} another application of Lemma 4 yields

ndeδ2d(n,l)|Veff,ω,L(n)ω(l)Veff,ωα,L(n)ω(l)|\displaystyle\sum_{n\in\mathbb{Z}^{d}}e^{\frac{\delta}{2}d(n,l)}\absolutevalue{\frac{\partial V_{{\mathrm{eff}},\omega,L}(n)}{\partial\omega(l)}-\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(n)}{\partial\omega(l)}}
λ2|αU(n0)|48FCaη2eνd(n0,l)Sδδ02Sν\displaystyle\leq\lambda^{2}\absolutevalue{\alpha-U(n_{0})}\frac{48\norm{F}_{\infty}C_{a}}{\eta^{2}}e^{-\nu d(n_{0},l)}S_{\frac{\delta-\delta_{0}}{2}}S_{-\nu}
+|g|48FCaη2λ|αU(n0)|eδ2d(n0,l)Sδ02Sδ0δ2C1\displaystyle+|g|\frac{48\norm{F}_{\infty}C_{a}}{\eta^{2}}\lambda\absolutevalue{\alpha-U(n_{0})}e^{-\frac{\delta}{2}d(n_{0},l)}S_{-\frac{\delta_{0}}{2}}S_{\frac{\delta_{0}-\delta}{2}}C_{1}
λ|αU(n0)|48FCaη2Sδδ02Sν(λ+|g|C1)eδ2d(n0,l)\displaystyle\leq\lambda\absolutevalue{\alpha-U(n_{0})}\frac{48\norm{F}_{\infty}C_{a}}{\eta^{2}}S_{\frac{\delta-\delta_{0}}{2}}S_{-\nu}(\lambda+|g|C_{1})e^{-\frac{\delta}{2}d(n_{0},l)}

with C1C_{1} as in (III.5).

V A pair of technical lemmas

Fix LL\in\mathbb{N} and ΛΛL\Lambda^{\prime}\subset\Lambda_{L}. Recall that in (III.23) we have denoted U(n)=ω(n)+gλVeff,ω,L(n)U(n)=\omega(n)+\frac{g}{\lambda}V_{{\mathrm{eff}},\omega,L}(n) for each nΛn\in\Lambda^{\prime} with Veff,ω,LV_{{\mathrm{eff}},\omega,L} given by (II.8). We also write 𝒯:|Λ||Λ|\mathcal{T}:\mathbb{R}^{|\Lambda^{\prime}|}\rightarrow\mathbb{R}^{|\Lambda^{\prime}|} the above change of variables, i.e

𝒯(ω(n1),,ω(n|Λ|))=(U(n1),,U(n|Λ|)).\mathcal{T}(\omega(n_{1}),\ldots,\omega(n_{|\Lambda^{\prime}|}))=(U(n_{1}),\ldots,U(n_{|\Lambda^{\prime}|})). (V.1)

In the sequel we will abbreviate this by writing

𝒯ω=Uorω=𝒯1U.\mathcal{T}\omega=U\,\,\text{or}\,\,\omega={\mathcal{T}}^{-1}U.

The first result on uniform control of the conditional density of U(n0)U(n_{0}) is given below.

Lemma 8

Under assumptions 1-6 whenever |g|Ca722FηSδγSδν<12\absolutevalue{g}C_{a}\frac{72\sqrt{2}\norm{F}_{\infty}}{\eta}S_{\delta-\gamma}S_{\delta-\nu}<\frac{1}{2} for some δ<δ0:=min{γa,ν}\delta<\delta_{0}:=\min\{\gamma_{a},\nu\} the conditional distribution of U(n0)=vU(n_{0})=v at specified values of {U(n)}nΛ{n0}\{U(n)\}_{n\in\Lambda^{\prime}\setminus\{n_{0}\}} has a density ρn0,Leff,Λ(v)\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v). Moreover, ρn0,Leff,Λ(v)\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v) is bounded:

M:=supωΩsupLsupΛΛLsupn0Λsupvρn0,Leff,Λ(v)<.M_{\infty}:=\sup_{\omega\in\Omega}\sup_{L\in\mathbb{N}}\sup_{\Lambda^{\prime}\subset\Lambda_{L}}\sup_{n_{0}\in\Lambda}\sup_{v\in\mathbb{R}}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)<\infty. (V.2)

Proof. We note that in the above setting ρn0,Leff,Λ\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L} is given by

ρn0,Leff,Λ(v)=ρ(vgλVeff,𝒯1U,L(n0))nΛ{n0}ρ(U(n)gλVeff,𝒯1U,L(n))JUρ(αgλVeff,𝒯1Uα,L(n0))nΛ{n0}ρ(Uα(n)gλVeff,𝒯1Uα,L(n))JUαdα\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)=\frac{\rho\left(v-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U,L}(n_{0})\right)\prod_{n\in\Lambda^{\prime}\setminus\{n_{0}\}}\rho\left(U(n)-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U,L}(n)\right)J_{U}}{\int^{\infty}_{-\infty}\rho\left(\alpha-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U_{\alpha},L}(n_{0})\right)\prod_{n\in\Lambda^{\prime}\setminus\{n_{0}\}}\rho\left(U^{\alpha}(n)-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U_{\alpha},L}(n)\right)J_{U_{\alpha}}\,d\alpha} (V.3)

Where

JU=det(()IgλVeff,𝒯1U,L(ni)U(nj))|Λ|×|Λ|JUα=det(()IgλVeff,𝒯1Uα,L(ni)U(nj))|Λ|×|Λ|J_{U}=\det\Big{(}I-\frac{g}{\lambda}\frac{\partial V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U,L}(n_{i})}{\partial U(n_{j})}\Big{)}_{|\Lambda^{\prime}|\times|\Lambda^{\prime}|}\,\,\,J_{U_{\alpha}}=\det\Big{(}I-\frac{g}{\lambda}\frac{\partial V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U_{\alpha},L}(n_{i})}{\partial U(n_{j})}\Big{)}_{|\Lambda^{\prime}|\times|\Lambda^{\prime}|} (V.4)

and we recall that

Uα(n):=U(n)+(αU(n0))δn=n0.U^{\alpha}(n):=U(n)+\left(\alpha-U(n_{0})\right)\delta_{n=n_{0}}.

Letting A=gλ(Veff,ω,L(ni)ω(nj))|Λ|×|Λ|A=-\frac{g}{\lambda}\left(\frac{\partial V_{{\mathrm{eff}},\omega,L}(n_{i})}{\partial\omega(n_{j})}\right)_{|\Lambda^{\prime}|\times|\Lambda^{\prime}|} and B=gλ(Veff,ωα,L(ni)ω(nj))|Λ|×|Λ|B=-\frac{g}{\lambda}\left(\frac{\partial V_{{\mathrm{eff}},\omega_{\alpha},L}(n_{i})}{\partial\omega(n_{j})}\right)_{|\Lambda^{\prime}|\times|\Lambda^{\prime}|} one has that

em,nΛ|((AB)(I+B)1)(m,n)||det(I+B)det(I+A)|em,nΛ|((BA)(I+A)1)(m,n)|.e^{-\sum_{m,n\in\Lambda^{\prime}}\absolutevalue{\left((A-B)(I+B)^{-1}\right)(m,n)}}\leq\absolutevalue{\frac{\det(I+B)}{\det(I+A)}}\leq e^{\sum_{m,n\in\Lambda^{\prime}}\absolutevalue{\left((B-A)(I+A)^{-1}\right)(m,n)}}. (V.5)

Indeed, (V.5) follows from the inequality det(I+M)eM1\det(I+M)\leq e^{\norm{M}_{1}} (c.f (Simon-trace, , Lemma 3.3)), see (M-S, , Lemma 22). We remark that it suffices to control ratios of the above determinants instead of the ones in (V.4) since the later arise from the inverse change of variables 𝒯1U=ω{\mathcal{T}}^{-1}U=\omega.

We are now ready to estimate the right-hand side of (V.5). Using Lemma 5 we see that whenever |g|λC1<14\frac{|g|}{\lambda}C_{1}<\frac{1}{4} we have that

B,:=supnΛlΛeδd(n,l)|B(n,l)|<14\norm{B}_{\infty,\infty}:=\sup_{n\in\Lambda^{\prime}}\sum_{l\in\Lambda^{\prime}}e^{\delta d(n,l)}\absolutevalue{B(n,l)}<\frac{1}{4} (V.6)

thus

|(I+B)1(n,l)|<4eδd(n,l)\absolutevalue{(I+B)^{-1}(n,l)}<4e^{-\delta d(n,l)} (V.7)

by the Combes-Thomas bound. Using Lemma 7 and the inequalities (V.5) and (V.7) we find that

e4C2Sδ22|αU(n0)|detJUαdetJUe4C2Sδ22|αU(n0)|.e^{-4C_{2}S^{2}_{-\frac{\delta}{2}}|\alpha-U(n_{0})|}\leq\frac{\det J_{U_{\alpha}}}{\det J_{U}}\leq e^{4C_{2}S^{2}_{-\frac{\delta}{2}}|\alpha-U(n_{0})|}. (V.8)

For each nn0n\neq n_{0}, writing ωα(n)=Uα(n)gλVeff,ω,Lα(n)\omega_{\alpha}(n)=U^{\alpha}(n)-\frac{g}{\lambda}V^{\alpha}_{{\mathrm{eff}},\omega,L}(n), one concludes from assumption 6 that

ec1|ωα(n)ω(n)|ρ(ωα(n))ρ(ω(n))ec1|ωα(n)ω(n)|.e^{-c_{1}\absolutevalue{\omega_{\alpha}(n)-\omega(n)}}\leq\frac{\rho(\omega_{\alpha}(n))}{\rho(\omega(n))}\leq e^{c_{1}\absolutevalue{\omega_{\alpha}(n)-\omega(n)}}. (V.9)

By Lemma 6 it then follows that for δ<min{γa,ν}\delta<\min\{\gamma_{a},\nu\}

e2|g|λc1C1((|αU(n0)|+2|g|λVeff)nn0ρ(ωα(n))ρ(ω(n))e2|g|λc1C1((|αU(n0)|+2|g|λVeff).e^{-2\frac{\absolutevalue{g}}{\lambda}c_{1}C_{1}\left((|\alpha-U(n_{0})|+2\frac{\absolutevalue{g}}{\lambda}\norm{V_{{\mathrm{eff}}}}_{\infty}\right)}\leq\prod_{n\neq n_{0}}\frac{\rho(\omega_{\alpha}(n))}{\rho(\omega(n))}\leq e^{2\frac{\absolutevalue{g}}{\lambda}c_{1}C_{1}\left((|\alpha-U(n_{0})|+2\frac{\absolutevalue{g}}{\lambda}\norm{V_{{\mathrm{eff}}}}_{\infty}\right)}. (V.10)

In particular, under assumptions 1-6 for each fixed λ\lambda we obtain for |g|F\absolutevalue{g}\norm{F}_{\infty} sufficiently small that if

ϑ=(2c1|g|λ+4c1C1|g|2λ2)Sγa182F\vartheta=(2c_{1}\frac{\absolutevalue{g}}{\lambda}+4c_{1}C_{1}\frac{\absolutevalue{g}^{2}}{\lambda^{2}})S_{-\gamma_{a}}18\sqrt{2}\norm{F}_{\infty} (V.11)

then

supvρn0,Leff,Λ(v)eϑsupvρ(v)ρ(α)eε1|vα|𝑑α<,\sup_{v\in\mathbb{R}}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\leq e^{\vartheta}\sup_{v\in\mathbb{R}}\frac{\rho(v)}{\int^{\infty}_{-\infty}\rho(\alpha)e^{-\varepsilon_{1}\absolutevalue{v-\alpha}}\,d\alpha}<\infty, (V.12)

finishing the proof of Lemma 8.

Now we shall see that under assumption 7 one may achieve a better control on the conditional densities.

Lemma 9

Under assumptions 1-7 there exits ε>0\varepsilon>0, ϑ=ϑ(F,g,λ,η0,γa,ν,ρ)\vartheta=\vartheta(\norm{F},g,\lambda,\eta_{0},\gamma_{a},\nu,\rho) and g1=g1(λ,c1,γ,ν,η0)g_{1}=g_{1}(\lambda,c_{1},\gamma,\nu,\eta_{0}) independent of Λ\Lambda^{\prime} and LL such that if |g|F<g1\absolutevalue{g}\norm{F}_{\infty}<g_{1} then

  1. (i)
    eϑ(cρε2)ecρε)|v|ρn0,Leff,Λ(v)eϑ(cρε2)e(cρ+ε)|v|.e^{-\vartheta}\left(\frac{c_{\rho}-\varepsilon}{2}\right)e^{-c_{\rho}-\varepsilon)|v|}\leq\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\leq e^{\vartheta}\left(\frac{c_{\rho}-\varepsilon}{2}\right)e^{(-c_{\rho}+\varepsilon)\absolutevalue{v}}. (V.13)
  2. (ii)
    ec1(1ϑ)|vv|ρn0,Leff,Λ(v)ρn0,Leff,Λ(v)ec1(1+ϑ)|vv|e^{-c_{1}(1-\vartheta)\absolutevalue{v-v^{\prime}}}\leq\frac{\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)}{\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v^{\prime})}\leq e^{c_{1}(1+\vartheta)\absolutevalue{v-v^{\prime}}} (V.14)

Moreover, ϑ0\vartheta\to 0 as |g|F0\absolutevalue{g}\norm{F}_{\infty}\to 0.

Proof. To reach the upper bound we follow most of the proof of Lemma 8, obtaining improvements at the very end with help of assumption 7. Observe that, with the choice δ<min{γa,ν}\delta<\min\{\gamma_{a},\nu\}, equations (V.3)-(V.8) imply the pointwise bound

ρn0,Leff,Λ(v)ρ(vgλVeff,𝒯1U,L(n0))ρ(αgλVeff,𝒯1Uα,L)e2|g|λc1C1((|αv|+2|g|λVeff,𝒯1U,L)e4C2Sδ22|αv|𝑑α\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\leq\frac{\rho(v-\frac{g}{\lambda}V_{{\mathrm{eff}},\mathcal{T}^{-1}U,L}(n_{0}))}{\int^{\infty}_{-\infty}\rho(\alpha-\frac{g}{\lambda}V_{{\mathrm{eff}},\mathcal{T}^{-1}U_{\alpha},L})e^{-2\frac{\absolutevalue{g}}{\lambda}c_{1}C_{1}\left((|\alpha-v|+2\frac{\absolutevalue{g}}{\lambda}\norm{V_{{\mathrm{eff}},\mathcal{T}^{-1}U,L}}_{\infty}\right)}e^{-4C_{2}S^{2}_{-\frac{\delta}{2}}|\alpha-v|}\,d\alpha} (V.15)

where we recall that C1C_{1} is given in (III.5) and is independent of |g|\absolutevalue{g}. The constant C2C_{2} is given in (IV.2) and is proportional to |g|\absolutevalue{g} when this number is sufficiently small. Note that by assumption 6 we have for any tt\in\mathbb{R} and n0Λn_{0}\in\Lambda^{\prime}:

ec1|g|λVeff,𝒯1U,Lρ(tgλVeff,𝒯1U,L(n0))ρ(t)ec1|g|λVeff,𝒯1U,L.e^{-c_{1}\frac{\absolutevalue{g}}{\lambda}\norm{V_{{\mathrm{eff}},\mathcal{T}^{-1}U,L}}_{\infty}}\leq\frac{\rho(t-\frac{g}{\lambda}V_{{\mathrm{eff}},\mathcal{T}^{-1}U,L}(n_{0}))}{\rho(t)}\leq e^{c_{1}\frac{\absolutevalue{g}}{\lambda}\norm{V_{{\mathrm{eff}},\mathcal{T}^{-1}U,L}}_{\infty}}. (V.16)

Hence from (V.15)

ρn0,Leff,Λ(v)e2c1(1+C12|g|λ)|g|λVeff,𝒯1U,Lρ(v)ρ(α)eθ|αU(n0)|𝑑α\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\leq e^{2c_{1}(1+C_{1}\frac{2\absolutevalue{g}}{\lambda})\frac{\absolutevalue{g}}{\lambda}\norm{V_{{\mathrm{eff}},\mathcal{T}^{-1}U,L}}_{\infty}}\frac{\rho(v)}{\int^{\infty}_{-\infty}\rho(\alpha)e^{-\theta|\alpha-U(n_{0})|}\,d\alpha} (V.17)

with θ=2|g|λc1C1+4C2Sδ22\theta=2\frac{\absolutevalue{g}}{\lambda}c_{1}C_{1}+4C_{2}S^{2}_{-\frac{\delta}{2}}. Now we make use of assumption 7 to write ρ(v)ρ(α)=h(v)h(α)ecρ(|v||α|)\frac{\rho(v)}{\rho(\alpha)}=\frac{h(v)}{h(\alpha)}e^{-c_{\rho}(|v|-|\alpha|)} with

eε2|vα|h(v)h(α)eε2|vα|e^{-\varepsilon_{2}\absolutevalue{v-\alpha}}\leq\frac{h(v)}{h(\alpha)}\leq e^{\varepsilon_{2}\absolutevalue{v-\alpha}} (V.18)

and observe that

Veff,ω,LSγa182F\norm{V_{{\mathrm{eff}},\omega,L}}_{\infty}\leq S_{-\gamma_{a}}18\sqrt{2}\norm{F}_{\infty} (V.19)

c.f. Theorem 3 in A-G and assumption 5. This yields, with ϑ=(2c1|g|λ+4c1C1|g|2λ2)Sγa182F\vartheta=(2c_{1}\frac{\absolutevalue{g}}{\lambda}+4c_{1}C_{1}\frac{\absolutevalue{g}^{2}}{\lambda^{2}})S_{-\gamma_{a}}18\sqrt{2}\norm{F}_{\infty},

ρn0,Leff,Λ(v)eϑecρ|v|ecρ|α|e(ε2+θ)|vα|𝑑α.\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\leq e^{\vartheta}\frac{e^{-c_{\rho}\absolutevalue{v}}}{\int^{\infty}_{-\infty}e^{-c_{\rho}\absolutevalue{\alpha}}e^{-(\varepsilon_{2}+\theta)\absolutevalue{v-\alpha}}\,d\alpha}. (V.20)

Pick g1g_{1} sufficiently small such that if |g|F<g1\absolutevalue{g}\norm{F}_{\infty}<g_{1} then θ<ε22\theta<\frac{\varepsilon_{2}}{2}

ρn0,Leff,Λ(v)eϑecρ|v|ecρ|α|e3ε22|vα|𝑑α.\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\leq e^{\vartheta}\frac{e^{-c_{\rho}\absolutevalue{v}}}{\int^{\infty}_{-\infty}e^{-c_{\rho}\absolutevalue{\alpha}}e^{-\frac{3\varepsilon_{2}}{2}\absolutevalue{v-\alpha}}\,d\alpha}. (V.21)

from which we readily obtain, for ε:=3ε22\varepsilon:=\frac{3\varepsilon_{2}}{2} and

ρn0,Leff,Λ(v)eϑ(cρε2)e(cρ+ε)|v|.\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\leq e^{\vartheta}\left(\frac{c_{\rho}-\varepsilon}{2}\right)e^{(-c_{\rho}+\varepsilon)\absolutevalue{v}}. (V.22)

The lower bound in (i) is analogous. One follows the above process using instead the upper bounds given in (V.8) and (V.10) along with assumptions 6, 7 and (V.3) to reach

ρn0,Leff,Λ(v)eϑ(cρε2)e(cρε)|v|.\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\geq e^{-\vartheta}\left(\frac{c_{\rho}-\varepsilon}{2}\right)e^{(-c_{\rho}-\varepsilon)\absolutevalue{v}}. (V.23)

finishing the proof of (i). To prove (ii) we use (V.3) to write

ρn0,Leff,Λ(v)ρn0,Leff,Λ(v)=ρ(vgλVeff,𝒯1U,L(n0))nΛ{n0}ρ(U(n)gλVeff,𝒯1U,L(n))JUρ(vgλVeff,𝒯1Uv,L(n0))nΛ{n0}ρ(Uv(n)gλVeff,𝒯1Uv,L(n))JUv\frac{\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)}{\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v^{\prime})}=\frac{\rho\left(v-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U,L}(n_{0})\right)\prod_{n\in\Lambda^{\prime}\setminus\{n_{0}\}}\rho\left(U(n)-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U,L}(n)\right)J_{U}}{\rho\left(v^{\prime}-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U_{v^{\prime}},L}(n_{0})\right)\prod_{n\in\Lambda^{\prime}\setminus\{n_{0}\}}\rho\left(U_{v^{\prime}}(n)-\frac{g}{\lambda}V_{{\mathrm{eff}},{\mathcal{T}}^{-1}U_{v^{\prime}},L}(n)\right)J_{U_{v^{\prime}}}} (V.24)

Where Uv(n)=U(n)+(vU(n0))δn0U_{v^{\prime}}(n)=U(n)+(v^{\prime}-U(n_{0}))\delta_{n_{0}} for nΛn\in\Lambda^{\prime}. The bounds in (ii) then follow as above from (V.10) and (V.8), both applied to α=v\alpha=v^{\prime}, along with assumption 6 and (V.19).

VI Self-avoiding walks and localization: Proof of Theorem 1

It is well known that the conclusion of Theorem 1 follows from the result below, see (A-S-F-H, , Appendix B).

Theorem 10

There exist λHF\lambda_{HF} and g0=g0(Ca,d,ρ,λ,γa,η0)g_{0}=g_{0}(C_{a},d,\rho,\lambda,\gamma_{a},\eta_{0}) (independent of LL and Λ\Lambda^{\prime}) such that whenever λ>λHF\lambda>\lambda_{HF} and |g|F<g0\absolutevalue{g}\norm{F}_{\infty}<g_{0} we have that for each s(0,1)s\in(0,1)

𝔼(|GLΛ(m,n;z)|s)Cseξs|mn|\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}}_{L}(m,n;z)}^{s}\right)\leq C_{s}e^{-\xi_{s}|m-n|} (VI.1)

for all zz\in\mathbb{C}\setminus\mathbb{R} and certain constants Cs>0C_{s}>0 and ξs>0\xi_{s}>0 independent of LL and Λ\Lambda^{\prime}. Moreover, λHF\lambda_{HF} solves (II.12).

Proof.

We closely follow the arguments of Schenkl but provide details for the sake of completeness since a few modifications are required to account for the Hartree-Fock setting. Let zz\in\mathbb{C}\setminus\mathbb{R}. We start from the depleted resolvent identity which is valid for mnΛm\neq n\in\Lambda^{\prime}:

GLΛ(m,n;z)=GLΛ(m,m;z)mΛ|mm|=1GLΛ{m}(m,n;z).G^{\Lambda^{\prime}}_{L}(m,n;z)=-G^{\Lambda^{\prime}}_{L}(m,m;z)\sum_{\begin{subarray}{c}m^{\prime}\in\Lambda^{\prime}\\ |m^{\prime}-m|=1\end{subarray}}G^{\Lambda^{\prime}\setminus\{m\}}_{L}(m^{\prime},n;z). (VI.2)

Note that by Lemma 8 we have the local fractional moment bound

𝔼U(mj)(|GLΛ′′(mj,mj;z)|s)(2M)s(1s)λs\mathbb{E}_{U(m_{j})}\left(\absolutevalue{G^{\Lambda^{\prime\prime}}_{L}(m_{j},m_{j};z)}^{s}\right)\leq\frac{(2M_{\infty})^{s}}{(1-s)\lambda^{s}} (VI.3)

which is valid for any Λ′′ΛL\Lambda^{\prime\prime}\subset\Lambda_{L} and s(0,1)s\in(0,1), see (A-W-B, , Theorem 8.1). Iterating (VI.2) along a sequence m0=m,m1,,mjm_{0}=m,m_{1},\ldots,m_{j} of distinct points in Λ\Lambda^{\prime} and applying (VI.3) we find that after NN iterations

𝔼(|GLΛ(m,n;z)|s)\displaystyle\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}}_{L}(m,n;z)}^{s}\right)\leq j=0N((2M)s(1s)λs)j{mk}k=1jSjΛ(n,m)𝔼(|GLΛ{m0,,mj}(n,n;z)|s)\displaystyle\sum^{N}_{j=0}\left(\frac{(2M_{\infty})^{s}}{(1-s)\lambda^{s}}\right)^{j}\sum_{\{m_{k}\}^{j}_{k=1}\in S^{\Lambda^{\prime}}_{j}(n,m)}\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}\setminus\{m_{0},\ldots,m_{j}\}}_{L}(n,n;z)}^{s}\right)
+((2M)s(1s)λs)N{mk}k=1NSNΛ(m)mknk=1,,N𝔼(|GLΛ{m0,,mk}(mk,n;z)|s)\displaystyle+\left(\frac{(2M_{\infty})^{s}}{(1-s)\lambda^{s}}\right)^{N}\sum_{\begin{subarray}{c}\{m_{k}\}^{N}_{k=1}\in S^{\Lambda^{\prime}}_{N}(m)\\ m_{k}\neq n\,\,k=1,\ldots,N\end{subarray}}\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}\setminus\{m_{0},\ldots,m_{k}\}}_{L}(m_{k},n;z)}^{s}\right)

where we denote by SjΛ(n,m)S^{\Lambda^{\prime}}_{j}(n,m) the set of self-avoiding walks in Λ\Lambda^{\prime} of length jj starting at mm and ending at nn and by SNΛ(m)=nΛSNΛ(n,m)S^{\Lambda^{\prime}}_{N}(m)=\cup_{n\in\Lambda^{\prime}}S^{\Lambda^{\prime}}_{N}(n,m) the set of all self-avoiding walks in Λ\Lambda^{\prime} of length NN starting at mm. Therefore, applying (VI.3) once more and denoting Γ(s):=(2M)s(1s)λs\Gamma(s):=\frac{(2M_{\infty})^{s}}{(1-s)\lambda^{s}} we have that

𝔼(|GLΛ(m,n;z)|s)j=0NΓ(s)j+1#SjΛ(n,m)+Γ(s)N#SNΛ(m)1|Imz|s.\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}}_{L}(m,n;z)}^{s}\right)\leq\sum^{N}_{j=0}\Gamma(s)^{j+1}\#S^{\Lambda^{\prime}}_{j}(n,m)+\Gamma(s)^{N}\#S^{\Lambda^{\prime}}_{N}(m)\frac{1}{\absolutevalue{\mathrm{Im}z}^{s}}. (VI.4)

We now make use of some facts about self-avoiding walks, see Schenkl and references therein for a more detailed discussion. Recall that the self-avoiding walk correlation function is defined by

Cγ(nm):=N=0γN#SN(n,m)C_{\gamma}(n-m):=\sum^{\infty}_{N=0}\gamma^{N}\#S_{N}(n,m) (VI.5)

whenever N=0|γ|N#SN(n,m)<\sum^{\infty}_{N=0}\absolutevalue{\gamma}^{N}\#S_{N}(n,m)<\infty. The self-avoiding walk susceptibility is defined by

χ(γ):=mdCγ(m)=N=0CNγN\chi(\gamma):=\sum_{m\in\mathbb{Z}^{d}}C_{\gamma}(m)=\sum^{\infty}_{N=0}C_{N}\gamma^{N} (VI.6)

where CNC_{N} denotes the number of self-avoiding walks of length NN starting at 0. We also recall that the conective constant of d\mathbb{Z}^{d} is

μd=limN(CN)1N.\mu_{d}=\lim_{N\to\infty}(C_{N})^{\frac{1}{N}}. (VI.7)

In particular, 1μd\frac{1}{\mu_{d}} is the radius of convergence of (VI.6). It is also well-known that 0<μd<2d10<\mu_{d}<2d-1. It is crucial for our argument that whenever 0<γ<1μd0<\gamma<\frac{1}{\mu_{d}} the self-avoiding walk correlation function Cγ(m)C_{\gamma}(m) decays exponentially as |m||m|\to\infty. This follows from the inequality

Cγ(m)Bε((μd+ε)γ)|m|C_{\gamma}(m)\leq B_{\varepsilon}\left((\mu_{d}+\varepsilon)\gamma\right)^{|m|} (VI.8)

valid for ε>0\varepsilon>0 and some constant BεB_{\varepsilon}.

Therefore, whenever Γ(s)<1μd\Gamma(s)<\frac{1}{\mu_{d}} we have that χ(Γ(s))N=0CNΓ(s)N<\chi(\Gamma(s))\leq\sum^{\infty}_{N=0}C_{N}\Gamma(s)^{N}<\infty. In particular, the remainder in (VI.4) satisfies

Γ(s)N#SNΛ(m)Γ(s)NCN0asN.\Gamma(s)^{N}\#S^{\Lambda^{\prime}}_{N}(m)\leq\Gamma(s)^{N}C_{N}\to 0\,\,\,\text{as}\,\,\,N\to\infty. (VI.9)

Thus, letting NN\to\infty in (VI.4) we find

𝔼(|GLΛ(m,n;z)|s)j=0Γ(s)j+1#SjΛ(n,m).\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}}_{L}(m,n;z)}^{s}\right)\leq\sum^{\infty}_{j=0}\Gamma(s)^{j+1}\#S^{\Lambda^{\prime}}_{j}(n,m). (VI.10)

from which we conclude that

𝔼(|GLΛ(m,n;z)|s)Γ(s)CΓ(s)(mn).\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}}_{L}(m,n;z)}^{s}\right)\leq\Gamma(s)C_{\Gamma(s)}(m-n). (VI.11)

Finally, to end the proof we determine for which values of s(0,1)s\in(0,1) one has that Γ(s)<1μd\Gamma(s)<\frac{1}{\mu_{d}}. Observe that whenever λ2M>e\frac{\lambda}{2M_{\infty}}>e the only critical point of Γ(s)\Gamma(s) is s0(λ)=11ln(λ2M)s_{0}(\lambda)=1-\frac{1}{\ln\left(\frac{\lambda}{2M_{\infty}}\right)} which yields

Γ(s0)=eln(λ2M)2Mλ.\Gamma(s_{0})=e\ln\left(\frac{\lambda}{2M_{\infty}}\right)\frac{2M_{\infty}}{\lambda}. (VI.12)

Thus Γ(s0)<1μd\Gamma(s_{0})<\frac{1}{\mu_{d}} if and only if

λ>2Mμdeln(λ2M)\lambda>{2M_{\infty}}\mu_{d}e\ln\left(\frac{\lambda}{2M_{\infty}}\right) (VI.13)

so the critical threshold is λHF=2Mln(λHF(2M))μde\lambda_{\mathrm{HF}}=2M_{\infty}\ln\left(\frac{\lambda_{\mathrm{HF}}}{(2M_{\infty})}\right)\mu_{d}e. For values of λ\lambda greater than λHF\lambda_{HF} we conclude that there exists ε>0\varepsilon>0 for which

𝔼(|GLΛ(m,n;z)|11ln(λ2M))eln(λ2M)2MλBε((μd+ε)eln(λ2M)2Mλ)|mn|.\mathbb{E}\left(\absolutevalue{G^{\Lambda^{\prime}}_{L}(m,n;z)}^{1-\frac{1}{\ln\left(\frac{\lambda}{2M_{\infty}}\right)}}\right)\leq e\ln\left(\frac{\lambda}{2M_{\infty}}\right)\frac{2M_{\infty}}{\lambda}B_{\varepsilon}\left((\mu_{d}+\varepsilon)e\ln\left(\frac{\lambda}{2M_{\infty}}\right)\frac{2M_{\infty}}{\lambda}\right)^{|m-n|}. (VI.14)

and (μd+ε)eln(λ2M)2Mλ<1(\mu_{d}+\varepsilon)e\ln\left(\frac{\lambda}{2M_{\infty}}\right)\frac{2M_{\infty}}{\lambda}<1. Applying Hölder’s inequality we conclude that (VI.1) holds for any s(0,1)s\in(0,1) and some Cs>0C_{s}>0 and ξs>0\xi_{s}>0. This is immediate if 0<s<11ln(λ2M)0<s<1-\frac{1}{\ln\left(\frac{\lambda}{2M_{\infty}}\right)} and follows from (off-diagonal) a-priori bounds for the Green’s function if 11ln(λ2M)<s<11-\frac{1}{\ln\left(\frac{\lambda}{2M_{\infty}}\right)}<s<1, see (A-S-F-H, , Lemma B2) and (A-W-B, , Theorem 8.3).

VII Proof of theorem 3 and Corollary 2

Similarly to how Theorem 10 implies Theorem 1, Theorem 3 follows from the result below.

Theorem 11

In the setting of Lemma 9, for each II\subset\mathbb{R} the exists g1(Ca,d,ρ,λ,γa,η0)g_{1}(C_{a},d,\rho,\lambda,\gamma_{a},\eta_{0}), ν′′>0\nu^{\prime\prime}>0, C>0C>0 and λ0\lambda_{0} (independent of Λ\Lambda^{\prime} and LL) such that whenever |g||F|<g1|g|\absolutevalue{F}_{\infty}<g_{1} and λ<λ0\lambda<\lambda_{0} we have that

𝔼(|GΛ(m,n,E)|s)Ceν′′|mn|\mathbb{E}\left(\absolutevalue{G_{\Lambda}(m,n,E)}^{s}\right)\leq Ce^{-\nu^{\prime\prime}|m-n|} (VII.1)

for some s(0,1)s\in(0,1). Moreover, we have that

λ0=sups(0,1)supμ>0infEIλ^s,μ(E)\lambda_{0}=\sup_{s\in(0,1)}\sup_{\mu>0}\inf_{E\in I}\widehat{\lambda}_{s,\mu}(E) (VII.2)

where

λ^s,μ(E)=(Ds,1supδ0supudvd|G0(u,v;E+iδ)|seμ|uv|)1s.\widehat{\lambda}_{s,\mu}(E)=\left(D_{s,1}\sup_{\delta\neq 0}\sup_{u\in\mathbb{Z}^{d}}\sum_{v\in\mathbb{Z}^{d}}|G_{0}(u,v;E+i\delta)|^{s}e^{\mu|u-v|}\right)^{\frac{-1}{s}}. (VII.3)

Theorem 11 in turn follows from Lemma 9 along with known results and thus we only provide an outline for how it is proven. Before doing so, we recall some notions of regularity for probability distributions, c.f. A-Molc ; A-W-B which will be relevant in the sequel.

Definition 12
  1. (i)

    A probability measure ρ(dv)\rho(dv) on the real line is τ\tau-regular, with τ(0,1]\tau\in(0,1], if for some v0v_{0}\in\mathbb{R} and C>0C>0

    ρ([vδ,v+δ])C|δ|τρ([vv0,v+v0])\rho\left([v-\delta,v+\delta]\right)\leq C|\delta|^{\tau}\rho\left([v-v_{0},v+v_{0}]\right) (VII.4)

    holds for all δ(0,1)\delta\in(0,1) and vv\in\mathbb{R}.

  2. (ii)

    A joint probability measure ρ(dV)\rho(dV) of a collection of random variables {Vn}\{V_{n}\} is conditionally τ\tau-regular if the conditional distributions of VnV_{n} at specified values of {Vm}mn\{V_{m}\}_{m\neq n} satisfy (VII.4) with uniform values of the constants appearing there.

  3. (iii)

    If, additionally, for some ε>0\varepsilon>0 the conditional expectations of |Vn|ε|V_{n}|^{\varepsilon} are uniformly bounded:

    𝔼(|Vn|ε|V{n}c)B,for someB>0,\mathbb{E}\left(|V_{n}|^{\varepsilon}|\,\,V_{\{n\}^{c}}\right)\leq B,\,\,\,\,\,\,\text{for some}\,\,\,B>0, (VII.5)

    then the joint probability measure ρ(dV)\rho(dV) is said to be conditionally (τ,ε)(\tau,\varepsilon)-regular.

  4. (iv)

    ρ\rho has regular qq-decay for q>0q>0 if

    ρ([u1,u+1])C1+|u|q,for someC>0.\rho\left([u-1,u+1]\right)\leq\frac{C}{1+\absolutevalue{u}^{q}},\,\,\,\,\,\,\text{for some}\,\,\,C>0. (VII.6)

Proof of Theorem 11: Lemma 9 (i) readily implies that ρn0,Leff,Λ(v)dv\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\,dv has regular qq decay for all q>0q>0 and that for all p>0p>0

|v|pρn0,Leff,Λ(v)𝑑v<,\int^{\infty}_{-\infty}|v|^{p}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\,dv<\infty,

i.e. ρn0,Leff,Λ(v)dv\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\,dv is conditionally (1,p)(1,p)-regular for all p>0p>0. Moreover, by Lemma 9 (ii), we have that for any δ(0,1]\delta\in(0,1] and uu\in\mathbb{R}

uδu+δρn0,Leff,Λ(v)𝑑v\displaystyle\int^{u+\delta}_{u-\delta}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\,dv (2δ)ec1(1+ϑ)ρn0,Leff,Λ(u)\displaystyle\leq(2\delta)e^{c_{1}(1+\vartheta)}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(u)
δe2c1(1+ϑ)u1u+1ρn0,Leff,Λ(v)𝑑v,\displaystyle\leq\delta e^{2c_{1}(1+\vartheta)}\int^{u+1}_{u-1}\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\,dv,

in particular we see that ρn0,Leff,Λ(v)dv\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L}(v)\,dv is (uniformly) 11-regular.

We then conclude from (A-W-B, , Theorem 8.7) that ρn0,Leff,Λ\rho^{{\mathrm{eff}},\Lambda^{\prime}}_{n_{0},L} is 11-moment regular, namely Ds,1<D_{s,1}<\infty with Ds,1D_{s,1} as in (II.17) for all s(0,1)s\in(0,1). In particular, Theorem 11 falls into the framework of (A-W-B, , Theorem 10.4) .

Proof of Corollary 2: Note that when |g|F0\absolutevalue{g}\norm{F}_{\infty}\to 0 then θ0\theta\to 0 in equation (V.17) ( which only requires assumptions 1-6). Thus, by dominated convergence, we may choose MM_{\infty} such that MρM_{\infty}\to{\norm{\rho}}_{\infty} as |g|F0\absolutevalue{g}\norm{F}_{\infty}\to 0. Corollary 2 now follows from (I.3) and (II.14) since these equations imply

(λHFλAnd)2M(ln(λHF)ln(λAnd))=\displaystyle(\lambda_{\mathrm{HF}}-\lambda_{\mathrm{And}})-2M_{\infty}(\ln(\lambda_{\mathrm{HF}})-\ln(\lambda_{\mathrm{And}}))= 2μdln(λAnd)(Mρ)\displaystyle 2\mu_{d}\ln(\lambda_{\mathrm{And}})(M_{\infty}-\norm{\rho}_{\infty})
2μd(ρln(2ρ)Mln(2M))\displaystyle 2\mu_{d}(\norm{\rho}_{\infty}\ln(2\norm{\rho}_{\infty})-M_{\infty}\ln(2M_{\infty}))

and by construction λHF2M>e>1.\frac{\lambda_{\mathrm{HF}}}{2M_{\infty}}>e>1.

Appendix A Appendix

We now provide some results on existence and uniqueness of the effective potentials as well as their regularity with respect to the random variables. Since the statements are mostly immediate generalizations from the ones given in M-S we skip most proofs. We formulate the first of these results for (d)\ell^{\infty}\left(\mathbb{Z}^{d}\right) but remark that its finite volume analogue holds similarly.

A.1 Contraction mapping arguments

Let Φ:(d)(d)\Phi:\ell^{\infty}\left(\mathbb{Z}^{d}\right)\rightarrow\ell^{\infty}\left(\mathbb{Z}^{d}\right) be given by

Φ(V)=mda(n,m)δm,F(A+λVω+gV)δm\Phi(V)=\sum_{m\in\mathbb{Z}^{d}}a(n,m)\langle{\delta_{m}},F(A+\lambda V_{\omega}+gV)\delta_{m}\rangle (A.1)

We wish to show that there is a unique solution VeffV_{{\mathrm{eff}}} to the equation Φ(V)=V\Phi(V)=V. For that purpose, we introduce a technical Lemma which may be found in (M-S, , Proposition 12)

Lemma 13
  1. (a)

    Let T=A+λVωT=A+\lambda V_{\omega} be as in assumptions 1-5. Given potentials V,W(d)V,W\in\ell^{\infty}\left(\mathbb{Z}^{d}\right), we have, for any ν\nu satisfying (III.3) and δ(0,ν)\delta\in(0,\nu), that

    |δm,(F(T+V)F(T+W))δn|722ηSδνFVWeνd(m,n).\Big{|}\langle\delta_{m},\left(F(T+V)-F(T+W)\right)\delta_{n}\rangle\Big{|}\leq\frac{72\sqrt{2}}{\eta}S_{\delta-\nu}\|F\|_{\infty}\|V-W\|_{\infty}e^{-\nu^{\prime}d(m,n)}. (A.2)
  2. (b)

    For any m,n,jdm,n,j\in\mathbb{Z}^{d}, the matrix elements δm,F(T+gV)δn\langle\delta_{m},F(T+gV)\delta_{n}\rangle are differentiable with respect to V(j)V(j) and

    |δm,F(T+gV)δnV(j)||g|722eν(|d(m,j)+d(n,j))ηFV.\Big{|}\frac{\partial\langle\delta_{m},F(T+gV)\delta_{n}\rangle}{\partial V(j)}\Big{|}\leq|g|\frac{72\sqrt{2}e^{-\nu\left(|d(m,j)+d(n,j)\right)}}{\eta}\|F\|_{\infty}\|V\|_{\infty}. (A.3)

From Lemma 13 and assumption 5 we obtain

Φ(V)Φ(W)\displaystyle\norm{\Phi(V)-\Phi(W)}_{\infty} |g|722ηSδνSγaCaFVW\displaystyle\leq|g|\frac{72\sqrt{2}}{\eta}S_{\delta-\nu}S_{-\gamma_{a}}C_{a}\|F\|_{\infty}\|V-W\|_{\infty}

thus we conclude the following.

Proposition 14

Whenever |g|722ηSδνSγaCaF<1|g|\frac{72\sqrt{2}}{\eta}S_{\delta-\nu}S_{-\gamma_{a}}C_{a}\|F\|_{\infty}<1 for some δ(0,ν)\delta\in(0,\nu) the map Φ:(d)(d)\Phi:\ell^{\infty}\left(\mathbb{Z}^{d}\right)\rightarrow\ell^{\infty}\left(\mathbb{Z}^{d}\right) is a contraction. In particular, there is a unique Veff(d)V_{{\mathrm{eff}}}\in\ell^{\infty}\left(\mathbb{Z}^{d}\right) such that Φ(Veff)=Veff\Phi(V_{{\mathrm{eff}}})=V_{{\mathrm{eff}}}. Moreover, the analogue effective potential in finite volume ΛL\Lambda_{L}, Veff,ω,LV_{{\mathrm{eff}},\omega,L}, is a smooth function of (ω(n1),,ω(n|ΛL|))(\omega(n_{1}),...,\omega(n_{\absolutevalue{\Lambda_{L}}})).

We also note that if a(n,m)a(n,m)\in\mathbb{R} for each n,mdn,m\in\mathbb{Z}^{d} then Veff(n)V_{{\mathrm{eff}}}(n)\in\mathbb{R} for each ndn\in\mathbb{Z}^{d}.

A.2 Norm resolvent convergence

Finally, we briefly comment on the convergence of resolvents which allows to extend the results of Theorems 1 and 3 to infinite volume operators. It will be useful to introduce the augumented boundary

ΛL={ud:dist(u,ΛL)=1ordist(u,ΛLc)=1}\partial\Lambda_{L}=\{u\in\mathbb{Z}^{d}:\,\,\mathrm{dist}(u,\Lambda_{L})=1\,\,\text{or}\,\,\mathrm{dist}(u,\Lambda^{c}_{L})=1\} (A.4)

with dist(u,X)\mathrm{dist}(u,X) calculated in the metric of d\mathbb{Z}^{d}.

Lemma 15
  1. (a)

    Given nΛLn\in\Lambda_{L} whenever 32|g|FSνη<12\frac{3\sqrt{2}\absolutevalue{g}\norm{F}_{\infty}S_{-\nu}}{\eta}<\frac{1}{2} we have that

    |Veff,ω(n)Veff,ω,L(n)|Ceδd(n,ΛL)\absolutevalue{V_{{\mathrm{eff}},\omega}(n)-V_{{\mathrm{eff}},\omega,L}(n)}\leq Ce^{-\delta d(n,\partial\Lambda_{L})} (A.5)

    for any δ<min{ν,γa}\delta<\min\{\nu,\gamma_{a}\} and C=432CaF|g|SνηC=\frac{432C_{a}\norm{F}_{\infty}\absolutevalue{g}S_{-\nu}}{\eta}, with d(n,ΛL)d(n,\partial\Lambda_{L}) calculated in the metric d(,)d(\cdot,\cdot) of assumption 5.

  2. (b)

    For any κ>0\kappa>0, with |g|\absolutevalue{g} and δ\delta as above

    |GΛL(m,n;t+iκ)GLΛL(m,n;t+iκ)|4Cκ2eνd(m,n)δd(n,ΛL))Sν\absolutevalue{G^{\Lambda_{L}}(m,n;t+i\kappa)-G^{\Lambda_{L}}_{L}(m,n;t+i\kappa)}\leq\frac{4C}{\kappa^{2}}e^{-\nu d(m,n)-\delta d(n,\partial\Lambda_{L}))}S_{-\nu} (A.6)

    In particular, for each fixed z+z\in\mathbb{C}^{+} and ψ2(d)\psi\in\ell^{2}\left(\mathbb{Z}^{d}\right) we have that

    (HΛLz)1ψ(HLΛLz)1ψ0asL.\norm{(H^{\Lambda_{L}}-z)^{-1}\psi-{(H^{\Lambda_{L}}_{L}-z)^{-1}\psi}}\to 0\,\,\,\text{as}\,\,\,L\to\infty. (A.7)

Proof. Using (III.1) and the analogous representation for Veff,ω(n)V_{{\mathrm{eff}},\omega}(n) we find

|Veff,ω(n)Veff,ω,L(n)|3F2π|K(n,ω;t)KL(n,ω;t)|𝑑t\absolutevalue{V_{{\mathrm{eff}},\omega}(n)-V_{{\mathrm{eff}},\omega,L}(n)}\leq\frac{3\norm{F}_{\infty}}{2\pi}\int^{\infty}_{-\infty}\absolutevalue{K(n,\omega;t)-K_{L}(n,\omega;t)}\,dt

where for z=tiηz=t-i\eta

|K(n,ω;t)KL(n,ω;t)|\displaystyle\absolutevalue{K(n,\omega;t)-K_{L}(n,\omega;t)} md|a(n,m)||G(m,m;z)GL(m,m;z)|\displaystyle\leq\sum_{m\in\mathbb{Z}^{d}}\absolutevalue{a(n,m)}\absolutevalue{G(m,m;z)-G_{L}(m,m;z)}
md|a(n,m)||G(m,m;z¯)GL(m,m;z¯)|\displaystyle\sum_{m\in\mathbb{Z}^{d}}\absolutevalue{a(n,m)}\absolutevalue{G(m,m;\bar{z})-G_{L}(m,m;\bar{z})}

Observe that letting ΛLo:=ΛLΛL\Lambda_{L}^{\mathrm{o}}:=\Lambda_{L}\setminus\partial\Lambda_{L} and (ΛLo)c=dΛLo(\Lambda_{L}^{\mathrm{o}})^{c}=\mathbb{Z}^{d}\setminus\Lambda_{L}^{\mathrm{o}}, for any mdm\in\mathbb{Z}^{d} we have that

|G(m,m;z)GL(m,m;z)|\displaystyle\absolutevalue{G(m,m;z)-G_{L}(m,m;z)} |g|kΛLo|G(m,k;z)||Veff,ω(k)Veff,ω,L(k)||GL(k,m;z)|\displaystyle\leq\absolutevalue{g}\sum_{k\in\Lambda_{L}^{\mathrm{o}}}\absolutevalue{G(m,k;z)}\absolutevalue{V_{{\mathrm{eff}},\omega}(k)-V_{{\mathrm{eff}},\omega,L}(k)}\absolutevalue{G_{L}(k,m;z)}
362FSγa|g|k(ΛLo)c|G(m,k;z)||GL(k,m;z)|\displaystyle 36\sqrt{2}\norm{F}_{\infty}S_{-\gamma_{a}}\absolutevalue{g}\sum_{k^{\prime}\in(\Lambda_{L}^{\mathrm{o}})^{c}}\absolutevalue{G(m,k^{\prime};z)}\absolutevalue{G_{L}(k^{\prime},m;z)}

where we have used that max{Veff,ω,L,Veff,ωSγa182F\max\{\norm{V_{{\mathrm{eff}},\omega,L}}_{\infty},\norm{V_{{\mathrm{eff}},\omega}}_{\infty}\leq S_{-\gamma_{a}}18\sqrt{2}\norm{F}_{\infty} c.f. Theorem 3 in A-G and assumption 5.

The result in (a) now follows from

|G(u,v;z)||GL(v,u;z)|𝑑t22πηe2νd(u,v)\int^{\infty}_{-\infty}\absolutevalue{G(u,v;z)}\absolutevalue{G_{L}(v,u;z)}\,dt\leq\frac{2\sqrt{2}\pi}{\eta}e^{-2\nu d(u,v)} (A.8)

combined with assumption 5 and another application of Lemma 4 with

φ(n)=|Veff,ω(n)Veff,ω,L(n)|,W(n)=eδd(n,ΛL)\varphi(n)=\absolutevalue{V_{{\mathrm{eff}},\omega}(n)-V_{{\mathrm{eff}},\omega,L}(n)},\,\,\,W(n)=e^{\delta d(n,{\partial\Lambda_{L}})} (A.9)

and

K(n,u)=(mdeγad(n,m)2νd(m,u))𝟙ΛLo(u)K(n,u)=\left(\sum_{m\in\mathbb{Z}^{d}}e^{-\gamma_{a}d(n,m)-2\nu d(m,u)}\right)\mathds{1}_{{\Lambda_{L}}^{\mathrm{o}}}(u) (A.10)

for which we have b1=216F|g|Sνηb_{1}=\frac{216\norm{F}_{\infty}\absolutevalue{g}S_{-\nu}}{\eta} and b2=32|g|FSνηb_{2}=\frac{3\sqrt{2}\absolutevalue{g}\norm{F}_{\infty}S_{-\nu}}{\eta}.

(b) now follows from (a) combined with the resolvent identity and another application Combes-Thomas bound.

Acknowledgements

Dedicated to Abel Klein in ocasion of his 78th birthday. This work was partially supported by NSF DMS-2000345 and DMS-2052572. R. Matos is thankful to the anonymous reviewer for several remarks which greatly improved the exposition in this note.

References

  • [1] H. Abdul-Rahman, B. Nachtergaele, R. Sims, and G. Stolz. Localization properties of the disordered XY spin chain: a review of mathematical results with an eye toward many-body localization. Ann. Phys., 529(7):201600280, 17, 2017.
  • [2] V. Acosta and A. Klein. Analyticity of the density of states in the Anderson model on the Bethe lattice. J. Statist. Phys., 69(1-2):277–305, 1992.
  • [3] M. Aizenman. Localization at weak disorder: some elementary bounds. volume 6, pages 1163–1182. 1994. Special issue dedicated to Elliott H. Lieb.
  • [4] M. Aizenman, A. Elgart, S. Naboko, J. H. Schenker, and G. Stolz. Moment analysis for localization in random Schrödinger operators. Invent. Math., 163(2):343–413, 2006.
  • [5] M. Aizenman and G. M. Graf. Localization bounds for an electron gas. J. Phys. A, 31(32):6783–6806, 1998.
  • [6] M. Aizenman and S. Molchanov. Localization at large disorder and at extreme energies: an elementary derivation. Comm. Math. Phys., 157(2):245–278, 1993.
  • [7] M. Aizenman, J. H. Schenker, R. M. Friedrich, and D. Hundertmark. Finite-volume fractional-moment criteria for Anderson localization. volume 224, pages 219–253. 2001. Dedicated to Joel L. Lebowitz.
  • [8] M. Aizenman and S. Warzel. Localization bounds for multiparticle systems. Comm. Math. Phys., 290(3):903–934, 2009.
  • [9] M. Aizenman and S. Warzel. Random operators, volume 168 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2015. Disorder effects on quantum spectra and dynamics.
  • [10] P. W. Anderson. Absence of diffusion in certain random lattices. Phys. Rev., 109:1492–1505, Mar 1958.
  • [11] V. Bach, E. H. Lieb, and J. P. Solovej. Generalized Hartree-Fock theory and the Hubbard model. J. Statist. Phys., 76(1-2):3–89, 1994.
  • [12] V. Beaud and S. Warzel. Low-energy Fock-space localization for attractive hard-core particles in disorder. Ann. Henri Poincaré, 18(10):3143–3166, 2017.
  • [13] N. Benedikter, M. Porta, and B. Schlein. Effective evolution equations from quantum dynamics, volume 7 of SpringerBriefs in Mathematical Physics. Springer, Cham, 2016.
  • [14] J. Bourgain and C. E. Kenig. On localization in the continuous Anderson-Bernoulli model in higher dimension. Invent. Math., 161(2):389–426, 2005.
  • [15] V. Bucaj, D. Damanik, J. Fillman, V. Gerbuz, T. VandenBoom, F. Wang, and Z. Zhang. Localization for the one-dimensional Anderson model via positivity and large deviations for the Lyapunov exponent. Trans. Amer. Math. Soc., 372(5):3619–3667, 2019.
  • [16] E. Cancès, S. Lahbabi, and M. Lewin. Mean-field electronic structure models for disordered materials. In XVIIth International Congress on Mathematical Physics, pages 549–557. World Sci. Publ., Hackensack, NJ, 2014.
  • [17] R. Carmona, A. Klein, and F. Martinelli. Anderson localization for Bernoulli and other singular potentials. Comm. Math. Phys., 108(1):41–66, 1987.
  • [18] I. Chenn and S. Zhang. On the reduced Hartree-Fock equations with a small anderson type background charge distribution. Journal of Functional Analysis, 283(12):109702, 2022.
  • [19] V. Chulaevsky and Y. Suhov. Multi-particle Anderson localisation: induction on the number of particles. Math. Phys. Anal. Geom., 12(2):117–139, 2009.
  • [20] V. Chulaevsky and Y. Suhov. Efficient Anderson localization bounds for large multi-particle systems. J. Spectr. Theory, 7(1):269–320, 2017.
  • [21] J. Ding and C. K. Smart. Localization near the edge for the Anderson Bernoulli model on the two dimensional lattice. Invent. Math., 219(2):467–506, 2020.
  • [22] R. Ducatez. Anderson localisation for infinitely many interacting particles in Hartree-Fock theory. J. Spectr. Theory, 8(3):1019–1050, 2018.
  • [23] A. Elgart. Lifshitz tails and localization in the three-dimensional Anderson model. Duke Math. J., 146(2):331–360, 2009.
  • [24] A. Elgart and A. Klein. Localization in the random XXZ quantum spin chain. https://arxiv.org/abs/2210.14873, 2022.
  • [25] A. Elgart, A. Klein, and G. Stolz. Manifestations of dynamical localization in the disordered XXZ spin chain. Comm. Math. Phys., 361(3):1083–1113, 2018.
  • [26] A. Elgart, A. Klein, and G. Stolz. Many-body localization in the droplet spectrum of the random XXZ quantum spin chain. J. Funct. Anal., 275(1):211–258, 2018.
  • [27] A. Elgart, M. Shamis, and S. Sodin. Localisation for non-monotone Schrödinger operators. J. Eur. Math. Soc. (JEMS), 16(5):909–924, 2014.
  • [28] A. Elgart, M. Tautenhahn, and I. Veselić. Anderson localization for a class of models with a sign-indefinite single-site potential via fractional moment method. Ann. Henri Poincaré, 12(8):1571–1599, 2011.
  • [29] J. Fröhlich, F. Martinelli, E. Scoppola, and T. Spencer. Constructive proof of localization in the Anderson tight binding model. Comm. Math. Phys., 101(1):21–46, 1985.
  • [30] J. Fröhlich and T. Spencer. Absence of diffusion in the Anderson tight binding model for large disorder or low energy. Comm. Math. Phys., 88(2):151–184, 1983.
  • [31] M. Gebert and C. Rojas-Molina. Lifshitz tails for the fractional Anderson model. J. Stat. Phys., 179(2):341–353, 2020.
  • [32] M. Gebert and C. Rojas-Molina. Lifshitz tails for random diagonal perturbations of Laurent matrices. Ann. Henri Poincaré, 23(11):4149–4170, 2022.
  • [33] F. Germinet, P. D. Hislop, and A. Klein. Localization for Schrödinger operators with Poisson random potential. J. Eur. Math. Soc. (JEMS), 9(3):577–607, 2007.
  • [34] F. Germinet and A. Klein. Bootstrap multiscale analysis and localization in random media. Comm. Math. Phys., 222(2):415–448, 2001.
  • [35] F. Germinet and A. Klein. Operator kernel estimates for functions of generalized Schrödinger operators. Proc. Amer. Math. Soc., 131(3):911–920, 2003.
  • [36] C. Hainzl, M. Lewin, and C. Sparber. Ground state properties of graphene in Hartree-Fock theory. J. Math. Phys., 53(9):095220, 27, 2012.
  • [37] E. Hamza, R. Sims, and G. Stolz. Dynamical localization in disordered quantum spin systems. Comm. Math. Phys., 315(1):215–239, 2012.
  • [38] S. Jitomirskaya and X. Zhu. Large deviations of the Lyapunov exponent and localization for the 1D Anderson model. Comm. Math. Phys., 370(1):311–324, 2019.
  • [39] W. Kirsch. An invitation to random Schrödinger operators. In Random Schrödinger operators, volume 25 of Panor. Synthèses, pages 1–119. Soc. Math. France, Paris, 2008. With an appendix by Frédéric Klopp.
  • [40] W. Kirsch, P. Stollmann, and G. Stolz. Anderson localization for random Schrödinger operators with long range interactions. Comm. Math. Phys., 195(3):495–507, 1998.
  • [41] A. Klein. Multiscale analysis and localization of random operators. In Random Schrödinger operators, volume 25 of Panor. Synthèses, pages 121–159. Soc. Math. France, Paris, 2008.
  • [42] A. Klein and J. F. Perez. Localization in the ground-state of the one-dimensional XX-YY model with a random transverse field. Comm. Math. Phys., 128(1):99–108, 1990.
  • [43] F. Klopp. Internal Lifshitz tails for Schrödinger operators with random potentials. J. Math. Phys., 43(6):2948–2958, 2002.
  • [44] F. Klopp and S. Nakamura. Spectral extrema and Lifshitz tails for non-monotonous alloy type models. Comm. Math. Phys., 287(3):1133–1143, 2009.
  • [45] S. Lahbabi. The reduced Hartree-Fock model for short-range quantum crystals with nonlocal defects. Ann. Henri Poincaré, 15(7):1403–1452, 2014.
  • [46] E. H. Lieb and B. Simon. The Hartree-Fock theory for Coulomb systems. Comm. Math. Phys., 53(3):185–194, 1977.
  • [47] R. Matos and J. Schenker. Localization and ids regularity in the disordered Hubbard model within Hartree-Fock theory. Comm. Math. Phys., 382(3):1725–1768.
  • [48] R. Mavi and J. Schenker. Localization in the disordered Holstein model. Comm. Math. Phys., 364(2):719–764, 2018.
  • [49] B. Nachtergaele, R. Sims, and G. Stolz. Quantum harmonic oscillator systems with disorder. J. Stat. Phys., 149(6):969–1012, 2012.
  • [50] J. Schenker. How large is large? Estimating the critical disorder for the Anderson model. Lett. Math. Phys., 105(1):1–9, 2015.
  • [51] B. Simon. Trace ideals and their applications, volume 120 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, second edition, 2005.
  • [52] G. Stolz. An introduction to the mathematics of Anderson localization. In Entropy and the quantum II, volume 552 of Contemp. Math., pages 71–108. Amer. Math. Soc., Providence, RI, 2011.
  • [53] H. von Dreifus and A. Klein. A new proof of localization in the Anderson tight binding model. Comm. Math. Phys., 124(2):285–299, 1989.
  • [54] W.-M. Wang. Localization and universality of Poisson statistics for the multidimensional Anderson model at weak disorder. Invent. Math., 146(2):365–398, 2001.