TRANSPORT PROPERTIES OF RANDOM SCHR ¨ODINGER OPERATORS ON CORRELATED ENVIRONMENTS. By Rodrigo Bezerra de Matos A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Mathematics – Doctor of Philosophy 2020 ABSTRACT TRANSPORT PROPERTIES OF RANDOM SCHR ¨ODINGER OPERATORS ON CORRELATED ENVIRONMENTS. By Rodrigo Bezerra de Matos This Ph.D. thesis presents recent developments in the theory of random Schr¨odinger op- erators. Differently from what is often studied in the subject, our main results consider potentials which are not independent at distinct sites but, rather, display some form of long range correlation. These are natural objects to investigate if one wishes to understand the long term behavior of a single particle which evolves in a disordered environment but also interacts with different members of this environment (other particles, spins, etc). In chapter 2 it is shown that, within the Hartree-Fock approximation for the disordered Hubbard Hamiltonian, weakly interacting Fermions at positive temperature exhibit localiza- tion, suitably defined as exponential decay of eigenfunction correlators. Our result holds in any dimension in the regime of large disorder and at any disorder in the one dimensional case. As a consequence of our methods, we are able to show H¨older continuity of the integrated density of states with respect to energy, disorder and interaction using known techniques. This is based on joint work with Jeffrey Schenker [46] Chapter 3 is based on joint work with Jeffrey Schenker and Rajinder Mavi. There we present simple, physically motivated, examples where small geometric changes on a two- dimensional graph G, combined with high disorder, have a significant impact on the spectral and dynamical properties of the random Schr¨odinger operator −AG +Vω obtained by adding a random potential to the graph’s adjacency operator. Differently from the standard An- derson model, the random potential will be constant along vertical line, hence the models exhibit long range correlations. Moreover, one of the models presented here is a natural example where the transient and recurrent components of the absolutely continuous spec- trum, introduced by Avron and Simon in [9], coexist and allow us to capture a sharp phase transition present in the system. Copyright by RODRIGO BEZERRA DE MATOS 2020 To Marsya, Sandra and Cory. v ACKNOWLEDGEMENTS I would like to thank Jeff Schenker for many things, in particular for supervising the projects which compose this thesis. The importance of his guidance during my Ph.D. goes well beyond research, travel support and recommendation letters. Jeff is an extremely generous and open-minded individual from whom I learned a lot. I am thankful to Russell Schwab, firstly for being my initial contact at Michigan State University and, secondly, for supervising my readings and research projects in partial differential equations. Russell has been very patient with me and encouraged me during the last five years. I am also thankful to Ilya Kachkovskiy for the reading groups and seminars organized by him and for the interesting discussions we have had in the last three years, from which I have certainly benefited a lot. I thank Jun Kitagawa and Willie Wong for the excellent graduate courses they have taught and for always being available and helpful. I would like to thank Tsveta Sendova, Rachael Lund and Andy Krause for supporting my teaching and making my experience as an instructor much easier. I have been lucky to make a number of friends at MSU, specially Abhishek Mallick, Leonardo Abbrescia, Christos Grego, Ioannis Zachos, Reshma Menon, Wenchuan Tian, Zak Tilocco and Andr´es Galindo. I have also met amazing individuals while traveling for confer- ences. By naming Wencai Liu I thank them all. I would like to express my gratitude to Fabio Montenegro, for his guidance during my undergraduate education and encouragement to study abroad. I have no words to describe how much I learned from him and his attitude towards mathematics. I thank Rafael Alves for ten years of friendship, starting from the first semester of college. Being able to often exchange ideas with him during this period was very important to me. This journey would not have happened without love and support from my parents, Sandra and Cory, and brothers Bruno and Pedro. By naming them I thank my whole family for everything they did to me. vi I am specially grateful to Marsya, who made my life brighter from the first day we met. Her love, patience and support during the final years of this program were crucial to me. Her intelligence, dedication and grace are a constant source of inspiration. vii TABLE OF CONTENTS LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . CHAPTER 1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 A short introduction to Anderson Localization . . . . . . . . . . . . . . . . . 1.2 Dynamical localization at large disorder . . . . . . . . . . . . . . . . . . . . 1.3 Green’s function decay: A step towards dynamical localization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4 First proof theorem 1.2: the self-avoiding walk expansion . . . . . . . . . . . 1.5 Second proof of theorem 1.2: the ’depleted’ resolvent identity . . . . . . . . . 1.5.1 Further Aspects of A-priori bounds . . . . . . . . . . . . . . . . . . . 1.6 From Green’s function decay to dynamical localization . . . . . . . . . . . . 1.6.1 Independence as a key tool . . . . . . . . . . . . . . . . . . . . . . . . Instructive consequences of dynamical localization . . . . . . . . . . . . . . . 1.7.1 From dynamical localization to finite averaged moments . . . . . . . 1.7.2 From dynamical localization to pure point spectrum . . . . . . . . . . Localization centers and Eigenfunction decay . . . . . . . . . . . . . 1.7.3 1.8 Spectral Measures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.7 CHAPTER 2 2.2 Definitions and Statement of the Main Result LOCALIZATION AND IDS REGULARITY IN THE HUBBARD MODEL WITH HARTREE-FOCK THEORY . . . . . . . . . . . . . 2.1 Chapter Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.1 Discussion of the results and main obstacles . . . . . . . . . . . . . . 2.1.2 Hartree-Fock theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1.3 Background on Localization for interacting systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2.2 Definition of the operators . . . . . . . . . . . . . . . . . . . . . . . . 2.2.3 Main Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.4 Outline of the Proof of theorem 2.2 . . . . . . . . . . . . . . . . . . . . . . . 2.5 Existence of the effective potential . . . . . . . . . . . . . . . . . . . . . . . . 2.6 Regularity of the effective potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.7 Decay estimates for the effective potential Improvements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.8 Proof of lemma 2.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.9 The Hartree approximation for the Hubbard model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.9.1 Existence of the Effective potential 2.9.2 Regularity of the effective potential . . . . . . . . . . . . . . . . . . . 2.9.3 Decay estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.10 One dimensional Aspects:proof of theorem 2.1 . . . . . . . . . . . . . . . . . 2.11 One dimensional aspects: strategy of the proof of theorem 2.1 . . . . . . . . 2.7.1 x 1 1 5 6 6 8 9 10 11 11 11 12 13 14 16 16 17 18 18 19 19 20 22 24 26 28 32 33 40 44 48 49 49 50 51 51 viii 2.11.1 Main ideas in the i.i.d case . . . . . . . . . . . . . . . . . . . . . . . . 2.11.2 Modifications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.11.3 Proof of Lemma 2.21 . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.11.4 Proof of lemma 2.22 . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.11.5 Proof of (2.11.22) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.12 H¨older Continuity for the integrated density of states at weak interaction . . CHAPTER 3 SPECTRAL AND DYNAMICAL CONTRAST ON HIGHLY COR- 3.1 RELATED ANDERSON-TYPE MODELS. . . . . . . . . . . . . . . Introduction and Main results . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.1 Background on Correlated Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.2 Main Results 3.1.3 Dynamical Contrast between HSym and HDiag . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1.4 Spectral contrast between HSym and HDiag 3.1.5 Phase transition within σ(cid:0)HSym (cid:0)HSym (cid:1) . . . . . . . . . . . . . . . . . . . . (cid:1) 3.1.6 Definition of the Models . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Lower Bound on the Averaged Moments for HSym: Proof of Theorem 3.1 . . 3.3 The absolute continuity of µδ(0,0) . . . . . . . . . . . . . . . . . . . 3.4 Floquet Theory for HSym . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4.1 Localization in the horizontal direction . . . . . . . . . . . . . . . . . 3.4.2 Transient and recurrent components: proof of theorem 3.7 . . . . . . 3.5 Absence of Diffusion for GDiag,γ: Proof of theorem 3.2 . . . . . . . . . . . . 3.6 Upper bounds for quantum dynamics revisited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6.1 Paking Dimension: Proof of Corollary 3.3 3.7 A Version of Boole’s Equality for level sets of Heglotz functions: proof of proposition 3.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.8 A priori bounds on the Green’s function . . . . . . . . . . . . . . . . . . . . 52 53 57 59 61 62 68 68 68 69 70 71 72 72 73 76 77 79 81 84 85 87 88 90 APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 ix LIST OF FIGURES Figure 2.1: The Hubbard Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 Figure 3.1: 1/2 spin and single-particle model . . . . . . . . . . . . . . . . . . . . . . 68 Figure 3.2: Highly correlated Anderson-type models . . . . . . . . . . . . . . . . . . 69 x CHAPTER 1 INTRODUCTION 1.1 A short introduction to Anderson Localization The effects of disorder on transport properties of quantum systems have drawn a signif- icant amount of attention in the mathematics community since their introduction in 1958 in the celebrated paper [8] by the Physicist Philip Anderson, who later in 1977 became a Nobel prize winner. The efforts to encode Anderson’s claim that randomness may prevent diffusion into a rigorous mathematical statement and to obtain a proof for it gave rise to a beautiful theory, incomplete up to these days. Absence of diffusion in the disordered context is now generically referred to as Anderson Localization which we shall briefly review in this chapter, which aims at introducing the main concepts and results relevant for the remainder of the thesis. For a more complete historical picture we refer to the survey [58] and the book [6]. The proofs for well-known results given in this introduction are deeply influenced by the presentation in these two sources. The Anderson model is the family of operators Hω = −∆ + λVω acting on (cid:96)2(cid:16)Zd(cid:17) • ∆ is the discrete Laplacian acting on ϕ ∈ (cid:96)2(cid:16)Zd(cid:17) through , the space of functions f : Zd → C which are square summable, where (cid:88) |m−n|1=1 (∆ϕ) (n) = (ϕ(m) − ϕ(n)) , n ∈ Zd, |n|1 = |n1| + ...|nd|. . • The on site potential Vω acts as a multiplication operator on (cid:96)2(cid:16)Zd(cid:17) via (Vωϕ) (n) = ω(n)ϕ(n) , n ∈ Zd. 1 • ω = {ω(n)} n∈Zd is a list of independent, identically distributed random variables. • λ > 0 denotes the disorder strength. For technical reasons one often includes the following regularity condition on the random variables P (ω(0) ∈ I) = (cid:90) I ρ(v) dv. (1.1.1) A typical assumption on the density ρ is requiring it to be bounded and of compact sup- port, which makes the random variable ω(n) bounded for every n ∈ Zd. Sometimes it is more convenient to work with distributions with unbounded support such as the Gaussian distribution given by ρ(v) = 1√ π e−v2 1 (cid:96)2(cid:16)Zd(cid:17) and the Cauchy distribution given by ρ(v) = The lattice structure of Zd allow us to make use of the canonical basis {δn} π(1+v2) ninZd of defined by δn(m) = δmn where δmn is the Kronecker delta. The probability of . finding a particle at position n and time t, given that it started at position 0 is given by |(cid:104)δn, e−itH δ0(cid:105)|2. Dynamical Localization is naturally defined as averaged decay of the matrix elements |(cid:104)δn, e−itH δ0(cid:105)|. which can be explicit in a bound as E(cid:16)|(cid:104)δn, e−itH δ0(cid:105)|(cid:17) ≤ r(n). (1.1.2) where (cid:88) n∈Zd r(n) < ∞. In the best scenario, the function r(n) was shown to have exponential decay. In dimension d ≥ 2, dynamical localization has been achieved at large disorder, meaning that λ is taken sufficiently large or at weak disorder and extreme energies, see [6, Theorem 10.4] for a precise 2 statement. We shall outline a proof of localization at large disorder in the following sections of this introduction. If one is only concerned whether diffusion takes place or not in a system, a useful tool is given by the averaged q-moments of the position operator defined as M q T (H) = 2 T −2t T E(cid:104)0|eitHω|X|qe−itHω|0(cid:105) dt. e (cid:90) ∞ 0 (cid:88) Where the position operator X is formally defined as (Xϕ)(n) = |n|ϕ(n). For instance, whenever |n|qr2(n) < C < ∞ dynamical localization in the above sense implies the bound n∈Zd T (H) ≤ C < ∞ M q The above inequality is a convenient signature of Anderson localization, whereas its coun- terpart T (H) ≥ CT α M q for α > 0 indicates diffusion. Finally, recall that associated to the action of H in the Hilbert space (cid:72) = (cid:96)2(cid:16)Zd(cid:17) is the decomposition (cid:72) = (cid:72)pp ⊕ (cid:72)sc ⊕ (cid:72)ac of (cid:72) into the pure point, singular continuous and absolutely continuous subspaces and the corresponding splitting of its spectrum into components σ(H) = σac(H)∪ σsc(H)∪ σpp(H). The so called RAGE theorem (after Ruelle, Amrein, Georgescu and Enss.[6, Theorem 2.6]) provides dynamical characterizations for the spectral components. In particular, one of its consequences is the fact that dynamical localization in the form of (1.1.2) implies that H has only pure point spectrum. In other words, dynamical localization implies that the spectral projection of H into the continuous component (cid:72)c = (cid:72)sc ⊕ (cid:72)ac vanishes. This means that the spectrum σ(H) is the closure of the set of eigenvalues for H. If the associated eigenfunctions decay exponentially, the operator 3 H is said to exhibit Anderson Localization. A sufficient condition is the exponential decay of the eigenfunction correlators E (|(cid:104)n|g(Hω)|0(cid:105)|) ≤ Ce−µ|n| sup |g|≤1 (1.1.3) where the above supremum is taken over all Borel measurable functions g : R → C and (cid:104)n|g(Hω)|0(cid:105) = (cid:104)δn, g(Hω)δ0(cid:105) (in this thesis we may alternate between these two notations for the inner product). For proofs for some of the above facts and more precise statements In the series of works by Jakˇsi´c and Molchanov [38], [37] and [39] the authors consider a with randomness being introduced only = Zd through independent, identically distributed random we refer to the next few sections of this introduction. (cid:16)Zd × Z+(cid:17) Schr¨odinger operator H = −∆+Vω on (cid:96)2(cid:16)Zd × Z+(cid:17)  0 if n2 (cid:54)= 0. along the surface ∂ variables {ω(m)} m∈Zd. More precisely, let Vω(n1, n2) = ω(n1) if n2 = 0. In [37] (see also remark (3) in [39]), for d = 1 a sharp dichotomy is shown as σpp(H) = σ(H) \ σ(−∆) , σac(H) = σ(−∆) (1.1.4) holds almost surely. An important fact, which will allow us to obtain an analogue of (1.1.4) in the correlated context later in this thesis, is that the absolutely continuous subspace can be further de- composed into its transient and recurrent subspaces (cid:72)ac = (cid:72)tac ⊕ (cid:72)rac, an idea which was introduced by Avron and Simon in [9]. As explained there, one of the motivations is that in case µψ = χC dx, where C is a Cantor-like set of positive Lebesgue measure, the measure µψ resembles a singular measure, despite its absolute continuity. This will be a typical situation in which ψ belongs to the recurrent subspace (cid:72)rac. The subspace (cid:72)tac is the closure of the set of all ψ ∈ (cid:72)ac such that, for all N ∈ N . (1.1.5) (cid:12)(cid:12)(cid:12)(cid:104)ψ|e−itH|ψ(cid:105)(cid:12)(cid:12)(cid:12) = O (cid:16) t−N(cid:17) 4 Recall that, by the Riemann-Lebesgue lemma, if ψ ∈ (cid:72)ac, (cid:12)(cid:12)(cid:12)(cid:104)ψ|e−itH|ψ(cid:105)(cid:12)(cid:12)(cid:12) = 0 lim t→∞ hence, considering the spectral measure µψ associated to ψ in (cid:72)tac, we see that its density with respect to Lebesgue measure dµψ dE must have additional regularity properties. In fact, according to [9, Proposition 3.1] if ψ is a transient vector then µψ = f (x) dx where f is a smooth function. The converse statement is also true under the additional assumption of f being compactly supported. We will show in theorem 3.7 below that the transient and recurrent spectrum can naturally arise and coexist in a situation of physical relevance. 1.2 Dynamical localization at large disorder We let H = −∆ + λVω for the remainder of this chapter, unless otherwise specified. In terms of background for this thesis, the following result is of great relevance. Strictly speak- ing, the version presented here was proven first by Aizenman in [1] although localization at large disorder was shown earlier by Aizenman and Molchanov [5]. The technical assumptions on the random potential Vω will be clear when we outline a proof of this result within a few paragraphs. Theorem 1.1. There is a λ > 0 such that for all λ > λ0 (cid:18) E sup t (cid:19) |(cid:104)δn, e−itH δ0(cid:105)| ≤ Ce−ν|n|. for all n ∈ Zd and some C < ∞ and ν > 0. As mentioned in the above section, theorem 1.1 means that, on average, the probability of finding the wavepacket at a site n at any time is exponentially small. This is the strongest form of single particle localization, known as exponential dynamical localization. We shall present a version of this theorem for the interacting case in chapter 2. A related statement can be achieved via multiscale analysis although with a subexponential rate, see [28]. Some general consequences of dynamical localization will be detailed in section 1.7. 5 1.3 Green’s function decay: A step towards dynamical localization. An object which is often helpful to the analysis of random operators is the Green’s function. It is defined at z ∈ C by G(m, n; z) = (cid:104)δm, (H − z)−1δn(cid:105) when this expression makes sense. In particular, since the operators object of our study are self-adjoint, this will be the case for all z ∈ C \ R. The first step to show dynamical localization as in theorem 1.1 is obtaining exponential decay for the above quantity as follows. Theorem 1.2. Given s ∈ (0, 1) and λ sufficiently large there exist positive constants CAnd and µAnd such that for all z ∈ C+ E (|G(m, n; z)|s) ≤ CAnde−µAnd|m−n| (1.3.1) We shall outline two different proofs for the above fact, from which the relationship between the aforementioned constants and the parameters λ and s will be more clear. 1.4 First proof theorem 1.2: the self-avoiding walk expansion The proposition below is the self-avoiding walk expansion of the Green’s function, also called Feenberg’s loop-erased expansion (since it seems to date back to Eugene Feenberg in the 1940s) or the locator expansion (as in Anderson’s seminal work [8]). It will be useful not only for this chapter but also on chapter 3, when applied to a graph which is different from Zd. For this reason we keep the general formulation as in [6], i.e, replacing Zd by an arbitrary vertex set G of a graph with finite degree, denoted henceforth by degG. Proposition 1.3. The Green’s function of a self-adjoint operator H = H0 + V on (cid:96)2 (G) admits the expansion (−1)|τ|H0(τ (0), τ (1))...H0(τ (|τ| − 1), τ (|τ|)) × Gk(z). On any graph, the above expansion converges whenever Imz > (cid:107)T(cid:107). On finite graphs it is convergent for all z ∈ C+. SAW(cid:88) τ :m(cid:55)→n G(m, n; z) = |τ|(cid:89) k=0 6 Here Gk(z) = (cid:104)δτ (k), (Hτ,k − z)−1δτ (k)(cid:105) where Hτ,k denotes the restriction of H to the set (cid:96)2(cid:0)G \ ∪j 0, s ∈ (0, 1) and z ∈ C+ we have E (|G(n, n; z)|s) ≤ CAP(s) λs . where CAP(s) = 2s(cid:107)ρ(cid:107)s∞ 1−s . Proof. Letting m = n in (1.4.3) we find that G(n, n; z) = 1 λω(n) − Σ(z) 7 (1.4.3) (1.4.4) (1.4.5) where Σ(z) = −(cid:16) ˆG(n, n; z) (cid:17)−1 is a complex number independent of ω(n). Let v = ω(n). Under the assumption that ρ ∈ L∞ (R) we have that for any interval I ⊂ R P (v ∈ I) ≤ (cid:107)ρ(cid:107)∞|I|. (1.4.6) Combining the above observations with the layer cake representation we discover that (cid:90) ∞ −∞ 1 0 0 P({|v − ξ| ≤ t−1/s})dt min{1, 2(cid:107)ρ(cid:107)∞t−1/s}dt (cid:90) ∞ (cid:90) ∞ |v − ξ|s ρ(v)dv = ≤ = 2s(cid:107)ρ(cid:107)s∞ + 2(cid:107)ρ(cid:107)∞ 2s(cid:107)ρ(cid:107)s∞ = 2s(cid:107)ρ(cid:107)s∞ + 2(cid:107)ρ(cid:107)∞ s 1 − s = 2s(cid:107)ρ(cid:107)s∞ + 2(cid:107)ρ(cid:107)∞ s 1 − s (cid:90) ∞ t−1/sdt s−1 s ((2(cid:107)ρ(cid:107)∞)s) (2(cid:107)ρ(cid:107)∞)s−1 2s(cid:107)ρ(cid:107)s∞ 1 − s . = finishing the proof of the a-priori bound. It readily follows from Feenberg’s expansion that whenever λ > CAP degG E (|G(m, n; z)|s) ≤ ∞(cid:88) degkG k=|m−n| (cid:18) CAP λs (cid:19)k+1 (cid:18)degG CAP λs (cid:19)|m−n| finishing the proof of theorem 1.2. = CAP λs 1 1 − degG CAP λs 1.5 Second proof of theorem 1.2: the ’depleted’ resolvent identity Let us now provide another proof of theorem 1.2, based on the so called ’depleted’ resol- vent identity which, for m (cid:54)= n, reads GΛ(m, n; z) = −GΛ(m, m; z) H0(m, n(cid:48))GΛ\{m}(n(cid:48), n; z). (cid:88) n(cid:48)(cid:54)=m 8 where (cid:104)m|(HΛ − z)−1|n(cid:105) and HΛ denotes the restriction of H to a box of size |Λ| centered at the origin. Using the a-priori bound and that GΛ\{m}(n(cid:48), n; z) is independent of m we reach E (|GΛ(m, n; z)|s) ≤ CAP λs (cid:88) n(cid:48) |H0(m, n(cid:48))|sE(cid:16)|GΛ\{m}(n(cid:48), n; z)|s(cid:17) Exponential decay is now achieved whenever with λs > CAP(cid:107)H0(cid:107)∞,∞ (cid:107)H0(cid:107)∞,∞ := sup m (cid:88) n |H0(m, n)|. . (1.5.1) (1.5.2) (1.5.3) through a suitable iteration of (1.5.1). A Couple of Remarks are in order: • Note that both proofs make use of independence but that the existence of a conditional density is enough. • The second of the above proofs also allows H0 to be of the general form |H0(m, n)|(cid:16) (cid:17) eν|m−n| − 1 sup m (cid:88) n < η < ∞. (1.5.4) In case the matrix elements of H0 only exhibit polynomial decay, the argument given above yields polynomial decay of the Green’s function. • In [47], a refined estimate for the large disorder threshold is presented. Remarkably, it coincides with Anderson’s original prediction of the large disorder regime. Further comments and details may be found on [6, Section 10.1]. 1.5.1 Further Aspects of A-priori bounds The estimates for the Green’s function given in this section heavily rely on the rank-one formula (1.4.3). An alternative, and more general, approach, originally due to [2], which is useful in the continuum and also for the operators studied on chapter 3 will be given in section 3.8 of that chapter. 9 1.6 From Green’s function decay to dynamical localization The next result we shall review explains why, in the context of the Anderson model, it is enough to shown Green’s function decay in order to obtain dynamical localization. A proof will not be given here, but related ideas are presented in chapter 3, section 3.4.1. The original argument is due to Aizenman in [1] with a streamlined version being presented in [3, Appendix A]. We will say that a random operator Hω has uniform Green’s function decay on an energy domain I ⊂ R when given some s ∈ (0, 1) there exist C > 0 and µ > 0 such that E(cid:16)|GΛ(m, n, E)|s(cid:17) ≤ Ce−µ|m−n| (1.6.1) holds for all Λ ⊂ Zd and every E ∈ I. Theorem 1.5. Let Hω = H0 +Vω be a random operator with the hopping term H0 satisfying (1.5.4). Assume that • E(|Vω(n)|δ) < ∞ holds for some δ > 0. • For every n0 ∈ Zd the conditional distribution of Vω(n0) = v at specified values of {Vω(n)}n(cid:54)=n0 has a density ρn0(v) and sup ω∈Ω sup n0∈Zd sup v∈R ρn0(v) < ∞. Suppose further that for some s ∈ (0, 1) the uniform Green’s function decay (1.6.1) is satisfied in some interval I ⊂ R. Then, there exist C(cid:48) > 0 and µ(cid:48) > 0 such that ≤ C(cid:48)e−µ(cid:48)|m−n| (cid:33) (cid:104)m|f (Hω)PI (Hω)|n(cid:105) (1.6.2) (cid:32) E sup |f|≤1 where the supremum is taken over all Borel measurable functions f : R → C bounded by one. 10 1.6.1 Independence as a key tool When establishing localization by the above arguments, independence of the random vari- ables {ω(n)} n∈Zd is important in many steps. Strict independence can, at most, be re- placed by an assumption of a weaker form of independence which cannot be removed. This is not simply a technical issue related to fractional moments. For instance, the robust multiscale analysis technique is usually formulated under the assumption of independence at distance, meaning that there exists a ρ > 0 such that events based on boxes ΛL1 are independent of events based on boxes ΛL2 ΛL(m) = {m(cid:48) ∈ Zd : |m − m(cid:48)|∞ < L (m), ΛL2 2 }, see [41] for a survey in the subject. (n) if dist ΛL1 (m) (n) > ρ. Here (cid:16) (cid:17) 1.7 Instructive consequences of dynamical localization Let us point out some consequences of theorem 1.1. More precisely, the facts presented in this section will follow from dynamical localization as in (1.6.2). 1.7.1 From dynamical localization to finite averaged moments Let M q T (H) = 2 T (cid:90) ∞ 0 T E(cid:16)(cid:104)0|eitHω|X|qe−itHω|0(cid:105)(cid:17) −2t e dt. be the averaged q-moments of the position operator X, formally defined as (Xϕ)(n) = |n|ϕ(n). The following proposition holds true. Proposition 1.6. Assume that dynamical localization in the sense of (1.6.2) holds. Then for some finite constant C. T (H) ≤ C M q 11 Proof. By definition and from (1.6.2), we have that M q T (H) = (cid:90) ∞ ≤ C(cid:48) (cid:88) 2 T 0 e −2t T (cid:88) n∈Zd |n|qe−µ(cid:48)|n| |n|q|E(cid:16)(cid:104)δn, e−itH δ0(cid:105)|2(cid:17) dt. n∈Zd where the above quantity is finite. 1.7.2 From dynamical localization to pure point spectrum For the convenience of the reader, we shall recall the definition of the spectral types below in section 1.8. The next consequence of dynamical localization we would like to point out is the following. Proposition 1.7. Suppose that dynamical localization as described by (1.6.2) holds on an interval I ⊂ R. Then, almost surely, Hω has only pure point spectrum in I. Proof. By the RAGE theorem (see [6, Theorem 2.6]), given ϕ ∈ (cid:96)2(cid:16)Zd(cid:17) and denoting by P c I the projection onto the continuous component of H within I we have that (cid:107)P c I ϕ(cid:107)2 = lim Let ϕ = δn for some n ∈ Zd. Then, L→∞ lim T→∞ 1 T (cid:90) T 0 (cid:107)1 Λc L e−itH PI (H)ϕ(cid:107)2 dt. (1.7.1) (cid:107)1 Λc L e−itH PI (H)ϕ(cid:107)2 ≤ (cid:107)1 ≤ (cid:107)1 Λc L e−itH PI (H)Pn(cid:107)2(cid:107)δn(cid:107)2 e−itH PI (H)Pn(cid:107) (cid:12)(cid:12)(cid:104)δm, e−itH PI (H)δn(cid:105)(cid:12)(cid:12). Λc L ≤ (cid:88) |m|>L 12 assumption (1.6.2) (cid:88) Taking expectations, using Fatou’s lemma and Fubini’s theorem, we conclude that under the (cid:107)P c I ϕ(cid:107)2 ≤ C(cid:48) e−µ(cid:48)|m−n| = 0. The general result follows from density of finitely supported functions on (cid:96)2(cid:16)Zd(cid:17) lim L→∞ |m|>L (1.7.2) . 1.7.3 Localization centers and Eigenfunction decay Lastly, let us discuss the phenomenon on engenfunction decay in the regime of dynamical localization. Definition 1.8. Given a Borel set I ⊂ R, the eigenfunction correlator on I is defined by Q(m, n; I) = sup |f|≤1 (cid:104)m|f (Hω)PI (Hω)|n(cid:105) . (1.7.3) with the supremum being taken over Borel measurable functions. If H has only pure point spectrum in I, we have Q(m, n; I) = |ϕE,m(m)||ϕE,m(n)| (1.7.4) (cid:88) E∈I where ϕE,m is the normalized eigenfunction within the cyclic subspace (cid:72)H,δm associated to the eigenvalue E. Recall that, in general, (cid:72)H,v = span{(H − z)−1v : z ∈ C \ R}. (1.7.5) In case H is a bounded operator, the above set coincides with the closure of the set spanned by vectors of the form q(H)v where q is a polynomial, see [19, lemma 2.4.3]. The expression (1.7.4) is obtained expanding (1.7.3) into the eigenfunction basis. Definition 1.9. Given ϕ ∈ (cid:96)2(cid:16)Zd(cid:17) with (cid:107)ϕ(cid:107)2 = 1, a point n0 ∈ Zd is said to be a localization center for ϕ if :=(cid:80) where C−1 d |ϕ(n0)|2 ≥ Cd (1 + |n0|)d+1 (1.7.6) n∈Zd 1 (1+|n0|)d+1 . 13 It is immediate to check that any such ϕ admits a localization center. Proposition 1.10. Assume that for some ν > 0. (cid:88) m Cd (1 + |m|)d+1 n (cid:88) eν|m−n|E(cid:16) (cid:17) Q2(m, n; I) < ∞. (1.7.7) Then, almost surely, for any simple eigenvalue E ∈ I of H there is a localization center nE(ω) such that the normalized eigenfunction ϕE,ω satisfies |ϕE,ω(n)| ≤ C(I, ω)(1 + |n|)d+1e − ν|n−nE| 2 . (1.7.8) for all n ∈ Zd. Moreover, ω (cid:55)→ C2(I, ω) belongs to L1 (Ω; P). Proof. Let (cid:32)(cid:88) m (cid:88) n Cd (1 + |m|)d+1 C(I, ω) := (cid:33)1/2 eν|m−n|Q2(m, n; I) . (1.7.9) By assumption, C2(I, ω) belongs to L1 (Ω; P) hence it is a finite quantity almost surely. By proposition 1.7, the spectrum of H is, almost surely, pure point in I. Moreover, by (1.7.4), any simple eigenfunction of H satisfies |ϕE,m(m)|2|ϕE,m(n)|2 ≤ C2(I, ω)C−1 d (1 + |n|)d+1e−ν|n−nE|. (1.7.10) recalling the definition of localization center and letting m = nE we reach |ϕE,m(n)|2 ≤ C2(I, ω)C−2 d (1 + |n|)2d+2e−ν|n−nE|. (1.7.11) which yields the desired result after taking square roots. 1.8 Spectral Measures Finally, we conclude this introduction with a few notions on spectral theory. Let (cid:72) be a Hilbert space and H a self-adjoint operator on (cid:72). Given ψ ∈ (cid:72) , the spectral measure µψ is the unique finite Borel measure on R satisfying, for all z ∈ C+, (cid:90) 1 t − z dµψ(t). R (cid:104)ψ, (H − z)−1ψ(cid:105) = 14 The existence and uniqueness of µψ follows from the representation theorem for Herglotz functions (see [6, Appendix B]) since F (z) = (cid:104)ψ, (H − z)−1ψ(cid:105) (1.8.1) defines a map F : C+ → C+ whenever ψ (cid:54)= 0. By the Radon-Nikodym theorem, we can decompose µψ (with respect to Lebesgue measure) according to its pure point, singular continuous and absolutely continuous components. µψ = µpp ψ + µsc ψ + µac ψ Associated to the above decomposition is the decomposition of the Hilbert space where and (cid:72) = (cid:72)pp ⊕ (cid:72)ac ⊕ (cid:72)sc (cid:72)pp = {ψ |µψ = µpp ψ } , (cid:72)sc = {ψ |µψ = µsc ψ } (cid:72)ac = {ψ |µψ = µac ψ }. The spectrum of H, denoted by σ(H), is defined as the set of z ∈ C such that H − z does not have a bounded inverse and the spectral components σpp(H), σsc(H) and σac(H) are defined by restricting the operator to the subspaces (cid:72)pp,(cid:72)sc and (cid:72)ac. 15 CHAPTER 2 LOCALIZATION AND IDS REGULARITY IN THE HUBBARD MODEL WITH HARTREE-FOCK THEORY 2.1 Chapter Outline Our main goal in this chapter is to study Anderson localization in the context of in- finitely many particles. We shall formulate our results for the disordered Hubbard model within Hartree-Fock theory. However, as the techniques involved are quite flexible, we expect that similar statements can be made in a more general framework, under appropriate modi- fications of the decorrelation estimates in section 2.7.1. The (deterministic) Hubbard model under Generalized Hartree-Fock Theory has been discussed (at zero and positive tempera- ture) by Lieb, Bach and Solovej in [11] however, to the best of our knowledge, the localization properties of the disordered version of this model remained unexplored, even in the context of restricted Hartree-Fock theory, up to the present work. The main difficulty lies in the addition of a self-consistent effective field, which will be random and non-local by nature, to a random Schr¨odinger operator. The conclusion of this chapter can be summarized as follows: under technical assumptions, the results on (single-particle) Anderson localization obtained in the non-interacting setting in the regimes of large disorder (in dimension d ≥ 2) and at any disorder (in dimension d = 1), remain valid under the presence of sufficiently weak interactions. More specifically, in the regime of strong disorder this is accomplished in any dimension by theorem 2.2 below. Theorem 2.1 contains the improvement in dimension one, where any disorder strength leads to localization, provided the interaction strength is taken sufficiently small. Our methods contain various bounds in the fluctuations of the ef- fective interaction which are interesting on their own right and potentially useful in different contexts. To exemplify this, we prove H¨older regularity of the integrated density of states (IDS) with respect to various parameters by adapting arguments of [29], which is the content 16 of theorem 2.3. 2.1.1 Discussion of the results and main obstacles Mathematically, our setting can be understood as an Anderson-type model Hω = H0 + λUω where the values of the random potential U at different sites are correlated in a highly non- local and self-consistent fashion. The correlations are governed by a nonlinear function of Hω, as explained in section 2.3. In comparison to the recent result on Hartree-Fock theory for lattice fermions of [23], achieved via multiscale analysis, we use the fractional moment method to establish exponential decay of the eigenfunction correlators at large disorder in any dimension but also at any disorder in dimension one. In particular, in the above regimes, we show exponential decay (on expectation) of the matrix elements of the Hamiltonian evolution at all times, which means that, on average, |(cid:104)m|e−itH|n(cid:105)| decays exponentially on |m − n| for all t > 0. The result of complete localization in dimension one in such interacting context is new and deserves attention on its own. There, the main technical difficulty also lies in the non- local correlations of the potential, which means that standard tools such as Furstenberg’s theorem and Kotani theory are not available. Moreover, a large deviation theory for the Green’s function is a further obstacle to establishing dynamical localization even if one obtains uniform positivity of the Lyapunov exponent. We overcome these challenges using ideas of [6, Chapter 12], where arguments reminiscent of the proof of the main result in [43] are presented. We then obtain positivity of the Lyapunov exponent at any disorder using uniform positivity for the Lyapunov exponent of the Anderson model, combined with an explicit bound on how this quantity depends on the interaction strength, see theorem 2.19. When it comes to establishing a large deviation theorem, our modification of the argument in [6, Theorem 12.8] relies on quantifying the decorrelations on the effective potential, which is presented on lemma 2.22 in the form of a strong mixing statement. It is worth clarifying that, since our proof is based on fractional moments, we have not established localization 17 in one dimension for rough potentials as in [34]. Moreover, the gap assumption in [23] is replaced by working at positive temperature thus our results do not apply to Hartree-Fock ground states. 2.1.2 Hartree-Fock theory Hartree-Fock theory has been widely applied in computational physics and chemistry. It also has a rich mathematical literature which goes well beyond the scope of random operators, see for instance [31],[32],[11],[12],[33] and references therein. 2.1.3 Background on Localization for interacting systems The main results of this chapter lie in between the vast literature on (non-interacting) single particle localization and the recent efforts to study many particle systems, as in the case of an arbitrary, but finite, number of particles in the series of works by Chulaevsky-Suhov [15],[16],[17] and Aizenman-Warzel [7]. In comparison to the later, we only seek for a single- particle localization result but allow for infinitely many interactions, which occur in the form of a mean field. In comparison to the recent developments on spin chains, as the study of the XY spin chain in [30] and the droplet spectrum of the XXZ quantum spin chain in [25] and [13], the notions of localization for a single-particle effective Hamiltonian are more clear and can be displayed from pure point spectrum to exponential decay of eigenfunctions and exponential decay of eigenfunction correlators. The later is agreed to be the strongest form of single particle localization and it is what we end up accomplishing. If fact, as we have explained in the introduction of this thesis, dynamical localization in the form of theorems 2.1 and 2.2 implies pure point spectrum via the RAGE theorem and exponential decay of eigenfunctions, see also [58, Proposition 5.3] and [6, Theorem 7.2 and Theorem 7.4]. 18 2.2 Definitions and Statement of the Main Result 2.2.1 Notation In what follows, Zd will be equipped with the norm |n| = |n1| + ... +|nd| for n = (n1, ..., nd). n∈Λ |ϕ(n)|2 < ∞} and, for . Throughout this chapter, η will Given a subset Λ ⊂ Zd, we define (cid:96)2(Λ) := {ϕ : Λ → C|(cid:80) ϕ ∈ (cid:96)2(Λ), we let (cid:107)ϕ(cid:107)(cid:96)2(Λ) := (cid:0)(cid:80) n∈Λ |ϕ(n)|2 < ∞}(cid:1)1/2 be a positive constant and Fβ,κ will denote the Fermi-Dirac function at inverse temperature β > 0 and chemical potential κ, meaning that Fβ,κ(z) = 1 1 + eβ(z−κ) . (2.2.1) We shall omit the dependence of F on the above parameters whenever it is clear from the context. For many of our bounds, the specific form of (2.2.1) is not important and F could denote an arbitrarily chosen function which is analytic in the strip (cid:83) = {z ∈ C : |Imz| < η} and continuous up to the boundary of (cid:83), in which case we let (cid:107)F(cid:107)∞ := supz∈(cid:83) |F (z)|. For the function Fβ,κ in (2.2.1) one can take η = π 2β . However, to obtain robust results which incorporate delicate fluctuations, further properties of the Fermi-Dirac function are necessary. Namely, in section 2.7.1 we use the the fact that tF (t) is bounded as t → ∞ and that t(1− F (t)) is bounded t → −∞. These properties will also play a role in the decoupling estimates needed in the proof of theorem 2.1 but could be relaxed if one is only interested in the large disorder proof of theorem 2.2 for a specific distribution with heavy tails (for instance, the Cauchy distribution). Our main goal is to study localization properties of non-local perturbations of the Ander- son model HAnd := −∆ + λVω which naturally arise in the context of Hartree-Fock theory for the Hubbard model. The random potential Vω is the multiplication operator on (cid:96)2(Zd) defined as for all n ∈ Zd and {ωn} n∈Zd are independent, identically distributed random variables (Vωϕ) (n) = ωnϕ(n) (2.2.2) 19 on which we impose technical assumptions described in the next paragraph. The hopping operator ∆ is the discrete Laplacian on Zd, defined via (∆ϕ) (n) = (ϕ(m) − ϕ(n)) . (2.2.3) (cid:88) |m−n|=1 The proofs of localization via fractional moments usually do not require the hopping to be dictated by the Laplacian and, indeed, we will replace ∆ by a more general operator H0 whose matrix elements decay sufficiently fast away from the diagonal. It is technically useful to formulate some of our results in finite volume, i.e, we will work with restrictions of the operators to (cid:96)2(Λ) but the estimates obtained will be volume independent, meaning that all the constants involved are independent of Λ ⊂ Zd. We will use 1Λ to denote the characteristic function of Λ as well as the natural projection PΛ : (cid:96)2(Zd) → (cid:96)2(Λ). With these preliminaries we are ready to define the Schr¨odinger operators studied in this chapter. 2.2.2 Definition of the operators Let HAnd = H0 + λVω be the Anderson model on (cid:96)2(cid:16)Zd(cid:17) (cid:17) eν|m−n| − 1 |H0(m, n)|(cid:16) (A1) ζ(ν) := sup m (cid:88) n∈Zd where: η 2 , for some ν > 0 fixed. < (cid:90) (A2) Vω is defined as in (2.2.2) and the random variables {ω(n)} n∈Zd are independent, identically distributed with a density ρ: P (ω(0) ∈ I) = ρ(x) dx, for I ⊂ R a Borel set . (A3) We also assume that supp ρ = R with I ≥ e−c1(ρ)|x−x(cid:48)|(1+c2(ρ) max{|x|,|x(cid:48)|}) ρ(x) ρ(x(cid:48)) (2.2.4) for some c1(ρ) > 0, c2(ρ) ≥ 0 and any x, x(cid:48) ∈ R. 20 Before stating the remaining assumptions on ρ, we need to introduce some notation. Assume that ρ satisfies (2.2.4). Let ρ(x) = where h(x) = (cid:82) ∞ −∞ ρ(α)h(x − α) dα ρ(x) e−cρ|x| if c2(ρ) = 0. e−cρ|x|2 if c2(ρ) > 0. (2.2.5) (2.2.6)  (A4) The function ρ is bounded for some cρ > 0. Remark 1. The technical assumptions (A3)−(A4) will be needed for the large disorder result of theorem 2.2 below. They include, for instance, the Cauchy distribution, the Gaussian, and the exponential distribution ρ(v) = m 2 e−m|v|. Further details an related comments will be provided on the appendix of this thesis, section A.1. The above requirements will suffice to show localization at large disorder on theorem 2.2 below. To show complete localization in dimension one, theorem 2.1 will also require a moment condition on ρ, which is the following. (A5) For some ε > 0 and some cρ > 0,(cid:82) ∞ −∞ |x|ερ(x) dx < ∞. Remark 2. The assumption (A5) covers, for example, the Gaussian and the exponential distributions but it does not cover the Cauchy or other distribution with heavy tails. It will be necessary for the one dimensional result of theorem 2.1 below. More specifically, this requirement will imply a moment condition which will be needed to relate the Green’s function to the Lyapunov exponent, see sections 2.11 and 2.11.5. Remark 3. The specific bound on ζ(ν) is necessary to ensure that the Combes-Thomas bound |G(m, n; E + iη)| ≤ 2 (cid:104)m|(H − z)−1|n(cid:105), whenever this quantity is defined. η e−ν|m−n| holds [6, Theorem 10.5], where G(m, n; z) denotes 21 0 HHub =  H↑(ω) Define the operator HHub, acting on (cid:96)2(cid:16)Zd(cid:17) ⊕ (cid:96)2(cid:16)Zd(cid:17)  H0 + λVω + gV↑(ω)  := where the operators H↑(ω) and H↓(ω) act on (cid:96)2(cid:16)Zd(cid:17)  V↑(ω)(n)  =  (cid:104)n|F (H↓)|n(cid:105)  . are defined via 0 H↓(ω) V↓(ω)(n) (cid:104)n|F (H↑)|n(cid:105) 0 H0 + λVω + gV↓(ω) and the so-called effective potentials by 0  (2.2.7) (2.2.8) Note that the above equations only define H↑(ω) and H↓(ω) implicitly. Existence and uniqueness of V↑ and V↓ will be shown in section 2.9.1 via a fixed point argument. The model (2.2.7) is usually referred to as the Hartree approximation, due to the absence of exchange terms. In section 2.3 below we will show that the terminology Hartree-Fock approximation is justified when g < 0, which represents a repulsive interaction. The Hubbard model is schematically represented in the following picture. The black (horizontal) edges represent hopping between sites and the red (vertical) edges represent the effective interaction between the two layers, which are identical copies of Zd. Figure 2.1: The Hubbard Model n↑ n↓ (0,0) 2.2.3 Main Theorems Fix an interval I ⊂ R and define the eigenfunction correlator through (cid:0)|(cid:104)m|ϕ(H↑)|n(cid:105)| + |(cid:104)m|ϕ(H↓)|n(cid:105)|(cid:1) . QI (m, n) := sup |ϕ|≤1 (2.2.9) 22 The operators H↑ and H↓ are defined as in (2.2.7) and the supremum is taken over Borel measurable functions bounded by one and supported on the interval I. In case I = R we simply write Q(m, n). Our first result is the following: Theorem 2.1. In dimension d = 1, let H0 = −∆ and assume that the conditions (A1)−(A5) hold. For any λ > 0 and any closed interval I ⊂ R , there is a constant g1 > 0 such that whenever |g| < g1 we have E (QI (m, n)) ≤ Ce−µ1|m−n|. (2.2.10) for any m, n ∈ Zd and positive constants µ1 = µ1(λ, ν, η, I), C(η, g, λ,(cid:107)F(cid:107)∞, I). Theorem 2.2. Suppose that the conditions (A1) − (A4) hold. For any dimension d ≥ 1, there exists a constant gd = g(d, η,(cid:107)F(cid:107)∞, ν) such that, whenever |g| < gd, there is a positive constant λ0(g) for which E (Q(m, n)) ≤ Ce−µd|m−n|. (2.2.11) holds for λ > λ0(g), any m, n ∈ Zd and some positive constants µd = µ(d, λ, g, ν, η), C(η, ν, d, g, λ,(cid:107)F(cid:107)∞). (cid:16) Remark 4. It will follow from the proof that the constant gd in theorem 2.2 can be taken 1−e−ν(cid:17)d (cid:107)F(cid:107)∞ . (cid:16) (cid:107)F(cid:107)∞ and the upper bound obtained in corollary 2.20, which also 1−e−ν(cid:17) factor proportional to depends on the lower bound for the Lyapunov exponent of the Anderson model on (cid:96)2 (Z). Remark 5. The constant g1 in theorem 2.1 can be taken equal to be the minimum among a proportional to η η Before proceeding to the next result, recall the definition of the integrated density of states for an ergodic operator H : NH (E) = lim|Λ|→∞ TrP(−∞,E)(1ΛH1Λ) |Λ| . (2.2.12) Recall also that H is said to be ergodic (on Zd) when for every ω ∈ Ω and j ∈ Zd HTj ω = UjHωU∗ j . 23 (2.2.13) In the above equality, T : Ω → Ω denotes the measure preserving transformation given by Tjω(n) = ω(n − j) and the unitary maps Uj are defined via (cid:0)Ujϕ(cid:1) (n) = ϕ(n − j). (2.2.14) In fact, one can relax the above conditions and allow the maps Uj to be of more general form, the interested reader may consult [6, Definition 3.4]. In what follows, we denote by N0(E) the corresponding quantity for the free operator H0 defined above, which is assumed to be ergodic for the result of theorem 2.3, where we shall be concerned with the small disorder regime and aim for bounds which do not depend upon λ as λ → 0. Theorem 2.3. Assume that (A1) − (A2) hold with x2ρ(x) bounded and that g2 < λ and g(cid:48)2 < λ(cid:48). Fix a interval I where E (cid:55)→ N0(E) is α0-H¨older continuous and a bounded interval J ⊂ R. The integrated density of states Nλ,g(E) of HHub is H¨older continuous with respect to E and with respect to the pair (λ, g). More precisely: • For E, E(cid:48) ∈ I |Nλ,g(E) − Nλ,g(E(cid:48))| ≤ C(α, I, g)|E − E(cid:48)|α (2.2.15) for α ∈ [0, α0 2+α0 ] and C(α, I, g) independent of λ. • If λ, λ(cid:48) ∈ J, we have that, for any E ∈ I, α ∈ [0, |Nλ,g(E) − Nλ(cid:48),g(cid:48)(E)| ≤ C(α0.d, I) α0 2+α0 ] and β ∈ [0, (cid:16)|λ − λ(cid:48)|β + |g − g(cid:48)|β(cid:17) 2α α+3d+4], . (2.2.16) 2.3 Motivation We shall explain the motivation for the above choice of the effective potential. We are only going to outline the derivation of the self-consistent equations later in section A.2 as this is a standard topic, see, for instance, [44, Chapter 3]. 24 Let Λ ⊂ Zd be a finite set. Following the notation of [11], we use Γ to denote a one particle density matrix, i.e, a 2 × 2 matrix whose entries are operators on (cid:96)2 (Λ) and which satisfies 0 ≤ Γ ≤ 1. We then write  Γ↑ Γ↑↓ Γ↓↑ Γ↑  Γ := † where Γ↓↑ = Γ ↑↓. As in [11, Equation 3a.8], the pressure functional (cid:80)(Γ) is defined as −(cid:80)(Γ) = (cid:69)(Γ) − β−1(cid:83)(Γ). The energy functional is (cid:69)(Γ) = Tr (H0 − κ + λVω) Γ + g (cid:104)n|Γ↑|n(cid:105) (cid:104)n|Γ↓|n(cid:105) , (cid:88)  H0 − κ + λVω n 0 (2.3.1) (2.3.2) . where we have identified H0 − κ + λVω with The entropy is given by 0 H0 − κ + λVω (cid:83)(Γ) = −Tr (Γ log Γ + (1 − Γ) log(1 − Γ)) . (2.3.3) Generally, the choice of energy functional (2.3.2) is referred to as Hartree approximation as exchange terms are neglected. However, in the case of a repulsive interaction among the particles, it is easy to prove that such exchange terms do not affect the choice of minimizer for −(cid:80)(Γ) and the process may be referred to as the Hartree-Fock approximation. Indeed, the Hartree-Fock energy for the repulsive interaction would incorporate the term −g|(cid:104)n|Γ↑↓|n(cid:105)|2, which is non-negative when g < 0, inside the summation. Thus, for repusive interactions, off-diagonal terms can be disregarded for minimization purposes, see the analogue discussion in [11, Section 4a]. The minimizer Γ of −(cid:80)(Γ) exists since Λ is a finite set. Moreover, it satisfies (cid:104)n|Γ↑|n(cid:105) = (cid:104)n| 1 β(H0−κ+λVω+Diag 1 + e (cid:16) (cid:17)|n(cid:105) . Γ↓) (2.3.4) 25 (cid:17)|n(cid:105) . Thus, the effective Hamiltonian on (cid:96)2 (Λ) ⊕ (cid:96)2 (Λ) is determined by β(H0−κ+λVω+Diag (cid:104)n|Γ↓|n(cid:105) = (cid:104)n| 1 + e 1 (cid:16) Γ↑)  H0 + λω(n) + gV Λ↑ (n) 0 HΛ ω := H0 + λω(n) + gV Λ↓ (n)  0 |n(cid:105) |n(cid:105) . (2.3.5) (2.3.6) (2.3.7) V Λ↑ (ω)(n) := (cid:104)n| V Λ↓ (ω)(n) := (cid:104)n| 1 1 β(H0−κ+λω+gV↓) 1 + e β(H0−κ+λω+gV↑) 1 + e It will follow from arguments given below that if ΛR is an increasing sequence with ∪R∈NΛR = Zd then, for fixed m ∈ Zd, lim R→∞ V ΛR eff (m) = Veff (m) (2.3.8) and this fact ensures that, for localization purposes in the Hubbard model, it suffices to study HHub and its finite volume restrictions. 2.4 Outline of the Proof of theorem 2.2 We now want to outline the proof of the theorem 2.2 in the related model where HHub is replaced by the operator acting on (cid:96)2(cid:16)Zd(cid:17) with H = H0 + λω(n) + gVeff (n) Veff (n) = (cid:104)n| 1 1 + eβ(H0+λω+gVeff ) |n(cid:105) . In this case, the eigenfunction correlator is defined as QI (m, n) := sup |ϕ|≤1 |(cid:104)m|ϕ(H)|n(cid:105)|. (2.4.1) (2.4.2) (2.4.3) where the supremum is taken over functions ϕ which are Borel measurable and supported on I. 26 The above operator exhibits the main mathematical features of the Hubbard model, namely: the effective potential is defined self-consistently as a non-local and non-linear function of H. Thus, it is natural to first illustrate our methods here. For now let’s assume the existence and uniqueness of Veff are proven, as well as its regularity with respect to {ω(n)} n∈Zd. Combined with estimates on the derivatives of Veff , the above facts form a significant portion of the proof which is developed in sections 2.5 and 2.6. The, somewhat straightforward, extension of the proof to HHub will be explained in section 2.9. A feature which theorem 2.1 and theorem 2.2 have in common is that the eigenfunction correlator decay will be achieved via analysis of the Green’s function of HΛ = 1ΛH1Λ, which is H restricted to a finite set Λ ⊂ Zd. Let GΛ(m, n, z) = (cid:104)m|(HΛ − z)−1|n(cid:105) . (2.4.4) Using the basics of the fractional moment method, which dates back to [5] and [1], we aim at showing that, for some s ∈ (0, 1), E(cid:16)(cid:12)(cid:12)(cid:12)GΛ(m, n; z) (cid:12)(cid:12)(cid:12)s(cid:17) ≤ Ce−µd|m−n| (2.4.5) holds uniformly in z ∈ C+, with positive constants C = C and µ depending on the param- eters (d, s, g, λ, ν, η,(cid:107)F(cid:107)∞) but independent of the volume |Λ|. In this context, the Green’s function decay expressed by equation (2.4.5) implies −µ(cid:48) E (Q(m, n)) ≤ C(cid:48)e d|m−n| (2.4.6) for some exponent µ(cid:48) d = µ(cid:48)(d, s, g, λ, ν, η,(cid:107)F(cid:107)∞) > 0 and C(cid:48) = C(cid:48)(η, ν, d, g, λ, s,(cid:107)F(cid:107)∞). This is well known and explained in the introduction of this thesis. Another aspect which is shared by the proofs of theorems 2.1 and 2.2 is that the starting point to obtain (2.4.5) will be the following a-priori bound. Lemma 2.4. Given a finite set Λ ⊂ Zd, there exist a constant CAP depending on the parameters (η, ν, d, g, λ, s,(cid:107)F(cid:107)∞), but independent of Λ, such that (2.4.7) E(cid:16)(cid:12)(cid:12)(cid:12)GΛ(m, n; z) (cid:12)(cid:12)(cid:12)s(cid:17) ≤ CAP 27 holds for any m, n ∈ Λ. The proof of lemma 2.4 will follow from lemma 2.5 below. Let Uω(n) = ω(n) + g λ Veff (n, ω). (2.4.8) be the “full” potential at site n. From now on, to keep the notation simple, we drop the dependence on ω in the new variables {U (n)}n∈Λ. Note that U (n) and U (m) are correlated for all values of m and n. The strategy is to show that, for g sufficiently small, they still behave as if they were independent in the following sense: Lemma 2.5. Fix Λ ⊂ Zd finite and n0 ∈ Λ. The conditional distribution of U (n0) = u at specified values of {U (n)}n∈Λ\{n0} has density ρΛ . Moreover, under assumptions (A1) − (A4) we have that n0 sup u∈R If, additionally, assumption (A5) holds then ρΛ n0 sup n0∈Λ sup Λ (u) < ∞. ρΛ n0 (u) ∈ L1 (R,|x|εdx). (2.4.9) The proof of the above result is detailed in section 2.8; it requires exponential decay of ∂ω(m) | and | ∂2Veff (n) | ∂Veff (n) ∂ω(m)ω(l)| with respect to |m− n| and |m− n| +|l− n|, respectively. The need for this decay is the main reason to require β > 0 or, in other words, to require analiticity of F in a strip. The intuitive explanation for lemma 2.5 is that the random variables U (n) and U (n0) decorrelate in a strong fashion as |n − n0| becomes large. As explained in the introduction of this manuscript, lemma 2.5 implies (2.4.5) for any 0 < s < 1 as long as λ is taken sufficiently large, see also [6, Theorem 10.2]. The proof of theorem 2.1 will require additional efforts involving tools which are specific to one dimension, which we shall comment on below. 2.5 Existence of the effective potential To justify the definition of the effective potential in (2.4.1), let Φ(V ) : (cid:96)∞(Zd) → (cid:96)∞(Zd) be given by Φ(V )(n) := (cid:104)n|F (T + λVω + gV )|n(cid:105) . Recall that F is analytic, bounded in the 28 strip S = {|Imz| < η} and continuous up to the boundary of S. Our goal is to check that Φ is a contraction in (cid:96)∞(cid:16)Zd(cid:17) , meaning that (cid:107)Φ(V ) − Φ(W )(cid:107) (cid:96)∞(Zd) < c(cid:107)V − W(cid:107) (cid:96)∞(Zd) (2.5.1) holds for some c < 1 and all V, W ∈ (cid:96)∞(Zd). Let R(z, H) = (H − z)−1 and T be a self- adjoint operator. Using the analiticity of F we have the following representation for F (T ) [4, Equation (D.2)] (cid:90) ∞ −∞ (R(t + iη, T ) − R(t − iη, T )) f (t) dt 1 2πi F (T ) = (2.5.2) for all V ∈ (cid:96)∞(Zd), where f = F+ + F− − D ∗ F for F±(u) = F (u ± iη) and D(u) = η2+u2(cid:17) is the Poisson kernel. As explained in [4], the above representation is a consequence η (cid:16) π of the identities D+ + D− = δ + D ∗ D (with δ denoting the Dirac delta) and (D+ + D−) ∗ F = D ∗ (F+ + F−). (2.5.3) (2.5.4) It follows immediately that (cid:107)f(cid:107)∞ ≤ 3(cid:107)F(cid:107)∞. This is a prelude for the following fixed point argument, where the operator T will be assumed to satisfy < η 2 . (2.5.5) Proposition 2.6. The following contraction estimates hold true. (cid:17) eν|m−n| − 1 m sup n (cid:88) |T (m, n)|(cid:16) • For any self-adjoint operator T on (cid:96)2(cid:16)Zd(cid:17) (cid:12)(cid:12)(cid:12)(cid:104)m|(F (T + V ) − F (T + W ))|n(cid:105)(cid:12)(cid:12)(cid:12) ≤ 72 (cid:16) V, W , we have, for any ν(cid:48) ∈ (0, ν), that satisfying (2.5.5) and bounded potentials √ 2e−ν(cid:48)|m−n| 1 − eν(cid:48)−ν(cid:17)d η (cid:107)F(cid:107)∞(cid:107)V − W(cid:107)∞. (2.5.6) 29 • For any self-adjoint operator T on (cid:96)2(cid:16)Zd(cid:17) ⊕ (cid:96)2(cid:16)Zd(cid:17) potentials V, W on (cid:96)2(cid:16)Zd(cid:17) ⊕ (cid:96)2(cid:16)Zd(cid:17) (cid:12)(cid:12)(cid:12)(cid:104)m|(F (T + V ) − F (T + W ))|n(cid:105)(cid:12)(cid:12)(cid:12) ≤ 144 (cid:16) 1 − eν(cid:48)−ν(cid:17)d 2e−ν(cid:48)|m−n| √ η satisfying (2.5.5) and bounded we have, for any ν(cid:48) ∈ (0, ν), that (cid:107)F(cid:107)∞(cid:107)V − W(cid:107)∞ (2.5.7) • For any m, n, j ∈ Zd, the matrix elements (cid:104)m|F (T + gV )|n(cid:105) are differentiable with respect to V (j) and (cid:12)(cid:12)(cid:12) ∂ (cid:104)m|F (T + gV )|n(cid:105) ∂V (j) (cid:12)(cid:12)(cid:12) ≤ |g|72 √ 2e−ν(|m−j|+|n−j|) η (cid:107)F(cid:107)∞(cid:107)V (cid:107)∞. (2.5.8) Proof. The resolvent identity gives (cid:104)m| 1 T + V − t − iη − = (cid:104)m|( 1 T + V − t − iη 1 T + W − t − iη − 1 |n(cid:105) + (cid:104)m| 1 T + W − t + iη − )(W − V ) 1 T + W − t − iη |n(cid:105) 1 T + V − t + iη |n(cid:105) (cid:18) − (cid:104)m| 1 T + W − t + iη (W − V ) 1 T + V − t + iη |n(cid:105) . (cid:19) T + V − t + iη − T + W − t − iη 1 Taking absolute values in the first term on the right-hand side we obtain (cid:19) |n(cid:105)(cid:12)(cid:12)(cid:12) (cid:18) (cid:12)(cid:12)(cid:12)(cid:104)m| ≤ (cid:88) (cid:88) l∈Zd ≤ 24 l × (cid:104)m| 1 − 1 T + V − t − iη |GV (m, l; t + iη) − GV (m, l; t − iη)||(W − V )(l)|GW (l, n; t + iη)| T + W − t − iη T + V − t + iη (W − V ) 1 |(V − W )(l)|e−ν(|l−n|+|l−m|) (cid:104)l| 1 (T + V − E)2 + η2/2 |l(cid:105)1/2 1 (T + V − t)2 + η2/2 |m(cid:105)1/2 . 30 In the last step we made use of the Combes-Thomas bound |GW (m, n; t + iη)| ≤ 2 η e−ν|m−n| as well as lemma 3 in [4, appendix D] to estimate the difference between the Green functions as |GV (m, l; t + iη) − GV (m, l; t − iη)| ≤ 12ηe−ν|m−l| (cid:104)m| 1 (T + V − t)2 + η2/2 |m(cid:105)1/2 (cid:104)l| 1 (T + V − E)2 + η2/2 |l(cid:105)1/2 . Integrating over t we conclude, using Cauchy-Schwarz and the spectral measure representa- tion, that (cid:82) ∞ −∞ (cid:12)(cid:12)(cid:12)(cid:104)m|(cid:16) (cid:17) T +W−E−iη|n(cid:105)(cid:12)(cid:12)(cid:12) dt 1 (cid:80) T +V −t−iη − 1 √ ≤ 24 η 2π (W − V ) T +V −t+iη l |(V − W )(l)|e−ν(|l−n|+|l−m|). 1 The above implies that (cid:90) ∞ 1 −∞ √ 2π ≤ 12 2 η √ 12 η = 2 1 − T + V − t − iη (cid:18) (cid:12)(cid:12)(cid:12)(cid:104)m| (cid:107)V − W(cid:107)∞e−ν(cid:48)|m−n| (cid:88) (cid:16) (cid:107)V − W(cid:107)∞e−ν(cid:48)|m−n| l∈Zd 1 T + V − t + iη e(ν(cid:48)−ν)|l−n| 1 − eν(cid:48)−ν(cid:17)d 1 . (cid:19) (W − V ) 1 T + W − t − iη (2.5.9) (2.5.10) |n(cid:105)(cid:12)(cid:12)(cid:12) dt As a similar bound holds for the term (cid:90) ∞ −∞ (cid:18) (cid:12)(cid:12)(cid:12)(cid:104)m| 1 2π 1 H0 + W − t + iη − 1 H0 + W − t − iη (cid:19) (V − W ) 1 H0 + V − t + iη |n(cid:105)(cid:12)(cid:12)(cid:12) dt we conclude the proof of the inequality (2.5.6) by recalling that (cid:107)f(cid:107)∞ ≤ 3(cid:107)F(cid:107)∞. The inequality (2.5.7) follows from the same argument with the only difference that one has to sum two geometric series, hence the modification of the upper bound by a factor of two. The e−ν|j−n|e−ν|m−j| is an upper bound bound (2.5.8) is proven similarly: note that |g||h| 24 2π √ η for the left-hand side of equation (2.5.9) with V replaced by gV and W = g(V + hPj), where 31 Pj denotes the projection onto Span{δj}. We also observe that this time there will be no summation over l, hence the introduction of the ν(cid:48) is unnecessary. We then conclude that (cid:12)(cid:12)(cid:12) (cid:104)m|F (T + gV + hPj)|n(cid:105) − (cid:104)m|F (T + gV )|n(cid:105) h (cid:12)(cid:12)(cid:12) ≤ |g|72 η √ 2 e−ν|j−n|e−ν|m−j|. (2.5.11) Letting h → 0 finishes the proof. Taking m = n we conclude that the contraction mapping condition (2.5.1) holds whenever η(cid:0)1 − e−ν(cid:1)d √ 2(cid:107)F(cid:107)∞ . 72 |g| < (2.5.12) This observation yields the following. Proposition 2.7. Let gd = potential Veff ∈ (cid:96)∞(cid:16)Zd(cid:17) (cid:16) 1−e−ν(cid:17)d 2(cid:107)F(cid:107)∞ . Then, for |g| < gd, there is a unique effective √ η 72 satisfying Veff (n) = (cid:104)n|F (H0 + λω + gVeff|n(cid:105) . (2.5.13) Moreover, for Λ ⊂ Zd, there is a unique V Λ HΛ = 1ΛH1Λ. Remark 6. Replacing H0 by H0− κI we can incorporate a chemical potential in our results. For simplicity, we shall make no further reference to κ during the proofs and assume it was eff in (cid:96)2 (Λ) satisfying (2.5.13) with H replaced by already incorporated to H0. 2.6 Regularity of the effective potential Our goal in this section is to conclude that, for a fixed finite subset Λ ⊂ Zd with |Λ| = n, the effective potential Veff is a smooth function of {ω(j)}j∈Λ. This will be of relevance for several resampling arguments later in the thesis. For that purpose, define a map ξ : (cid:96)∞(cid:16)Zd(cid:17) × Rn → (cid:96)∞(cid:16)Zd(cid:17) by ξ(V, ω)(j) = V (j) − (cid:104)j|F (H0 + λω + gV )|j(cid:105) 32 (2.6.1) (2.6.2) Then, Veff is the unique solution of ξ(V, ω) = 0. Thus, its regularity can inferred via the implicit function theorem once we check that the derivative Dξ(·, ω) is non-singular. Note that ∂ξ(V, ω)(j) ∂V (l) = δjl − ∂ (cid:104)j|F (H0 + λω + gV )|j(cid:105) ∂V (l) . (2.6.3) Using proposition 2.6, we have that (cid:12)(cid:12)(cid:12)∂ (cid:104)j|F (H0 + λω + gV )|j(cid:105) (cid:12)(cid:12)(cid:12) ≤ |g|72 √ 2e−2ν|j−l| (cid:107)F(cid:107)∞. ∂V (l) √ η(1−e−2ν )d < 1 we have that the operator Dξ(ω, .) : (cid:96)∞ (Λ) → In particular, whenever |g| 72 (cid:96)∞ (Λ) is invertible since it has the form I + gM where gM has operator norm less than one. It is worth observing that the smallness condition on g is independent of Λ ⊂ Zd. It is a 2(cid:107)F(cid:107)∞ η (2.6.4) consequence of the implicit function theorem that V is a smooth function of (ω(1), ..., ω(n)). 2.7 Decay estimates for the effective potential We start this section with the following lemma, which will be useful in order to formulate the decay of correlations between U (n) and U (m) as |m − n| → ∞. √ 2|g|(cid:107)F(cid:107)∞ η(1−e−ν)d < 1, there exist constants C1(d, λ, g, η,(cid:107)F(cid:107)∞, ν) and Lemma 2.8. Whenever 72 C2(d, λ, g, η,(cid:107)F(cid:107)∞, ν) such that eν|n−m|(cid:12)(cid:12)(cid:12)∂Veff (n) eν(|l−n|+|n−m|+|l−m|)(cid:12)(cid:12)(cid:12) ∂2Veff (n) eν|n−m|(cid:12)(cid:12)(cid:12) ∂Veff (n) (cid:12)(cid:12)(cid:12)(cid:111) ≤ C1 (cid:12)(cid:12)(cid:12) ≤ C2. (cid:110)(cid:88) (cid:88) (cid:88) (cid:12)(cid:12)(cid:12), ∂ω(m) ∂ω(m) (2.7.2) (2.7.1) max m n ∂ω(m)∂ω(l) l,m,n Moreover C1 and C2 can be bounded from above by a constant of the form λD 1−gθ with D and θ independent of g and these constants are explicit in the proof. Proof. For convenience of notation we let Veff = V in this proof. As in section 2.5 we write (cid:90) ∞ (cid:18) −∞ F (H) = 1 2πi (cid:19) f (t) dt 1 H − t − iη − 1 H − t + iη 33 where f is bounded by 3(cid:107)F(cid:107)∞. Thus V (n, ω) = (cid:90) ∞ −∞ K(n, t, ω)f (t) dt 1 2πi (2.7.3) (cid:96)2(cid:16)Zd(cid:17) where K(n, t, ω) = G(n, n; t − iη) − G(n, n; t + iη). Denote by Pm the projection mapping onto (cid:96)2 (Span{δm}). Using difference quotients, it is easy to check ∂ ∂ω(m) 1 H − z + g 1 H − z ∂V ∂ω(m) 1 H − z = −λ 1 H − z Pm 1 H − z . (2.7.4) Taking matrix elements we obtain ∂K(n, t, ω) ∂ω(m) = −g (cid:88) l ˜G(l, n) ∂V (l) ∂ω(m) + λr(m, n). ˜G(l, n) := G(l, n; t + iη)G(n, l; t + iη) − G(l, n; t − iη)G(n, l; t − iη). r(m, n) := G(n, m; t + iη)G(m, n; t + iη) − G(n, m; t − iη)G(m, n; t − iη). Note ˜G(l, n) = (G(l, n; t + iη) − G(l, n; t − iη)) G(n, l; t + iη) + (G(n, l; t + iη) − G(n, l; t − iη)) G(l, n; t − iη). (2.7.5) We now make use of [4, Lemma 3]: |G(l, n; t + iη) − G(l, n; t − iη)| ≤ 12ηe−ν|l−n| (cid:104)n| |n(cid:105)1/2 (2.7.6) 1 (H−t)2+η2/2 |l(cid:105)1/2 . 1 (H−t)2+η2/2 ×(cid:104)l| This, together with the Combes-Thomas bound |G(l, n, t ± iη)| ≤ 2 η e−ν|l−n| and (2.7.5) implies | ˜G(l, n)| ≤ 48e−2ν|l−n| (cid:104)n| 1 (H − t)2 + η2/2 |n(cid:105)1/2 (cid:104)l| 1 (H − t)2 + η2/2 |l(cid:105)1/2 . |r(m, n)| ≤ 48e−2ν|m−n| (cid:104)m| 1 (H − t)2 + η2/2 |m(cid:105)1/2 (cid:104)n| 1 (H − t)2 + η2/2 |n(cid:105)1/2 . 34 To summarize, we have shown the following inequality −∞ ˜r(m, n) := ˜K(l, n) := (cid:90) ∞ (cid:90) ∞ (cid:32) (cid:12)(cid:12)(cid:12) ∂V (n) (cid:12)(cid:12)(cid:12) ≤ 3(cid:107)F(cid:107)∞ |g|(cid:88) 1−e−2ν(cid:17)d < 1 we have that ∂ω(m) −∞ 2π l √ 2|g| 72 √ | ˜G(l, n)| dt ≤ 48 2π η √ |r(m, n)| dt ≤ 48 2π η e−2ν|l−n| e−2ν|m−n|. (cid:12)(cid:12)(cid:12) ∂V (l) ∂ω(m) (cid:33) (cid:12)(cid:12)(cid:12) + λ˜r(m, n) . ˜K(l, n) |g|(cid:107) ˜K(cid:107)∞,∞ < 1 Thus Whenever (cid:16) η where (cid:107) ˜K(cid:107)∞,∞ = sup Considering the weight W (n) := eν|m−n| we let l θ := 3(cid:107)F(cid:107)∞ 2π sup n By the triangle inequality, (cid:88) m ˜K(l, m). (cid:88) l W (n) W (l) ˜K(n, l). eν|n−l| ˜K(n, l) (cid:88) θ ≤ 3(cid:107)F(cid:107)∞ sup 2π n √ 2(cid:107)F(cid:107)∞ ≤ 72 η (1 − e−ν)d l . hence, whenever √ 72 2|g| η(1−e−ν)d < 1, we have that |g|θ < 1. 35 (2.7.7) (2.7.8) (2.7.9) (2.7.10) Moreover, with the choice we have (cid:88) n D1 := W (n)˜r(m, n) √ D1 ≤ 72 2(cid:107)F(cid:107)∞ η (1 − e−ν)d . (2.7.11) (2.7.12) After conditions (2.7.7), (2.7.10) and (2.7.12) have been verified, the general result [6, Theorem 9.2] applies, yielding eν|n−m|(cid:12)(cid:12)(cid:12) ∂V (n) ∂ω(m) (cid:88) m Differentiating (2.7.4) with respect to ω(l), (cid:12)(cid:12)(cid:12) < (cid:18) ∂ λD1 1 − gθ := C1(d,(cid:107)F(cid:107)∞, λ, g, η, ν). (2.7.13) ∂2 ∂ω(m)∂ω(l) H − z 1 (cid:18) ∂ ∂ω(m) 1 + g = −λ 1 H − z ∂V + g (cid:18) ∂ (cid:19) ∂ω(l) ∂ω(l) H − z Pm (cid:19) ∂V ∂ω(m) 1 H − z 1 (cid:19) + g H − z 1 H − z − λ ∂ω(l) 1 H − z 1 H − z 1 H − z ∂2V (cid:18) ∂ 1 (cid:19) H − z 1 ∂ω(m)∂ω(l) Pm ∂ω(l) H − z Repeating the previous argument and using the established decay of ∂V (n) ∂ω(m) we reach (2.7.2), finishing the proof. Given a finite set Λ ⊂ Zd, let us define (cid:84) : R|Λ| → R|Λ| by ((cid:84)ω) (n) = ω(n) + g λ Veff (n). (2.7.14) Let U (n) := ((cid:84)ω) (n) be the new coordinates in the probability space. The bound (2.7.1) implies that, for |g| sufficiently small, (cid:84) is a differentiable perturbation of the identity by an operator with norm less than one hence (cid:84)−1 is well defined. Fix n0 ∈ Λ and denote by Uα = U + (α − U (n0)) δn0 the new potential obtained from U by changing its value at n0 to α ∈ R. Let ωα(n) =(cid:0)(cid:84)−1Uα (cid:1) (n). The variables ωα(n) correspond to the change in ω(n) when a resampling argument is applied to the new proba- bility space at the point n0. Intuitively, the exponential decay shown above guarantees that this change is not too large if n and n0 are far away. This is the content of the result below. 36 Lemma 2.9. Given α ∈ R and |g| < λC−1 (cid:88) n(cid:54)=n0 eν|n−n0|(cid:12)(cid:12)ωα(n) − ω(n)(cid:12)(cid:12) ≤ C1|g| 1 , we have λ (cid:16)|α − U (n0)| + 2 (cid:16) 1 − |g| λ C1 (cid:17) . |g|(cid:107)F(cid:107)∞ λ (cid:17) where C1 is the upper bound in equation (2.7.1). Proof. Using the given definitions and the mean value inequality we obtain, for n (cid:54)= n0, λ |V (n, ω) − V (n, ωα)| |ω(n) − ωα(n)| ≤ |g| (cid:12)(cid:12)(cid:12) ∂Veff (n, ˆωα) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ωα(l) − ω(l) (cid:12)(cid:12)(cid:12) (cid:88) ≤ |g| (cid:12)(cid:12)(cid:12)(cid:18) (cid:12)(cid:12)(cid:12)∂Veff (n, ˆωα) ≤ |g| (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ωα(l) − ω(l) (cid:12)(cid:12)(cid:12) ∂Veff (n, ˆωα) (cid:12)(cid:12)(cid:12). (cid:88) |g| λ |α − U (n0)| + 2 ∂ω(n0) l∈Zd ∂ω(l) ∂ω(l) λ λ l(cid:54)=n0 (cid:19) + |g|(cid:107)F(cid:107)∞ λ Where ˆωα denotes some configuration in the probability space with ˆωα(l) in the interval connecting ω(l) to ωα(l). Let W (n) = eν|n−n0|. According to (2.7.1), (cid:88) l sup n W (n) W (l) (cid:12)(cid:12)(cid:12) ∂Veff (n, ˆωα) ∂ω(l) (cid:12)(cid:12)(cid:12) ≤ sup n (cid:88) l eν|n−l|(cid:12)(cid:12)(cid:12)∂Veff (n, ˆωα) ∂ω(l) (cid:12)(cid:12)(cid:12) ≤ C1. Once again, the conditions of [6, Theorem 9.2] are satisfied for |g| < λC−1 |g|(cid:107)F(cid:107)∞ eν|n−n0|(cid:12)(cid:12)ωα(n) − ω(n)(cid:12)(cid:12) ≤ C1|g| (cid:16)|α − U (n0)| + 2 (cid:88) λ 1 (cid:17) . (cid:16) (cid:17) λ therefore 1 − |g| λ C1 n(cid:54)=n0 Since another application of the mean value theorem gives, after a possible correction on ˆωα that (cid:12)(cid:12)(cid:12) ∂Veff (n, ω) ∂ω(m) − ∂Veff (n, ωα) ∂ω(m) (cid:12)(cid:12)(cid:12) ≤ (cid:88) l∈Zd (cid:12)(cid:12)(cid:12) ∂2Veff (n, ˆωα) ∂ω(l)∂ω(m) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ω(l) − ωα(l) (cid:12)(cid:12)(cid:12) 37 we obtain that, for any ν(cid:48) ∈ (0, ν), the difference above by C2|g| λ C1 (cid:16)|α − U (n0)| + 2 (cid:16) (cid:17) (cid:17) |g|(cid:107)F(cid:107)∞ (1 − eν(cid:48)−ν)d 1 − |g| λ λ C1 ∂ω(m) (cid:12)(cid:12)(cid:12) is bounded from ∂ω(m) − ∂Veff (n,ωα) (cid:12)(cid:12)(cid:12) ∂Veff (n,ω)  e−ν(cid:48)(|m−n|+|n−n0|+|m−n0|). + 2(cid:107)F(cid:107)∞ where C2 is the constant in (2.7.2). In particular, letting ν(cid:48) = ν/2 and with the choices we have(cid:88) (m,n)∈Λ×Λ |(A − B)m,n| ≤ C2|g|2 A = g λ (cid:18) ∂Veff (ni, ωα) ∂ω(nj) (cid:19) C1 (cid:19) (cid:18) ∂Veff (ni,ωα) λ2 Lemma 2.10. Let A = g λ |g| λ C1 < 1 we have (cid:88) (m,n)∈Λ×Λ We summarize the above observation as a lemma. ∂ω(nj) (cid:19) (cid:18) ∂Veff (ni, ω) (cid:17) |g|(cid:107)F(cid:107)∞ (1 − e−ν/2)3d λ , B = g λ |Λ|×|Λ| (cid:16)|α − U (n0)| + 2 (cid:16) (cid:17) 1 − |g| λ C1 |Λ|×|Λ|  . 2(cid:107)F(cid:107)∞ (1 − e−ν/2)2d + (2.7.15) (cid:18) ∂Veff (ni,ω) ∂ω(nj ) (cid:19) . Whenever |Λ|×|Λ| ∂ω(nj ) |Λ|×|Λ| and B = g λ |(A − B)m,n| ≤ |g|2 (C3|α − U (n0)| + C4) . (2.7.16) Moreover, the constant C3 can be chosen independent of λ and C4 is proportional to 1 λ. Finally, we analyze how the effective potential varies with respect to disorder and inter- action. This will be relevant to the Integrated Density of States regularity with respect to disorder and interaction strengths developed in chapter four. Lemma 2.11. For a fixed ω ∈ Ω |Vλ,g(n) − Vλ(cid:48),g(cid:48)(n)| ≤ C5(d,(cid:107)F(cid:107)∞, g, η, ν, ω) 1 − gC6(d,(cid:107)F(cid:107)∞, g, η, ν) |λ − λ(cid:48)| + C7(d,(cid:107)F(cid:107)∞, g, η, ν)|g − g(cid:48)|. (2.7.17) 38 Note when λ (cid:54)= λ(cid:48) the bound depends on ω through the constant C5. Proof. Let Rλ,g(z) = H0+λω+gVλ,g−z and Rλ(cid:48),g(cid:48)(z) = 1 1 H0+λ(cid:48)ω+g(cid:48)Vλ(cid:48),g(cid:48)−z for z = t + iη. Similarly as in the above proofs, it is immediate to check that Rλ,g(z) − Rλ(cid:48),g(cid:48)(z) =(λ(cid:48) − λ)Rλ,g(z)VωRλ(cid:48),g(cid:48)(z) + (g(cid:48) − g)Rλ,g(z)Vλ(cid:48),g(cid:48)Rλ(cid:48),g(cid:48)(z) (cid:16) (cid:17) + gRλ,g(z) Vλ,g − Vλ(cid:48),g(cid:48) Rλ(cid:48),g(cid:48)(z). Replacing z by ¯z and subtracting the resulting equations: (cid:17) (cid:0)Rλ,g(z) − Rλ,g(¯z)(cid:1) −(cid:16) (cid:0)Rλ,g(z) − Rλ,g(¯z)(cid:1)(cid:16) Rλ(cid:48),g(cid:48)(z) − Rλ(cid:48),g(cid:48)(¯z) (cid:16) (cid:17)(cid:16) (λ(cid:48) − λ)Vω + (g(cid:48) − g)Vλ(cid:48),g(cid:48) (cid:17) + g(cid:0)Rλ,g(z) − Rλ,g(¯z)(cid:1)(cid:16) (λ(cid:48) − λ)Vω + (g(cid:48) − g)Vλ(cid:48),g(cid:48) Rλ(cid:48),g(cid:48)(z) − Rλ(cid:48),g(cid:48)(¯z) (cid:16) (cid:17)(cid:16) (cid:17) Vλ,g − Vλ(cid:48),g(cid:48) Rλ(cid:48),g(cid:48)(z) Rλ(cid:48),g(cid:48)(z) − Rλ(cid:48),g(cid:48)(¯z) Vλ,g − Vλ(cid:48),g(cid:48) Rλ(cid:48),g(cid:48)(z) + gRλ,g(z) + Rλ,g(¯z) (cid:17) = . (cid:17) Taking matrix elements, multiplying by f (t), integrating with respect to t and taking absolute values we can read from the representation (2.7.3) that, denoting Kλ,g(n, l) = |Gλ,g(n, l; z) − Gλ,g(n, l; ¯z)|, |ω(l)| (cid:90) ∞ (cid:16) (cid:90) ∞ (cid:16)|Gλ,g(n, l)|Kλ(cid:48),g(cid:48)(l, n) + Gλ(cid:48),g(cid:48)(n, l)|Kλ,g(l, n) Kλ,g(n, l)Kλ(cid:48),g(cid:48)(l, n) + |Gλ,g(n, l)| ˜Kλ(cid:48),g(cid:48)(l, n) (cid:17) −∞ dt. −∞ (2.7.18) (cid:17) dt |Vλ,g(n) − Vλ(cid:48),g(cid:48)(n)| ≤ 3(cid:107)F(cid:107)∞ |λ − λ(cid:48)|(cid:88) |g − g(cid:48)|(cid:88) 2π 3(cid:107)F(cid:107)2∞ (cid:90) ∞ (cid:88) l∈Zd l∈Zd 2π + + 2g −∞ l∈Zd |Gλ,g(n, l)||Vλ,g(l) − Vλ(cid:48),g(cid:48)(l)|Kλ(cid:48),g(cid:48)(l, n) dt. Using equation (2.7.6) together with [6, Theorem 9.2] we conclude the proof. 39 2.7.1 Improvements We will now improve upon the previous bounds. More specifically, we need robust estimates which also reflect the decay of the derivatives of Veff (n) when the local potential ω(n) is large. The content of this section will be important for a general fluctuation analysis on section 2.8 and for complete localization in the one dimensional setting. Before stating the main result of the section we start with the following deterministic estimate, which incorporates ideas from [4, Lemma 3]. Lemma 2.12. |G(m, l; t + iη)| ≤ √ 2(cid:104)l| 1 (H − t)2 + η2|l(cid:105)1/2 e−ν|m−l| (2.7.19) Proof. To keep the notation simple, we set t = 0 without loss of generality. Let Hf = ef He−f where f (n) = ν min{|n − l|, R} for a fixed l ∈ Zd and R > 0. By choosing R sufficiently large, we may assume that |m − l| < R. We then have eν|m−l|G(m, l; iη) = (cid:104)m|(Hf − iη)−1|l(cid:105) . ||(Hf − iη)−1(H2 + η2)1/2|| ≤ √ 2. (2.7.20) We claim that Indeed, ||(Hf − iη)−1(H2 + η2 2 )1/2||2 = ||(H2 + = ||(H2 + η2 2 η2 2 )1/2(H∗ )1/2 f + iη)−1(Hf − iη)−1(H2 + η2 2 (H2 + η2 2 )1/2|| 1 f + iη) )1/2|| where by [4, Eq D.9] (with f replaced by −f ) we have (Hf − iη)(H∗ (cid:19) (cid:18) H2 + η2 2 (Hf − iη)(H∗ f + iη) ≥ 1 2 showing the claim in (2.7.20). The inequality (2.7.19) will now follow from |(cid:104)m|(Hf − iη)−1|l(cid:105)| ≤ (cid:107)(Hf − iη)−1(H2 + )1/2(cid:107)|(H2 + η2 2 )−1/2δl| η2 2 √ ≤ 2(cid:104)l|(H2 + )−1|l(cid:105)1/2 . η2 2 40 Lemma 2.13. There exists C7(λ, η, d, g,(cid:107)F(cid:107)∞, ν) > 0 such that, for m (cid:54)= n, max{|ω(n)|,|ω(m)|}(cid:12)(cid:12)(cid:12) ∂V (n) ∂ω(m) (cid:12)(cid:12)(cid:12) ≤ C7e−2ν|m−n| and, for n (cid:54)= n0, (cid:12)(cid:12)ω(n)(ωα(n) − ω(n))(cid:12)(cid:12) ≤ C7|g| λ − |g|C1 (cid:32) |α − U (n0)| + 2 |g|(cid:107)F(cid:107)∞ λ + 1 (1 − e−ν)d (2.7.21) (cid:33) e−ν|n−n0|. (2.7.22) Moreover, whenever |g| λ C1 < 1, C7 can be chosen to be uniformly bounded as a function of the parameters λ and g. Proof. Recall that U (n) = ω(n) + g λ Veff (n) denotes the “full” potential at site n. We divide the proof into two cases. • Case one: U (n) ≥ 0. Let us start by noting that lemma 2.7.19 implies that for n, l ∈ Zd √ |G(n, l; t + iη)G(l, n; t + iη)| dt ≤ 2 From the previous section we already know that ∂V (n) ∂ω(m) = r(n, l) ∂V (l) ∂ω(m) + λr(m, n) f (t) dt (2.7.23) where f (t) = F+(t + iη) + F−(t − iη) − D ∗ F (t) and r(m, n) = G(n, m; t + iη)G(m, n; t + iη) − G(n, m; t − iη)G(m, n; t − iη). Observe that, for z = t + iη and n (cid:54)= m, λ|U (n)G(n, m; z)G(m, n; z)| (cid:88) l λ|U (n)| |λU (n) − z| |z| (cid:18) = ≤ 1 + |λU (n) − z| (cid:19)(cid:88) l 41 |H0(n, l)G(l, m; z)G(m, n, z)| |H0(n, l)G(l, m; z)G(m, n, z)| (cid:90) ∞ −∞ (cid:32) (cid:90) ∞ −∞ −g (cid:88) l 2π η e−2ν|n−l|. (cid:33) where we made use of the identity (λU (n) − z)G(n, m; z) = δmn −(cid:88) Note that if U (n) ≥ 0 and t = Rez < 0, then l |z| |λU (n) − z| ≤ 1. (cid:90) ∞ Using the fact that tf (t) goes to zero as t → ∞ we conclude that |U (n)G(n, m; t+iη)G(m, n; t+iη)||f (t)| dt ≤ C(ν, η,(cid:107)F(cid:107)∞) −∞ H0(n, l)G(l, m; z). (2.7.24) (2.7.25) e−2ν|m−n|. (2.7.26) λ Since a similar equation holds with m replaced by l, we can proceed as in the previous section and, using the exponential decay of ∂V (n) ∂ω(m), conclude the proof. • Case two: U (n) < 0. In this case, the argument given above must be modified to take into account that the inequality |z| |λU (n) − z| ≤ 1. (2.7.27) is satisfied when t = Re > 0. In this case, the immediate use of (2.7.23) would result in a problem as tf (t) is unbounded as t → −∞. This can be addressed by observing that the Fermi-Dirac function F (z) has the following symmetry 1 1 + eβ(z−µ) = 1 − 1 1 + eβ(−z+µ) . Hence we can make use of the representation (2.7.23) corresponding to − 1 1 + eβ(−z+µ) := F∗ µ(z) (2.7.28) (2.7.29) since, for m (cid:54)= n, the constant term does not affect the calculation of ∂V (n) ∂ω(m). Denoting by f∗(t) = F∗ +(t + iη) + F∗−(t − iη) − D ∗ F∗(t) 42 we reach (cid:32) (cid:90) ∞ −∞ −g (cid:88) l ∂V (n) ∂ω(m) = r(n, l) ∂V (l) ∂ω(m) (cid:33) + r(m, n) f∗(t) dt (2.7.30) where now tf∗(t) → 0 as t → −∞. Proceeding as in the first case the proof is finished, showing (2.7.21). Following the proof of lemma 2.9 and using (2.7.21) we conclude (2.7.22) Lemma 2.14. Let Λ1 and Λ2 be subsets of Zd with dist(Λ1, Λ2) ≥ r. Let V c potential defined by 2 be the effective 2 (n) = (cid:104)n|F (HΛc V c 2)|n(cid:105) for n ∈ Zd where HΛc 2 denotes the restriction of H to the complement of Λ2. Then, for any n ∈ Λ1 (cid:12)(cid:12)(cid:12) ∂V (n) ∂ω(m) (cid:12)(cid:12)(cid:12) ≤ C(η, d, λ, g,(cid:107)F(cid:107)∞, ν)e−ν(|m−n|+r) − ∂V c 2 (n) ∂ω(m) (2.7.31) Proof. The proof follows the same steps as in the previous results. The only modification which is required comes when comparing the quantities r(m, n) and rΛc 2(m, n) given by r(m, n) = G(n, m; t + iη)G(m, n; t + iη) − G(n, m; t − iη)G(m, n; t − iη) rΛc 2(m, n) = GΛc 2(n, m; t + iη)GΛc 2(m, n; t + iη) − GΛc 2(n, m; t − iη)GΛc 2(m, n; t − iη). We observe that G(n, m; z)G(m, n; z) − GΛc G(m, n; z) − GΛc G(n, m; z) (cid:16) 2(n, m; z)GΛc 2(m, n; z) = 2(m, n; z) + G(n, m; z) − GΛc 2(n, m; z) GΛc 2(m, n; z). (cid:17) Moreover, G(m, n; z) − GΛc 2(m, n; z) = −λ G(m, l; z)ω(l)GΛc 2(l, n; z) l∈Λ2 G(m, l; z) (Veff (l) − V c + g λ 2 (l)) GΛc 2(l, n; z). (cid:17) (cid:16) (cid:88) (cid:88) l 43 Since Veff (l)− V c is now finished using arguments identical to the proof of lemma 2.8 and the improvement on 2 (l) can be expressed in terms of Green’s functions of H and HΛc 2 the proof lemma 2.13. 2.8 Proof of lemma 2.5 We now verify the existence of the density ρΛ n0 . Fix Λ ⊂ Zd finite. Recall that we have defined U (n, ω) = ω(n) + g λ Veff (n, ω). (2.8.1) Until the end of this section we suppress the ω dependence on U (n) and Veff . Note that, for m, n ∈ Λ, We have denoted the above change of variables by (cid:84) : R|Λ| → R|Λ|, which reads ∂U (m) ∂ω(n) = δmn + g λ ∂Veff (m) ∂ω(n) . (2.8.2) (cid:84)(ω(n1), ..., ω(n|Λ|)) = (U (n1), ..., U (n|Λ|)). (2.8.3) It is now simple compute the joint distribution of the variables {U (n)} n∈Zd. Using the n∈Zd have a common density ρ we conclude that for −1U (nk) ρ (cid:84) dU (n1)...dU (n|Λ|) ∂Veff (ni, (cid:84)−1U ) ∂U (nj) ρ U (nk) − g λ Veff (nk, (cid:84) (cid:17) −1U ) dU (nk). (cid:17) = fact that the random variables {ω(n)} all Borel sets I1, ..., IN in R: (cid:17) |Λ|(cid:89) P(cid:16) U (n1) ∈ I1, ... , U (n|Λ|) ∈ I|Λ| (cid:90) (cid:84)−1(cid:16) (cid:90) (cid:90) I1×...×I|Λ| I1×...×I|Λ| |Λ|(cid:89) k=1 | det J (cid:12)(cid:12) det (cid:84)−1| (cid:16) I + = = k=1 (cid:16) g λ I1×...×I|Λ| ρ(ω(nk)) dω(nk) (cid:16) (cid:17) (cid:17)(cid:12)(cid:12) |Λ|(cid:89) k=1 44 Therefore the joint distribution of {U (nk)}|Λ| k=1 is given by the measure (cid:16) (cid:12)(cid:12)(cid:12) det ∂Veff (ni, (cid:84)−1U ) ∂U (nj) g λ I + (cid:16) (cid:17)(cid:12)(cid:12)(cid:12) |Λ|(cid:89) k=1 (cid:17) −1U ) ρ U (nk) − g λ Veff (nk, (cid:84) dU (n1)...dU (n|Λ|). (2.8.4) It follows that for each n0 ∈ Λ the conditional distribution of U (n0) at specified values of {U (n)}n(cid:54)=n0 (cid:82) ∞ (cid:17)(cid:12)(cid:12)(cid:81)|Λ| k=1 ρ(cid:0)U (nk) − g (cid:17)(cid:12)(cid:12)(cid:81)|Λ| k=1 ρ(cid:0)U α(nk) − g λ Veff (nk, (cid:84)−1U )(cid:1) λ Veff (nk, (cid:84)−1U α)(cid:1) dα ∂Veff (ni,(cid:84)−1U ) ∂Veff (ni,(cid:84)−1U α) (cid:16) (cid:12)(cid:12) det (cid:16) (cid:12)(cid:12) det has a density given by I + g λ I + g λ ∂U (nj ) ρΛ n0 −∞ = ∂U (nj ) Where U α(n) := U (n) + (α − U (n0)) δn=n0. This strategy naturally leads to the analysis of ratios of determinants. A sufficient condition for finding an upper bound to the right-hand (2.8.5) side of (2.8.5) is to obtain positive constants C = Cfluct(U (n0)) and D = D(α) which are independent of |Λ| and such that the following estimates hold true (cid:90) ∞ −∞ D(α)ρ where dπα := (cid:16) (cid:12)(cid:12) det (cid:16) (cid:12)(cid:12) det (cid:16) α − g λ (cid:17)(cid:12)(cid:12) (cid:17)(cid:12)(cid:12) ≥ D(α). ∂U (nj ) ∂U (nj ) I + g λ I + g λ Veff (n0, (cid:84) dπα ≥ Cfluct(U (n0)) ∂Veff (ni,(cid:84)−1U α) ∂Veff (ni,(cid:84)−1U ) (cid:17) −1U α) λ Veff (n, (cid:84)−1U α)(cid:1) ρ(cid:0)U α(n) − g (cid:89) ρ(cid:0)U (n) − g λ Veff (n, (cid:84)−1U )(cid:1) dα. λ Veff (n0, (cid:84)−1U )(cid:1) (u) ≤ ρ(cid:0)u − g . n∈|Λ|\{n0} ρΛ n0 Cfluct(u) (2.8.6) (2.8.7) (2.8.8) (2.8.9) The bounds (2.8.7) and (2.8.6) readily imply that letting U (n0) = u Lemma 2.5 will follow from a precise control of the right-hand side of equation (2.8.9). We now execute the strategy which was outlined above. The ratio of determinants can be controlled through the following bound, where (cid:107)M(cid:107)1 denotes the trace norm of a matrix M . 45 Lemma 2.15. Let A, B be square matrices with I + B invertible. Then, (cid:12)(cid:12)(cid:12)(cid:12) det (I + A) det (I + B) (cid:12)(cid:12)(cid:12)(cid:12) ≤ e(cid:107)(A−B)(I+B)−1(cid:107)1. Proof. We make use of the elementary identities det(I + A) det(I + B) = det(I + A)(I + B)−1 and (I + A)(1 + B)−1 = I + (A − B)(I + B)−1. The proof is now finished with the inequality |det(1 + M )| ≤ e(cid:107)M(cid:107)1 which holds in the general setting of trace class operators, see [49, Lemma 3.3] The triangle inequality for the trace norm implies the following. (2.8.10) (2.8.11) (2.8.12) Corollary 2.16. Under the above conditions (cid:12)(cid:12)(cid:12)det(I + B) (cid:18) ∂V (ni,ω) (cid:19) det(I + A) (cid:12)(cid:12)(cid:12) ≥ e −(cid:80) (A−B)(I+B)−1(cid:17) m,n |(cid:16) (cid:18) ∂V (ni,ωα) (cid:19) | (2.8.13) mn Letting A = g λ ∂U (nj ) see that, for |g| < λC−1 1 , (1 + B)−1 has uniformly bounded operator norm. Using lemma 2.10 and corollary 2.16 we conclude that (2.8.6) holds with D(α) = e−|g|2C3(|α−U (n0)|+C4). ∂U (nj ) |Λ|×|Λ| |Λ|×|Λ| and using lemma 2.8 we , B = g λ We now check that equation (2.8.7) holds when ρ satisfies the fluctuation bound (2.2.4). We divide the proof in two cases: • Suppose that c2(ρ) > 0. Let cρ = max{c1(ρ), c2(ρ)}. The left-hand side of (2.8.7) is bounded from below by (cid:90) ∞ −∞ D(α)ρ e−cρ|ω(n)−ωα(n)|(1+|ω(n)|+|ωα(n)|) dα (cid:17) (cid:89) α − g λ −1U α) Veff (n0, (cid:84) (cid:16) n∈Λ\{n0} 46 (cid:16) α − g λ Veff (n0, (cid:84) (cid:17) −1U α) eS(n) dα which equals for S(n) = (cid:90) ∞ −∞ D(α)ρ (cid:88) n∈Λ\{n0} −cρ|ω(n) − ωα(n)|(1 + |ω(n)| + |ωα(n)| ). (2.8.14) Due to the triangle inequality and lemmas 2.9 and 2.13, we conclude that there is a positive constant C = C(d,(cid:107)F(cid:107)∞, g, η, ν) with limg→0 C(d,(cid:107)F(cid:107)∞, g, η, ν) < ∞ such that for n (cid:54)= n0 − cρ|ω(n) − ωα(n)| (1 + |ω(n)| + |ωα(n)|) ≥ − cρ|ω(n) − ωα(n)| (1 + 2|ω(n)| + |ωα(n) − ω(n)|) C2|α − U (n0)|2 + 2C|α − U (n0)|(cid:17) ≥ −|g|cρe−ν|n−n0|(cid:16) . Therefore, − cρ − |g| (cid:88) n∈Λ\{n0} 2cρ (1 − e−ν)d |ω(n) − ωα(n)| (1 + |ω(n)| + |ωα(n)|) ≥ C2|α − U (n0)|2 + 2C|α − U (n0)|(cid:17) (cid:16) ). Thus, by choosing |g| sufficiently small so that |g| (1−e−ν)d C2 < cρ and using the assumption (A4) we obtain that the integral below is finite and bounded from below 4cρ by a positive constant independent of Λ and n0. (cid:90) ∞ −∞ D(α) ρ(cid:0)α − g ρ(cid:0)U (n0) − g λ Veff (n0, (cid:84)−1U α)(cid:1) λ Veff (n0, (cid:84)−1U )(cid:1) eS(n) dα. where S(n) was defined in (2.8.14). This, together with (2.8.9), verifies lemma 2.5 (u) ∈ when c2(ρ) > 0. If ρ satisfies the assumption (A5) the above argument yields ρΛ n0 L1 (R,|u|εdu). 47 • Assume that c2(ρ) = 0: Similarly to the above argument, the left-hand side of (2.8.7) is bounded from below by Where, from (2.2.4) (cid:90) ∞ −∞ D(α) −|g| λ Veff (n0, (cid:84)−1U α)(cid:1) ρ(cid:0)α − g ρ(cid:0)U (n0) − g λ Veff (n0, (cid:84)−1U )(cid:1) e λ Veff (n0, (cid:84)−1U α)(cid:1) ρ(cid:0)α − g ρ(cid:0)U (n0) − g λ Veff (n0, (cid:84)−1U )(cid:1) ≥ e λ Veff (n0, (cid:84)−1U α)(cid:1) ρ(cid:0)α − g λ Veff (n0, (cid:84)−1U )(cid:1) (cid:89) ρ(cid:0)U (n0) − g (cid:18) −c1(ρ) (cid:90) ∞ −∞ D(α) 0 < Again, choosing |g| sufficiently small we conclude that n∈Λ\{n0} c1(ρ) (1−e−ν)d C|α−U (n0)| . (2.8.15) (cid:19) . |g| λ (2.8.16) |α−U (n0)|+2 e−c1(ρ)|ω(n)−ωα(n)| dα < ∞. finishing the proof. 2.9 The Hartree approximation for the Hubbard model Let us now explain how the techniques from the previous sections can adapted to the  0 0 H0 + λVω + gV↓(ω) 0 0 H↓(ω)  :=  H↑(ω) Hubbard model. Recall that HHub is defined as  H0 + λVω + gV↑(ω) acting on (cid:96)2(cid:16)Zd(cid:17) ⊕ (cid:96)2(cid:16)Zd(cid:17) is the standard Anderson model acting on (cid:96)2(cid:16)Zd(cid:17)  V↑(ω)(n)  =  (cid:104)n|F (H↓)|n(cid:105)  . (cid:104)n|F (H↑)|n(cid:105) V↓(ω)(n) . The operators H0 and Vω are defined as before, i.e; H0 + λVω . The effective potentials are given by (2.9.1) Mathematically, the treatment of the above model is very similar to the one explained above, therefore most details are skipped and we just indicate the required modifications. 48 2.9.1 Existence of the Effective potential Let Φ(X, Y ) : (cid:96)∞(cid:16)Zd(cid:17) ⊕ (cid:96)∞(cid:16)Zd(cid:17) → (cid:96)∞(cid:16)Zd(cid:17) ⊕ (cid:96)∞(cid:16)Zd(cid:17) be given by Φ(X, Y )(m, n) := ((cid:104)n|F (H0 + Vω + gY )|n(cid:105) , (cid:104)m|F (H0 + Vω + gX)|m(cid:105)) . using proposition 2.6, we immediately reach |g| √ 72 (cid:96)∞(cid:16)Zd(cid:17)⊕(cid:96)∞(cid:16)Zd(cid:17) ≤ (cid:96)∞(cid:16)Zd(cid:17)(cid:33) (cid:32) (cid:107)Φ(X1, Y1) − Φ(X2, Y2)(cid:107) (cid:16) 1 − eν(cid:48)−ν(cid:17)d (cid:107)X1 − X2(cid:107) (cid:16) 1−eν(cid:48)−ν(cid:17)d(cid:107)F(cid:107)∞ < 1 we conclude Φ has a unique fixed point (cid:96)∞(cid:16)Zd(cid:17) + (cid:107)Y1 − Y2(cid:107) (cid:107)F(cid:107)∞ 2 √ 72 2 η η Veff =(cid:0)V↑, V↓(cid:1) . (2.9.2) Therefore, if |g| belonging to (cid:96)∞(cid:16)Zd(cid:17) ⊕ (cid:96)∞(cid:16)Zd(cid:17) . 2.9.2 Regularity of the effective potential Fix Λ ⊂ Zd finite and define functions ξ : ((cid:96)∞ (Λ) ⊕ (cid:96)∞ (Λ)) × Rn → (cid:96)∞ (Λ) ⊕ (cid:96)∞ (Λ) through ξ↑(V, ω)(j) = V ↑(j) − (cid:104)j|F (H0 + λω + gV↓)|j(cid:105) . ξ↓(V, ω)(j) = V ↓(j) − (cid:104)j|F (H0 + λω + gV↑)|j(cid:105) . (2.9.3) (2.9.4) Our goal is to conclude V ↑,V ↓ are smooth functions of an arbitrary, but finite, list {ω(j)}j∈Λ. Again, this can be done via implicit function theorem once we check that the derivative ∂ξ(ω, V )(j) ∂V (l) = δjl − ∂ (cid:104)j|F (H0 + λω + gV )|j(cid:105) ∂V (l) . is non-singular. Using lemma 2.6, we have that for (cid:93) ∈ {↑,↓} (cid:12)(cid:12)(cid:12) ∂ (cid:104)j|F (H0 + λω + gV(cid:93))|j(cid:105) ∂V (l) (cid:12)(cid:12)(cid:12) ≤ |g|72 √ 2e−2ν|j−l| η (cid:107)F(cid:107)∞. (2.9.5) (2.9.6) 49 In particular, whenever |g| 144 √ η(1−e−2ν )d < 1 we have that the operator Dξ(·, ω) : (cid:96)∞ (Λ)⊕ (cid:96)∞ (Λ) → (cid:96)∞ (Λ) ⊕ (cid:96)∞ (Λ) has an inverse. From the implicit function theorem it follows that V is a smooth function of (ω(1), ..., ω(n)) for n = |Λ|. 2(cid:107)F(cid:107)∞ 2.9.3 Decay estimates The decay rate in the case of the Hubbard model is dictated by ∂ω(m) (cid:12)(cid:12)(cid:12) ∂V↑(n) (cid:12)(cid:12)(cid:12) ∂V↓(n) ∂ω(m) (cid:12)(cid:12)(cid:12) ≤ 3|g|(cid:107)F(cid:107)∞(cid:88) (cid:12)(cid:12)(cid:12) ≤ 3|g|(cid:107)F(cid:107)∞(cid:88) l l ∂ω(m) (cid:12)(cid:12)(cid:12) ∂V↓(l) (cid:12)(cid:12)(cid:12) ∂V↑(l) ∂ω(m) (cid:12)(cid:12)(cid:12) + ˜r↓(n). (cid:12)(cid:12)(cid:12) + ˜r↑(n). ˜K↓(l, m) ˜K↑(l, m) where, for (cid:93) ∈ {↑,↓} (2.9.7) (2.9.8) ˜G(cid:93)(l, m) := G(cid:93)(l, n; t + iη)G(cid:93)(n, l; t + iη) − G(cid:93)(l, n; t − iη)G(cid:93)(n, l; t − iη). r(cid:93)(m, n) := G(cid:93)(n, m; t + iη)G(cid:93)(m, n; t + iη) − G(cid:93)(n, m; t − iη)G(cid:93)(m, n; t − iη). ˜K(cid:93)(l, m) := ˜r(cid:93)(n) := −∞ | ˜G(cid:93)(l, m)| dt. (cid:90) ∞ (cid:90) ∞ (cid:16) ˜K↑(l, m) + ˜K↓(l, m) |r(cid:93)(n)| dt. −∞ In particular, (cid:12)(cid:12)(cid:12)∂V↑(n) ∂ω(m) (cid:12)(cid:12)(cid:12) + (cid:12)(cid:12)(cid:12)∂V↓(n) ∂ω(m) (cid:12)(cid:12)(cid:12) ≤ 3|g|(cid:107)F(cid:107)∞(cid:88) +(cid:0) ˜r↑(n, m) + ˜r↓(n, m)(cid:1) . l (cid:17)(cid:18)(cid:12)(cid:12)(cid:12) ∂V↑(l) ∂ω(m) (cid:12)(cid:12)(cid:12) + ∂ω(m) (cid:12)(cid:12)(cid:12)(cid:19) (cid:12)(cid:12)(cid:12) ∂V↓(l) M11 . The M21 The analysis from the previous sections applies and we obtain lemmas 2.8,2.9,2.10 and 2.13 with |.| being replaced by the matrix norm |M| = |M11| +|M21| for M = effective potential and its derivatives are to be interpreted as follows: 50 Recall in this case H0 = −∆ hence, we define HHub acting on(cid:0)(cid:96)2 (Z) ⊕ (cid:96)2 (Z)(cid:1) by (cid:76)(z) = −E(cid:0)ln|G+(0, 0; z)|(cid:1) And(0, 0; z)|(cid:1) . (cid:76)And(z) = −E(cid:0)ln|G+   H↑(ω)  HAnd + gV↑(ω)  := HHub = H↓(ω) 0 0 0  . 0 (2.10.1) (2.10.2) (2.10.3) (2.10.4) V ↑ eff (n) ↓ eff (n) V  , Veff (n) = ∂Veff (n) ∂ω(m) =  ∂V ↑ eff ↓ eff (n) ∂ω(m) ∂V (n) ∂ω(m)  and ∂2Veff (n) ∂ω(m)∂ω(l) =  ∂2V ↑ eff ↓ eff (n) ∂ω(m)ω(l) ∂2V (n) ∂ω(m)ω(l)  . 2.10 One dimensional Aspects:proof of theorem 2.1 In this section we will prove theorem 2.1. We let H+ = H[0,∞)∩Z and denote by G+(m, n; z) the Green’s function of H+. Recall the definition of the Lyapunov exponent: where, denoting by HAnd the standard Anderson model −∆ + Vω on (cid:96)2 (Z),  H↑(ω) 0 H↓(ω) 0 HAnd + gV↓(ω) The effective potentials are defined as (2.2.8). In the theorem below, we will use an abbre- viation and (cid:76)(z) will refer to the Lyapunov exponent of either H↑ or H↓ whereas (cid:76)And(z) will denote the Lyapunov exponent of the Anderson model on (cid:96)2 (Z). 2.11 One dimensional aspects: strategy of the proof of theorem 2.1 The argument for proving theorem 2.1 follows closely the approach in the proof of theorem 12.11 in [6], which we now recall. 51 2.11.1 Main ideas in the i.i.d case In the reference [6, Chapter 12] the decay rate for the Green’s function is described in terms of the moment generating function, defined by ϕ(s, z) = lim|n|→∞ ln E(|G(0, n; z)|s) |n| . (2.11.1) The existence of the above quantity for all z ∈ C+ and s ∈ (0, 1) and its relationship to the Lyapunov exponent are a consequence of Fekete’s lemma: Lemma 2.17 (Fekete). Let {an}n∈N be a sequence of real numbers such that, for every pair (m, n) of natural numbers, an+m ≤ an + am. (2.11.2) Then, α = limn→∞ an n exists and equals infn∈N an n . It is an elementary observation that if, instead, the sequence {an}n∈N satisfies the in- equality an+m ≤ an + am + C then the above result applies to bn := an + C and that an ana- logue statement holds for superadditive sequences, i.e, sequences which satisfy (2.11.2) but with the reversed inequality sign. In the i.i.d. context, the sequence an = ln E (|G(0, n; z)|s) is shown to be both subbaditive and superadditive, meaning that there exist constants C−(s, z) and C+(s, z) for which an + am + C− ≤ an+m ≤ an + am + C+. (2.11.3) holds for all m, n ∈ N, see [6, Lemma 12.10]. A consequence of this fact, together with a precise control of the arising constants, is stated below where we operate under the assump- tion that the random variables {ω(n)}n∈Z have a density ρ which satisfies ρ ∈ L1 (R;|u|εdu) with ε ∈ (0, 1). Lemma 2.18. [6, Theorem 12.8] For any z ∈ C+, there are cs(z), Cs(z) ∈ (0,∞) such that for all n ∈ Z c−1 s (z)eϕ(s,z)|n| ≤ E (|GAnd(0, n; z)|s) ≤ Cs(z)eϕ(s,z)|n|. (2.11.4) 52 Moreover, for any compact set K ⊂ R and S ⊂ [−ε, 1), we have the local uniform bound sup s∈S sup z∈K+i(0,1] max{cs(z), Cs(z)} < ∞ (2.11.5) and the same result holds with z replaced by its boundary value E + i0 for Lebesgue almost every E. On the other hand, for fixed z ∈ C+, ϕ(s, z) is shown to be a convex function of s and ∂s = −(cid:76)(z). It is a non-increasing in [−ε, +∞), with its derivative at s = 0 satisfying ∂ϕ(0,z) consequence of these facts that for almost every E ∈ R there exists a value s = s(E) ∈ (0, 1) such that ϕ(s, E) ≤ − s 2 (cid:76)(E). (2.11.6) The above is the content of [6, Equation (12.86)]. Dynamical localization is shown to hold locally as a consequence of the inequality (2.11.3) along with lemma 2.18, the inequality (2.11.6) and Kotani theory, which establishes that (cid:76)(E) is positive for almost every E ∈ R. 2.11.2 Modifications In this section we will outline the proof theorem 2.1 with HHub again replaced by the operator defined in (2.4.1). For simplicity we set λ = 1 since the disorder strenght does H on (cid:96)2(cid:16)Zd(cid:17) (cid:96)2(cid:0)Z+(cid:1) and denote by G+(m, n; z) the Green’s function of H+. Recall the definition of the not play an important role in theorem 2.1. Let H+ = H[0,∞)∩Z be the restriction of H to Lyapunov exponent: initially, for z ∈ C+, we let (cid:76)(z) = −E(cid:0)ln|G+(0, 0; z)|(cid:1) . (2.11.7) By Herglotz theory (see, for instance, [6, Appendix B] and references therein) it is seen that, for Lebesgue almost every E ∈ R, (cid:76)(E) is well defined as limδ→0+ (cid:76)(E + iδ). Finally, recall the uniform positivity of the Lyapunov exponent for the Anderson model on (cid:96)2 (Z): ess inf E∈R (cid:76)And(E) > (cid:76)And (2.11.8) 53 for some (cid:76)And > 0. The first step towards Green’s function decay (2.4.5) will be showing uniform positivity of (cid:76)(E), which is accomplished by the following. Theorem 2.19. There exists a constant CLyap(s, η, g,(cid:107)F(cid:107)∞) > 0 such that |(cid:76)(z) − (cid:76)And(z)| ≤ CLyap|g|s (2.11.9) for all z ∈ C+. Proof. From the resolvent identity we obtain |G+(0, 0; z)| |G+ And(0, 0; z)| ≤ 1 + |g|(cid:107)F(cid:107)∞(cid:88) |G+(0, 0; z)| ≤ 1 + |g|(cid:107)F(cid:107)∞(cid:88) (cid:32) |G+(0, 0; z)| (cid:88) And(0, 0; z)| (cid:33) |G+ n n ln |G+ And(0, 0; z)| s n Using the bound ln(1 + x) ≤ xs And(n, 0; z)| |G+(0, n; z)||G+ |G+ And(0, 0; z)| And(0, n; z)||G+(n, 0; z)| |G+ |G+(0, 0; z)| (2.11.10) (2.11.11) s for 0 < s < 1 and x > 0 we reach, for 0 < s < 1/2, ≤ |g|s . (cid:107)F(cid:107)s∞ (2.11.12) And(n, 0; z)|s |G+(0, n; z)|s|G+ And(0, 0; z)|s |G+ (cid:18)|G+ (cid:19)1/2 (n,0;z)|2s (0,0;z)|2s And |G+ And := CLyap(s, η, ν,(cid:107)F(cid:107)∞)|g|s. (2.11.13) Taking expectations, using the definition of the Lyapunov exponents and the Cauchy-Schwarz inequality (cid:76)And(z) − (cid:76)(z) ≤ |g|s s (cid:107)F(cid:107)s∞ supn E(cid:0)|G+(0, n; z)|2s(cid:1)1/2(cid:80) E n The fact that CLyap is a finite quantity follows from a couple of remarks. Firstly, by Feen- berg’s expansion [6, Theorem 6.2] we have the identity |G+ And(n, 0; z)| = |G+ And(0, 0; z)||G+ And(1, n; z)| (2.11.14) where G+ And(1, n; z) denotes the Green’s function of HAnd restricted to (cid:96)2 (Z)∩ [1,∞). From the a-priori fractional moment bound on lemma 2.4 combined with the Green’s function decay for dimensional Anderson model E(cid:16)|G+ And(1, n; z)|2s(cid:17) < C(s)e−µAnd|n| (2.11.15) we conclude that that CLyap < ∞. The estimate for (cid:76)(z) − (cid:76)And(z) is similar. 54 In principle one might worry that the pre-factor CLyap on the above bound will depend on g. However, it is easy to see from the arguments in the proof of lemma 2.4, that CAP converges to a finite quantity as g → 0, thus we shall disregard its dependence on g. (cid:19)1/s (cid:18) (cid:76)And CLyap Corollary 2.20. Whenever |g| < holds for some s ∈ (0, 1/2), we have (cid:76)0 := ess inf E∈R (cid:76)(E) > 0. (2.11.16) We can now proceed to the second step of the proof of theorem 2.1, which consists of establishing Green’s function decay from corollary 2.20. For that purpose, an important detail to keep in mind is that, in the correlated context, if we choose an = log E (|G(0, n; z)|s), the condition (2.11.3) will not be fulfilled for all pairs (m, n) due to the lack of independence between the potentials. This means that Fekete’s lemma is not applicable. Moreover, its well-studied modifications (for instance by P. Erd¨os and N. G. de Bruijn [24]) do not seem to suffice either. To the best of our knowledge the result given below is new. Its formulation takes into account the strong decorrelation between the potentials in the Hubbard model and introduces a notion of approximate subbaditivity. Lemma 2.21 (Fekete-type lemma for approximately subbaditive sequences). Let δ > 0 be given and {an}n∈N be a sequence of real numbers such that, for every triplet m, n, r of natural numbers with r ≥ δ max{log m, log n}, the inequality an+m+r ≤ an + am + C holds with a constant C independent of m, n and r. Then, α = lim n→∞ an n exists and equals infn∈N an+C n . Moreover, α ∈ [−∞, 0]. Note that, as a consequence, we have an ≥ nα − C 55 (2.11.17) (2.11.18) (2.11.19) for all n ∈ N, where C is the same constant as in (2.11.17). The following decoupling estimate guarantees the applicability of the above lemma with the choice (cid:16)| ˆG(0, n; z)|s(cid:17) an = log E [0,n] where ˆG(0, n; z) = (cid:104)0|(H[0,n] − z)−1|n(cid:105) is the Green’s function of the operator H restricted to (cid:96)2 ([0, n] ∩ Z) and E [0,n] denotes the expectation with respect to U (0), ..., U (n). Lemma 2.22. [Strong mixing decoupling] There exist constants CDec(s, ν, η, g,(cid:107)F(cid:107)∞) and δ = δ(η, ν, g,(cid:107)F(cid:107)∞) such that the inequality (cid:16)| ˆG(0, n + m + r; z)|s(cid:17) ≤ CDecE (cid:16)| ˆG(0, n; z)|s(cid:17) E (cid:16)| ˆG(0, m; z)|s(cid:17) [0,n] [0,m] E [0,n+m+r] holds whenever r ≥ δ log max{m, n}. (2.11.20) A combination of lemmas 2.22, 2.21 and equation (2.11.19) yields the lower bound C−1 Deceϕ(s,z)n ≤ E [0,n] for all n ∈ N. (2.11.21) (cid:16)| ˆG(0, n; z)|s(cid:17) As we shall see in section 2.11.5 below, an application of the lower bound (2.11.21) in combination with the superadditive version of lemma 2.21 applied to the sequence bn = −ϕ(s, E)n + log E is enough to establish an upper bound (cid:16)| ˆG(0, n; z)|s(cid:17) E(cid:16)| ˆG(0, n; z)|s(cid:17) ≤ C(s, z)eϕ(s,z)n for all n ∈ N. [0,n] (2.11.22) where the constant C(s, z) is locally uniform in (s, z) ∈ (0, 1) × C+. After obtaining an analogue of lemma 2.18, the final step will be to relate the moment- generating function to the Lyapunov exponent through an inequality of the type ϕ(s, E) ≤ − s 2 (cid:76)0. (2.11.23) In reference [6], the bound (2.11.23) is stated with (cid:76)0 replaced by (cid:76)(E) and with s depending on E. However, it is easy to see from the from the arguments given there that s can be 56 chosen locally uniformly in E, see [6, Equations (12.79) and (12.80)]. Moreover, by making use uniform positivity of the Lyapunov exponent obtained in corollary 2.20 we reach the inequality (2.11.23). The Green’s function decay follows from the bounds (2.11.22) and (2.11.23). 2.11.3 Proof of Lemma 2.21 For simplicity we set C = 0. The general statement will follow by considering the related sequence bn := an + C. Given integers L and (cid:96) with L >> (cid:96), our goal is to bound above in terms of a(cid:96) (cid:96) . As a initial step, observe that by (2.11.17) we have aL L from aL ≤ aL−δ log L−(cid:96) + a(cid:96). Iterating the above procedure k + 1 times for k = k(cid:96),L := (cid:98) L − 2(cid:96) − δ log L δ log L + (cid:96) (cid:99) (2.11.24) we obtain aL ≤ (k + 2) a(cid:96) In the above iteration we have made use of the fact that in the assumption (2.11.17) the remainder r can be adjusted as long as it satisfies the inequality given there. Thus, aL L ≤ (cid:96)(k + 2) L a(cid:96) (cid:96) (2.11.25) Before proceeding with the proof, a few remarks are in order. Firstly, nothing is achieved by holding (cid:96) fixed and letting L → ∞ directly on equation (2.11.25) since this only yields the upper bound of zero. A second attempt would be showing that letting (cid:96) → ∞ (hence, L → ∞ as well) implies that the ratio k(cid:96) L converges to one. However, as q(cid:96),L − (cid:96) L ≤ k(cid:96) L ≤ q(cid:96),L for the choice q(cid:96),L = 1 − 2 (cid:96) L − δ log L L 1 + δ log L (cid:96) 57 (2.11.26) (2.11.27) we see that k(cid:96) L converges to one as (cid:96) → ∞ only along a subsequence where → 0 and (cid:96) L log L (cid:96) → 0. (2.11.28) Taking this into account, let ε > 0 be given and (cid:96)1 be an initial scale to be determined. Let L >> (cid:96)1 be a positive integer to be interpreted as a larger scale. Iterating equation (2.11.25) throughout a sequence of scales (cid:96)1 < (cid:96)2 < ... < (cid:96)NL satisfying, for some p > 0, p log(cid:0)(cid:96)j and ≤ L < (cid:96)NL+1 < ... (cid:1) . (cid:1) ≤ log (cid:96)j+1 < p2 log(cid:0)(cid:96)j ∞(cid:88) (cid:33) NL−1(cid:89) (cid:96)j (cid:96)j+1 < ∞ (cid:18) q(cid:96)j ,(cid:96)j+1 + 2(cid:96)j (cid:96)j+1 j=1 (2.11.29) (2.11.30) (2.11.31) (2.11.32) (cid:19) a(cid:96)1 (cid:96)1 . j=1 we reach, for q(cid:96),L defined in (2.11.27), (cid:32) ≤ aL L q(cid:96)NL ,L + 2(cid:96)NL L Since q(cid:96)j ,(cid:96)j+1 chosen (independently of L) so that → 1 as j → ∞, due to (2.11.31), we conclude that the value of (cid:96)1 can be NL−1(cid:88) (cid:33) (cid:32) (cid:19) (cid:18) log q(cid:96)j ,(cid:96)j+1 + 2(cid:96)j (cid:96)j+1 + log q(cid:96)NL ,L + 2(cid:96)NL L < ε. (2.11.33) j=1 Thus aL L ≤ eε a(cid:96)1 (cid:96)1 . (2.11.34) Moreover, the above conclusion holds for any integer (cid:96)1 sufficiently large, as long as L >> (cid:96)1. In particular, we can also require that Combining equations (2.11.34) and (2.11.35) the proof is finished letting ε → 0. a(cid:96)1 (cid:96)1 ≤ inf n an n + ε. (2.11.35) 58 2.11.4 Proof of lemma 2.22 E [0,n+m+r] We will show that the following inequality holds (cid:16)| ˆG(0, n + m + r; z)|s(cid:17) ≤ CDecE where CDec = CAPeC(η,g,(cid:107)F(cid:107)∞)e−ν(cid:48)r(m2+n2), E [0,n] (cid:16)| ˆG(0, n; z)|s(cid:17) E [0,m] (cid:16)| ˆG(0, m; z)|s(cid:17) (2.11.36) [0,n] denotes the expectation with respect to the variables U (0), ..., U (n) and CAP is, up to a multiplication by a constant independent of m, n and r, the constant obtained on the a priori from lemma 2.4. Denote ˆω = (ω(1), ..., ω(n + 1), ω(n + r), ..., ω(n + r + m)) (2.11.37) and let us change variables according to ˆω (cid:55)→ ˆU . (2.11.38) We remark that the variables ω(n + 2), ..., ω(n + r − 1) are fixed in this process. Note that by lemma 2.4 and a geometric resolvent expansion we have E(cid:16) ˆG(0, n + m + r; z)|s(cid:17) ≤ CAPE(cid:54)=n+1,n+r (cid:16)| ˆG(0, n; z)|s| ˆG(n + r + 1, n + r + m; z)|s(cid:17) . (2.11.39) where E(cid:54)=n+1,n+r indicates that the variables U (n + 1) and U (n + r) were integrated out. Observe that the corresponding Jacobian has the structure  (cid:65)n×n (cid:74) = (cid:66)n×(m+r) (cid:67)(m+r)×n (cid:68)(m+r)×(m+r)  where Moreover, (cid:65)jk = δjk + g ∂Veff (j) ∂U (k) , (cid:67)jk = g ∂Veff (n + j) ∂U (k) . , (cid:66)jk = g ∂Veff (j) ∂U (n + k)  (cid:73)r×r (cid:81)r×m (cid:79)m×r (cid:80)m×m  (cid:68) = 59 where (cid:73)r×r is the identity matrix and (cid:80)jk = δjk + g ∂Veff (n + r + j) ∂U (n + r + k) , (cid:81)jk = g ∂Veff (n + j) ∂U (n + r + k) , (cid:79)jk = g ∂Veff (n + r + j) ∂U (n + k) . By the Schur complement formula det (cid:74) = det (cid:65) det (cid:80) det (cid:73)m+r×m+r − (cid:68) −1(cid:67)(cid:65) −1(cid:66) (cid:16) (cid:17) (cid:16) det (cid:73)m×m − (cid:80) −1(cid:79)(cid:81) (cid:17) . (2.11.40) where, according to the estimate (2.7.1), the matrices (cid:66) and (cid:81) have entries which decay exponentially away from their lower-left corner. Likewise, the entries of (cid:67), (cid:79) decay expo- nentially away from their upper-right corner. It readily follows from lemma 2.15 that, (cid:16) (cid:73)m+r×m+r − (cid:68) −1(cid:67)(cid:65) −1(cid:66) det (cid:73)m×m − (cid:80) −1(cid:79)(cid:81) det (cid:17) (cid:16) (cid:17) ≤ C(η, ν, g,(cid:107)F(cid:107)∞). therefore, for C as above, det (cid:74) ≤ C det (cid:65) det (cid:80). Let ρ(l) = ρ (U (l) − gVeff (l)). We obtain a decoupling estimate by observing that, setting U (l) = 0 for l ≥ n + r + 1 would only alter Veff (j) and ∂Veff (j) ∂U (k) by at most a factor which decays as C(η, g,(cid:107)F(cid:107)∞)e−ν(r+|j−k|) for 1 ≤ j, k ≤ n. This follows from the exponential decay on lemmas 2.8, 2.13 and 2.14. Similarly, we can set U (l) = 0 for l ≤ n and this only changes by at most a factor which decays as Ce−ν(r+|j−k|) for n + r ≤ j, k ≤ m + n + r. The above process yields two independent measures Veff (j) and ∂Veff (j) ∂U (k) (cid:18) (cid:19) (cid:89) [0,n] 0≤l≤n dπ0 [0,n] := det0 I + g ∂Veff (j) ∂U (k) ρ0(l) dU (l) (2.11.41) and dπ0 [n+r+1,n+m+r] := det0 (cid:18) I + g ∂Veff (j) ∂U (k) (cid:19) (cid:89) [n+r+1,n+m+r] n+r+1≤n+r+m 60 ρ0(l) dU (l). (2.11.42) Using lemma 2.15 we then arrive at an inequality of the type (cid:16)| ˆG(0, n + m + r; z)|s(cid:17) ≤ CAPeC(η,g,(cid:107)F(cid:107)∞)e−νr(m2+n2) E [0,m+n+r] (cid:90) | ˆG(0, n; z)|s dπ0 [0,n] (cid:90) × |G+(n + r, n + r + m; z)|s dπ0 [n+r+1,n+m+r]. Rewriting the above conclusion in terms of expectations we obtain (cid:16)| ˆG(0, n + m + r; z)|s(cid:17) ≤ CDecE [0,n] (cid:16)| ˆG(0, n; z)|s(cid:17)| E [0,m+n+r] × E [n+r+1,n+m+r] (G+(n + r, n + r + m; z)|s) . where CDec = CAPeC(η,g,(cid:107)F(cid:107)∞)e−νr(m2+n2). At the expense of increasing CAP the above expectations can be replaced by expectations over the full probability space, which yields the desired conclusion by translation invariance. 2.11.5 Proof of (2.11.22) We shall modify the proof of lemma 2.22 to obtain a “super-additive” estimate of the form (cid:16)| ˆG(0, n + m + r; z)|s(cid:17) ≥ C(s, E)eϕ(s,z)re−C(η,g,(cid:107)F(cid:107)∞)e−ν(cid:48)r(m2+n2) (cid:16)| ˆG(0, m; z)|s(cid:17) (cid:16)| ˆG(0, n; z)|s(cid:17) E ×E [0,n] [0,m] E [0,n+m+r] where the constant C(s, z) can be chosen locally uniform in z and s ∈ (0, 1). Since the argu- ment is very similar to the one in the proof of (2.11.36), we only explain the key modification which consists in obtaining a lower bound for E as follows. [n+1,n+r] (cid:16)| ˆG(n + 1, n + r; z)|s(cid:17) We start by writing | ˆG(n + 1, n + r; z)|s = | ˆG(n + 1, n + 1; z)|s| ˆG(n + 2, n + r; z)|s. (2.11.43) Using Jensen’s inequality we have that, for any ε ∈ (0, s), (cid:16)| ˆG(n + 1, n + 1; z)|s(cid:17) ≥ En+1 (cid:16)| ˆG(n + 1, n + 1; z)|−ε(cid:17)−s ε . En+1 (2.11.44) 61 where En+1 denotes the conditional expectation with respect to U (n + 1). Making use of the discrete Riccati equation [6, Proposition 12.1] we obtain Equations (2.11.43), (2.11.44) and (2.11.45) together with lemma 2.5 yield a lower bound . (2.11.45) (2.11.46) (2.11.47) (cid:16)| ˆG(n + 1, n + 1; z)|−ε(cid:17) En+1 = En+1 (cid:16)| ˆG(n + 1, n + r; z)|s(cid:17) ≥ C(z, s)E E [n+1,n+r] (cid:16)|U (n + 1) − z − ˆG(n + 2, n + 2; z)|ε(cid:17) (cid:16)| ˆG(n + 2, n + r; z)|s(cid:17) [n+2,n+r] which in combination with (2.11.21) implies that, after a suitable adjustment of the constant C(s, z), E [n+1,n+r] (cid:16)| ˆG(n + 1, n + r; z)|s(cid:17) ≥ C(z, s)eϕ(s,z)r. Equation (2.11.43) follows from the above inequality combined with a decoupling esti- mate analogous to the one in the proof of (2.11.36). Again, choosing r comparable to max{log m, log n} we obtain (cid:16)| ˆG(0, n + m + r; z)|s(cid:17) ≥ C(s, z)eϕ(s,z)r E [0,n+m+r] (cid:16)| ˆG(0, n; z)|s(cid:17) (cid:16)| ˆG(0, m; z)|s(cid:17) E [0,n] ×E [0,m] multiplying both sides of the above inequality by e−ϕ(s,z)(m+n) and taking logarithms we conclude that the sequence bn = log e−ϕ(s,z)n(cid:16)E (cid:16)| ˆG(0, n; z)|s(cid:17)(cid:17) [0,n] satisfies bn+m+r ≥ log(C(s, z)) + bn + bm. (2.11.48) The bound (2.11.22) now follows from an application of the supperaditive version of lemma 2.21. 2.12 H¨older Continuity for the integrated density of states at weak interaction In this section we shall address the problem of H¨older continuity for the integrated density of states for the Hubbard model with respect to energy, disorder and interaction. Our 62 results follow from modifications of the methods in [29] and references therein after we have established the existence of a suitable conditional density as in lemma 2.5. Let’s now prove theorem 2.3, starting from H¨older continuity with respect to energy, equation (2.2.15). We proceed as in [29, Section 2]. For simplicity, we replace HHub by H defined in (2.4.1). The arguments given below will be applicable to H↑ and H↓ and, therefore, suffice to show the same result for HHub. Fix an energy interval I of length ε > 0 centered at E ∈ R. The idea is to use the H¨older continuity of N0 and the resolvent identity to reach the following inequality for ε << 1 and |I| = ε, where we denote by PΛ(I) the spectral projection of HΛ on the interval I. (1 − o(ε))E (TrPΛ(I)) ≤ C(I, ρ)εα|Λ|. (2.12.1) Dividing both sides of (2.12.1) by |Λ| and letting |Λ| → ∞ gives (2.2.15). To obtain (2.12.1) (cid:16) we fix an interval J containing I with |J| to be determined. We then write, with P0,Λ(J) = P (cid:17) (J), HΛ 0 Tr(PΛ(I)) = Tr(PΛ(I)P0,Λ(J)) + Tr(PΛ(I)P0,Λ(J c)). (2.12.2) Note (2.12.3) The above inequality combined with to the H¨older continuity of N0 with respect to E ∈ R Tr(PΛ(I)P0,Λ(J)) ≤ Tr(P0,Λ(J)). |N0(E) − N0(E(cid:48))| ≤ C(I, d)|E − E(cid:48)|α0. yields, for |Λ| sufficiently large depending only on J, Tr(PΛ(I)P0,Λ(J)) ≤ C(J, d)|J|α0|Λ|. (2.12.4) (2.12.5) We now estimate the second term on the left-hand side of equation (2.12.2). By the resolvent identity, Tr(cid:0)PΛ(I)P0,Λ(J c)(cid:1) = Tr(cid:0)PΛ(I)(H − E)P0,Λ(J c)(H0,Λ − E)−1(cid:1) PΛ(I)U ΛP0,Λ(J c)(H0,Λ − E)−1(cid:17) −λTr (cid:16) . (2.12.6) 63 Where we have written U = Vω + g λ Veff . Moreover, using using functional calculus and that E is the center of I, we estimate the first term on the left-hand side of equation (2.12.6) by Now, the second term in equation (2.12.6) can be controlled by means of (cid:16) Tr |J| − |I|Tr(PΛ(I)). |I| (PΛ(I))(HΛ − E)P0,Λ(J c)(H0,Λ − E)−1(cid:17) ≤ (cid:16) PΛ(I)U ΛP0,Λ(J c)(H0,Λ − E)−1(cid:17) (cid:16) (HΛ − E)(PΛ(I))U ΛP0,Λ(J c)(H0,Λ − E)−2(cid:17) U Λ(PΛ(I))U ΛP0,Λ(J c)(H0,Λ − E)−2(cid:17) (cid:16) A = −λTr B = λ2Tr −λTr = A + B . for (2.12.7) (2.12.8) (2.12.9) Now, because U Λ is unbounded, we continue a slight modification of the argument in [29]. The only difference is that we bound term (A) above (which corresponds to [29, (iii) in equation (2.6)] as (cid:16) (HΛ − E)(PΛ(I))U ΛP0,Λ(J c)(H0,Λ − E)−2(cid:17)| ≤ |Tr (cid:16) PΛ(I)U Λ(cid:17)|. |I| (|J| − |I|)2|Tr (2.12.10) At this point, with an estimate analogous to the one in the proof of Proposition 3.2 in [14] (cid:110)(cid:90) (m+1)ε (cid:111)|Λ| ε = |I|. (2.12.11) we reach E (|TrPΛ(I)Vω|) ≤ λ−1 sup (cid:110)(cid:82) (m+1)ε m∈N mε Thus, with M1(ε) := supm∈N mε ωjρ(ωj) dωj λ|Tr(HΛ−E)PΛ(I)U ΛP0,Λ(J c)(HΛ 0 −E)−2| ≤ (cid:110)(cid:82) (m+1)ε λ2|TrU Λ(PΛ(I))U ΛP0,Λ(J c)(H0,Λ − E)−2| ≤ Similarly, with M2(ε) := supm∈N mε ωjρ(ωj) dωj (cid:111) , λ|I| (|J| − |I|)2 ( (cid:111) M1(ε) λ + g(cid:107)F(cid:107)∞ λ )|Λ|. (2.12.12) (cid:19) . (cid:18) M2(ε) λ ω2 j ρ(ωj) dωj , we estimate term (B) through 4λ2 (|J| − |I|)2 |Λ| + g2 λ2 Tr(PΛ(I)) (2.12.13) 64 Due lemma 2.5 and the Wegner estimate (see [6, theorem 4.1]) we conclude that Tr(PΛ(I)) ≤ C λ |I||Λ|. (2.12.14) Choosing the interval J such that |J| = εδ for δ < 1, keeping in mind the assumption g2 < λ and combining the bounds (2.12.5), (2.12.7), (2.12.12), (2.12.13), (2.12.14) and optimizing over δ gives δ = 1 2+α0 therefore we reach (2.12.1) for α ∈ [0, α0 2+α0 ] and (2.2.15) is proven. To show 2.3 we follow the proof of theorem 1.2 in [29]. We fix λ, λ(cid:48) ∈ J, g and g(cid:48) satisfying the assumptions of theorem 2.3 and E ∈ I. As explained in [29], using H¨older continuity with respect to energy given by equation (2.2.15), trace identities and ergodicity of Hλ,g and TrP0ϕ(Hλ,g)(ϕ(Hλ,g) − ϕ(Hλ(cid:48),g(cid:48)))P0 (cid:17) where ϕ is a smooth (2.12.15) function such that Hλ(cid:48),g(cid:48), it suffices to estimate E(cid:16)  ϕ ≡ 1 on (−∞, E], ϕ ≡ 0 on (−∞, E + |λ − λ(cid:48)|δ + |g − g(cid:48)|δ)c, (cid:107)ϕ(j)(cid:107)∞ ≤ C (cid:16)|λ − λ(cid:48)|δ + |g − g(cid:48)|δ(cid:17)−j , j = 1, 2..., 3d + 4 with δ > 0 to be determined. The need for a high regularity of ϕ is due to the fact that the random potential Vω may be unbounded. Let ˜ϕ be an almost analytic extension of ϕ of order 3 + 3d. In particular, ˜ϕ is defined in a complex neighborhood of the support of ϕ and if z = E + iη we have that |∂z ˜ϕ(z)| ≤ |η|3d+3|ϕ(3d+4)(E)|. (2.12.16) By the Helffer-Sj¨ostrand formula, the quantity (cid:16) Tr (cid:17) P0ϕ(Hλ,g)(ϕ(Hλ,g) − ϕ(Hλ(cid:48),g(cid:48)))P0 65 equals ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z) (cid:16) λ(cid:48)Uλ(cid:48),g(cid:48) − λUλ,g (cid:17) Rλ(cid:48),g(cid:48)(z)P0 d2z = ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)VωRλ(cid:48),g(cid:48)(z)P0 d2z ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)Veff,λ(g)Rλ(cid:48),g(cid:48)(z)P0 d2z (cid:16) (cid:17) Veff,λ(cid:48)(g(cid:48)) − Veff,λ(g) ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z) Rλ(cid:48),g(cid:48)(z)P0 d2z. (cid:90) 1 π (λ(cid:48) − λ) C (cid:90) (cid:90) C π (g(cid:48) − g) g(cid:48) π C π + + (cid:90) C Since the last two terms enjoy a better modulus of H¨older continuity (since they do not involve Vω) and can be treated as in [29], we shall only estimate the first of the above integrals. By the resolvent identity, Rλ,g(z)VωRλ(cid:48),g(cid:48)(z) =Rλ,g(z)VωRλ,g(z) + (λ − λ(cid:48))Rλ,g(z)VωRλ(cid:48),g(cid:48)(z)VωRλ,g(z)+ (g − g(cid:48))Rλ,g(z)VωRλ(cid:48),g(cid:48)(z)Veff,λ(cid:48)(g(cid:48))Rλ,g(z) − gRλ,g(z)VωRλ(cid:48),g(cid:48)(z)(Veff,λ(cid:48)(g(cid:48)) − Veff,λ(g))Rλ,g(z). (cid:90) ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)VωRλ(cid:48),g(cid:48)(z)P0 d2z (λ(cid:48) − λ) π C The above considerations lead to a perturbative expansion of into four terms. We will show below that each of them can be bounded in terms of powers of either |λ − λ(cid:48)| or |g − g(cid:48)|. We start by estimating E C π (cid:90) (cid:18)(cid:12)(cid:12)(cid:12)(λ − λ(cid:48))2 (cid:18)(cid:12)(cid:12)(cid:12) (λ−λ(cid:48))2 E π ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)VωRλ(cid:48),g(cid:48)(z)VωRλ,g(z)P0 d2z (cid:12)(cid:12)(cid:12)(cid:19) (2.12.17) with a slight modification of equation (3.15) in [29] since Vω is unbounded. By the Combes- Thomas bound, equation (2.12.16) and the choice of ϕ (cid:82)C ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)VωRλ(cid:48),g(cid:48)(z)VωRλ,g(z)P0 d2z ≤ C(d)(cid:0)1 + E2(|Vω|)(cid:1) |λ−λ(cid:48)|2 (|λ−λ(cid:48)|δ+|g−g(cid:48)|δ)3d+4 . (cid:12)(cid:12)(cid:12)(cid:19) 66 Similarly, E (cid:90) (cid:18)(cid:12)(cid:12)(cid:12)(λ − λ(cid:48))(g − g(cid:48)) π C ≤ C(d) (1 + E(|Vω|)) ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)Veff,λ(cid:48)(g(cid:48))Rλ(cid:48),g(cid:48)(z)VωRλ,g(z)P0 d2z |λ − λ(cid:48)||g − g(cid:48)| (|λ − λ(cid:48)|δ + |g − g(cid:48)|δ)3d+4 . Moreover, using lemma 2.11 with the the explicit dependence on ω given there, we obtain (cid:12)(cid:12)(cid:12)(cid:19) (cid:12)(cid:12)(cid:12) that the expected value of (cid:90) (cid:12)(cid:12)(cid:12)g(λ − λ(cid:48)) π C ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)VωRλ(cid:48),g(cid:48)(z)(Veff,λ(cid:48)(g(cid:48)) − Veff,λ(g))Rλ,g(z)P0 d2z is bounded from above by C(d) (1 + E(|Vω|)) |g||λ − λ(cid:48)|(|g − g(cid:48)| + |λ − λ(cid:48)|) (|λ − λ(cid:48)|δ + |g − g(cid:48)|δ)3+3d . (cid:90) Using the same arguments as in [29, Equations 3.17 and 3.18] we see that (cid:12)(cid:12)(cid:12)(λ(cid:48) − λ) (cid:12)(cid:12)(cid:12) |λ − λ(cid:48)||E(cid:0)Tr(P0ϕ(Hλ,g)Rλ,g(z)VωRλ,g(z)P0)(cid:1)| ≤ C|λ − λ(cid:48)|E (|Vω|) ∂z ˜ϕ TrP0ϕ(Hλ,g)Rλ,g(z)VωRλ,g(z)P0 d2z can be bounded from above by C π (|λ − λ(cid:48)|δ + |g − g(cid:48)|δ) . (2.12.18) Finally, we conclude that |Nλ,g(E) − Nλ(cid:48),g(cid:48)(E)| ≤ C(α0, d, I)m(|λ − λ(cid:48)|)m(|g − g(cid:48)|) (2.12.19) where Choosing δ = m(x) = xδα + x2−(3d+4)δ + x1−δ. 2 α+3d+4 we obtain, for any β ∈ [0, |Nλ,g(E) − Nλ(cid:48),g(cid:48)(E)| ≤ C(α0, I) 2α α+3d+4], (cid:16)|λ − λ(cid:48)|β + |g − g(cid:48)|β(cid:17) finishing the proof of theorem 2.3. 67 CHAPTER 3 SPECTRAL AND DYNAMICAL CONTRAST ON HIGHLY CORRELATED ANDERSON-TYPE MODELS. 3.1 Introduction and Main results This chapter aims at presenting and analysing examples of random Schr¨odinger operators where contrasting dynamical and spectral behaviors can be observed. In comparison to the well established theory of Anderson localization, discussed below in detail, the systems studied here will exhibit some form of long range correlations. Depending on their geometry, the spectral properties of the models can change significantly. More surprisingly, we present a model which exhibits purely absolutely continuous spectrum but where a ’phase transition’ can still be observed. Such phenomenon can be captured utilizing the notion of transient and recurrent absolutely continuous spectrum due to Avron and Simon [9], which we shall also review in the subsequent discussion. Before stating our models and main results precisely, we shortly discuss the relevant background on correlated models. 3.1.1 Background on Correlated Models The effects of strong correlations on localization properties of a given lattice were recently studied in [50] where the authors consider a system consisting of a particle and a 1/2 spin and the particle flips the spin only when it visits the origin. Figure 3.1: 1/2 spin and single-particle model There it is shown that resonant tunneling is compatible with correlated pure point spec- trum the since model exhibits Green’s function decay on the graph metric, hence pure point 68 spectrum, but has eigenfunctions which are only localized in the particle position [50, The- orems II.2 and II.5]. 3.1.2 Main Results Our goal in this chapter is to study infinite volume analogues of the model in [50]. These will be obtained by connecting infinitely many copies of the Anderson model in two ways. We will present the contrast between two random Schr¨odinger operators HSym,ω = −ASym,γ +Vω and HDiag,ω = −ADiag,γ + Vω, where the random potential only depends on the first coordinate and is defined through Vω(n) = ω(n1) for n = (n1, n2). By ASym,γ and ADiag,γ we denote weighted adjacency operators of graphs which are infinite (in the horizontal and vertical directions) subsets of Z2 obtained by connecting infinitely many copies of Z along a walk which is either entirely vertical or alternates between vertical components of fixed length (cid:96) > 0 and diagonal components. Figure 3.2: Highly correlated Anderson-type models (0,0) (0,0) The operators HSym,ω and HDiag,ω act in a similar fashion if one fixes n ∈ N and considers the horizontal layer Z × {n} separately, when the models are easily seen to be distinct samples of the Anderson model −∆ + Vω acting on (cid:96)2 (Z). However, as we shall make precise below, by connecting these horizontal components in distinct ways a fundamental difference is introduced. To state the quantum dynamical contrast, let q > 0 be given. The 69 averaged q-moments of a self-adjoint operator H are defined as M q where we recall that the position operator |X| acts as a multiplication operator on (cid:96)2(cid:0)Z2(cid:1) T (H) = (3.1.1) 0 −2t T E(cid:104)0|eitHω|X|qe−itHω|0(cid:105) dt e (cid:90) ∞ 2 T via (|X|qφ) (n) := |n|qφ(n) and |n| = |n1| + |n2| for n = (n1, n2) ∈ Z2. Our first result concerns HSym,ω where a combination of the symmetry and localization in the horizontal direction induces ballistic transport in the vertical direction, up to a logarithmic correction. For the sake of simplicity, we assume that the random variables {ω(n)}n∈Z are nonnegative and bounded throughout this chapter so ω(n) ∈ [0, ωmax] for all n ∈ Z2. 3.1.3 Dynamical Contrast between HSym and HDiag Theorem 3.1. There exists T0 > 0 for which the averaged moments satisfy (cid:18) T for all times T ≥ T0 and some positive constant Cδ0 (cid:1) ≥ Cδ0 (cid:0)HSym M q T (cid:19)q log T which depends on (cid:107)ρ(cid:107)∞. (3.1.2) Our second result concerns HDiag,ω,γ = −ADiag,γ + Vω. Inspired by the analysis of [40] (cid:1) cannot grow faster than a logarithmic power of T , which is an (cid:0)HDiag,γ we show that M q T analogue of [40, Theorem 1]. Theorem 3.2. There are γ0 > 0 and T0 > 0 such that whenever γ < γ0, the averaged moments of the operator HDiag,ω,γ = −ASym,γ + Vω satisfy (cid:0)HDiag,γ (cid:1) ≤ (log T )q + C (q) (3.1.3) M q T for all times T ≥ T0. We remark that γ0 will be explicit in the proof. It depends on (cid:96) and the decay rate of the Green’s function of the one-dimensional Anderson model. 70 3.1.4 Spectral contrast between HSym and HDiag The following is a simple consequence of theorem 3.2. Its proof and the relevant definitions of packing measure and dimension will be given in section 3.6.1. Corollary 3.3. Whenever γ < γ0, the packing dimension dim+ P (HDiag,ω) vanishes almost surely. The spectral contrast between the two models will be evident from the result below. (cid:0)HSym (cid:1) is purely absolutely con- Theorem 3.4. For every ω ∈ Ω the spectral measure µδ0 tinuous and supported on a set of Lebesgue measure 4γ. (cid:0)HSym An interesting feature of theorem 3.4 is that, for small values of the parameter γ, the support of µδ0 since the later contains [0, 2 + ωmax]. As we shall see below, this phenomenon is linked to (cid:1) has Lebesgue measure much smaller than the spectrum of HSym, (cid:1) inside [0, 2 + ωmax] is recurrent. To prove theorem 3.4 the fact that the portion of σ(cid:0)HSym we have shown the following abstract result which, to the best of our knowledge, is new. (cid:82) Proposition 3.5. Let µ be a finite Borel measure which is purely singular and F (z) = 1 u−z dµ(u) its Borel transform, defined whenever z ∈ C+. Then, the limit F (E + i0) = limδ→0+ F (E + iδ) exists for Lebesgue almost every E and (cid:12)(cid:12){E ∈ R : α < E + F (E + i0) < β}(cid:12)(cid:12) = β − α. (3.1.4) As we could not find a reference in the literature with the exact statement needed, the details of the argument are carried out in section 3.7. Proposition 3.5 is liked to a beautiful equality by Boole [10], although with a slightly different statement. Boole’s equality [10] and its extensions have been rediscovered or studied in various contexts by different authors ([45], [57] [20],[21],[36],[22],[52]). For the sake of completeness, we state it below Proposition 3.6. Under the above assumptions (cid:12)(cid:12){E ∈ R : F (E + i0) > t}(cid:12)(cid:12) = C t . (3.1.5) 71 Further historical notes on Boole’s equality can be found in [55, Chapter 5] and [6, Chapter 8]. 3.1.5 Phase transition within σ(cid:0)HSym (cid:1) Our next result sheds light on theorem 3.4 and provides further information on the dynamics −itHSym. Before proceeding with the statement we recall that the notions of transient and e recurrent subspaces where given in the introduction of the thesis. The main point of theorem 3.7 below is that the transient and recurrent spectrum can naturally arise and coexist in a situation of physical relevance. Theorem 3.7. For all values of γ the spectrum of HSym is a random set with [−2, 2 + ωmax] ⊂ σ(cid:0)HSym (cid:1) \ [−2, 2 + ωmax) = σtac(cid:0)HSym (cid:1) ⊂ [−2 − 2γ, 2 + ωmax + 2γ]. (cid:1) and σ(cid:0)HSym (3.1.6) (cid:1) ∩ [−2, 2 + ωmax) = Moreover, σ(cid:0)HSym σrac(cid:0)HSym (cid:1). 3.1.6 Definition of the Models We are going to define the graphs GSym and GDiag by specifying the non-zero matrix el- ements of their adjacency operators on the basis {δn}n∈Z2. In both cases the adjacency operators are symmetric. Let Z+ = N ∪ {0} and IX be the indicator function of a set X. The adjacency operator ASym,γ is defined by  ASym,γ(x, y) = γ if x = (0, n2) and y ∈ {(0, n2 − 1)I{n2>0}, (0, n2 + 1)} (3.1.7) 1 if x = (n1, n2) and y ∈ {(n1 − 1, n2)I{n1>0}, (n1 + 1, n2)} where n1, n2 ∈ Z+ in the above definition. In other words, two vertices x and y satisfying d(x, y) = 1 are connected only if their second coordinate is the same or their first coordinate is zero. The constant γ > 0 is the 72 hopping strength on the vertical direction and it will play a key role later in this chapter. On the other hand, the non-zero matrix elements of ADiag,γ are defined via  ADiag,γ(x, y) = γ if x = (n1, (n1 + 1)(cid:96) − 1) and y = (n1 + 1, (n1 + 1)(cid:96)) , n1 ∈ Z+ 1 if x = (n1, n2) and y ∈ {(n1 − 1, n2)I, (n1 + 1, n2), (n1, n2 + 1)I(cid:96) (3.1.8) where I = I{n1>0} and I(cid:96) = I{n1(cid:96)≤n2<(n1+1)(cid:96)} for n1, n2 ∈ Z+. Note that ADiag,γ is obtained by modifying ASym,γ at points whose second coordinate is a multiple of the fixed length L > 0. 3.2 Lower Bound on the Averaged Moments for HSym: Proof of Theorem 3.1 The proof of Theorem 3.1 consists of rigorously implementing the following ideas. Firstly, the fact that HSym exhibits some form of transport has its roots on the symmetry of the model. We shall prove that the spectral measure µδ(0,0) tinuous and this implies diffusion by the Guarnieri bound, which we shall revisit below. (cid:0)HSym (cid:1) is purely absolutely con- Secondly, the disordered nature of the system and its connection to the Anderson model allow us to show a horizontal localization result. In other words, transport can only occur in the vertical direction. In particular, this will imply a rate of wavepacket spreading which is faster than what one would expected generically. A careful analysis will then imply the almost ballistic lower bound. The Guarnieri bound states that if the spectral measure µψ (H) of a self-adjoint operator H is α-H¨older continuous then T (H) ≥ CψT M q αq 2 . (3.2.1) Recall that µ is said to be α-H¨older continuous if there exists a constant C < ∞ such that for all intervals I with |I| < 1 we have µ(I) ≤ C|I|α. In particular, if µψ is purely q absolutely continuous we can directly reach M q 2 . However, the proof of (3.2.1) T (H) ≥ CψT 73 can be adapted to incorporate the improvements due to the disorder which are specific to our context, therefore we reproduce the main arguments below. The starting point is the following estimate on the averaged quantum dynamics. Theorem 3.8. (Strichartz-Last) Let H be a self-adjoint operator on a Hilbert space (cid:72) and assume the spectral measure of H with respect to ψ is uniformly α-H¨older continuous for some α ∈ (0, 1). Then, there exists a constant Cψ < ∞ such that for all φ ∈ (cid:72) and all T > 0 (cid:90) T 0 1 T |(cid:104)φ|eitH|ψ(cid:105)|2 ≤ Cψ(cid:107)φ(cid:107)2 T α . (3.2.2) The above result, which is found in [6, Theorem 2.3], can be used to reach (3.2.1) as follows. For simplicity, let HSym = H throughout this section. We firstly rewrite M q T (H) as (cid:88) n 2 T |n|q (cid:90) ∞ 0 T |(cid:104)n|e−itHω|0(cid:105)|2 dt ≥ 2e−2 −2t T e |n|>N (cid:90) T N q (cid:88) 1 − (cid:88) (cid:18) 0 |n|≤N 1 T 1 − N (N + 1) 2 0 Cδ0 T α |(cid:104)n|e−itHω|0(cid:105)|2 dt (cid:90) T |(cid:104)n|e−itHω|0(cid:105)|2 dt (cid:19)  = 2e−2N q ≥ 2e−2N q ≥ Cδ0 qα 2 T where on the last step we choose N 2 comparable to T α and the constant Cδ0 was adjusted. In our particular setup, one can improve upon the Guarnieri bound using the localization (cid:90) T 1 T (cid:88) statement for the horizontal direction in lemma 3.11 which combined with theorem 3.8 yields E|(cid:104)n|e−itHω|0(cid:105)|2 dt ≤ (cid:88) 2−s |n1|(cid:111) (cid:17) 1 (cid:19) 2−s . Thus, it is where C = C(γ, ωmax,(cid:107)ρ(cid:107)∞, s) = 2 natural to split the above sum into two contributions according to |n1| ≤ 2−s µAnd (2γ)s(cid:107)ρ(cid:107)∞ (4 + 4γ + ωmax) CAnd(s) (cid:110)Cδ0 − µAnd |n|≤N |n|≤N 0 (3.2.3) (cid:18) min , Ce (cid:16) T π . log CT Cδ0 , 74 in which case min T , Ce (cid:110) Cδ0 2−s |n1|(cid:111) (cid:16) C, s, Cδ0 − µAnd − µAnd 2−s |n1|(cid:111) (cid:17) − µAnd = 2−s |n1| . = Ce , µAnd (cid:110) Cδ0 min T , Ce For T > T0 T , and |n1| > 2−s Cδ0 µAnd log (cid:19) (cid:18) CT Cδ0 which implies the first contribution can be estimated from above by Cδ0 T (cid:93){n ∈ GSym : |n| < N ,|n1| < 2 log T}. (3.2.4) where N must be chosen bigger than 2 log T , in which case there are two possibilities for n to be in the above set. The first one is that n2 ≤ N − 2 log T , when the condition |n| < N is automatically satisfied whenever |n1| < 2 log T . Secondly, if (cid:98)N − 2 log T + j(cid:99) = n2 for some j ∈ {1, ...,(cid:98)2 log T(cid:99)} then |n| < N is only satisfied when |n1| < (cid:98)2 log T(cid:99)− j. This reasoning immediately translates into the bound 2 log T(cid:88) (cid:93){n ∈ GSym : |n| < N ,|n1| < 2 log T} ≤ 2 log T (N − 2 log T ) + (2 log T − j) j=1 = 2 log T (N − 2 log T ) + log T (2 log T − 1) ≤ 2 log T (N − log T ) . Therefore, the first contribution in the splitting of (3.2.3) is bounded from above by log T (N − log T ) 2Cδ0 T . The second contribution is estimated by (cid:88) |n| log T 2 2 T N(cid:88) − µAnd 2−s |n1| ≤ 2C Ce − µAnd 2−s j e T j=(cid:98) log T 2 (cid:99) 2C(N − log T 2 ) = T 1 − e ) (N − log T 2 (cid:16) − µAnd 2−s 1 − e−1 N−(cid:98) log T 2 (cid:99)−1 (3.2.5) (cid:17) . In particular, the choice N = (cid:16) log T + η T log T (cid:17) ensures that, by picking η sufficiently small and eventually changing T0 (but with the same dependence on C, s, Cδ0 , µAnd ), the 75 same argument as in the proof of the Guarnieri bound yields the inequality (cid:0)HSym (cid:1) ≥ Cδ0 M q T (cid:18) T (cid:19)q log T whenever T > T0. Theorem 3.1 is proven. 3.3 The absolute continuity of µδ(0,0) (HSym) (cid:0)HSym (cid:1) follows from recursive We shall explain how the absolute continuity of µδ(0,0) relations for the Green’s function which are available due to the symmetry of GSym. For z ∈ C \ R, let ϕ := (HSym − z)−1δ(0,0). By definition of HSym we have −ϕ(1, 0) − γϕ(0, 1) + (ω(0) − z)ϕ(0, 0) = 1. (3.3.1) hence −GSym ((0, 0), (1, 0); z) − γGSym ((0, 0), (0, 1); z) + (ω(0) − z)GSym ((0, 0), (0, 0); z) − 1 = 0. (3.3.2) Thus, denoting by GAnd the Green’s function of the Anderson model on (cid:96)2 (Z), it follows from Feenberg’s expansion, explained in the introduction of this thesis, that Sym ((0, 1), (0, 1); z) + (ω(0) − z)−(cid:17) − 1 = 0. GSym ((0, 0), (0, 0); z) And(1, 1; z) − γ2G+ (cid:16)−G+ Where in a path expansion of the form G(u, v; z) = G(u, u; z)G+(v, v; z) the term G+(v, v; z) denotes the Green’s function of the operator HSym restricted to the component of GSym containing v and obtained from GSym by removing the edge which connects the sites u and v. By the vertical symmetry of GSym we have that (3.3.3) G+ Sym ((0, 1), (0.1); z) = GSym ((0, 0), (0, 0); z) (3.3.4) thus we can rewrite equation (3.3.3) as (cid:16)−G+ GSym ((0, 0), (0, 0); z) And(1, 1; z) − γ2GSym ((0, 0), (0, 0); z) + (ω(0) − z) (cid:17) − 1 = 0. (3.3.5) 76 Therefore, letting w = −ω(0) + z + G+ And(1, 1; z) we have 2γ2GSym ((0, 0), (0, 0); z) = −w + (3.3.6) where we have chosen the branch of the square root onto the upper half-plane and the reason for the positive sign on the second of the above terms is that GSym ((0, 0), (0, 0); z) necessarily has non-negative imaginary part since HSym is self-adjoint. Equation (3.3.6) implies that 2γGSym ((0, 0), (0, 0); z) = −w + w . (3.3.7) Therefore, using either equation (3.3.6) or (3.3.7) depending on whether |w| is small or large, we conclude that GSym ((0, 0), (0, 0); z) remains bounded as Imz → 0 for any z ∈ C+ thus the spectral measure µδ(0,0) proposition B.4 in [6, Appendix B]. (cid:1) is purely absolutely continuous, see, for instance, (cid:0)HSym 3.4 Floquet Theory for HSym The vertical symmetry of the graph GSym and the definition of the operator HSym suggests that the use of a Fourier transform may be helpful when studying the dynamics (cid:104) w2 − 4γ2(cid:105)1/2 (cid:19)1/2 (cid:18) 1 − 4γ2 w2 −itHSym. Let us define e through (cid:1) → (cid:96)2 (Z+) ⊗ L2 ([0, π]) (cid:70) : (cid:96)2(cid:0)GSym (cid:114) 2 ∞(cid:88) π n2=0 ((cid:70)ψ) (n1, p) := ψ(n1, n2) sin (p(n2 + 1)) . (3.4.1) For simplicity of notation we let ˆψ(n1, p) = (cid:70) (ψ) (n1, p). It is immediate to check that the following version of Plancherel’s identity is satisfied (cid:104) ˆϕ, ˆψ(cid:105)(cid:96)2(Z+)⊗L2([0,π]) = (cid:104)ϕ, ψ(cid:105) (cid:96)2(cid:16)G Sym (cid:17). (3.4.2) 77 Indeed, (cid:104) ˆϕ, ˆψ(cid:105)(cid:96)2(Z+)⊗L2([0,π]) = = × = ˆϕ(n1, p) ˆψ(n1, p) dp ϕ(n1, n2)ψ(n1, n(cid:48) 2) 0 2 π 2 π n1∈Z+ (cid:90) π (cid:88) (cid:88) (cid:90) π sin ((n2 + 1)p) sin(cid:0)(n(cid:48) n1,n2,n(cid:48) (cid:88) (cid:96)2(cid:16)G n1,n2∈Z+ 2∈Z+ (cid:17). 0 Sym ϕ(n1, n2)ψ(n1, n2) = (cid:104)ϕ, ψ(cid:105) 2 + 1)p(cid:1) dp A calculation which is almost identical to the one above gives that, for every ψ ∈ (cid:96)2(cid:0)GSym (cid:1) where Using that(cid:8)(cid:113) 2 −1( ˆψ) = ψ. (cid:90) π (cid:70) (cid:114) 2 (cid:70) (cid:70) = g. (cid:16) (cid:114) 2 π (cid:17) −1(g) (cid:90) π (cid:16) 0 −1(g)(n1, n2) = (cid:70) : n ∈ N(cid:9) is a complete orthonormal system in [0, π] one may also g(n1, p) sin (p(n2 + 1)) dp. π 0 π sin(np) check that for g ∈ (cid:96)2 (Z+) ⊗ L2 ([0, π]) Therefore, (cid:70) is a unitary map with inverse given by −1(g)(n1, n2) = (cid:70) g(n1, p) sin (p(n2 + 1)) dp. (3.4.3) (cid:17) ˆψ(n1, p). From the definition of HSym one readily sees that (cid:92)HSymψ(n1, p) = −∆ ˆψ(n1, p) + (3.4.4) where we have committed a slight abuse of notation by letting ∆ ˆψ(n1, p) = ˆψ(n1 − 1, p) + ˆψ(n1 + 1, p). In particular, equation (3.4.4) allow us to conclude that HSym is unitarily ω(n1) − 2γ cos pδn1=0 equivalent to an operator acting on (cid:96)2 (Z+) ⊗ L2 ([0, π]) 78 (3.4.5) with action on each fiber given by (cid:104)p where (cid:104)p is a rank-one perturbation of the one di- mensional Anderson model given by (3.4.4). A consequence of this fact is the following result Lemma 3.9. δ(0,0) is a cyclic vector for HSym. Proof. Firstly, we claim that given n1 ∈ N, the vector δ(n1,0) belongs to the cyclic subspace generated by δ(0,0), denoted henceforth by (cid:72)0. Indeed, by definition of the Fourier transform (3.4.1), ˆδ(n1,0) = (cid:114) 2 π sin p δn1. (3.4.6) By a result of Simon [54] the vector δ0 is a cyclic vector for (cid:104)p thus we conclude from (3.4.6) that ˆδ(n1,0) belongs to the cyclic subspace, relative to (cid:104)p, generated by ˆδ(0,0). By taking the inverse Fourier transform, it follows that δ(n1,0) ∈ (cid:72)0. Since HSymδ(0,0) = −γδ(0,1) + δ(1,0) + ω(0)δ(0,0) we easily conclude that δ(0,1) ∈ (cid:72)0 as well. Proceeding by induction one shows that δ(n1,n2) ∈ (cid:72)0 for every pair (n1, n2) ∈ GSym, finishing the proof. (cid:0)HSym (cid:1) is absolutely continuous we readily obtain (cid:1) is purely absolutely continuous. Corollary 3.10. σ(cid:0)HSym Since µδ(0,0) 3.4.1 Localization in the horizontal direction The above considerations are particularly effective to translate dynamical localization for the one dimensional Anderson model into horizontal localization for HSym. We start with the following technical result. Lemma 3.11. Given s ∈ (0, 1) we have, for all m, n ∈ Z+ (cid:32) E sup |f|≤1 |(cid:104)m|f(cid:0)(cid:104)p (cid:33) (cid:1)|n(cid:105)| | ≤ (2γ) 2 2−s ((cid:107)ρ(cid:107)∞ (4 + 4γ + ωmax) CAnd(s)) 1 2−s e − µAnd 2−s |m−n| . (3.4.7) 79 The above supremum is taken over all Borel measurable functions bounded by one. Proof. It suffices to show that for every L ∈ N (3.4.7) holds with (cid:104)p replaced by its restriction to (cid:96)2 (Z+ ∩ [0, L]), denoted henceforth by (cid:104)L p . Let vp = −2γ cos p. From rank-one perturbation formulas (see, for instance, [6, Theorem 5.3] and [3, Equation(A.7)] ) we find that the spectral measure of (cid:104)L p is given by dµp,L m,n(E) = −vpGL And (m, n; E) δ(Σ(E) − vp) dE (3.4.8) (3.4.9) where Σ(E) := − 1 GL And (0, 0; E) and GL And(m, n; E) denotes the Green’s function for the Anderson model on (cid:96)2 (Z+ ∩ [0, L]). Therefore, by general structure of the spectral measures, (cid:90) (cid:12)(cid:12)(cid:12)GL |vp|2 (cid:12)(cid:12)(cid:12)2 And (m, n; E) δ(Σ(E) − vp) dE ≤ 1. (3.4.10) A second observation is that the analogue of equation (3.4.8) for m = n reads In particular dµp,L m,m(E) = δ(Σ(E) − vp) dE. (cid:90) ∞ −∞ δ(Σ(E) − vp) dE = 1. (3.4.11) (3.4.12) Combining equations (3.4.8), (3.4.10) and (3.4.12) with H¨older’s inequality (applied to the exponents (p, q) = (2 − s, 2−s that for all intervals I ⊂ R |vp|sE (cid:20) m,n E(cid:16)(cid:12)(cid:12)(cid:12)µp,L (cid:12)(cid:12)(cid:12) (I) (cid:17) ≤ (cid:12)(cid:12)(cid:12) (I) E(cid:16)(cid:12)(cid:12)(cid:12)µp,L (cid:17) ≤ (2γ) m,n Thus, 1−s)) and Jensen’s inequality for expectations, we conclude (cid:18)(cid:90) (cid:12)(cid:12)(cid:12)GL I And (m, n; E) (cid:19)(cid:21) 1 2−s δ(Σ(E) − vp) dE . (3.4.13) (cid:12)(cid:12)(cid:12)s (cid:18)(cid:90) I s 2−s(cid:107)ρ(cid:107) 1 2−s∞ E (|GAnd (m, n; E)|s) dE (cid:19) 1 2−s . (3.4.14) 80 Since the operator (cid:104)p is bounded, with operator norm less or equal than 2 + 2γ + ωmax, the inequality (3.4.14) suffices to conclude the proof of lemma 3.11. We mention that by introducing an integrable weight, one could also handle the case of unbounded operators. For further details we refer to [3, Equations (A.13)-(A.18)]. The localization statement for (cid:104)p described on Lemma 3.11 can be immediately translated into a (horizontal) localization statement for HSym with the use of Plancherel’s identity (3.4.2) and the fact that ˆδ(m1,m2)(n1, p) = The precise statement reads (cid:114) 2 π δm1(n1) sin (p(m2 + 1)) . (3.4.15) Lemma 3.12. E (cid:32) (cid:12)(cid:12)(cid:104)δm1,m2, f(cid:0)HSym (cid:1) δn1,n2(cid:105)(cid:12)(cid:12)(cid:33) (cid:0)4γ2(cid:107)ρ(cid:107)∞ (4 + 4γ + ωmax) CAnd(s)(cid:1) 1 sup |f|≤1 2−s . with C = 2 π ≤ Ce − µAnd 2−s |m1−n1| . (3.4.16) 3.4.2 Transient and recurrent components: proof of theorem 3.7 According to [9, Proposition 3.1] in order for an absolutely continuous measure µϕ to be transient it is necessary that µϕ = f (E) dE where f ∈ C∞ (R). Moreover, if f is also assumed to be compactly supported this condition also sufficient. Since the operators studied here are bounded, this distinction will not be relevant to us. Let We claim that S cannot be the support of a smooth function and this implies that σ(cid:0)HSym S = supp(µδ(0,0) ) ∩ [−2, 2 + ωmax). (cid:1)∩ [−2, 2 + ωmax) is recurrent. The claim is a consequence of the lemma below. Lemma 3.13. Let I ⊂ [−2, 2+ωmax] be a closed interval. Then, I∩Sc has positive Lebesgue measure. 81 Proof. Let µ+ 1 be the spectral measure of the Anderson model on the half line [+1,∞) ∩ Z, And, associated to δ1. Since, almost surely, H+ And has dense point spectrum in 1 (I) > 0. It follows from [53, Theorem 1.5] (theorem one in denoted by H+ [−2, 2 + ωmax) we have that µ+ [52], when restated in terms of Borel measures on the real line, would also suffice) that In particular, by choosing t sufficiently large we may conclude that lim And(1, 1; E)| > t}(cid:12)(cid:12) > 0. t→∞ t(cid:12)(cid:12){E ∈ I : |G+ (cid:12)(cid:12){E ∈ I : |ω(0) − E − G+ And(1, 1; E)| > 2γ}(cid:12)(cid:12) > 0. (cid:18) (cid:19) (3.4.17) (3.4.18) since the above set does not belong to supp µδ(0,0) (this follows from the formula given at the end of section 3.3. An alternative argument is provided by equation (3.4.21) below) the lemma is proven. To show the remaining portion of theorem 3.7, having a formula for the spectral measure will be useful. Let HL µδ(0,0) explained in the previous section, for any L > 0 we have Sym denote the restriction of HSym to (cid:96)2 ({0, ..., L} × Z+). As (cid:17) (cid:17) (cid:16) (cid:16) (cid:104)δ(0,0), f HL Sym δ(0,0)(cid:105) = (cid:90) π (cid:90) π 0 0 2 π 2 π = (cid:104)L p δ0(cid:105) sin2 p dp. (cid:90) ∞ (cid:104)δ0, f −∞ f (E)dµp,L(E) sin2 p dp. dµp,L(E) = δ (ΣL(E) + 2γ cos p) dE (3.4.19) where and ΣL(E) = − . In particular, 1 (0,0;E) GL And (cid:16) (cid:17) (cid:104)δ(0,0), f HL where we have used the fact that, in this case,(cid:82) 1−1 δ(Σ(E)+v) dv holds for E ∈ supp(µδ(0,0) dE. Sym (3.4.20) ). δ(0,0)(cid:105) = 1 πγ 1 − ΣL(E)2 4γ2 Letting L → ∞ in the above equation we conclude that dµδ(0,0) (E) = 1 πγ 1 − Σ2(E) 4γ2 dE. (3.4.21) (cid:115) (cid:90) {|ΣL(E)|<2γ} f (E) (cid:115) 1{|Σ(E)|<2γ} 82 Since Σ(E) is a smooth function on [−2, 2 + ωmax]c, it follows that σ(cid:0)HSym such that σ(cid:0)HSym (cid:1) ∩ [−2, 2 + (cid:1) ∩ [−2, 2 + ωmax]c (cid:54)= ∅ whenever γ > γω. Indeed, this follows from the The above analysis also shows that for any ω ∈ Ω there exist a hopping parameter γω ωmax]c is transient, finishing the proof of theorem 3.7. formula (3.4.21) since Σ(E) is increasing on [−2, 2 + ωmax]c. The critical values is then where Ec = 2 + ωmax. For completeness we provide a proof that this number γω = |Σ(Ec)| 2 is non-zero almost surely. Lemma 3.14. The Green’s function for the half-line Anderson model satisfies −Σ(E) = ω(0) − E − G+ And(1, 1; E). Moreover, at the edge Ec = 2 + ωmax we have Σ(Ec) (cid:54)= 0 almost surely. (3.4.22) (3.4.23) Proof. By definition of the half-line Anderson model H+ And, letting ϕ = (H+ And − z)−1δ0 we obtain −ϕ(1) + (ω(0) − z) ϕ(0) = 1. Since ϕ(0) = G+ And(0, 0; z), ϕ(1) = G+ And(0, 1; z) and, by Feenberg’s expansion, G+ And(0, 1; z) = G+ And(0, 0; z)G+ And(1, 1; z) and equation (3.4.22) is proven. For the sake of contradiction, assume that Then, by (3.4.22) P (Σ(Ec) = 0) = δ > 0. P(cid:0)ω(0) − E − G+ And(1, 1; E) = 0(cid:1) = δ > 0 (3.4.24) And(1, 1; Ec) is independent of ω(0) this would contradict the fact that the distribu- Since G+ tion of ω(0) has a density. We conclude that P (Σ(Ec) = 0) = 0. 83 3.5 Absence of Diffusion for GDiag,γ: Proof of theorem 3.2 A key step to show absence of diffusion for HDiag is proving that, on expectation, the fractional moments of the Green’s function of this operator decay exponentially. We start with a particular case. Lemma 3.15. Let s ∈ (0, 1) be given. For all m ∈ Z+, r ∈ {0, 1, ..., (cid:96)} and z ∈ C+ we have (cid:0)|GDiag ((0, 0), (m, m((cid:96) + 1) + r); z)|s(cid:1) ≤ CAP (CAPγs)m . E ω(0),...,ω(m) (3.5.1) where CAP is the constant in the a priori bound (3.8.4). Proof. Fix r ∈ {0, 1, ..., (cid:96)}, z ∈ C+ and let xm,r = (m, m((cid:96) + 1) + r). Let us proceed by induction in m. When m = 0 the statement reduces to the a priori bound verified in the appendix, equation (3.8.4). Suppose that the desired conclusion holds for some m ∈ Z+ and recall the factorization G(cid:0)0, xm+1,r; z(cid:1) = γG(cid:0)0, xm,(cid:96); z(cid:1) G+ (cid:0)xm+1,0, xm+1,r; z(cid:1) . (3.5.2) Taking absolute values on both sides of (3.5.2), raising to the power s and taking expec- Integrating out the variables ω(0), ..., ω(m), using the inductive assumption and the fact that tations we reach E(cid:0)|G(cid:0)0, xm+1,r; z(cid:1)|s(cid:1) = γsE(cid:0)|G(cid:0)0, xm,(cid:96); z(cid:1)|s|G+ (cid:0)xm+1,0, xm+1,r; z(cid:1)|s(cid:1) . (cid:0)xm+1,0, xm+1,r; z(cid:1) only depends on ω(j) for j > m, we obtain from (3.5.3) that E(cid:0)|G(cid:0)0, xm+1,r; z(cid:1)|s(cid:1) ≤ (CAPγs)m+1 E(cid:0)G+ (cid:0)xm+1,0, xm+1,r; z(cid:1)|s(cid:1) . G+ (3.5.3) (3.5.4) Making use of the a priori bound (3.8.4) one more time finishes the proof. Making use of the Green’s function decay for the one-dimensional Anderson model we readily obtain that 84 E (|G ((0, 0), (m + j, m((cid:96) + 1) + r); z)|s) ≤ CAPCAnd (CAPγs)m e−µAndj. (3.5.5) Pick γ small enough so that CAPγs < e−µAnd((cid:96)+2). (3.5.6) Then, (3.5.5) implies Lemma 3.16. For any s ∈ (0, 1) there exist a constant CDiag such that for all n ∈ GDiag and z ∈ C \ R E(cid:0)|GDiag (0, n; z)|s(cid:1) ≤ CDiage −µDiag|n| (3.5.7) with CDiag = eµAnd(cid:96)CAPCAnd and µDiag = µAnd. Proof. The above argument immediately yields the desired conclusion for z ∈ C+. By symmetry, reasoning along the same lines we obtain (3.5.7) with |GDiag (0, n; z)| replaced by |GDiag (n, 0; z)| and z ∈ C+. Since GDiag (n, 0; z) = GDiag (0, n; z) the proof is finished. 3.6 Upper bounds for quantum dynamics revisited We are now ready to prove theorem 3.2. Our approach is a combination of the fractional moment method and techniques used by Jitomirskaya and Schulz-Baldes in [40]. Let q > 0 and, to simplify the notation, let H = HDiag throughout this section. Recall that −2t T E(cid:104)0|eitHω|X|qe−itHω|0(cid:105) dt. e (3.6.1) (cid:90) ∞ 0 M q T = 2 T The main advantage of working with time-averaged moments is that they can be directly related to the Green’s function via the following formula. M q T = 1 πT |n|q E|G(n, 0; E + )|2 dE. i T (3.6.2) R To verify that (3.6.2) holds, following the proof of lemma 3.2 in [42], one can observe that (cid:90) (cid:88) (cid:90) ∞ n∈Z −2t T |(cid:104)0|eitHω|n(cid:105)|2 dt e 0 85 is the L2 norm of the function − t T g(t) = e restricted to the positive real line. Let (cid:90) R eiE(cid:48)t dµ0,n(E(cid:48)) (cid:90) e−iEtg(t) dt ˆg(E) = 1√ 2π R (cid:90) be the Fourier transform of g. It is easy to check that dµ0,n(E(cid:48)) E(cid:48) − E + i T By definition of the spectral measure µ0,n we discover that i√ 2π ˆg(E) = R (3.6.3) . (3.6.4) ˆg(E) = i√ 2π G(0, n; E − i T ). (3.6.5) Equation (3.6.2) then readily follows from Plancherel’s identity. Returning to the proof theorem 3.2, another useful observation is that, as in to [40, Proposition 1], it suffices to show the desired upper bound for the related quantity M q T (E0, E1) = whenever α > 1 and σ(cid:0)HDiag (cid:90) E1 E0 1 πT (cid:88) (cid:1) ⊂ (E0, E1). |n|N E|G(n, 0; E + i T i T )|2 ≤ CDiage −µDiag|n| T 2−s. )|2 ≤ CDiagT 2−s (cid:88) |n|>N −µDiag|n| e . (3.6.6) By following exactly the same steps in [40] we can now finish the proof. Namely, we start with the general result 86 |n|q 1 πT E|G(n, 0; E + )|2 dE ≤ N q. i T Lemma 3.17. ([40, Lemma 5])(cid:88) (cid:90) E1 E0 applied with N = (cid:98)log T(cid:99) to obtain |n|≤N (cid:88) |n|≤N (cid:90) E1 E0 |n|q 1 πT E|G(n, 0; E + i T )|2 dE ≤ (log T )q. (3.6.7) On the other hand, using equation (3.6.6) followed by the explicit bound on the exponentially decaying sum [40, Lemma 2] we obtain, with E+ := E1 − E0 (cid:88) |n|>log T (cid:90) E1 E0 1 πT )|2 dE ≤ CDiagE+T 1−s (cid:88) E|G(n, 0; E + i T |n|qe −µDiagN |n|>log T −µDiag(log T ) ((log T ) + (µDiag)−1)q+1 ≤ CDiagE+T 1−se × 2(q + 1)! µDiag ≤ C(s, q). whenever s is chosen so that 1 − s < µDiag (we remark that this choice is possible since µDiag = µAnd and the later can be taken proportional to s, see [6, Theorem 12.11 and Equation (12.86)] ). Combining the above estimate with (3.6.7) we finish the proof of theorem 3.2 3.6.1 Paking Dimension: Proof of Corollary 3.3 Before proving corollary 3.3, recall that the upper packing dimension of a Borel probability measure µ is defined as where, for E ∈ supp µ, dim+ P (µ) := µ − ess sup E∈R dµ(E) (3.6.8) dµ(E) := lim sup ε→0 log (µ[E − ε, E + ε]) log(2ε) . (3.6.9) 87 If E /∈ supp µ we set dµ(E) = ∞. We shall denote by dim+ dim+ P (HDiag,ω) the packing dimension is the spectral measure associated to HDiag,ω and the vector P (µHDiag ) where µHDiag δ(0,0). Proof. (of corollary 3.3) From [35, Theorem 1] we have that, for each ω ∈ Ω, dimP (HDiag,ω) ≤ lim sup T→∞ log(cid:0)M q T (HDiag,ω)(cid:1) . q log T Taking expectations and a subsequence Tk → ∞ which realizes the lim sup we obtain, by Fatou’s lemma and Jensen’s inequality, E(cid:0)dimP (HDiag)(cid:1) ≤ lim sup Tk→∞ log E(cid:16) (cid:17) M q Tk (HDiag,ω) q log Tk where, by theorem 3.2, the limit in the right-hand side equals zero. 3.7 A Version of Boole’s Equality for level sets of Heglotz func- tions: proof of proposition 3.5 Consider F (E) =(cid:80)N pn un−E for real numbers {un}N n=1,{pn}N n=1. This is the specific n=1 form taken by the diagonal elements of the Green’s function in finite volume. For a real number α, let Qα be a polynomial of degree N + 1 given by Qα(E) = (α − E − F (E)) (E − un). Qα. Therefore, the coefficient of EN in the expansion Qα(E) = −(cid:81)N +1 n=1 vn. On the other hand, by definition of Qα, this coefficient must be α +(cid:80)N +1 (cid:80)N +1 Note that the solutions v1, ..., vN +1 of the equation E + F (E) = α coincide the roots of n=1 (E − vn) equals n=1 un, thus Replacing α by β we conclude the solutions w1, ..., wN +1 of E + F (E) = β satisfy N(cid:89) n=1 N(cid:88) N(cid:88) N +1(cid:88) N +1(cid:88) vn = α + un. n=1 n=1 wn = β + un. n=1 n=1 88 (3.7.1) (3.7.2) The set {E ∈ R : α ≤ E +F (E) ≤ β} is a disjoint union of intervals ∪N +1 n=1 [vn, wn] therefore, we may conclude from equations (3.7.1) and (3.7.2) that (cid:12)(cid:12){E ∈ R : α ≤ E + F (E) ≤ β}(cid:12)(cid:12) = N +1(cid:88) (wn − vn) n=1 = β − α. The following argument is inspired by the analysis in [6, Proposition 8.2] and provides a proof which is also valid in the infinite-volume context. Proof. Since µ is assumed to be purely singular, the boundary values F (E + i0) are real numbers for almost every E ∈ R (see [6, Proposition B3]). Note that the indicator function 1{E : α < E + F (E + i0) < β} can be represented as φα,β (E+F (E+i0)) for π φα,β(z) = ImLog(z − β) − ImLog(z − α) (3.7.3) and where Log denotes the principal branch of the logarithm. Let ψη(E) = η2 E2 + η2 . We then have, by dominated convergence and definition of φα,β, |{E ∈ R : α ≤ E + F (E + i0) < β}| = lim η→∞ (cid:90) ∞ −∞ π−1ψη(E)φα,β(E + F (E + i0)) dE. Using dominated convergence one more time and recovering φα,β from its boundary values, (cid:90) ∞ −∞ π−1ψη(E)φα,β(E + F (E + i0)) dE = lim η→∞ ε→0+ lim η→∞ (cid:90) ∞ −∞ π−1ψη(E) × φα,β(E + F (E + iε)) dE η ηφα,β(iη + F (iη + iε)) = lim η→∞ ε→0+ = lim η→∞ ηφα,β(iη + F (iη)). 89 On the other hand, by definition of φα,β we know φα,β(iη + F (iη)) = lim η→∞ Im 1 E − iη − F (iη) dE (3.7.4) (cid:90) β α hence lim η→∞ ηφα,β(iη + F (iη)) = lim η→∞ Im η E − iη − F (iη) dE η2 + ηImF (iη) (E − ReF (iη))2 + (η + ImF (iη))2 dE (cid:90) β (cid:90) β α = lim η→∞ = β − α α where we have made use of the simple fact that limη→∞ F (iη) = 0 and limη→∞ ηImF (iη) = µ(R). 3.8 A priori bounds on the Green’s function To obtain a-priori bounds for the Green’s function, we will make use of the weak L1 bound [2, Lemma 3.1]. This bound is valid for any maximally dissipative operator A on a separable Hilbert space (cid:72) and Hilbert-Schmidt operators M1 : (cid:72) → (cid:72)1 and M2 : (cid:72)1 → (cid:72) where (cid:72)1 is another separable Hilbert space. Recall that a densely defined operator A is said to be dissipative if Im(cid:104)ϕ, Aϕ(cid:105) ≥ 0 for every ϕ ∈ D(A). A is said to be maximally dissipative when it is dissipative and has no proper dissipative extension. Denoting by | · | Lebesgue measure and by (cid:107) · (cid:107)HS the Hilbert-Schmidt norm, we have Lemma 3.18.(cid:12)(cid:12)(cid:12){v : (cid:107)M1 M2(cid:107)HS > t}(cid:12)(cid:12)(cid:12) ≤ CW(cid:107)M1(cid:107)HS(cid:107)M2(cid:107)HS 1 t 1 A − v + i0 (3.8.1) where the constant CW is independent of A,M1 and M2. For a proof, we refer to [2, Lemma 3.1]. Another fact, which can be found in [2, Propo- sition 3.2], is 90 Lemma 3.19. Let A,M1 and M2 be as above and let U1, U2 be nonnegative operators. (cid:12)(cid:12)(cid:12){(v1, v2) ∈ [0, 1]2 : (cid:107)M1U 1/2 1 1 A − v + i0 U 2 M2(cid:107)HS > t}(cid:12)(cid:12)(cid:12) ≤ 2CW(cid:107)M1(cid:107)HS(cid:107)M2(cid:107)HS 1/2 We also make use of the Birman-Schwinger relation [2, Lemma B.1] P (z − H − vP )−1 P = P (z − H)−1P − v (cid:16) (cid:17)−1 for a projection P onto a subspace of (cid:72) and the fact that P (z − H)−1P is maximally dissipative on (KerP )⊥, where the equality in question holds. It immediately follows that (cid:12)(cid:12)(cid:12){v : (cid:107)M1P (z − H − vP )−1 P M2(cid:107)HS > t}(cid:12)(cid:12)(cid:12) ≤ CW(cid:107)M1(cid:107)HS(cid:107)M2(cid:107)HS 1 t . equations (3.8.1), (3.8.2) and (3.8.3) together easily imply following apriori bound, for details we refer the reader to [51, appendix A] 1 t (3.8.2) (3.8.3) Em,n Λ (cid:16)|GDiag (cid:110)(2CW ωmax(cid:107)ρ(cid:107)∞)s (m, n; z)|s(cid:17) ≤ CAP(s) (cid:0)4CW ω2 1 − s , (3.8.4) (cid:111) . max(cid:107)ρ(cid:107)2∞(cid:1)s 1 − s CAP(s) = max and we remark that the constant can be taken as 91 APPENDIX 92 APPENDIX TECHNICAL COMMENTS A.1 Conditions on the density ρ imposed in chapter 2 Assume that for some M > 0 and all v ∈ R (cid:12)(cid:12) d dv log ρ(v)(cid:12)(cid:12) ≤ M. This condition is satisfied, for instance, by the Cauchy and exponential distributions. Then Letting h(v) = e−c|v| with c > 0 to be determined. The following estimate holds: ρ(x) ρ(v) = elog ρ(x)−log ρ(v) ≥ e−M|x−v|. (cid:90) ∞ −∞ e−(M +c)|x−v| dx (cid:90) ∞ −∞ h(x − v) dx ≥ ρ(x) ρ(v) therefore, letting ¯ρ(v) = (cid:82) ∞−∞ ρ(x)h(x−v) dx ρ(v) = 2 M + c . , we have shown that ¯ρ(v) ≤ M +c 2 for all v ∈ R. Let us now verify that ¯ρ decays in two cases, starting with when ρ is an exponential. For simplicity, let us ignore normalization factors and let ρ(v) = e−|v|. Then, by the triangle inequality, −∞ ρ(x)h(x − v) dx ≥ ec|v|(cid:90) ∞ (cid:90) ∞ −∞ e−(1+c)|x| dx 2ec|v| 1 + c . = Thus, in this situation ¯ρ(v) ≤ 1+c h(v) = e−cv2 with c > 0 to be determined. In this case, we have that 2 e−(1+c)|v|. Let us now assume that ρ(v) = e−v2 and ρ(v)h(x − v) ≥ e−(1+2c)x2−2cv2 93 thus (cid:90) ∞ −∞ ρ(v)h(x − v) dx ≥ e−2cv2(cid:90) ∞ −∞ e−(1+2c)x2 dx √ πe−2cv2 √ 1 + 2c . = In particular, whenever 2c < 1 we may conclude that ¯ρ decays according to ¯ρ(v) (cid:46) e−(1−2c)v2 . A.2 Derivation of the self-consistent equations Let f (Γ) = Tr (Γ log Γ) where 0 < Γ < 1 is a symmetric matrix. Then, given H which is also symmetric with 0 < H < 1 we have (cid:16) (cid:16) (cid:17) (cid:17) + H log Γ(I + Γ−1H) . I + Γ−1H f (Γ + H) = f (Γ) + Tr Γ log Using the power expansion of the logarithm log(A) = A − 1 2 A2 + ... we may conclude that f (Γ + H) = f (Γ) + Tr (H(I + log Γ) + O((cid:107)H(cid:107)2). therefore, we have the following expression for the Frech´et derivative of f f(cid:48)(Γ) = log Γ + I. Similarly, the derivative of the map Γ (cid:55)→ f (1 − Γ) equals −(I − Γ) log(I − Γ) − I. We may conclude that the entropy (cid:83)(Γ) = −Tr (Γ log Γ + (1 − Γ) log(1 − Γ)) satisfies (cid:48)(Γ) = − (log Γ − log(I − Γ)) . (cid:83) Moreover, the energy functional (cid:69)(Γ) = Tr (H0 − κ + λVω) Γ + g (cid:88) n (cid:104)n|Γ↑|n(cid:105) (cid:104)n|Γ↓|n(cid:105) , can be more succinctly expressed as (cid:69)(Γ) = Tr(cid:0)(H0 + λVω) Γ + gDiag(Γ↑)Diag(Γ↓)(cid:1) 94 from which we see that (cid:48)(Γ) = (cid:69)  H0 + λVω + gDiag(Γ↓) 0  . 0 H0 + λVω + gDiag(Γ↓) In particular, we see from the above equations that whenever (cid:69)(cid:48)(Γ) − β−1(cid:83)(cid:48)(Γ) = 0 the exp(cid:0)−β(H0 + λVω + gDiag(Γ↓)(cid:1) 0  . matrix Γ must be such that Γ(I − Γ)−1 = from which we conclude that  exp(cid:0)−β(H0 + λVω + gDiag(Γ↓)(cid:1) (cid:0)1 + exp(cid:0)βH↓(cid:1)(cid:1)−1 Γ = 0 0  (cid:0)1 + exp(cid:0)βH↑(cid:1)(cid:1)−1 0 where H↓ = H0 + λVω + gDiag(Γ↓) and H↑ = H0 + λVω + gDiag(Γ↑). 95 BIBLIOGRAPHY 96 BIBLIOGRAPHY [1] Aizenman M. Localization at Weak Disorder: Some Elementary Bounds Reviews in Mathematical Physics, Volume 6, Issue 5a, pp. 1163-1182 (1994). [2] Aizenman M., Elgart A. , Naboko S., Schenker J. and Stolz G. Moment analysis for localization in random Schr¨odinger operators Invent. math. 163, 343-413 (2006). [3] Aizenman M., J.H Schenker, R.M. Friedrich and D. Hundertmark. Finite-Volume Fractional-Moment Criteria for Anderson Localization Comm. Math. Phys. 224 (2001), 219-254. [4] Aizenman, M and Graf G.M. Localization Bounds for an Electron Gas (1998) J.Phys.A, Math. Gen., No 32, 6783-6806. [5] Aizenman, M. and Molchanov ,S. Localization at large disorder and extreme energies: an elementary derivation. (1993) Comm. Math. Phys 157(2), 245-278. [6] Aizenman, M. and Warzel, S. Random Operators: Disorder Effects on Quantum Spectra and Dynamics Graduate Studies in Mathematics 18. [7] Aizenman, M. and Warzel, S. Localization Bounds for Multiparticle Systems (2009) Commun Math. Phys 290, 903-934. [8] Anderson, P. Absence of Difusion in Certain Random Lattices (1958) Phys. Rev. 109, 1492-1505. [9] Avron, Simon. Transient and Recurrent Spectrum. Journal of Functional Analysis. 43. 1-31. (1981). [10] Boole, G. On the comparison of transcendents, with certain applications to the theory of definite integrals Philos. Trans. Roy. Soc. London 147 (1857), 745–803. [11] V Bach, E. Lieb and J.P. Solovej.: Generalized Hartree-Fock Theory and the Hubbard Model Journal of Statistical Physics Vol 76, Nos 1/2, 1994. [12] V Bach, E. Lieb, M. Loss and J.P. Solovej.: There are no unfilled shells in unrestricted Hartree-Fock theory Phys. Rev. Lett. 72, 2981 (1994) [13] V. Beaud and S. Warzel.: Low-energy Fock-space localization for attractive hard-core particles in disorder Ann. Henri Poincar´e (2017) [14] J.M Combes, P Hislop, F Klopp An Optimal Wegner estimate and its application to the global continuity of the integrated density of states for random Schr¨odinger operators Duke Mathematical Journal Vol 140 No 3, (2007). [15] Chulaevsky, V and Suhov, Y.: Anderson localisation for an interacting two-particle quantum system on Z, preprint, http://arxiv.org/abs/0705.0657) (2007) 97 [16] Chulaevsky, V and Suhov, Y.: Wegner bounds for a two-particle tight binding model Comm. Math. Phys. 283, 479-489, 2008. [17] Chulaevsky, V and Suhov, Y.: Eigenfunctions in a two-particle Anderson tight-binding model Comm. Math. Phys. July 2009, Volume 289, Issue 2, pp 701–723. [18] Chulaevsky, V. and Suhov, Y.: Multi-Scale Analysis for random quantum systems with interaction (2014). Progress in Mathematical Physics 65, Birkh¨auser/Springer, New York [19] Davies, E.B. Spectral Theory and Differential Operators Cambridge University Press (1995). [20] Davis, B. On the distributions of conjugate functions of nonnegative measures, Duke Math. J. 40 (1973), 695–700. [21] Davis, B. On the weak type (1, 1) inequality for conjugate functions, Proc. Amer. Math. Soc. 44 (1974), 307–311. [22] Del Rio R., Jitomirskaya S., Last Y, Simon, B. Operators with singular continuous spectrum IV Hausdorff dimension, rank one perturbations and localization [23] Ducatez, Raphael. Anderson Localization for infinitely many interacting particles in the Hartree-Fock theory Journal Of Spectral Theory. Volume 8, Issue 3, 2018, pp. 1019–105 [24] De Bruijn, N. and Erd¨os. Some linear and some quadratic recursion formulas, II Nederl. Akad. Wetensch. Proc. Ser. A. 55 (1952), 152–163 . [25] Elgart,A. Klein, A and Stolz, G. Many-body localization in the droplet spectrum of the random XXZ quantum spin chain (2018) Journal of Functional Analysis. Vol 275 Issue 1 ,211-258 [26] Elgart,A. Klein ,A. and Stolz G. Manifestations of dynamical localization in the disor- dered XXZ spin chain Elgart, A., Commun. Math. Phys. (2018) 361: 1083-1113 [27] Fr¨ohlich, J. and T. Spencer Absence of diffusion in the Anderson tight binding model for large disorder or low energy.(1983) Comm. Math. Phys 88(2), 151-184. [28] Germinet, F., Klein, A.Bootstrap Multiscale Analysis and Localization¶in Random Me- dia. Commun. Math. Phys. 222, 415–448 (2001). [29] P. Hislop, F Klopp, J. Schenker. Continuity with respect to disorder of the integrated density of states, Illinois Journal of Mathematics. Vol 49 Number 3 893-904. [30] E. Hamza, R. Sims and G. Stolz. Dynamical localization in disordered quantum spin systems, Comm. Math. Phys. 315 (2012) 215239 [31] C. Ha´ınzl, Mathiew Lewin and C. Sparber. Ground State Properties of Graphene in Hartree-Fock theory. Journal of Mathematical Physics 53, 095220 (2012) 98 [32] D. Gontier, C. Hainzl, and M. Lewin. Lower bound on the Hartree-Fock energy of the electron gas. Phys. Rev. A, 99:052501, 2019. [33] E. Lieb and B. Simon. The Hartree-Fock theory for Coulomb Systems Comm Math Phys. (1977) 53(2), 185-194. [34] R. Carmona, A. Klein, F. Martinelli. Anderson Localization for Bernoulli and other singular potentials (1987) Comm Math Phys (108) 41-66. [35] Guarnieri, I., Schulz-Baldes, H. Lower Bounds on Wave Packet Propagation By Packing Dimensions of Spectral Measures Mathematical Physics Electronic Journal Vol 5 (1999) [36] Hruˇscev,S. V. and Vinogradov,S. A. Free interpolation in the space of uniformly con- vergent Taylor series, Complex Analysis and Spectral Theory (Leningrad, 1979/1980), pp. 171–213, Lecture Notes in Math., Springer, Berlin–New York, 1981. [37] Jakˇsi´c, V., Molchanov, S. On the Surface Spectrum in Dimension Two. Helv. Phys. Acta 71, 629 (1999) [38] Jakˇsi´c, V., Molchanov, S. On the Spectrum of the Surface Maryland Model. Lett. Math. Phys. 45, 185 (1998) [39] Molchanov S., Jakˇsi´c. Localization of Surface Spectra Comm. Math. Physics 208, 153- 172 (1999). [40] S. Jitomirskaya, H. Schulz-Baldes Upper Bounds for Wavepacket spreading for random Jacobi matrices (2007) Comm Math Phys. 273(3), 601-618. [41] Klein, Abel. Multiscale Analysis and Localization of Random Operators arXiv:0708.2292 [42] Killip B. , Kiselev A. and Last, Y. Dynamical Upper Bounds for Wavepacket Spreading American Journal of Mathematics. Vol. 125, No. 5 (Oct., 2003), pp. 1165-1198 [43] Kunz, H. and B Souillard , Sur le spectre des operateurs aux differences finies aleatories (1981) Comm Math Phys. 78(2), 201-246. [44] Kurig, Carolin.: Random Lattice Models, D77 Dissertation Johannes Gutenberg Uni- versit¨at Mainz (2013) [45] L. H. Loomis A note on the Hilbert transform, Bull. Amer. Math. Soc. 52 (1946), 1082–1086. [46] Localization and IDS regularity for the Hubbard model within the Hartree-Fock Theory. arXiv:1906.10800, Matos, R. and Schenker, J. [47] Schenker, J. How large is large? Estimating the Critical Disorder for the Anderson Model (2015 ) Lett Math Phys 105, 1-9. [48] Schenker, J. H¨older equicontinuity of the density of states at weak disorder (2004) Lett. Math Physics 70, 195-209 99 [49] B Simon.: Trace Ideals and Their Applications, Mathematical Surveys and Monographs vol 120. [50] Mavi,R. and Schenker,J. Resonant Tunneling In A System With Correlated Pure Point Spectrum. J. Math. Phys. 60, 052103 (2019). [51] Mavi, R and Schenker, J. Localization in the Disordered Holstein model Commun Math Phys . 365 719-764 (2018) [52] Poltoratski,A. On the distributions of boundary values of Cauchy integrals, Proc. Amer. Math. Soc. 124 (1996), 2455–2463. [53] Poltoratski, A. Simon, B. and Zinchenko, M. The Hilbert Transform of a Measure Journal d’Analyse Mathematique, Vol 111 (2010), 247-265. [54] Simon,B. Cyclic vectors in the Anderson model Reviews in Mathematical Physics Vol. 06, No. 05a, pp. 1183-1185 (1994). [55] Simon,B. Harmonic Analysis:Comprehensive Course in Analysis,part 3 American Math- ematical Society (2015). [56] Simon, B. Operators With Singular Continuous Spectrum, VI. Graph Laplacians And Laplace-Beltrami Operators. Proceedings of the AMS, Vol 124, Num 4 (1996). [57] Stein,E. M. and Wiess, G. An extension of a theorem of Marcinkiewicz and some of its applications, J. Math. Mech. 8 (1959), 263–284. [58] Stolz, G.An introduction to the mathematics of Anderson Localization arXiv:1104.2317. 100