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

Magic angle (in)stability and mobility edges in disordered Chern insulators

Simon Becker [email protected] ETH Zurich, Institute for Mathematical Research, Raemistrasse 101, 8092 Zurich, Switzerland Izak Oltman [email protected] University of California, Department of Mathematics, Berkeley, CA 94720, USA  and  Martin Vogel [email protected] Université de Strasbourg, Institut de Recherche Mathématique Avancée, 67084 Strasbourg Cedex, France
Abstract.

Why do experiments only exhibit one magic angle if the chiral limit of the Bistritzer-MacDonald Hamiltonian suggest a plethora of them? - In this article, we investigate the remarkable stability of the first magic angle in contrast to higher (smaller) magic angles. More precisely, we examine the influence of disorder on magic angles and the Bistritzer-MacDonald Hamiltonian. We establish the existence of a mobility edge near the energy of the flat band for small disorder. We also show that the mobility edges persist even when all global Chern numbers become zero, leveraging the C2zTC_{2z}T symmetry of the system to demonstrate non-trivial sublattice transport. This effect is robust even beyond the chiral limit and in the vicinity of perfect magic angles, as is expected from experiments.

1. Introduction

Twisted bilayer graphene is a highly tunable material that exhibits approximately flat bands at special twisting angles, the so-called magic angles [TKV19, BiMa11].

In this article, we study the question of why the largest magic angle, as predicted by the chiral model, is more robust than smaller magic angles within the chiral limit. We analyze why the chiral model’s largest magic angle exhibits greater resilience compared to smaller magic angles within the chiral limit. This finding potentially elucidates why, up until now, only the first magic angle has been experimentally observed. We also study the impact of disorder in the chiral limit and its interplay with the flat bands at magic angles. In quantum systems, disorder-induced dynamical localization is a well-known phenomenon, wherein spatially localized wavepackets do not significantly diffuse under time evolution. While the underlying mechanisms behind this phenomenon are relatively well understood, the opposite behavior—diffusive behavior in disordered systems—is only understood in specific cases [AS19, AW13, BH22, GKS07, JSS03].

It is widely believed that large classes of two-dimensional quantum systems, even under minor disorder, exclusively exhibit localization, as conjectured by Problem 2 on Simon’s list of open problems for Schrödinger operators [Si00]. As we will show below, the Hamiltonian describing twisted bilayer graphene at a magic angle or other related materials is an exception. These new classes of materials, so-called Chern insulators, exhibit non-zero Chern numbers in the absence of external magnetic fields [Li21]. In particular, in twisted bilayer graphene, for wavepackets localized sufficiently close to the perturbed flat band at zero energy, the time-evolution is, in a suitable sense, ballistic. Our argument here is an adaptation of an argument by Germinet, Klein, and Schenker showing a form of delocalization for the Landau Hamiltonian [GKS07]. The physical intuition behind this delocalization argument is straightforward: the Landau Hamiltonian exhibits non-zero Hall conductivity at each Landau level. Moreover, as the Hall conductivity, a topological quantity, remains invariant under minor disorder, the existence of substantial spectral gaps between the Landau levels prevents strong localization across the spectrum.

The properties of the flat bands for twisted bilayer graphene are analogous to Landau levels with the crucial difference that no magnetic field is required. At the first magic angle, the two flat bands exhibit only a Chern number zero. However, the two flat bands individually carry non-zero Chern numbers ±1\pm 1, allowing for an anomalous quantum Hall effect when the TBG substrate is e.g. aligned with hexagonal boron nitride [Li21]. Mathematically, the effect of aligning the substrate with hBN is modeled by adding an effective mass term to the Hamiltonian thereby splitting the two flat bands each carrying a non-zero Chern number. In addition, the flat bands are gapped from the rest of the spectrum.

We also establish a localized regime that rests on the multi-scale analysis framework of Germinet-Klein [GK01, GK03]. Here, the only difficulty is to allow for a sufficiently large class of random perturbations which requires us to extend the estimate on the number of eigenvalues (NE) and thus the Wegner estimate (W).

The chiral limit of the massive continuum model for twisted bilayer graphene, which can also be thought of as a model Hamiltonian for twisted transition metal dichalcogenides (TMDs) [CRQ23], is the Hamiltonian H(m,α)H(m,\alpha) acting on L2(;4)L^{2}({\mathbb{C}};{\mathbb{C}}^{4}) with domain given by the Sobolev space H1(;4)H^{1}({\mathbb{C}};{\mathbb{C}}^{4})

H(m,α)=(mD(α)D(α)m) with D(α)=(2Dz¯αU(z)αU(z)2Dz¯),H(m,\alpha)=\begin{pmatrix}m&D(\alpha)^{*}\\ D(\alpha)&-m\end{pmatrix}\text{ with }D(\alpha)=\begin{pmatrix}2D_{\bar{z}}&\alpha U(z)\\ \alpha U(-z)&2D_{\bar{z}}\end{pmatrix}, (1.1)

where Dz¯=iz¯D_{\bar{z}}=-i\partial_{\bar{z}} , α{0}\alpha\in\mathbb{C}\setminus\{0\} is an effective parameter that is inversely proportional to the twisting angle and m0m\geq 0 a mass parameter. Let Γ:=4πiω(ω)\Gamma:=4\pi i\omega({\mathbb{Z}}\oplus\omega{\mathbb{Z}}) be a triangular lattice with ω=e2πi/3\omega=e^{2\pi i/3}. The tunnelling potentials UU are Γ\Gamma-periodic functions satisfying for 𝐚=4πia1ω/3+4πia2ω2/3\mathbf{a}=4\pi ia_{1}\omega/3+4\pi ia_{2}\omega^{2}/3 with aia_{i}\in{\mathbb{Z}}, i.e. 𝐚Γ3:=Γ/3\mathbf{a}\in\Gamma_{3}:=\Gamma/3

U(z+𝐚)=ω¯a1+a2U(z),U(ωz)=ωU(z),U(z)¯=U(z¯).U(z+\mathbf{a})=\bar{\omega}^{a_{1}+a_{2}}U(z),\quad U(\omega z)=\omega U(z),\quad\overline{U(z)}=U(\bar{z}). (1.2)

The central object in the one-particle picture of twisted bilayer graphene are the so-called magic angles. We say that α{0}\alpha\in\mathbb{C}\setminus\{0\} is magic if and only if the Bloch-Floquet transformed Hamiltonian, see [Be*22, (2.11)], with mass parameter m0m\geq 0 exhibits a flat band at energy ±m\pm m, i.e.

±mkSpecL2(/Γ)(Hk(m,α)) with Hk(m,α)=(mD(α)+k¯D(α)+km),\pm m\in\bigcap_{k\in{\mathbb{C}}}\operatorname{Spec}_{L^{2}({\mathbb{C}}/\Gamma)}(H_{k}(m,\alpha))\text{ with }H_{k}(m,\alpha)=\begin{pmatrix}m&D(\alpha)^{*}+\bar{k}\\ D(\alpha)+k&-m\end{pmatrix}, (1.3)

with SpecX(S)\operatorname{Spec}_{X}(S) denoting the spectrum of the linear operator SS on the Hilbert space XX on a suitable dense domain, where Hk(m,α):H1(/Γ;2)L2(/Γ;2).H_{k}(m,\alpha):H^{1}({\mathbb{C}}/\Gamma;{\mathbb{C}}^{2})\to L^{2}({\mathbb{C}}/\Gamma;{\mathbb{C}}^{2}). The set of α\alpha under which there exists a flat band at energy ±m\pm m is independent of mm111Observe that SpecHk(m,α)=±SpecHk(0,α)2+m2\operatorname{Spec}H_{k}(m,\alpha)=\pm\sqrt{\operatorname{Spec}H_{k}(0,\alpha)^{2}+m^{2}} [T92, (5.66)].. In the sequel, we shall suppress the mass parameter m0m\geq 0 in the notation.

For the study of magic angles we also introduce a translation operator

𝐚w(z):=(ωa1+a2001)w(z+𝐚),𝐚Γ3,\mathscr{L}_{\mathbf{a}}w(z):=\begin{pmatrix}\omega^{a_{1}+a_{2}}&0\\ 0&1\end{pmatrix}w(z+\mathbf{a}),\ \ \ \mathbf{a}\in\Gamma_{3}, (1.4)

and a rotation operator 𝒞u(z)=u(ωz).\mathscr{C}u(z)=u(\omega z). We can then define subspaces

Lk,p2:={uL2(/Γ;2);𝐚u(z)=ωku(z) and 𝒞u(z)=ω¯pu(z)}L^{2}_{k,p}:=\{u\in L^{2}({\mathbb{C}}/\Gamma;{\mathbb{C}}^{2});\mathscr{L}_{\mathbf{a}}u(z)=\omega^{k}u(z)\text{ and }\mathscr{C}u(z)=\bar{\omega}^{p}u(z)\} (1.5)

and similarly Lk2:=p3Lk,p2L^{2}_{k}:=\bigoplus_{p\in{\mathbb{Z}}_{3}}L^{2}_{k,p}.

The set of such magic parameters α\alpha, that we shall denote by 𝒜,\mathcal{A}, is characterized by [Be*22, Theo.22]

α1SpecL02(Tk) with Tk=(2Dz¯k)1(0U(z)U(z)0) for some kΓ\alpha^{-1}\in\operatorname{Spec}_{L^{2}_{0}}(T_{k})\text{ with }T_{k}=(2D_{\bar{z}}-k)^{-1}\begin{pmatrix}0&U(z)\\ U(-z)&0\end{pmatrix}\text{ for some }k\notin\Gamma^{*} (1.6)

with Γ\Gamma^{*} the dual lattice.

We then define the set of generic magic angles. The terminology generic is motivated by [BHZ23, Theo.33] which shows that for a generic choice of tunnelling potentials all magic angles are of the following form.

Definition 1.1 (Generic magic angles).

We say that α𝒜\alpha\in\mathcal{A} is a simple or two-fold degenerate magic angle if 1/αSpecL02(Tk)1/\alpha\in\operatorname{Spec}_{L^{2}_{0}}(T_{k}) and dimkerL02(Tk1/α)=ν\dim\ker_{L^{2}_{0}}(T_{k}-1/\alpha)=\nu with ν=1,2\nu=1,2, respectively. In the following sections, we will denote the combination of these magic angles as the collection of generic magic angles.

1.1. Magic angle (in)stability

The first aim of this article is to study perturbations of the operator TkT_{k}, i.e. for δ>0\delta>0

Tk,δ=(2Dz¯k)1(0U(z)+δV+U(z)+δV0)T_{k,\delta}=(2D_{\bar{z}}-k)^{-1}\begin{pmatrix}0&U(z)+\delta V_{+}\\ U(-z)+\delta V_{-}&0\end{pmatrix}

with bounded linear perturbations V±V_{\pm}. We then obtain in Theorem 6 a bound on the spread of the magic angles under such perturbations. This is a non-trivial result as the operator TkT_{k}, whose eigenvalues are the magic angles, is a non-normal operator. In particular, small perturbations in norm can lead to substantial perturbations of the spectrum, see [TE05]. On the other hand, we show that even simple rank 11-perturbations of exponentially small size in 1/|μ|1/|\mu| suffice to generate eigenvalues μ\mu in the spectrum of TkT_{k}.

Theorem 1 (Instability).

Let μ\mu\in\mathbb{C} and kΓk\notin\Gamma^{*}, then there exists a rank-11 operator RR with R=𝒪(ec/|μ|)\|R\|=\mathcal{O}(e^{-c/|\mu|}) and c(k)>0c(k)>0 independent of μ\mu such that μSpec(Tk+R)\mu\in\operatorname{Spec}(T_{k}+R). Here, TkT_{k} characterizes the set of magic parameters as explained in (1.6).

1.2. Anderson model and IDS

One consequence of having a flat band is the occurrence of jump discontinuities in the integrated density of states (IDS). The integrated density of states is defined as follows, see [Sj89] and others:

Definition 1.2.

The integrated density of states (IDS) for energies E2>E1E_{2}>E_{1} and I=[E1,E2]I=[E_{1},E_{2}] is defined by

N(I):=limLtr(1lI(HΛL(α)))L2N(I):=\lim_{L\to\infty}\frac{\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{I}(H_{\Lambda_{L}}(\alpha)))}{L^{2}}

with ΛL=/(LΓ)\Lambda_{L}={\mathbb{C}}/(L\Gamma) and HΛLH_{\Lambda_{L}} has periodic boundary conditions, i.e. HΛL:H1(ΛL)L2(HΛL)L2(HΛL).H_{\Lambda_{L}}:H^{1}(\Lambda_{L})\subset L^{2}(H_{\Lambda_{L}})\to L^{2}(H_{\Lambda_{L}}).

For ergodic random operators, the almost sure existence of this limit is shown using the subadditive ergodic theorem, see for instance [K89, Sec. 7.3]. Alternatively, one may define for fCc()f\in C_{c}^{\infty}({\mathbb{R}}) the regularized trace

tr~(f(H(α)))=limLtr(1lΛLf(H(α)))|ΛL|.\widetilde{\operatorname{tr}}(f(H(\alpha)))=\lim_{L\to\infty}\frac{\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{{\Lambda_{L}}}f(H(\alpha)))}{|\Lambda_{L}|}. (1.7)

By Riesz’s theorem on the representation of positive functionals, one has that

tr~(f(H(α)))=f(λ)𝑑ρ(λ),\widetilde{\operatorname{tr}}(f(H(\alpha)))=\int_{{\mathbb{R}}}f(\lambda)\ d\rho(\lambda),

where ρ\rho is the density of states (DOS) measure of H(α).H(\alpha). This way, N(I)=I𝑑ρ(λ).N(I)=\int_{I}\ d\rho(\lambda).

Remark 1.

For Schrödinger operators it is common to consider Dirichlet approximations of the finite-size truncation. It is known that Dirac operators do in general not have any self-adjoint Dirichlet realizations. However, self-adjoint Neumann-type boundary conditions are possible, see [BM87] and for instance the introduction of [SV19] for a mathematical discussion. The independence of the definition of the IDS of the boundary conditions can then be shown using spectral shift function techniques if the operator contains a gap in the spectrum, see for instance the work by Nakamura [N01] on Schrödinger operators.

A periodic Hamiltonian that exhibits a flat band at energy EE possesses a jump discontinuity in the IDS at EE. In particular, the Lebesgue decomposition of ρ\rho has a pure point contribution at E.E. As a consequence, if we define the associated cumulative distribution function NE0:(E0,)N_{E_{0}}:(E_{0},\infty)\to\mathbb{R} by NE0(E):=N([E0,E])N_{E_{0}}(E):=N([E_{0},E]), then this function is monotonically increasing and right-continuous (càdlàg). At a magic angle, the function NE0N_{E_{0}} for E0<±mE_{0}<\pm m exhibits a jump discontinuity at E=±m.E=\pm m. Indeed, this can easily be seen from the following formula which shows that for a periodic Hamiltonian one just has [Sj89, (1.29)]

N(I)=/Γ(kλSpecL2(/Γ)(Hk(α))1lI(λ))dk4π2.N(I)=\int_{\mathbb{C}/\Gamma^{*}}\Bigg{(}k\mapsto\sum_{\lambda\in\operatorname{Spec}_{L^{2}({\mathbb{C}}/\Gamma)}(H_{k}(\alpha))}\operatorname{1\hskip-2.75ptl}_{I}(\lambda)\Bigg{)}\ \frac{dk}{4\pi^{2}}.

Let α𝒜\alpha\in\mathcal{A} be a generic magic angle, as in Def. 1.1, then we define the energy gap between the flat bands and the rest of the spectrum

Egap:=infλSpec(H(α)2){0}λ>0.E_{\operatorname{gap}}:=\inf_{\lambda\in\operatorname{Spec}(H(\alpha)^{2})\setminus\{0\}}\sqrt{\lambda}>0. (1.8)

That this quantity is non-zero by [BHZ22, Theo.22] for simple and by [BHZ23, Theo.44] for two-fold degenerate magic angles and is illustrated in Figure 1. In particular, the following union of intervals is in the resolvent set of the Hamiltonians

(Egap2+m2,m)(m,m)(m,Egap2+m2)Spec(H).\Big{(}-\sqrt{E_{\operatorname{gap}}^{2}+m^{2}},-m\Big{)}\cup\big{(}-m,m\big{)}\cup\Big{(}m,\sqrt{E_{\operatorname{gap}}^{2}+m^{2}}\Big{)}\subset\mathbb{R}\setminus\operatorname{Spec}(H). (1.9)

Let PXP_{X} be the orthogonal projection onto a closed subspace XX. For α𝒜\alpha\in\mathcal{A} generic it has been shown in [BHZ22b, Theo. 44] and [BHZ23, Theo. 55] that the Chern number of the flat band at energy ±m\pm m is 1\mp 1 or more generally (including m=0m=0) for the Hamiltonian in (1.1)

Cher(Pker(D(α)))=1 and Cher(Pker(D(α)))=1.\operatorname{Cher}(P_{\ker(D(\alpha))})=-1\text{ and }\operatorname{Cher}(P_{\ker(D(\alpha)^{*})})=1. (1.10)

The Chern number can be computed from the Hall conductivity Ω(P)\Omega(P), see (4.4), by using that

Cher(P)=2πiΩ(P).\operatorname{Cher}(P)=-2\pi i\Omega(P).

In particular, the net Chern number of the flat bands for m=0m=0 is zero

Cher(Pker(D(α))Pker(D(α)))=Cher(Pker(H(α)))=0.\operatorname{Cher}(P_{\ker(D(\alpha))}\oplus P_{\ker(D(\alpha)^{*})})=\operatorname{Cher}(P_{\ker(H(\alpha))})=0.
Refer to captionRefer to captionFlat band (Chern number +1)Flat band (Chern number -1)Flat band (Chern number +1)Flat band (Chern number -1)EEEEEgapE_{\text{gap}}EgapE_{\text{gap}}mmmmEgap2+m2\sqrt{E_{\text{gap}}^{2}+m^{2}}Egap2+m2\sqrt{E_{\text{gap}}^{2}+m^{2}}k1k_{1}k2k_{2}k1k_{1}k2k_{2}
Figure 1. Band structure of non-disordered twisted bilayer graphene (1.1) at the first real positive magic angle α0.58566\alpha\approx 0.58566 with zero effective mass (top) and non-zero effective mass (bottom).
Assumption 1 (Anderson model).

We introduce the Anderson-type Hamiltonian with alloy-type potentials and (possible) lattice relaxation effects for λ>0\lambda>0 and uCc(;4).u\in C^{\infty}_{c}({\mathbb{C}};{\mathbb{C}}^{4}).

Hλ=H+λVω where Vω=γΓωγu(γξγ),H_{\lambda}=H+\lambda V_{\omega}\text{ where }V_{\omega}=\sum_{\gamma\in\Gamma}\omega_{\gamma}u(\bullet-\gamma-\xi_{\gamma}), (1.11)

where (ωγ)γ(\omega_{\gamma})_{\gamma} and (ξγ)γ(\xi_{\gamma})_{\gamma} are families of i.i.d. random variables with absolutely continuous bounded densities. For (ωγ)(\omega_{\gamma}) the density is a function gg with supp(g)[1,1]\operatorname{supp}(g)\subset[-1,1] and in case of (ξγ)γ(\xi_{\gamma})_{\gamma} a density hh supported within a compact domain DD\subset{\mathbb{C}} where we allow ξconst\xi\equiv\operatorname{const}. Random variables ξγ\xi_{\gamma} model small inhomogeneities of the moiré lattice due to relaxation effects. We assume that either

  1. (1)

    Case 11: The disorder uu in (1.11) is of the form

    u(z)=(Y(z)Z(z)Z(z)Y(z))Cc(;4)u(z)=\begin{pmatrix}Y(z)&Z(z)^{*}\\ Z(z)&-Y(z)\end{pmatrix}\in C^{\infty}_{c}({\mathbb{C}};{\mathbb{C}}^{4}) (1.12)

    where infξDΓinfzγΓY(zγξγ)>0.\inf_{\xi\in D^{\Gamma}}\inf_{z\in{\mathbb{C}}}\sum_{\gamma\in\Gamma}Y(z-\gamma-\xi_{\gamma})>0.

  2. (2)

    Case 22: The disorder uCc(;4)u\in C^{\infty}_{c}({\mathbb{C}};{\mathbb{C}}^{4}) with u0u\geq 0 and infzBε(z0),γΓ,ξDu(zγξ)>0\inf_{z\in B_{\varepsilon}(z_{0}),\gamma\in\Gamma,\xi\in D}u(z-\gamma-\xi)>0 for some Bε(z0)B_{\varepsilon}(z_{0}).

For normalization purposes, we assume that supξDΓγΓu(γξγ)1\sup_{\xi\in D^{\Gamma}}\|\sum_{\gamma\in\Gamma}u(\bullet-\gamma-\xi_{\gamma})\|_{\infty}\leq 1 and suppuΛR(0)\operatorname{supp}u\subset\Lambda_{R}(0) for some fixed R>0R>0 where ΛL:=/(LΓ)\Lambda_{L}:={\mathbb{C}}/(L\Gamma) and ΛL(z)=ΛL+z.\Lambda_{L}(z)=\Lambda_{L}+z.

We emphasize that under assumption (1), the matrix uu is neither positive nor negative definite. This usually poses an obstruction to proving Wegner estimates as the eigenvalues are not monotone in the noise parameter. We can overcome this obstacle here by using the off-diagonal structure of the Hamiltonian. The probability space is the Polish space Ω=(supp(g))Γ×(supp(h))Γ\Omega=(\operatorname{supp}(g))^{\Gamma}\times(\operatorname{supp}(h))^{\Gamma} with the product measure. Then (Hλ)(H_{\lambda}) is an ergodic (with respect to lattice translations) family of self-adjoint operators with continuous dependence Ω(ω,ξ)(Hλ+i)1.\Omega\ni(\omega,\xi)\mapsto(H_{\lambda}+i)^{-1}. Thus, there is Σ\Sigma\subset\mathbb{R} closed such that

SpecL2()(Hλ)=Σ a.s.,\operatorname{Spec}_{L^{2}({\mathbb{C}})}(H_{\lambda})=\Sigma\text{ a.s.}, (1.13)

see [CFKS87, Pa80]. In addition, using ergodicity arguments, see e.g. [W95], the density of states measure for the random operator, ρHλ\rho^{H_{\lambda}} exists almost surely and is almost surely non-random, i.e. is almost surely equal to a non-random measure ρHλ=ρ\rho^{H_{\lambda}}=\rho a.s. with ρ\rho non-random. An extension of our work to unbounded disorder is possible. In the context of Schrödinger operators this extension has been shown for magnetic Landau Hamiltonians [GKS09, GKM09]. Related proofs of localization for Dirac operators under a spectral gap assumption have also been obtained in [BCZ19].

For λ0\lambda\neq 0, the infinitely-degenerate point spectrum of HH at energy zero, i.e. the flat band, gets non-trivially perturbed and expands in energy. To capture this, we then introduce constants K±:=Egap2+m2±λsupωΩVωK_{\pm}:=\sqrt{E_{\operatorname{gap}}^{2}+m^{2}}\pm\lambda\sup_{\omega\in\Omega}\|V_{\omega}\|_{\infty} and k±:=m±|λ|supωΩVω.k_{\pm}:=m\pm|\lambda|\sup_{\omega\in\Omega}\|V_{\omega}\|_{\infty}. One thus finds analogously to (1.9) for the disordered Hamiltonian

(K,k+)(k,k)(k+,K)Σ,(-K_{-},-k_{+})\cup(-k_{-},k_{-})\cup(k_{+},K_{-})\subset\mathbb{R}\setminus\Sigma, (1.14)

where all three intervals are non-trivial for λ>0\lambda>0 sufficiently small and m>0m>0. We then also define

J:=[k+,k] and J+:=[k,k+].J_{-}:=[-k_{+},-k_{-}]\text{ and }J_{+}:=[k_{-},k_{+}]. (1.15)

When perturbing away from perfect magic angles, we may do so by either using a random potential or perturbing α\alpha slightly. In both cases, for sufficiently small perturbations, this leaves the spectral gap to the remaining bands open.

Given a finite domain ΛL:=/(LΓ)\Lambda_{L}:={\mathbb{C}}/(L\Gamma)\subset{\mathbb{C}}, we introduce the Hamiltonian

Hλ,ΛL=HΛL+λVω,ΛL,H_{\lambda,\Lambda_{L}}=H_{\Lambda_{L}}+\lambda V_{\omega,\Lambda_{L}},

with periodic boundary conditions where γΛ~Lωγ(u1lΛL)(γξγ) with Λ~L:=ΛLΓ\sum_{\gamma\in\tilde{\Lambda}_{L}}\omega_{\gamma}(u\operatorname{1\hskip-2.75ptl}_{\Lambda_{L}})(\bullet-\gamma-\xi_{\gamma})\text{ with }\widetilde{\Lambda}_{L}:=\Lambda_{L}\cap\Gamma. In general we shall denote by SΛLS_{\Lambda_{L}} the restriction of an operator SS to the domain ΛL\Lambda_{L} with periodic boundary conditions in case that SS is a differential operator.

While the occurrence of a flat band for the unperturbed Hamiltonian (1.1) leads to a jump discontinuity in the IDS, one has that the random Hamiltonian (1.11) has a Lipschitz continuous IDS for all λ0\lambda\neq 0. Since the randomly perturbed Hamiltonian is no longer periodic, it is customary to measure the destruction of the flat band by studying the regularity of the IDS.

Theorem 2 (Continuous IDS).

Consider the Anderson Hamiltonian as in Assumption (1) with m0m\geq 0 and coupling constant λ(ε(m),ε(m)){0}\lambda\in(-\varepsilon(m),\varepsilon(m))\setminus\{0\} with ε(m)>0\varepsilon(m)>0 sufficiently small, then the integrated density of states (IDS) is a.s. Hölder continuous in Hausdorff distance dHd_{H} for all β(0,1)\beta\in(0,1), i.e.

|N(I)N(I)|β,I,IdH(I,I)β,|N(I)-N(I^{\prime})|\lesssim_{\beta,I,I^{\prime}}d_{\text{H}}(I,I^{\prime})^{\beta},
  • Case 1 disorder: for intervals I,I[k+,k+]I,I^{\prime}\subset[-k_{+},k_{+}] with m>0m>0 and

  • Case 2 disorder: arbitrary bounded intervals I,II,I^{\prime}\Subset{\mathbb{R}}, λ{0}\lambda\in{\mathbb{R}}\setminus\{0\} and m0.m\geq 0. If we assume in addition that uu is globally positive, i.e.

    infξDΓinfzγΓu(zγξγ)>0,\inf_{\xi\in D^{\Gamma}}\inf_{z\in{\mathbb{C}}}\sum_{\gamma\in\Gamma}u(z-\gamma-\xi_{\gamma})>0, (1.16)

    then the IDS is a.s. Lipschitz continuous

    |N(I)N(I)|I,IdH(I,I).|N(I)-N(I^{\prime})|\lesssim_{I,I^{\prime}}d_{\text{H}}(I,I^{\prime}).

    In particular, the IDS is a.s. differentiable and its Radon-Nikodym derivative, the density of states (DOS), exists a.s. and is a.s. bounded.

The above results follow directly from the following estimate on the number of eigenvalues (NE) that directly lead to Wegner estimates (4.2).

Proposition 1.3 (NE).

Under the assumptions of Theorem 2, we have that there is β(0,1)\beta\in(0,1)

𝐄tr(1lI(Hλ,ΛL))β|I|β|ΛL|.\mathbf{E}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{I}(H_{\lambda,\Lambda_{L}}))\lesssim_{\beta}|I|^{\beta}|\Lambda_{L}|.

If in Case 2 we assume in addition that (1.16) holds, then we may take β=1\beta=1

𝐄tr(1lI(Hλ,ΛL))|I||ΛL|.\mathbf{E}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{I}(H_{\lambda,\Lambda_{L}}))\lesssim|I||\Lambda_{L}|.

1.3. Mobility edges

In the works of Germinet–Klein [GK01, GK03, GK04] dynamical measures of transport have been introduced. The dynamical localization implies a strong form of decaying eigenfunctions, see Def. 4.1. To measure dynamical localization/delocalization one introduces the following Hilbert-Schmidt norm

Mλ(p,χ,t)=p/2eitHλχ(Hλ)1l/Γ322,M_{\lambda}(p,\chi,t)=\left\lVert\langle\bullet\rangle^{p/2}e^{-itH_{\lambda}}\chi(H_{\lambda})\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}, (1.17)

where Γ3:=Γ/3\Gamma_{3}:=\Gamma/3, for some non-negative χCc\chi\in C_{c}^{\infty} with time average

λ(p,χ,T)=1T0𝐄(Mλ(p,χ,t))et/T𝑑t.\mathcal{M}_{\lambda}(p,\chi,T)=\frac{1}{T}\int_{0}^{\infty}\mathbf{E}\Big{(}M_{\lambda}(p,\chi,t)\Big{)}e^{-t/T}\ dt.

Recall that 1T0tnet/T𝑑t=TnΓ(n+1),\frac{1}{T}\int_{0}^{\infty}t^{n}e^{-t/T}\ dt=T^{n}\Gamma(n+1), to see that λ(p,χ,T)\mathcal{M}_{\lambda}(p,\chi,T) indicates a time-averaged power scaling of Mλ(p,χ,t).M_{\lambda}(p,\chi,t). Here, Mλ(p,χ,t)M_{\lambda}(p,\chi,t) measures the spread of mass in a spectral energy window of the Hamiltonian from the origin under the free Schrödinger evolution.

We shall then show that the random Hamiltonian (1.11) exhibits diffusive behavior in the vicinity of magic angles.

Theorem 3 (Dynamical delocalization).

Let α\alpha_{*} be a generic magic angles as in Definition 1.1. We consider a coupling constant λ(ε(m,α),ε(m,α))\lambda\in(-\varepsilon(m,\alpha_{*}),\varepsilon(m,\alpha_{*})), α(αδ(m,α),α+δ(m,α))\alpha\in(\alpha_{*}-\delta(m,\alpha_{*}),\alpha_{*}+\delta(m,\alpha_{*})), mass m0m\geq 0 and sufficiently small ε(m,α),δ(m,α)>0\varepsilon(m,\alpha_{*}),\delta(m,\alpha_{*})>0. The random Hamiltonian HλH_{\lambda} exhibits diffusive behavior for m>0m>0 at at least two energies E±(λ)E_{\pm}(\lambda) located close to energies ±m\pm m, respectively, and at at least one energy E(λ)E(\lambda) for m=0m=0. In particular, for every χCc\chi\in C_{c}^{\infty} that equals to one on an open interval JJ containing at least one of E±(λ)E_{\pm}(\lambda) and p>0p>0 we have for all T>0T>0

λ(p,χ,T)p,JTp46.\mathcal{M}_{\lambda}(p,\chi,T)\gtrsim_{p,J}T^{\frac{p}{4}-6}.

We do not have a very precise understanding how close E±(λ)E_{\pm}(\lambda) are to ±m\pm m. By choosing suitable disorder (of fixed support but rescaled probability), one can show that E±E_{\pm} and ±m\pm m can get arbitrarily close, see Theorem 7, at least when α𝒜\alpha\in\mathcal{A} is a magic angle, as the bands of the unperturbed Hamiltonian are perfectly flat.

Remark 2.

Transport behavior can also be characterized by the pp-dependence of the estimate in the previous theorem and using a local transport exponent

βλ(E)=supp>0infIEI open supχCc(I;[0,))lim infTlog+λ(p,χ,T)plog(T).\beta_{\lambda}(E)=\sup_{p>0}\inf_{\begin{subarray}{c}I\ni E\\ I\text{ open }\end{subarray}}\sup_{\chi\in C_{c}^{\infty}(I;[0,\infty))}\liminf_{T\to\infty}\frac{\log_{+}\mathcal{M}_{\lambda}(p,\chi,T)}{p\log(T)}.

The region of dynamical localization is then defined as the open set

ΣDL:={E;βλ(E)=0}\Sigma^{\operatorname{DL}}:=\{E\in{\mathbb{R}};\beta_{\lambda}(E)=0\} (1.18)

whereas the region of dynamical delocalization ΣDD\Sigma^{\operatorname{DD}} is defined as its complement. A mobility edge is an energy EΣDDΣDLΣ¯.E\in\Sigma^{\operatorname{DD}}\cap\overline{\Sigma^{\operatorname{DL}}\cap\Sigma}. It follows from [GK04, Theo. 2.102.10, 2.112.11] that Theorem 3 implies βλ(E)>1/4.\beta_{\lambda}(E)>1/4. Theorem 7 then proves the existence of mobility edges for the disordered Hamiltonian.

While Theorem 3 describes the dynamical features of the Hamiltonian, it is equally valid to ask for a spectral theoretic interpretation of transport and localization. The interpretation of the nature of the spectrum in the dynamically localized phase is captured by the concept of SUDEC, see Def. 4.1.

The existence of a dynamical delocalization, in the above sense, does not imply the existence of a.c. or s.c. spectrum. Given that at magic angles the Hamiltonian H0(α)H_{0}(\alpha) only exhibits (infinitely degenerate) point spectrum at energies ±m\pm m, it is unknown if such phases can occur for our disordered Hamiltonian in a neighborhood of the flat bands. We conjecture that this is not the case.

As we will explain below, see Remark 3, the point spectrum of the Hamiltonian within an energy window, cannot be too localized.

This can be made precise within the framework of generalized Wannier functions [CMM19, MMP]:

Definition 1.4 (Wannier basis).

Let PP be an orthogonal projection onto L2()L^{2}({\mathbb{C}}). We say an orthonormal basis (ψβ)βIL2()(\psi_{\beta})_{\beta\in I}\in L^{2}({\mathbb{C}}) for an index set II\subset\mathbb{N} is an ss-localized generalized Wannier basis for PP for some s>0s>0 if:

  • span¯(ψβ)=ran(P).\overline{\operatorname{span}}(\psi_{\beta})=\operatorname{ran}(P).

  • There exists M<M<\infty and a collection of localization centers (μβ)(\mu_{\beta})\subset{\mathbb{C}} such that for all βI\beta\in I

    zμβ2s|ψβ(z)|2𝑑λ(z)M, with λ Lebesgue measure.\int_{{\mathbb{C}}}\langle z-\mu_{\beta}\rangle^{2s}|\psi_{\beta}(z)|^{2}d\lambda(z)\leq M,\text{ with }\lambda\text{ Lebesgue measure.}

One then has for the random Hamiltonian Hλ:H_{\lambda}:

Theorem 4 (Slow decay; m>0m>0).

Under the assumptions of Theorem 3, we define the orthogonal projection Pλ:=1lJ±(Hλ)P_{\lambda}:=\operatorname{1\hskip-2.75ptl}_{J_{\pm}}(H_{\lambda}) on L2()L^{2}({\mathbb{C}}) with J±J_{\pm} as in (1.15) for m>0.m>0. For any δ>0\delta>0 and for any λ(ε(m),ε(m))\lambda\in(-\varepsilon(m),\varepsilon(m)) with ε(m)>0\varepsilon(m)>0 sufficiently small and independent of δ>0\delta>0, PλP_{\lambda} does not admit a 1+δ1+\delta-localized generalized Wannier basis.

However, the projection admits a 1δ1-\delta-localized generalized Wannier basis for small disorder.

In this article, we have not considered disorder that only perturbs the off-diagonal entries of the Hamiltonian (1.1), since no techniques to show Wegner estimates for such disorder are known on which the multi-scale analysis rests.

Wegner estimates are however not needed to study the decay of Wannier functions and thus we shall consider such perturbations now, by looking at the Hamiltonian

Hλ=(m(D(α)+λW)D(α)+λWm)H_{\lambda}=\begin{pmatrix}m&(D(\alpha)+\lambda W)^{*}\\ D(\alpha)+\lambda W&-m\end{pmatrix} (1.19)

where WL(;2×2)W\in L^{\infty}({\mathbb{C}};{\mathbb{C}}^{2\times 2}) is a (possibly random) potential which we assume without loss of generality to satisfy W1.\|W\|_{\infty}\leq 1. The result of Theorem 4 cannot be directly extended to m=0m=0, since the net Chern number of the Hamiltonian is zero. However, the square of the Hamiltonian (1.19) exhibits a diagonal form

Hλ2=diag((D(α)+λW)(D(α)+λW)+m2,(D(α)+λW)(D(α)+λW)+m2).H_{\lambda}^{2}=\operatorname{diag}((D(\alpha)+\lambda W)^{*}(D(\alpha)+\lambda W)+m^{2},(D(\alpha)+\lambda W)(D(\alpha)+\lambda W)^{*}+m^{2}). (1.20)

Thus, to capture the low-lying spectrum, we may study the projections

P+,λ:=1l[0,μ]((D(α)+λW)(D(α)+λW)) and P,λ:=1l[0,μ]((D(α)+λW)(D(α)+λW)),\begin{split}P_{+,\lambda}&:=\operatorname{1\hskip-2.75ptl}_{[0,\mu]}((D(\alpha)+\lambda W)^{*}(D(\alpha)+\lambda W))\text{ and }\\ P_{-,\lambda}&:=\operatorname{1\hskip-2.75ptl}_{[0,\mu]}((D(\alpha)+\lambda W)(D(\alpha)+\lambda W)^{*}),\end{split} (1.21)

separately, where we dropped the m0m\geq 0, dependence as it does not affect the spectrum apart from a constant shift. We then have

Theorem 5 (Slow decay; m0m\geq 0).

Let μ<Egap2/2\mu<E_{\text{gap}}^{2}/2 with EgapE_{\text{gap}} as in (1.8) and P±,λP_{\pm,\lambda} be as in (1.21). For any δ>0\delta>0 and for any λ(ε,ε)\lambda\in(-\varepsilon,\varepsilon) with ε>0\varepsilon>0 sufficiently small and independent of δ>0\delta>0, the projection P±,λP_{\pm,\lambda} does not admit a 1+δ1+\delta-localized generalized Wannier basis. However, the projections admit a 1δ1-\delta-localized generalized Wannier basis for small disorder.

We make a few observations related to Theorem 4 and the notion of Wannier bases. First, these theorems imply a lower bound on the uniform decay of eigenfunctions for the random Hamiltonian. In particular, it implies that if the random Hamiltonian exhibits pure point spectrum, then the decay is not too fast in a uniform sense which should be compared with the notion of SUDEC, see Def. 4.1 which one obtains by applying the multiscale analysis. In particular, one has

Remark 3 (Lower bound on uniform eigenfunction decay).

If the Hamiltonian only exhibits point spectrum in the interval II, for which the associated spectral projections does not admit a 1+δ1+\delta generalized Wannier basis, then we can choose an orthonormal basis of eigenfunctions (ψβ)(\psi_{\beta}) such that span¯(ψβ)=ran(P)\overline{\operatorname{span}}(\psi_{\beta})=\operatorname{ran}(P) and any sequence of localization centers μβ\mu_{\beta}

supβzμβ2+δ|ψβ(z)|2𝑑z=.\sup_{\beta}\int_{{\mathbb{C}}}\langle z-\mu_{\beta}\rangle^{2+\delta}|\psi_{\beta}(z)|^{2}\ dz=\infty.

In this sense, Theorem 4 gives a lower-bound on the decay of eigenfunctions in case that the random Hamiltonian exhibits only pure point spectrum.

Outline of article.

  • In Section 2, we study the stability of the first magic angle under small perturbations.

  • In Section 3, we study the regularity of the integrated density of states by stating a estimate on the number of eigenvalues (NE) under Assumption 1.

  • In Section 4, we derive the existence of a mobility edge in a neighborhood of the perturbed flat bands.

  • In Section 5, we prove Theorem 4.

2. (In)stability of magic angles

In this section, we obtain stability bounds on magic angles with respect to perturbations.

We recall the definition of the compact Birman-Schwinger operator TkT_{k}, with k=(ω2k1ωk2)/3k=(\omega^{2}k_{1}-\omega k_{2})/\sqrt{3}, where (k1,k2)2(32+{(0,0),(1,1)})(k_{1},k_{2})\in\mathbb{R}^{2}\setminus(3{\mathbb{Z}}^{2}+\{(0,0),(-1,-1)\}). This operator is defined by

Tk:=(2Dz¯k)1(0U(z)U(z)0):L02(/Γ;2)(H1L02)(/Γ;2),T_{k}:=(2D_{\bar{z}}-k)^{-1}\begin{pmatrix}0&U(z)\\ U(-z)&0\end{pmatrix}:L^{2}_{0}(\mathbb{C}/\Gamma;\mathbb{C}^{2})\to(H^{1}\cap L^{2}_{0})(\mathbb{C}/\Gamma;\mathbb{C}^{2}),

where

Lp2(/Γ;2):={uL2(/Γ,2):𝐚u(z)=e2πi(a1p+a2p)u(z+𝐚),aj13},L^{2}_{p}(\mathbb{C}/\Gamma;\mathbb{C}^{2}):=\Big{\{}u\in L^{2}(\mathbb{C}/\Gamma,{\mathbb{C}}^{2}):\mathscr{L}_{\mathbf{a}}u(z)=e^{2\pi i(a_{1}p+a_{2}p)}u(z+\mathbf{a}),\ a_{j}\in\tfrac{1}{3}\mathbb{Z}\Big{\}},

for 𝐚=4πi(ωa1+ω2a2).\mathbf{a}=4\pi i(\omega a_{1}+\omega^{2}a_{2}). For scalar functions, we also define spaces Lp2(/Γ;2)L^{2}_{p}({\mathbb{C}}/\Gamma;{\mathbb{C}}^{2}) where we replace the translation operator by its first component (1.4). As described in (1.6), α0\alpha\neq 0 is magic if and only if 1/αSpecL02(Tk){0}.1/\alpha\in\operatorname{Spec}_{L^{2}_{0}}(T_{k})\setminus\{0\}. One can then show that 1/αSpecL02(Tk){0}1/\alpha\in\operatorname{Spec}_{L^{2}_{0}}(T_{k})\setminus\{0\} if and only if 1/αSpecL12(Tk){0},1/\alpha\in\operatorname{Spec}_{L^{2}_{1}}(T_{k})\setminus\{0\}, see [BHZ22b]. By squaring the operator, we define new compact operators Ak,BkA_{k},B_{k}

Tk2=:3diag(Ak,Bk) with Ak:=(2Dz¯k)1U(z)(2Dz¯k)1U(z) and Bk:=(2Dz¯k)1U(z)(2Dz¯k)1U(z).\begin{split}&T_{k}^{2}=:3\operatorname{diag}(A_{k},B_{k})\text{ with }A_{k}:=(2D_{\bar{z}}-k)^{-1}U(z)(2D_{\bar{z}}-k)^{-1}U(-z)\text{ and }\\ &B_{k}:=(2D_{\bar{z}}-k)^{-1}U(-z)(2D_{\bar{z}}-k)^{-1}U(z).\end{split} (2.1)

Since the operator (2Dz¯k)1(2D_{\bar{z}}-k)^{-1} is compact, it follows that

SpecL12(Tk2){0}=3SpecL12(Ak){0}=3SpecL22(Bk){0}.\operatorname{Spec}_{L^{2}_{1}}(T_{k}^{2})\setminus\{0\}=3\operatorname{Spec}_{L^{2}_{1}}(A_{k})\setminus\{0\}=3\operatorname{Spec}_{L^{2}_{2}}(B_{k})\setminus\{0\}.

This implies that tr((Tk2)n)=23ntr(Akn)\operatorname{tr}((T_{k}^{2})^{n})=2\cdot 3^{n}\operatorname{tr}(A_{k}^{n}) for n>1n>1 which is well-defined as AkA_{k} is a Hilbert-Schmidt operator. In particular, the spectrum of AkA_{k} is independent of kk, see [Be*22, Prop.3.1.] and we have SpecL12(A0)=SpecL12(Ak)\operatorname{Spec}_{L^{2}_{1}}(A_{0})=\operatorname{Spec}_{L^{2}_{1}}(A_{k}) for any kΓ.k\notin\Gamma^{*}. The traces of powers of AkA_{k} are illustrated in Table 1. We thus have that

{0}α is magic 1/αSpecL12(Tk)1/(3α2)SpecL12(Ak),\mathbb{C}\setminus\{0\}\ni\alpha\text{ is magic }\Leftrightarrow 1/\alpha\in\operatorname{Spec}_{L^{2}_{1}}(T_{k})\Leftrightarrow 1/(3\alpha^{2})\in\operatorname{Spec}_{L^{2}_{1}}(A_{k}),

where the right-hand side depends only on α\alpha and the unperturbed operator A.A.

pp σp3π\sigma_{p}\frac{\sqrt{3}}{\pi}
1 2/92/9
2 4/94/9
3 32/6332/63
4 40/8140/81
pp σp3π\sigma_{p}\frac{\sqrt{3}}{\pi}
5 9560/20007{9560}/{20007}
6 245120/527877{245120}/{527877}
7 1957475168/4337177481{1957475168}/{4337177481}
8 13316086960/30360242367{13316086960}/{30360242367}
Table 1. Traces of AkA_{k}, σp=tr(Akp)\sigma_{p}=\operatorname{tr}(A_{k}^{p}), where σ1{\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}\sigma_{1}} is not absolutely summable as AkA_{k} is not of trace-class.

We then consider a perturbation of the potentials which gives us a new operator

Tk,δ=(2Dz¯k)1(0U(z)+δV+U(z)+δV0):L12L12,T_{k,\delta}=(2D_{\bar{z}}-k)^{-1}\begin{pmatrix}0&U(z)+\delta V_{+}\\ U(-z)+\delta V_{-}&0\end{pmatrix}:L^{2}_{1}\to L^{2}_{1},

with bounded potentials V±C(/Γ)V_{\pm}\in C^{\infty}({\mathbb{C}}/\Gamma) and δ>0\delta>0 where V±V_{\pm} satisfies the same symmetries as U(±)U(\pm\bullet), respectively, cf. (1.2). By squaring the operator, similar to (LABEL:eq:Tk), we define

Tk,δ2=:3diag(Ak,δ,Bk,δ)T_{k,\delta}^{2}=:3\operatorname{diag}(A_{k,\delta},B_{k,\delta}) (2.2)

such that SpecL12(Tk,δ2){0}=3SpecL12(Ak,δ){0}=3SpecL22(Bk,δ){0}.\operatorname{Spec}_{L^{2}_{1}}(T_{k,\delta}^{2})\setminus\{0\}=3\operatorname{Spec}_{L^{2}_{1}}(A_{k,\delta})\setminus\{0\}=3\operatorname{Spec}_{L^{2}_{2}}(B_{k,\delta})\setminus\{0\}.

To describe the spectral (in)-stability of non-normal operators one resorts to the pseudospectrum, see also the book [ET05].

Definition 2.1.

Let PP be a bounded linear operator. We denote the ε\varepsilon-pseudospectrum of PP, for every ε>0\varepsilon>0, by

Specε(P):=KL(H);KεSpec(P+K),\operatorname{Spec}_{\varepsilon}(P):=\bigcup_{K\in L(H);\|K\|\leq\varepsilon}\operatorname{Spec}(P+K), (2.3)

with L(H)L(H) the space of bounded linear operators. It is equivalently characterized by

Specε(P):=Spec(P){zSpec(P);(zP)1>1/ε}.\operatorname{Spec}_{\varepsilon}(P):=\operatorname{Spec}(P)\cup\{z\notin\operatorname{Spec}(P);\|(z-P)^{-1}\|>1/\varepsilon\}.

2.1. Stability of magic angles

In order to study the stability of small magic angles, characterized by the eigenvalues of AkA_{k} (α\alpha is magic if and only if (3α2)1SpecL12(Ak)(3\alpha^{2})^{-1}\in\operatorname{Spec}_{L^{2}_{1}}(A_{k})), we start with a resolvent bound and recall the definition of the regularized determinant for a Hilbert-Schmidt operator TT [Si77]

det2(1+T):=λSpec(T)(1+λ)eλ.\det_{2}(1+T):=\prod_{\lambda\in\operatorname{Spec}(T)}(1+\lambda)e^{-\lambda}.

The following estimate is non-trivial, as the operator AkA_{k} is non-normal:

Lemma 2.2.

Let A=A0A=A_{0} be as above, then for α\alpha\in{\mathbb{C}} such that 1SpecL12(3α2A)1\notin\operatorname{Spec}_{L^{2}_{1}}(3\alpha^{2}A)

(13α2A)11+e(6|α|2+e)|det2(13α2A)|.\|(1-3\alpha^{2}A)^{-1}\|\leq 1+\frac{e^{(6|\alpha|^{2}+e)}}{|\det_{2}(1-3\alpha^{2}A)|}.

Before stating the proof of this lemma we state a perturbation estimate that limits by how much the eigenvalues of AδA_{\delta} can spread. This bound is illustrated in Fig. 2.

Refer to captionMagic angle
Figure 2. This figure shows the right-hand side of equation (2.4) close to the first magic angle.
Theorem 6.

Let A:=A0A:=A_{0}, as in (LABEL:eq:Tk), with Spec(A){0}\sqrt{\operatorname{Spec}(A)}\setminus\{0\} the magic angles, and define Aδ:=A0,δA_{\delta}:=A_{0,\delta} as in (2.2). The perturbed operator AδA_{\delta} does not have any eigenvalues 1/(3α2)1/(3\alpha^{2}) with α{0}\alpha\in{\mathbb{C}}\setminus\{0\} as long as the size of the perturbation satisfies

AδA|det2(13α2A)|3|α2|(|det2(13α2A)|+e(6|α|2+e)).\|A_{\delta}-A\|\leq\frac{|\det_{2}(1-3\alpha^{2}A)|}{3|\alpha^{2}|\Big{(}|\det_{2}(1-3\alpha^{2}A)|+e^{(6|\alpha|^{2}+e)}\Big{)}}. (2.4)

Before stating the proof of this result, we shall briefly discuss the interpretation of (2.4). The right hand side of (2.4) is small for large |α||\alpha|, i.e. small twisting angles as well as for 1/(3α2)1/(3\alpha^{2}) close to SpecL12(A)\operatorname{Spec}_{L^{2}_{1}}(A), i.e. for α\alpha that are almost magic. This means that for such α\alpha even small perturbations of the potential may generate eigenvalues of the form 1/(3α2)1/(3\alpha^{2}) of the perturbed operator AδA_{\delta}. This shows that such α\alpha are inherently unstable, as small perturbations can generate and destroy them. Conversely, for large twisting angles, i.e. small α,\alpha, it is in general impossible to generate spectrum 1/(3α2)1/(3\alpha^{2}) of the perturbed operator. In particular, this bound implies a spectral stability for small α\alpha, i.e. large magic angles, since they cannot move by much. The regularized determinant in (2.4) can be controlled (from above and below) by Lemma 2.3.

Proof of Theo. 6.

On the one hand by (2.3), we find

Spec1(3α2Aδ)Spec1,3α2(AδA)(3α2A).\operatorname{Spec}_{1}(3\alpha^{2}A_{\delta})\subset\operatorname{Spec}_{1,3\|\alpha^{2}(A_{\delta}-A)\|}(3\alpha^{2}A).

This implies that if 1Spec1(3α2Aδ),1\in\operatorname{Spec}_{1}(3\alpha^{2}A_{\delta}), then by the characterization of the pseudo-spectrum and Lemma 2.2

13α2(AδA)(13α2A)11+e(6|α|2+e)|det2(13α2A)|.\frac{1}{3\|\alpha^{2}(A_{\delta}-A)\|}\leq\|(1-3\alpha^{2}A)^{-1}\|\leq 1+\frac{e^{(6|\alpha|^{2}+e)}}{|\operatorname{det}_{2}(1-3\alpha^{2}A)|}.

Rearranging this estimate implies the result. ∎

We now give the proof of the auxiliary Lemma 2.2.

Proof of Lemma 2.2.

We recall from [Si77, Theo.5.15.1] that together with SS1S22\|S^{*}S\|_{1}\leq\|S\|_{2}^{2} for SS a Hilbert-Schmidt operator and KK a trace-class operator with ν=1+e1/2\nu=1+e^{1/2}

|det2(1+S+K)|eS222+νK1.|\det_{2}(1+S+K)|\leq e^{\frac{\|S\|_{2}^{2}}{2}+\nu\|K\|_{1}}. (2.5)

Assuming 1+S1+S is invertible and SS a finite rank operator, we have for the usual determinant

det(1+S+μK)=det(1+S)det(1+μ(1+S)1K)=det(1+S)(1+μtr((1+S)1K))+𝒪(μ2).\begin{split}\det(1+S+\mu K)&=\det(1+S)\det(1+\mu(1+S)^{-1}K)\\ &=\det(1+S)(1+\mu\operatorname{tr}((1+S)^{-1}K))+\mathcal{O}(\mu^{2}).\end{split}

This shows that

μ|μ=0det(1+S+μK)=det(1+S)tr((1+S)1K))\partial_{\mu}|_{\mu=0}\det(1+S+\mu K)=\det(1+S)\operatorname{tr}((1+S)^{-1}K))

which shows

μ|μ=0logdet(1+S+μK)=tr((1+S)1K)).\partial_{\mu}|_{\mu=0}\log\det(1+S+\mu K)=\operatorname{tr}((1+S)^{-1}K)).

Using that

logdet2(1+S+μK)=logdet(1+S+μK)tr(S+μK),\log\det_{2}(1+S+\mu K)=\log\det(1+S+\mu K)-\operatorname{tr}(S+\mu K), (2.6)

we find the log-derivative of the regularized 2-determinant

μ|μ=0log(det2(1+S+μK))=tr((1+S)1K)tr(K).\partial_{\mu}|_{\mu=0}\log(\det_{2}(1+S+\mu K))=\operatorname{tr}((1+S)^{-1}K)-\operatorname{tr}(K).

By using a density argument it follows that this formula also holds for SS Hilbert-Schmidt, i.e. we can drop the assumption that SS is of finite rank. Thus, one finds from (2.6) by specializing to K=ϕ,ψK=\langle\phi,\bullet\rangle\psi, with ϕ=ψ=1\|\phi\|=\|\psi\|=1 and multiplying by det2(1+S)\det_{2}(1+S)

det2(1+S)ϕ,(1+S)1ψ=μ|μ=0det2(1+S+μK)det2(1+S)ϕ,ψ.\det_{2}(1+S)\langle\phi,(1+S)^{-1}\psi\rangle=\partial_{\mu}\Big{|}_{\mu=0}\det_{2}(1+S+\mu K)-\det_{2}(1+S)\langle\phi,\psi\rangle.

Hence, using a Cauchy estimate |μ|μ=0f(μ)|sup|μ|=1|f(μ)||\partial_{\mu}|_{\mu=0}f(\mu)|\leq\sup_{|\mu|=1}|f(\mu)| for f(μ):=det2(1+S+μK)f(\mu):=\det_{2}(1+S+\mu K), we find

det2(1+S)(1+S)1sup|μ|=1|det2(1+S+μK)|+|det2(1+S)|\begin{split}\|\det_{2}(1+S)(1+S)^{-1}\|&\leq\sup_{|\mu|=1}|\det_{2}(1+S+\mu K)|+|\det_{2}(1+S)|\end{split}

it thus follows together with (2.5) that

(1+S)11+sup|μ|=1|det2(1+S+μK)||det2(1+S)|1+eS22/2+ν|det2(1+S)|.\begin{split}\|(1+S)^{-1}\|&\leq 1+\sup_{|\mu|=1}\frac{|\det_{2}(1+S+\mu K)|}{|\det_{2}(1+S)|}\leq 1+\frac{e^{\|S\|_{2}^{2}/2+\nu}}{|\det_{2}(1+S)|}\end{split}.

Specializing the estimate to S=3α2AS=-3\alpha^{2}A, we find by using that A022\|A_{0}\|_{2}\leq 2, see [BHZ22, Lemma 4.14.1] and ν<e\nu<e, see [Si77],

(13α2A)11+e(6|α|2+e)|det2(13α2A)|\|(1-3\alpha^{2}A)^{-1}\|\leq 1+\frac{e^{(6|\alpha|^{2}+e)}}{|\det_{2}(1-3\alpha^{2}A)|}

which was to be shown. ∎

Consequently, if α\alpha_{*} is a magic angle, we can estimate det2(13α2A0)\det_{2}(1-3\alpha^{2}A_{0}) in (2.4) by using [BHZ22b, Lemma 5.15.1], which in a reduced version states that

Lemma 2.3.

The entire function αdet2(13α2Ak){\mathbb{C}}\ni\alpha\mapsto\operatorname{det}_{2}(1-3\alpha^{2}A_{k}) satisfies for any n0n\geq 0

|det2(13α2Ak)k=0nμk(3)kα2kk!|j=n+1(eA023|α|2j)j\begin{split}\left\lvert\operatorname{det}_{2}(1-3\alpha^{2}A_{k})-\sum_{k=0}^{n}\mu_{k}\frac{(-3)^{k}\alpha^{2k}}{k!}\right\rvert&\leq\sum_{j=n+1}^{\infty}\Bigg{(}\frac{\sqrt{e}\|A_{0}\|_{2}3|\alpha|^{2}}{\sqrt{j}}\Bigg{)}^{j}\end{split}

with A022\|A_{0}\|_{2}\leq 2, where

μj=det(0j100σ20j20σj1σj201σjσj1σj20), with σj=trAkj.{\mu}_{j}=\operatorname{det}\begin{pmatrix}0&j-1&0&\cdots&0\\ \sigma_{2}&0&j-2&\cdots&0\\ \vdots&\vdots&\ddots&\ddots&\vdots\\ \sigma_{j-1}&\sigma_{j-2}&\cdots&0&1\\ \sigma_{j}&\sigma_{j-1}&\sigma_{j-2}&\cdots&0\end{pmatrix},\text{ with }\sigma_{j}=\operatorname{tr}A_{k}^{j}. (2.7)

The first few traces σj\sigma_{j} are summarized in Table 1.

2.2. Instability of magic angles

Refer to captionRefer to captionRe(α)\operatorname{Re}(\alpha)Re(α)\operatorname{Re}(\alpha)Im(α)\operatorname{Im}(\alpha)zlog((Tk1/z)1)z\mapsto\log(\|(T_{k}-1/z)^{-1}\|)Refer to captionRefer to captionIm(α)\operatorname{Im}(\alpha)Re(α)\operatorname{Re}(\alpha)Re(α)\operatorname{Re}(\alpha)
Figure 3. Upper row: Magic angles (left) and resolvent norm of operator TkT_{k} (right).
Lower row: 1000 realizations of random perturbations of tunnelling potential U+δVU+\delta V with new magic angles (black dots) superimposed on resolvent norm figure. δ=1/100\delta=1/100 (left) and δ=1/10\delta=1/10 (right).

We shall now give the proof of Theorem 1. Arbitrary low-lying eigenvalues of TkT_{k}, which correspond to large magic angles in the unperturbed case, can be produced by rank 11 perturbations of TkT_{k} that are exponentially small in the spectral parameter. Let μ\mu be one such low-lying eigenvalue of TkT_{k}. On the Hamiltonian side, this indicates that zero modes with quasi-momentum kk and α=1/μ\alpha=1/\mu can be generated by rank one perturbations of the Bloch-Floquet Hamiltonian, Hk(α)H_{k}(\alpha)

Proof of Theo. 1.

We recall that by [Be*22, Theo 44] there exists for each kk\in\mathbb{C} an L2L^{2}-normalized uμCc(;2)u_{\mu}\in C_{c}^{\infty}({\mathbb{C}};{\mathbb{C}}^{2}) such that the operator

P(μ)=(2μDz¯U(z)U(z)2μDz¯)P(\mu)=\begin{pmatrix}2\mu D_{\bar{z}}&U(z)\\ U(-z)&2\mu D_{\bar{z}}\end{pmatrix}

satisfies (P(μ)μk)uμ=𝒪(ec/|μ|)\|(P(\mu)-\mu k)u_{\mu}\|=\mathcal{O}(e^{-c/|\mu|}) with uμL2=1\|u_{\mu}\|_{L^{2}}=1 and c>0.c>0. This implies that there is a constant K>0K>0, which we allow to change throughout this proof, such that (P(μ)μk)1Kec/|μ|.\|(P(\mu)-\mu k)^{-1}\|\geq Ke^{c/|\mu|}. Hence, we define the normalized vμ:=(P(μ)μk)uμ(P(μ)μk)uμ,v_{\mu}:=\frac{(P(\mu)-\mu k)u_{\mu}}{\|(P(\mu)-\mu k)u_{\mu}\|}, then (P(μ)μk)1vμ>Kec/|μ|.\|(P(\mu)-\mu k)^{-1}v_{\mu}\|>Ke^{c/|\mu|}. We recall that

(P(μ)μk)1=(Tkμ)1(2Dz¯k)1.(P(\mu)-\mu k)^{-1}=-(T_{k}-\mu)^{-1}(2D_{\bar{z}}-k)^{-1}.

This implies that, since (2Dz¯k)1=1/d(k,Γ)\|(2D_{\bar{z}}-k)^{-1}\|=1/d(k,\Gamma^{*}), where dd denotes the Hausdorff distance

(Tkμ)1Kec/|μ|.\left\lVert(T_{k}-\mu)^{-1}\right\rVert\geq Ke^{c/|\mu|}.

Hence, for the normalized sμ:=(2Dz¯k)1vμ(2Dz¯k)1vμs_{\mu}:=\frac{(2D_{\bar{z}}-k)^{-1}v_{\mu}}{\|(2D_{\bar{z}}-k)^{-1}v_{\mu}\|}, we have

(Tkμ)1sμ=tμ with tμKec/|μ|.(T_{k}-\mu)^{-1}s_{\mu}=t_{\mu}\text{ with }\|t_{\mu}\|\geq Ke^{c/|\mu|}.

Thus, we can define Rφ:=φ,tμtμ2sμR\varphi:=\frac{\langle\varphi,t_{\mu}\rangle}{\|t_{\mu}\|^{2}}s_{\mu} with norm R=𝒪(ec/|μ|)\|R\|=\mathcal{O}(e^{-c/|\mu|}) such that

μSpec(TkR).\mu\in\operatorname{Spec}(T_{k}-R).

3. Integrated DOS and Wegner estimate

In this section we prove Theorem 2 by stating a proof of Prop. 1.3, i.e. study the regularity of the integrated density of states and prove a corresponding estimate on the number of eigenvalues of the disordered Hamiltonian. This then also implies a Wegner estimate by (4.2). We start by giving the proof of Hölder continuity, which uses the spectral shift function, see [CHK07, CHK03, CHN01], and then subsequently explain the modifications to obtain Lipschitz continuity, which uses spectral averaging. In the following, we will write χx,L:=1lΛL(x)\chi_{x,L}:=\operatorname{1\hskip-2.75ptl}_{\Lambda_{L}}(x) with χx:=χx,1\chi_{x}:=\chi_{x,1}, ΛL:=/(LΓ)\Lambda_{L}:={\mathbb{C}}/(L\Gamma), and ΛL(z):=z+ΛL\Lambda_{L}(z):=z+\Lambda_{L} and often drop subscripts to simplify the notation. Let AA be a compact operator then Ak\|A\|_{k} denotes the kk-th Schatten class norm.

3.1. Proof of Prop. 1.3

In this subsection we shall give the proof of Prop. 1.3, up to two crucial estimates that are provided in different subsections, namely the Hölder estimate (3.13) in Subsection 3.2 and the Lipschitz estimate (3.15) in Subsection 3.3.

Proof of Prop. 1.3.

In the proof, we shall focus on Case 1 disorder as Case 2 disorder follows along the same lines but more care is needed since the potential uu is not positive in Case 1. We shall emphasize the differences of the two cases in our proof. Since the spectrum in Case 1 exhibits a spectral gap, see (1.14), we may focus without loss of generality on the spectrum around mm. The argument around m-m is analogous. In Case 2, we do not have to restrict ourselves to those neighborhoods. Let E0ΔΔ~(k,k+)E_{0}\in\Delta\subset\tilde{\Delta}\subset(k_{-},k_{+}) for two closed bounded intervals Δ,Δ~\Delta,\tilde{\Delta}, with Δ\Delta of non-empty interior, centered at E0E_{0} and d0:=d(E0,Δ~)d_{0}:=d(E_{0},{\mathbb{R}}\setminus\tilde{\Delta}). We decompose

tr(1lΔ(Hλ,ΛL))=tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))+tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)).\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))=\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}}))+\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})). (3.1)

We then write for the second term in (3.1)

tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))=tr(1lΔ(Hλ,ΛL)(Hλ,ΛLE0)(H0,ΛLE0)11lΔ~(H0,ΛL))tr(1lΔ(Hλ,ΛL)λVω,ΛL(H0,ΛLE0)11lΔ~(H0,ΛL)).\begin{split}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))&=\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})(H_{\lambda,\Lambda_{L}}-E_{0})(H_{0,\Lambda_{L}}-E_{0})^{-1}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))\\ &-\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\lambda V_{\omega,\Lambda_{L}}(H_{0,\Lambda_{L}}-E_{0})^{-1}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})).\end{split} (3.2)

The first term in (3.2) satisfies by Hölder’s inequality and the definition of d0d_{0}

|tr(1lΔ(Hλ,ΛL)(Hλ,ΛLE0)(H0,ΛLE0)11lΔ~(H0,ΛL))||Δ|2d0tr(1lΔ(Hλ,ΛL)).|\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})(H_{\lambda,\Lambda_{L}}-E_{0})(H_{0,\Lambda_{L}}-E_{0})^{-1}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))|\leq\frac{|\Delta|}{2d_{0}}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})).

We then use the inequality

tr(1lΔ(Hλ,ΛL)λVω,ΛL(H0,ΛLE0)11lΔ~(H0,ΛL))|λ|d01lΔ(Hλ,ΛL)Vω,ΛL21lΔ~(H0,ΛL)1lΔ(Hλ,ΛL)2ζtr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))2d0+λ2tr(1lΔ(Hλ,ΛL)Vω,ΛL2)2ζd0,\begin{split}&\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\lambda V_{\omega,\Lambda_{L}}(H_{0,\Lambda_{L}}-E_{0})^{-1}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))\\ &\leq\frac{|\lambda|}{d_{0}}\|\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})V_{\omega,\Lambda_{L}}\|_{2}\|\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\|_{2}\\ &\leq\frac{\zeta\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))}{2d_{0}}+\frac{\lambda^{2}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})V_{\omega,\Lambda_{L}}^{2})}{2\zeta d_{0}},\end{split}

with ζ>0\zeta>0. We can then bound (3.2), in terms of

V~ω,ΛL:=γΛ~Lu(γξγ),\tilde{V}_{\omega,\Lambda_{L}}:=\sum_{\gamma\in\tilde{\Lambda}_{L}}u(\bullet-\gamma-\xi_{\gamma}),

by choosing ζ>0\zeta>0 sufficiently small

tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))|Δ|d0tr(1lΔ(Hλ,ΛL))+λ2tr(1lΔ(Hλ,ΛL)Vω,ΛL2)ζd0|Δ|d0tr(1lΔ(Hλ,ΛL))+λ2tr(1lΔ(Hλ,ΛL)V~ω,ΛL)ζd0.\begin{split}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))&\leq\frac{|\Delta|}{d_{0}}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))+\frac{\lambda^{2}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})V_{\omega,\Lambda_{L}}^{2})}{\zeta d_{0}}\\ &\lesssim\frac{|\Delta|}{d_{0}}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))+\frac{\lambda^{2}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}})}{\zeta d_{0}}.\end{split} (3.3)

Notice that while we do not have that Vω,ΛL2V~ω,ΛLV_{\omega,\Lambda_{L}}^{2}\lesssim\tilde{V}_{\omega,\Lambda_{L}}, at least in case (1), since V~ω,ΛL\tilde{V}_{\omega,\Lambda_{L}} is not positive, we still have that

tr(1lΔ(Hλ,ΛL)Vω,ΛL2)tr(1lΔ(Hλ,ΛL)V~ω,ΛL).\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})V_{\omega,\Lambda_{L}}^{2})\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}). (3.4)

To see this, observe that

1lΔ(H0,ΛL)=Pker(D(α)ΛL)0.\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{0,\Lambda_{L}})=P_{\ker(D(\alpha)_{\Lambda_{L}})}\oplus 0. (3.5)

Indeed, since ΛL=/(LΓ)\Lambda_{L}={\mathbb{C}}/(L\Gamma), we have by periodicity of the Hamiltonian H0H_{0} that Spec(H0,ΛL)Spec(H0).\operatorname{Spec}(H_{0,\Lambda_{L}})\subset\operatorname{Spec}(H_{0}). Since the spectrum of Hλ,ΛLH_{\lambda,\Lambda_{L}} is uniformly gapped for λ\lambda small, it follows that the spectral projection λ1lΔ(Hλ,ΛL)\lambda\mapsto\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}) is norm-continuous. We conclude from (3.5) that for φ=(φ1,φ2)\varphi=(\varphi_{1},\varphi_{2})

φ=1lΔ(Hλ,ΛL)φφ2ε(λ)φ1.\varphi=\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\varphi\Rightarrow\|\varphi_{2}\|\leq\varepsilon(\lambda)\|\varphi_{1}\|. (3.6)

with ε(0)=0\varepsilon(0)=0 and λε(λ)0\lambda\mapsto\varepsilon(\lambda)\geq 0 continuous. Indeed, applying norms to (3.6), we find by substituting

1lΔ(Hλ,ΛL)=1lΔ(H0,ΛL)+(1lΔ(Hλ,ΛL)1lΔ(H0,ΛL))\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})=\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{0,\Lambda_{L}})+(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})-\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{0,\Lambda_{L}}))

in (3.6) that, since 1lΔ(Hλ,ΛL)1lΔ(H0,ΛL)=𝒪(|λ|)\|\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})-\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{0,\Lambda_{L}})\|=\mathcal{O}(|\lambda|), there is C>0C>0 such that

φ12+φ22Pker(D(α)ΛL)φ1+Cλφ12+φ22.\sqrt{\|\varphi_{1}\|^{2}+\|\varphi_{2}\|^{2}}\leq\|P_{\ker(D(\alpha)_{\Lambda_{L}})}\varphi_{1}\|+C\lambda\sqrt{\|\varphi_{1}\|^{2}+\|\varphi_{2}\|^{2}}.

Rearranging this, we find

(1Cλ)φ12+φ22Pker(D(α)ΛL)φ1φ1(1-C\lambda)\sqrt{\|\varphi_{1}\|^{2}+\|\varphi_{2}\|^{2}}\lesssim\|P_{\ker(D(\alpha)_{\Lambda_{L}})}\varphi_{1}\|\leq\|\varphi_{1}\|

which implies that, since

(1Cλ)1+φ22/φ121 that φ2λ/(1Cλ)φ1(1-C\lambda)\sqrt{1+\|\varphi_{2}\|^{2}/\|\varphi_{1}\|^{2}}\leq 1\text{ that }\|\varphi_{2}\|\lesssim\lambda/(1-C\lambda)\|\varphi_{1}\|

showing (3.6). This then directly implies (3.4), since in the notation of (1.12)

tr(1lΔ(Hλ,ΛL)V~ω,ΛL)=φ ONB of ran(1lΔ(Hλ,ΛL))(φ1,Yφ1φ2,Yφ2+2Re(φ1,Zφ2))φ ONB of ran(1lΔ(Hλ,ΛL))(φ12infYφ22supYφ1φ2supZ)φ ONB of ran(1lΔ(Hλ,ΛL))φ12(infYλ2supYλsupZ)φ ONB of ran(1lΔ(Hλ,ΛL))φ12infY for λ small enough.\begin{split}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}})&=\sum_{\begin{subarray}{c}\varphi\text{ ONB of }\\ \operatorname{ran}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\end{subarray}}\Big{(}\langle\varphi_{1},Y\varphi_{1}\rangle-\langle\varphi_{2},Y\varphi_{2}\rangle+2\operatorname{Re}(\langle\varphi_{1},Z\varphi_{2}\rangle)\Big{)}\\ &\geq\sum_{\begin{subarray}{c}\varphi\text{ ONB of }\\ \operatorname{ran}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\end{subarray}}\Big{(}\|\varphi_{1}\|^{2}\inf Y-\|\varphi_{2}\|^{2}\sup Y-\|\varphi_{1}\|\|\varphi_{2}\|\sup Z\Big{)}\\ &\gtrsim\sum_{\begin{subarray}{c}\varphi\text{ ONB of }\\ \operatorname{ran}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\end{subarray}}\|\varphi_{1}\|^{2}(\inf Y-\lambda^{2}\sup Y-\lambda\sup Z)\\ &\gtrsim\sum_{\begin{subarray}{c}\varphi\text{ ONB of }\\ \operatorname{ran}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\end{subarray}}\|\varphi_{1}\|^{2}\inf Y\text{ for }\lambda\text{ small enough}.\end{split}

We can easily obtain, along the same lines, an upper bound on the left-hand side of (3.4)

tr(1lΔ(Hλ,ΛL)Vω,ΛL2)φ ONB of ran(1lΔ(Hλ,ΛL))φ12sup(Vω,ΛL2)11\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})V_{\omega,\Lambda_{L}}^{2})\lesssim\sum_{\begin{subarray}{c}\varphi\text{ ONB of }\\ \operatorname{ran}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\end{subarray}}\|\varphi_{1}\|^{2}\sup(V_{\omega,\Lambda_{L}}^{2})_{11}

to see that (3.4) holds.

Finally, we have for the first term in (3.1)

tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL)1lΔ(Hλ,ΛL))tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL(1l1lΔ~(H0,ΛL)))tr(1lΔ(Hλ,ΛL)(1l1lΔ~(H0,ΛL))V~ω,ΛL1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL))tr(1lΔ(Hλ,ΛL)(V~ω,ΛL1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL))).\begin{split}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}&(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}}))\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\\ &\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}(\operatorname{1\hskip-2.75ptl}-\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})))\\ &\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})(\operatorname{1\hskip-2.75ptl}-\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))\tilde{V}_{\omega,\Lambda_{L}}-\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))\\ &\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})(\tilde{V}_{\omega,\Lambda_{L}}-\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})\\ &\qquad-\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})-\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))).\end{split} (3.7)

Here, we used in the first line of (3.7) the following identity that we shall verify below

1lΔ~(H0,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL).\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\lesssim\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}}). (3.8)

Since H0,ΛLH_{0,\Lambda_{L}} is the unperturbed Hamiltonian, the eigenvectors associated to spectrum in Δ~\tilde{\Delta} are supported in the first two entries of the wavefunction, cf. (3.5). Let π1:=diag(1,1,0,0)\pi_{1}:=\operatorname{diag}(1,1,0,0) be the projection onto the first two entries.

We can then define another auxiliary potential V^ΛL(z):=infξDΓγΛ~Lπ1u(zγξ)π1.\hat{V}_{\Lambda_{L}}(z):=\inf_{\xi\in D^{\Gamma}}\sum_{\gamma\in\tilde{\Lambda}_{L}}\pi_{1}u(z-\gamma-\xi)\pi_{1}. Thus, one has that 0V^ΛLπ1V~ω,ΛLπ10\leq\hat{V}_{\Lambda_{L}}\leq\pi_{1}\tilde{V}_{\omega,\Lambda_{L}}\pi_{1}. The projection onto the first two components is redundant for case 22 disorder since u0u\geq 0 in that case.

Finally to show (3.8) it suffices to show that

1lΔ~(H0,ΛL)1lΔ~(H0,ΛL)V^ΛL1lΔ~(H0,ΛL).\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\lesssim\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\hat{V}_{\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}}).

Since H0H_{0} is a periodic Hamiltonian with respect to any lattice LΓL\Gamma it suffices by Bloch-Floquet theory to prove the estimate for the Bloch functions of the full Hamiltonian H0.H_{0}. Indeed, let (vi(k))iI(k)(v_{i}(k))_{i\in I(k)} be the Bloch functions associated with the spectral projection 1lΔ~(H0),\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0}), where I(k)I(k) is the set of Bloch eigenvalues inside Δ~\tilde{\Delta} with quasimomentum kk, where H0,ΛLH_{0,\Lambda_{L}} has a finite subset (in kk) of those as eigenvectors. It then suffices to show that M(k):=(vi(k),V^ΛLvj(k)L2())i,jM(k):=(\langle v_{i}(k),\hat{V}_{\Lambda_{L}}v_{j}(k)\rangle_{L^{2}({\mathbb{C}})})_{i,j} is strictly positive definite for all kk. If not, then there is k0k_{0}\in{\mathbb{C}} and w(k0):=jβjvjw(k_{0}):=\sum_{j}\beta_{j}v_{j} with βj\beta_{j} not all zero, such that M(k0)w(k0)=0M(k_{0})w(k_{0})=0 and by strict positivity of V^ΛL\hat{V}_{\Lambda_{L}}, see (1.12), we find w(k0)|Bε(z0)0,w(k_{0})|_{B_{\varepsilon}(z_{0})}\equiv 0, but this implies that w0w\equiv 0 by real-analyticity of w(k0),w(k_{0}), since H0H_{0} is elliptic with real-analytic coefficients, which is a contradiction. Thus MkM_{k} is a strictly positive matrix and using continuity in kk and compactness of /Γ{\mathbb{C}}/\Gamma^{*}, we also see that Mk>c0>0M_{k}>c_{0}>0 for all kk.

For the second term in (3.7) observe that by the boundedness of the potential

|tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL)1lΔ(Hλ,ΛL))|tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))|\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))|\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))

where the last term can be estimated using (3.3).

We shall now estimate the third and fourth term at the end of (3.7) for δ>0\delta>0, using Young’s, the Cauchy-Schwarz inequality, and that A2=A2\|A\|_{2}=\|A^{*}\|_{2}

|tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL))|tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))2δ+δ21lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL22tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))2δ+δ21lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)22\begin{split}&|\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}}))|\\ &\leq\frac{\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))}{2\delta}+\frac{\delta}{2}\|\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\|_{2}^{2}\\ &\lesssim\frac{\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))}{2\delta}+\frac{\delta}{2}\|\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\|_{2}^{2}\end{split}

and similarly

|tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)V~ω,ΛL1lΔ~(H0,ΛL))|tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))2δ+δ21lΔ(Hλ,ΛL)1lΔ~(H0,ΛL)22.\begin{split}&|\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))|\\ &\lesssim\frac{\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))}{2\delta}+\frac{\delta}{2}\|\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}})\|_{2}^{2}.\end{split}

Inserting the last two estimates into (3.7) and choosing δ>0\delta>0 small enough

tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))tr(1lΔ(Hλ,ΛL)V~ω,ΛL)+tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))δ.\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{\tilde{\Delta}}(H_{0,\Lambda_{L}}))\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}})+\frac{\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))}{\delta}.

Inserting this estimate into (3.1) yields

tr(1lΔ(Hλ,ΛL))tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))+tr(1lΔ(Hλ,ΛL)V~ω,ΛL)+tr(1lΔ(Hλ,ΛL)1lΔ~(H0,ΛL))δ.\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))+\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}})+\frac{\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\operatorname{1\hskip-2.75ptl}_{{\mathbb{R}}\setminus\tilde{\Delta}}(H_{0,\Lambda_{L}}))}{\delta}.

Thus, by choosing |Δ||\Delta| sufficiently small in (3.3)

tr(1lΔ(Hλ,ΛL))tr(1lΔ(Hλ,ΛL)V~ω,ΛL).\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\lesssim\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}}).

Applying expectation values and using (3.13), which we show below, we find for q(0,1)q\in(0,1)

𝐄tr(1lΔ(Hλ,ΛL))𝐄tr(1lΔ(Hλ,ΛL)V~ω,ΛL)q|Δ|q|ΛL|\mathbf{E}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\lesssim\mathbf{E}\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\tilde{V}_{\omega,\Lambda_{L}})\lesssim_{q}|\Delta|^{q}|\Lambda_{L}| (3.9)

which shows the result by using a partition of small intervals Δ\Delta covering II. ∎

3.2. Spectral shift function and Hölder continuity

To obtain the Hölder estimate, used to show (3.9), we recall the definition of the spectral shift function, first. Let H0H_{0} and H1H_{1} be two self-adjoint operator such that H1H0H_{1}-H_{0} is trace-class, then the spectral shift function is defined as, see [Y92, Ch. 88, Sec. 22, Theo. 11]

ξ(λ,H1,H0):=1πlimε0argdet(id+(H1H0)(H0λiε)1).\xi(\lambda,H_{1},H_{0}):=\frac{1}{\pi}\lim_{\varepsilon\downarrow 0}\operatorname{arg}\det(\operatorname{id}+(H_{1}-H_{0})(H_{0}-\lambda-i\varepsilon)^{-1}).

In particular for any p1p\geq 1 one has the LpL^{p} bound [CHN01, Theorem 2.12.1]

ξ(,H1,H0)LpH1H01/p1/p\|\xi(\bullet,H_{1},H_{0})\|_{L^{p}}\leq\|H_{1}-H_{0}\|_{1/p}^{1/p} (3.10)

where the right-hand side is defined as the generalized Schatten norm

Tq=(λSpec(TT)λq/2)1/q.\|T\|_{q}=\Big{(}\sum_{\lambda\in\operatorname{Spec}(T^{*}T)}\lambda^{q/2}\Big{)}^{1/q}.

We then start by setting φ(x):=arctan(xn)\varphi(x):=\arctan(x^{n}), with n20+1n\in 2\mathbb{N}_{0}+1 sufficiently large, such that h1h0h_{1}-h_{0} is trace-class, with h0:=φ(H0)h_{0}:=\varphi(H_{0}) and h1:=φ(H1)h_{1}:=\varphi(H_{1}). Then, we have the Birman-Krein formula, see [Y92, Ch.8, Sec. 1111, Lemma 33] stating that for absolutely continuous ff

tr(f(H1)f(H0))=ξ(φ(λ),h1,h0)𝑑f(λ).\operatorname{tr}(f(H_{1})-f(H_{0}))=\int_{{\mathbb{R}}}\xi(\varphi(\lambda),h_{1},h_{0})\ df(\lambda).

Let Δ=[a,b]\Delta=[a,b] then we start by defining

s(x):={0x03x22x30x111xs(x):=\begin{cases}0&x\leq 0\\ 3x^{2}-2x^{3}&0\leq x\leq 1\\ 1&1\leq x\\ \end{cases}

and

fΔ(t):=1s(ta+12|Δ|2|Δ|).f_{\Delta}(t):=1-s\Bigg{(}\tfrac{t-a+\tfrac{1}{2}|\Delta|}{2|\Delta|}\Bigg{)}. (3.11)

We observe that this function satisfies inftΔ(fΔ(t))=9/8\inf_{t\in\Delta}(-f_{\Delta}^{\prime}(t))=9/8.

Thus, we have for C>0C>0

1lΔ(Hλ,ΛL)C|Δ|fΔ(Hλ,ΛL)\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\leq-C|\Delta|f_{\Delta}^{\prime}(H_{\lambda,\Lambda_{L}})

which implies

tr(λV~ω,ΛL1lΔ(Hλ,ΛL))C|Δ|tr(λV~ω,ΛLfΔ(Hλ,ΛL))=C|Δ|γΛ~Lωγtr(fΔ(Hλ,ΛL)).\begin{split}\operatorname{tr}(\lambda\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))&\leq-C|\Delta|\operatorname{tr}(\lambda\tilde{V}_{\omega,\Lambda_{L}}f_{\Delta}^{\prime}(H_{\lambda,\Lambda_{L}}))\\ &=-C|\Delta|\sum_{\gamma\in\tilde{\Lambda}_{L}}\partial_{\omega_{\gamma}}\operatorname{tr}(f_{\Delta}(H_{\lambda,\Lambda_{L}})).\end{split}

Applying the expectation value to this inequality, we find by positivity of gg, the density of ωγ,\omega_{\gamma}, that for 𝐄γ\mathbf{E}_{\gamma} the expectation value with respect to all random variables (ξγ)(\xi_{\gamma^{\prime}}) and all ωγ\omega_{\gamma^{\prime}} apart from γ=γ\gamma^{\prime}=\gamma

𝐄tr(λV~ω,ΛL1lΔ(Hλ,ΛL))γΛ~L𝐄γC|Δ|01g(ωγ)ωγtr(fΔ(Hλ,ΛL))dωγγΛ~L𝐄γC|Δ|g01ωγtr(fΔ(Hλ,ΛL))dωγC|Δ|gγΛ~L𝐄γtr(fΔ(Hλ,ΛL(ωγ=1))fΔ(Hλ,ΛL(ωγ=0)))=C|Δ|gγΛ~Lsupp(fΔ)fΔ(t)𝐄γξ(φ(t),φ(Hλ,ΛL(ωγ=1)),φ(Hλ,ΛL(ωγ=0)))𝑑t,\begin{split}\mathbf{E}\operatorname{tr}(\lambda&\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\leq-\sum_{\gamma\in\tilde{\Lambda}_{L}}\mathbf{E}_{\gamma}C|\Delta|\int_{0}^{1}g(\omega_{\gamma})\partial_{\omega_{\gamma}}\mathbf{\operatorname{tr}}(f_{\Delta}(H_{\lambda,\Lambda_{L}}))\ d\omega_{\gamma}\\ &\leq-\sum_{\gamma\in\tilde{\Lambda}_{L}}\mathbf{E}_{\gamma}C|\Delta|\|g\|_{\infty}\int_{0}^{1}\partial_{\omega_{\gamma}}\mathbf{\operatorname{tr}}(f_{\Delta}(H_{\lambda,\Lambda_{L}}))\ d\omega_{\gamma}\\ &\leq-C|\Delta|\|g\|_{\infty}\sum_{\gamma\in\tilde{\Lambda}_{L}}\mathbf{E}_{\gamma}\operatorname{tr}(f_{\Delta}(H_{\lambda,\Lambda_{L}}(\omega_{\gamma}=1))-f_{\Delta}(H_{\lambda,\Lambda_{L}}(\omega_{\gamma}=0)))\\ &=C|\Delta|\|g\|_{\infty}\sum_{\gamma\in\tilde{\Lambda}_{L}}\int_{\operatorname{supp}(f_{\Delta})}f_{\Delta}^{\prime}(t)\mathbf{E}_{\gamma}\xi(\varphi(t),\varphi(H_{\lambda,\Lambda_{L}}(\omega_{\gamma}=1)),\varphi(H_{\lambda,\Lambda_{L}}(\omega_{\gamma}=0)))\ dt,\end{split} (3.12)

where Hλ,ΛL(ωγ=ζ)H_{\lambda,\Lambda_{L}}(\omega_{\gamma}=\zeta) is the Hamiltonian Hλ,ΛLH_{\lambda,\Lambda_{L}} with ωγ\omega_{\gamma} replaced by the constant ζ\zeta and |supp(fΔ)|=𝒪(|Δ|).|\operatorname{supp}(f_{\Delta})|=\mathcal{O}(|\Delta|). Thus, using Hölder’s inequality, we find for any β(0,1)\beta\in(0,1) with (3.10) and nn in the arctan\arctan regularization φ\varphi sufficiently large222using φ(t)φ(t0)=t0tnsn11+s2n𝑑s\varphi(t)-\varphi(t_{0})=\int_{t_{0}}^{t}\frac{ns^{n-1}}{1+s^{2n}}\ ds we can create, by choosing nn sufficiently large, arbitrarily large powers of the resolvent. This yields the desired trace-class condition.

𝐄tr(λV~ω,ΛL1lΔ(Hλ,ΛL))|Δ|β|ΛL|\mathbf{E}\operatorname{tr}(\lambda\tilde{V}_{\omega,\Lambda_{L}}\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}}))\lesssim|\Delta|^{\beta}|\Lambda_{L}| (3.13)

which is the identity used to obtain (3.9).

3.3. Spectral averaging and Lipschitz continuity

We now complete the proof of Lipschitz continuity and follow an argument developed initially by Combes and Hislop [CH94, Corr. 4.24.2] for Schrödinger operators.

Proof of Theorem 2 (Lipschitz continuity).

Let E=max{|E1|,|E2|}E=\max\{|E_{1}|,|E_{2}|\} where Δ=[E1,E2],\Delta=[E_{1},E_{2}], then

𝐄(tr(1lΔ(Hλ,ΛL)))eE2𝐄(tr(1lΔ(Hλ,ΛL)eHλ,ΛL2))eE2jΛ~L(𝐄(χj1lΔ(Hλ,ΛL)χj)supωΩtr(χjeHλ,ΛL2))eE2jΛ~L𝐄(χj1lΔ(Hλ,ΛL)χj),\begin{split}\mathbf{E}(\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})))&\leq e^{E^{2}}\mathbf{E}(\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})e^{-H^{2}_{\lambda,\Lambda_{L}}}))\\ &\leq e^{E^{2}}\sum_{j\in\tilde{\Lambda}_{L}}\Bigg{(}\|\mathbf{E}(\chi_{j}\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\chi_{j})\|\sup_{\omega\in\Omega}\operatorname{tr}\Big{(}\chi_{j}e^{-H^{2}_{\lambda,\Lambda_{L}}}\Big{)}\Bigg{)}\\ &\lesssim e^{E^{2}}\sum_{j\in\tilde{\Lambda}_{L}}\|\mathbf{E}(\chi_{j}\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})\chi_{j})\|,\end{split} (3.14)

where we used that supωΩtr(χjeHλ,ΛL2)\sup_{\omega\in\Omega}\operatorname{tr}\Big{(}\chi_{j}e^{-H^{2}_{\lambda,\Lambda_{L}}}\Big{)} is uniformly bounded in all parameters. Under the assumptions of Case 22, we know that uju_{j} are strictly positive on supp(χj)\operatorname{supp}(\chi_{j}) thus also 0χj2uj0\leq\chi_{j}^{2}\lesssim u_{j} which is the necessary condition [CH94, (4.2)] to apply spectral averaging which readily implies together with (3.14) that

𝐄(tr(1lΔ(Hλ,ΛL)))|Δ||ΛL|\mathbf{E}(\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{\Delta}(H_{\lambda,\Lambda_{L}})))\lesssim|\Delta||\Lambda_{L}| (3.15)

which is the identity (3.9) with β=1\beta=1 for Case 22 disorder. ∎

4. Mobility edge

To prove Theorem 3 we recall Germinet and Klein’s notion of summable uniform decay of correlations (SUDEC), see [GK06].

Definition 4.1 (SUDEC).

The Hamiltonian HλH_{\lambda} is said to exhibit a.e. SUDEC in an interval JJ if its spectrum is pure point and for every closed IJI\subset J, for {φn}\{\varphi_{n}\} an orthonormal set of eigenfunctions of HλH_{\lambda} with eigenvalues EnIE_{n}\in I, we define βn:=z2φn2.\beta_{n}:=\|\langle z\rangle^{-2}\varphi_{n}\|^{2}. Then for ζ(0,1)\zeta\in(0,1) there is CI,ζ<C_{I,\zeta}<\infty such that

χz(φnφn)χwCI,ζβnz2w2e|zw|ζ for w,z\|\chi_{z}(\varphi_{n}\otimes\varphi_{n})\chi_{w}\|\leq C_{I,\zeta}\beta_{n}\langle z\rangle^{2}\langle w\rangle^{2}e^{-|z-w|^{\zeta}}\text{ for }w,z\in{\mathbb{C}}

and in addition one has 𝐏\mathbf{P}-almost surely

nβn<.\sum_{n\in\mathbb{N}}\beta_{n}<\infty. (4.1)

The strategy to establish delocalization is to show that if the Hamiltonian would exhibit only SUDEC-type localization (SUDEC), then this would contradict the non-vanishing Chern numbers of the flat bands.

4.1. The ingredients to the multi-scale analysis

For the applicability of the multi-scale analysis à la Germinet-Klein we require six ingredients of our Hamiltonian often referred to in their works by the acronyms, as introduced in [GK01],

  • Strong generalized eigenfunction expansion SGEE (Lemma 4.2),

  • Simon-Lieb inequality SLI and exponential decay inequality EDI (both Lemma 4.3),

  • Number of eigenvalues estimate NE and Wegner estimate W (both (4.2) and Prop. 1.3), and

  • Independence at a distance IAD.

Here, independence at a distance (IAD) just follows from the choice of Anderson-type randomness and means that the disorder of the potentials at a certain distance are independent of each other.

We then start with the strong generalized eigenfunction expansion (SGEE). Therefore, we introduce Hilbert spaces

±:=L2(,4;z±4νdz).\mathcal{H}_{\pm}:=L^{2}({\mathbb{C}},{\mathbb{C}}^{4};\langle z\rangle^{\pm 4\nu}\ dz).
Lemma 4.2 (SGEE).

Let ν1/2\nu\geq 1/2. The set D+ω:={ϕD(Hλ)+;Hλϕ+}D^{\omega}_{+}:=\{\phi\in D(H_{\lambda})\cap\mathcal{H}_{+};H_{\lambda}\phi\in\mathcal{H}_{+}\} is dense in +\mathcal{H}_{+} and a core of Hλ.H_{\lambda}. Moreover, for μ{0}\mu\in\mathbb{R}\setminus\{0\} we have

𝐄(tr(z2ν(Hλiμ)41lI(Hλ)z2ν)2)<.\mathbf{E}\Bigg{(}\operatorname{tr}\Big{(}\langle z\rangle^{-2\nu}(H_{\lambda}-i\mu)^{-4}\operatorname{1\hskip-2.75ptl}_{I}(H_{\lambda})\langle z\rangle^{-2\nu}\Big{)}^{2}\Bigg{)}<\infty.
Proof.

The statement about the core is immediate, as Cc(;4)C_{c}^{\infty}({\mathbb{C}};{\mathbb{C}}^{4}) is a core, see for instance Theorem 8. The second statement follows as z2ν(Hλiμ)2\langle z\rangle^{-2\nu}(H_{\lambda}-i\mu)^{-2} is a uniformly bounded (in ω\omega) Hilbert-Schmidt operator. This is easily seen by ∎

The Simon-Lieb inequality (SLI), relating resolvents at different scales, and the eigenfunction decay inequality (EDI), relating decay of finite-volume resolvents to the decay of generalized eigenfunctions and thus Anderson localization, are discussed in the next Lemma. We thus define ΞΛL(z)\Xi_{\Lambda_{L}(z)} to be the characteristic function of the belt

ΥL(z):=ΛL1(z)ΛL3(z).\Upsilon_{L}(z):=\Lambda_{L-1}(z)\setminus\Lambda_{L-3}(z).

For zΓz\in\Gamma and l>4l>4, we define smooth cut-off functions χ~Λl(z)Cc(;[0,1])\tilde{\chi}_{\Lambda_{l}(z)}\in C_{c}^{\infty}({\mathbb{C}};[0,1]) that are equal to one on Λl3(z)\Lambda_{l-3}(z) and 0 on Λl5/2(z).{\mathbb{C}}\setminus\Lambda_{l-5/2}(z).

Lemma 4.3 (SLI & EDI).

Let JJ be a compact interval. For L,l,l′′2L,l^{\prime},l^{\prime\prime}\in 2\mathbb{N} and x,y,y′′Γx,y^{\prime},y^{\prime\prime}\in\Gamma with Λl′′(y)Λl(y)ΛL(x)\Lambda_{l^{\prime\prime}}(y)\subsetneq\Lambda_{l^{\prime}}(y^{\prime})\subsetneq\Lambda_{L}(x), then 𝐏\mathbf{P}-a.s.: If EJ(Spec(Hλ,ΛL(x))Spec(Hλ,Λl(y)))cE\in J\cap(\operatorname{Spec}(H_{\lambda,\Lambda_{L}(x)})\cap\operatorname{Spec}(H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime})}))^{c} then the Simon-Lieb inequality holds

ΞΛL(x)(Hλ,ΛL(x)E)1χΛl′′(y)JΞΛl(y)(Hλ,Λl(y)E)1χΛl′′(y)×ΞΛL(x)(Hλ,ΛL(x)E)1ΞΛl(y).\begin{split}\|\Xi_{\Lambda_{L}(x)}(H_{\lambda,\Lambda_{L}(x)}-E)^{-1}\chi_{\Lambda_{l^{\prime\prime}}(y)}\|&\lesssim_{J}\|\Xi_{\Lambda_{l^{\prime}}(y^{\prime})}(H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime})}-E)^{-1}\chi_{\Lambda_{l^{\prime\prime}}(y)}\|\\ &\times\|\Xi_{\Lambda_{L}(x)}(H_{\lambda,\Lambda_{L}(x)}-E)^{-1}\Xi_{\Lambda_{l^{\prime}}(y^{\prime})}\|.\end{split}

In addition, we have 𝐏\mathbf{P}-a.s. that any zΓz\in\Gamma, and any generalized eigenfunction ψ\psi, i.e. ψ\psi solving (HλE)ψ=0(H_{\lambda}-E)\psi=0 and growing at most polynomially, with EJSpec(Hλ,ΛL(x))cE\in J\cap\operatorname{Spec}(H_{\lambda,\Lambda_{L}(x)})^{c} one has the eigenfunction decay inequality

χzψJΞΛL(x)(Hλ,ΛL(x)E)1χzΞΛL(x)ψ.\begin{split}\|\chi_{z}\psi\|&\lesssim_{J}\|\Xi_{\Lambda_{L}(x)}(H_{\lambda,\Lambda_{L}(x)}-E)^{-1}\chi_{z}\|\|\Xi_{\Lambda_{L}(x)}\psi\|.\end{split}
Proof.
  1. (1)

    The proof of the SLI can be streamlined for linear differential operators with disorder of Anderson-type. We start from the following resolvent identity

    (HλE)χ~Λl(y)(Hλ,ΛL(x)E)1=[(HλE),χ~Λl(y)](Hλ,ΛL(x)E)1+χ~Λl(y)(HλE)(Hλ,ΛL(x)E)1.\begin{split}(H_{\lambda}-E)\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}(H_{\lambda,\Lambda_{L}(x)}-E)^{-1}&=[(H_{\lambda}-E),\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}](H_{\lambda,\Lambda_{L}(x)}-E)^{-1}\\ &\ +\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}(H_{\lambda}-E)(H_{\lambda,\Lambda_{L}(x)}-E)^{-1}.\end{split}

    Using that by assumption Λl(y)ΛL(x)\Lambda_{l^{\prime}}(y^{\prime})\subset\Lambda_{L}(x) we have χΛl(y)Hλ=χΛl(y)Hλ,ΛL(x)\chi_{\Lambda_{l^{\prime}}(y^{\prime})}H_{\lambda}=\chi_{\Lambda_{l^{\prime}}(y^{\prime})}H_{\lambda,\Lambda_{L}(x)} and find by substituting χΛl(y)Hλ\chi_{\Lambda_{l^{\prime}}(y^{\prime})}H_{\lambda} in the last line above

    (HλE)χ~Λl(y)(Hλ,ΛL(x)E)1=[Hλ,χ~Λl(y)](Hλ,ΛL(x)E)1+χ~Λl(y).(H_{\lambda}-E)\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}(H_{\lambda,\Lambda_{L}(x)}-E)^{-1}=[H_{\lambda},\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}](H_{\lambda,\Lambda_{L}(x)}-E)^{-1}+\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}.

    Since Hλχ~Λl(y)=Hλ,Λl(y)χ~Λl(y)H_{\lambda}\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}=H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime})}\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})} we find by multiplying the previous line by (Hλ,Λl(y)E)1(H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime})}-E)^{-1} that

    χ~Λl(y)(Hλ,ΛL(x)E)1=(Hλ,Λl(y)E)1[Hλ,Λl(y),,χ~Λl(y)](Hλ,ΛL(x)E)1+(Hλ,Λl(y)E)1χ~Λl(y).\begin{split}\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}(H_{\lambda,\Lambda_{L}(x)}-E)^{-1}&=(H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime})}-E)^{-1}[H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime}),},\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}](H_{\lambda,\Lambda_{L}(x)}-E)^{-1}\\ &\quad+(H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime})}-E)^{-1}\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}.\end{split}

    Multiplying this equation from the left by χΛl′′(y)\chi_{\Lambda_{l^{\prime\prime}}(y)} and from the right by ΞΛL(x)\Xi_{\Lambda_{L}(x)}, the SLI ready follows from the boundedness of [Hλ,Λl(y),,χ~Λl(y)][H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime}),},\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}] and submultiplicativity of the operator norm, as χ~Λl(y)ΞΛL(x)=0\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}\Xi_{\Lambda_{L}(x)}=0 implies that the second term on the right vanishes and

    [Hλ,Λl(y),,χ~Λl(y)]=ΞΛl(y)[Hλ,Λl(y),,χ~Λl(y)]ΞΛl(y).[H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime}),},\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}]=\Xi_{\Lambda_{l^{\prime}}(y^{\prime})}[H_{\lambda,\Lambda_{l^{\prime}}(y^{\prime}),},\tilde{\chi}_{\Lambda_{l^{\prime}}(y^{\prime})}]\Xi_{\Lambda_{l^{\prime}}(y^{\prime})}. (4.1)
  2. (2)

    For the proof of the EDI, it suffices to choose ψ\psi as in the Lemma and observe the resolvent identity (Hλ,x,LE)1[Hλ,χ~ΛL(x)]ψ=χ~ΛL(x)ψ(H_{\lambda,x,L}-E)^{-1}[H_{\lambda},\tilde{\chi}_{\Lambda_{L}(x)}]\psi=\tilde{\chi}_{\Lambda_{L}(x)}\psi which is easily verified by using that (VωVω,ΛL(x))χ~ΛL(x)=0(V_{\omega}-V_{\omega,\Lambda_{L}(x)})\tilde{\chi}_{\Lambda_{L}(x)}=0 as well as Hλψ=Eψ.H_{\lambda}\psi=E\psi. Using then an analogue of (4.1), [Hλ,χ~ΛL(x)]=ΞΛL(x)[Hλ,χ~ΛL(x)]ΞΛL(x),[H_{\lambda},\tilde{\chi}_{\Lambda_{L}(x)}]=\Xi_{\Lambda_{L}(x)}[H_{\lambda},\tilde{\chi}_{\Lambda_{L}(x)}]\Xi_{\Lambda_{L}(x)}, together with the boundedness of the commutator shows the claim.

We complete our preparations by discussing the estimate on the number of eigenvalues (NE) and the Wegner estimate (W). The estimate on the number of eigenvalues (NE) is the estimate stated in Proposition 1.3. The Wegner estimate is then obtained by applying the estimate in Proposition 1.3 to the last expression in this set of inequalities

𝐏(d(Spec(Hλ,ΛL),E)<η)=𝐏(rank1l(Eη,E+η)(Hλ,ΛL)1)𝐄(tr(1l(Eη,E+η)(Hλ,ΛL))).\begin{split}\mathbf{P}(d(\operatorname{Spec}(H_{\lambda,\Lambda_{L}}),E)<\eta)&=\mathbf{P}(\operatorname{rank}\operatorname{1\hskip-2.75ptl}_{(E-\eta,E+\eta)}(H_{\lambda,\Lambda_{L}})\geq 1)\\ &\leq\mathbf{E}(\operatorname{tr}(\operatorname{1\hskip-2.75ptl}_{(E-\eta,E+\eta)}(H_{\lambda,\Lambda_{L}}))).\end{split} (4.2)

4.2. Dynamical delocalization

In this subsection we prove Theorem 3. To imitate the proof of delocalization in [GKS07], we shall study the third power of the random Hamiltonian (1.11), since H3(M)L2(M)H^{3}(M)\hookrightarrow L^{2}(M), for MM a two-dimensional compact manifold, is a trace-class embedding333Recall that λkMk\lambda_{k}\sim_{M}k is the Weyl asymptotics of the negative Laplacian in dimension 2; thus kk3/2<\sum_{k}k^{-3/2}<\infty and xx3x\mapsto x^{3} is bijective, by defining

Sλ:=Hλ3.S_{\lambda}:=H_{\lambda}^{3}.

Let 𝒞±:=B|λ|supωΩVω(±m)\mathcal{C}_{\pm}:=\partial B_{|\lambda|\sup_{\omega\in\Omega}\|V_{\omega}\|_{\infty}}(\pm m) such that 𝒞±\mathcal{C}_{\pm} encircles the spectrum of the random perturbation of a single flat band, but nothing else (if m=0m=0, then 𝒞±\mathcal{C}_{\pm} both coincide, we shall explain the modifications of this case at the end of this section). This is possible for sufficiently small noise λ>0\lambda>0 as the flat band at energies ±m\pm m are strictly gapped (1.8) from all higher bands, in the absence of disorder. We then define the L2(;4)L^{2}({\mathbb{C}};\mathbb{C}^{4})-valued spectral projection

Pλ,±=12πi𝒞±3(Sλz)1𝑑z,P_{\lambda,\pm}=-\frac{1}{2\pi i}\int_{\mathcal{C}_{\pm}^{3}}(S_{\lambda}-z)^{-1}\ dz, (4.3)

where by 𝒞±3\mathcal{C}_{\pm}^{3} we just mean the set of elements in 𝒞±\mathcal{C}_{\pm} raised to the third power. The delocalization argument rests on the following two pillars:

  • If the random Hamiltonian exhibits only dynamical localization close to ±m\pm m, then this implies that the partial Chern numbers of Pλ,±P_{\lambda,\pm}, defined in section B, have to vanish, see Prop. B.2.

  • The partial Chern numbers of Pλ,±P_{\lambda,\pm} are invariant under disorder as well as small perturbations in α\alpha away from perfect magic angles.

As a consequence, the Hamiltonian exhibits dynamical delocalization at energies close to ±m\pm m. To simplify the notation, we drop the ±\pm and just focus on +m+m, since m-m can be treated analogously.

The central object in this discussion is the Hall conductance. Assuming P[[P,Θ1],[P,Θ2]]1<\|P[[P,\Theta_{1}],[P,\Theta_{2}]]\|_{1}<\infty for a spectral projection PP and multiplication operators Θ1(z):=1l[1/2,)(Rez)\Theta_{1}(z):=\operatorname{1\hskip-2.75ptl}_{[1/2,\infty)}(\operatorname{Re}z) and Θ2(z):=1l[1/2,)(Imz)\Theta_{2}(z):=\operatorname{1\hskip-2.75ptl}_{[1/2,\infty)}(\operatorname{Im}z) the Hall conductance is defined by

Ω(P):=tr(P[[P,Θ1],[P,Θ2]])=tr([PΘ1P,PΘ2P]).\Omega(P):=\operatorname{tr}(P[[P,\Theta_{1}],[P,\Theta_{2}]])=\operatorname{tr}([P\Theta_{1}P,P\Theta_{2}P]). (4.4)

Here, κ=i[PΘ1P,PΘ2P]\kappa=-i[P\Theta_{1}P,P\Theta_{2}P] is also called the adiabatic curvature with Hall charge transport Q=2πtr(κ).Q=-2\pi\operatorname{tr}(\kappa). That QQ is an integer is shown for example in [ASS94, Theorem 8.28.2] or [BES94, (49),(58)] where it is related to Chern characters and Fredholm indices, respectively, and then [BES94, Theorem 11] where this quantity is discussed for periodic and quasi-periodic operators.

Proof of Theo. 3.

Since H3(M)L2(M)extensionL2()H^{3}(M)\hookrightarrow L^{2}(M)\xrightarrow[]{\text{extension}}L^{2}({\mathbb{C}}) is a trace-class embedding, for bounded open sets MM, it follows that there is a universal constant K1>0K_{1}>0 such that for sufficiently small disorder λ\lambda and μ𝒞3\mu\in\mathcal{C}^{3} with 𝒞3\mathcal{C}^{3} as above in trace norm

(Sλμ)1χz1K1 for all zΓ.\|(S_{\lambda}-\mu)^{-1}\chi_{z}\|_{1}\leq K_{1}\text{ for all }z\in\Gamma. (4.5)

Next, we are going to construct an analogue of the Combes-Thomas estimate (CTE) for the operator SλS_{\lambda}:

By conjugating the operator SλS_{\lambda} with efe^{f} where ff is some smooth function, we find

efSλef=Sλ+Rf,e^{f}S_{\lambda}e^{-f}=S_{\lambda}+R_{f},

where

RfL(H3,L2)ε if βfε1 for all 1|β|3.\|R_{f}\|_{L(H^{3},L^{2})}\lesssim\varepsilon\text{ if }\|\partial^{\beta}f\|_{\infty}\leq\varepsilon\ll 1\text{ for all }1\leq|\beta|\leq 3.

This implies that for zSpec(Sλ)z\notin\operatorname{Spec}(S_{\lambda})

ef(Sλz)ef=(id+Rf(Sλz)1)(Sλz).e^{f}(S_{\lambda}-z)e^{-f}=(\operatorname{id}+R_{f}(S_{\lambda}-z)^{-1})(S_{\lambda}-z).

Thus, for zSpec(Sλ)z\notin\operatorname{Spec}(S_{\lambda}) and ε>0\varepsilon>0 sufficiently small such that Rf(Sλz)1<1\|R_{f}(S_{\lambda}-z)^{-1}\|<1,

ef(Sλz)1efL(L2,H3)=𝒪(d(Spec(Sλ),z)1).\|e^{-f}(S_{\lambda}-z)^{-1}e^{f}\|_{L(L^{2},H^{3})}=\mathcal{O}(\langle d(\operatorname{Spec}(S_{\lambda}),z)^{-1}\rangle).

We conclude that for f(z):=εzw0f(z):=\varepsilon\langle z-w_{0}\rangle with w0w_{0}\in{\mathbb{C}} fixed, we have for all ww\in{\mathbb{C}}

χw0(Sλz)1χw=χw0ef(ef(Sλz)1ef)efχw=𝒪(eεww0d(Spec(Sλ),z)),\begin{split}\|\chi_{w_{0}}(S_{\lambda}-z)^{-1}\chi_{w}\|&=\|\chi_{w_{0}}e^{f}(e^{-f}(S_{\lambda}-z)^{-1}e^{f})e^{-f}\chi_{w}\|=\mathcal{O}\Big{(}\tfrac{e^{-\varepsilon\langle w-w_{0}\rangle}}{d(\operatorname{Spec}(S_{\lambda}),z)}\Big{)},\end{split} (CTE)

as well as

χw0(SλS0)(Sλz)1χw=χw0efef(SλS0)efL(H3,L2)×ef(Sλz)1efL(L2,H3)efχw=𝒪(efχwd(Spec(Sλ),z))=𝒪(eεww0d(Spec(Sλ),z)).\begin{split}\|\chi_{w_{0}}(S_{\lambda}-S_{0})(S_{\lambda}-z)^{-1}\chi_{w}\|&=\|\chi_{w_{0}}e^{f}\|\|e^{-f}(S_{\lambda}-S_{0})e^{f}\|_{L(H^{3},L^{2})}\\ &\quad\times\|e^{-f}(S_{\lambda}-z)^{-1}e^{f}\|_{L(L^{2},H^{3})}\|e^{-f}\chi_{w}\|\\ &=\mathcal{O}(\tfrac{\|e^{-f}\chi_{w}\|}{d(\operatorname{Spec}(S_{\lambda}),z)})=\mathcal{O}\Big{(}\tfrac{e^{-\varepsilon\langle w-w_{0}\rangle}}{d(\operatorname{Spec}(S_{\lambda}),z)}\Big{)}.\end{split} (4.6)

From the Combes-Thomas estimate (CTE) and (4.3) we find the exponential estimate

χw0Pλχweε|ww0|.\|\chi_{w_{0}}P_{\lambda}\chi_{w}\|\lesssim e^{-\varepsilon|w-w_{0}|}. (4.7)

By [GKS07, Lemma 3.13.1], this implies that

Pλ[[Pλ,Θ1],[Pλ,Θ2]]1<,\|P_{\lambda}[[P_{\lambda},\Theta_{1}],[P_{\lambda},\Theta_{2}]]\|_{1}<\infty,

which implies that the Hall conductance is well-defined. In fact, using (4.5) we have

χwPλχw01=𝒪(1) and χwPλχw022χwPλχw01χwPλχw0=𝒪(eε|ww0|).\begin{split}\|\chi_{w}P_{\lambda}\chi_{w_{0}}\|_{1}=\mathcal{O}(1)\text{ and }\|\chi_{w}P_{\lambda}\chi_{w_{0}}\|^{2}_{2}\leq\|\chi_{w}P_{\lambda}\chi_{w_{0}}\|_{1}\|\chi_{w}P_{\lambda}\chi_{w_{0}}\|=\mathcal{O}(e^{-\varepsilon|w-w_{0}|}).\end{split} (4.8)

To obtain the invariance of the Chern number under small disorder, we now define

Qλ,ζ:=PζPλ=ζλ2πi𝒞3(Sλz)1(SζSλ)(ζλ)(Sζz)1𝑑z.Q_{\lambda,\zeta}:=P_{\zeta}-P_{\lambda}=\frac{\zeta-\lambda}{2\pi i}\int_{\mathcal{C}^{3}}(S_{\lambda}-z)^{-1}\frac{(S_{\zeta}-S_{\lambda})}{(\zeta-\lambda)}(S_{\zeta}-z)^{-1}\ dz. (4.9)

then by (4.8) we find

χwQλ,ζχw022=𝒪(eε|ww0|).\|\chi_{w}Q_{\lambda,\zeta}\chi_{w_{0}}\|_{2}^{2}=\mathcal{O}(e^{-\varepsilon|w-w_{0}|}). (4.10)

If the random potential has compact support, i.e. HλH_{\lambda} in (1.11) is replaced by

Hλ(L)=H+λVω where Vω=γΛ~Lωγu(γξγ),H_{\lambda}(L)=H+\lambda V_{\omega}\text{ where }V_{\omega}=\sum_{\gamma\in\tilde{\Lambda}_{L}}\omega_{\gamma}u(\bullet-\gamma-\xi_{\gamma}), (4.11)

for some L>0L>0, then by using a partition of unity and (4.5), we find Qλ,ζ1<\|Q_{\lambda,\zeta}\|_{1}<\infty and consequently the traces of all commutators vanish

Ω(Pζ)Ω(Pλ)=tr([Qλ,ζΘ1Pζ,PζΘ2Pζ]+[PλΘ1Qλ,ζ,PζΘ2Pζ]+[PλΘ1Pλ,Qλ,ζΘ2Pζ]+[PλΘ1Pλ,PλΘ2Qλ,ζ])=0.\begin{split}\Omega(P_{\zeta})-\Omega(P_{\lambda})=&\operatorname{tr}([Q_{\lambda,\zeta}\Theta_{1}P_{\zeta},P_{\zeta}\Theta_{2}P_{\zeta}]+[P_{\lambda}\Theta_{1}Q_{\lambda,\zeta},P_{\zeta}\Theta_{2}P_{\zeta}]\\ &+[P_{\lambda}\Theta_{1}P_{\lambda},Q_{\lambda,\zeta}\Theta_{2}P_{\zeta}]+[P_{\lambda}\Theta_{1}P_{\lambda},P_{\lambda}\Theta_{2}Q_{\lambda,\zeta}])=0.\end{split} (4.12)

So the integer-valued map λΩ(Pλ)\lambda\mapsto\Omega(P_{\lambda}) is constant for λ\lambda small around zero, under the assumption of a compactly supported random potential in (4.11).

It remains now to drop the compact support constraint on the random potential in (4.11). Let Sλ(L)=Hλ(L)3S_{\lambda}(L)=H_{\lambda}(L)^{3}, then we define

Qλ,>L:=PλPλ(L)=λ2πi𝒞3(Sλz)1(SλSλ(L))(Sλ(L)z)1𝑑z,Q_{\lambda,>L}:=P_{\lambda}-P_{\lambda}(L)=\frac{\lambda}{2\pi i}\int_{\mathcal{C}^{3}}(S_{\lambda}-z)^{-1}(S_{\lambda}-S_{\lambda}(L))(S_{\lambda}(L)-z)^{-1}\ dz, (4.13)

where Pλ(L)P_{\lambda}(L) is the corresponding spectral projection associated with Sλ(L).S_{\lambda}(L). By the Combes-Thomas estimates (CTE) and the resolvent identity (4.13), we find

χwQλ,>Lχw0ecε((LR|w|)++(LR|w0|)++|ww0|)\|\chi_{w}Q_{\lambda,>L}\chi_{w_{0}}\|\lesssim e^{-c\varepsilon((L-R-|w|)_{+}+(L-R-|w_{0}|)_{+}+|w-w_{0}|)}

for some c>0c>0, where we used that SλSλ(L)S_{\lambda}-S_{\lambda}(L) is zero on ΛLR(0).\Lambda_{L-R}(0). Thus, writing the difference of Hall conductivities yields the desired limit

Ω(Pλ)Ω(Pλ(L))=tr(Qλ,>L[[Pλ,Θ1],[Pζ,Θ2]]+Pλ,L[[Qλ,>L,Θ1],[Pλ,Θ2]]+Pλ,L[[Pλ,Θ1],[Qλ,>L,Θ2]])0 as L.\begin{split}\Omega(P_{\lambda})-\Omega(P_{\lambda}(L))=&\operatorname{tr}(Q_{\lambda,>L}[[P_{\lambda},\Theta_{1}],[P_{\zeta},\Theta_{2}]]+P_{\lambda,L}[[Q_{\lambda,>L},\Theta_{1}],[P_{\lambda},\Theta_{2}]]\\ &+P_{\lambda,L}[[P_{\lambda},\Theta_{1}],[Q_{\lambda,>L},\Theta_{2}]])\to 0\text{ as }L\to\infty.\end{split} (4.14)

Here, one uses the strong limit slimLQλ,>L=0s-\lim_{L\to\infty}Q_{\lambda,>L}=0 to show the non-vanishing of the first term on the right-hand side in (4.14) and that

|tr(Pλ,L[[Qλ,>L,Θ1],[Pλ,Θ2]])|2γ,γΓχγ[Qλ,>L,Θ1]χγ2χγ[Pλ,Θ2]χγ2|\operatorname{tr}(P_{\lambda,L}[[Q_{\lambda,>L},\Theta_{1}],[P_{\lambda},\Theta_{2}]])|\leq 2\sum_{\gamma,\gamma^{\prime}\in\Gamma}\|\chi_{\gamma}[Q_{\lambda,>L},\Theta_{1}]\chi_{\gamma^{\prime}}\|_{2}\|\chi_{\gamma^{\prime}}[P_{\lambda},\Theta_{2}]\chi_{\gamma}\|_{2}

with a similar estimate for the last term in (4.14). The last bound converges to zero for LL\to\infty by using (4.8) and (4.10), see [GKS07, Lemma 3.1 (i)] for details. Thus, the conductivity derived from PλP_{\lambda} is locally constant in λ\lambda and α\alpha, see (4.12), which shows using (1.10) that Chern numbers stay ±1\pm 1, for m>0m>0, respectively.

For m=0m=0 we repeat the previous computation with our modified Ωi\Omega_{i} (B.3) to arrive at the same conclusion. Thus, if, in the notation of (1.14), Σ(K,K)ΣDL\Sigma\cap(-K_{-},K_{-})\subset\Sigma^{\operatorname{DL}} then this would contradict the non-vanishing of the (partial) Chern number, see (B.6), in regions of full localization as shown in Prop. B.2.

The bound in the statement of Theorem 3 follows then from [GK04, Theo 2.102.10].

4.3. Dynamical localization

Working under assumptions (1), we shall now study the localized phase of the Anderson model of the form

Hλ=H+Vω where Vω=γΓωγu(γξγ).H_{\lambda}=H+V_{\omega}\text{ where }V_{\omega}=\sum_{\gamma\in\Gamma}\omega_{\gamma}u(\bullet-\gamma-\xi_{\gamma}). (4.15)

Here we got rid of the λ\lambda parameter but instead consider random variables ωγ\omega_{\gamma} that are distributed according to a bounded density gλg_{\lambda} of compact support in [δ,δ][-\delta,\delta] with δ<min(m,Egap)\delta<\min(m,E_{\operatorname{gap}}) for m>0m>0 and δ<Egap\delta<E_{\operatorname{gap}} for m=0m=0. Here gλg_{\lambda} is a rescaled distribution gλ(u)=cλg(u/λ)/λ1l[δ,δ]g_{\lambda}(u)=c_{\lambda}g(u/\lambda)/\lambda\operatorname{1\hskip-2.75ptl}_{[-\delta,\delta]}, with g>0g>0, such that as λ0\lambda\downarrow 0 the mass gets concentrated near zero and cλCc_{\lambda}\leq C, uniformly in λ\lambda, is the normalizing constant.

By (1.13), with probability 11, the spectrum Σ\Sigma is independent of λ.\lambda. Our next theorem shows that the mobility edges can be shown to be located arbitrarily close to the original flat bands, by choosing λ\lambda small, while keeping the support of the disorder fixed, within the interval [δ,δ][-\delta,\delta] which is the motivation for our modifications of the Hamiltonian.

Theorem 7 (Mobility edge).

Let γg\langle\bullet\rangle^{\gamma}g be bounded for some γ>3\gamma>3 and let τ(0,γ3γ+1)\tau\in(0,\tfrac{\gamma-3}{\gamma+1}). Let HλH_{\lambda} be as in Assumption 1 with the modification that λ\lambda is incorporated in the rescaled density, as described in (4.15) and DD\subset\mathbb{C} small enough. Then for any m>0m>0 there exist at least two distinct dynamical mobility edges, denoted by +(λ)>(λ)\mathscr{E}_{+}(\lambda)>\mathscr{E}_{-}(\lambda) such that

|+(λ)m|+|(λ)+m|λ14γ+1τλ00.\left\lvert\mathscr{E}_{+}(\lambda)-m\right\rvert+\left\lvert\mathscr{E}_{-}(\lambda)+m\right\rvert\lesssim\lambda^{1-\frac{4}{\gamma+1}-\tau}\xrightarrow[\lambda\downarrow 0]{}0.

In particular,

{E(Egap2/2+m2,Egap2/2+m2);|E±m|λ14γ+1τ}ΣDL,\left\{E\in\Big{(}-\sqrt{E_{\operatorname{gap}}^{2}/2+m^{2}},\sqrt{E_{\operatorname{gap}}^{2}/2+m^{2}}\Big{)};|E\pm m|\gtrsim\lambda^{1-\frac{4}{\gamma+1}-\tau}\right\}\subset\Sigma^{\operatorname{DL}},

where the region of dynamical localization ΣDL\Sigma^{\operatorname{DL}} has been defined in (1.18). In the case of m=0m=0 the same result holds but with only at least one guaranteed mobility edge.

Proof.

We start by observing that using the LL^{\infty} bound on γg\langle\bullet\rangle^{\gamma}g, we have for any ε>0\varepsilon>0 and XgλX\sim g_{\lambda}

𝐏(|X|ε)=δ|x|εgλ(x)𝑑xδ/λ|x|ε/λg(x)𝑑xλ/εγ1.\begin{split}\mathbf{P}(|X|\geq\varepsilon)&=\int_{\delta\geq|x|\geq\varepsilon}g_{\lambda}(x)\ dx\lesssim\int_{\delta/\lambda\geq|x|\geq\varepsilon/\lambda}g(x)\ dx\lesssim\langle\lambda/\varepsilon\rangle^{\gamma-1}.\end{split} (4.16)

Thus, for the probability of the low-lying spectrum to be contained in a small interval [ε,ε][-\varepsilon,\varepsilon], we find for L01L_{0}\gg 1 fixed

𝐏(Spec(Hλ,ΛL0)(Egap2/2+m2,Egap2/2+m2)±m+[ε,ε])union bound𝐏(|ωγ|ε/2;γΛ~L0+R(x))(4.16)(1C(λ/ε)γ1)(L0+R)2Bernoulli1C(λ/ε)γ1L02 for small λ/ε,\begin{split}\mathbf{P}&\Big{(}\operatorname{Spec}(H_{\lambda,\Lambda_{L_{0}}})\cap\Big{(}-\sqrt{E_{\operatorname{gap}}^{2}/2+m^{2}},\sqrt{E_{\operatorname{gap}}^{2}/2+m^{2}}\Big{)}\subset\pm m+[-\varepsilon,\varepsilon]\Big{)}\\ \overset{\text{union bound}}{\geq}&\mathbf{P}(|\omega_{\gamma}|\leq\varepsilon/2;\gamma\in\tilde{\Lambda}_{L_{0}+R}(x))\\ \overset{\eqref{eq:flambda}}{\geq}&(1-C(\lambda/\varepsilon)^{\gamma-1})^{(L_{0}+R)^{2}}\\ \overset{\text{Bernoulli}}{\geq}&1-C(\lambda/\varepsilon)^{\gamma-1}L_{0}^{2}\text{ for small }\lambda/\varepsilon,\end{split}

for small enough λ/ε\lambda/\varepsilon. We recall that R>0R>0 is such that suppuΛR(0).\operatorname{supp}u\subset\Lambda_{R}(0). This probability is large, if we choose

ε=CλL02γ1\varepsilon=C\lambda L_{0}^{\frac{2}{\gamma-1}} (4.17)

for C1.C\gg 1. To prove localization, one chooses L01L_{0}\gg 1 large enough, as specified in [GK03, (2.16)] and λ1\lambda\ll 1. We now fix an energy Egap2/2+m2|E|\sqrt{E_{\operatorname{gap}}^{2}/2+m^{2}}\geq|E| such that |E±m|2ε|E\pm m|\geq 2\varepsilon with EΣE\in\Sigma . Then EE is, with high probability, a distance ε>0\varepsilon>0 away from the spectrum of the finite-size Hamiltonian Hλ,ΛL0H_{\lambda,\Lambda_{L_{0}}}.

In order to show localization, we shall satisfy the finite-size criterion of [GK03, Theorem 2.42.4]. This will give us another condition aside from (4.17). Indeed, in our setting the finite-size criterion stated in [GK03, Theorem 2.42.4] takes the following form

C1L025/3λεeC2εL0<1\frac{C_{1}L_{0}^{25/3}}{\lambda\varepsilon}e^{-C_{2}\varepsilon L_{0}}<1 (4.18)

for two constants C1,C2>0.C_{1},C_{2}>0. The term L025/3L_{0}^{25/3} is obtained from [GK03, Theorem 2.42.4] by choosing (in the notation of [GK03]) b=1,d=2b=1,d=2, and performing a union bound over a partition of Γ0\Gamma_{0} and χ0,L0/3\chi_{0,L_{0}/3} which accounts for another L03.L_{0}^{3}. The λ\lambda in the denominator is due to the scaling of the constant in the Wegner estimate which for us is proportional to the supremum norm of the density, which for us is gλ=𝒪(1/λ)\|g_{\lambda}\|_{\infty}=\mathcal{O}(1/\lambda).

By [GK03, Theorem 2.42.4] one concludes localization if both (4.17) and (4.18) hold.

Setting then ε:=C3λL02γ1\varepsilon:=C_{3}\lambda L_{0}^{\frac{2}{\gamma-1}} with C3>0C_{3}>0 sufficiently large to satisfy (4.17), we find that (4.18) becomes

C1L025/3C3λ2L02γ1eC2C3λ2L0γ+1γ1<1.\frac{C_{1}L_{0}^{25/3}}{C_{3}\lambda^{2}L_{0}^{\frac{2}{\gamma-1}}}e^{-C_{2}C_{3}\lambda^{2}L_{0}^{\frac{\gamma+1}{\gamma-1}}}<1.

We now also set L02γ1=λ4γ+1τL_{0}^{\frac{2}{\gamma-1}}=\lambda^{-\frac{4}{\gamma+1}-\tau} with τ(γ)>0\tau(\gamma)>0 small such that 4γ+1τ>1.-\frac{4}{\gamma+1}-\tau>-1. This means that L0γ+1γ1=λ2τγ+12L_{0}^{\frac{\gamma+1}{\gamma-1}}=\lambda^{-2-\tau\frac{\gamma+1}{2}} which implies that for λ\lambda small enough, (4.18) holds as well.

The characterization of the localized regime then follows from [GK03, Theorem 2.42.4], the existence of a mobility edge follows together with Theorem 3. ∎

5. Decay of point spectrum and Wannier bases

We now give the proof of Theo. 4 and 5. We shall mainly focus on the first case and only explain the modifications for the second result at the very end.

Proof of Theo.4 & 5.

We first reduce the analysis to λ=0.\lambda=0. By λ\lambda-continuity of the random perturbation, the spectral projections P0=1lJ±(H0)P_{0}=\operatorname{1\hskip-2.75ptl}_{J_{\pm}}(H_{0}) and Pλ=1lJ±(Hλ)P_{\lambda}=\operatorname{1\hskip-2.75ptl}_{J_{\pm}}(H_{\lambda}) with J±J_{\pm} as in (1.15)

P0Pλ=𝒪(|λ|)\|P_{0}-P_{\lambda}\|=\mathcal{O}(|\lambda|)

by using e.g. the resolvent identity and holomorphic functional calculus and the spectral gap of the Hamiltonian. Thus, for λ\lambda small enough there is an isometry [BES94, Lemma 1010] UU such that UU=P0UU^{*}=P_{0} and UU=Pλ.U^{*}U=P_{\lambda}. In particular P0U=UPλ.P_{0}U=UP_{\lambda}. It then follows that UU has a Schwartz kernel KK that is exponentially close to the identity, cf. [CMM19, Lemma 8.58.5]. By this we mean that there is γ>0\gamma>0 such that

|K(z,z)1|=𝒪(eγ|zz|).|K(z,z^{\prime})-1|=\mathcal{O}(e^{-\gamma|z-z^{\prime}|}).

The Schur test for integral operators implies that U~:=z0Uz01\tilde{U}:=\langle\bullet-z_{0}\rangle U\langle\bullet-z_{0}\rangle^{-1} is a family of operators uniformly bounded in z0z_{0}\in{\mathbb{C}}. This implies that for any φL2(;4)\varphi\in L^{2}({\mathbb{C}};{\mathbb{C}}^{4})

z01+δP0Uφ=z01+δUz01δz01+δPλφ.\langle\bullet-z_{0}\rangle^{1+\delta}P_{0}U\varphi=\langle\bullet-z_{0}\rangle^{1+\delta}U\langle\bullet-z_{0}\rangle^{-1-\delta}\langle\bullet-z_{0}\rangle^{1+\delta}P_{\lambda}\varphi.

Taking norms, we find, using that z01+δPλφ<\|\langle\bullet-z_{0}\rangle^{1+\delta}P_{\lambda}\varphi\|<\infty by assumption, that

z01+δP0Uφ<.\|\langle\bullet-z_{0}\rangle^{1+\delta}P_{0}U\varphi\|<\infty.

This implies, by choosing for UφU\varphi an orthonormal basis of ran¯(P0)\overline{\operatorname{ran}}(P_{0}), i.e. (ψn)(\psi_{n}) is an orthonormal basis of ran¯(P0)\overline{\operatorname{ran}}(P_{0}), then φn:=Uψn\varphi_{n}:=U^{*}\psi_{n}, that P0P_{0} exhibits a (1+δ)(1+\delta)-localized generalized Wannier basis. Since P0P_{0} is precisely the projection onto ker(D(α))\ker(D(\alpha)), we deduce that P0P_{0} exhibits a non-zero Chern number, see (1.10), and therefore do not possess a (1+δ)(1+\delta)-localized Wannier basis, see [LS21] which gives a contradiction.

Conversely, let Pk(α)=(2πi)1γ(zHk(0,α))1𝑑zP_{k}(\alpha)=(2\pi i)^{-1}\oint_{\gamma}(z-H_{k}(0,\alpha))^{-1}\ dz, where γ\gamma is a sufficiently small circle around zero encircling only the flat band eigenvalue but nothing else in the spectrum of Hk(0,α)H_{k}(0,\alpha). Then Pk(α)P_{k}(\alpha) is the spectral projection onto the flat band eigenfunction of HkH_{k}. Since kHkk\mapsto H_{k} is real-analytic, this implies that kPkk\mapsto P_{k} is real-analytic. Moreover, since Hkγ(α)=τ(γ)Hk(α)τ(γ)1H_{k-\gamma^{*}}(\alpha)=\tau(\gamma^{*})H_{k}(\alpha)\tau(\gamma^{*})^{-1} with τγ(z):=eiRe(zγ¯)\tau_{\gamma}(z):=e^{i\operatorname{Re}(z\overline{\gamma^{*}})} with γΓ3,\gamma^{*}\in\Gamma_{3}^{*}, the spectral projection satisfies the covariance relation

Pkγ(α)=τ(γ)Pk(α)τ(γ)1.P_{k-\gamma}(\alpha)=\tau(\gamma^{*})P_{k}(\alpha)\tau(\gamma^{*})^{-1}.

It then follows from [MPPT18, Theo. 2.42.4] that there exists an associated Wannier basis satisfying p/2wγL2()2C<\|\langle\bullet\rangle^{p/2}w_{\gamma}\|^{2}_{L^{2}({\mathbb{C}})}\leq C<\infty for p<1p<1 and all γΓ\gamma\in\Gamma for the unperturbed periodic problem. Reversing the argument provided in the first part of the proof, it follows that the randomly perturbed problem also exhibits such a Wannier basis.

To show Theorem 5 one proceeds analogously and notices that P±,λ=0P_{\pm,\lambda=0} correspond to the projections onto ker(D(α))\ker(D(\alpha)) and ker(D(α))\ker(D(\alpha)^{*}), each one exhibiting a non-zero Chern number. ∎

With this result at hand, we are able to evaluate the quantity (1.17) for the unperturbed Hamiltonian providing a link between the dynamical and spectral theoretic notion of (de)-localization.

Proposition 5.1.

Let α\alpha be a simple magic angle, as in Def. 1.1, then for all p1p\geq 1

p/2eitH(α)Pker(H(α))1l/Γ322=,\left\lVert\langle\bullet\rangle^{p/2}e^{-itH(\alpha)}P_{\ker(H(\alpha))}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}=\infty,

while the left-hand side is finite for p<1.p<1.

Proof.

We start by observing that for an orthonormal basis (fn)(f_{n}) of L2(/Γ3)L^{2}({\mathbb{C}}/\Gamma_{3}) and (ei)(e_{i}) the standard basis of 4{\mathbb{C}}^{4}

p/2eitH(α)Pker(H(α))1l/Γ322=p/2Pker(H(α))1l/Γ322=p/2Pker(D(α))ker(D(α))1l/Γ322=i=14np/2Pker(D(α))ker(D(α))fnei2=p/2Pker(D(α))1l/Γ322+p/2Pker(D(α))1l/Γ322.\begin{split}\left\lVert\langle\bullet\rangle^{p/2}e^{-itH(\alpha)}P_{\ker(H(\alpha))}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}&=\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(H(\alpha))}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}\\ &=\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha))\oplus\ker(D(\alpha)^{*})}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}\\ &=\sum_{i=1}^{4}\sum_{n\in\mathbb{N}}\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha))\oplus\ker(D(\alpha)^{*})}f_{n}\otimes e_{i}\right\rVert^{2}\\ &=\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha))}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}+\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha)^{*})}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}.\end{split}

Without loss of generality, we shall focus on the first summand. Consider the unitary Bloch-Floquet transform u(z,k):=γΓ3eiz+γ,kγu(z)\mathcal{B}u(z,k):=\sum_{\gamma\in\Gamma_{3}}e^{i\langle z+\gamma,k\rangle}\mathscr{L}_{\gamma}u(z), where LγL_{\gamma} has been defined in (1.4), with the convention that z,z0:=Re(zz¯0),\langle z,z_{0}\rangle:=\operatorname{Re}(z\bar{z}_{0}), and its inverse/adjoint 𝒞v(z):=/Γ3v(z,k)eiz,kdm(k)|/Γ3|\mathcal{C}v(z):=\int_{{\mathbb{C}}/\Gamma_{3}^{*}}v(z,k)e^{-i\langle z,k\rangle}\ \frac{dm(k)}{|{\mathbb{C}}/\Gamma_{3}^{*}|}. We then find that

γ𝒞v(z):=/Γ3v(z,k)eiz+γ,kdm(k)|/Γ3|=𝒞(eiγ,kv(z,k)).\mathscr{L}_{\gamma}\mathcal{C}v(z):=\int_{{\mathbb{C}}/\Gamma_{3}^{*}}v(z,k)e^{-i\langle z+\gamma,k\rangle}\ \frac{dm(k)}{|{\mathbb{C}}/\Gamma_{3}^{*}|}=\mathcal{C}(e^{-i\langle\gamma,k\rangle}v(z,k)). (5.1)

Since by assumption ker(D(α)+k)=span{φ(,k)}\ker(D(\alpha)+k)=\operatorname{span}\{\varphi(\bullet,k)\}, we see that

(eiγ,kφ(z,k))γΓ, for φ(,k)L2(/Γ3) normalized,(e^{-i\langle\gamma,k\rangle}\varphi(z,k))_{\gamma\in\Gamma}\text{, for }\varphi(\bullet,k)\in L^{2}({\mathbb{C}}/\Gamma_{3})\text{ normalized,} (5.2)

forms a basis of the space /Γ3ker(D(α)+k)𝑑k\int^{\oplus}_{{\mathbb{C}}/\Gamma_{3}^{*}}\ker(D(\alpha)+k)dk. Indeed, orthonormality just follows from

eiγ,kφ(z,k),eiγ,kφ(z,k)=/Γ3/Γ3|φ(z,k)|2eiγγ,kdzdk|/Γ3|=/Γ3eiγγ,kdk|/Γ3|=δγ,γ\begin{split}\langle e^{-i\langle\gamma,k\rangle}\varphi(z,k),e^{i\langle\gamma^{\prime},k\rangle}\varphi(z,k)\rangle&=\int_{{\mathbb{C}}/\Gamma_{3}^{*}}\int_{{\mathbb{C}}/\Gamma_{3}}|\varphi(z,k)|^{2}e^{-i\langle\gamma-\gamma^{\prime},k\rangle}\frac{dz\ dk}{|{\mathbb{C}}/\Gamma_{3}^{*}|}\\ &=\int_{{\mathbb{C}}/\Gamma_{3}^{*}}e^{-i\langle\gamma-\gamma^{\prime},k\rangle}\frac{dk}{|{\mathbb{C}}/\Gamma_{3}^{*}|}=\delta_{\gamma,\gamma^{\prime}}\end{split} (5.3)

and completeness from the completeness of the regular Fourier expansion, i.e. a general element in this subspace is of the form

γΓ3f(γ)eiγ,kv(z,k) for f2(Γ3).\sum_{\gamma\in\Gamma_{3}^{*}}f(\gamma)e^{-i\langle\gamma,k\rangle}v(z,k)\text{ for }f\in\ell^{2}(\Gamma_{3}^{*}).

We then have that D(α)𝒞φ(x,k)=(D(α)+k)φ(x,k).\mathcal{B}D(\alpha)\mathcal{C}\varphi(x,k)=(D(\alpha)+k)\varphi(x,k). Recall the trivial decomposition of L2L^{2} given by L2()=L2(/Γ3)L2((/Γ3)).L^{2}({\mathbb{C}})=L^{2}({\mathbb{C}}/\Gamma_{3})\oplus L^{2}({\mathbb{C}}\setminus({\mathbb{C}}/\Gamma_{3})).

We then find for the Hilbert-Schmidt norm using an orthonormal basis (en)(e_{n}) of L2(/Γ3)L^{2}({\mathbb{C}}/\Gamma_{3})

p/2Pker(D(α))1l/Γ322=p/2Pker(D(α))1l/Γ322=np/2Pker(D(α))enL2()2=np/2𝒞Pker(D(α)+k)1l/Γ3enL2()2.\begin{split}&\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha))}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}=\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha))}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}\\ &=\sum_{n\in{\mathbb{Z}}}\|\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha))}e_{n}\|^{2}_{L^{2}({\mathbb{C}})}=\sum_{n\in{\mathbb{Z}}}\|\langle\bullet\rangle^{p/2}\mathcal{C}P_{\ker(D(\alpha)+k)}\mathcal{B}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}e_{n}\|^{2}_{L^{2}({\mathbb{C}})}.\end{split}

Since by assumption Pker(D(α)+k)=φ(,k)φ(,k)P_{\ker(D(\alpha)+k)}=\varphi(\bullet,k)\otimes\varphi(\bullet,k) is a rank 11 projection, we have

p/2𝒞Pker(D(α)+k)1l/Γ3enL2()2=p/2𝒞φL2()2|φ,(1l/Γ3en)L2(/Γ3×/Γ3)|2=p/2𝒞φL2()2|𝒞φ,1l/Γ3enL2()|2.\begin{split}\|\langle\bullet\rangle^{p/2}\mathcal{C}P_{\ker(D(\alpha)+k)}\mathcal{B}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}e_{n}\|^{2}_{L^{2}({\mathbb{C}})}&=\|\langle\bullet\rangle^{p/2}\mathcal{C}\varphi\|^{2}_{L^{2}({\mathbb{C}})}|\langle\varphi,\mathcal{B}(\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}e_{n})\rangle_{L^{2}({\mathbb{C}}/\Gamma_{3}\times{\mathbb{C}}/\Gamma_{3}^{*})}|^{2}\\ &=\|\langle\bullet\rangle^{p/2}\mathcal{C}\varphi\|^{2}_{L^{2}({\mathbb{C}})}|\langle\mathcal{C}\varphi,\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}e_{n}\rangle_{L^{2}({\mathbb{C}})}|^{2}.\end{split}

This implies that

p/2Pker(D(α))1l/Γ322=p/2𝒞φL2()2n|𝒞φ,enL2(/Γ3)|2=p/2𝒞φL2()2𝒞φL2(/Γ3).\begin{split}\left\lVert\langle\bullet\rangle^{p/2}P_{\ker(D(\alpha))}\operatorname{1\hskip-2.75ptl}_{{\mathbb{C}}/\Gamma_{3}}\right\rVert_{2}^{2}&=\|\langle\bullet\rangle^{p/2}\mathcal{C}\varphi\|^{2}_{L^{2}({\mathbb{C}})}\sum_{n\in\mathbb{Z}}|\langle\mathcal{C}\varphi,e_{n}\rangle_{L^{2}({\mathbb{C}}/\Gamma_{3})}|^{2}\\ &=\|\langle\bullet\rangle^{p/2}\mathcal{C}\varphi\|^{2}_{L^{2}({\mathbb{C}})}\|\mathcal{C}\varphi\|_{L^{2}({\mathbb{C}}/\Gamma_{3})}.\end{split}

However, a Wannier basis is obtained from 𝒞φ\mathcal{C}\varphi by defining wγ:=γ𝒞φw_{\gamma}:=\mathscr{L}_{\gamma}\mathcal{C}\varphi. Indeed, using (5.3) functions wγw_{\gamma} are an orthonormal basis of ker(D(α))\ker(D(\alpha)) as

wγ,wγL2()=γ𝒞φ,γ𝒞φL2()=δγ,γ\langle w_{\gamma},w_{\gamma^{\prime}}\rangle_{L^{2}({\mathbb{C}})}=\langle\mathscr{L}_{\gamma}\mathcal{C}\varphi,\mathscr{L}_{\gamma^{\prime}}\mathcal{C}\varphi\rangle_{L^{2}({\mathbb{C}})}=\delta_{\gamma,\gamma^{\prime}}

and span ker(D(α))\ker(D(\alpha)) due to (5.1) and (5.2). Thus, we obtain since γ\mathscr{L}_{\gamma} is an isometry that

p/2w0L2()2=γp/2w0L2()2=+γp/2wγL2().\|\langle\bullet\rangle^{p/2}w_{0}\|^{2}_{L^{2}({\mathbb{C}})}=\|\mathscr{L}_{\gamma}\langle\bullet\rangle^{p/2}w_{0}\|^{2}_{L^{2}({\mathbb{C}})}=\|\langle\bullet+\gamma\rangle^{p/2}w_{\gamma}\|_{L^{2}({\mathbb{C}})}.

From the non-existence of a 11-localized Wannier basis and the existence of a (1δ)(1-\delta) Wannier basis, for any δ>0\delta>0, see for instance [MPPT18], we find that p/2w0L2()2=\|\langle\bullet\rangle^{p/2}w_{0}\|^{2}_{L^{2}({\mathbb{C}})}=\infty for p1p\geq 1 and is finite for p<1.p<1.

Appendix A Essential self-adjointness

In this appendix, we recall the essential self-adjointness of our Hamiltonian with even possibly unbounded disorder on Cc()C_{c}^{\infty}({\mathbb{C}}).

Theorem 8.

The Hamiltonian Hλ(α)H_{\lambda}(\alpha) (1.11) is, under the more general assumptions, with L()L^{\infty}({\mathbb{R}})-bounded density gg for random variables (ωγ)(\omega_{\gamma}) and arbitrary density hh is almost surely essentially self-adjoint on Cc().C_{c}^{\infty}({\mathbb{C}}).

Proof.

To see that Hλ(α)H_{\lambda}(\alpha) is essentially self-adjoint, we first observe that it is symmetric on Cc().C_{c}^{\infty}({\mathbb{C}}). It thus suffices to show that for any L2L^{2}-normalized ψ\psi

(Hλ(α)±i)ψ=0 implies ψ0,(H_{\lambda}(\alpha)\pm i)\psi=0\text{ implies }\psi\equiv 0,

i.e. the deficiency indices are zero. Elliptic regularity and the assumption that uLu\in L^{\infty} implies that ψC()\psi\in C^{\infty}({\mathbb{C}}). We then pick a cut-off function ηn(z):=η(z/n)\eta_{n}(z):=\eta(z/n) with ηCc()\eta\in C_{c}^{\infty}({\mathbb{C}}) and η|B1(0)1\eta|_{B_{1}(0)}\equiv 1 and find

(Hλ(α)±i)ηnψ=(02Dzηnid22Dz¯ηnid20)ψ.(H_{\lambda}(\alpha)\pm i)\eta_{n}\psi=\begin{pmatrix}0&2D_{z}\eta_{n}\cdot\operatorname{id}_{\mathbb{C}^{2}}\\ 2D_{\bar{z}}\eta_{n}\cdot\operatorname{id}_{\mathbb{C}^{2}}&0\end{pmatrix}\psi.

We conclude that

ηnψ22+Hλ(α)ηnψ22=(Hλ(α)±i)ηnψ22ηn2=𝒪(1/n2)n0.\|\eta_{n}\psi\|_{2}^{2}+\|H_{\lambda}(\alpha)\eta_{n}\psi\|_{2}^{2}=\|(H_{\lambda}(\alpha)\pm i)\eta_{n}\psi\|_{2}^{2}\lesssim\|\nabla\eta_{n}\|_{\infty}^{2}=\mathcal{O}(1/n^{2})\xrightarrow[n\to\infty]{}0.

Since ηnψψ\eta_{n}\psi\to\psi by dominated convergence, we conclude that ψ0.\psi\equiv 0.

Appendix B Partial Chern numbers

Let PP be an orthogonal projection on L2(;2n)L^{2}({\mathbb{C}};{\mathbb{C}}^{2n}) such that for some ξ(0,1)\xi\in(0,1), κ>0\kappa>0, and KP<K_{P}<\infty we have

χz0Pχz12KPz0κz1κe|z0z1|ξ for all z0,z1Γ.\|\chi_{z_{0}}P\chi_{z_{1}}\|_{2}\leq K_{P}\langle z_{0}\rangle^{\kappa}\langle z_{1}\rangle^{\kappa}e^{-|z_{0}-z_{1}|^{\xi}}\text{ for all }z_{0},z_{1}\in\Gamma. (B.1)

This condition is satisfied for spectral projections of Hamiltonians under the assumption of (SUDEC), cf. Def. 4.1. Let π1:=diag(idn,0)\pi_{1}:=\operatorname{diag}(\operatorname{id}_{{\mathbb{C}}^{n}},0) and π2:=diag(0,idn)\pi_{2}:=\operatorname{diag}(0,\operatorname{id}_{{\mathbb{C}}^{n}}). By the definition of the Hilbert-Schmidt norm one finds for all i,ji,j

χz0πiPπjχz12χz0Pχz12KPz0κz1κe|z0z1|ξ for all z0,z1Γ.\|\chi_{z_{0}}\pi_{i}P\pi_{j}\chi_{z_{1}}\|_{2}\leq\|\chi_{z_{0}}P\chi_{z_{1}}\|_{2}\leq K_{P}\langle z_{0}\rangle^{\kappa}\langle z_{1}\rangle^{\kappa}e^{-|z_{0}-z_{1}|^{\xi}}\text{ for all }z_{0},z_{1}\in\Gamma. (B.2)

We define the new Θ^j,i:=πiΘj=Θjπi\hat{\Theta}_{j,i}:=\pi_{i}\Theta_{j}=\Theta_{j}\pi_{i} and replace (4.4) by

Ωi(P):=tr(P[[P,Θ^1(i)],[P,Θ^2(i)]])\Omega_{i}(P):=\operatorname{tr}(P[[P,\hat{\Theta}_{1}(i)],[P,\hat{\Theta}_{2}(i)]]) (B.3)

which is well-defined for

|Ωi(P)|:=P[[P,Θ^1(i)],[P,Θ^2(i)]]1<.|\Omega_{i}(P)|:=\|P[[P,\hat{\Theta}_{1}(i)],[P,\hat{\Theta}_{2}(i)]]\|_{1}<\infty. (B.4)
Remark 4.

It is convenient to modify Θ^i\hat{\Theta}_{i} rather than PP in the definition of Ω\Omega, since πiPπj\pi_{i}P\pi_{j} is in general no longer a projection, even for i=j.i=j.

Since we still have that [Θ^i,Θ^j]=0[\hat{\Theta}_{i},\hat{\Theta}_{j}]=0 we find the equivalent formulation of (B.3)

Ωi(P)=tr[PΘ^1(i)P,PΘ^2(i)P].\Omega_{i}(P)=\operatorname{tr}[P\hat{\Theta}_{1}(i)P,P\hat{\Theta}_{2}(i)P]. (B.5)

In particular, if PP is a finite-rank projection then, we always find Ωi(P)=0\Omega_{i}(P)=0, as (B.5) is a commutator of trace-class operators.

To provide further motivation for the above definition (B.3), we shall consider the unperturbed Hamiltonian H0=(mDDm)H_{0}=\begin{pmatrix}m&D^{*}\\ D&-m\end{pmatrix} then H02=diag(DD+m2,DD+m2)H_{0}^{2}=\operatorname{diag}(D^{*}D+m^{2},DD^{*}+m^{2}) and consequently any spectral projection of H02H_{0}^{2} is also diagonal and thus of the form P0=diag(P0(1),P0(2)).P_{0}=\operatorname{diag}(P_{0}(1),P_{0}(2)). Thus, we have

Ωi(P0)=tr([P0Θ^1(i)P0,P0Θ^2(i)P0])=Ω(P0(i)),\Omega_{i}(P_{0})=\operatorname{tr}([P_{0}\hat{\Theta}_{1}(i)P_{0},P_{0}\hat{\Theta}_{2}(i)P_{0}])=\Omega(P_{0}(i)),

where we recall from (1.10) that for a generic magic angle and P0=1l[0,μ](H02)P_{0}=\operatorname{1\hskip-2.75ptl}_{[0,\mu]}(H_{0}^{2}) with μ(0,Egap2)\mu\in(0,E_{\text{gap}}^{2})

Ω1(P0)=i2π and Ω2(P0)=i2π.\Omega_{1}(P_{0})=\frac{i}{2\pi}\text{ and }\Omega_{2}(P_{0})=-\frac{i}{2\pi}. (B.6)

Thus, while Ω(P0)=0\Omega(P_{0})=0 for m=0m=0, we have Ω1(P0),Ω2(P0)0.\Omega_{1}(P_{0}),\Omega_{2}(P_{0})\neq 0. The definition of Ωi\Omega_{i} captures the non-trivial sublattice Chern numbers of twisted bilayer graphene while the total Chern number vanishes.

One also readily verifies the usual properties of Chern characters, see for instance [GKS07, Lemma 3.1], [BES94]:

Proposition B.1.

Let PP be an orthogonal projection satisfying (B.1), then

  1. (1)

    |Ωi(P)|κ,ξKP2.|\Omega_{i}(P)|\lesssim_{\kappa,\xi}K_{P}^{2}.

  2. (2)

    Let ss\in\mathbb{R} and define Θ^j,i(s)(t):=πiΘj(ts)\hat{\Theta}^{(s)}_{j,i}(t):=\pi_{i}\Theta_{j}(t-s), then

    Ωir,s(P):=tr(P[[P,Θ^1,i(s)],[P,Θ^2,i(r)]]) for r,s.\Omega_{i}^{r,s}(P):=\operatorname{tr}(P[[P,\hat{\Theta}_{1,i}^{(s)}],[P,\hat{\Theta}_{2,i}^{(r)}]])\text{ for }r,s\in{\mathbb{R}}.

    In particular,

    Ωir,s=Ωi.\Omega_{i}^{r,s}=\Omega_{i}. (B.7)
  3. (3)

    Let P,QP,Q be two orthogonal projections, each satisfying (B.1), such that PQ=QP=0PQ=QP=0, then

    Ωi(P+Q)=Ωi(P)+Ωi(Q).\Omega_{i}(P+Q)=\Omega_{i}(P)+\Omega_{i}(Q).
Proof.

The first property follows readily from combining (B.2) with the proof [GKS07, Lemma 3.1 (i)].

The second property follows from a direct computation, see [GKS07, Lemma 3.1 (ii)].

The last property follows from P[Q,Θ^i]=PΘ^iQP[Q,\hat{\Theta}_{i}]=-P\hat{\Theta}_{i}Q and evaluating (B.3) since one finds for the cross-terms

tr(PΘ^1QΘ^2+PΘ^1QΘ^2PQΘ^1PΘ^2+QΘ^1PΘ^2Q)=0.\begin{split}\operatorname{tr}\Big{(}-P\hat{\Theta}_{1}Q\hat{\Theta}_{2}+P\hat{\Theta}_{1}Q\hat{\Theta}_{2}P-Q\hat{\Theta}_{1}P\hat{\Theta}_{2}+Q\hat{\Theta}_{1}P\hat{\Theta}_{2}Q\Big{)}=0.\end{split}

We also want to mention reference [ASS94, Sec.6] showing full details on how to obtain the second point and [BES94, Lemma 88] for the third point.

The independence of switch functions Θ^j,i(s)\hat{\Theta}^{(s)}_{j,i} in Prop. B.1 implies that Ωi\Omega_{i} is an almost surely constant quantity

Ωi(P)=𝐄Ωi(P) for 𝐏a.s.\Omega_{i}(P)=\mathbf{E}\Omega_{i}(P)\text{ for }\mathbf{P}-a.s.

The purpose of the first and last point in Prop. B.1 is to conclude that in regions of SUDEC\operatorname{SUDEC}, cf. Definition 4.1, all Ωi\Omega_{i} vanish.

Proposition B.2.

Let HλH_{\lambda} exhibit SUDEC\operatorname{SUDEC} in an interval JJ, then for all closed IJI\subset J we have

Ωi(1lI(Hλ))=0 for 𝐏a.s.\Omega_{i}(\operatorname{1\hskip-2.75ptl}_{I}(H_{\lambda}))=0\text{ for }\mathbf{P}-a.s.
Proof.

Let MM\subset\mathbb{N} be a (finite or infinite) enumeration (counting multiplicities) of all point spectrum of HλH_{\lambda}. We can then write the spectral projection as

1lI(Hλ)=mMPm\operatorname{1\hskip-2.75ptl}_{I}(H_{\lambda})=\sum_{m\in M}P_{m}

where PmP_{m} are rank one projections. In addition, we have KP:=mMαmK_{P}:=\sum_{m\in M}\alpha_{m} where αm\alpha_{m} are defined in (4.1). Using the third item in Prop. B.1 we then have for any {1,,N}M\{1,...,N\}\subset M

Ωi(1lI(Hλ))=m=1NΩi(Pm)=0+Ωi(mM{1,..,N}Pm)=Ωi(mM{1,..,N}Pm).\Omega_{i}(\operatorname{1\hskip-2.75ptl}_{I}(H_{\lambda}))=\sum_{m=1}^{N}\underbrace{\Omega_{i}(P_{m})}_{=0}+\Omega_{i}\Bigg{(}\sum_{m\in M\setminus\{1,..,N\}}P_{m}\Bigg{)}=\Omega_{i}\Bigg{(}\sum_{m\in M\setminus\{1,..,N\}}P_{m}\Bigg{)}.

Invoking then the first item in Prop. B.1, we find that as we make NN arbitrarily large or equal to |M||M|, if MM is finite, we obtain Ωi(mM{1,..,N}Pm)0.\Omega_{i}(\sum_{m\in M\setminus\{1,..,N\}}P_{m})\to 0.

Acknowledgements:. We thank Jie Wang for suggesting the relevance of different disorder types for TBG and Maciej Zworski for initial discussions on the project. M. Vogel was partially funded by the Agence Nationale de la Recherche, through the project ADYCT (ANR-20-CE40-0017). I. Oltman was jointly funded by the National Science Foundation Graduate Research Fellowship under grant DGE-1650114 and by grant DMS-1901462.

References

  • [AS19] N. Anantharaman, M. Sabri, Quantum ergodicity on graphs: From spectral to spatial delocalization, Ann. of Math. (2) 189(3): 753-835 (2019).
  • [AW13] M. Aizenman, S. Warzel, Resonant delocalization for random Schrödinger operators on tree graphs, J. Eur. Math. Soc 15 (4), 1167-1222, (2013).
  • [ASS94] J.  Avron, R.  Seiler, and B. Simon, Charge Deficiency, Charge Transport and Comparison of Dimensions, Commun. Math. Phys. 159, 399-422, (1994).
  • [An58] P.  Anderson, Absence of diffusion in certain random lattices. Phys. Rev. 109, 1492- 1505, (1958).
  • [BCZ19] J. -M. Barbaroux, H. Cornean, and S. Zalczer, Documenta mathematica, 24:65-93, (2019).
  • [BES94] J. Bellissard, A. van Elst, and H. Schulz‐-Baldes, The noncommutative geometry of the quantum Hall effect, J. Math. Phys. 35, 5373 (1994).
  • [Be*21] S. Becker, M. Embree, J. Wittsten and M. Zworski, Spectral characterization of magic angles in twisted bilayer graphene, Phys. Rev. B 103, 165113, (2021).
  • [Be*22] S. Becker, M. Embree, J. Wittsten and M. Zworski, Mathematics of magic angles in a model of twisted bilayer graphene, Probab. Math. Phys. 3,69–103, (2022)
  • [BH22] S. Becker, R. Han, Density of States and Delocalization for Discrete Magnetic Random Schrödinger Operators, International Mathematics Research Notices 2022 (17), 13447-13504.
  • [BHZ22] S. Becker, T. Humbert and M. Zworski, Fine structure of flat bands in a chiral model of magic angles, arXiv:2208.01628, (2022).
  • [BHZ22b] S. Becker, T. Humbert and M. Zworski, Integrability in the chiral model of magic angles, arXiv:2208.01620, (2022).
  • [BHZ23] S. Becker, T. Humbert and M. Zworski, Degenerate flat bands in twisted bilayer graphene, arXiv:2306.02909, (2023).
  • [BM87] M.V. Berry and R.J. Mondragon, Neutrino billiards:time-reversal symmetry-breaking without magnetic fields.Proc. Roy. Soc. London Ser.A,412(1842):53–74, (1987).
  • [BiMa11] R. Bistritzer and A. MacDonald, Moiré bands in twisted double-layer graphene. PNAS, 108, 12233–12237, (2011).
  • [CH94] J. M. Combes and P. D. Hislop, Localization for some continuous, random Hamiltonians in d dimensions, J. Funct. Anal. 12, 149-180, (1994).
  • [CHK07] J.-M. Combes, P. D. Hislop, and F. Klopp, An optimal Wegner estimate and its application to the global continuity of the integrated density of states for random Schrödinger operators, Duke Math. J. 140(3), 469–498, (2007).
  • [CHN01] J.-M. Combes, P. D. Hislop, and S. Nakamura, The LpL^{p}-theory of the spectral shift function, the Wegner estimate, and the integrated density of states for some random operators, Comm. Math. Phys. 218, no. 1, 113–130,(2001).
  • [CHK03] J.-M. Combes, P. D. Hislop, and F. Klopp, Hölder Continuity of the Integrated Density of States for Some Random Operators at All Energies, International Mathematics Research Notices, No. 4, (2003).
  • [CMM19] H.D.Cornean, D. Monaco,and M. Moscolari, Parseval Frames of Exponentially Localized Magnetic Wannier Functions. Commun. Math. Phys. 371, 1179-1230, (2019).
  • [CRQ23] V. Crépel, N. Regnault, R. Queiroz, The chiral limits of moiré semiconductors: origin of flat bands and topology in twisted transition metal dichalcogenides homobilayers, arXiv:2305.10477, (2023).
  • [CFKS87] H.L Cycon, R.G. Froese, W. Kirsch, W., B. Simon, Schrödinger Operators. Lecture Notes in Physics, Berlin-Heidelberg-New York: Springer Verlag, (1987).
  • [ET05] M. Embree, L. Trefethen, Spectra and Pseudospectra: The Behavior of Nonnormal Matrices and Operators, Princeton University Press, (2005).
  • [GK01] F. Germinet, A. Klein, Bootstrap Multiscale Analysis and Localization in Random Media. Commun. Math. Phys. 222, 415–448, (2001).
  • [GK03] F. Germinet, A. Klein, Explicit finite volume criteria for localization in continuous random media and applications. Geom. Funct. Anal. 13 1201-1238, (2003).
  • [GK04] F. Germinet, A. Klein, A characterization of the Anderson metal-insulator transport transition. Duke Math. J. 124, 309-351, (2004).
  • [GK06] F. Germinet, A. Klein, New characterizations of the region of complete localization for random Schrödinger operators, J. Statist. Phys. 122, 73–94, (2006).
  • [GKS07] F. Germinet, A. Klein, and J. Schenker, Delocalization in random Landau Hamiltonians, Annals of Mathematics, 166, 215-244, (2007).
  • [GKS09] F. Germinet, A. Klein, and J. Schenker, Quantization of the Hall conductance and delocalization in ergodic Landau Hamiltonians, Reviews in Mathematical Physics, Vol. 21, No. 08, pp. 1045-1080, (2009).
  • [GKM09] F. Germinet, A. Klein, and B. Mandy, Delocalization for random Landau Hamiltonians with unbounded random variables, Contemporary Mathematics, Volume 500, (2009).
  • [JSS03] S. Jitomirskaya, H. Schulz-Baldes, G. Stolz, Delocalization in random polymer models, Commun. Math. Phys. 233, 27-48, (2003).
  • [K89] W. Kirsch, Random Schrödinger operators, in ”Schrödinger Operators” (H. Holdenand A. Jensen, Eds.),Lecture Notes in Physics,Vol. 345, Springer-Verlag, Berlin New York, (1989).
  • [Li21] T. Li, S. Jiang, B. Shen. et al. Quantum anomalous Hall effect from intertwined moiré bands. Nature 600, 641–646, (2021).
  • [LS21] J. Lu and K. D Stubbs, Algebraic localization of Wannier functions implies Chern triviality in non-periodic insulators, arXiv:2107.10699, (2021).
  • [MMP] G. Marcelli, M. Moscolari, and G. Panati, Localization implies Chern triviality in non periodic insulators, arXiv:2012.14407, (2020).
  • [MPPT18] D. Monaco, G. Panati, A. Pisante, S. Teufel, Optimal Decay of Wannier functions in Chern and Quantum Hall Insulators. Commun. Math. Phys. 359, 61-100, (2018).
  • [N01] S.  Nakamura, A Remark on the Dirichlet-Neumann Decoupling and the Integrated Density of States, Journal of Functional Analysis 179,136152, (2001).
  • [Pa80] L. Pastur, Spectral properties of disordered systems in one-body approximation. Commun. Math. Phys. 75, (1980).
  • [Si77] B. Simon, Notes on infinite determinants of Hilbert space operators, Adv. in Math. 24, 244-273, (1977).
  • [Si00] B. Simon, Schrödinger Operators in the Twenty-First Century. Mathematical Physics. Imperial College London. pp. 283–288, (2000).
  • [Sj89] J. Sjöstrand, Microlocal analysis for periodic magnetic Schrödinger equation and related questions, in Microlocal Analysis and Applications, J.-M. Bony, G. Grubb, L. Hörmander, H. Komatsu and J. Sjöstrand eds. Lecture Notes in Mathematics 1495, Springer, (1989).
  • [SV19] E. Stockmeyer, S. Vugalter, Infinite mass boundary conditions for Dirac operators. J. Spectr. Theory 9, no. 2, pp. 569–600, (2019).
  • [T92] B. Thaller, The Dirac Equation, Springer-Verlag, (1992).
  • [TE05] L. Trefethen and M. Embree, Spectra and Pseudospectra: The Behavior of Nonnormal Matrices and Operators, Princeton: Princeton University Press, (2005).
  • [TKV19] G. Tarnopolsky, A.J. Kruchkov and A. Vishwanath, Origin of magic angles in twisted bilayer graphene, Phys. Rev. Lett. 122, 106405, (2019).
  • [W95] W.-M. Wang, Asymptotic expansion for the density of states of the magnetic Schrödinger operator with a random potential. Commun. Math. Phys. 172, 401–425, (1995).
  • [Y92] D. R. Yafaev, Mathematical Scattering Theory. General Theory, Translations of Mathematical Monographs, vol. 105, American Mathematical Society, Rhode Island, (1992).