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

Eigenvalue Asymptotics near a flat band in presence of a slowly decaying potential

Pablo Miranda Departamento de Matemática y Ciencia de la Computación, Universidad de Santiago de Chile, Las Sophoras 173, Santiago, Chile. pablo.miranda.r@usach.cl  and  Daniel Parra Departamento de Matemática y Estadística, Universidad de La Frontera, Avenida Francisco Salazar 01145, Temuco, Chile daniel.parra@ufrontera.cl
Abstract.

We provide eigenvalue asymptotics for a Dirac–type operator on n\mathbb{Z}^{n}, n2n\geq 2, perturbed by multiplication operators that decay as |μ|γ|\mu|^{-\gamma} with γ<n\gamma<n. We show that the eigenvalues accumulate near the value of the flat band at a “semiclassical” rate with a constant that encodes the structure of the flat band. Similarly, we show that this behaviour can be obtained also for a Laplace operator on a periodic graph.

Keywords: Discrete Dirac operator, Eigenvalue Asymptotics, Flat Band.


Mathematics Subject Classification: 35P20, 47A10, 81Q10, 47A55.

1. Introduction

In this article, we consider an operator whose decomposition into a direct integral presents a flat band. We are interested in the accumulation of eigenvalues near the value of the flat band when a perturbation is added. We start by briefly explaining the setting and then discuss our motivation for entertaining such an analysis.

Let us denote by 𝒳\mathcal{X} the standard graph structure in n\mathbb{Z}^{n} and consider the Dirac type operator on 2(𝒳)\ell^{2}(\mathcal{X}) defined by

H0=(mddm),H_{0}=\begin{pmatrix}m&d^{*}\\ d&-m\end{pmatrix},

where dd is the discrete version of the exterior derivative and mm a positive constant. We refer to Section˜2.1 for the precise definition but one can readily notice that by construction H0H_{0} satisfies the supersymmetry condition making it into an abstract Dirac operator as in [Tha92]. Moreover, from the analysis of its band functions, see (8) below, we obtain that the spectrum of H0H_{0} is

(1) σ(H0)=σess(H0)=σac(H0)=[m2+4n,m][m,m2+4n].\sigma(H_{0})=\sigma_{ess}(H_{0})=\sigma_{ac}(H_{0})=[-\sqrt{m^{2}+4n},-m]\bigcup[m,\sqrt{m^{2}+4n}]\ .

An essential observation pertinent to this study is that if n2n\geq 2, m-m is an embedded infinite dimensional eigenvalue of H0H_{0}. Throughout this article, we will assume that m>0m>0 and n2n\geq 2.

Let us now consider a perturbation by a multiplication operator V:𝒳V:\mathcal{X}\to\mathbb{R} decaying at infinity. Hence we define

(2) H:=H0+V.H:=H_{0}+V\ .

Since VV is a compact operator, σess(H)=σess(H0)\sigma_{\rm ess}(H)=\sigma_{\rm ess}(H_{0}). Moreover, since m>0m>0, equality Eq.˜1 tell us that (m,m)(-m,m) is a gap in the essential spectrum of HH. Then, for λ(0,m)\lambda\in(0,m) we consider the function

𝒩(λ)=Rank𝟙(m+λ,0)(H),\mathcal{N}(\lambda)={\rm Rank}\mathds{1}_{(-m+\lambda,0)}(H)\ ,

with 𝟙Ω\mathds{1}_{\Omega} being the characteristic function over the Borel set Ω\Omega. Clearly, this function count the number of eigenvalues of HH (with multiplicity) on the interval (m+λ,0)(-m+\lambda,0).

Our primary objective is to analyze the asymptotic behaviour of 𝒩\mathcal{N} as λ0\lambda\downarrow 0 for a specific class of perturbations that decay slowly at infinity. Further details can be found in ˜3.1, while our main result is presented in ˜3.2.

One motivation for studying 𝒩\mathcal{N} stems from our previous work on the distribution of eigenvalues as presented in [MPR23]. This article extends our prior research in three significant ways: it encompasses the general nn-dimensional scenario, incorporates the potential for non-definite perturbations, and addresses potentials with slower rates of decay at infinity. Moreover, we employ a distinct method to derive the effective Hamiltonian, drawing inspiration from the analysis of eigenvalue distributions for magnetic Schrödinger operators, see [Rai90, IT98, PR11]. This approach yields an effective Hamiltonian with a “typical” structure denoted as PVPPVP, where PP is a projection.

Another motivation arises from the recent surge in interest surrounding the study of flat bands in the discrete setting. Unlike the common assumption in the continuous case, periodic Schrödinger operators in periodic graphs often exhibit flat bands, as discussed in [SY23]. While these configurations have long been studied by the physics community, see [BL13, Kol+20] and references therein, recent attention from the spectral theory community has also emerged, see for instance [KTW23, PS23, Zwo24, GZ23]. Remarkably, we demonstrate a striking similarity between the results obtained for our Dirac operator and those for the Laplacian on a specific periodic graph showcasing such a flat band, see ˜5.3.

We finish this introduction by briefly describing the structure of the article. In Section˜2 we give the precise definition of H0H_{0} and study its main spectral characteristics. In Section˜3 we introduce the class of admissible perturbations and state our main result, which we prove in Section˜4. Finally, in Section˜5, we show a similar result for the standard graph–Laplacian in a particular 2\mathbb{Z}^{2}–periodic graph.

2. Spectral Theory for a Dirac operator on n\mathbb{Z}^{n}

In this section we provide the definition of H0H_{0} taking most notations from [Par17], see also [AT15], recall its integral decomposition, and show the explicit expression of its resolvent as a fibered operator that will be central to our investigations.

2.1. Discrete Dirac operator

We denote by 𝒳=(𝒱,𝒜)\mathcal{X}=(\mathcal{V},\mathcal{A}) the standard graph structure in n\mathbb{Z}^{n}. That is, the set of vertices 𝒱\mathcal{V} consists of points μn\mu\in\mathbb{Z}^{n} and the set of oriented edges 𝒜\mathcal{A} is composed of pairs (μ,ν)(\mu,\nu) such that ν=μ±δj\nu=\mu\pm\delta_{j}, where {δj}j=1n\{\delta_{j}\}_{j=1}^{n} denotes the canonical basis of n\mathbb{Z}^{n}. An edge in 𝒜\mathcal{A} is written e=(μ,ν)\mathrm{e}=(\mu,\nu) and its transpose e¯:=(ν,μ)\overline{\mathrm{e}}:=(\nu,\mu). Let us consider the vector spaces of 00-cochains C0(𝒳)C^{0}(\mathcal{X}) and 11-cochains C1(𝒳)C^{1}(\mathcal{X}) given by

C0(𝒳):={f:𝒱} ; C1(𝒳):={g:𝒜g(e)=g(e¯)}.C^{0}(\mathcal{X}):=\{f:\mathcal{V}\to\mathbb{C}\}\text{ ; }\qquad C^{1}(\mathcal{X}):=\{g:\mathcal{A}\to\mathbb{C}\mid g(\mathrm{e})=-g(\overline{\mathrm{e}})\}.

The Hilbert spaces 2(𝒱)\ell^{2}(\mathcal{V}) and 2(𝒜)\ell^{2}(\mathcal{A}) are naturally defined by the inner products of cochains: f1,f20=μ𝒱f1(μ)f2(μ)¯\langle f_{1},f_{2}\rangle_{0}=\sum_{\mu\in\mathcal{V}}f_{1}(\mu)\overline{f_{2}(\mu)} and g1,g21=12e𝒜g1(e)g2(e)¯\langle g_{1},g_{2}\rangle_{1}=\frac{1}{2}\sum_{e\in\mathcal{A}}g_{1}(e)\overline{g_{2}(e)}, respectively.

The coboundary operator d:2(𝒱)2(𝒜)d:\ell^{2}(\mathcal{V})\to\ell^{2}(\mathcal{A}) is defined by

(3) df(e):=f(ν)f(μ), for e=(μ,ν)𝒜.df(\mathrm{e}):=f(\nu)-f(\mu),\quad\text{ for }\mathrm{e}=(\mu,\nu)\in\mathcal{A}\ .

This is the discrete version of the exterior derivative and its adjoint d:2(𝒜)2(𝒱)d^{*}:\ell^{2}(\mathcal{A})\to\ell^{2}(\mathcal{V}) is given at each edge by the finite sum

(4) dg(μ)=j=1ng(μ,μ±δj), for μ𝒱.d^{*}g(\mu)=\sum_{j=1}^{n}g(\mu,\mu\pm\delta_{j}),\quad\text{ for }\mu\in\mathcal{V}.

Let us define the Hilbert space 2(𝒳)=2(𝒱)2(𝒜)\ell^{2}(\mathcal{X})=\ell^{2}(\mathcal{V})\oplus\ell^{2}(\mathcal{A}) and denote by P𝒱P_{\mathcal{V}} and P𝒜P_{\mathcal{A}} the corresponding projections. Further, we introduce the involution τ\tau on 2(𝒳)\ell^{2}(\mathcal{X}) by

τ(f,g)=(f,g).\tau(f,g)=(f,-g)\ .

Then, for a strictly positive constant mm let us consider the free Dirac operator

H0=d+d+mτ=(0dd0)+m(1001)=(mddm)H_{0}=d+d^{*}+m\tau=\begin{pmatrix}0&d^{*}\\ d&0\end{pmatrix}+m\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}=\begin{pmatrix}m&d^{*}\\ d&-m\end{pmatrix}

where we have slightly abused notation by considering dd and dd^{*} acting on 2(𝒳)\ell^{2}(\mathcal{X}). Note that H0H_{0} is a Dirac-type operator in the sense that

H02=(Δ0+m200Δ1+m2)H_{0}^{2}=\begin{pmatrix}\Delta_{0}+m^{2}&0\\ 0&\Delta_{1}+m^{2}\end{pmatrix}

where Δ0\Delta_{0} is the Laplacian on vertices and Δ1\Delta_{1} is the (1-down) Laplacian on edges.

2.2. Integral decomposition

Let us denote by :=L2(𝕋n,n+1)\mathcal{H}:=L^{2}(\mathbb{T}^{n},\mathbb{C}^{n+1}). In this section we construct a unitary operator 𝒰:l2(𝒳)\mathscr{U}:l^{2}(\mathcal{X})\to\mathcal{H}. Consider the action of n\mathbb{Z}^{n} on 𝒳\mathcal{X} given for μn\mu\in\mathbb{Z}^{n}, x𝒱x\in\mathcal{V}, and e=(x,y)A\mathrm{e}=(x,y)\in A by

μx=μ+x and μe:=(μ+x,μ+y).\mu x=\mu+x\text{ and }\mu\mathrm{e}:=(\mu+x,\mu+y)\ .

Then, a natural class of representatives of the orbits of such action is given by 𝟎𝒱\mathbf{0}\in\mathcal{V} together with the edges ej=(𝟎,δj)\mathrm{e}_{j}=(\mathbf{0},\delta_{j}) and ej=(𝟎,δj)\mathrm{e}_{j}^{-}=(\mathbf{0},-\delta_{j}).

Let us denote 𝕋n=n/[0,1]n\mathbb{T}^{n}=\mathbb{R}^{n}/[0,1]^{n} and set Cc(𝒳)C_{c}(\mathcal{X}) to be the set of cochains with compact support, i.e. , fCc(𝒳)f\in C_{c}(\mathcal{X}) if and only if it vanishes except for a finite number of vertices and edges. We define 𝒰:Cc(𝒳)L2(𝕋n,n+1)\mathscr{U}:C_{c}(\mathcal{X})\to L^{2}(\mathbb{T}^{n},\mathbb{C}^{n+1}) by setting, for fCc(𝒳)f\in C_{c}(\mathcal{X}) and ξ𝕋n\xi\in\mathbb{T}^{n},

(𝒰f)(ξ)=(μne2πiξμf(μ),μne2πiξμf(μe1),,μne2πiξμf(μen)).(\mathscr{U}f)(\xi)=\left(\sum_{\mu\in\mathbb{Z}^{n}}e^{-2\pi i\xi\cdot\mu}f(\mu),\sum_{\mu\in\mathbb{Z}^{n}}e^{-2\pi i\xi\cdot\mu}f(\mu\mathrm{e}_{1}),\ldots,\sum_{\mu\in\mathbb{Z}^{n}}e^{-2\pi i\xi\cdot\mu}f(\mu\mathrm{e}_{n})\right).

Then 𝒰\mathscr{U} extends to a unitary operator, still denoted by 𝒰\mathscr{U}, from 2(𝒳)\ell^{2}(\mathcal{X}) to \mathcal{H}. Further, we set 𝒱:=𝒰P𝒱(2(𝒳))L2(𝕋n)\mathcal{H}_{\mathcal{V}}:=\mathscr{U}P_{\mathcal{V}}(\ell^{2}(\mathcal{X}))\cong L^{2}(\mathbb{T}^{n}) and 𝒜:=𝒰P𝒜(2(𝒳))L2(𝕋n,n)\mathcal{H}_{\mathcal{A}}:=\mathscr{U}P_{\mathcal{A}}(\ell^{2}(\mathcal{X}))\cong L^{2}(\mathbb{T}^{n},\mathbb{C}^{n}).

We draw the reader’s attention to the fact that this definition of 𝒰\mathscr{U} correspond to the following choice of the Fourier transform in n\mathbb{Z}^{n}:

:l2(n)L2(𝕋n);(f)(ξ):=μne2πiξμf(μ).\mathcal{F}:l^{2}(\mathbb{Z}^{n})\to L^{2}(\mathbb{T}^{n})\quad;\quad(\mathcal{F}f)(\xi):=\sum_{\mu\in\mathbb{Z}^{n}}e^{-2\pi i\xi\cdot\mu}f(\mu)\ .

Finally, let us define the functions

aj(ξ):=1+e2πiξj.a_{j}(\xi):=-1+e^{-2\pi i\xi_{j}}\ .

The following Section shows that through conjugation by 𝒰\mathscr{U}, the operator H0H_{0} becomes a multiplication operator, enabling the study of its spectral properties through the examination of characteristics of its band functions.

Proposition 2.1 ([Par17, Prop. 3.5]).

The operator H0H_{0} satisfy that

𝒰H0𝒰=h0\mathscr{U}H_{0}\mathscr{U}^{*}=h_{0}

where h0h_{0} denotes the multiplication operator by the real analytic function

h0:𝕋nMn+1×n+1()h_{0}:\mathbb{T}^{n}\to M_{n+1\times n+1}(\mathbb{C})

on L2(𝕋n,n+1)L^{2}(\mathbb{T}^{n},\mathbb{C}^{n+1}) given by

(5) h0(ξ)=(ma1(ξ)an(ξ)a1(ξ)¯m0an(ξ)¯0m).h_{0}(\xi)=\begin{pmatrix}m&a_{1}(\xi)&\ldots&a_{n}(\xi)\\ \overline{a_{1}(\xi)}&-m&\ldots&0\\ \vdots&&\ddots&\vdots\\ \overline{a_{n}(\xi)}&0&\ldots&-m\end{pmatrix}\ .

2.3. Spectrum and resolvent of H0H_{0}

The band functions of h0h_{0} have an explicit expression so we are able to compute σ(H0)=jλj(𝕋n)\sigma(H_{0})=\bigcup_{j}\lambda_{j}(\mathbb{T}^{n}). Indeed, from Eq.˜5 one can see that for ξ𝕋n\xi\in\mathbb{T}^{n} the characteristic polynomial associated to h0(ξ)h_{0}(\xi) is given by

(6) pz(ξ)=(1)n(m+z)n1(m2z2+j=1n|ai(ξ)|2).p_{z}(\xi)=(-1)^{n}(m+z)^{n-1}\Big{(}m^{2}-z^{2}+\sum_{j=1}^{n}|a_{i}(\xi)|^{2}\Big{)}\ .

For convenience we define r:𝕋n+r:\mathbb{T}^{n}\to\mathbb{R}^{+} and ri:𝕋n+r_{i}:\mathbb{T}^{n}\to\mathbb{R}^{+} for i{1,n}i\in\{1,\dots n\} by

(7) r(ξ):=j=1n|aj(ξ)|2 and ri(ξ)=r(ξ)|ai(ξ)|2=ji|aj|2.r(\xi):=\sum_{j=1}^{n}|a_{j}(\xi)|^{2}\text{ and }r_{i}(\xi)=r(\xi)-|a_{i}(\xi)|^{2}=\sum_{j\neq i}|a_{j}|^{2}\ .

Thus, there are three band functions:

(8) z0(ξ)=m,z±(ξ)=±m2+r(ξ).z_{0}(\xi)=-m\ ,\quad z_{\pm}(\xi)=\pm\sqrt{m^{2}+r(\xi)}.

From the identities

(9) |aj(ξ)|2=2(1cos(2πξj))=4sin2(πξj)|a_{j}(\xi)|^{2}=2(1-\cos(2\pi\xi_{j}))=4\sin^{2}(\pi\xi_{j})

we easily see that the spectrum of H0H_{0} satisfies Eq.˜1. Moreover, as is shown in Figure˜1, we can observe that the threshold m-m correspond to both the maximum of zz_{-} and the constant value of the flat band. Note from Equations˜9 and 8 that zz_{-} attains its maximum only at 𝟎𝕋n\mathbf{0}\in\mathbb{T}^{n} for every nn and hence Figure˜1 is generic.

Refer to caption
Refer to caption
Figure 1. Two views of the three band functions for n=2n=2. The negative band and the flat band only touch at (0,0)(0,0).

From Equation˜5 and for zσ(H0)z\notin\sigma(H_{0}) on can check that

(h0z)1=(1)npz×((z+m)na1(z+m)n1an(z+m)n1a1¯(z+m)n1(z+m)n2(z2m2r1)ana1¯(z+m)n2an¯(z+m)n1a1an¯(z+m)n2(z+m)n2(z2m2rn))(h_{0}-z)^{-1}=\frac{(-1)^{n}}{p_{z}}\times\\ \begin{pmatrix}(z+m)^{n}&a_{1}(z+m)^{n-1}&\cdots&a_{n}(z+m)^{n-1}\\[5.0pt] \overline{a_{1}}(z+m)^{n-1}&(z+m)^{n-2}(z^{2}-m^{2}-r_{1})&\cdots&a_{n}\overline{a_{1}}(z+m)^{n-2}\\[5.0pt] \vdots&\vdots&\ddots&\vdots\\ \overline{a_{n}}(z+m)^{n-1}&a_{1}\overline{a_{n}}(z+m)^{n-2}&\cdots&(z+m)^{n-2}(z^{2}-m^{2}-r_{n})\end{pmatrix}

from where one can obtain

(10) (h0z)1=1(m+z)(m2z2+r)((z+m)2a1(z+m)an(z+m)a1¯(z+m)z2m2r1ana1¯an¯(z+m)a1an¯z2m2rn).(h_{0}-z)^{-1}\hskip-2.0pt=\frac{1}{(m+z)(m^{2}-z^{2}+r)}\begin{pmatrix}(z+m)^{2}&a_{1}(z+m)&\cdots&a_{n}(z+m)\\[5.0pt] \overline{a_{1}}(z+m)&z^{2}-m^{2}-r_{1}&\cdots&a_{n}\overline{a_{1}}\\[5.0pt] \vdots&\vdots&\ddots&\vdots\\ \overline{a_{n}}(z+m)&a_{1}\overline{a_{n}}&\cdots&z^{2}-m^{2}-r_{n}\end{pmatrix}\ .

Notice that from the particular form of h0(ξ)+mh_{0}(\xi)+m, that can be obtained directly from Equation˜5, we can check that Ker(h0+m)𝒜\operatorname{Ker}(h_{0}+m)\leq\mathcal{H}_{\mathcal{A}}. Indeed, one can prove directly that Ker(H0+m){𝟎}×2(A)\operatorname{Ker}(H_{0}+m)\leq\{\mathbf{0}\}\times\ell^{2}(A) by constructing for each μn\mu\in\mathbb{Z}^{n} a closed path over which we define a cochain alternating the values 11 and 1-1, see [MPR23, Sec. 2] for an explicit construction for the 2\mathbb{Z}^{2} case. We stress that the flat bands of discrete periodic graphs are known to generate finitely supported eigenfunctions [Kuc91].

3. Perturbed operator and Main Result

We turn now our attention to the concrete class of perturbations VV that we will treat in this article. A symmetric multiplication operator on 2(𝒳)\ell^{2}(\mathcal{X}) is defined by V:𝒳V:\mathcal{X}\to\mathbb{R} such that V(e)=V(e¯)V(\mathrm{e})=V(\overline{\mathrm{e}}) for every eA\mathrm{e}\in A. Given such a VV, our full hamiltonian is defined by Equation˜2.

Further, we define the following real-valued functions on n\mathbb{Z}^{n}

(11) v0(μ):=V(μ);vj(μ):=V(μej),1jn.v_{0}(\mu):=V(\mu);\quad v_{j}(\mu):=V(\mu\mathrm{e}_{j}),\quad 1\leq j\leq n\ .

This choice allows us to further specify the decay of VV at infinity, but other choices of representatives would give the same type of decay.

Let us consider the class of Symbols Sγ(n)S^{\gamma}(\mathbb{Z}^{n}) given by the functions v:nv:\mathbb{Z}^{n}\to\mathbb{C} that satisfies for any multi-index α=(α1,,αn)n\alpha=(\alpha_{1},\cdots,\alpha_{n})\in\mathbb{N}^{n}

(12) |Dαv(μ)|Cαμγ|α|,|{\rm D}^{\alpha}v(\mu)|\leq C_{\alpha}\langle\mu\rangle^{-\gamma-|\alpha|},

where Djv(μ):=v(μ+δj)v(μ){\rm D_{j}}v(\mu):=v(\mu+\delta_{j})-v(\mu), |α|:=j=1nαj|\alpha|:=\sum_{j=1}^{n}\alpha_{j}, and Dα:=D1α1Dnαn{\rm D}^{\alpha}:={\rm D}_{1}^{\alpha_{1}}...{\rm D}_{n}^{\alpha_{n}}.

Definition 3.1.

We call a perturbation VV admissible of order γ\gamma, with n>γ>0n>\gamma>0, if {vj}j=0nSγ(n)\{v_{j}\}_{j=0}^{n}\in S^{\gamma}(\mathbb{Z}^{n}) and for j=1,,nj=1,\dots,n

(13) vj(μ)=μγ(Γj+o(1)) as μ,v_{j}(\mu)=\langle\mu\rangle^{-\gamma}(\Gamma_{j}+o(1))\text{ as }\mu\to\infty,

with Γj0\Gamma_{j}\neq 0 for at least one jj.

This condition may look to be restrictive, but simplifies the presentation of the results. Naturally, alternative classes of symbols and asymptotic behaviours at infinity of the vjv_{j}’s could be addressed using akin methods to those employed in this article.

For an admissible perturbation we define the diagonal (n+1)×(n+1)(n+1)\times(n+1) matrix Γ\Gamma by

(14) Γll={Γl1 if Γl1>0,0 otherwise .\Gamma_{ll}=\begin{cases}\Gamma_{l-1}&\text{ if }\Gamma_{l-1}>0\ ,\\ 0&\text{ otherwise }.\end{cases}

We define as well the function M:𝕋nM(n+1)×(n+1)()M:\mathbb{T}^{n}\to M_{(n+1)\times(n+1)}(\mathbb{C}) by

(15) M(ξ):=1r(ξ)(00000r1(ξ)a2(ξ)a1(ξ)¯an(ξ)a1(ξ)¯0a1(ξ)a2(ξ)¯r2(ξ)an(ξ)a2(ξ)¯0a1(ξ)an(ξ)¯a2(ξ)an(ξ)¯rn(ξ)).M(\xi):=\frac{1}{r(\xi)}\begin{pmatrix}0&0&0&\cdots&0\\[5.0pt] 0&r_{1}(\xi)&-a_{2}(\xi)\overline{a_{1}(\xi)}&\cdots&-a_{n}(\xi)\overline{a_{1}(\xi)}\\[5.0pt] 0&-a_{1}(\xi)\overline{a_{2}(\xi)}&r_{2}(\xi)&\cdots&-a_{n}(\xi)\overline{a_{2}(\xi)}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&-a_{1}(\xi)\overline{a_{n}(\xi)}&-a_{2}(\xi)\overline{a_{n}(\xi)}&\cdots&r_{n}(\xi)\end{pmatrix}\ .
Theorem 3.2.

Assume that VV is an admissible perturbation of order γ\gamma. Define the constant 𝒞\mathcal{C} by

(16) 𝒞:=𝕋nTr((ΓM(ξ))nγ)𝑑ξ.\mathcal{C}:=\int_{\mathbb{T}^{n}}{\rm Tr}\left((\Gamma M(\xi))^{\frac{n}{\gamma}}\right)d\xi\ .

Let τn\tau_{n} denotes the volume of the unitary sphere in n\mathbb{R}^{n}. Then, the eigenvalue counting function satisfies

(17) 𝒩(λ)=λnγ(𝒞τn+o(1)),λ0.\mathcal{N}(\lambda)=\lambda^{-\frac{n}{\gamma}}\,\,(\mathcal{C}\,\tau_{n}+o(1)),\quad\lambda\downarrow 0\ .
Remark 3.3.

The best–known case of degenerate eigenvalues in the continuous setting is the Landau Hamiltonian on 2\mathbb{R}^{2}. Although they are not usually thought of as flat bands, the direct integral decomposition obtained from the Landau gauge gives us that each Landau level is the image of a constant band function in \mathbb{R}. In this sense, it is somewhat natural that the asymptotic order obtained in Equation˜17 coincides with the result in [Rai90, Theo. 2.6], see also Eq.˜24. However, the constants differ in both cases. For the Landau Hamiltonian, the constant depends only on the multiplicity of the corresponding Landau level and the intensity of the magnetic field, whereas for the discrete Dirac operator, the perturbation interacts with the associated eigenspace non–trivially as encoded by 𝒞\mathcal{C}.

4. Proof

In this section we will prove our main result ˜3.2. Before that, we start by recalling some known results on compact operators in order to settle notation and then reduce the study of 𝒩\mathcal{N} to the study of the eigenvalue counting function of an effective Hamiltonian.

4.1. Some notation and results on compact operators

Given the Hilbert spaces 1\mathcal{H}_{1} and 2\mathcal{H}_{2}, we denote by 𝔖(1,2)\mathfrak{S}_{\infty}(\mathcal{H}_{1},\mathcal{H}_{2}) the class of compact operators from 1\mathcal{H}_{1} to 2\mathcal{H}_{2}. When 1=2=\mathcal{H}_{1}=\mathcal{H}_{2}=\mathcal{H} we will just write 𝔖()\mathfrak{S}_{\infty}(\mathcal{H}). For K=K𝔖()K=K^{*}\in\mathfrak{S}_{\infty}(\mathcal{H}) and s>0s>0 we set

n±(s;K):=Rank𝟙(s,)(±K).n_{\pm}(s;K):=\operatorname{Rank}\mathds{1}_{(s,\infty)}(\pm K)\ .

Thus, the functions n±(;K)n_{\pm}(\cdot;K) are respectively the counting functions of the positive and negative eigenvalues of the operator KK. For K𝔖(1,2)K\in\mathfrak{S}_{\infty}(\mathcal{H}_{1},\mathcal{H}_{2}) we define

n(s;K):=n+(s2;KK),s>0;n_{*}(s;K):=n_{+}(s^{2};K^{*}K),\quad s>0;

thus n(;K)n_{*}(\cdot;K) is the counting function of the singular values of KK which, when ordered non–increasingly, we denote by {sj(K)}\{s_{j}(K)\}. Let KjK_{j}, j=1,2j=1,2 be self-adjoint compact operators. For s1,s2>0s_{1},s_{2}>0, we have the Weyl inequalities (see e.g. [BS87, Theorem 9.2.9])

(18) n±(s1+s2;K1+K2)n±(s1;K1)+n±(s2;K2).n_{\pm}(s_{1}+s_{2};K_{1}+K_{2})\leq n_{\pm}(s_{1};K_{1})+n_{\pm}(s_{2};K_{2})\ .

If instead we only have {K1,K2}𝔖(1,2)\{K_{1},K_{2}\}\subset\mathfrak{S}_{\infty}(\mathcal{H}_{1},\mathcal{H}_{2}) the Ky Fan inequality (see e.g. [BS87, Subsection 11.1.3]) gives

(19) n(s1+s2;K1+K2)n(s1;K1)+n(s2;K2).n_{*}(s_{1}+s_{2};K_{1}+K_{2})\leq n_{*}(s_{1};K_{1})+n_{*}(s_{2};K_{2})\ .

Further, for 0<p<0<p<\infty we define the class of compact operators 𝔖p,w\mathfrak{S}_{p,w} by

𝔖p,w:={K𝔖:sj(K)=O(j1/p)},\mathfrak{S}_{p,w}:=\{K\in\mathfrak{S}_{\infty}:s_{j}(K)=O(j^{-1/p})\}\ ,

together with the quasi-norm

||K||p,w:=supj{j1/psj(K)}=(sups>0{spn(s;K))1/p||K||_{p,w}:=\sup_{j}\{j^{1/p}s_{j}(K)\}=\left(\sup_{s>0}\{s^{p}n_{*}(s;K)\right)^{1/p}

that satisfies the "weakened triangle inequality"

K1+K2p,w21/p(Kp,w+Kp,w)||K_{1}+K_{2}||_{p,w}\leq 2^{1/p}(||K||_{p,w}+||K||_{p,w})

and the "weakened Hölder inequality"

(20) K1K2r,wc(p,q)K1p,wK2q,w,||K_{1}K_{2}||_{r,w}\leq c(p,q)||K_{1}||_{p,w}||K_{2}||_{q,w}\ ,

for r1=p1+q1r^{-1}=p^{-1}+q^{-1} and c(p,q)=(p/r)1/p(q/r)1/qc(p,q)=(p/r)^{1/p}(q/r)^{1/q} (see [BS87, Chapter 11]).

Finally, consider the set lp,wl_{p,w} of functions v:nv:\mathbb{Z}^{n}\to\mathbb{C} such that

#{μn:|v(μ)|>λ}=O(λp).\#\{\mu\in\mathbb{Z}^{n}:|v(\mu)|>\lambda\}=O(\lambda^{-p})\ .

Let us finish this section by considering the following result, which is a particular case of [BKS91, Theorem 4.8(ii)]

Proposition 4.1 (Cwikel-Birman-Solomyak).

Let p>2p>2 and assume vlp,wv\in l_{p,w} and fLp(𝕋n)f\in L^{p}(\mathbb{T}^{n}). Then fv𝔖p,w(2(n),L2(𝕋n))f\mathcal{F}v\in\mathfrak{S}_{p,w}(\ell^{2}(\mathbb{Z}^{n}),L^{2}(\mathbb{T}^{n})), and there exists a positive constant C(n)C(n) such that

fvp,wC(n)vlp,wfLp(𝕋n).||f\mathcal{F}v||_{p,w}\leq C(n)\|v\|_{l_{p,w}}\|f\|_{L^{p}(\mathbb{T}^{n})}.

4.2. Effective Hamiltomian

In this section we will use the notation

𝒩((a,b);T):=Rank𝟙(a,b)(T),\mathcal{N}((a,b);T):=\operatorname{Rank}\mathds{1}_{(a,b)}(T),

where a<ba<b and TT is a self-adjoint operator without essential spectrum in (a,b)(a,b). Following the approach coming from the study of magnetic Schrödinger operators, our aim is to study PVPPVP where PP stands for the projection on the flat band, i.e.,

(21) P:=𝟙{m}(H0).P:=\mathds{1}_{\{-m\}}(H_{0})\ .
Lemma 4.2.

Recall that M:𝕋nM(n+1)×(n+1)M:\mathbb{T}^{n}\to M_{(n+1)\times(n+1)} was defined in Equation˜15. Then

P=𝒰M𝒰.P=\mathscr{U}^{*}M\mathscr{U}\ .
Proof.

By Stone formula one can check that

P=slimκ01iπm0((H0siκ)1(H0s+iκ)1)𝑑s.P=\operatorname*{s\hskip 0.1pt-\hskip 0.1ptlim}_{\kappa\downarrow 0}\frac{1}{i\pi}\int_{-m}^{0}\bigl{(}(H_{0}-s-i\kappa)^{-1}-(H_{0}-s+i\kappa)^{-1}\bigr{)}ds\ .

Then, from Equation˜10 we need only to check that

limκ0m0(1m+λ+iκ1m+λiκ)𝑑λ=iπδm,\lim_{\kappa\downarrow 0}\int_{-m}^{0}\biggl{(}\frac{1}{m+\lambda+i\kappa}-\frac{1}{m+\lambda-i\kappa}\biggr{)}d\lambda=-i\pi\delta_{-m},

where δm\delta_{-m} is the Dirac delta function on m-m. ∎

Now, set P:=IPP^{\perp}:=I-P and for κ>0\kappa>0 define

Hκ±:=H0+P(V±κ|V|)P+P(V±κ1|V|)PH_{\kappa}^{\pm}:=H_{0}+P(V\pm\kappa|V|)P+P^{\perp}(V\pm\kappa^{-1}|V|)P^{\perp}

Then, by [PR11, Lemma 4.2]

(22) HκHHκ+.H_{\kappa}^{-}\leq H\leq H_{\kappa}^{+}.

Then, arguing as in the proof of [PR11, Theorem 4.1(ii)] we obtain that:

±𝒩(λ)\displaystyle\pm\mathcal{N}(\lambda)\leq ±𝒩((m+λ,0);mP+P(V±κ|V|)P)\displaystyle\pm\mathcal{N}((-m+\lambda,0);-mP+P(V\pm\kappa|V|)P)
(23) ±𝒩((m+λ,0);P(H0+(V±κ1|V|))P)+O(1).\displaystyle\pm\mathcal{N}((-m+\lambda,0);P^{\perp}(H_{0}+(V\pm\kappa^{-1}|V|))P^{\perp})+O(1).

In the next Lemma we treat the second term on the right of Equation˜23, showing that the perturbation VV interacts with the complement of the degenerated eigenspace only at a lesser order.

Lemma 4.3.
𝒩((m+λ,0);P(H0+V±κ1|V|)P)=o(λn/γ),λ0.\mathcal{N}((-m+\lambda,0);P^{\perp}(H_{0}+V\pm\kappa^{-1}|V|)P^{\perp})=o(\lambda^{-n/\gamma}),\quad\lambda\downarrow 0.
Proof.

Define the function W:𝒳+W:\mathcal{X}\to\mathbb{R}^{+} by wj(μ)=μγw_{j}(\mu)=\langle\mu\rangle^{-\gamma}, where we are using the notation of Eq.˜11. From Eq.˜12 there exist a constant C>0C>0 such that |V|CW|V|\leq CW. Denote by Wκ:=C(1+κ1)WW_{\kappa}:=C(1+\kappa^{-1})W. Then it can been seen that (again as in the proof of [Theorem 4.1(ii)][PR11])

𝒩((m+λ,0);P(H0+V±κ1|V|)P)𝒩((m+λ,0);P(H0+Wκ)P)+O(1).\mathcal{N}((-m+\lambda,0);P^{\perp}(H_{0}+V\pm\kappa^{-1}|V|)P^{\perp})\leq\mathcal{N}((-m+\lambda,0);P^{\perp}(H_{0}+W_{\kappa})P^{\perp})+O(1).

Now, by the Birman-Schwinger principle (see for instance [Kla83, Pus09]), we get for λ(0,m)\lambda\in(0,m)

𝒩((m+λ,0);P(H0+Wκ)P)=n+(1;PWκ1/2P(H0+mλ)1PWκ1/2P)+O(1).\mathcal{N}((-m+\lambda,0);P^{\perp}(H_{0}+W_{\kappa})P^{\perp})=n_{+}(1;P^{\perp}W_{\kappa}^{1/2}P^{\perp}(H_{0}+m-\lambda)^{-1}P^{\perp}W_{\kappa}^{1/2}P^{\perp})+O(1).

Define the (n+1)×(n+1)(n+1)\times(n+1) matrix

MR:=(λa1ana1¯λ2m0an¯0λ2m).M_{R}:=\begin{pmatrix}\lambda&a_{1}&\cdots&a_{n}\\[5.0pt] \overline{a_{1}}&\lambda-2m&\cdots&0\\[5.0pt] \vdots&\vdots&\ddots&\vdots\\ \overline{a_{n}}&0&\cdots&\lambda-2m\end{pmatrix}.

Then, from Eq.˜10 and Lemma˜4.2 it is not difficult to see that for λ(0,m)\lambda\in(0,m)

(H0+mλ)1P=𝒰MRr+λ(2mλ)(IdM)𝒰,(H_{0}+m-\lambda)^{-1}P^{\perp}=\mathscr{U}^{*}\frac{{M}_{R}}{r+\lambda(2m-\lambda)}({\rm Id}-M)\mathscr{U},

where Id{\rm Id} denotes the identity (n+1)×(n+1)(n+1)\times(n+1) matrix. Furthermore, the operator Wκ1/2P(H0+mλ)1PWκ1/2W_{\kappa}^{1/2}P^{\perp}(H_{0}+m-\lambda)^{-1}P^{\perp}W_{\kappa}^{1/2} is obviously compact and from Eq.˜20

Wκ1/2P(H0+mλ)1PWκ1/2n/γ,wCWκ1/2𝒰1r+λ(2mλ)2n/γ,w×MR(IdM)𝒰Wκ1/22n/γ,w.\|W_{\kappa}^{1/2}P^{\perp}(H_{0}+m-\lambda)^{-1}P^{\perp}W_{\kappa}^{1/2}\|_{n/\gamma,w}\leq C\|W_{\kappa}^{1/2}\mathscr{U}^{*}\frac{1}{r+\lambda(2m-\lambda)}\|_{2n/\gamma,w}\\ \times\|{M}_{R}({\rm Id}-M)\mathscr{U}W_{\kappa}^{1/2}\|_{2n/\gamma,w}.

Consider the operator Wκ1/2𝒰1r+λ(2mλ)W_{\kappa}^{1/2}\mathscr{U}^{*}\frac{1}{r+\lambda(2m-\lambda)}. Since 1r+λ(2mλ)\frac{1}{r+\lambda(2m-\lambda)} is bounded, it is in Lp(𝕋n)L^{p}(\mathbb{T}^{n}) for any p>1p>1. Further, each component of the multiplication operator Wκ1/2W_{\kappa}^{1/2} is in l2n/γ,wl_{2n/\gamma,w}. Then, since 2n/γ>22n/\gamma>2, by ˜4.1,

Wκ1/2𝒰1r+λ(2mλ)2n/γ,wC1r+λ(2mλ)L2nγ(𝕋n)Wκ1/2l=0nl2n/γ,w.\left\|W_{\kappa}^{1/2}\mathscr{U}^{*}\frac{1}{r+\lambda(2m-\lambda)}\right\|_{2n/\gamma,w}\leq C\left\|\frac{1}{r+\lambda(2m-\lambda)}\right\|_{L^{\frac{2n}{\gamma}}(\mathbb{T}^{n})}\|W_{\kappa}^{1/2}\|_{\oplus_{l=0}^{n}l_{2n/\gamma},w}.

To estimate the L2nγL^{\frac{2n}{\gamma}} norm we use the coarea formula

𝕋n|(r+λ(2mλ))1|2n/γ=\displaystyle\int_{\mathbb{T}^{n}}|(r+\lambda(2m-\lambda))^{-1}|^{2n/\gamma}= 01/21(ρ+λ(2mλ))2n/γr(ξ)=ρ1|r(ξ)|𝑑ξ𝑑ρ\displaystyle\int_{0}^{1/2}\frac{1}{(\rho+\lambda(2m-\lambda))^{2n/\gamma}}\int_{r(\xi)=\rho}\frac{1}{|\nabla r(\xi)|}d\xi d\rho
\displaystyle\leq C01/21(ρ+λ(2mλ))2n/γr(ξ)=ρ1r(ξ)1/2𝑑ξ𝑑ρ\displaystyle C\int_{0}^{1/2}\frac{1}{(\rho+\lambda(2m-\lambda))^{2n/\gamma}}\int_{r(\xi)=\rho}\frac{1}{r(\xi)^{1/2}}d\xi d\rho
\displaystyle\leq C01/21(ρ+λ(2mλ))2n/γ1ρ1/2r(ξ)=ρ𝑑ξ𝑑ρ\displaystyle C\int_{0}^{1/2}\frac{1}{(\rho+\lambda(2m-\lambda))^{2n/\gamma}}\frac{1}{\rho^{1/2}}\int_{r(\xi)=\rho}d\xi d\rho
\displaystyle\leq C01/2ρn/21(ρ+λ(2mλ))2n/γ𝑑ρ\displaystyle C\int_{0}^{1/2}\frac{\rho^{n/2-1}}{(\rho+\lambda(2m-\lambda))^{2n/\gamma}}d\rho
\displaystyle\leq Cλ2n/γ+n/2,\displaystyle C\lambda^{-2n/\gamma+n/2},

where in the first and third inequalities have used Eq.˜9. Analogously,

MR(IdM)𝒰Wκ1/22n/γ,wCWκ1/2l=0nl2n/γ,w,\left\|{M}_{R}({\rm Id}-M)\mathscr{U}W_{\kappa}^{1/2}\right\|_{2n/\gamma,w}\leq C\|W_{\kappa}^{1/2}\|_{\oplus_{l=0}^{n}l_{2n/\gamma},w},

since the matrix MR(IdM){M}_{R}({\rm Id}-M) is bounded with uniform bound in λ\lambda. Putting all this together we obtain

Wκ1/2P(H0+mλ)1PWκ1/2n/γ,wCλγ/41,\|W_{\kappa}^{1/2}P^{\perp}(H_{0}+m-\lambda)^{-1}P^{\perp}W_{\kappa}^{1/2}\|_{n/\gamma,w}\leq C\lambda^{\gamma/4-1},

which is equivalent to say that

n(s;Wκ1/2P(H0+mλ)1PWκ1/2)\displaystyle n_{*}(s;W_{\kappa}^{1/2}P^{\perp}(H_{0}+m-\lambda)^{-1}P^{\perp}W_{\kappa}^{1/2})\leq Cλn/γ+n/4\displaystyle C\lambda^{-n/\gamma+n/4}
=\displaystyle= o(λn/γ).\displaystyle o(\lambda^{-n/\gamma}).\qed

4.3. Eigenvalue counting function for the effective Hamiltonian

From Eqs.˜23 and 4.3

(24) ±𝒩(λ)±𝒩((λ,m);P(V±κ|V|)P)+o(λn/γ),λ0.\pm\mathcal{N}(\lambda)\leq\pm\mathcal{N}((\lambda,m);P(V\pm\kappa|V|)P)+o(\lambda^{-n/\gamma}),\quad\lambda\downarrow 0\ .

Then, we are led to study the distribution of positives eigenvalues of the compact operator P(V±κ|V|)PP(V\pm\kappa|V|)P.

For ease of notation, for any κ>0\kappa>0 we define Tκ±T_{\kappa}^{\pm} in 𝔖()\mathfrak{S}_{\infty}(\mathcal{H}) by

Tκ±:=𝒰P(V±κ|V|)P𝒰=M𝒰(V±κ|V|)𝒰M.T^{\pm}_{\kappa}:=\mathscr{U}P(V\pm\kappa|V|)P\mathscr{U}^{*}=M\mathscr{U}(V\pm\kappa|V|)\mathscr{U}^{*}M\ .
Proposition 4.4.

For an admisible VV we have

n+(λ;Tκ±)=(1±κλ)n/γτn𝕋nTr((M(ξ)ΓM(ξ))n/γ)𝑑ξ(1+o(1)),λ0.n_{+}(\lambda;T^{\pm}_{\kappa})=\Big{(}\frac{1\pm\kappa}{\lambda}\Big{)}^{n/\gamma}\tau_{n}\int_{\mathbb{T}^{n}}{\rm Tr}\left({(M(\xi)\Gamma{M(\xi)})^{n/\gamma}}\right)d\xi\,(1+o(1)),\quad\lambda\downarrow 0\ .

In order to proof this Section we follow the ideas of [MPR23, Theorem 6.1], which in turn are inspired by the proof of [BS70, Theorem 1]. By analogy, we denote :=[0,1)nn\Box:=[0,1)^{n}\subset\mathbb{R}^{n} and hence j=0nL2()\mathcal{H}\cong\oplus^{n}_{j=0}L^{2}(\Box). Finally, for ease of notation, let us set V^κ±=𝒰(V±κ|V|)𝒰\hat{V}^{\pm}_{\kappa}=\mathscr{U}(V\pm\kappa|V|)\mathscr{U}^{*}.

Remark 4.5.

The statement of ˜4.4 is particular to our effective Hamiltonian and problem. However, in the proof we use only that MLp(𝕋d;n+1)M\in L^{p}(\mathbb{T}^{d};\mathbb{C}^{n+1}) for p>2p>2 and we could also replace V±κ|V|V\pm\kappa|V| with another potential satisfying Equations˜12 and 13. A similar statement holds for nn_{-}.

Lemma 4.6.

Let XX and YY be two subsets of \Box with no interior points in common. Then

n(r;𝟙XV^κ±𝟙Y)=o(rn/γ),r0.n_{*}(r;\mathds{1}_{X}\hat{V}^{\pm}_{\kappa}\mathds{1}_{Y})=o(r^{-n/\gamma}),\quad r\downarrow 0.
Proof.

The proof uses ˜4.1 and is almost equal to the proof of [MPR23, Lemma 6.4]. ∎

Lemma 4.7.

Let {j}\{\Box_{j}\} be a partition of \Box into cubes of equal size 1/qn1/q^{n}, q+q\in\mathbb{Z}_{+}, and let {Bj}j=1qn\{B_{j}\}_{j=1}^{q^{n}} be matrices in M(n+1)×(n+1)()M_{(n+1)\times(n+1)}(\mathbb{C}). Let Tˇκ±:j=0nL2()j=0nL2()\check{T}_{\kappa}^{\pm}:\oplus^{n}_{j=0}L^{2}(\Box)\to\oplus^{n}_{j=0}L^{2}(\Box) be the operator defined by

Tˇκ±=jBj𝟙jV^κ±𝟙jBj.\check{T}^{\pm}_{\kappa}=\sum_{j}B_{j}\mathds{1}_{\Box_{j}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{j}}B_{j}^{*}.

Then, for any δ(0,1)\delta\in(0,1),

τnqnjTr((BjΓ(1±κ)Bj)n/γ)(λ(1+δ))n/γ(1+o(1))\displaystyle\frac{\tau_{n}}{q^{n}}\sum_{j}{\rm Tr}((B_{j}\Gamma(1\pm\kappa)B_{j}^{*})^{n/\gamma})\,(\lambda(1+\delta))^{-n/\gamma}(1+o(1))
\displaystyle\leq n+(λ;Tˇκ±)\displaystyle n_{+}(\lambda;\check{T}^{\pm}_{\kappa})
\displaystyle\leq τnqnjTr((BjΓ(1±κ)Bj)n/γ)(λ(1δ))n/γ(1+o(1)),λ0.\displaystyle\frac{\tau_{n}}{q^{n}}\sum_{j}{\rm Tr}((B_{j}\Gamma(1\pm\kappa)B_{j}^{*})^{n/\gamma})\,(\lambda(1-\delta))^{-n/\gamma}(1+o(1)),\quad\lambda\downarrow 0.
Proof.

We will show the proof of the upper bound. The lower bound is similar. Let B0B_{0} be a constant (n+1)×(n+1)(n+1)\times(n+1) matrix. Then for any δ(0,1)\delta\in(0,1)

n+(λ;B0V^κ±B0)\displaystyle n_{+}(\lambda;B_{0}\hat{V}^{\pm}_{\kappa}B_{0}^{*})\geq jn+(λ(1+δ);B0𝟙jV^κ±𝟙jB0)n(λδ;jlB0𝟙jV^κ±𝟙lB0)\displaystyle\sum_{j}n_{+}(\lambda(1+\delta);B_{0}\mathds{1}_{\Box_{j}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{j}}B_{0}^{*})-n_{-}(\lambda\delta;\sum_{j\neq l}B_{0}\mathds{1}_{\Box_{j}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{l}}B_{0}^{*})
=\displaystyle= qnn+(λ(1+δ);B0𝟙0V^κ±𝟙0B0)+o(λn/γ),\displaystyle q^{n}n_{+}(\lambda(1+\delta);B_{0}\mathds{1}_{\Box_{0}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{0}}B_{0}^{*})+o(\lambda^{-n/\gamma}),

where for the inequality we used Equation˜18. For the equality we used first the fact that each operator B0𝟙jV^κ±𝟙jB0B_{0}\mathds{1}_{\Box_{j}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{j}}B_{0}^{*} is unitary equivalent to B0𝟙0V^κ±𝟙0B0B_{0}\mathds{1}_{\Box_{0}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{0}}B_{0}^{*}, for 0=(0,1/q)n\Box_{0}=(0,1/q)^{n}. Then we used and Lemmas˜4.6 and 19. It follows that

(25) n+(λ;B0𝟙0V^κ±𝟙0B0)1qnn+(λ(1δ);B0V^κ±B0)+o(λn/γ).n_{+}(\lambda;B_{0}\mathds{1}_{\Box_{0}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{0}}B_{0}^{*})\leq\frac{1}{q^{n}}n_{+}(\lambda(1-\delta);B_{0}\hat{V}^{\pm}_{\kappa}B_{0}^{*})+o(\lambda^{-n/\gamma})\ .

Define v(μ)=μγv_{*}(\mu)=\langle\mu\rangle^{-\gamma}. One can check that

(26) #{μn:v(μ)>λ}=τmλn/γ(1+o(1)),λ0,\#\{\mu\in\mathbb{Z}^{n}:v_{*}(\mu)>\lambda\}=\tau_{m}\lambda^{-n/\gamma}(1+o(1)),\quad\lambda\downarrow 0\ ,

see for instance [RS78, Prop. 2 XIII.15].

From Eq.˜13 set V^0±:=𝒰Γ(1±κ)μγ𝒰\hat{V}^{\pm}_{0}:=\mathscr{U}\Gamma(1\pm\kappa)\langle\mu\rangle^{-\gamma}\mathscr{U}^{*} and use the Weyl inequalities Eq.˜18 to obtain that for δ~(0,1)\tilde{\delta}\in(0,1)

n+(λ;B0V^κ±B0)n+(λ(1δ~);B0V^0±B0)+n+(λδ~;B0(V^κ±V^0±)B0).n_{+}(\lambda;B_{0}\hat{V}^{\pm}_{\kappa}B_{0}^{*})\\ \leq n_{+}(\lambda(1-\tilde{\delta});B_{0}\hat{V}^{\pm}_{0}B_{0}^{*})+n_{+}(\lambda\tilde{\delta};B_{0}(\hat{V}^{\pm}_{\kappa}-\hat{V}^{\pm}_{0})B_{0}^{*}).

Now, denote by {β0,l}\{\beta_{0,l}\} the eigenvalues of the matrix B0ΓB0B_{0}\Gamma B_{0}^{*}. We have that the eigenvalues of B0V^0±B0B_{0}\hat{V}_{0}^{\pm}B_{0}^{*} are given by

{β0,lv(μ):1lk,μd}.\{\beta_{0,l}\,v_{*}(\mu):1\leq l\leq k,\mu\in\mathbb{Z}^{d}\}.

Thus Eq.˜26 implies that

n+(λ;B0V^0±B0)\displaystyle n_{+}(\lambda;B_{0}\hat{V}_{0}^{\pm}B_{0}^{*}) =#{1lk,μn:β0,lv(μ)>λ}\displaystyle=\#\{1\leq l\leq k,\mu\in\mathbb{Z}^{n}:\beta_{0,l}\,v_{*}(\mu)>\lambda\}
=β0,l>0n+(λ/β0,l;v)\displaystyle=\sum_{\beta_{0,l}>0}n_{+}(\lambda/\beta_{0,l};v_{*})
=(τnβ0,l>0β0,ln/γ)λn/γ(1+o(1)),λ0.\displaystyle=\Big{(}\tau_{n}\sum_{\beta_{0,l}>0}\beta_{0,l}^{n/\gamma}\Big{)}\lambda^{-n/\gamma}(1+o(1)),\quad\lambda\downarrow 0.

The same reasoning can be used to show that n+(B0(V^κ±V^0)B0)=o(λn/γ)n_{+}(B_{0}(\hat{V}^{\pm}_{\kappa}-\hat{V}_{0})B^{*}_{0})=o(\lambda^{-n/\gamma}). Putting the previous inequalities together, for all δ,δ~(0,1)\delta,\tilde{\delta}\in(0,1)

n+(λ;Tˇκ±)\displaystyle n_{+}(\lambda;\check{T}^{\pm}_{\kappa}) =jn+(λ;Bj𝟙jV^κ±𝟙jBj)\displaystyle=\sum_{j}n_{+}(\lambda;B_{j}\mathds{1}_{\Box_{j}}\hat{V}^{\pm}_{\kappa}\mathds{1}_{\Box_{j}}B_{j}^{*})
1qnjn+(λ(1δ);BjV^κ±Bj)+o(λn/γ)\displaystyle\leq\frac{1}{q^{n}}\sum_{j}n_{+}(\lambda(1-\delta);B_{j}\hat{V}^{\pm}_{\kappa}B_{j}^{*})+o(\lambda^{-n/\gamma})
1qnjn+(λ(1δ)(1+δ~);BjV^0±Bj)+o(λn/γ)\displaystyle\leq\frac{1}{q^{n}}\sum_{j}n_{+}\Big{(}\frac{\lambda(1-\delta)}{(1+\tilde{\delta})};B_{j}\hat{V}_{0}^{\pm}B_{j}^{*}\Big{)}+o(\lambda^{-n/\gamma})
=τnqnjTr((BjΓ(1±κ)Bj)n/γ)(λ(1δ)(1+δ~))n/γ(1+o(1)),λ0.\displaystyle=\frac{\tau_{n}}{q^{n}}\sum_{j}{\rm Tr}\Big{(}(B_{j}\Gamma(1\pm\kappa)B_{j}^{*})^{n/\gamma}\Big{)}\,\left(\frac{\lambda(1-\delta)}{(1+\tilde{\delta})}\right)^{-n/\gamma}(1+o(1)),\quad\lambda\downarrow 0\ .\qed
Proof of ˜4.4.

Let ε>0\varepsilon>0, and take Bε=jBε,j𝟙ε,jB_{\varepsilon}=\sum_{j}B_{\varepsilon,j}\mathds{1}_{\Box_{\varepsilon,j}} a step matrix function such that MBεLp(𝕋n)<ε\|M-B_{\varepsilon}\|_{L^{p}(\mathbb{T}^{n})}<\varepsilon. Assume that the size of each cube ε,j\Box_{\varepsilon,j} is 1/qn1/q^{n} as in the previous lemma.

Take Sε:=BεV^κ±BεS_{\varepsilon}:=B_{\varepsilon}\hat{V}^{\pm}_{\kappa}B_{\varepsilon}^{*}. Then by ˜4.1 Tκ±Sεn/γ,w<Cε\|T_{\kappa}^{\pm}-S_{\varepsilon}\|_{n/\gamma,w}<C\varepsilon, which means that

(27) n(λ;Tκ±Sε)(Cε)n/γλn/γ.n_{*}(\lambda;T_{\kappa}^{\pm}-S_{\varepsilon})\leq(C\varepsilon)^{n/\gamma}\lambda^{-n/\gamma}.

Also, let Tˇε,κ±=jBε,j𝟙ε,jV^κ±Bε,j𝟙ε,j\check{T}_{\varepsilon,\kappa}^{\pm}=\sum_{j}B_{\varepsilon,j}\mathds{1}_{\Box_{\varepsilon,j}}\hat{V}^{\pm}_{\kappa}B_{\varepsilon,j}\mathds{1}_{\Box_{\varepsilon,j}}. Thus, by Lemma˜4.6

(28) n(s;SεTˇε,κ±)=o(sn/γ).n_{*}(s;S_{\varepsilon}-\check{T}_{\varepsilon,\kappa}^{\pm})=o(s^{-n/\gamma}).

Now, using Lemma˜4.7, we have that for any δ(0,1)\delta\in(0,1)

τnqnjTr((Bε,jΓ(1±κ)Bε,j)n/γ)(λ(1+δ))n/γ(1+o(1))\displaystyle\frac{\tau_{n}}{q^{n}}\sum_{j}{\rm Tr}((B_{\varepsilon,j}\Gamma(1\pm\kappa)B_{\varepsilon,j}^{*})^{n/\gamma})\,(\lambda(1+\delta))^{-n/\gamma}(1+o(1))
(29) \displaystyle\leq n+(λ;Tˇε,κ±)\displaystyle n_{+}(\lambda;\check{T}_{\varepsilon,\kappa}^{\pm})
\displaystyle\leq τnqnjTr((Bε,jΓ(1±κ)Bε,j)n/γ)(λ(1δ))n/γ(1+o(1)),λ0.\displaystyle\frac{\tau_{n}}{q^{n}}\sum_{j}{\rm Tr}((B_{\varepsilon,j}\Gamma(1\pm\kappa)B_{\varepsilon,j}^{*})^{n/\gamma})\,(\lambda(1-\delta))^{-n/\gamma}(1+o(1)),\quad\lambda\downarrow 0\ .

Finally, putting together Eqs.˜27, 28, 4.3, 18 and 19, and making λ\lambda, δ\delta and ε\varepsilon goes to 0, we finish the proof. ∎

Proof of ˜3.2.

The result follows from ˜4.4 by taking κ0\kappa\downarrow 0, Equation˜24 and using the cyclicity of the trace. ∎

5. The Laplacian on a particular 2\mathbb{Z}^{2}-periodic graph

5.1. A simple example of a 2\mathbb{Z}^{2}-periodic graph with a flat band

Let us start by briefly recalling some notions from the periodic graph theory, we refer to [Sun13, KS14, PR18] for more details. We say that a graph is d\mathbb{Z}^{d}– periodic if it admits an action of d\mathbb{Z}^{d} by graph–automorphisms. By fixing representatives of each orbit of vertices for this action we can define the entire part of a vertex by xxˇ=x\lfloor x\rfloor\check{x}=x where xˇ\check{x} is the representative of the orbit of xx. Then, the index of an oriented edge e=(x,y)\mathrm{e}=(x,y) is just η(e)=yx\eta(\mathrm{e})=\lfloor y\rfloor-\lfloor x\rfloor. Note that η\eta is d\mathbb{Z}^{d}–periodic and hence we can refer to the index of an edge in the quotient graph.

Let us now denote by 𝒳~=(𝒱~,𝒜~)\tilde{\mathcal{X}}=(\tilde{\mathcal{V}},\tilde{\mathcal{A}}) the graph obtained from 2\mathbb{Z}^{2} by adding a vertex on each edge with trivial weights (see Figure˜2(a)). The quotient graph obtained by the action of 2\mathbb{Z}^{2} is composed by three vertices and four edges as presented in Figure˜2(b). If we takes a representatives the vertices (0,0)(0,0), (0,12)(0,\tfrac{1}{2}) and (12,0)(\tfrac{1}{2},0) One can easily check that η(e1)=η(e2)=(0,0)\eta(\mathrm{e}_{1})=\eta(\mathrm{e}_{2})=(0,0) while η(e3)=(1,0)\eta(\mathrm{e}_{3})=(1,0) and η(e4)=(0,1)\eta(\mathrm{e}_{4})=(0,1).

(a) The periodic graph obtained from 2\mathbb{Z}^{2} by adding a vertex to each edge.
x0,0x_{0,0}x1,0x_{1,0}x0,1x_{0,1}e1\mathrm{e}_{1}e4\mathrm{e}_{4}e3\mathrm{e}_{3}e2\mathrm{e}_{2}
(b) The quotient graph by the usual action of 2\mathbb{Z}^{2}.

Set H~0=Δ0\tilde{H}_{0}=-\Delta_{0}, where Δ0\Delta_{0} is the usual graph Laplacian, i.e. , for f2(𝒱~)f\in\ell^{2}(\tilde{\mathcal{V}}) and x𝒱~x\in\tilde{\mathcal{V}}:

(H~0f)(x)=e𝒜~,e=(x,y)f(x)f(y).(\tilde{H}_{0}f)(x)=\sum_{\mathrm{e}\in\tilde{\mathcal{A}},\mathrm{e}=(x,y)}f(x)-f(y)\ .

Hence, by defining a~j=1+e2πiξj\tilde{a}_{j}=1+e^{2\pi i\xi_{j}}, for j=1,2j=1,2, we obtain the following representation of the graph Laplacian as a matrix-valued multiplication operator.

Proposition 5.1 ([PR18, Prop. 4.7]).

There exists a unitary operator 𝒰~:(𝒱~)L2(𝕋2;n)\tilde{\mathscr{U}}:\ell(\tilde{\mathcal{V}})\to L^{2}(\mathbb{T}^{2};\mathbb{C}^{n}) such that

𝒰~(H~0)𝒰~=h~0\tilde{\mathscr{U}}(\tilde{H}_{0})\tilde{\mathscr{U}}^{*}=\tilde{h}_{0}

where h0h_{0} denotes the multiplication operator by the real analytic function

h~0:𝕋2M3×3()\tilde{h}_{0}:\mathbb{T}^{2}\to M_{3\times 3}(\mathbb{C})

on L2(𝕋2,3)L^{2}(\mathbb{T}^{2},\mathbb{C}^{3}) given by

(30) h~0(ξ)=(4a~1(ξ)a~2(ξ)a~1(ξ)¯20a~2(ξ)¯02).\tilde{h}_{0}(\xi)=\begin{pmatrix}4&-\tilde{a}_{1}(\xi)&-\tilde{a}_{2}(\xi)\\ -\overline{\tilde{a}_{1}(\xi)}&2&0\\ -\overline{\tilde{a}_{2}(\xi)}&0&2\end{pmatrix}\ .

Setting as before r~(ξ)=|a~1(ξ)|2+|a~2(ξ)|2\tilde{r}(\xi)=|\tilde{a}_{1}(\xi)|^{2}+|\tilde{a}_{2}(\xi)|^{2}, and noticing

|a~j(ξ)|2=2+2cos(2πξj)=4cos2(πξj)|\tilde{a}_{j}(\xi)|^{2}=2+2\cos(2\pi\xi_{j})=4\cos^{2}(\pi\xi_{j})

we can obtain the associated characteristic polynomial to h~0\tilde{h}_{0}

p~z(ξ)=(2z)(z26z+8r~(ξ))\tilde{p}_{z}(\xi)=(2-z)(z^{2}-6z+8-\tilde{r}(\xi))

and the corresponding non constant band functions

z~±(ξ)=3±1+r~(ξ).\tilde{z}_{\pm}(\xi)=3\pm\sqrt{1+\tilde{r}(\xi)}\ .

It follows that the spectrum satisfies

(31) σ(H~0)=σess(H~0)=σac(H~0)=[0,2][4,6]\sigma(\tilde{H}_{0})=\sigma_{ess}(\tilde{H}_{0})=\sigma_{ac}(\tilde{H}_{0})=[0,2]\bigcup[4,6]

with 22 an embedded degenerated eigenvalue. Given V~:𝒱~\tilde{V}:\tilde{\mathcal{V}}\to\mathbb{R} we define the Schrödinger operator

H~=H~0+V~\tilde{H}=\tilde{H}_{0}+\tilde{V}

and the corresponding eigenvalue counting function by

𝒩~(λ)=Rank𝟙(2+λ,3)(H),\tilde{\mathcal{N}}(\lambda)={\rm Rank}\mathds{1}_{(2+\lambda,3)}(H)\ ,

for λ(0,1)\lambda\in(0,1). As before, by taking the limit λ0\lambda\downarrow 0 we will be able to study the accumulation of eigenvalues near the perturbed flat band.

Remark 5.2.

An attentive reader can wonder why this Laplacian operator show the same spectral properties than the Dirac operator studied in previous sections. From a purely computational point of view, the similarities with H0H_{0} can be deduced from the fact that the symbol on 𝕋2\mathbb{T}^{2} of H~03\tilde{H}_{0}-3 correspond to the symbol of H0H_{0} with m=1m=1 by replacing aja_{j} with a~j-\tilde{a}_{j}. In general, one can say that the clear distinction of the order of a differential operator gets muddy in the discrete case, see for instance the discussion related to the continuum limit of discrete Dirac operators [Nak24, CGJ22].

5.2. Admissible perturbations and eigenvalue asymptotics.

Let us start by noticing that for every μ2\mu\in\mathbb{Z}^{2} we can define fμ2(𝒱~)f_{\mu}\in\ell^{2}(\tilde{\mathcal{V}}) by

fμ(x)={1 if x=μ+(12,0),1 if x=μ+(0,12),0 else.f_{\mu}(x)=\begin{cases}1&\text{ if }x=\mu+(\tfrac{1}{2},0)\ ,\\ -1&\text{ if }x=\mu+(0,\tfrac{1}{2})\ ,\\ 0&\text{ else.}\end{cases}

and it satisfies H0fμ=2fμH_{0}f_{\mu}=2f_{\mu}. Hence, if we decompose 2(𝒱~)\ell^{2}(\tilde{\mathcal{V}}) by

2(𝒱~)2(2)2(2+(12,0))2(2+(0,12))\ell^{2}(\tilde{\mathcal{V}})\cong\ell^{2}(\mathbb{Z}^{2})\oplus\ell^{2}(\mathbb{Z}^{2}+(\tfrac{1}{2},0))\oplus\ell^{2}(\mathbb{Z}^{2}+(0,\tfrac{1}{2}))

we have that

Ker(H02){𝟎}2(2+(12,0))2(2+(0,12)).\operatorname{Ker}(H_{0}-2)\leq\{\mathbf{0}\}\oplus\ell^{2}(\mathbb{Z}^{2}+(\tfrac{1}{2},0))\oplus\ell^{2}(\mathbb{Z}^{2}+(0,\tfrac{1}{2}))\ .

Then, if we define v~j:2\tilde{v}_{j}:\mathbb{Z}^{2}\to\mathbb{R}, for j{0,1,2}j\in\{0,1,2\} by

v~0(μ)=V~(μ),v~1(μ)=V~(μ+(12,0)) and v~2(μ)=V~(μ+(0,12)),\tilde{v}_{0}(\mu)=\tilde{V}(\mu)\quad,\quad\tilde{v}_{1}(\mu)=\tilde{V}(\mu+(\tfrac{1}{2},0))\quad\text{ and }\quad\tilde{v}_{2}(\mu)=\tilde{V}(\mu+(0,\tfrac{1}{2}))\ ,

we can apply ˜3.1 to V~\tilde{V}.

Let us now observe that for any zσ(H~0)z\notin\sigma(\tilde{H}_{0})

(h~0z)1=1pz((2z)2a~1(2z)a~2(2z)a~1¯(2z)(2z)(4z)|a2|2a~1¯a~2a~2¯(2z)a~1a~2¯(2z)(4z)|a1|2)=1(z26z+8r~)((2z)a~1a~2a~1¯(4z)0a~2¯0(4z))+1pz(0000|a2|2a~1¯a~20a~1a~2¯|a1|2).(\tilde{h}_{0}-z)^{-1}=\frac{1}{p_{z}}\begin{pmatrix}(2-z)^{2}&\tilde{a}_{1}(2-z)&\tilde{a}_{2}(2-z)\\ \overline{\tilde{a}_{1}}(2-z)&(2-z)(4-z)-|a_{2}|^{2}&\overline{\tilde{a}_{1}}\tilde{a}_{2}\\ \overline{\tilde{a}_{2}}(2-z)&\tilde{a}_{1}\overline{\tilde{a}_{2}}&(2-z)(4-z)-|a_{1}|^{2}\end{pmatrix}\\ =\frac{1}{(z^{2}-6z+8-\tilde{r})}\begin{pmatrix}(2-z)&\tilde{a}_{1}&\tilde{a}_{2}\\ \overline{\tilde{a}_{1}}&(4-z)&0\\ \overline{\tilde{a}_{2}}&0&(4-z)\end{pmatrix}+\frac{1}{p_{z}}\begin{pmatrix}0&0&0\\ 0&-|a_{2}|^{2}&\overline{\tilde{a}_{1}}\tilde{a}_{2}\\ 0&\tilde{a}_{1}\overline{\tilde{a}_{2}}&-|a_{1}|^{2}\end{pmatrix}\ .

Hence, we define M~:𝕋2M3×3()\tilde{M}:\mathbb{T}^{2}\to M_{3\times 3}(\mathbb{C}) by

(32) M~:=1r~(0000|a~2|2a~2a~1¯0a~1a~2¯|a~1|2).\tilde{M}:=\frac{1}{\tilde{r}}\begin{pmatrix}0&0&0\\ 0&|\tilde{a}_{2}|^{2}&-\tilde{a}_{2}\overline{\tilde{a}_{1}}\\ 0&-\tilde{a}_{1}\overline{\tilde{a}_{2}}&|\tilde{a}_{1}|^{2}\end{pmatrix}\ .
Theorem 5.3.

Assume that V~\tilde{V} is an admissible perturbation of order γ\gamma and associate 3×33\times 3 matrix Γ~\tilde{\Gamma}. Define the constant 𝒞~\tilde{\mathcal{C}} by

(33) 𝒞~:=𝕋2Tr((ΓM~(ξ))nγ)𝑑ξ.\tilde{\mathcal{C}}:=\int_{\mathbb{T}^{2}}{\rm Tr}\left((\Gamma\tilde{M}(\xi))^{\frac{n}{\gamma}}\right)d\xi\ .

Then, the eigenvalue counting function satisfies

(34) 𝒩~(λ)=λnγ(𝒞~τn+o(1)),λ0.\tilde{\mathcal{N}}(\lambda)=\lambda^{-\frac{n}{\gamma}}\,\,(\tilde{\mathcal{C}}\,\tau_{n}+o(1)),\quad\lambda\downarrow 0\ .

Acknowledgments

P. Miranda was supported by the Chilean Fondecyt Grant 1201857. D. Parra was partially supported by the Chilean Fondecyt Grant 3210686 and Universidad de La Frontera, Apoyo PF24-0027. Both authors gratefully acknowledge the hospitality of the Institut de Mathématiques de Bordeaux were the final draft of this manuscript was prepared.

References

  • [AT15] Colette Anné and Nabila Torki-Hamza “The Gauss-Bonnet operator of an infinite graph” In Anal. Math. Phys. 5.2, 2015, pp. 137–159 DOI: 10.1007/s13324-014-0090-0
  • [BL13] Emil J. Bergholtz and Zhao Liu “Topological flat band models and fractional Chern insulators” In Internat. J. Modern Phys. B 27.24, 2013, pp. 1330017\bibrangessep43 DOI: 10.1142/S021797921330017X
  • [BKS91] M.. Birman, G.. Karadzhov and M.. Solomyak “Boundedness conditions and spectrum estimates for the operators b(X)a(D)b(X)a(D) and their analogs” In Estimates and asymptotics for discrete spectra of integral and differential equations (Leningrad, 1989–90) 7, Adv. Soviet Math. Amer. Math. Soc., Providence, RI, 1991, pp. 85–106
  • [BS70] M.. Birman and M.. Solomjak “Asymptotics of the spectrum of weakly polar integral operators” In Izv. Akad. Nauk SSSR Ser. Mat. 34, 1970, pp. 1142–1158
  • [BS87] M.. Birman and M.. Solomjak “Spectral theory of selfadjoint operators in Hilbert space” Translated from the 1980 Russian original by S. Khrushchëv and V. Peller, Mathematics and its Applications (Soviet Series) D. Reidel Publishing Co., Dordrecht, 1987, pp. xv+301
  • [CGJ22] Horia Cornean, Henrik Garde and Arne Jensen “Discrete approximations to Dirac operators and norm resolvent convergence” In J. Spectr. Theory 12.4, 2022, pp. 1589–1622 DOI: 10.4171/jst/438
  • [GZ23] Jeffrey Galkowski and Maciej Zworski “An abstract formulation of the flat band condition”, 2023 eprint: arXiv:2307.04896
  • [IT98] Akira Iwatsuka and Hideo Tamura “Asymptotic distribution of eigenvalues for Pauli operators with nonconstant magnetic fields” In Duke Math. J. 93.3, 1998, pp. 535–574 DOI: 10.1215/S0012-7094-98-09319-X
  • [KTW23] Joachim Kerner, Matthias Täufer and Jens Wintermayr “Robustness of Flat Bands on the Perturbed Kagome and the Perturbed Super-Kagome Lattice” In Annales Henri Poincaré, 2023 DOI: 10.1007/s00023-023-01399-7
  • [Kla83] Martin Klaus “Some applications of the Birman-Schwinger principle” In Helv. Phys. Acta 55.1, 1982/83, pp. 49–68
  • [Kol+20] Alicia J. Kollár, Mattias Fitzpatrick, Peter Sarnak and Andrew A. Houck “Line-graph lattices: Euclidean and non-Euclidean flat bands, and implementations in circuit quantum electrodynamics” In Comm. Math. Phys. 376.3, 2020, pp. 1909–1956 DOI: 10.1007/s00220-019-03645-8
  • [KS14] Evgeny Korotyaev and Natalia Saburova “Schrödinger operators on periodic discrete graphs” In J. Math. Anal. Appl. 420.1, 2014, pp. 576–611 DOI: 10.1016/j.jmaa.2014.05.088
  • [Kuc91] P.. Kuchment “On the Floquet theory of periodic difference equations” In Geometrical and algebraical aspects in several complex variables (Cetraro, 1989) 8, Sem. Conf. EditEl, Rende, 1991, pp. 201–209
  • [MPR23] Pablo Miranda, Daniel Parra and Georgi Raikov “Spectral asymptotics at thresholds for a Dirac-type operator on 2\mathbb{Z}^{2} In J. Funct. Anal. 284.2, 2023, pp. Paper No. 109743 DOI: 10.1016/j.jfa.2022.109743
  • [Nak24] Shu Nakamura “Remarks on discrete Dirac operators and their continuum limits” In J. Spectr. Theory, 2024, pp. 1–15 DOI: 10.4171/JST/478
  • [Par17] D. Parra “Spectral and scattering theory for Gauss-Bonnet operators on perturbed topological crystals” In J. Math. Anal. Appl. 452.2, 2017, pp. 792–813 DOI: 10.1016/j.jmaa.2017.03.002
  • [PR18] D. Parra and S. Richard “Spectral and scattering theory for Schrödinger operators on perturbed topological crystals” In Rev. Math. Phys. 30.4, 2018, pp. 1850009\bibrangessep39 DOI: 10.1142/S0129055X18500095
  • [Pus09] Alexander Pushnitski “Operator theoretic methods for the eigenvalue counting function in spectral gaps” In Ann. Henri Poincaré 10.4, 2009, pp. 793–822 DOI: 10.1007/s00023-009-0422-z
  • [PR11] Alexander Pushnitski and Grigori Rozenblum “On the spectrum of Bargmann-Toeplitz operators with symbols of a variable sign” In J. Anal. Math. 114, 2011, pp. 317–340 DOI: 10.1007/s11854-011-0019-6
  • [PS23] Alexander Pushnitski and Alexander Sobolev “Hankel operators with band spectra and elliptic functions”, 2023 eprint: arXiv:2307.09242
  • [Rai90] Georgi Raikov “Eigenvalue asymptotics for the Schrödinger operator with homogeneous magnetic potential and decreasing electric potential. I. Behaviour near the essential spectrum tips” In Comm. Partial Differential Equations 15.3, 1990, pp. 407–434 DOI: 10.1080/03605309908820690
  • [RS78] Michael Reed and Barry Simon “Methods of modern mathematical physics. IV. Analysis of operators” Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1978, pp. xv+396
  • [SY23] Mostafa Sabri and Pierre Youssef “Flat bands of periodic graphs” In J. Math. Phys. 64.9, 2023, pp. Paper No. 092101\bibrangessep23 DOI: 10.1063/5.0156336
  • [Sun13] Toshikazu Sunada “Topological crystallography” With a view towards discrete geometric analysis 6, Surveys and Tutorials in the Applied Mathematical Sciences Springer, Tokyo, 2013, pp. xii+229 DOI: 10.1007/978-4-431-54177-6
  • [Tha92] Bernd Thaller “The Dirac equation”, Texts and Monographs in Physics Springer-Verlag, Berlin, 1992, pp. xviii+357 DOI: 10.1007/978-3-662-02753-0
  • [Zwo24] Maciej Zworski “Mathematical results on the chiral models of twisted bilayer graphene” With an appendix by Mengxuan Yang and Zhongkai Tao In J. Spectr. Theory 14.3, 2024, pp. 1063–1107 DOI: 10.4171/jst/521