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

Double Dirac Cone in Band Structures of Periodic Schrödinger Operators

Ying Cao Yau Mathematical Sciences Center, Tsinghua Unversity, Beijing, 100084, China ([email protected]).    Yi Zhu Yau Mathematical Sciences Center, Tsinghua Unversity, Beijing, 100084, China, and Yanqi Lake Beijing Institute of Mathematical Sciences and Applications, Beijing, 101408, China([email protected]).
Abstract

Dirac cones are conical singularities that occur near the degenerate points in band structures. Such singularities result in enormous unusual phenomena of the corresponding physical systems. This work investigates double Dirac cones that occur in the vicinity of a fourfold degenerate point in the band structures of certain operators. It is known that such degeneracy originates in the symmetries of the Hamiltonian. We use two dimensional periodic Schrödinger operators with novel designed symmetries as our prototype. First, we characterize admissible potentials, termed as super honeycomb lattice potentials. They are honeycomb lattices potentials with a key additional translation symmetry. It is rigorously justified that Schrödinger operators with such potentials almost guarantee the existence of double Dirac cones on the bands at the Γ\Gamma point, the origin of the Brillouin zone. We further show that the additional translation symmetry is an indispensable ingredient by a perturbation analysis. Indeed, the double cones disappear if the additional translation symmetry is broken. Many numerical simulations are provided, which agree well with our analysis.

MSCcodes: 35Q40, 35Q60, 35P99

Keywords: periodic Schrödinger operator, super honeycomb lattice, fourfold degeneracy, double Dirac cone

1 Introduction

The Dirac cone is a conical structure near the degenerate point on the energy bands. It is deeply rooted in the symmetries of operators. Single Dirac cones near a twofold degenerate point have been found in many physical systems. It is a hallmark and reveals the underlying mechanism of versatile electronic or photonic properties of topological materials [7, 10, 15, 26, 19]. A typical system possessing the single Dirac cone is the two-dimensional material–graphene, which has the atomic honeycomb lattice made of carbon atoms [21, 20, 25]. Its great success in many fields has brought the blooming time for both experimental and theoretical understanding of such degenerate points on spectral bands. Meanwhile, other types of conically degenerate points were reported. Among those, double Dirac cones have attracted considerable attention [31, 34]. Such conical structures consist of two cones that share a fourfold degenerate apex. Due to the higher degeneracy, the corresponding wave patterns and physical properties are different from those in systems possessing single Dirac cones [30, 24, 33, 22]. It is known that the nontrivial topology of energy bands and corresponding significant properties of materials are born from the singularity. Thus, investigating the underlying symmetries and degeneracy related to the double Dirac cone will help us better understand the unusual physical properties.

Regarding single Dirac cones, a lot of analyses about the related time reversal symmetric operators have been done through different models and vehicles, especially when the material has a honeycomb structure. The tight-binding approximation was first developed by Wallace to describe the band structure of graphite [29], and later used systematically by others [27, 18]. The perturbation theory and multiscale analysis help to solve shallow potential cases successfully [1, 16]. One pioneering rigorous result on characterizing the honeycomb potentials and demonstrations of the existence of Dirac points was given by Fefferman and Weinstein [13]. They paved the way to rigorously analyzing such degenerate spectral points by combining Lyapunov-Schmidt reduction, perturbation theories, and multidimensional complex analysis. Based on their results, many other problems were solved such as the evolution of wave packets spectrally concentrated near Dirac points [14], edge states and valley Hall effect [8], lower dimensional degenerate points [12] and threefold Weyl points in the three-dimensional problems [17]. Ammari and collaborators did a lot of work on the Dirac cone and edge states using layer potential theory in the subwavelength regime [3, 4, 2]. Besides, the group representation theory has been used by Berkolaiko and Comech to describe the symmetric structure [6]. Despite the aforementioned progress in this blooming area, there is rarely any rigorous result on double Dirac cones as those in single Dirac cones.

In this paper, we investigate the two dimensional Schrödinger operator HV=Δ+V(𝐱)H_{V}=-\Delta+V({\bf x}) with V(𝐱)V({\bf x}) specially structured such that HVH_{V} has a double Dirac cone on its energy surfaces. Our goal is to find the precise mathematical description of this special kind of V(𝐱)V({\bf x}), and establish the rigorous proof about the existence of the double Dirac cone. We first define a class of potentials V(𝐱)V({\bf x}), termed as the super honeycomb lattice potentials. They are honeycomb lattice potentials equipped with an additional translation symmetry. Then we prove that such V(𝐱)V({\bf x}) is enough for the existence of a double Dirac cone at Γ\Gamma point, the origin of the Brillouin zone, as is stated in the main theorem Theorem 3.1. To achieve our goal, we utilize Lyapunov-Schmidt reduction, perturbation theory, and spectral theories about infinite dimensional linear operators. The rigorous analysis is inspired by pioneering works on single Dirac cones by Fefferman and Weinstein [13]. However, due to higher multiplicity and the additional symmetry, we need a more delicate decomposition of the working function spaces and the bifurcation matrix, see Sections 2.3 and 3.2. We also show that the extra translation symmetry is indispensable. Namely, a small perturbation that breaks this symmetry leads to the separation of the fourfold degeneracy and disappearance of the double cone in the band structure, see Section 4. Besides, we give two typical examples of potentials that are in the class of admissible potentials. Numerical simulations are provided to support our analysis. Our results will shine a light on the study of more complicated symmetries of operators and higher degeneracy on energy bands.

The rest of the paper is organized as follows. Section 2 provides the preliminaries. The definition of super honeycomb lattice potentials and the decomposition of the working function space are given based on symmetries. In Section 3, we state and prove the main theorem–the existence of the double Dirac cone at the Γ\Gamma point of the Schrödinger operator with a super honeycomb lattice. Inspired by [13], the proof is divided into two main parts. First, we show that the fourfold degeneracy at the Γ\Gamma point leads to a double cone in the vicinity under proper assumptions. Secondly, we justify the assumptions for shallow potentials and then extend the shallow potentials to generic potentials. In Section 4, we discuss the band structures under perturbations which break the additional translation symmetry. The double Dirac cone separates into two parts and a local energy gap appears near the Γ\Gamma point. At the end, corresponding numerical simulations for the two typical potentials are given in Section 5.

2 Super honeycomb lattice potential and symmetries

Symmetries of an operator are the origin of many novel properties of its spectrum. In this section, we introduce a large class of potentials, termed as super honeycomb lattice potentials, which are characterized by several symmetries. Their properties and corresponding spectral theory are discussed.

2.1 Super honeycomb lattice potentials

We first introduce the parity, complex-conjugation, and rotation operators for a function f(x)f(x) defined in 2{\mathbb{R}}^{2} as below:

𝒫[f](𝐱)=f(𝐱),𝒞[f](𝐱)=f(𝐱)¯,θ[f](𝐱)=f(Rθ𝐱),{\mathcal{P}}[f]({\bf x})=f(-{\bf x}),\quad{\mathcal{C}}[f]({\bf x})=\overline{f({\bf x})},\quad{\mathcal{R}}_{\theta}[f]({\bf x})=f(R_{\theta}^{*}{\bf x}),

where

Rθ=(cosθsinθsinθcosθ)R_{\theta}=\begin{pmatrix}\cos{\theta}&\sin{\theta}\\ -\sin{\theta}&\cos{\theta}\end{pmatrix} (2.1)

represents the clockwise rotation by an angle of θ[0,2π]\theta\in[0,2\pi] in 2{\mathbb{R}}^{2} and its Hermitian Rθ=Rθ1R_{\theta}^{*}=R_{\theta}^{-1} represents the anticlockwise rotation. A function f(x)f(x) is called 𝒫{\mathcal{P}}-invariant (or parity symmetric) if 𝒫[f](𝐱)=f(𝐱){\mathcal{P}}[f]({\bf x})=f({\bf x}), and similar for 𝒞{\mathcal{C}}- invariant (or conjugation symmetric) and θ{{\mathcal{R}}_{\theta}}-invariant (or rotation symmetric). Besides, being 𝒫𝒞{\mathcal{P}}{\mathcal{C}} invariant is also called having the time reversal symmetry. In this work, we are interested in R23πR_{\frac{2}{3}\pi}, so we omit the subscript, i.e., R=R23πR=R_{\frac{2}{3}\pi} and =23π{\mathcal{R}}={\mathcal{R}}_{\frac{2}{3}\pi} for simplicity.

We also use the following notation for the translation operator of a nonzero vector 𝐮{\bf u} in this article:

𝒯𝐮[f](𝐱)=f(𝐱+𝐮).{\mathcal{T}}_{{\bf u}}[f]({\bf x})=f({\bf x}+{\bf u}).

In this work we are interested in the spectra of HV=Δ+V(𝐱)H_{V}=-\Delta+V({\bf x}) with the potential V(𝐱)V({\bf x}) being equipped with the above symmetries. Namely, we have the following definitions.

Definition 2.1.

V(𝐱)L(2)V({\bf x})\in L^{\infty}({\mathbb{R}}^{2}) is called a honeycomb lattice potential, if

  1. 1.

    V(𝐱)V({\bf x}) is real and even,

  2. 2.

    V(𝐱)V({\bf x}) is {\mathcal{R}}-invariant,

  3. 3.

    V(𝐱)V({\bf x}) has a period 𝐮10{\bf u}_{1}\neq 0, and thus 𝐮2periodic{\bf u}_{2}-periodic with 𝐮2=R𝐮1{\bf u}_{2}=-R^{*}{\bf u}_{1}.

Remark 2.2.

All these three properties are discussed in L(2)L^{\infty}({\mathbb{R}}^{2}), that is to say, V(𝐱)=𝒞V(𝐱)=𝒫V(𝐱)=V(𝐱)=𝒯𝐮1V(𝐱)=𝒯𝐮2V(𝐱)V({\bf x})={\mathcal{C}}V({\bf x})={\mathcal{P}}V({\bf x})={\mathcal{R}}V({\bf x})={\mathcal{T}}_{{\bf u}_{1}}V({\bf x})={\mathcal{T}}_{{\bf u}_{2}}V({\bf x}) is valid almost everywhere. The following discussion about super honeycomb lattice potential is also in L(2)L^{\infty}({\mathbb{R}}^{2}) in the same way.

Therefore, the non-relativistic Schödinger operator HV(𝐱)=Δ+V(𝐱)H_{V}({\bf x})=-\Delta+V({\bf x}) has time reversal symmetry, rotation symmetry, and translation symmetry if V(𝐱)V({\bf x}) is a honeycomb lattice potential. The honeycomb lattice can refer to the blue lattice in Figure 1. As shown in this Figure, the black lattice is a honeycomb lattice, too. But it has an extra translation symmetry with periods 𝐯1{\bf v}_{1} and 𝐯2{\bf v}_{2}, where

𝐯1=13(2𝐮1𝐮2),𝐯2=13(𝐮1+𝐮2).{\bf v}_{1}=\frac{1}{3}(2{\bf u}_{1}-{\bf u}_{2}),{\qquad}{\bf v}_{2}=\frac{1}{3}({\bf u}_{1}+{\bf u}_{2}). (2.2)

We call it a super honeycomb lattice because of this additional symmetry.

Refer to caption
Refer to caption
Figure 1: (a) the figure of lattice, and (b) the figure of dual lattice. The blue lattices are the cases of honeycomb lattice, with periods 𝐮1{\bf u}_{1} and 𝐮2{\bf u}_{2} and dual periods 𝐤1{\bf k}_{1} and 𝐤2{\bf k}_{2}. The black lattices are the cases of corresponding super honeycomb lattice, with periods 𝐯1{\bf v}_{1} and 𝐯2{\bf v}_{2} and dual periods 𝐪1{\bf q}_{1} and 𝐪2{\bf q}_{2}. Obviously, the super honeycomb lattice is a kind of honeycomb lattice, but not vice versa.
Definition 2.3.

A honeycomb lattice potential V(𝐱)L(2)V({\bf x})\in L^{\infty}({\mathbb{R}}^{2}) is called a super honeycomb potential if

  1. 4.

    V(x) is 𝐯1{\bf v}_{1} and 𝐯2{\bf v}_{2} periodic, where 𝐯1{\bf v}_{1} and 𝐯2{\bf v}_{2} are as in (2.2) and the non-degeneracy condition holds:

    1|Ω|Ωei𝐪1𝐲V(𝐲)𝑑𝐲0,\frac{1}{|\Omega|}\int_{\Omega}e^{-i{\bf q}_{1}\cdot{\bf y}}V({\bf y})d{\bf y}\neq 0, (2.3)

    where Ω\Omega is a unit cell of honeycomb lattice as in (2.7).

By definitions, a super honeycomb lattice potential is a honeycomb lattice potential with an additional translation symmetry. In other words, it has a smaller lattice structure. Since in applications we need to break this symmetries to obtain bands with different topological indices [5, 24], see also Figure 2. We still consider the super honeycomb lattice potential in the bigger lattice structure.

We remark that the non-degeneracy condition (2.3) in the definition ensures the lowest Fourier coefficients of V(𝐱)V({\bf x}) do not vanish. As a consequence, the degeneracy occurs at 2nd7th2^{nd}-7^{th} bands. While higher bands should be considered if the non-degeneracy condition does not hold, which will be investigated in future works.

Here we give a typical example of super honeycomb lattice potentials which is a dimerization of a honeycomb lattice potential. It is a more general mathematical construction of the case studied in [31]. Our approach is first dimerizing in one direction, and then applying rotations to get the final results. Beginning from a honeycomb lattice potential, there should be three directions for dimerization: 𝐮1{\bf u}_{1}, 𝐮2{\bf u}_{2}, and 𝐮3=𝐮2𝐮1{\bf u}_{3}={\bf u}_{2}-{\bf u}_{1}. Thus, the following steps are needed to construct the dimer model.

Assume that f(𝐱)f({\bf x}) is a function such that f(𝐱+12𝐮3)f({\bf x}+\frac{1}{2}{\bf u}_{3}) is a honeycomb lattice potential, to be specific:

f(𝐱+𝐮1)=f(𝐱),f(𝐱+𝐮2)=f(𝐱),f(𝐱+12𝐮3)=f(𝐱+12𝐮3),x2.f({\bf x}+{\bf u}_{1})=f({\bf x}),{\quad}f({\bf x}+{\bf u}_{2})=f({\bf x}),{\quad}{\mathcal{R}}f({\bf x}+\frac{1}{2}{\bf u}_{3})=f({\bf x}+\frac{1}{2}{\bf u}_{3}),{\quad}\forall x\in{\mathbb{R}}^{2}. (2.4)

First dimerize in the 𝐮3{\bf u}_{3} direction:

g(𝐱,r)=f(𝐱12r𝐮3)+f(𝐱+12r𝐮3).g({\bf x},r)=f({\bf x}-\frac{1}{2}r{\bf u}_{3})+f({\bf x}+\frac{1}{2}r{\bf u}_{3}). (2.5)

Here r[0,1]r\in[0,1] is the distance ratio for dimers. Then rotates the obtained g(𝐱)g({\bf x}) :

W(𝐱,r)=g(𝐱,r)+g(𝐱,r)+2g(𝐱,r).W({\bf x},r)=g({\bf x},r)+{\mathcal{R}}g({\bf x},r)+{\mathcal{R}}^{2}g({\bf x},r). (2.6)

W(𝐱,r)W({\bf x},r) is the dimer model we want. Using (2.4)-(2.6), it is easy to check the conclusion below.

Proposition 2.4.

Any W(𝐱,13)W({\bf x},\frac{1}{3}) constructed by the above steps is a super honeycomb lattice potential.

Figure 2 shows a discrete example.

Refer to caption
Refer to caption
Refer to caption
Figure 2: Discrete honeycomb-dimer model examples with rr taking values (a) 1360\frac{13}{60}, (b) 13\frac{1}{3}, and (c) 12\frac{1}{2}. The blue hexagon is a unit cell of honeycomb lattice, and the black hexagon is a unit cell of the super honeycomb lattice. Blue points are elements before dimerization and rotation, red points are dimers. Elements in a dimer are connected by a red line.

Before all the detailed analysis of the energy surfaces of Schrödinger operators with super honeycomb lattice potentials, we try to first explain our method using symmetries. If V(𝐱)V({\bf x}) is a super honeycomb lattice potential, the operator HVH_{V} has following properties.

Lemma 2.5.

Assume that V(𝐱)V({\bf x}) is a super honeycomb lattice potential, then

[HV,𝒯𝐯1]=[HV,𝒯𝐯2]=[HV,]=0{[H_{V},{\mathcal{T}}_{{\bf v}_{1}}]=[H_{V},{\mathcal{T}}_{{\bf v}_{2}}]=[H_{V},{\mathcal{R}}]=0}
[HV,𝒫]=[HV,𝒞]=[HV,𝒫𝒞]=0[H_{V},{\mathcal{P}}]=[H_{V},{\mathcal{C}}]=[H_{V},{\mathcal{P}}{\mathcal{C}}]=0

Proof  Take [HV,𝒫]=0[H_{V},{\mathcal{P}}]=0 as an example and others are the same. For any ϕ(𝐱)\phi({\bf x}),

HV𝒫ϕ(𝐱)=(Δ+V(𝐱))ϕ(𝐱)=(Δx+V(𝐱))ϕ(𝐱)=𝒫HVϕ(𝐱),H_{V}{\mathcal{P}}\phi({\bf x})=(-\Delta+V({\bf x}))\phi(-{\bf x})=(-\Delta_{-x}+V(-{\bf x}))\phi(-{\bf x})={\mathcal{P}}H_{V}\phi({\bf x}),

because V(𝐱)V({\bf x}) is even, and the Laplace operator is rotation invariant. \Box

The spectrum of HVH_{V} on L2(2)L^{2}({\mathbb{R}}^{2}) is equivalent to such a union by Floquet-Bloch theorem:

σL2(2)(HV)=𝐤Ωσχ(HV(𝐤)),\sigma_{L^{2}({\mathbb{R}}^{2})}(H_{V})=\bigcup_{{\bf k}\in\Omega^{*}}\sigma_{\chi}(H_{V}({\bf k})),

where the notations are introduced in the following two subsections. χ\chi is as in (2.12). Thus, finding eigenvalues of HVH_{V} on L2(2)L^{2}({\mathbb{R}}^{2}) is transformed into finding eigenvalues of HV(𝐤)H_{V}({\bf k}) on χ\chi. It is clear that HV(𝟎)=HVH_{V}(\mathbf{0})=H_{V} - the operator corresponding to Γ\Gamma point are commutative with 𝒯𝐯1{\mathcal{T}}_{{\bf v}_{1}} and {\mathcal{R}}. 𝒯𝐯1{\mathcal{T}}_{{\bf v}_{1}} and {\mathcal{R}} are unitary and have three eigen-subspaces corresponding to different eigenvalues ξ1=1\xi_{1}=1, ξ2=e23πi\xi_{2}=e^{\frac{2}{3}\pi i}, and ξ3=e23πi\xi_{3}=e^{-\frac{2}{3}\pi i} on χ\chi. Suppose ϕl(𝐱)\phi_{l}({\bf x}) and ϕj(𝐱)\phi_{j}({\bf x}) are any normalized eigenfunctions of {\mathcal{R}} with eigenvalues ξl\xi_{l} and ξj\xi_{j}, then:

HVϕl(𝐱),ϕj(𝐱)\displaystyle\langle H_{V}\phi_{l}({\bf x}),\phi_{j}({\bf x})\rangle =HVϕl(𝐱),ϕj(𝐱)\displaystyle=\langle{\mathcal{R}}H_{V}\phi_{l}({\bf x}),{\mathcal{R}}\phi_{j}({\bf x})\rangle
=ξl¯ξjHVϕl(𝐱),ϕj(𝐱)=δl,jHVϕl(𝐱),ϕj(𝐱),\displaystyle=\overline{\xi_{l}}\xi_{j}\langle H_{V}\phi_{l}({\bf x}),\phi_{j}({\bf x})\rangle=\delta_{l,j}\langle H_{V}\phi_{l}({\bf x}),\phi_{j}({\bf x})\rangle,

where the inner product is as in (LABEL:eqn-innerPro). This tells that eigen-subspaces of {\mathcal{R}} are invaraint spaces of HVH_{V}, and similar for 𝒯𝐯1{\mathcal{T}}_{{\bf v}_{1}}. Thus, we can further decompose the spectrum of HVH_{V} on χ\chi to the spectrum of HVH_{V} on these subspaces. Finally, we associate these subspaces by 𝒫{\mathcal{P}}, 𝒞{\mathcal{C}}, and 𝒫𝒞{\mathcal{P}}{\mathcal{C}} and construct the degeneracy by the fact that they are commutative with HVH_{V}.

2.2 A quick review of Floquet-Bloch theory

In this subsection, we review the Floquet-Bloch theory briefly. Let {𝐮1,𝐮2}\{{\bf u}_{1},{\bf u}_{2}\} be linear independent vectors in 2{\mathbb{R}}^{2}. Its corresponding equilateral lattice is 𝐔=𝐮𝟏𝐮𝟐\bf{U}={\mathbb{Z}}\bf{u}_{1}\oplus{\mathbb{Z}}\bf{u}_{2}. Denote a unit cell

Ω={𝐮=c1𝐮1+c2𝐮2,c1,c2[0,1]}.\Omega=\{{\bf u}=c_{1}{\bf u}_{1}+c_{2}{\bf u}_{2},{\quad}c_{1},c_{2}\in[0,1]\}. (2.7)

Dual lattice of 𝐔\bf{U} is

𝐔=𝐤1𝐤2𝐮l𝐤j=2πδl,jl,j=1,2.{\bf{U}}^{*}={\mathbb{Z}}{{\bf k}}_{1}\oplus{\mathbb{Z}}{{\bf k}}_{2}{\quad}{{\bf u}}_{l}\cdot{\bf k}_{j}=2\pi\delta_{l,j}{\quad}l,j=1,2.

Ω=2/𝐔\Omega^{*}={\mathbb{R}}^{2}/\bf{U}^{*} is the Brillouin zone. We can divide the eigenvalue problem of HVH_{V} on L2(2)L^{2}({\mathbb{R}}^{2}) into the following eigenvalue problems of HVH_{V} traversing all the 𝐤{\bf k} in the Brillouin zone [9, 28].

For 𝐤Ω{\bf k}\in\Omega^{*}, consider the eigenvalue problem

HVΦ(𝐱,𝐤)=μ(𝐤)Φ(𝐱,𝐤),𝐱2,H_{V}\Phi({\bf x},{\bf k})=\mu({\bf k})\Phi({\bf x},{\bf k}),{\quad}{\bf x}\in{\mathbb{R}}^{2}, (2.8)
Φ(𝐱+𝐮,𝐤)=ei𝐤𝐮Φ(𝐱,𝐤),𝐮𝐔,𝐱𝟐\Phi({\bf x}+{\bf u},{\bf k})=e^{i{\bf k}\cdot{\bf u}}\Phi({\bf x},{\bf k}),{\quad}{\bf u}\in\bf{U},{\quad}{\bf x}\in{\mathbb{R}}^{2} (2.9)

Let Φ(𝐱,𝐤)=ei𝐤𝐱ϕ(𝐱,𝐤)\Phi({\bf x},{\bf k})=e^{i{\bf k}\cdot{\bf x}}\phi({\bf x},{\bf k}), where ϕ(𝐱)\phi({\bf x}) is periodic. Then (2.8) is equal to:

(Δ+V(𝐱))(ei𝐤𝐱ϕ(𝐱,𝐤))=μei𝐤𝐱ϕ(𝐱,𝐤)(-\Delta+V({\bf x}))(e^{i{\bf k}\cdot{\bf x}}\phi({\bf x},{\bf k}))=\mu e^{i{\bf k}\cdot{\bf x}}\phi({\bf x},{\bf k})

that is

((+i𝐤)(+i𝐤)+V(𝐱))ϕ(𝐱,𝐤)=μϕ(𝐱,𝐤).(-(\nabla+i{\bf k})\cdot(\nabla+i{\bf k})+V({\bf x}))\phi({\bf x},{\bf k})=\mu\phi({\bf x},{\bf k}){\quad}.

Let HV(𝐤)=(+i𝐤)(+i𝐤)+V(𝐱)H_{V}({{\bf k}})=-(\nabla+i{\bf k})\cdot(\nabla+i{\bf k})+V({\bf x}). Thus, the eigenvalue problem can be rewritten as:

HV(𝐤)ϕ(𝐱,𝐤)=μϕ(𝐱,𝐤),𝐱2H_{V}({{\bf k}})\phi({\bf x},{\bf k})=\mu\phi({\bf x},{\bf k}),{\quad}{\bf x}\in{\mathbb{R}}^{2} (2.10)
ϕ(𝐱+𝐮,𝐤)=ϕ(𝐱),𝐮𝐔\phi({\bf x}+{\bf u},{\bf k})=\phi({\bf x}),{\quad}{\bf u}\in\bf{U} (2.11)

(2.8)-(2.9), or equivalently (2.10)-(2.11), has a real, discrete and lower bounded spectrum [11]:

μ1(𝐤)μ2(𝐤)μ3(𝐤)\mu_{1}({\bf k})\leq\mu_{2}({\bf k})\leq\mu_{3}({\bf k})\leq...

These energy bands have the following property [13].

Lemma 2.6.

Any μb(𝐤)\mu_{b}({\bf k}) is periodic and Lipschitz continuous. Its periods are 𝐤1{\bf k}_{1} and 𝐤2{\bf k}_{2}.

2.3 Function spaces and symmetries at Γ\Gamma point

For honeycomb lattice potentials, let 𝐮2=R𝐮1{\bf u}_{2}=-R{\bf u}_{1}. We concern about the spectrum at Γ\Gamma point most, so define

χ={f(𝐱)Lloc2(2),f(𝐱+𝐮)=f(𝐱),𝐮𝐔}.\chi=\{f({\bf x})\in L^{2}_{loc}({\mathbb{R}}^{2}),{\quad}f({\bf x}+{\bf u})=f({\bf x}),{\quad}\forall{\bf u}\in\bf{U}\}. (2.12)

χ\chi is a Hilbert space under the inner product:

f(𝐱),g(𝐱)=1|Ω|Ωf(𝐱)¯g(𝐱)𝑑𝐱\langle f({\bf x}),g({\bf x})\rangle=\frac{1}{|\Omega|}\int_{\Omega}\overline{f({\bf x})}g({\bf x})d{\bf x} (2.13)

Our aim is to find the fourfold degeneracy of HVH_{V} on χ\chi and the double Dirac cone in the vicinity of this highly degenerate point Γ\Gamma. We also define the limitation of χ\chi in Hloc1(2)H^{1}_{loc}({\mathbb{R}}^{2}):

Hper1={f(𝐱)Hloc1(2),f(𝐱+𝐮)=f(𝐱),𝐮𝐔}.H^{1}_{per}=\{f({\bf x})\in H^{1}_{loc}({\mathbb{R}}^{2}),{\quad}f({\bf x}+{\bf u})=f({\bf x}),{\quad}\forall{\bf u}\in\bf{U}\}. (2.14)

A super honeycomb lattice potential V(x)V(x) is not only real and in χ\chi, but also in

χs={fLloc2(2),f(𝐱+𝐯)=f(𝐱),𝐱2,𝐯𝐯1𝐯2},\chi_{s}=\{f\in L^{2}_{loc}({\mathbb{R}}^{2}),{\quad}f({\bf x}+{\bf v})=f({\bf x}),{\quad}{\bf x}\in{\mathbb{R}}^{2},{\bf v}\in{\mathbb{Z}}{\bf v}_{1}\oplus{\mathbb{Z}}{\bf v}_{2}\}, (2.15)

where 𝐯1{\bf v}_{1} and 𝐯2{\bf v}_{2} are in the form of (2.2). It is easy to see that

𝐮1=𝐯1+𝐯2,𝐮2=𝐯1+2𝐯2,{\bf u}_{1}={\bf v}_{1}+{\bf v}_{2},{\qquad}{\bf u}_{2}=-{\bf v}_{1}+2{\bf v}_{2}, (2.16)

which means χsχ\chi_{s}\subset\chi, or functions in χs\chi_{s} have smaller periods than those in χ\chi, as we have mentioned before. Dual vectors for {𝐯1,𝐯2}\{{\bf v}_{1},{\bf v}_{2}\} such that 𝐯i𝐪j=δi,j{\bf v}_{i}{\bf q}_{j}=\delta_{i,j} are:

𝐪1=𝐤1𝐤2,𝐪2=𝐤1+2𝐤2.{\bf q}_{1}={\bf k}_{1}-{\bf k}_{2},{\qquad}{\bf q}_{2}={\bf k}_{1}+2{\bf k}_{2}. (2.17)

First, we claim a decomposition of χ\chi. This decomposition associates the extra translation symmetry with parity symmetry perfectly.

Proposition 2.7.

χ=χsχ𝐤1χ𝐤1\chi=\chi_{s}\oplus\chi_{{\bf k}_{1}}\oplus\chi_{-{\bf k}_{1}}, where

χ𝐤1={f(𝐱)=ei𝐤1𝐱p(𝐱),p(𝐱)χs},\chi_{{\bf k}_{1}}=\{f({\bf x})=e^{i{\bf k}_{1}\cdot{\bf x}}p({\bf x}),{\quad}p({\bf x})\in\chi_{s}\}, (2.18)
χ𝐤1={f(𝐱)=ei𝐤1𝐱p(𝐱),p(𝐱)χs}.\chi_{-{\bf k}_{1}}=\{f({\bf x})=e^{-i{\bf k}_{1}\cdot{\bf x}}p({\bf x}),{\quad}p({\bf x})\in\chi_{s}\}. (2.19)

And χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}} and χ𝐤1\chi_{-{\bf k}_{1}} are eigen-spaces of 𝒯𝐯1{\mathcal{T}}_{{\bf v}_{1}} with eigenvalues 11, e23πie^{-\frac{2}{3}\pi i} and e23πie^{\frac{2}{3}\pi i}.

Proof  It is obvious that χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}}, and χ𝐤1\chi_{-{\bf k}_{1}} are subsets of χ\chi. Also, it is easy to verify they are eigenspaces after some very simple calculations. Thus, only need to prove that they are orthogonal to each other and χχsχ𝐤1χ𝐤1\chi\subset\chi_{s}\oplus\chi_{{\bf k}_{1}}\oplus\chi_{-{\bf k}_{1}}.

According to Fourier analysis,

{ei(m1𝐤1+m2𝐤2)𝐱}(m1,m2)2\{e^{i(m_{1}{\bf k}_{1}+m_{2}{\bf k}_{2}){\cdot}{\bf x}}\}_{(m_{1},m_{2})\in{\mathbb{Z}}^{2}} (2.20)

forms a Hilbert basis of χ\chi. Similarly,

χs=span{ei(n1𝐪1+n2𝐪2)𝐱,(n1,n2)2}¯\chi_{s}=\overline{span\{e^{i(n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}},{\quad}(n_{1},n_{2})\in{\mathbb{Z}}^{2}\}} (2.21)

From (2.17), (2.21) is equivalent to

χs=span{ei((n1+n2)𝐤1+(2n2n1)𝐤2)𝐱,(n1,n2)2}¯\chi_{s}=\overline{span\{e^{i((n_{1}+n_{2}){\bf k}_{1}+(2n_{2}-n_{1}){\bf k}_{2}){\cdot}{\bf x}},{\quad}(n_{1},n_{2})\in{\mathbb{Z}}^{2}\}} (2.22)

And therefore

χ𝐤1=span{ei((n1+n2+1)𝐤1+(2n2n1)𝐤2)𝐱,(n1,n2)2}¯\chi_{{\bf k}_{1}}=\overline{span\{e^{i((n_{1}+n_{2}+1){\bf k}_{1}+(2n_{2}-n_{1}){\bf k}_{2}){\cdot}{\bf x}},{\quad}(n_{1},n_{2})\in{\mathbb{Z}}^{2}\}} (2.23)
χ𝐤1=span{ei((n1+n21)𝐤1+(2n2n1)𝐤2)𝐱,(n1,n2)2}¯\chi_{-{\bf k}_{1}}=\overline{span\{e^{i((n_{1}+n_{2}-1){\bf k}_{1}+(2n_{2}-n_{1}){\bf k}_{2}){\cdot}{\bf x}},{\quad}(n_{1},n_{2})\in{\mathbb{Z}}^{2}\}} (2.24)

The 2{\mathbb{Z}}^{2} solutions of equations

{n1+n2=m12n2n1=m2\left\{\begin{aligned} n_{1}+n_{2}=m_{1}\\ 2n_{2}-n_{1}=m_{2}\end{aligned}\right.

exists: (n1,n2)=(2m1m23,m1+m23)(n_{1},n_{2})=(\frac{2m_{1}-m_{2}}{3},\frac{m_{1}+m_{2}}{3}), which means ei(m1𝐤1+m2𝐤2)𝐱e^{i(m_{1}{\bf k}_{1}+m_{2}{\bf k}_{2}){\cdot}{\bf x}} is in χs\chi_{s}, if and only if (m1+m2)0(mod3)(m_{1}+m_{2})\equiv 0\pmod{3}. It is easy to check, similarly, ei(m1𝐤1+m2𝐤2)𝐱e^{i(m_{1}{\bf k}_{1}+m_{2}{\bf k}_{2}){\cdot}{\bf x}} is in χ𝐤1\chi_{{\bf k}_{1}} if and only if (m1+m2)1(mod3)(m_{1}+m_{2})\equiv 1\pmod{3} and ei(m1𝐤1+m2𝐤2)𝐱e^{i(m_{1}{\bf k}_{1}+m_{2}{\bf k}_{2}){\cdot}{\bf x}} is in χ𝐤1\chi_{-{\bf k}_{1}} if and only if (m1+m2)2(mod3)(m_{1}+m_{2})\equiv 2\pmod{3}. It follows that ei(m1𝐤1+m2𝐤2)𝐱e^{i(m_{1}{\bf k}_{1}+m_{2}{\bf k}_{2}){\cdot}{\bf x}} must be in one and only one of χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}} and χ𝐤1\chi_{-{\bf k}_{1}}. This completes the proof. \Box

Obviously, multiplications by elements in χs\chi_{s} will map χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}}, and χ𝐤1\chi_{-{\bf k}_{1}} into themselves. To some extent, this shows functions in χs\chi_{s}, or to be specific, our potential V(𝐱)V({\bf x}) possesses higher symmetry. Besides, the decomposition above has the following symmetry:

Proposition 2.8.

If f(𝐱)f({\bf x}) is in χ𝐤1\chi_{{\bf k}_{1}}, then f(𝐱)f(-{\bf x}) is in χ𝐤1\chi_{-{\bf k}_{1}}, and vice versa.

Thus, the transformation 𝒫[f](𝐱)=f(𝐱){\mathcal{P}}[f]({\bf x})=f(-{\bf x}) takes χ𝐤1\chi_{{\bf k}_{1}} and χ𝐤1\chi_{-{\bf k}_{1}} exactly to each other.

Secondly, we introduce rotation eigen-subspaces of χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}}, and χ𝐤1\chi_{-{\bf k}_{1}} according to Fourier analysis. It is easy to get

R𝐤1=𝐤2,R𝐤2=𝐤1𝐤2,R(𝐤1𝐤2)=𝐤1,R{\bf k}_{1}={\bf k}_{2},{\quad}R{\bf k}_{2}=-{\bf k}_{1}-{\bf k}_{2},{\quad}R(-{\bf k}_{1}-{\bf k}_{2})={\bf k}_{1}, (2.25)

and

R𝐪1=𝐪2,R𝐪2=𝐪1𝐪2,R(𝐪1𝐪2)=𝐪1.R{\bf q}_{1}={\bf q}_{2},{\quad}R{\bf q}_{2}=-{\bf q}_{1}-{\bf q}_{2},{\quad}R(-{\bf q}_{1}-{\bf q}_{2})={\bf q}_{1}. (2.26)

Note the fact that:

Lemma 2.9.

χs,χ𝐤1,χ𝐤1\chi_{s},\chi_{{\bf k}_{1}},\chi_{-{\bf k}_{1}} are invariant function spaces of {\mathcal{R}}.

Proof  From (2.26), clearly {\mathcal{R}} maps χs\chi_{s} to itself. Since

(ei𝐤1𝐱)=ei𝐤1R𝐱=eiR𝐤1𝐱=ei𝐤2𝐱=ei𝐤1𝐱ei𝐪1𝐱{\mathcal{R}}(e^{i{\bf k}_{1}{\cdot}{\bf x}})=e^{i{\bf k}_{1}{\cdot}R^{*}{\bf x}}=e^{iR{\bf k}_{1}{\cdot}{\bf x}}=e^{i{\bf k}_{2}{\cdot}{\bf x}}=e^{i{\bf k}_{1}{\cdot}{\bf x}}e^{-i{\bf q}_{1}{\cdot}{\bf x}}

is in χ𝐤1\chi_{{\bf k}_{1}}, {\mathcal{R}} maps χ𝐤1\chi_{{\bf k}_{1}} to itself, too. It is similar for χ𝐤1\chi_{-{\bf k}_{1}} with

(ei𝐤1𝐱)=ei𝐤2𝐱=ei𝐤1𝐱ei𝐪1𝐱.{\mathcal{R}}(e^{-i{\bf k}_{1}{\cdot}{\bf x}})=e^{-i{\bf k}_{2}{\cdot}{\bf x}}=e^{-i{\bf k}_{1}{\cdot}{\bf x}}e^{i{\bf q}_{1}{\cdot}{\bf x}}.

\Box

The next proposition gives detailed properties of these spaces with respect to the transformation {\mathcal{R}} by separating χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}}, and χ𝐤1\chi_{-{\bf k}_{1}} into eigenspaces of {\mathcal{R}}. This decomposition helps to associate the rotation symmetry with conjugation symmetry, as in Proposition 2.11.

Proposition 2.10.

The {\mathcal{R}}-invariant spaces χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}} and χ𝐤1\chi_{-{\bf k}_{1}} have the following decomposition:

χs=χs,1χs,τχs,τ¯\chi_{s}=\chi_{s,1}\oplus\chi_{s,\tau}\oplus\chi_{s,\overline{\tau}} (2.27)
χ𝐤1=χ𝐤1,1χ𝐤1,τχ𝐤1,τ¯\chi_{{\bf k}_{1}}=\chi_{{\bf k}_{1},1}\oplus\chi_{{\bf k}_{1},\tau}\oplus\chi_{{\bf k}_{1},\overline{\tau}} (2.28)
χ𝐤1=χ𝐤1,1χ𝐤1,τχ𝐤1,τ¯\chi_{-{\bf k}_{1}}=\chi_{-{\bf k}_{1},1}\oplus\chi_{-{\bf k}_{1},\tau}\oplus\chi_{-{\bf k}_{1},\overline{\tau}} (2.29)

where τ=e2πi/3\tau=e^{2\pi i/3} and χσ,ξ={fχσ,[f](𝐱)=ξf(𝐱)}\chi_{\sigma,\xi}=\{f\in\chi_{\sigma},{\mathcal{R}}[f]({\bf x})=\xi f({\bf x})\}, for σ=s,𝐤1,𝐤1,\sigma=s,{\bf k}_{1},-{\bf k}_{1}, and ξ=1,τ,τ¯\xi=1,\tau,\overline{\tau}.

Proof  Take (2.28) as an example, others are similar.

According to (2.23), let (n1,n2)(n_{1},n_{2}) denote

ei(𝐤1+n1𝐪1+n2𝐪2)𝐱=ei((n1+n2+1)𝐤1+(2n2n1)𝐤2)𝐱e^{i({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}=e^{i((n_{1}+n_{2}+1){\bf k}_{1}+(2n_{2}-n_{1}){\bf k}_{2}){\cdot}{\bf x}}

, then

(n1,n2)(n21,n1n2)(n2n11,n11)(n_{1},n_{2})\stackrel{{\scriptstyle{\mathcal{R}}}}{{\longrightarrow}}(-n_{2}-1,n_{1}-n_{2})\stackrel{{\scriptstyle{\mathcal{R}}}}{{\longrightarrow}}(n_{2}-n_{1}-1,-n_{1}-1) (2.30)

are all in χ𝐤1\chi_{{\bf k}_{1}}. Note that

(n1,n2)=(n1,n2)=(n21,n1n2),(n_{1},n_{2})={\mathcal{R}}(n_{1},n_{2})=(-n_{2}-1,n_{1}-n_{2}),
(n1,n2)=2(n1,n2)=(n2n11,n11)(n_{1},n_{2})={\mathcal{R}}^{2}(n_{1},n_{2})=(n_{2}-n_{1}-1,-n_{1}-1)

have no integer solutions. Thus, we can define the equivalence relation: (n1,n2)(n21,n1n2)(n2n11,n11)(n_{1},n_{2})\sim(-n_{2}-1,n_{1}-n_{2})\sim(n_{2}-n_{1}-1,-n_{1}-1) and equivalence class S+=2/S_{+}={\mathbb{Z}}^{2}/\sim.

Now it is clear that:

χ𝐤1,1={f(𝐱)χ:f(𝐱)=(n1,n2)S+c(n1,n2)(ei(𝐤1+n1𝐪1+n2𝐪2)𝐱+\displaystyle{\chi_{{\bf k}_{1},1}=\{f({\bf x})\in\chi:f({\bf x})=\sum_{(n_{1},n_{2})\in S_{+}}c(n_{1},n_{2})(e^{i({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}+} (2.31)
eiR(𝐤1+n1𝐪1+n2𝐪2)𝐱+eiR2(𝐤1+n1𝐪1+n2𝐪2)𝐱),c(n1,n2)}\displaystyle{e^{iR({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}+e^{iR^{2}({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}),{\quad}c(n_{1},n_{2})\in{\mathbb{C}}\}}
χ𝐤1,τ={f(𝐱)χ:f(𝐱)=(n1,n2)S+c(n1,n2)(ei(𝐤1+n1𝐪1+n2𝐪2)𝐱+\displaystyle\chi_{{\bf k}_{1},\tau}=\{f({\bf x})\in\chi:f({\bf x})=\sum_{(n_{1},n_{2})\in S_{+}}c(n_{1},n_{2})(e^{i({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}+ (2.32)
τ¯eiR(𝐤1+n1𝐪1+n2𝐪2)𝐱+τeiR2(𝐤1+n1𝐪1+n2𝐪2)𝐱),c(n1,n2)}\displaystyle\overline{\tau}e^{iR({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}+\tau e^{iR^{2}({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}),{\quad}c(n_{1},n_{2})\in{\mathbb{C}}\}
χ𝐤1,τ¯={f(𝐱)χ:f(𝐱)=(n1,n2)S+c(n1,n2)(ei(𝐤1+n1𝐪1+n2𝐪2)𝐱+\displaystyle\chi_{{\bf k}_{1},\overline{\tau}}=\{f({\bf x})\in\chi:f({\bf x})=\sum_{(n_{1},n_{2})\in S_{+}}c(n_{1},n_{2})(e^{i({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}+ (2.33)
τeiR(𝐤1+n1𝐪1+n2𝐪2)𝐱+τ¯eiR2(𝐤1+n1𝐪1+n2𝐪2)𝐱),c(n1,n2)}\displaystyle\tau e^{iR({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}+\overline{\tau}e^{iR^{2}({\bf k}_{1}+n_{1}{\bf q}_{1}+n_{2}{\bf q}_{2}){\cdot}{\bf x}}),{\quad}c(n_{1},n_{2})\in{\mathbb{C}}\}

\Box

Again, given below is some information about symmetry properties between these subspaces.

Proposition 2.11.

If f(𝐱)f({\bf x}) is in χ±𝐤1,τ\chi_{\pm{\bf k}_{1},\tau}, then 𝒫𝒞[f](𝐱)=f(𝐱)¯{\mathcal{P}}{\mathcal{C}}[f]({\bf x})=\overline{f(-{\bf x})} is in χ±𝐤1,τ¯\chi_{\pm{\bf k}_{1},\overline{\tau}}, and vice versa.

3 Double Dirac cone in the band structure

In this section, we shall state the main theorem of the fourfold degeneracy and doubly conical structures at the Γ\Gamma point for the two-dimensional non-relativistic Schödinger operator HVH_{V} with super honeycomb lattice potentials. And a rigorous proof follows. Our proof is mainly inspired by the pioneering works [13, 12, 17, 23]. The proof is divided into two parts. First, we show that the fourfold degeneracy and a particular non-degenerate condition about eigenfunction are sufficient for the existence of double Dirac cones. Due to the higher degeneracy, we need deal with a more complicated bifurcation matrix. However, taking advantage of the higher symmetries and corresponding novel decomposition of working space in last section, we find that the bifurcation matrix can be pretty concise. Then we establish the existence of the degenerate point and the prescribed condition to complete the proof. Specifically, we first justify the prescribed conditions for shallow potentials, and then extend the justification for generic potentials.

3.1 Main theorem of double Dirac cone

The main theorem of our paper is as follows.

Theorem 3.1.

Let V(𝐱)V({\bf x}) be a super honeycomb lattice potential in the sense of Definition 2.3. H(ϵ)=Δ+ϵV(𝐱)H^{(\epsilon)}=-\Delta+\epsilon V({\bf x}) is the corresponding Schrödinger operator. Then the following is true for all ϵA\epsilon\in{\mathbb{R}}\setminus A, where A is a discrete subset of {\mathbb{R}}:

  1. 1.

    there exists a real number μD\mu_{{}_{D}} such that μD\mu_{D} is an eigenvalue of H(ϵ)H^{(\epsilon)} on χ\chi of multiplicity four. Namely, there exists bb\in{\mathbb{N}} such that

    μb(𝟎)<μD=μb+1(𝟎)=μb+2(𝟎)=μb+3(𝟎)=μb+4(𝟎)<μb+5(𝟎),\mu_{b}(\bm{0})<\mu_{{}_{D}}=\mu_{b+1}(\bm{0})=\mu_{b+2}(\bm{0})=\mu_{b+3}(\bm{0})=\mu_{b+4}(\bm{0})<\mu_{b+5}(\bm{0}), (3.1)
  2. 2.

    there exists a number vF>0v_{F}>0, such that for 𝐤{\bf k} sufficiently small, the four spectral bands are of the form:

    μb+1(𝐤)=μDvF|𝐤|(1+η1(𝐤))\displaystyle\mu_{b+1}({\bf k})=\mu_{{}_{D}}-v_{F}|{\bf k}|(1+\eta_{1}({\bf k})) (3.2)
    μb+2(𝐤)=μDvF|𝐤|(1+η2(𝐤))\displaystyle\mu_{b+2}({\bf k})=\mu_{{}_{D}}-v_{F}|{\bf k}|(1+\eta_{2}({\bf k}))
    μb+3(𝐤)=μD+vF|𝐤|(1+η3(𝐤))\displaystyle\mu_{b+3}({\bf k})=\mu_{{}_{D}}+v_{F}|{\bf k}|(1+\eta_{3}({\bf k}))
    μb+4(𝐤)=μD+vF|𝐤|(1+η4(𝐤))\displaystyle\mu_{b+4}({\bf k})=\mu_{{}_{D}}+v_{F}|{\bf k}|(1+\eta_{4}({\bf k}))

    where ηj(𝐤)=O(|𝐤|)\eta_{j}({\bf k})=O({|}{\bf k}{|}), j=1,2,3,4j=1,2,3,4.

(3.2) is the strict description of the double Dirac cone mathematically. Thus, this theorem tells that there always exist two tangent cones with the same apex for the non-degenerate super honeycomb lattice potentials.

The rest of this section is proof of this theorem. First, we give the conclusion that the fourfold degeneracy under some conditions always yields the double Dirac cone for super honeycomb lattice potentials. Thus, our attention should be paid mainly to the existence of this condition and fourfold degeneracy.

3.2 Fourfold degeneracy and double Dirac cone

Due to (2.5), if ϕ(𝐱)ker(HVμI)\phi({\bf x})\in ker(H_{V}-\mu I), then 𝒫ϕ(𝐱),𝒞ϕ(𝐱),𝒫𝒞ϕ(𝐱)ker(HVμI){\mathcal{P}}\phi({\bf x}),{\mathcal{C}}\phi({\bf x}),{\mathcal{P}}{\mathcal{C}}\phi({\bf x})\in ker(H_{V}-\mu I). That is to say, if HVH_{V} does have an eigenvalue on χ𝐤1,τ\chi_{{\bf k}_{1},\tau} of multiplicity one, then the fourfold degeneracy of HVH_{V} on χ\chi can be realized by symmetries. In this subsection, we want to verify that energy bands intersect conically in a small vicinity of this kind of eigenvalues under some necessary assumptions.

Proposition 3.2.

Let V(𝐱)V({\bf x}) be a super honeycomb lattice potential. Assume that:

  1. 1.

    there exists a real number μD\mu_{{}_{D}} and bb\in{\mathbb{N}} such that (3.1) holds for energy bands of HVH_{V},

  2. 2.

    μD\mu_{{}_{D}} is a simple eigenvalue on χ𝐤1,τ\chi_{{\bf k}_{1},\tau} of HVH_{V} with eigenfunction ϕ1(𝐱)\phi_{1}({\bf x}) and

    <ϕ1(𝐱),ϕ1(𝐱)>=1,<\phi_{1}({\bf x}),\phi_{1}({\bf x})>=1,
  3. 3.

    𝒗=<ϕ1(𝐱),𝒫𝒞[ϕ1](𝐱)>𝟎\bm{v}_{\sharp}=<\phi_{1}({\bf x}),\nabla{\mathcal{P}}{\mathcal{C}}[\phi_{1}]({\bf x})>\neq\bf{0}.

Then (3.2) holds for energy bands of HVH_{V}.

Remark 3.3.

The assumption that μD\mu_{{}_{D}} is of multiplicity four is necessary, because for higher degeneracy, obviously more branches are included, and thus the bands near the Γ\Gamma point will be more complex. Hence, we need to verify this condition in the later subsections.

Proof  Using Proposition 2.8, 2.11 and 2.5, we easily deduce that μD\mu_{{}_{D}} is also a simple eigenvalue on χ𝐤1,τ¯\chi_{{\bf k}_{1},\overline{\tau}}, χ𝐤1,τ\chi_{-{\bf k}_{1},{\tau}}, χ𝐤1,τ¯\chi_{-{\bf k}_{1},\overline{\tau}} with eigenfunctions ϕ2(𝐱)\phi_{2}({\bf x}), ϕ3(𝐱)\phi_{3}({\bf x}), and ϕ4(𝐱)\phi_{4}({\bf x}) :

ϕ2(𝐱)=𝒫𝒞ϕ1(𝐱),ϕ3(𝐱)=𝒫ϕ1(𝐱),ϕ4(𝐱)=𝒞ϕ1(𝐱),\phi_{2}({\bf x})={\mathcal{P}}{\mathcal{C}}\phi_{1}({\bf x}),{\quad}\phi_{3}({\bf x})={\mathcal{P}}\phi_{1}({\bf x}),{\quad}\phi_{4}({\bf x})={\mathcal{C}}\phi_{1}({\bf x}), (3.3)

Based on this, let us observe how the dispersion surfaces develop. First rewrite down the eigenvalue problem near 𝐤=𝟎{\bf k}=\bf{0}. Assume 𝐤{\bf k} is sufficiently small, the 𝐤{\bf k} quasi-momentum eigenproblem (2.10)-(2.11) is:

HV(𝐤)ϕ(𝐱)=μ(𝐤)ϕ(𝐱),xΩH_{V}({\bf k})\phi({\bf x})=\mu({\bf k})\phi({\bf x}),{\quad}x\in\Omega (3.4)
ϕ(𝐱+𝐮)=ϕ(𝐱),𝐮𝐔.\phi({\bf x}+{\bf u})=\phi({\bf x}),{\quad}{\bf u}\in\bf{U}. (3.5)

Now let μ(𝐤)=μD+λ\mu({\bf k})=\mu_{{}_{D}}+\lambda and

ϕ(𝐱)=αjϕj(𝐱)+ψ(𝐱)\phi({\bf x})=\alpha^{j}\phi_{j}({\bf x})+\psi({\bf x}) (3.6)

with ϕ(𝐱)χ\phi({\bf x})\in\chi and ψ(𝐱)ker(HVμDI)\psi({\bf x})\perp ker(H_{V}-\mu_{{}_{D}}I). We use the same superscript and subscript to represent summation over this script throughout the article. Since ker(HVμDI)ker(H_{V}-\mu_{{}_{D}}I) is a closed subspace of χ\chi, there exists projection operator 𝒬{\mathcal{Q}}_{\parallel} from χ\chi to ker(HVμDI)ker(H_{V}-\mu_{{}_{D}}I) and 𝒬{\mathcal{Q}}_{\perp} from χ\chi to ker(HVμDI)ker(H_{V}-\mu_{{}_{D}}I)^{\perp}. It is trivial that (HVμDI)ψ(𝐱)ker(HVμDI)(H_{V}-\mu_{{}_{D}}I)\psi({\bf x})\in ker(H_{V}-\mu_{{}_{D}}I)^{\perp}.

Substitute (3.6) into (3.4):

(HVμDI)ψ(𝐱)=(2i𝐤|𝐤|2+λ)(αjϕj(𝐱)+ψ(𝐱)).(H_{V}-\mu_{{}_{D}}I)\psi({\bf x})=(2i{\bf k}\cdot\nabla-|{\bf k}|^{2}+\lambda)(\alpha^{j}\phi_{j}({\bf x})+\psi({\bf x})). (3.7)

Project (3.7) by 𝒬{\mathcal{Q}}_{\parallel},

0=𝒬(2i𝐤|𝐤|2+λ)αjϕj(𝐱)+𝒬2i𝐤ψ(𝐱);0={\mathcal{Q}}_{\parallel}(2i{\bf k}\cdot\nabla-|{\bf k}|^{2}+\lambda)\alpha^{j}\phi_{j}({\bf x})+{\mathcal{Q}}_{\parallel}2i{\bf k}\cdot\nabla\psi({\bf x}); (3.8)

by 𝒬{\mathcal{Q}}_{\perp},

(HVμDI)ψ(𝐱)=𝒬2i𝐤αjϕj(𝐱)+𝒬(2i𝐤|𝐤|2+λ)ψ(𝐱).(H_{V}-\mu_{{}_{D}}I)\psi({\bf x})={\mathcal{Q}}_{\perp}2i{\bf k}\cdot\nabla\alpha^{j}\phi_{j}({\bf x})+{\mathcal{Q}}_{\perp}(2i{\bf k}\cdot\nabla-|{\bf k}|^{2}+\lambda)\psi({\bf x}). (3.9)

Eigenvalue problem (3.4)-(3.5) is equivalent to (3.8)-(3.9). First, solve ψ(𝐱)\psi({\bf x}) from (3.9, and then go back to (3.8) to obtain λ\lambda by linear approximation to complete the proof. This is exactly a Lyapunov-Schmidt reduction strategy.

Consider (3.9). Note that the resolvent operator (HVμDI)1(H_{V}-\mu_{{}_{D}}I)^{-1} is a bounded map from ker(HVμDI)ker(H_{V}-\mu_{{}_{D}}I)^{\perp} to 𝒬(Hper1){\mathcal{Q}}_{\perp}(H^{1}_{per}). Accordingly,

(I(HVμDI)1𝒬(2i𝐤|𝐤|2+λ))ψ(𝐱)=(HVμDI)1𝒬2ikαjϕj(𝐱)\big{(}I-(H_{V}-\mu_{{}_{D}}I)^{-1}{\mathcal{Q}}_{\perp}(2i{\bf k}\cdot\nabla-|{\bf k}|^{2}+\lambda)\big{)}\psi({\bf x})=(H_{V}-\mu_{{}_{D}}I)^{-1}{\mathcal{Q}}_{\perp}2ik\cdot\nabla\alpha^{j}\phi_{j}({\bf x})

Let 𝒜=(HVμDI)1𝒬(2i𝐤|𝐤|2+λ){\mathcal{A}}=(H_{V}-\mu_{{}_{D}}I)^{-1}{\mathcal{Q}}_{\perp}(2i{\bf k}\cdot\nabla-|{\bf k}|^{2}+\lambda). With |𝐤|+|λ||{\bf k}|+|\lambda| sufficiently small, operator norm of 𝒜{\mathcal{A}} should be less than 1, which means I𝒜I-{\mathcal{A}} is invertible, and (I𝒜)1(I-{\mathcal{A}})^{-1} is bounded and we have

ψ(𝐱)=(I𝒜)1(HVμDI)1𝒬2i𝐤αjϕj(𝐱).\psi({\bf x})=(I-{\mathcal{A}})^{-1}(H_{V}-\mu_{{}_{D}}I)^{-1}{\mathcal{Q}}_{\perp}2i{\bf k}\cdot\nabla\alpha^{j}\phi_{j}({\bf x}). (3.10)

Let 𝒯j[𝐤,λ](𝐱)=(I𝒜)1(HVμDI)1𝒬2i𝐤ϕj(𝐱){\mathcal{T}}_{j}[{\bf k},\lambda]({\bf x})=(I-{\mathcal{A}})^{-1}(H_{V}-\mu_{{}_{D}}I)^{-1}{\mathcal{Q}}_{\perp}2i{\bf k}\cdot\nabla\phi_{j}({\bf x}). It is a bounded map from a sufficient small neighborhood of (𝐤,λ)=(𝟎,𝟎)({\bf k},\lambda)=(\bf{0},0) in 2×{\mathbb{R}}^{2}\times{\mathbb{C}} to Hper1H^{1}_{per}. Its norm is less than C(|𝐤|+|λ|)C(|{\bf k}|+|\lambda|) in the small neighborhood. Thus, 𝒯j[𝐤,λ](𝐱))L2C(|𝐤|+|λ|)\parallel\nabla{\mathcal{T}}_{j}[{\bf k},\lambda]({\bf x}))\parallel_{L^{2}}\leq C(|{\bf k}|+|\lambda|). Rewrite (3.10) as

ψ(𝐱)=αj𝒯j[𝐤,λ](𝐱)\psi({\bf x})=\alpha^{j}{\mathcal{T}}_{j}[{\bf k},\lambda]({\bf x}) (3.11)

Substituting into (3.8),

𝒬((2i𝐤|𝐤|2+λ)ϕj(𝐱)+2i𝐤𝒯j[𝐤,λ](𝐱))αj=0{\mathcal{Q}}_{\parallel}\bigg{(}(2i{\bf k}\cdot\nabla-|{\bf k}|^{2}+\lambda)\phi_{j}({\bf x})+2i{\bf k}\cdot\nabla{\mathcal{T}}_{j}[{\bf k},\lambda]({\bf x})\bigg{)}\alpha^{j}=0 (3.12)

This equation has a nonzero solution {αj}\{\alpha^{j}\} when the matrix

M(𝐤,λ)=(<ϕl(𝐱),(2i𝐤|𝐤|2+λ)ϕj(𝐱)+2i𝐤𝒯j[𝐤,λ](𝐱)>)l,jM({\bf k},\lambda)=(<\phi_{l}({\bf x}),(2i{\bf k}\cdot\nabla-|{\bf k}|^{2}+\lambda)\phi_{j}({\bf x})+2i{\bf k}\cdot\nabla{\mathcal{T}}_{j}[{\bf k},\lambda]({\bf x})>)_{l,j} (3.13)

has a nonzero solution, which is equivalent to

det(M(𝐤,λ))=0det(M({\bf k},\lambda))=0 (3.14)

Our aim is to solve λ\lambda from this equation. Divide M(𝐤,λ)M({\bf k},\lambda) into two parts M0M_{0} and M1M_{1}:

M0(𝐤,λ)=(<ϕl(𝐱),(λ+2i𝐤)ϕj(𝐱)>)l,jM_{0}({\bf k},\lambda)=(<\phi_{l}({\bf x}),(\lambda+2i{\bf k}\cdot\nabla)\phi_{j}({\bf x})>)_{l,j} (3.15)
M1(𝐤,λ)=(|𝐤|2<ϕl(𝐱),ϕj(𝐱)>)l,j+(<ϕl(𝐱),2i𝐤𝒯j[𝐤,λ](𝐱)>)l,j.M_{1}({\bf k},\lambda)=(-|{\bf k}|^{2}<\phi_{l}({\bf x}),\phi_{j}({\bf x})>)_{l,j}+(<\phi_{l}({\bf x}),2i{\bf k}\cdot\nabla{\mathcal{T}}_{j}[{\bf k},\lambda]({\bf x})>)_{l,j}. (3.16)

Fix 𝐤{\bf k}, let (𝐤,λ(𝐤))({\bf k},\lambda({\bf k})) be the solution of (3.14). Taking advantage of (2.6), λ(𝐤)\lambda({\bf k}) should be Lipschitz continuous with 𝐤{\bf k} on each branch of the solutions. Thus, M1M_{1} is of order O(|𝐤|2)O(|{\bf k}|^{2}). That is to say, M1M_{1} will only contribute a term of order (|𝐤|8)(|{\bf k}|^{8}) to det(M(𝐤,λ))det(M({\bf k},\lambda)) when det(M(𝐤,λ))=0det(M({\bf k},\lambda))=0. We first solve the truncated equation of (3.14):

det(M0(𝐤,λ))=0det(M_{0}({\bf k},\lambda))=0 (3.17)

The M0M_{0} here, is the linear truncation of the original question, and is called a bifurcation matrix.

We already have ϕ1(𝐱),ϕ1(𝐱)=1\langle\phi_{1}({\bf x}),\phi_{1}({\bf x})\rangle=1. Because that the subspaces are orthogonal, ϕl(𝐱),ϕj(𝐱)=0\langle\phi_{l}({\bf x}),\phi_{j}({\bf x})\rangle=0 for ljl\neq j. And it is also obvious that ϕl(𝐱),ϕl(𝐱)=1\langle\phi_{l}({\bf x}),\phi_{l}({\bf x})\rangle=1 for all ll due to (3.3) and some simple calculations. Note that

<ϕ1(𝐱),2i𝐤ϕ2(𝐱)>\displaystyle<\phi_{1}({\bf x}),2i{\bf k}\cdot\nabla\phi_{2}({\bf x})> =2i𝐤𝒗\displaystyle=2i{\bf k}\cdot\bm{v}_{\sharp} (3.18)
<ϕ2(𝐱),2i𝐤ϕ1(𝐱)>\displaystyle<\phi_{2}({\bf x}),2i{\bf k}\cdot\nabla\phi_{1}({\bf x})> =<ϕ1(𝐱),2i𝐤ϕ2(𝐱)>¯=2i𝐤𝒗¯,\displaystyle=\overline{<\phi_{1}({\bf x}),2i{\bf k}\cdot\nabla\phi_{2}({\bf x})>}=\overline{2i{\bf k}\cdot\bm{v}_{\sharp}},
<ϕ3(𝐱),2i𝐤ϕ4(𝐱)>\displaystyle<\phi_{3}({\bf x}),2i{\bf k}\cdot\nabla\phi_{4}({\bf x})> =<ϕ1(𝐱),2i𝐤ϕ2(𝐱)>=2i𝐤𝒗,\displaystyle=<\phi_{1}({\bf x}),-2i{\bf k}\cdot\nabla\phi_{2}({\bf x})>=-2i{\bf k}\cdot\bm{v}_{\sharp},
<ϕ4(𝐱),2i𝐤ϕ3(𝐱)>\displaystyle<\phi_{4}({\bf x}),2i{\bf k}\cdot\nabla\phi_{3}({\bf x})> =<ϕ2(𝐱),2i𝐤ϕ1(𝐱)>=2i𝐤𝒗¯.\displaystyle=<\phi_{2}({\bf x}),-2i{\bf k}\cdot\nabla\phi_{1}({\bf x})>=-\overline{2i{\bf k}\cdot\bm{v}_{\sharp}}.

Besides, we also have the following results.

Proposition 3.4.

For ψj(𝐱)\psi_{j}({\bf x}) satisfies ψj(𝐱)=ξjψj(𝐱){\mathcal{R}}\psi_{j}({\bf x})=\xi_{j}\psi_{j}({\bf x}), ξj{1,τ,τ¯}\xi_{j}\in\{1,\tau,\overline{\tau}\}, j=1,2j=1,2

ξ1=ξ2<ψ1(𝐱),ψ2(𝐱)>=𝟎\xi_{1}=\xi_{2}{\quad}\Rightarrow{\quad}<\psi_{1}({\bf x}),\nabla\psi_{2}({\bf x})>=\bf{0}

Proof  Since the transformation {\mathcal{R}} is unit, if ξ1=ξ2\xi_{1}=\xi_{2}, we can attain that

<ψ1(𝐱),ψ2(𝐱)>\displaystyle<\psi_{1}({\bf x}),\nabla\psi_{2}({\bf x})> =<ψ1(𝐱),ψ2(𝐱)>=ξ1¯<ψ1(𝐱),R𝐱ψ2(R𝐱)>\displaystyle=<{\mathcal{R}}\psi_{1}({\bf x}),{\mathcal{R}}\nabla\psi_{2}({\bf x})>=\overline{\xi_{1}}<\psi_{1}({\bf x}),\nabla_{R^{*}{\bf x}}\psi_{2}(R^{*}{\bf x})>
=ξ1¯<ψ1(𝐱),Rψ2(𝐱)>=ξ1¯ξ2<ψ1(𝐱),Rψ2(𝐱)>\displaystyle=\overline{\xi_{1}}<\psi_{1}({\bf x}),R^{*}\nabla{\mathcal{R}}\psi_{2}({\bf x})>=\overline{\xi_{1}}\xi_{2}<\psi_{1}({\bf x}),R^{*}\nabla\psi_{2}({\bf x})>
=ξ1¯ξ2R<ψ1(𝐱),ψ2(𝐱)>=R<ψ1(𝐱),ψ2(𝐱)>\displaystyle=\overline{\xi_{1}}\xi_{2}R^{*}<\psi_{1}({\bf x}),\nabla\psi_{2}({\bf x})>=R^{*}<\psi_{1}({\bf x}),\nabla\psi_{2}({\bf x})>

Here RR^{*} is the rotation matrix. Because 11 is not an eigenvalue of RR^{*},

<ψ1(𝐱),ψ2(𝐱)>=𝟎.<\psi_{1}({\bf x}),\nabla\psi_{2}({\bf x})>=\bf{0}.

Besides, this equation tells us that <ψ1(𝐱),ψ2(𝐱)><\psi_{1}({\bf x}),\nabla\psi_{2}({\bf x})> must be 𝟎\bf{0} or an eigenvector of RR^{*}. \Box

From the proof of this lemma, we know 𝒗\bm{v}_{\sharp} is an eigenvector of RR^{*} with eigenvalue τ\tau. Thus it can be written as :

𝒗=vF2(1i)eiθ.\bm{v}_{\sharp}=\frac{v_{F}}{2}\begin{pmatrix}1\\ i\end{pmatrix}e^{i\theta}. (3.19)

We can choose an appropriate θ\theta such that vF>0v_{F}>0 because 𝒗𝟎\bm{v}_{\sharp}\neq\bf{0}.

Proposition 3.5.

For ϕ(𝐱)χ±𝐤1\phi({\bf x})\in\chi_{\pm{\bf k}_{1}}, each component of ϕ(𝐱)\nabla\phi({\bf x}) is in χ±𝐤1\chi_{\pm{\bf k}_{1}} too.

Proof  If ϕ(𝐱)χ±𝐤1\phi_{(}{\bf x})\in\chi_{\pm{\bf k}_{1}}, it can be written as ϕ(𝐱)=e±ik1𝐱p(𝐱)\phi_{(}{\bf x})=e^{\pm ik_{1}\cdot{\bf x}}p({\bf x}), where p(𝐱)χsp({\bf x})\in\chi_{s}. Therefore each component of ϕ(𝐱)=±i𝐤1e±ik1𝐱p(𝐱)+e±ik1𝐱p(𝐱)\nabla\phi({\bf x})=\pm i{\bf k}_{1}e^{\pm ik_{1}\cdot{\bf x}}p({\bf x})+e^{\pm ik_{1}\cdot{\bf x}}\nabla p({\bf x}) is in χ±𝐤1\chi_{\pm{\bf k}_{1}} too. \Box

Now using (3.4), (3.5), and again the orthogonality of subspaces, we can obtain the bifurcation matrix

M0(𝐤,λ)=(λ2i𝐤𝒗002i𝐤𝒗¯λ0000λ2i𝐤𝒗002i𝐤𝒗¯λ).M_{0}({\bf k},\lambda)=\begin{pmatrix}\lambda&2i{\bf k}\cdot\bm{v}_{\sharp}&0&0\\ \overline{2i{\bf k}\cdot\bm{v}_{\sharp}}&\lambda&0&0\\ 0&0&\lambda&-2i{\bf k}\cdot\bm{v}_{\sharp}\\ 0&0&-\overline{2i{\bf k}\cdot\bm{v}_{\sharp}}&\lambda\\ \end{pmatrix}. (3.20)

And det(M0(𝐤,λ))=(λ24|𝐤𝒗|2)2det(M_{0}({\bf k},\lambda))=(\lambda^{2}-4|{\bf k}\cdot\bm{v}_{\sharp}|^{2})^{2}. Now solve the whole equation (3.14). Thanks to all the discussion above, it can be written in the form that:

(λ24|𝐤𝒗|2)2+O(|𝐤|8)=0(\lambda^{2}-4|{\bf k}\cdot\bm{v}_{\sharp}|^{2})^{2}+O(|{\bf k}|^{8})=0 (3.21)

This equation’s solution (𝐤,λ(𝐤))({\bf k},\lambda({\bf k})) gives four branches of the dispersion surfaces by μ(𝐤)=μ+λ(𝐤)\mu({\bf k})=\mu+\lambda({\bf k}). Due to (3.19), for 𝐤2{\bf k}\in{\mathbb{R}}^{2}, 2|𝐤𝒗|=vF|𝐤|2|{\bf k}\cdot\bm{v}_{\sharp}|=v_{F}|{\bf k}|. Thus, for |𝐤||{\bf k}| sufficiently small, they are exactly:

μb+1(𝐤)=μDvF|𝐤|(1+η1(𝐤))\displaystyle\mu_{b+1}({\bf k})=\mu_{{}_{D}}-v_{F}|{\bf k}|(1+\eta_{1}({\bf k})) (3.22)
μb+2(𝐤)=μDvF|𝐤|(1+η2(𝐤))\displaystyle\mu_{b+2}({\bf k})=\mu_{{}_{D}}-v_{F}|{\bf k}|(1+\eta_{2}({\bf k}))
μb+3(𝐤)=μD+vF|𝐤|(1+η3(𝐤))\displaystyle\mu_{b+3}({\bf k})=\mu_{{}_{D}}+v_{F}|{\bf k}|(1+\eta_{3}({\bf k}))
μb+4(𝐤)=μD+vF|𝐤|(1+η4(𝐤))\displaystyle\mu_{b+4}({\bf k})=\mu_{{}_{D}}+v_{F}|{\bf k}|(1+\eta_{4}({\bf k}))

where ηj(𝐤)=O(|𝐤|)\eta_{j}({\bf k})=O({|}{\bf k}{|}), j=1,2,3,4j=1,2,3,4. \Box

3.3 Fourfold degeneracy with shallow potentials

Due to the discussion in section 2.1 , the left thing to do is finding a single eigenvalue of H(ϵ)H^{(\epsilon)} on χ𝐤1,τ\chi_{{\bf k}_{1},\tau} which is not an eigenvalue on χ±𝐤1,1\chi_{\pm{\bf k}_{1},1} and χs\chi_{s} for all ϵ\epsilon except a discrete set. In this subsection, we discuss the situation of ϵ\epsilon sufficiently small.

First, take ϵ=0\epsilon=0. The eigenvalue problem is

Δϕ(𝐱)=μϕ(𝐱),𝐱2,-\Delta\phi({\bf x})=\mu\phi({\bf x}),{\quad}{\bf x}\in{\mathbb{R}}^{2}, (3.23)
ϕ(𝐱+𝐮)=ϕ(𝐱),𝐮𝐔.\phi({\bf x}+{\bf u})=\phi({\bf x}),{\quad}{\bf u}\in\bf{U}. (3.24)

The operator Δ-\Delta is positive semi-definite on Hper1H^{1}_{per}. It is quite easy to know its spectrum from Sturm-Liouville Theorem and Fourier series’ presentations. Pay attention to the first eight eigenvalues:

μ1=0<μ2=μ3=μ4=μ5=μ6=μ7=|𝐤1|2<μ8\mu_{1}=0<\mu_{2}=\mu_{3}=\mu_{4}=\mu_{5}=\mu_{6}=\mu_{7}=|{\bf k}_{1}|^{2}<\mu_{8}

Obviously, a group of eigenfunctions for μ2\mu_{2}-μ7\mu_{7} are:

{ei𝐤1𝐱,eiR𝐤1𝐱,eiR2𝐤1𝐱,ei𝐤1𝐱,eiR𝐤1𝐱,eiR2𝐤1𝐱}\{e^{i{\bf k}_{1}\cdot{\bf x}},{\quad}e^{iR{\bf k}_{1}\cdot{\bf x}},{\quad}e^{iR^{2}{\bf k}_{1}\cdot{\bf x}},{\quad}e^{-i{\bf k}_{1}\cdot{\bf x}},{\quad}e^{-iR{\bf k}_{1}\cdot{\bf x}},{\quad}e^{-iR^{2}{\bf k}_{1}\cdot{\bf x}}\}

After some linear combinations, it is equivalent to:

{ϕ1(0)(𝐱)=13(ei𝐤1𝐱+τ¯eiR𝐤1𝐱+τeiR2𝐤1𝐱)χ𝐤1,τ,ϕ2(0)(𝐱)=13(ei𝐤1𝐱+τeiR𝐤1𝐱+τ¯eiR2𝐤1𝐱)χ𝐤1,τ¯,ϕ3(0)(𝐱)=13(ei𝐤1𝐱+τ¯eiR𝐤1𝐱+τeiR2𝐤1𝐱)χ𝐤1,τ,ϕ4(0)(𝐱)=13(ei𝐤1𝐱+τeiR𝐤1𝐱+τ¯eiR2𝐤1𝐱)χ𝐤1,τ¯,ϕ5(0)(𝐱)=13(ei𝐤1𝐱+eiR𝐤1𝐱+eiR2𝐤1𝐱)χ𝐤1,1,ϕ6(0)(𝐱)=13(ei𝐤1𝐱+eiR𝐤1𝐱+eiR2𝐤1𝐱)χ𝐤1,1,}.\begin{aligned} \{{\quad}&\phi_{1}^{(0)}({\bf x})=\frac{1}{\sqrt{3}}(e^{i{\bf k}_{1}\cdot{\bf x}}+\overline{\tau}e^{iR{\bf k}_{1}\cdot{\bf x}}+\tau e^{iR^{2}{\bf k}_{1}\cdot{\bf x}})&\in{\quad}&\chi_{{\bf k}_{1},\tau},\\ &\phi_{2}^{(0)}({\bf x})=\frac{1}{\sqrt{3}}(e^{i{\bf k}_{1}\cdot{\bf x}}+\tau e^{iR{\bf k}_{1}\cdot{\bf x}}+\overline{\tau}e^{iR^{2}{\bf k}_{1}\cdot{\bf x}})&\in{\quad}&\chi_{{\bf k}_{1},\overline{\tau}},\\ &\phi_{3}^{(0)}({\bf x})=\frac{1}{\sqrt{3}}(e^{-i{\bf k}_{1}\cdot{\bf x}}+\overline{\tau}e^{-iR{\bf k}_{1}\cdot{\bf x}}+\tau e^{-iR^{2}{\bf k}_{1}\cdot{\bf x}})&\in{\quad}&\chi_{-{\bf k}_{1},\tau},\\ &\phi_{4}^{(0)}({\bf x})=\frac{1}{\sqrt{3}}(e^{-i{\bf k}_{1}\cdot{\bf x}}+\tau e^{-iR{\bf k}_{1}\cdot{\bf x}}+\overline{\tau}e^{-iR^{2}{\bf k}_{1}\cdot{\bf x}})&\in{\quad}&\chi_{-{\bf k}_{1},\overline{\tau}},\\ &\phi_{5}^{(0)}({\bf x})=\frac{1}{\sqrt{3}}(e^{i{\bf k}_{1}\cdot{\bf x}}+e^{iR{\bf k}_{1}\cdot{\bf x}}+e^{iR^{2}{\bf k}_{1}\cdot{\bf x}})&\in{\quad}&\chi_{{\bf k}_{1},1},\\ &\phi_{6}^{(0)}({\bf x})=\frac{1}{\sqrt{3}}(e^{-i{\bf k}_{1}\cdot{\bf x}}+e^{-iR{\bf k}_{1}\cdot{\bf x}}+e^{-iR^{2}{\bf k}_{1}\cdot{\bf x}})&\in{\quad}&\chi_{-{\bf k}_{1},1},{\quad}\}\end{aligned}. (3.25)

Secondly, for ϵ\epsilon small enough, the 2nd7th2^{nd}-7^{th} eigenvalues of H(ϵ)H^{(\epsilon)} on χ\chi must also be eigenvalues on χ𝐤1\chi_{{\bf k}_{1}} or χ𝐤1\chi_{-{\bf k}_{1}} and separated from other eigenvalues by the continuity of the eigenvalues in ϵ\epsilon. Thus, whether the fourfold degeneracy on χ±𝐤1,τ\chi_{\pm{\bf k}_{1},\tau} and χ±𝐤1,τ¯\chi_{\pm{\bf k}_{1},\overline{\tau}} given by parity and conjugation symmetry and twofold degeneracy on χ±𝐤1,1\chi_{\pm{\bf k}_{1},1} given by conjugation symmetry will separate is most concerned about in this subsection.

Proposition 3.6.

Assume that V(𝐱)V({\bf x}) is a super honeycomb lattice potential. And H(ϵ)=Δ+ϵV(𝐱)H^{(\epsilon)}=-\Delta+\epsilon V({\bf x}) is the corresponding Schrödinger operator. Then there exists an ϵ0>0\epsilon_{0}>0 such that, for all ϵ(ϵ0,ϵ0){0}\epsilon\in(-\epsilon_{0},\epsilon_{0})\setminus\{0\}, there exists μD(ϵ)\mu_{{}_{D}}^{(\epsilon)} and 𝐯(ϵ)𝟎\bm{v}_{\sharp}^{(\epsilon)}\neq\bf{0} satisfies:

  1. 1.

    μD(ϵ)\mu_{{}_{D}}^{(\epsilon)} is an eigenvalue of H(ϵ)H^{(\epsilon)} on χ\chi of multiplicity four,

  2. 2.

    μD(ϵ)\mu_{{}_{D}}^{(\epsilon)} is a simple eigenvalue on χ𝐤1,τ\chi_{{\bf k}_{1},\tau} with eigenfunction ϕ1(ϵ)(𝐱)\phi_{1}^{(\epsilon)}({\bf x}),

  3. 3.

    𝒗(ϵ)=<ϕ1(ϵ)(𝐱),𝒫𝒞ϕ1(ϵ)(𝐱)>.\bm{v}_{\sharp}^{(\epsilon)}=<\phi_{1}^{(\epsilon)}({\bf x}),\nabla{\mathcal{P}}{\mathcal{C}}\phi_{1}^{(\epsilon)}({\bf x})>.

Remark 3.7.

From Lemma 2.5, we know that the conclusion 1 and 2 can deduce that μD(ϵ)\mu_{{}_{D}}^{(\epsilon)} is not an eigenvalue on χ±𝐤1,1\chi_{\pm{\bf k}_{1},1}.

Proof  The eigenvalue problem with shallow potentials on χ\chi is

(Δ+ϵV(𝐱))ϕ(ϵ)(𝐱)=μ(ϵ)ϕ(ϵ)(𝐱),𝐱2(-\Delta+\epsilon V({\bf x}))\phi^{(\epsilon)}({\bf x})=\mu^{(\epsilon)}\phi^{(\epsilon)}({\bf x}),{\quad}{\bf x}\in{\mathbb{R}}^{2} (3.26)

with ϕ(ϵ)(𝐱)χ\phi^{(\epsilon)}({\bf x})\in\chi. Due to discussion above and Lemma 2.6, only need to concern about the second to seventh eigenvalues.

Taking advantage of the discussion in section 2.1, we can limit this problem to each space χσ,ξ\chi_{\sigma,\xi}, where σ{𝐤1,𝐤1}\sigma\in\{{\bf k}_{1},-{\bf k}_{1}\} and ξ{1,τ,τ¯}\xi\in\{1,\tau,\overline{\tau}\}. Now ϕ(ϵ)(𝐱)χσ,ξ\phi^{(\epsilon)}({\bf x})\in\chi_{\sigma,\xi}. Analogous to the process in the proof of the last theorem, we rewrite the eigenvalue problem and divide it into two parts by projections.

Let μ(ϵ)=μ0+λ\mu^{(\epsilon)}=\mu^{0}+\lambda, ϕ(ϵ)(𝐱)=ϕ(0)(𝐱)+ψ(𝐱)\phi^{(\epsilon)}({\bf x})=\phi^{(0)}({\bf x})+\psi({\bf x}). Here μ0=|𝐤1|2\mu^{0}=|{\bf k}_{1}|^{2} and ϕ(0)(𝐱)\phi^{(0)}({\bf x}) is the corresponding eigenfunction on χσ,ξ\chi_{\sigma,\xi} in (3.25). ψ(𝐱)span{ϕ(0)(𝐱)}=ker(H(0)μ0I)\psi({\bf x})\in span\{\phi^{(0)}({\bf x})\}^{\perp}=ker(H^{(0)}-\mu^{0}I)^{\perp}. Introduce new projection operators 𝒬{\mathcal{Q}}_{\parallel} from χσ,ξ\chi_{\sigma,\xi} to ker(H(0)μ0I)ker(H^{(0)}-\mu^{0}I) and 𝒬{\mathcal{Q}}_{\perp} from χσ,ξ\chi_{\sigma,\xi} to ker(H(0)μ0I)ker(H^{(0)}-\mu^{0}I)^{\perp}. Perform 𝒬{\mathcal{Q}}_{\parallel} and 𝒬{\mathcal{Q}}_{\perp} on (3.26) to obtain

ϵ𝒬V(𝐱)ψ(𝐱)=𝒬(λϵV(𝐱))ϕ(0)(𝐱),\epsilon{\mathcal{Q}}_{\parallel}V({\bf x})\psi({\bf x})={\mathcal{Q}}_{\parallel}(\lambda-\epsilon V({\bf x}))\phi^{(0)}({\bf x}), (3.27)

and

(H(0)μ0I)ψ(𝐱)=ϵ𝒬V(𝐱)ϕ(0)(𝐱)𝒬(ϵV(𝐱)λ)ψ(𝐱).(H^{(0)}-\mu^{0}I)\psi({\bf x})=\epsilon{\mathcal{Q}}_{\perp}V({\bf x})\phi^{(0)}({\bf x})-{\mathcal{Q}}_{\perp}(\epsilon V({\bf x})-\lambda)\psi({\bf x}). (3.28)

These two equations are equivalent to the original equation, so only need to solve them. Note that (H(0)μ0I)1(H^{(0)}-\mu^{0}I)^{-1} is a bounded operator from ker(H(0)μ0I)ker(H^{(0)}-\mu^{0}I)^{\perp} to Hper1H^{1}_{per}. Thus,

(I+(H(0)μ0I)1𝒬(ϵV(𝐱)λ))ψ(𝐱)=ϵ(H(0)μ0I)1𝒬V(𝐱)ϕ(0)(𝐱)\big{(}I+(H^{(0)}-\mu^{0}I)^{-1}{\mathcal{Q}}_{\perp}(\epsilon V({\bf x})-\lambda)\big{)}\psi({\bf x})=\epsilon(H^{(0)}-\mu^{0}I)^{-1}{\mathcal{Q}}_{\perp}V({\bf x})\phi^{(0)}({\bf x}) (3.29)

Assume |ϵ|+|λ||\epsilon|+|\lambda| is small enough, (I+(H(0)μ0I)1𝒬(ϵV(𝐱)λ))1\big{(}I+(H^{(0)}-\mu^{0}I)^{-1}{\mathcal{Q}}_{\perp}(\epsilon V({\bf x})-\lambda)\big{)}^{-1} exists and is bounded. Use notation

𝒯[ϵ,λ](𝐱)=(I+(H(0)μ0I)1𝒬(ϵV(𝐱)λ))1(H(0)μ0I)1𝒬V(𝐱)ϕ(0)(𝐱).{\mathcal{T}}[\epsilon,\lambda]({\bf x})=\big{(}I+(H^{(0)}-\mu^{0}I)^{-1}{\mathcal{Q}}_{\perp}(\epsilon V({\bf x})-\lambda)\big{)}^{-1}(H^{(0)}-\mu^{0}I)^{-1}{\mathcal{Q}}_{\perp}V({\bf x})\phi^{(0)}({\bf x}).

𝒯{\mathcal{T}} is bounded in Hper1H^{1}_{per} when |ϵ|+|λ||\epsilon|+|\lambda| is small enough. Substituting ψ(𝐱)=ϵ𝒯[ϵ,λ](𝐱)\psi({\bf x})=\epsilon{\mathcal{T}}[\epsilon,\lambda]({\bf x}) into (3.27), we have

ϵ2𝒬V(𝐱)𝒯[ϵ,λ](𝐱)=𝒬(λϵV(𝐱))ϕ(0)(𝐱)\epsilon^{2}{\mathcal{Q}}_{\parallel}V({\bf x}){\mathcal{T}}[\epsilon,\lambda]({\bf x})={\mathcal{Q}}_{\parallel}(\lambda-\epsilon V({\bf x}))\phi^{(0)}({\bf x}) (3.30)

It is the same with:

ϵ2<V(𝐱)𝒯[ϵ,λ](𝐱),ϕ0(𝐱)>=<(λϵV(𝐱))ϕ(0)(𝐱),ϕ(0)(𝐱)>\epsilon^{2}<V({\bf x}){\mathcal{T}}[\epsilon,\lambda]({\bf x}),\phi^{0}({\bf x})>=<(\lambda-\epsilon V({\bf x}))\phi^{(0)}({\bf x}),\phi^{(0)}({\bf x})>

Solve it to attain

λ(ϵ)=<V(𝐱)ϕ(0)(𝐱),ϕ(0)(𝐱)>ϵ+<V(𝐱)𝒯[ϵ,λ](𝐱),ϕ(0)(𝐱)>ϵ2\lambda(\epsilon)=<V({\bf x})\phi^{(0)}({\bf x}),\phi^{(0)}({\bf x})>\epsilon+<V({\bf x}){\mathcal{T}}[\epsilon,\lambda]({\bf x}),\phi^{(0)}({\bf x})>\epsilon^{2} (3.31)

Thus, we solve out a λ(ϵ)\lambda(\epsilon) of order O(|ϵ|)O(|\epsilon|). This means there exactly exists a Floquet-Bloch eigenpair (μD(ϵ),ϕ(ϵ)(𝐱))(\mu_{{}_{D}}{(\epsilon)},\phi^{(\epsilon)}({\bf x})) with μD(ϵ)\mu_{{}_{D}}^{(\epsilon)} changing in order O(|ϵ|)O(|\epsilon|) on χσ,ξ\chi_{\sigma,\xi}.

First we take σ=𝐤1\sigma={\bf k}_{1} and ξ=τ\xi=\tau, consider ϕ1(ϵ)(𝐱)=ϕ1(0)(𝐱)+ϵ𝒯[ϵ,λ](𝐱)\phi^{(\epsilon)}_{1}({\bf x})=\phi^{(0)}_{1}({\bf x})+\epsilon{\mathcal{T}}[\epsilon,\lambda]({\bf x}). Provided |ϵ||\epsilon| sufficiently small, 𝒯[ϵ,λ(ϵ)]Hper1\parallel{\mathcal{T}}[\epsilon,\lambda(\epsilon)]\parallel_{H^{1}_{per}} is of order O(|ϵ)|O(|\epsilon)|, therefore

𝒗(ϵ)=<ϕ1(0)(𝐱),𝒫𝒞ϕ1(0)(𝐱)>+O(|ϵ|)0.\bm{v}_{\sharp}^{(\epsilon)}=<\phi^{(0)}_{1}({\bf x}),\nabla{\mathcal{P}}{\mathcal{C}}\phi^{(0)}_{1}({\bf x})>+O(|\epsilon|)\neq 0.

Secondly, traversing all σ\sigma and ξ\xi, the six eigenpairs on χσ,ξ\chi_{\sigma,\xi} give the second to seventh eigenvalues for the original problem on χ\chi when ϵ\epsilon is quite small. As mentioned above, four on χ𝐤1,τ\chi_{{\bf k}_{1},\tau}, χ𝐤1,τ¯\chi_{{\bf k}_{1},\overline{\tau}}, χ𝐤1,τ\chi_{-{\bf k}_{1},\tau} and χ𝐤1,τ¯\chi_{{\bf k}_{1},\overline{\tau}} are bound to each other due to symmetry. The same for χ𝐤1,1\chi_{{\bf k}_{1},1} and χ𝐤1,1\chi_{-{\bf k}_{1},1}. After some basic calculations, the λ(ϵ)\lambda(\epsilon) on χσ,ξ\chi_{\sigma,\xi} is

λσ,ξ(ϵ)=(c1+(ξ+ξ¯)c2)ϵ+O(|ϵ|2),\lambda_{\sigma,\xi}(\epsilon)=(c_{1}+(\xi+\overline{\xi})c_{2})\epsilon+O(|\epsilon|^{2}), (3.32)

where c1=<1,V(𝐱)>c_{1}=<1,V({\bf x})> and c2=<ei𝐪1𝐱,V(𝐱)>c_{2}=<e^{i{\bf q}_{1}\cdot{\bf x}},V({\bf x})>. c20c_{2}\neq 0 is the non-degeneracy condition of super honeycomb lattice potentials. It is obvious that λ+,τ=λ+,τ¯=λ,τ=λ,τ¯λ+,1=λ,1\lambda_{+,\tau}=\lambda_{+,\overline{\tau}}=\lambda_{-,\tau}=\lambda_{-,\overline{\tau}}\neq\lambda_{+,1}=\lambda_{-,1}. This explains these six branches will separate into fourfold and twofold for super honeycomb lattice potentials. \Box

Proposition 3.8.

With conditions in Proposition 3.6 and ϵ\epsilon sufficiently small, for the eigenvalue problem (3.26) on χ\chi, if <ei𝐪1𝐱,V(𝐱)><e^{i{\bf q}_{1}\cdot{\bf x}},V({\bf x})> is positive, then

μ2ϵ=μ3ϵ=μ4ϵ=μ5ϵ<μ6ϵ=μ7ϵ,\mu^{\epsilon}_{2}=\mu^{\epsilon}_{3}=\mu^{\epsilon}_{4}=\mu^{\epsilon}_{5}<\mu^{\epsilon}_{6}=\mu^{\epsilon}_{7},

and if <ei𝐪1𝐱,V(𝐱)><e^{i{\bf q}_{1}\cdot{\bf x}},V({\bf x})> is negative, then

μ2ϵ=μ3ϵ<μ4ϵ=μ5ϵ=μ6ϵ=μ7ϵ\mu^{\epsilon}_{2}=\mu^{\epsilon}_{3}<\mu^{\epsilon}_{4}=\mu^{\epsilon}_{5}=\mu^{\epsilon}_{6}=\mu^{\epsilon}_{7}

Proof  It is easy to verify using (3.32). \Box

3.4 Proof of the main theorem

Last subsection gives the result of shallow potentials, and this subsection briefly shows the method to verify Theorem 3.1, the generic case. The key strategy is constructing an analytic function (μ,ϵ){\mathcal{E}}(\mu,\epsilon) on each χσ,ξ\chi_{\sigma,\xi}, whose zero points are eigenvalues of HVH_{V} on function spaces we concern. This function is actually the determinant of infinite dimensional linear operator H(ϵ)=Δ+ϵV(𝐱)H^{(\epsilon)}=-\Delta+\epsilon V({\bf x}) by some renormalization using trace class. Because of the symmetry in Lemma 2.5 again, obviously the eigenvalue should exist simultaneously on χ𝐤1,τ\chi_{{\bf k}_{1},\tau}, χ𝐤1,τ¯\chi_{{\bf k}_{1},\overline{\tau}}, χ𝐤1,τ\chi_{-{\bf k}_{1},\tau} and χ𝐤1,τ¯\chi_{-{\bf k}_{1},\overline{\tau}}. And this eigenvalue should be different from those on other subspaces. The main work is focused on establishing this analytic function and prove the three conditions in Proposition 3.2.

Without loss of generality, let us assume that 0V(𝐱)Vm0\leq V({\bf x})\leq V_{m}. On each χσ,ξ\chi_{\sigma,\xi}, the eigenvalue problem is:

(Δ+ϵV(𝐱))Φ(𝐱)=μΦ(𝐱).(-\Delta+\epsilon V({\bf x}))\Phi({\bf x})=\mu\Phi({\bf x}).

In this subsection, we consider ϵ\epsilon\in{\mathbb{C}}. First consider ϵ\epsilon with nonnegative real part Re(ϵ)0Re(\epsilon)\geq 0. Our aim is to derive an operator whose determinant can be defined. Hence, we rewrite it as

(IΔ+ϵV(𝐱))Φ(𝐱)=(μ+1)Φ(𝐱),(I-\Delta+\epsilon V({\bf x}))\Phi({\bf x})=(\mu+1)\Phi({\bf x}), (3.33)

Note that (IΔ+ϵV(𝐱))(I-\Delta+\epsilon V({\bf x})) is invertible. Let 𝒜(ϵ)=(IΔ+ϵV(𝐱))1{\mathcal{A}}(\epsilon)=(I-\Delta+\epsilon V({\bf x}))^{-1}, then

(I(μ+1)𝒜(ϵ))Φ(𝐱)=0.(I-(\mu+1){\mathcal{A}}(\epsilon))\Phi({\bf x})=0. (3.34)

The jthj^{th} eigenvalue λj\lambda_{j} on χ\chi of 𝒜(ϵ){\mathcal{A}}(\epsilon) is asymptotic to j1j^{-1}. Thus, it is a Hilbert-Schmidt operator:

𝒜2(ϵ)=j|λj|2j|j|2<.\parallel{\mathcal{A}}^{2}(\epsilon)\parallel=\sum_{j}|\lambda_{j}|^{-2}\sim\sum_{j}|j|^{-2}<\infty.

Now a determinant for (I(μ+1)𝒜(ϵ))(I-(\mu+1){\mathcal{A}}(\epsilon)) can be defined through a regularized way:

det2(I(μ+1)𝒜(ϵ))=det(I+R2((μ+1)𝒜(ϵ))).det_{2}(I-(\mu+1){\mathcal{A}}(\epsilon))=det(I+R_{2}((\mu+1){\mathcal{A}}(\epsilon))). (3.35)

The right-hand side is Fredholm determinant. It is well-defined because the regularized form:

R2((μ+1)𝒜(ϵ))=(I+(μ+1)𝒜(ϵ))e(μ+1)𝒜(ϵ)IR_{2}((\mu+1){\mathcal{A}}(\epsilon))=(I+(\mu+1){\mathcal{A}}(\epsilon))e^{-(\mu+1){\mathcal{A}}(\epsilon)}-I (3.36)

is a trace class.

We already have the following lemma.

Lemma 3.9.

For all σ{s,𝐤1,𝐤1}\sigma\in\{s,{\bf k}_{1},-{\bf k}_{1}\} and ξ{1,τ,τ¯}\xi\in\{1,\tau,\overline{\tau}\}, the following is true:

  1. 1.

    ϵ𝒜(ϵ)\epsilon\to{\mathcal{A}}(\epsilon) is an analytic mapping from {ϵ,Re(ϵ)0}\{\epsilon\in{\mathbb{C}},Re(\epsilon)\geq 0\} to the space of Hilbert-Schmidt operators on χσ,ξ\chi_{\sigma,\xi},

  2. 2.

    The regularized determinant on χσ,ξ\chi_{\sigma,\xi}

    σ,ξ(μ,ϵ)=det2(I(μ+1)𝒜(ϵ)){\mathcal{E}}_{\sigma,\xi}(\mu,\epsilon)=det_{2}(I-(\mu+1){\mathcal{A}}(\epsilon))

    is analytic for both μ\mu and ϵ\epsilon with Re(ϵ)0Re(\epsilon)\geq 0,

  3. 3.

    For ϵ\epsilon real and nonnegative, μ\mu is a χσ,ξ\chi_{\sigma,\xi} eigenvalue for HVH_{V} of multiplicity mm if and only if it is a zero of σ,ξ(μ,ϵ){\mathcal{E}}_{\sigma,\xi}(\mu,\epsilon) of multiplicity mm.

Therefore we first need to prove that except a discrete set in {\mathbb{R}} for ϵ\epsilon, 𝐤1,τ{\mathcal{E}}_{{\bf k}_{1},\tau} has a simple zero (ϵ,μ)(\epsilon,\mu) which is not a zero of ±𝐤1,1{\mathcal{E}}_{\pm{\bf k}_{1},1} and s,ξ{\mathcal{E}}_{s,\xi} for all ξ\xi. Then prove that the eigenfunction on χ𝐤1,τ\chi_{{\bf k}_{1},\tau} corresponding to this μ\mu has nonzero 𝒗\bm{v}_{\sharp}, which is defined in Proposition 3.2. This is totally the same with the proof in [13], based on the previous subsection’s conclusion about shallow potentials. The only difference is about symmetry, which will not influence the proof.

For ϵ\epsilon with real part negative, just replace (3.33) with

(IΔ+ϵ(V(𝐱)Vm))Φ(𝐱)=(μ+1ϵVm)Φ(𝐱)(I-\Delta+\epsilon(V({\bf x})-V_{m}))\Phi({\bf x})=(\mu+1-\epsilon V_{m})\Phi({\bf x}) (3.37)

and replace (3.34) with

(I(μ+1ϵVm)𝒜~(ϵ)Φ(𝐱)=0,(I-(\mu+1-\epsilon V_{m})\tilde{{\mathcal{A}}}(\epsilon)\Phi({\bf x})=0, (3.38)

where 𝒜~(ϵ)=(IΔ+ϵ(V(𝐱)Vm))1\tilde{{\mathcal{A}}}(\epsilon)=(I-\Delta+\epsilon(V({\bf x})-V_{m}))^{-1}. The rest is similar.

4 Instability under symmetry breaking perturbations

We already show the existence of double cones for the operator HVH_{V} with a super honeycomb lattice potential V(𝐱)V({\bf x}). This section focuses on what will happen if the additional translation symmetry of the potential is broken. In other words, we investigate the behaviour of the band structures of HVH_{V} after some perturbations.

4.1 Perturbations breaking additional translation symmetry

Let us observe small perturbations which break this additional translation symmetry as in Figure 3. Shrinking and expanding the hexagons in super honeycomb lattice obtain the new lattices with red vertices. These perturbed lattices are not super honeycomb lattices any longer, and the four branches of energy bands intersecting under super honeycomb lattice potentials’ cases separate into two parts as in Figure 4.

Generally, let W(𝐱)W({\bf x}) be a bounded real function that can be written in the form

W(𝐱)=ei𝐤1𝐱p(𝐱)+ei𝐤1𝐱p(𝐱),ei𝐤1𝐱p(𝐱)χ𝐤1,1W({\bf x})=e^{i{\bf k}_{1}\cdot{\bf x}}p({\bf x})+e^{-i{\bf k}_{1}\cdot{\bf x}}p(-{\bf x}),{\quad}e^{i{\bf k}_{1}\cdot{\bf x}}p({\bf x})\in\chi_{{\bf k}_{1},1} (4.1)

Obviously, W(𝐱)W({\bf x}) is even, and it is {\mathcal{R}}-invariant.

Refer to caption
Refer to caption
Figure 3: Perturbed lattice based on a super honeycomb lattice: (a) the shrunk hexagonal lattice, and (b) the expanded hexagonal lattice. Grey points show the structure of super honeycomb lattice, and red points show the perturbed case. The scalar rr is the ratio of the distance of the components in a dimer to the length of periodicity of honeycomb lattice.

After adding a perturbation of W(𝐱)W({\bf x}), the Schrödinger operator is

Hδ=HV+δW(𝐱)=Δ+V(𝐱)+δW(𝐱)H^{\delta}=H_{V}+\delta W({\bf x})=-\Delta+V({\bf x})+\delta W({\bf x}) (4.2)

The potential V(𝐱)+δW(𝐱)V({\bf x})+\delta W({\bf x}) here has {\mathcal{R}}-symmetry, 𝒫𝒞{\mathcal{P}}{\mathcal{C}}-symmetry and translation symmetry. It is still a honeycomb lattice potential, but it is no more a super honeycomb lattice potential .

Remark 4.1.

Different ways to break the additional translation symmetry shown above represented by shrinking and expanding the lattices give different topological properties, which can be characterised by the Chern number. And gluing these two different topological materials together, say, the shrunk one and expanded one, by a domain wall, generates a new material with interesting edge states. This phenomenon will be researched in our forthcoming paper.

4.2 Band Structures after perturbations

Consider the perturbed eigenvalue problem:

HδΦ(𝐱)=μδ(𝐤)Φ(𝐱),𝐱2H^{\delta}\Phi({\bf x})=\mu^{\delta}({\bf k})\Phi({\bf x}),{\quad}{\bf x}\in{\mathbb{R}}^{2} (4.3)
Φ(𝐱+𝐮)=ei𝐤𝐮Φ(𝐱),𝐮𝐔\Phi({\bf x}+{\bf u})=e^{i{\bf k}\cdot{\bf u}}\Phi({\bf x}),{\quad}{\bf u}\in\bf{U} (4.4)

Here 𝐤{\bf k} is in Ω\Omega^{*}. Again, let Φ(𝐱)=ei𝐤𝐱ϕ(𝐱)\Phi({\bf x})=e^{i{\bf k}\cdot{\bf x}}\phi({\bf x}) and denote Hδ(𝐤)=(+i𝐤)(+i𝐤)+V(𝐱)+δW(𝐱)H^{\delta}({\bf k})=-(\nabla+i{\bf k})\cdot(\nabla+i{\bf k})+V({\bf x})+\delta W({\bf x}). We deduce that

Hδ(𝐤)ϕ(𝐱)=μδ(𝐤)ϕ(𝐱),𝐱2,H^{\delta}({\bf k})\phi({\bf x})=\mu^{\delta}({\bf k})\phi({\bf x}),{\quad}{\bf x}\in{\mathbb{R}}^{2}, (4.5)
ϕ(𝐱+𝐮)=ϕ(𝐱),𝐮𝐔.\phi({\bf x}+{\bf u})=\phi({\bf x}),{\quad}{\bf u}\in\bf{U}. (4.6)
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Top: honeycomb dimer potentials W(𝐱,r)W({\bf x},r) defined by (5.1), (2.5), and (2.6) with rr taking value (a) 1.053\frac{1.05}{3}, (b) 13\frac{1}{3}, and (c) 0.9753\frac{0.975}{3}. We paint three periods in both 𝐮1{\bf u}_{1} and 𝐮2{\bf u}_{2} directions. Middle: corresponding 2nd5th2^{nd}-5^{th} energy surfaces of H(r)=Δ+W(𝐱,r)H(r)=-\Delta+W({\bf x},r). Bottom: corresponding 2nd7th2^{nd}-7^{th} energy bands along λ𝐤1\lambda{\bf k}_{1}.

The following theorem shows that the fourfold degeneracy and double Dirac cone are not protected after the symmetry breaking perturbations we discuss in last subsection.

Theorem 4.2.

Let V(𝐱)V({\bf x}) be a super honeycomb lattice potential and W(x) be as in (4.1). HδH^{\delta} is as in (4.2). Assume that:

  1. 1.

    μD\mu_{{}_{D}} is an eigenvalue of HVH_{V} on χ\chi of multiplicity four:

    μb0(𝟎)<μD=μb+10(𝟎)=μb+20(𝟎)=μb+30(𝟎)=μb+40(𝟎)<μb+50(𝟎),\mu_{b}^{0}(\bm{0})<\mu_{{}_{D}}=\mu_{b+1}^{0}(\bm{0})=\mu_{b+2}^{0}(\bm{0})=\mu_{b+3}^{0}(\bm{0})=\mu_{b+4}^{0}(\bm{0})<\mu_{b+5}^{0}(\bm{0}),
  2. 2.

    μD\mu_{{}_{D}} is a simple eigenvalue on χ𝐤1,τ\chi_{{\bf k}_{1},\tau} with normalized eigenfunction ϕ1(𝐱)\phi_{1}({\bf x}),

  3. 3.

    𝒗=<ϕ1(𝐱),𝒫𝒞[ϕ1](𝐱)>𝟎\bm{v}_{\sharp}=<\phi_{1}({\bf x}),\nabla{\mathcal{P}}{\mathcal{C}}[\phi_{1}]({\bf x})>\neq\bf{0},

  4. 4.

    c=<ϕ1(𝐱),W(𝐱)ϕ3(𝐱)>0.c_{\sharp}=<\phi_{1}({\bf x}),W({\bf x})\phi_{3}({\bf x})>\neq 0.

Then due to Proposition 3.2, there exists double Dirac cone in the vicinity of the Γ\Gamma point for HVH_{V}. We claim that exists δ0>0\delta_{0}>0, such that for all δ(δ0,δ0){0}\delta\in(-\delta_{0},\delta_{0})\setminus\{0\}, the (b+1)th(b+1)^{th} to (b+4)th(b+4)^{th} energy bands of HδH^{\delta} will open a gap in a small neighbourhood of the Γ\Gamma point.

Proof  Again, employ the Lyapunov- Schmidt reduction strategy. Without loss of generality, assume <ϕl(𝐱),ϕj(𝐱)>=δl,j<\phi_{l}({\bf x}),\phi_{j}({\bf x})>=\delta_{l,j}.

Now settle (4.5)-(4.5). Let μδ(𝐤)=μD+λ\mu^{\delta}({\bf k})=\mu_{{}_{D}}+\lambda and ϕ(𝐱)=αjϕj(𝐱)+ψ(𝐱)\phi({\bf x})=\alpha^{j}\phi_{j}({\bf x})+\psi({\bf x}). ψ(𝐱)ker(HVμDI)\psi({\bf x})\in ker(H_{V}-\mu_{{}_{D}}I)^{\perp}. Then (4.5) is:

(HVμDI)ψ(𝐱)=(2i𝐤|𝐤|2δW(𝐱)+λ)(αjϕj(𝐱)+ψ(𝐱))(H_{V}-\mu_{{}_{D}}I)\psi({\bf x})=(2i{\bf k}\nabla-|{\bf k}|^{2}-\delta W({\bf x})+\lambda)(\alpha^{j}\phi_{j}({\bf x})+\psi({\bf x}))

Apply the projection 𝒬{\mathcal{Q}}_{\parallel} from χ\chi to ker(HVμDI)ker(H_{V}-\mu_{{}_{D}}I) and 𝒬{\mathcal{Q}}_{\perp} from χ\chi to ker(HVμDI)ker(H_{V}-\mu_{{}_{D}}I)^{\perp} to obtain the equation:

𝒬(δW(𝐱)2ik)ψ(𝐱)=𝒬(2i𝐤|𝐤|2δW(𝐱)+λ)αjϕj(𝐱),{\mathcal{Q}}_{\parallel}(\delta W({\bf x})-2ik\nabla)\psi({\bf x})={\mathcal{Q}}_{\parallel}(2i{\bf k}\nabla-|{\bf k}|^{2}-\delta W({\bf x})+\lambda)\alpha^{j}\phi_{j}({\bf x}), (4.7)

and

(HVμDI)ψ(𝐱)=\displaystyle(H_{V}-\mu_{{}_{D}}I)\psi({\bf x})= 𝒬(2i𝐤δW(𝐱))αjϕj(𝐱)\displaystyle{\mathcal{Q}}_{\perp}(2i{\bf k}\nabla-\delta W({\bf x}))\alpha^{j}\phi_{j}({\bf x}) (4.8)
+𝒬(2i𝐤|𝐤|2δW(𝐱)+λ)ψ(𝐱).\displaystyle+{\mathcal{Q}}_{\perp}(2i{\bf k}\nabla-|{\bf k}|^{2}-\delta W({\bf x})+\lambda)\psi({\bf x}).

After the analogous procedure to Proposition 3.2, Let

𝒜=(I(HVμDI)1𝒬(2i𝐤|𝐤|2δW(𝐱)+λ))1,{\mathcal{A}}=\big{(}I-(H_{V}-\mu_{{}_{D}}I)^{-1}{\mathcal{Q}}_{\perp}(2i{\bf k}\nabla-|{\bf k}|^{2}-\delta W({\bf x})+\lambda)\big{)}^{-1},
𝒯j[𝐤,δ,λ](𝐱)=𝒜(HVμD)1𝒬(2i𝐤δW(𝐱))ϕj(𝐱).{\mathcal{T}}_{j}[{\bf k},\delta,\lambda]({\bf x})={\mathcal{A}}(H_{V}-\mu_{{}_{D}})^{-1}{\mathcal{Q}}_{\perp}(2i{\bf k}\nabla-\delta W({\bf x}))\phi_{j}({\bf x}).

These operators exist and 𝒯j{\mathcal{T}}_{j} a bounded operator from 3{\mathbb{R}}^{3} to Hper1H^{1}_{per} of order O(|𝐤|+|δ|)O(|{\bf k}|+|\delta|) for (|𝐤|+|δ|+|λ||{\bf k}|+|\delta|+|\lambda|) small enough. Now ψ(𝐱)=αj𝒯j[𝐤,δ,λ](𝐱)\psi({\bf x})=\alpha^{j}{\mathcal{T}}_{j}[{\bf k},\delta,\lambda]({\bf x}). Put it into (4.7), and apply inner production with {ϕl(𝐱)}l\{\phi_{l}({\bf x})\}_{l}. Finally, we get linear equations M(𝐤,δ,λ)(αj)=0M({\bf k},\delta,\lambda)(\alpha^{j})=0, where

M(𝐤,δ,λ)=(<ϕl(𝐱),(λ+2i𝐤δW(𝐱)|𝐤|2)ϕj(𝐱)\displaystyle M({\bf k},\delta,\lambda)=(<\phi_{l}({\bf x}),(\lambda+2i{\bf k}\nabla-\delta W({\bf x})-|{\bf k}|^{2})\phi_{j}({\bf x})
+(2i𝐤δW(𝐱))𝒯j[𝐤,δ,λ](𝐱)>)l,j.\displaystyle+(2i{\bf k}\nabla-\delta W({\bf x})){\mathcal{T}}_{j}[{\bf k},\delta,\lambda]({\bf x})>)_{l,j}.

The truncated linear part of M(𝐤,δ,λ)M({\bf k},\delta,\lambda), or the bifurcation matrix is

M0(𝐤,δ,λ)=(<ϕl(𝐱),(λ+2i𝐤δW(𝐱))ϕj(𝐱)>)l,jM_{0}({\bf k},\delta,\lambda)=(<\phi_{l}({\bf x}),(\lambda+2i{\bf k}\nabla-\delta W({\bf x}))\phi_{j}({\bf x})>)_{l,j} (4.9)

The rest part is of order O(|𝐤|2+|𝐤||δ|+|δ|2)O(|{\bf k}|^{2}+|{\bf k}||\delta|+|\delta|^{2}) when |λ||\lambda| is sufficiently small:

M1(𝐤,δ,λ)=(<ϕl(𝐱),|𝐤|2ϕj(𝐱)+(2i𝐤δW(𝐱))𝒯j[𝐤,δ,λ](𝐱)>)l,jM_{1}({\bf k},\delta,\lambda)=(<\phi_{l}({\bf x}),-|{\bf k}|^{2}\phi_{j}({\bf x})+(2i{\bf k}\nabla-\delta W({\bf x})){\mathcal{T}}_{j}[{\bf k},\delta,\lambda]({\bf x})>)_{l,j} (4.10)

Thus,

det(M(𝐤,δ,λ))\displaystyle det(M({\bf k},\delta,\lambda)) =det(M0(𝐤,δ,λ)+M1(𝐤,δ,λ))\displaystyle=det(M_{0}({\bf k},\delta,\lambda)+M_{1}({\bf k},\delta,\lambda))
=det(M0(𝐤,δ,λ))+O((|𝐤|2+|𝐤||δ|+|δ|2)4)\displaystyle=det(M_{0}({\bf k},\delta,\lambda))+O((|{\bf k}|^{2}+|{\bf k}||\delta|+|\delta|^{2})^{4})

To calculate M0(𝐤,δ,λ)M_{0}({\bf k},\delta,\lambda), first note that:

Proposition 4.3.

There exists a real number cc_{\sharp} such that

(<ϕl(𝐱),W(𝐱)ϕj(𝐱)>)l,j=(00c0000cc0000c00).(<\phi_{l}({\bf x}),W({\bf x})\phi_{j}({\bf x})>)_{l,j}=\begin{pmatrix}0&0&c_{\sharp}&0\\ 0&0&0&c_{\sharp}\\ c_{\sharp}&0&0&0\\ 0&c_{\sharp}&0&0\\ \end{pmatrix}. (4.11)

Proof𝒫{\mathcal{P}}, 𝒞{\mathcal{C}} and {\mathcal{R}} are unit transformations. Because W(𝐱)W({\bf x}) is even , real and {\mathcal{R}}-invariant, it is true that for any ff, gg χ\in\chi:

<f(𝐱),W(𝐱)g(𝐱)>\displaystyle<f({\bf x}),W({\bf x})g({\bf x})> =<𝒫f(𝐱),𝒫(W(𝐱)g(𝐱)))>\displaystyle=<{\mathcal{P}}f({\bf x}),{\mathcal{P}}(W({\bf x})g({\bf x})))> (4.12)
=<𝒫f(𝐱),W(𝐱)𝒫g(𝐱))>,\displaystyle=<{\mathcal{P}}f({\bf x}),W({\bf x}){\mathcal{P}}g({\bf x}))>,
<f(𝐱),W(𝐱)g(𝐱)>\displaystyle<f({\bf x}),W({\bf x})g({\bf x})> =<f(𝐱),(W(𝐱)g(𝐱)))>\displaystyle=<{\mathcal{R}}f({\bf x}),{\mathcal{R}}(W({\bf x})g({\bf x})))> (4.13)
=<f(𝐱),W(𝐱)g(𝐱))>.\displaystyle=<{\mathcal{R}}f({\bf x}),W({\bf x}){\mathcal{R}}g({\bf x}))>.

First use (4.12) to obtain cc_{\sharp} is real:

c¯=<ϕ1(𝐱),W(𝐱)ϕ3(𝐱)>¯=<ϕ3(𝐱),W(𝐱)ϕ1(𝐱)>=<ϕ1(𝐱),W(𝐱)ϕ3(𝐱)>=c.\overline{c_{\sharp}}=\overline{<\phi_{1}({\bf x}),W({\bf x})\phi_{3}({\bf x})>}=<\phi_{3}({\bf x}),W({\bf x})\phi_{1}({\bf x})>=<\phi_{1}({\bf x}),W({\bf x})\phi_{3}({\bf x})>=c_{\sharp}.

In addition,

<ϕ2(𝐱),W(𝐱)ϕ4(𝐱)>\displaystyle<\phi_{2}({\bf x}),W({\bf x})\phi_{4}({\bf x})> =<𝒞ϕ3(𝐱),𝒞(W(𝐱)ϕ1(𝐱))>=c¯\displaystyle=<{\mathcal{C}}\phi_{3}({\bf x}),{\mathcal{C}}(W({\bf x})\phi_{1}({\bf x}))>=\overline{c_{\sharp}}
=<ϕ4(𝐱),W(𝐱)ϕ2(𝐱)>=c.\displaystyle=<\phi_{4}({\bf x}),W({\bf x})\phi_{2}({\bf x})>=c_{\sharp}.

Then take f(𝐱)=ϕ1(𝐱)f({\bf x})=\phi_{1}({\bf x}) and g(𝐱)=ϕ2(𝐱)g({\bf x})=\phi_{2}({\bf x}). Since ϕ1(𝐱)\phi_{1}({\bf x}) and ϕ2(𝐱)\phi_{2}({\bf x}) are eigenfunctions with different eigenvalues for {\mathcal{R}}, <ϕ1(𝐱),W(𝐱)ϕ2(𝐱)>=0<\phi_{1}({\bf x}),W({\bf x})\phi_{2}({\bf x})>=0. The same for <ϕ3(𝐱),W(𝐱)ϕ1(𝐱)><\phi_{3}({\bf x}),W({\bf x})\phi_{1}({\bf x})>, <ϕ2(𝐱),W(𝐱)ϕ4(𝐱)><\phi_{2}({\bf x}),W({\bf x})\phi_{4}({\bf x})> and <ϕ4(𝐱),W(𝐱)ϕ2(𝐱)><\phi_{4}({\bf x}),W({\bf x})\phi_{2}({\bf x})>. Then taking f(𝐱)=g(𝐱)=ϕ1(𝐱)f({\bf x})=g({\bf x})=\phi_{1}({\bf x}), using (4.1), due to the orthogonality of χs\chi_{s}, χ𝐤1\chi_{{\bf k}_{1}} and χ𝐤1\chi_{-{\bf k}_{1}}, we have

<ϕ1(𝐱),W(𝐱)ϕ1(𝐱)>=<ϕ1(𝐱),(ei𝐤1𝐱p(𝐱)+ei𝐤1𝐱p(𝐱))ϕ1(𝐱)>=0.<\phi_{1}({\bf x}),W({\bf x})\phi_{1}({\bf x})>=<\phi_{1}({\bf x}),(e^{i{\bf k}_{1}\cdot{\bf x}}p({\bf x})+e^{-i{\bf k}_{1}\cdot{\bf x}}p(-{\bf x}))\phi_{1}({\bf x})>=0.

It is the same that <ϕl(𝐱),W(𝐱)ϕl(𝐱)>=0<\phi_{l}({\bf x}),W({\bf x})\phi_{l}({\bf x})>=0 for all ll. Taking advantage of the discussion above, it is trivial to verify the result. \Box

Apply the equations (3.18) and the proposition above to get the bifurcation matrix after perturbations:

M0(𝐤,δ,λ)=(λ2i𝐤𝒗δc02i𝐤𝒗¯λ0δcδc0λ2i𝐤𝒗0δc2i𝐤𝒗¯λ)M_{0}({\bf k},\delta,\lambda)=\begin{pmatrix}\lambda&2i{\bf k}\cdot\bm{v}_{\sharp}&-\delta c_{\sharp}&0\\ \overline{2i{\bf k}\cdot\bm{v}_{\sharp}}&\lambda&0&-\delta c_{\sharp}\\ -\delta c_{\sharp}&0&\lambda&-2i{\bf k}\cdot\bm{v}_{\sharp}\\ 0&-\delta c_{\sharp}&-\overline{2i{\bf k}\cdot\bm{v}_{\sharp}}&\lambda\end{pmatrix} (4.14)

Thus, det(M0(𝐤,δ,λ))=(λ2(δc)24|𝐤𝒗|2)2det(M_{0}({\bf k},\delta,\lambda))=(\lambda^{2}-(\delta c_{\sharp})^{2}-4|{\bf k}\cdot\bm{v}_{\sharp}|^{2})^{2}. Now solve det(M(𝐤,δ,λ))=(λ2(δc)24|𝐤𝒗|2)2+O((|𝐤|2+|𝐤||δ|+|δ|2)4)det(M({\bf k},\delta,\lambda))=(\lambda^{2}-(\delta c_{\sharp})^{2}-4|{\bf k}\cdot\bm{v}_{\sharp}|^{2})^{2}+O((|{\bf k}|^{2}+|{\bf k}||\delta|+|\delta|^{2})^{4}). Use notation (3.19) here. For δ\delta small, with |𝐤||{\bf k}| small enough too, this gives four branches of this eigenvalue problem are:

μb+1δ(𝐤)=μD(δ|c|)2+|vF𝐤|2+O(|δ|2+|δ||𝐤|+|𝐤|2),\displaystyle\mu_{b+1}^{\delta}({\bf k})=\mu_{{}_{D}}-\sqrt{(\delta|c_{\sharp}|)^{2}+|v_{F}\cdot{\bf k}|^{2}}+O(|\delta|^{2}+|\delta||{\bf k}|+|{\bf k}|^{2}),
μb+2δ(𝐤)=μD(δ|c|)2+|vF𝐤|2+O(|δ|2+|δ||𝐤|+|𝐤|2),\displaystyle\mu_{b+2}^{\delta}({\bf k})=\mu_{{}_{D}}-\sqrt{(\delta|c_{\sharp}|)^{2}+|v_{F}\cdot{\bf k}|^{2}}+O(|\delta|^{2}+|\delta||{\bf k}|+|{\bf k}|^{2}),
μb+3δ(𝐤)=μD+(δ|c|)2+|vF𝐤|2+O(|δ|2+|δ||𝐤|+|𝐤|2),\displaystyle\mu_{b+3}^{\delta}({\bf k})=\mu_{{}_{D}}+\sqrt{(\delta|c_{\sharp}|)^{2}+|v_{F}\cdot{\bf k}|^{2}}+O(|\delta|^{2}+|\delta||{\bf k}|+|{\bf k}|^{2}),
μb+4δ(𝐤)=μD+(δ|c|)2+|vF𝐤|2+O(|δ|2+|δ||𝐤|+|𝐤|2).\displaystyle\mu_{b+4}^{\delta}({\bf k})=\mu_{{}_{D}}+\sqrt{(\delta|c_{\sharp}|)^{2}+|v_{F}\cdot{\bf k}|^{2}}+O(|\delta|^{2}+|\delta||{\bf k}|+|{\bf k}|^{2}).

Therefore, it is obvious that if c0c_{\sharp}\neq 0, these four bands will open a gap near 𝐤=𝟎{\bf k}=\bf{0} for δ\delta sufficiently small but nonzero, which means the fourfold degeneracy and double cone are not protected. \Box

5 Numerical results

In this section, we numerically compute the band structures for a smooth and a piecewise constant potentials to illustrate our analysis in above sections. We use the Fourier collocation method [32] to solve eigenvalue problems.

In our numerical simulations, we take

𝐮1=(3212),𝐮2=(3212).{\bf u}_{1}=\begin{pmatrix}\frac{\sqrt{3}}{2}\\ \frac{1}{2}\end{pmatrix},{\qquad}{\bf u}_{2}=\begin{pmatrix}\frac{\sqrt{3}}{2}\\ -\frac{1}{2}\end{pmatrix}.

𝐯1{\bf v}_{1} and 𝐯2{\bf v}_{2} are as in (2.2). 𝐤1{\bf k}_{1} and 𝐤2{\bf k}_{2} are dual periods of 𝐮1{\bf u}_{1} and 𝐮2{\bf u}_{2}. 𝐪1{\bf q}_{1} and 𝐪2{\bf q}_{2} are dual periods of 𝐯1{\bf v}_{1} and 𝐯2{\bf v}_{2}.

The first super honeycomb lattice potential is of the form:

V(𝐱)=cos(𝐪1𝐱)+cos(𝐪2𝐱)+cos(𝐪3𝐱).V({\bf x})=\cos({\bf q}_{1}\cdot{\bf x})+\cos({\bf q}_{2}\cdot{\bf x})+\cos({\bf q}_{3}\cdot{\bf x}).

Also we introduce the perturbation, which violates the additional translation symmetry in Definition 2.3:

W(𝐱)=cos(𝐤1𝐱)+cos(𝐤2𝐱)+cos(𝐤3𝐱).W({\bf x})=\cos({\bf k}_{1}\cdot{\bf x})+\cos({\bf k}_{2}\cdot{\bf x})+\cos({\bf k}_{3}\cdot{\bf x}).

We compute the lowest seven bands of Hδ=Δ+V(𝐱)+δW(𝐱)H^{\delta}=-\Delta+V({\bf x})+\delta W({\bf x}) for δ=0.3\delta=-0.3, 0, and 0.30.3. The 2nd7th2^{nd}-7^{th} bands along the 𝐤1{\bf k}_{1} direction are displayed in Figure 5. Apparently, with the additional translation symmetry, a double Dirac cone occurs at the Γ\Gamma point. When the additional translation symmetry is broken, the fourfold degeneracy disappears and a local gap opens between the 3rd3^{rd} and 4th4^{th} bands.

Refer to caption
Refer to caption
Refer to caption
Figure 5: The 2nd2^{nd}-7th7^{th} bands for the operator Hδ(𝐤)=HV(𝐤)+δW(𝐱)H^{\delta}({\bf k})=H_{V}({\bf k})+\delta W({\bf x}) with 𝐤=λ𝐤1{\bf k}=\lambda{\bf k}_{1}, where δ\delta takes the value (a) 0.3-0.3, (b) 0, and (c) 0.30.3.

We construct a piecewise constant potential as follows. In the unit cell Ω\Omega, take

f(𝐱)={1,if|𝐱12(𝐮1+𝐮2)|<0.130,elsef({\bf x})=\begin{cases}1,&if{\quad}|{\bf x}-\frac{1}{2}({\bf u}_{1}+{\bf u}2)|<0.1\\ 30,&else\end{cases} (5.1)

And construct f(𝐱)f({\bf x}) in 2{\mathbb{R}}^{2} by translation along 𝐮1{\bf u}_{1} and 𝐮2{\bf u}_{2}. Let g(𝐱,r)g({\bf x},r) and W(𝐱,r)W({\bf x},r) be in the form of (2.5) and (2.6). We compute the bands of H(r)=Δ+W(𝐱,r)H(r)=-\Delta+W({\bf x},r) for r=1.053r=\frac{1.05}{3}, r=13r=\frac{1}{3}, and r=0.9753r=\frac{0.975}{3} respectively. The potentials W(𝐱,r)W({\bf x},r) are displayed in the top panel of Figure 4. The corresponding bands are displayed in the bottom panel accordingly. We remark that (b) corresponds to the super honeycomb case, i.e, possessing the additional translation symmetry. Apparently, this potential admits a fourfold degeneracy at the Γ\Gamma point and a double Dirac cone in the vicinity, but they disappear for other two cases.

Acknowledgments

We would like to acknowledge the assistance of Borui Miao for interesting discussions and precious suggestions.

References

  • [1] Mark J. Ablowitz and Yi Zhu. Nonlinear waves in shallow honeycomb lattices. SIAM Journal on Applied Mathematics, 72(1):240–260, 2012.
  • [2] Habib Ammari, Bryn Davies, Erik Orvehed Hiltunen, and Sanghyeon Yu. Topologically protected edge modes in one-dimensional chains of subwavelength resonators. Journal de Mathématiques Pures et Appliquées, 144:17–49, 2020.
  • [3] Habib Ammari, Brian Fitzpatrick, Erik Orvehed Hiltunen, Hyundae Lee, and Sanghyeon Yu. Honeycomb-lattice minnaert bubbles. SIAM Journal on Mathematical Analysis, 52(6):5441–5466, 2020.
  • [4] Habib Ammari, Brian Fitzpatrick, Erik Orvehed Hiltunen, and Sanghyeon Yu. Subwavelength localized lodes for acoustic waves in bubbly crystals with a defect. SIAM Journal on Applied Mathematics, 78(6):3316–3335, 2018.
  • [5] Sabyasachi Barik, Hirokazu Miyake, Wade DeGottardi, Edo Waks, and Mohammad Hafezi. Two-dimensionally confined topological edge states in photonic crystals. New Journal of Physics, 18(11):113013, 2016.
  • [6] Gregory Berkolaiko and Andrew Comech. Symmetry and dirac points in graphene spectrum. Journal of Spectral Theory, 8, 2014.
  • [7] Markus Büttiker. Edge-state physics without magnetic fields. Science, 325(5938):278–279, 2009.
  • [8] A. Drouot and M. I. Weinstein. Edge states and the valley hall effect. Advances in Mathematics, 368, 2020.
  • [9] M.S.P. Eastham. The spectral theory of periodic differential equations. Texts in mathematics. Scottish Academic Press, 1973.
  • [10] Toshiaki Enoki and Kazuyuki Takai. The edge state of nanographene and the magnetism of the edge-state spins. Solid State Communications, 149(27):1144–1150, 2019.
  • [11] L.C. Evans. Partial differential equations. Graduate studies in mathematics. American Mathematical Society, 2010.
  • [12] Charles L. Fefferman, James P. Lee-Thorp, and Michael I. Weinstein. Topologically protected states in one-dimensional continuous systems and dirac points. Proceedings of the National Academy of Sciences, 111(24):8759–8763, 2014.
  • [13] Charles L. Fefferman and Michael I. Weinstein. Honeycomb lattice potentials and dirac points. Journal of the American Mathematical Society, 25(4):1169–1220, 2012.
  • [14] Charles L. Fefferman and Michael I. Weinstein. Wave packets in honeycomb structures and two-dimensional dirac equations. Communications in Mathematical Physics, 326(1):251–286, 2013.
  • [15] Hélène Feldner, Zi Yang Meng, Thomas C. Lang, Fakher F. Assaad, Stefan Wessel, and Andreas Honecker. Dynamical signatures of edge-state magnetism on graphene nanoribbons. Physical Review Letters, 106(22):226401, 2011.
  • [16] V. V. Grushin. Multiparameter perturbation theory of fredholm operators applied to bloch functions. Mathematical Notes, 86(5):767–774, 2009.
  • [17] Haimo Guo, Meirong Zhang, and Yi Zhu. Threefold Weyl points for the Schrödinger operator with periodic potentials. SIAM Journal on Mathematical Analysis, 54(3):3654–3695, 2022.
  • [18] V. P. Gusynin, S. G. Sharapov, and J. P. Carbotte. AC conductivity of graphene: from tight-binding model to 2 + 1-dimensional quantum electrodynamics. International Journal of Modern Physics B, 21(27):4611–4658, 2007.
  • [19] F. D. M. Haldane and S. Raghu. Possible realization of directional optical waveguides in photonic crystals with broken time-reversal symmetry. Physical Review Letters, 100(1):013904, 2008.
  • [20] M. Z. Hasan and C. L. Kane. Colloquium: topological insulators. Reviews of Modern Physics, 82(4):3045–3067, 2010.
  • [21] C. L. Kane and E. J. Mele. Z2{Z}_{2} topological order and the quantum spin hall effect. Physical Review Letters, 95(14):146802, 2005.
  • [22] Alexander B. Khanikaev and Gennady Shvets. Two-dimensional topological photonics. Nature Photonics, 11(12):763–773, 2017.
  • [23] J. P. Lee-Thorp, M. I. Weinstein, and Y. Zhu. Elliptic operators with honeycomb symmetry: dirac points, edge states and applications to photonic graphene. Archive for Rational Mechanics and Analysis, 232(1):1–63, 2019.
  • [24] Samuel J. Palmer and Vincenzo Giannini. Berry bands and pseudo-spin of topological photonic phases. Physical Review Research, 3(2):L022013, 2021.
  • [25] Xiao-Liang Qi and Shou-Cheng Zhang. Topological insulators and superconductors. Reviews of Modern Physics, 83(4):1057–1110, 2011.
  • [26] S. Raghu and F. D. M. Haldane. Analogs of quantum-hall-effect edge states in photonic crystals. Physical Review A, 78(3):033834, 2008.
  • [27] S. Reich, J. Maultzsch, C. Thomsen, and P. Ordejón. Tight-binding description of graphene. Physical Review B, 66(3):035412, 2002.
  • [28] E. C. Titchmarsh and T. Teichmann. Eigenfunction expansions associated with second order differential equations part II. Oxford University Press, 1959.
  • [29] P. R. Wallace. The band theory of graphite. Physical Review, 71(9):622–634, 1947.
  • [30] Congjun Wu, B. Andrei Bernevig, and Shou-Cheng Zhang. Helical liquid and the edge of quantum spin hall systems. Physical Review Letters, 96(10):106401, 2006.
  • [31] Long-Hua Wu and Xiao Hu. Scheme for achieving a topological photonic crystal by using dielectric material. Physical Review Letters, 114(22):223901, 2015.
  • [32] Jianke Yang. Nonlinear waves in integrable and nonintegrable systems. Society for Industrial and Applied Mathematics, 2010.
  • [33] Y. Yang, Y. F. Xu, T. Xu, H. X. Wang, J. H. Jiang, X. Hu, and Z. H. Hang. Visualization of a unidirectional electromagnetic waveguide using topological photonic crystals made of dielectric materials. Phys Rev Lett, 120(21):217401, 2018.
  • [34] S. Yves, R. Fleury, T. Berthelot, M. Fink, F. Lemoult, and G. Lerosey. Crystalline metamaterials for topological properties at subwavelength scales. Nat Commun, 8:16023, 2017.