Localized and extended phases in square moiré patterns

C. Madronẽro Instituto de Física, Universidad Nacional Autónoma de México, Apartado Postal 20-364, México City 01000, México    G. A. Domínguez-Castro Instituto de Física, Universidad Nacional Autónoma de México, Apartado Postal 20-364, México City 01000, México Institut für Theoretische Physik, Leibniz Universität Hannover, Appelstr. 2, D-30167 Hannover, Germany    R. Paredes Instituto de Física, Universidad Nacional Autónoma de México, Apartado Postal 20-364, México City 01000, México
(May 1, 2024)
Abstract

Random defects do not constitute the unique source of electron localization in two dimensions. Lattice quasidisorder generated from two inplane superimposed rotated, main and secondary, square lattices, namely monolayers where moiré patterns are formed, leads to a sharp localized to delocalized single-particle transition. This is demostrated here for both, discrete and continuum models of moiré patterns that arise as the twisting angle θ𝜃\thetaitalic_θ between main and secondary lattices is varied in the interval [0,π/4]0𝜋4[0,\pi/4][ 0 , italic_π / 4 ]. Localized to delocalized transition is recognized as the moiré patterns depart from being perfect square crystals to non-crystalline structures. Extended single-particle states were found for rotation angles associated with Pythagorean triples that produce perfectly periodic structures. Conversely, angles not arising from such Pythagorean triples lead to non-commensurate or quasidisordered structures, thus originating localized states. These conclusions are drawn from a stationary analysis where the standard IPR parameter measuring localization allowed us to detect the transition. While both, ground state and excited states were analyzed for the discrete model, where the secondary lattice was considered as a perturbation of the main one, the sharp transition was tracked back for the fundamental state in the continuous scenario where the secondary lattice is not a perturbation any more.

pacs:

I Introduction

Since the achievement of its experimental assembly, rotated graphene bilayers have attracted the attention not just because of the diversity of quantum phases accessible in them, but also by virtue of its fabrication simplicity [1, 2, 3, 4, 5], as well as for the control achievable with respect to other multicomponent materials. The quantum phases that emerge as a result of the angle tuning include the non-conventional superconducting, metal, and Mott insulating states as a function of the carrier density [6], correlated ferromagnetic phases [7, 8, 9], and also the arrive of anomalous optical properties [10, 11]. Localization of charge carriers is other phenomenon that can be explored in stacked layers of graphene [12, 13, 14]. In fact, it is thought that a central mechanism affecting the electron transport in moiré heterostructures is the appearance of flat bands. Among other remarkable properties of electromagnetic character, is the negative magnetoresistance appearing on twisted double bilayer graphene super lattices, as a result of weak localization [15].

Single particle localization induced by disorder, was first explained in the seminal work of Anderson within the single electron theory [16]. Because of its intrinsic nature, namely, resulting from destructive wave interference, the localization phenomenon manifests not just in the degenerate quantum regime [17, 18], but in any scenario in which the wave nature plays a main role [19]. The understanding of the opposite localized and extended states has been addresed in many different ways and schemes, both theoretically and experimentally, incorporating elements proper of the matter constituents. Among these elements it can be recognized, length and dimensionality of the lattice, inter-particle interactions [20, 21, 22, 23], long-range tunneling [24, 25, 26, 27], and of course structural disorder created either, by random potentials or quasicrystalline designs [28, 29, 30]. Because of the fact that the lattice spatial dicretization must play a fundamental role, a natural question is to explore the influence of the recently created moiré patterns on the emergence of localized vs. extended states.

Ultracold bosonic atoms, largely known as quantum simulators, are schemes were the superfluid to Mott insulator transition in twisted-bilayer square lattices based on atomic Bose-Einstein condensates loaded into spin-dependent optical lattices can be tested [31, 32]. Several theoretic proposals difficult to realize with crystals have been planned with cold atoms to observe analogous physics to its condensed matter counterpart [33]. The opportunity of preparing macroscopic clouds of weakly interacting atoms in two dimensions relies basically on two facts, one is the possibility of having a strong confinement along a spatial direction, which in turns creates a 2D Bose-Einstein condensate starting from a 3D cloud [34, 35], and the other, is the chance of varying the atom-atom interaction by tuning externally a magnetic field where the scattering length is nearly zero [36], usually occurring near a Feshbach resonance.

In this investigation we analyze the emergence of localized vs extended states of a single particle in square moiré lattices in two dimensions. For this purpose we consider both, a discrete lattice as well as its continuum counterpart in which a particle is confined in a two-dimensional potential composed by a principal and a secondary square lattices. Moiré patterns come up when the lattices lying one on top of the other are rotated by an angle θ𝜃\thetaitalic_θ. Our analysis comprises the study of stationary localization properties.

This paper is organized as follows. First in section II we present the model for both, continuum and discrete cases. Since the lattice model is used to analyze moiré patterns where the secondary lattice is considered as a perturbation, being such a perturbative term an onsite shift, a brief discusion on Wannier functions is included. In Section III we focus on the analysis of shallow moiré lattices and study both, ground and excited states. Then, in Section IV we concentrate in deep moiré patterns where the main and secondary lattices has the same depth. Finally in section V a summary of our investigation is given.

II Model

Our starting point is the two-dimensional continuum single-particle Hamiltonian:

H^cont=22m2+Vopt(x,y),subscript^𝐻contsuperscriptPlanck-constant-over-2-pi22𝑚superscriptsubscriptperpendicular-to2subscript𝑉opt𝑥𝑦\hat{H}_{\text{cont}}=\frac{-\hbar^{2}}{2m}\nabla_{\perp}^{2}+V_{\text{opt}}(x% ,y),over^ start_ARG italic_H end_ARG start_POSTSUBSCRIPT cont end_POSTSUBSCRIPT = divide start_ARG - roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ∇ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_V start_POSTSUBSCRIPT opt end_POSTSUBSCRIPT ( italic_x , italic_y ) , (1)

2=x2+y2superscriptsubscriptperpendicular-to2superscriptsubscript𝑥2superscriptsubscript𝑦2\nabla_{\perp}^{2}=\partial_{x}^{2}+\partial_{y}^{2}∇ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ∂ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ∂ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT being the Laplacian operator in 2D and m𝑚mitalic_m the particle mass. The optical potential Vopt(x,y)=V1(x,y)+V2(x,y)subscript𝑉opt𝑥𝑦subscript𝑉1𝑥𝑦subscript𝑉2𝑥𝑦V_{\text{opt}}(x,y)=V_{1}(x,y)+V_{2}(x,y)italic_V start_POSTSUBSCRIPT opt end_POSTSUBSCRIPT ( italic_x , italic_y ) = italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_y ) + italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x , italic_y ) is given by the superposition of two square optical lattices with a relative rotation angle of θ𝜃\thetaitalic_θ among them. Throughout the manuscript, we call V1(x,y)subscript𝑉1𝑥𝑦V_{1}(x,y)italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_y ) the primary potential and V2(x,y)subscript𝑉2𝑥𝑦V_{2}(x,y)italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x , italic_y ) the secondary potential, each one given as follows:

V1(x,y)=s1ER(sin2πx1a+sin2πy1a),V2(x,y)=s2ER(sin2πx2a+sin2πy2a),formulae-sequencesubscript𝑉1𝑥𝑦subscript𝑠1subscript𝐸𝑅superscript2𝜋subscript𝑥1𝑎superscript2𝜋subscript𝑦1𝑎subscript𝑉2𝑥𝑦subscript𝑠2subscript𝐸𝑅superscript2𝜋subscript𝑥2𝑎superscript2𝜋subscript𝑦2𝑎\begin{split}V_{1}(x,y)&=s_{1}E_{R}\left(\sin^{2}\frac{\pi x_{1}}{a}+\sin^{2}% \frac{\pi y_{1}}{a}\right),\\ V_{2}(x,y)&=s_{2}E_{R}\left(\sin^{2}\frac{\pi x_{2}}{a}+\sin^{2}\frac{\pi y_{2% }}{a}\right),\end{split}start_ROW start_CELL italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_y ) end_CELL start_CELL = italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG italic_π italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_a end_ARG + roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG italic_π italic_y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_a end_ARG ) , end_CELL end_ROW start_ROW start_CELL italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x , italic_y ) end_CELL start_CELL = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG italic_π italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_a end_ARG + roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG italic_π italic_y start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_a end_ARG ) , end_CELL end_ROW (2)

where a𝑎aitalic_a is the lattice constant, ER=2π2/2ma2subscript𝐸𝑅superscriptPlanck-constant-over-2-pi2superscript𝜋22𝑚superscript𝑎2E_{R}=\hbar^{2}\pi^{2}/2ma^{2}italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT = roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_m italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the recoil energy, and s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and s2subscript𝑠2s_{2}italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the depths of the optical potentials. In terms of an unrotated frame of reference (see Fig. 1a), the coordinates (x1,y1)subscript𝑥1subscript𝑦1(x_{1},y_{1})( italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) and (x2,y2)subscript𝑥2subscript𝑦2(x_{2},y_{2})( italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_y start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) are given by:

x1=xcosθ/2ysinθ/2,y1=xsinθ/2+ycosθ/2,x2=xcosθ/2+ysinθ/2,y2=xsinθ/2+ycosθ/2.formulae-sequencesubscript𝑥1𝑥𝜃2𝑦𝜃2formulae-sequencesubscript𝑦1𝑥𝜃2𝑦𝜃2formulae-sequencesubscript𝑥2𝑥𝜃2𝑦𝜃2subscript𝑦2𝑥𝜃2𝑦𝜃2\begin{split}x_{1}&=x\cos\theta/2-y\sin\theta/2,\\ y_{1}&=x\sin\theta/2+y\cos\theta/2,\\ x_{2}&=x\cos\theta/2+y\sin\theta/2,\\ y_{2}&=-x\sin\theta/2+y\cos\theta/2.\end{split}start_ROW start_CELL italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL = italic_x roman_cos italic_θ / 2 - italic_y roman_sin italic_θ / 2 , end_CELL end_ROW start_ROW start_CELL italic_y start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL = italic_x roman_sin italic_θ / 2 + italic_y roman_cos italic_θ / 2 , end_CELL end_ROW start_ROW start_CELL italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL start_CELL = italic_x roman_cos italic_θ / 2 + italic_y roman_sin italic_θ / 2 , end_CELL end_ROW start_ROW start_CELL italic_y start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL start_CELL = - italic_x roman_sin italic_θ / 2 + italic_y roman_cos italic_θ / 2 . end_CELL end_ROW (3)

Depending on the value of θ𝜃\thetaitalic_θ, the resulting optical potential Vopt(x,y)subscript𝑉opt𝑥𝑦V_{\text{opt}}(x,y)italic_V start_POSTSUBSCRIPT opt end_POSTSUBSCRIPT ( italic_x , italic_y ) gives rise to periodic (commensurable) or aperiodic (incommensurable) structures. Crucially, these structures, called moiré lattices, always feature the rotational symmetry of the underlying sublattices except for θ=π/4𝜃𝜋4\theta=\pi/4italic_θ = italic_π / 4, where a quasicrystal with octagonal rotation symmetry is obtained. Commensurable moiré patterns result from angles that come from a Pythagorean triple, that is, cosθ=m/n𝜃𝑚𝑛\cos\theta=m/nroman_cos italic_θ = italic_m / italic_n, sinθ=b/c𝜃𝑏𝑐\sin\theta=b/croman_sin italic_θ = italic_b / italic_c and m2+n2=l2superscript𝑚2superscript𝑛2superscript𝑙2m^{2}+n^{2}=l^{2}italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT with m,n𝑚𝑛m,nitalic_m , italic_n and l𝑙litalic_l integers. For all other rotation angles, Vopt(x,y)subscript𝑉opt𝑥𝑦V_{\text{opt}}(x,y)italic_V start_POSTSUBSCRIPT opt end_POSTSUBSCRIPT ( italic_x , italic_y ) leads to an incommensurable but not necessarily a disordered random lattice. In Figs. 1(b)-(f), we illustrate the resulting moiré patterns for several rotation angles.

Refer to caption
Figure 1: (a) Schematic representation of two square optical lattices (blue and red) with a relative rotation angle of θ𝜃\thetaitalic_θ among them. Black lines correspond to the axes of an unrotated frame of reference. (b)-(f) Moiré patterns for several twist angles.

Formation of moiré patterns as a function of rotation angle gives rise to a particular spatial distribution of the local energy minima. Thus, the localized Wanneir functions, fundamental in the tight-binding approximation for nearest- or next nearest- neighbors, must be adapted for moiré lattices. For instance, such a picture modifies the inherent physics of band structure. Also, the transport across the lattice must be modified by the presence of the structured patterns.

II.1 Review of basic theory for Wannier functions

With the purpose of establishing the discrete version of Hamiltonian (1), in this section we briefly review the basic concepts of a single particle subject to a periodic potential. In particular, we focus on the generation of localized Wannier functions (WFs) and their connection in the construction of a tight-binding description of a periodic Hamiltonian. A thorough discussion can be found in any standard solid state reference [37, 38, 39]. For simplicity, we consider a one-dimensional potential, and at the end of this section, we discuss the generalization to higher dimensions. The quantum problem of a particle of mass m𝑚mitalic_m in a periodic potential V(x+a)=V(x)𝑉𝑥𝑎𝑉𝑥V(x+a)=V(x)italic_V ( italic_x + italic_a ) = italic_V ( italic_x ) with periodicity a𝑎aitalic_a is described by the following Schrödinger equation:

[22md2dx2+V(x)]ψp(n)(x)=εp(n)ψp(n)(x),delimited-[]superscriptPlanck-constant-over-2-pi22𝑚superscript𝑑2𝑑superscript𝑥2𝑉𝑥superscriptsubscript𝜓𝑝𝑛𝑥superscriptsubscript𝜀𝑝𝑛superscriptsubscript𝜓𝑝𝑛𝑥\left[-\frac{\hbar^{2}}{2m}\frac{d^{2}}{dx^{2}}+V(x)\right]\psi_{p}^{(n)}(x)=% \varepsilon_{p}^{(n)}\psi_{p}^{(n)}(x),[ - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_V ( italic_x ) ] italic_ψ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ) = italic_ε start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ) , (4)

where n𝑛nitalic_n labels the band number and p[π/a,π/a]𝑝𝜋𝑎𝜋𝑎p\in[-\pi/a,\pi/a]italic_p ∈ [ - italic_π / italic_a , italic_π / italic_a ] is the quasi-momentum which is restricted to the first Brillouin zone. According to Bloch’s theorem, the solutions of the above equation can be written as the product of a plane wave times a periodic function with the same periodicity of the potential, ψp(n)(x)=eipxup(n)(x)superscriptsubscript𝜓𝑝𝑛𝑥superscript𝑒𝑖𝑝𝑥superscriptsubscript𝑢𝑝𝑛𝑥\psi_{p}^{(n)}(x)=e^{ipx}u_{p}^{(n)}(x)italic_ψ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ) = italic_e start_POSTSUPERSCRIPT italic_i italic_p italic_x end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ). Introducing this ansatz into Eq. (4) yields to the Schrödinger equation for up(n)(x)superscriptsubscript𝑢𝑝𝑛𝑥u_{p}^{(n)}(x)italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ):

[22m(iddx+p)2+V(x)]up(n)(x)=εp(n)up(n)(x).delimited-[]superscriptPlanck-constant-over-2-pi22𝑚superscript𝑖𝑑𝑑𝑥𝑝2𝑉𝑥superscriptsubscript𝑢𝑝𝑛𝑥superscriptsubscript𝜀𝑝𝑛superscriptsubscript𝑢𝑝𝑛𝑥\begin{split}\left[\frac{\hbar^{2}}{2m}\left(-i\frac{d}{dx}+p\right)^{2}+V(x)% \right]&u_{p}^{(n)}(x)=\\ &\varepsilon_{p}^{(n)}u_{p}^{(n)}(x).\end{split}start_ROW start_CELL [ divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ( - italic_i divide start_ARG italic_d end_ARG start_ARG italic_d italic_x end_ARG + italic_p ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_V ( italic_x ) ] end_CELL start_CELL italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ) = end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL italic_ε start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ) . end_CELL end_ROW (5)

The above equation can be seen as a set of eigenvalue problems, one for each p𝑝pitalic_p, with an infinite number of solutions. The collection of all eigenvalues gives rise to the corresponding band structure. Because Bloch functions extend over the entire lattice, they are not helpful for constructing a lattice Hamiltonian. Nevertheless, a convenient alternative is to use the so-called WFs. In terms of the Bloch functions, the Wannier functions are given as follows:

wn(xxi)=1Lpeipxiψp(n)(x),subscript𝑤𝑛𝑥subscript𝑥𝑖1𝐿subscript𝑝superscript𝑒𝑖𝑝subscript𝑥𝑖superscriptsubscript𝜓𝑝𝑛𝑥w_{n}(x-x_{i})=\frac{1}{\sqrt{L}}\sum_{p}e^{-ipx_{i}}\psi_{p}^{(n)}(x),italic_w start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x - italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_L end_ARG end_ARG ∑ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_i italic_p italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT ( italic_x ) , (6)

where xi=iasubscript𝑥𝑖𝑖𝑎x_{i}=iaitalic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_i italic_a is the position of the ilimit-from𝑖i-italic_i -th lattice site and L𝐿Litalic_L is the number of lattice sites. A typical feature of WFs is their relatively strong concentration around each lattice minimum. Thus, they provide an attractive option for building a lattice version of a periodic Hamiltonian. Additionally, as it is easy to show, the collection of Wannier functions form an orthonormal set

𝑑xwn(xxi)wm(xxj)=δn,mδi,jdifferential-d𝑥superscriptsubscript𝑤𝑛𝑥subscript𝑥𝑖subscript𝑤𝑚𝑥subscript𝑥𝑗subscript𝛿𝑛𝑚subscript𝛿𝑖𝑗\int dx\ w_{n}^{*}(x-x_{i})w_{m}(x-x_{j})=\delta_{n,m}\delta_{i,j}∫ italic_d italic_x italic_w start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_x - italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_w start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_x - italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = italic_δ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT (7)

An important subtlety of Eq. (6) is the presence of a gauge freedom that exists in the definition of Bloch functions. That is, one can replace ψ~p(n)eiφpψp(n)superscriptsubscript~𝜓𝑝𝑛superscript𝑒𝑖subscript𝜑𝑝superscriptsubscript𝜓𝑝𝑛\tilde{\psi}_{p}^{(n)}\rightarrow e^{i\varphi_{p}}\psi_{p}^{(n)}over~ start_ARG italic_ψ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT → italic_e start_POSTSUPERSCRIPT italic_i italic_φ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_n ) end_POSTSUPERSCRIPT without modifying the band structure. Nevertheless, such a gauge transformation will change the spatial behavior of the associated Wannier function. In one-dimension, one can always choose the phase of the Bloch waves in such a way that the corresponding WFs are real, have a well defined parity, and decay exponentially away from the central site (see Fig. 2(a)). In two and three dimensions this is generally not possible [40]. However, for separable potentials, the corresponding Wannier function is simply the product of the one-dimensional Wannier functions associated with each direction (see Fig. 2(b)). It is important to mention that even for non-separable potentials, WFs can be generated using more elaborate procedures. Recently, Wannier functions have been generated in a quasicrystal structure, where the Bloch theorem is no longer valid [41]. In section III Wannier functions will be a key element to analyze the effect of lattice localization.

Refer to caption
Figure 2: (a) Exponential decay of a one-dimensional Wannier function, notice the log scale in the y-axis. (b) Two-dimensional Wannier function w(x,y)=w(x)w(y)𝑤𝑥𝑦𝑤𝑥𝑤𝑦w(x,y)=w(x)w(y)italic_w ( italic_x , italic_y ) = italic_w ( italic_x ) italic_w ( italic_y ) associated with a square lattice.

III Shallow moiré lattices

In this section, we study the case in which the secondary potential is weak in comparison with the primary potential, i.e. s2s1much-less-thansubscript𝑠2subscript𝑠1s_{2}\ll s_{1}italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≪ italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. For the sake of simplicity, we consider the secondary lattice to be the only one that is rotated by an angle θ𝜃\thetaitalic_θ. Since s2s1much-less-thansubscript𝑠2subscript𝑠1s_{2}\ll s_{1}italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≪ italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, the minima of the main lattice are not considerably affected by the presence of the rotated lattice. Hence, we can safely use the associated Wannier functions w(𝐫)𝑤𝐫w(\mathbf{r})italic_w ( bold_r ) of the primary lattice to build a discrete version of the Hamiltonian in Eq. (1). In this scenario, the lowest-band lattice Hamiltonian reads

H^=𝐢,𝐣J𝐢𝐣(b^𝐢b^𝐣+h.c)+𝐢ϵ𝐢n^𝐢,\hat{H}=-\sum_{\mathbf{i},\mathbf{j}}J_{\mathbf{i}\mathbf{j}}(\hat{b}_{\mathbf% {i}}^{\dagger}\hat{b}_{\mathbf{j}}+h.c)+\sum_{\mathbf{i}}\epsilon_{\mathbf{i}}% \hat{n}_{\mathbf{i}},over^ start_ARG italic_H end_ARG = - ∑ start_POSTSUBSCRIPT bold_i , bold_j end_POSTSUBSCRIPT italic_J start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT ( over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_b end_ARG start_POSTSUBSCRIPT bold_j end_POSTSUBSCRIPT + italic_h . italic_c ) + ∑ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT , (8)

where 𝐢=(ix,iy)𝐢subscript𝑖𝑥subscript𝑖𝑦\mathbf{i}=(i_{x},i_{y})bold_i = ( italic_i start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT ) stands for a two-dimensional space coordinate in which ix,iysubscript𝑖𝑥subscript𝑖𝑦i_{x},i_{y}italic_i start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_i start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT run along the positions in a given two-dimensional lattice of size Nsites=L×Lsubscript𝑁𝑠𝑖𝑡𝑒𝑠𝐿𝐿N_{sites}=L\times Litalic_N start_POSTSUBSCRIPT italic_s italic_i italic_t italic_e italic_s end_POSTSUBSCRIPT = italic_L × italic_L and J𝐢𝐣subscript𝐽𝐢𝐣J_{\mathbf{i}\mathbf{j}}italic_J start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT is the usual hop** amplitude

J𝐢𝐣=d2rw(𝐫𝐫i)(22m2+V1(𝐫))w(𝐫𝐫j)subscript𝐽𝐢𝐣superscript𝑑2𝑟superscript𝑤𝐫subscript𝐫𝑖superscriptPlanck-constant-over-2-pi22𝑚subscriptsuperscript2perpendicular-tosubscript𝑉1𝐫𝑤𝐫subscript𝐫𝑗J_{\mathbf{i}\mathbf{j}}=\int d^{2}r\ w^{*}(\mathbf{r}-\mathbf{r}_{i})\left(-% \frac{\hbar^{2}}{2m}\nabla^{2}_{\perp}+V_{1}(\mathbf{r})\right)w(\mathbf{r}-% \mathbf{r}_{j})italic_J start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT = ∫ italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_r - bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ( - divide start_ARG roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m end_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT + italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_r ) ) italic_w ( bold_r - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) (9)

The rotated potential V2(𝐫)subscript𝑉2𝐫V_{2}(\mathbf{r})italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( bold_r ) can be considered as a site-dependent energy term ϵ𝐢𝐣subscriptitalic-ϵ𝐢𝐣\epsilon_{\mathbf{i}\mathbf{j}}italic_ϵ start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT:

ϵ𝐢𝐣=s2ERd2rw(𝐫𝐫i)V2(𝐫)w(𝐫𝐫j).subscriptitalic-ϵ𝐢𝐣subscript𝑠2subscript𝐸𝑅superscript𝑑2𝑟superscript𝑤𝐫subscript𝐫𝑖subscript𝑉2𝐫𝑤𝐫subscript𝐫𝑗\epsilon_{\mathbf{i}\mathbf{j}}=s_{2}E_{R}\int d^{2}r\ w^{*}(\mathbf{r}-% \mathbf{r}_{i})V_{\text{2}}(\mathbf{r})w(\mathbf{r}-\mathbf{r}_{j}).italic_ϵ start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ∫ italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_r italic_w start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( bold_r - bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( bold_r ) italic_w ( bold_r - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) . (10)

For deep enough lattices, one can rely on the so-called tight-binding approximation. In this limit, due to the exponential decay of the Wannier functions, the hop** amplitudes J𝐢𝐣subscript𝐽𝐢𝐣J_{\mathbf{i}\mathbf{j}}italic_J start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT strongly decay with the distance. Therefore, we can ignore the tunneling terms beyond the nearest neighbors. Furthermore, due to the translational symmetry of the primary lattice, the hop** parameter becomes a constant J𝐽Jitalic_J independent of the lattice site. Analogously, the leading contribution of ϵ𝐢𝐣subscriptitalic-ϵ𝐢𝐣\epsilon_{\mathbf{i}\mathbf{j}}italic_ϵ start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT is the 𝐢=𝐣𝐢𝐣\mathbf{i}=\mathbf{j}bold_i = bold_j term, which corresponds to an on-site energy shift ϵ𝐢𝐣=δ𝐢,𝐣ϵ𝐢subscriptitalic-ϵ𝐢𝐣subscript𝛿𝐢𝐣subscriptitalic-ϵ𝐢\epsilon_{\mathbf{i}\mathbf{j}}=\delta_{\mathbf{i},\mathbf{j}}\ \epsilon_{% \mathbf{i}}italic_ϵ start_POSTSUBSCRIPT bold_ij end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT bold_i , bold_j end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT. After some straightforward algebra, one can find the following expression for the on-site term:

ϵ𝐢=Δ[cos2π(ixcosθ+iysinθ)+cos2π(ixsinθ+iycosθ)],subscriptitalic-ϵ𝐢Δdelimited-[]2𝜋subscript𝑖𝑥𝜃subscript𝑖𝑦𝜃2𝜋subscript𝑖𝑥𝜃subscript𝑖𝑦𝜃\begin{split}\epsilon_{\mathbf{i}}&=\Delta[\cos 2\pi(i_{x}\cos\theta+i_{y}\sin% \theta)+\\ &\cos 2\pi(-i_{x}\sin\theta+i_{y}\cos\theta)],\end{split}start_ROW start_CELL italic_ϵ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT end_CELL start_CELL = roman_Δ [ roman_cos 2 italic_π ( italic_i start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_cos italic_θ + italic_i start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT roman_sin italic_θ ) + end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL roman_cos 2 italic_π ( - italic_i start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT roman_sin italic_θ + italic_i start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT roman_cos italic_θ ) ] , end_CELL end_ROW (11)

where ΔΔ\Deltaroman_Δ is the amplitude of the on-site potential and is given as follows:

Δ=s2ER2dxdy|w(x,y)|2×cos(2πxasinθ)cos(2πyacosθ).Δsubscript𝑠2subscript𝐸𝑅2𝑑𝑥𝑑𝑦superscript𝑤𝑥𝑦22𝜋𝑥𝑎𝜃2𝜋𝑦𝑎𝜃\begin{split}\Delta&=-\frac{s_{2}E_{R}}{2}\int dxdy\ |w(x,y)|^{2}\times\\ &\cos\left(\frac{2\pi x}{a}\sin\theta\right)\cos\left(\frac{2\pi y}{a}\cos% \theta\right).\end{split}start_ROW start_CELL roman_Δ end_CELL start_CELL = - divide start_ARG italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ∫ italic_d italic_x italic_d italic_y | italic_w ( italic_x , italic_y ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT × end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL roman_cos ( divide start_ARG 2 italic_π italic_x end_ARG start_ARG italic_a end_ARG roman_sin italic_θ ) roman_cos ( divide start_ARG 2 italic_π italic_y end_ARG start_ARG italic_a end_ARG roman_cos italic_θ ) . end_CELL end_ROW (12)
Refer to caption
Figure 3: (a) Absolute value of the on-site potential vs the primary potential depth s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for fixed s2=0.2subscript𝑠20.2s_{2}=0.2italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.2. Panels (b)-(d) illustrate the on-site potential in Eq. (11) as a function of the lattice coordinates for different twist angles. In all panels we fix |Δ|/J=1Δ𝐽1|\Delta|/J=1| roman_Δ | / italic_J = 1.

In Fig. 3(a), we plot the parameter Δ/JΔ𝐽\Delta/Jroman_Δ / italic_J as a function of the primary potential depth s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for two different twist angles. We consider a fixed value of s2=0.2subscript𝑠20.2s_{2}=0.2italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.2 of the depth of the secondary potential. Notice that Δ/JΔ𝐽\Delta/Jroman_Δ / italic_J can be tuned to a wide range of values without changing the depth of the secondary lattice. Furthermore, the amplitude of the on-site potential depends very weakly on the angle of rotation. It is important to mention that the considered depths of the optical potentials are easily achievable in current experiments. Figs. 3(b)-3(c) illustrate the on-site potential ϵ𝐢subscriptitalic-ϵ𝐢\epsilon_{\mathbf{i}}italic_ϵ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT as a function of the lattice coordinates for different twist angles.

Having established the Hamiltonian that takes into account the secondary lattice as a perturbation term, we investigate the ground state localization properties of it. A customary parameter that is used as a measure of localization is the inverse participation ratio (IPR), given a normalized wave function ψ𝜓\psiitalic_ψ its IPR is defined as IPR=i=1Nsites|ψ(i)|4IPRsuperscriptsubscript𝑖1subscript𝑁𝑠𝑖𝑡𝑒𝑠superscript𝜓𝑖4\text{IPR}=\sum_{i=1}^{N_{sites}}|\psi(i)|^{4}IPR = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_s italic_i italic_t italic_e italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT | italic_ψ ( italic_i ) | start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT. For extended states, the inverse participation ratio vanishes in the thermodynamic limit as IPRNsites1proportional-toIPRsuperscriptsubscript𝑁𝑠𝑖𝑡𝑒𝑠1\text{IPR}\propto N_{sites}^{-1}IPR ∝ italic_N start_POSTSUBSCRIPT italic_s italic_i italic_t italic_e italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, whereas for localized profiles is always finite. In Fig. 4(a), we plot the IPR associated with the ground state as a function of the twist angle θ𝜃\thetaitalic_θ and the potential strength Δ/JΔ𝐽\Delta/Jroman_Δ / italic_J. As shown in Fig. 4(a), there is a sharp localized-delocalized transition (LDT) at Δc/J2similar-to-or-equalssubscriptΔ𝑐𝐽2\Delta_{c}/J\simeq 2roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_J ≃ 2. Below this threshold, the ground states are extended regardless of the angle of rotation. In contrast, for Δ/J>Δc/JΔ𝐽subscriptΔ𝑐𝐽\Delta/J>\Delta_{c}/Jroman_Δ / italic_J > roman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_J, the spatial behavior of the the ground state becomes angle dependent. In particular, for angles given by a Pythagorean triple, identified here and henceforth as θPsubscript𝜃𝑃\theta_{P}italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, the fundamental mode is extended. In the next section we shall go back to this point.

To show the accuracy of the lattice Hamiltonian (8), in Fig. 4(b), we illustrate the IPR associated with the ground state of the continuum Hamiltonian in Eq. (1) as a function of the twist angle θ𝜃\thetaitalic_θ and the depth of the primary potential s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. The depth of the secondary potential is considered the same as for the calculations in the lattice Hamiltonian, that is we consider s2=0.2subscript𝑠20.2s_{2}=0.2italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.2. Details about the numerical calculations in the continuum model can be found in Appendix A. As one can notice, the results of the discrete and continuum models are in reasonable qualitative agreement, extended vs localized states appear for the same rotation angles, and definite critical values Δc/JsubscriptΔ𝑐𝐽\Delta_{c}/Jroman_Δ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_J and s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT separates the opposite phases above θπ/16greater-than-or-equivalent-to𝜃𝜋16\theta\gtrsim\pi/16italic_θ ≳ italic_π / 16. The quantitative difference between both calculations is a consequence of the different values that the IPR can take in a lattice and in continuous space, see Appendix A.

Refer to caption
Figure 4: (a) IPR associated with the ground state of the lattice Hamiltonian in Eq. (8) as a function of the rotation angle θ𝜃\thetaitalic_θ and the amplitude Δ/JΔ𝐽\Delta/Jroman_Δ / italic_J. (b) Inverse participation ratio associated with the ground state of the continuum Hamiltonian in Eq. (1) as a function of the twist angle and the depth of the primary lattice. Both panels consider s2=0.2subscript𝑠20.2s_{2}=0.2italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.2.

The building of the lattice Hamiltonian allows not only to study the localization properties of the ground state but also of the entire spectrum. In Fig. 5, we plot in a color scheme the inverse participation ratio as a function of the rotation angle θ𝜃\thetaitalic_θ and position in the spectrum. As one can notice, the eigenstates display a rich localization structure. In particular, localized eigenstates start to appear at the edges of the spectrum leaving half of the spectrum with extended states.

Refer to caption
Figure 5: IPR as a function of the rotation angle θ𝜃\thetaitalic_θ for the spectrum associated with Hamiltonian (8).

IV Deep moiré lattices

In this section, our focus is on the case where the potential amplitude of the secondary lattice is comparable or larger than that of the main lattice. In this particular scenario, the secondary potential ceases to be a mere perturbation, leading to a significant impact on the minima of the principal potential. Consequently, the Wannier functions associated with the principal potential are no longer suitable for constructing a lattice Hamiltonian.

For the analysis we solve the continuous Schrödinger equation for Hamiltonian (1) considering lattices having Nsites=50×50subscript𝑁𝑠𝑖𝑡𝑒𝑠5050N_{sites}=50\times 50italic_N start_POSTSUBSCRIPT italic_s italic_i italic_t italic_e italic_s end_POSTSUBSCRIPT = 50 × 50. As in previous section, we investigate the IPR as a function of the rotation angle for the ground state. In figure 6 we plot in a density color scheme the IPR as a function of both, the rotation angle and the lattice depth s1=s2subscript𝑠1subscript𝑠2s_{1}=s_{2}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. As can be apreciated from this figure, the behavior is similar, qualitatively, to that found for shallow moiré lattices analyzed in the previous section. Starting from a certain angle, a sharp LDT for the ground state again occurs at a given potential amplitude. Instead of finding the transition at s18subscript𝑠18s_{1}\approx 8italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 8 as in the shallow moiré lattice case, one observes that the LDT appears at s12.0subscript𝑠12.0s_{1}\approx 2.0italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 2.0. The behavior of the IPR for small angles near the LDT, say θπ/16less-than-or-similar-to𝜃𝜋16\theta\lesssim\pi/16italic_θ ≲ italic_π / 16, signals the presence of an extended ground state disregarding the twisting angle (see dark blue region). The dark blue region (for small rotation angles) suggesting extended states nearly above s18subscript𝑠18s_{1}\approx 8italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 8 for shallow moiré lattices, and s12.0subscript𝑠12.0s_{1}\approx 2.0italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 2.0 for deep moiré lattices, is, as we argue below, a result of the impossibility of exploring moiré lattices arising from small rotation angles. In other words, limitations imposed by numerical calculations for small twisting angles prevent us to have reliable predictions. Certainly, as the potential depth grows, it is expected that the ground state be a localized one because tunneling across sites must suffer a reduction.

Refer to caption
Figure 6: IPR as a function of the rotation angle θ𝜃\thetaitalic_θ for the ground state of Hamiltonian (1). The values of the potential depths in main and secondary lattices are equal (see Eq. (2)), s1=s2subscript𝑠1subscript𝑠2s_{1}=s_{2}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

As described previously, rotation angles associated with Pythagorean triples θPsubscript𝜃𝑃\theta_{P}italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT produce crystalline moiré patterns, being the lattice constant of those periodic arrays given by aMC=a/sinθsubscript𝑎𝑀𝐶𝑎𝜃a_{MC}=a/\sin{\theta}italic_a start_POSTSUBSCRIPT italic_M italic_C end_POSTSUBSCRIPT = italic_a / roman_sin italic_θ [42, 43], where a𝑎aitalic_a is the lattice constant of the main square lattice. Since the lattice constant grows as the rotation angle becomes smaller, a larger number of sites is needed to allow an appropriate exploration of the lattice space, and thus determine if the fundamental state is localized or extended. We analyzed lattices having a larger number of sites, as well as considering a smallest grid for the rotation angle between main and secondary lattices. In figure 7 we plot the IPR for θ<π/8𝜃𝜋8\theta<\pi/8italic_θ < italic_π / 8 for 2D lattices having Nsites=110×110subscript𝑁𝑠𝑖𝑡𝑒𝑠110110N_{sites}=110\times 110italic_N start_POSTSUBSCRIPT italic_s italic_i italic_t italic_e italic_s end_POSTSUBSCRIPT = 110 × 110 sites. As can be seen from this figure, several new sharp regions appear, indicating extended states not advised previously. These extended states correspond to other Pythagorean triads as described above. It appears in figure 7 that the LDT occurs at s1=s21.75subscript𝑠1subscript𝑠21.75s_{1}=s_{2}\approx 1.75italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≈ 1.75, but such an appearance must be attributed to the color scheme, as can be appreciated from the color bar. In figure 8 we illustrate the LDT for as a function of the potential depth s2subscript𝑠2s_{2}italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The points correspond to the value of s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT at which the ground state becomes localized, such a critical value was chosen as the point at which the IPR is no longer zero. The blue line corresponds to an exponential fit s1c=AeBs2+Csuperscriptsubscript𝑠1𝑐𝐴superscript𝑒𝐵subscript𝑠2𝐶s_{1}^{c}=Ae^{-Bs_{2}}+Citalic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT = italic_A italic_e start_POSTSUPERSCRIPT - italic_B italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT + italic_C [44]. At the inset of this figure we include several IPR curves associated with given values of s2subscript𝑠2s_{2}italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT indicated in colors.

Refer to caption
Figure 7: IPR as a function of the rotation angle 0<θ<π/80𝜃𝜋80<\theta<\pi/80 < italic_θ < italic_π / 8 for the ground state of Hamiltonian (1). The values of the potential depths in main and secondary lattices are equal (see Eq. (2)), s1=s2subscript𝑠1subscript𝑠2s_{1}=s_{2}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. The lattice has Nsites=110×110subscript𝑁𝑠𝑖𝑡𝑒𝑠110110N_{sites}=110\times 110italic_N start_POSTSUBSCRIPT italic_s italic_i italic_t italic_e italic_s end_POSTSUBSCRIPT = 110 × 110 sites.
Refer to caption
Figure 8: In (a) black dots are the critical value of the amplitude s1csuperscriptsubscript𝑠1𝑐s_{1}^{c}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT that signals the LDT vs. the amplitude of the secondary lattice amplitude appears. Blue curve in this inset fits the black dots. (b) IPR as a function of potential depth s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for several laticce sizes is plotted. The lattice size used is Nsites=50×50subscript𝑁𝑠𝑖𝑡𝑒𝑠5050N_{sites}=50\times 50italic_N start_POSTSUBSCRIPT italic_s italic_i italic_t italic_e italic_s end_POSTSUBSCRIPT = 50 × 50.

The reason behind the appearance of extended states for Pythagorean triples, namely crystalline moiré structures, is a direct consequence of its periodicity. As stated by the Bloch theorem, all the single-particle eigenstates are extended, particularly the ground state. Thus one reaches the conclusion that all of the moiré lattices coming from rotation angles associated with Phytagorean triples lead to extended states. It is worth to mention here that for θ[0,π/4)𝜃0𝜋4\theta\in[0,\pi/4)italic_θ ∈ [ 0 , italic_π / 4 ) one can find an infinite number of Phytagorean triples [45]. The Pythagorean angles appearing in figures 4 and 6 are those captured by our numerical grid. Conversely, deviation of commensurate structures, leads to obtain localized states above a certain treshold value of s1subscript𝑠1s_{1}italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. This result is reminiscent of the findings exhibited by the Aubry-André model for quasicrystalline lattices, which establishes that above a critical disorder strenght all the states are localized, and in particular the ground state [46]. This behavior can be understood from the potential landscape where the single-particle moves. As shown in figure 9, non-crystalline structures are such that consequtive potential minima are quite appart with respect to those associated with perfectly commensurate lattices (see figures 9(c) and 9(d)). In summary, a moiré lattice that departs from a perfect crystal, and becomes a non periodic structure, shows in good agreement with the seminal result of Anderson, localized states starting from a certain threshold. At this point one can mention that our quasidisordered moiré lattices do not show critical states as those exhibited in the 2D generalized Aubry-André model [47].

Refer to caption
Figure 9: Contour curves of potential in Eq. (2) for a) θP(1)=arctan3/4superscriptsubscript𝜃𝑃134\theta_{P}^{(1)}=\arctan 3/4italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = roman_arctan 3 / 4, (b) θ=arctan3/4+0.01𝜃340.01\theta=\arctan 3/4+0.01italic_θ = roman_arctan 3 / 4 + 0.01, (c) θP(8)=arctan33/56superscriptsubscript𝜃𝑃83356\theta_{P}^{(8)}=\arctan 33/56italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 8 ) end_POSTSUPERSCRIPT = roman_arctan 33 / 56, and (d) θ=arctan33/56+0.01𝜃33560.01\theta=\arctan 33/56+0.01italic_θ = roman_arctan 33 / 56 + 0.01. The values of the potential depths are the same for each panel s1=s2=1.5subscript𝑠1subscript𝑠21.5s_{1}=s_{2}=1.5italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_s start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 1.5

Not less important is the particular feature exhibited by certain twisting angles generating crystalline moiré lattices. As can be seen from figures 4 and 6, the white line indicating θP(1)=arctan3/4superscriptsubscript𝜃𝑃134\theta_{P}^{(1)}=\arctan 3/4italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = roman_arctan 3 / 4 is the center of a wider region signaling extended states, in comparison with other angles associated to different Pythagorean triads, as θP(2)=arctan5/12superscriptsubscript𝜃𝑃2512\theta_{P}^{(2)}=\arctan 5/12italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT = roman_arctan 5 / 12 and θP(3)=arctan8/15superscriptsubscript𝜃𝑃3815\theta_{P}^{(3)}=\arctan 8/15italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 3 ) end_POSTSUPERSCRIPT = roman_arctan 8 / 15, for which the region of extended states is narrower. The explanation is the following. The minima of moiré crystals, that is the spatial regions where the potential Vopt(x,y)subscript𝑉opt𝑥𝑦V_{\text{opt}(x,y)}italic_V start_POSTSUBSCRIPT opt ( italic_x , italic_y ) end_POSTSUBSCRIPT takes its minumum values, are separated among them by a distance determined by the Pythagorean triad that defines a given structure. Such a separation is the hypotenuse of the right triangle m2+n2=l2superscript𝑚2superscript𝑛2superscript𝑙2m^{2}+n^{2}=l^{2}italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, being m𝑚mitalic_m, n𝑛nitalic_n and l𝑙litalic_l integer multiples of the lattice constant a𝑎aitalic_a. Some examples of these Pythagorean triples and its associated angles are θP(1)(3,4,5)superscriptsubscript𝜃𝑃1345\theta_{P}^{(1)}\to(3,4,5)italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT → ( 3 , 4 , 5 ), θP(2)(5,12,13)superscriptsubscript𝜃𝑃251213\theta_{P}^{(2)}\to(5,12,13)italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT → ( 5 , 12 , 13 ), θP(2)(8,15,17)superscriptsubscript𝜃𝑃281517\theta_{P}^{(2)}\to(8,15,17)italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT → ( 8 , 15 , 17 ), and θP(11)(33,56,65)superscriptsubscript𝜃𝑃11335665\theta_{P}^{(11)}\to(33,56,65)italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 11 ) end_POSTSUPERSCRIPT → ( 33 , 56 , 65 ), where the label in parenthesis has been asigned considering the increasing size of the hypotenuse. Let us consider the origin and the rotation center coincide with one of the potential minima associated to a given Pythagorean angle θPsubscript𝜃𝑃\theta_{P}italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT. When a small rotation around θPsubscript𝜃𝑃\theta_{P}italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT is performed, namely θP±δθplus-or-minussubscript𝜃𝑃𝛿𝜃\theta_{P}\pm\delta\thetaitalic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ± italic_δ italic_θ, the nearest potential minima apart each other by a distance ±lδθsimilar-toabsentplus-or-minus𝑙𝛿𝜃\sim\pm l\delta\theta∼ ± italic_l italic_δ italic_θ. Therefore, for the same δθ𝛿𝜃\delta\thetaitalic_δ italic_θ, the minima will be further away as the hypotenuse grows, which breaks the potential periodicity. In the neigborhood of the rotation center appreciable variations of the potential can be appreciated. This is illustrated in a density color scheme in Figure 9. There, we show the contour curves of Vopt(x,y)subscript𝑉opt𝑥𝑦V_{\text{opt}}(x,y)italic_V start_POSTSUBSCRIPT opt end_POSTSUBSCRIPT ( italic_x , italic_y ) of Eq. (2) for a couple of angles θPsubscript𝜃𝑃\theta_{P}italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT. As can be seen from panels in this figure, detectable changes are seen between (c) and (d) corresponding to θP(11)=arctan33/56superscriptsubscript𝜃𝑃113356\theta_{P}^{(11)}=\arctan 33/56italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 11 ) end_POSTSUPERSCRIPT = roman_arctan 33 / 56 and θ=arctan33/56+0.01𝜃33560.01\theta=\arctan 33/56+0.01italic_θ = roman_arctan 33 / 56 + 0.01 respectively, while imperceptible changes appear for the cases θP(1)=arctan3/4superscriptsubscript𝜃𝑃134\theta_{P}^{(1)}=\arctan 3/4italic_θ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = roman_arctan 3 / 4 and θ=arctan3/4+0.01𝜃340.01\theta=\arctan 3/4+0.01italic_θ = roman_arctan 3 / 4 + 0.01 illustrated in figures (a) and (b) respectively.

V Conclusions

This investigation has deal with the identificaction of extended vs. localized phases of a single particle confined in two superimposed square lattices rotated one with respect to the other by an angle in the interval (0,π/4]0𝜋4(0,\pi/4]( 0 , italic_π / 4 ]. The lattices lying in the same plane are such that the amplitude of the secondary lattice can be either, shallow or equal to that of the main lattice. The patterns that result from the superposition, the so called moiré structures, can be classified in two types; crystalline and quasidisorderdered lattices. While the former result from twisting angles associated with the so called Pythagorean triples (satisfying a diphantine equation) and lead to perfect commensurability among main and secondary lattices, the later correspond to arbitrary angles and produce non-commensurate or quasidisordered structures. As described below, the structure plays a crucial role on the localized to delocalized transition.

Localized and delocalized phases were detected from stationary properties, considering two approaches: the discrete or lattice model, and the continuous one. In the case of the lattice model we constructed a Hamiltonian that considers the secondary lattice as an onsite perturbation term, obtained from the Wannier functions of the main lattice. Localized to delocalized transition was tracked from the standard IPR parameter as a function of the rotation angle and the potential depth of main and secondary laticces. Properties of both, fundamental and excited states were investigated for the lattice model. Regarding the continuous model, we concentrate in studying the ground state for both, shallow and deep moiré lattices.

The information provided by the IPR parameter leads us to reach the following conclusions. Extended phases emerge for crystalline moiré structures, while localized ones are identified for non-commensurate structures starting from certanin potential depths. The identification of such a critical potential amplitude, for both shallow and deep moiré lattices, at which a sharp localized-delocalized transitiosn occurs suggest that moiré patterns are the generalization of the one-dimensional Aubry-André model. Finaly, it is worth to mention that the existence of particular twisting Pythagorean angles for which robust extended states emerge, as for instance θ=arctan3/4𝜃34\theta=\arctan 3/4italic_θ = roman_arctan 3 / 4. As the Pythagorean triad associated with a given structure is such that the hypotenuse becomes larger, the narrower becomes the line signaling the extended phase.

Similarly to light propagation phenomena in moiré arrays, one finds that matter confined in moiré patterns offers also new possibilities regarding the transport and localization properties. For instance, moiré heterostructures greatly enhance the appearance of new types of phenomena that are under current investigation in 2D materials.

Acknowledgements

This work was partially funded by Grant No. IN108620 from DGAPA (UNAM). C.J.M.C acknowledges CONACYT scholarship. G.A.D.-C. acknowledge support of the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy –  EXC-2123 QuantumFrontiers  –  390837967.

Appendix A Continuum Calculations

In this appendix, we provide further details on the continuum calculations. For the ground state and determination of IPR we consider the parameters that appear in the following tables.

Name Symbol Value
Number of grid points in the x𝑥xitalic_x direction Nxsubscript𝑁𝑥N_{x}italic_N start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT 512-2048
Number of grid points in the y𝑦yitalic_y direction Nysubscript𝑁𝑦N_{y}italic_N start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT 512-2048
Spatial extension of the numerical grid in the x𝑥xitalic_x direction Lxsubscript𝐿𝑥L_{x}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT 50-200 a𝑎aitalic_a
Spatial extension of the numerical grid in the y𝑦yitalic_y direction Lysubscript𝐿𝑦L_{y}italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT 50-200 a𝑎aitalic_a
Step size used in real time evolution dτ𝑑𝜏d\tauitalic_d italic_τ 0.0050.0050.0050.005
Table 1: Parameters for the numerical simulation
Name Symbol Value
Lattice constant a𝑎aitalic_a 532532532532 nm
Potential depth V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 0-11 ERsubscript𝐸𝑅E_{R}italic_E start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT
Table 2: Physical parameters used in the numerical simulation

A brief comment regarding our numerical calculations is in order. While the number of commensurate lattices that arise from two rotated square lattices is infinite, namely angles originating crystalline structures, numerical restrictions impede both, to employ lower angle grids, and the use of larger lattices.

References