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

thanks: These authors contributed equally to this work.thanks: These authors contributed equally to this work.

Electronic structure of biased alternating-twist multilayer graphene

Kyungjin Shin Department of Physics and Astronomy, Seoul National University, Seoul 08826, Korea    Yunsu Jang Department of Physics and Astronomy, Seoul National University, Seoul 08826, Korea    Jiseon Shin Department of Physics, University of Seoul, Seoul 02504, Korea LG Electronics, CTO Division, Seocho R&D Campus, Seoul 06772, Korea    Jeil Jung [email protected] Department of Physics, University of Seoul, Seoul 02504, Korea Department of Smart Cities, University of Seoul, Seoul 02504, Korea    Hongki Min [email protected] Department of Physics and Astronomy, Seoul National University, Seoul 08826, Korea
Abstract

We theoretically study the energy and optical absorption spectra of alternating twist multilayer graphene (ATMG) under a perpendicular electric field. We obtain analytically the low-energy effective Hamiltonian of ATMG up to pentalayer in the presence of the interlayer bias by means of first-order degenerate-state perturbation theory, and present general rules for constructing the effective Hamiltonian for an arbitrary number of layers. Our analytical results agree to an excellent degree of accuracy with the numerical calculations for twist angles θ2.2\theta\gtrsim 2.2^{\circ} that are larger than the typical range of magic angles. We also calculate the optical conductivity of ATMG and determine its characteristic optical spectrum, which is tunable by the interlayer bias. When the interlayer potential difference is applied between consecutive layers of ATMG, the Dirac cones at the two moiré Brillouin zone corners K¯\bar{K} and K¯\bar{K}^{\prime} acquire different Fermi velocities, generally smaller than that of monolayer graphene, and the cones split proportionally in energy resulting in a step-like feature in the optical conductivity.

I Introduction

Twisted graphene systems have attracted widespread attention after the discovery of superconductivity and correlated insulating states Cao2018a ; Cao2018b ; Yankowitz2019 ; Lu2019 in magic-angle twisted bilayer graphene (TBG). By twisting two graphene layers, a new long-period structure, called a moiré superlattice, emerges due to spatially varying interlayer coupling, generating a unique band structure and associated electronic properties which strongly depend on the twist angle. Especially at the so-called magic angles, the Fermi velocity vanishes and nearly flat bands are formed Laissardiere2010 ; Morell2010 ; Bistritzer2011 ; Tarnopolsky2019 , providing an ideal platform to study correlated electron phenomena where electron-electron interactions are dominant over the kinetic energy.

Studies beyond TBG have been extended to systems like twisted double-bilayer graphene Koshino2019 ; Chebrolu2019 ; Lee2019 ; Shen2020 ; Liu2020 ; He2021 and twisted triple-bilayer graphene Shin2022 , and even to other two-dimensional moiré material systems Naik2018 ; Kariyado2019 ; Xian2019 ; Chen2019 ; Arora2020 ; Wang2020 ; An2020 , that revealed interesting interaction-driven phenomena such as correlated insulating Lee2019 ; Shen2020 ; Liu2020 ; He2021 ; Xian2019 and topological Koshino2019 phases that are in situ tunable.

Among them, the alternating twist multilayer graphene (ATMG) has been studied intensively both theoretically Khalaf2019 ; Carr2020 ; Bezanilla2020 ; Lei2021 ; Phong2021 ; Lake2021 ; Choi2021 ; Nguyen2022 ; Leconte2022 and experimentally Park2021 ; Hao2021 ; Cao2021 ; Kim2022 ; Park2022 ; Zhang2022 ; Shen2023 whose larger magic angle gives them an advantage over TBG. In particular, ATMG has attracted much attention due to its robust superconductivity observed from bilayer to pentalayer samples Cao2018b ; Yankowitz2019 ; Lu2019 ; Park2021 ; Hao2021 ; Cao2021 ; Kim2022 ; Park2022 ; Zhang2022 , while reports of superconductivity in other two-dimensional moiré systems are regarded as ambiguous Liu2020 ; He2021 ; Chen2019 ; Arora2020 ; Wang2020 ; An2020 ; Balents2020 .

In this paper, we theoretically analyze the effect in the electronic structure of a perpendicular electric field on ATMG. Using first-order degenerate-state perturbation theory treating the interlayer potential as a perturbation, we analytically investigate the low-energy effective Hamiltonian and its energy spectrum that becomes more accurate as the twist angle is increased. Then, we calculate the optical conductivity of biased ATMG that reveals a step-like feature arising from the splitting of Dirac nodes and their Fermi velocity renormalization introduced by the applied electric field.

The paper is organized as follows. In Sec. II, we introduce a model of ATMG in the presence of the interlayer potential difference between consecutive layers and derive the corresponding low-energy effective Hamiltonian analytically up to pentalayer. We also present general rules for constructing the effective Hamiltonian of biased ATMG with an arbitrary number of layers. In Sec. III, we calculate the longitudinal optical conductivity of biased ATMG, and explain their characteristic optical absorption spectrum. Finally, in Sec. IV, we discuss the interlayer coupling strength range for which our model is valid and summarize our main results.

II Electronic structure

II.1 Model

Refer to caption
Figure 1: Schematic illustration of the biased alternating twist multilayer graphene with N=3N=3 layers.

We consider a model of vertically stacked NN graphene layers with the \ellth layer alternatingly twisted by an angle θ=(1)θ/2\theta_{\ell}=(-1)^{\ell}\theta/2, as shown in Fig. 1. For a perpendicular electric field, we assume that the interlayer potential difference UU is the same between the two adjacent layers. Following Leconte et al. Leconte2022 , the Hamiltonian of ATMG in the presence of an interlayer potential difference can be expressed as

H=(H𝒌(1)T(𝒓)0T(𝒓)H𝒌(2)T(𝒓)0T(𝒓)H𝒌(3))+VH=\begin{pmatrix}H_{\boldsymbol{k}}^{(1)}&T(\boldsymbol{r})&0&\cdots\\ T^{\dagger}(\boldsymbol{r})&H_{\boldsymbol{k}}^{(2)}&T^{\dagger}(\boldsymbol{r})&\cdots\\ 0&T(\boldsymbol{r})&H_{\boldsymbol{k}}^{(3)}&\cdots\\ \vdots&\vdots&\vdots&\ddots\end{pmatrix}+V (1)

where the diagonal blocks H𝒌()=v0(𝒌𝝈θ)H_{\boldsymbol{k}}^{(\ell)}=\hbar v_{0}(\boldsymbol{k}\cdot\boldsymbol{\sigma}_{\theta_{\ell}}) with 𝝈θ=ei2θσz𝝈ei2θσz\boldsymbol{\sigma}_{\theta_{\ell}}=e^{\frac{i}{2}\theta_{\ell}\sigma_{z}}\boldsymbol{\sigma}e^{-\frac{i}{2}\theta_{\ell}\sigma_{z}} contain the Dirac cones of twisted graphene layers, T(𝒓)T(\boldsymbol{r}) is the interlayer tunneling matrix, and V=diag(V(1)𝕀2,V(2)𝕀2,,V(N)𝕀2)V={\rm{diag}}(V^{(1)}\mathbb{I}_{2},V^{(2)}\mathbb{I}_{2},...,V^{(N)}\mathbb{I}_{2}) is a diagonal matrix that captures the interlayer potential differences. The Fermi velocity of the monolayer graphene is set as v0=3a|t|/2106v_{0}=\sqrt{3}a\lvert t\rvert/2\hbar\simeq 10^{6} m/s with the lattice constant a=2.46a=2.46 Å\rm\AA and the nearest-neighbor intralayer hopping parameter t=3.1t=-3.1 eV.

In our model, the electric potential sequence V()V^{(\ell)} is defined to satisfy both V(+1)V()=UV^{(\ell+1)}-V^{(\ell)}=U and V(1)=V(N)V^{(1)}=-V^{(N)}, and the interlayer tunneling at the twisted interface takes the form

T(𝒓)=j=0,±ei𝒒j𝒓Tj,T(\boldsymbol{r})=\sum_{j=0,\pm}e^{i\boldsymbol{q}_{j}\cdot\boldsymbol{r}}T^{j}, (2)

where 𝒒0\boldsymbol{q}_{0}, 𝒒±\boldsymbol{q}_{\pm} are given by 𝒒0=2kDsin(θ/2)(0,1)\boldsymbol{q}_{0}=2k_{\rm D}\sin(\theta/2)(0,-1), 𝒒±=2kDsin(θ/2)(±3/2,1/2)\boldsymbol{q}_{\pm}=2k_{\rm D}\sin(\theta/2)(\pm\sqrt{3}/2,1/2) with the Dirac momentum for monolayer graphene kD=4π/3ak_{\rm D}=4\pi/3a. Our model also considers the corrugated lattice structure due to the effect of out-of-plane relaxation, leading to the larger interlayer spacing in AA-stacking region than the AB/BA-stacking region Uchida2014 ; Wijk2015 , thus resulting in unequal intra/inter sublattice hopping terms ww and ww^{\prime}, respectively. Following the convention of an initial AA stacking configuration Jung2014 , the interlayer tunneling matrices are given by

T0=(wwww),T±=(wwei2π/3we±i2π/3w),T^{0}=\begin{pmatrix}w^{\prime}&w\\ w&w^{\prime}\end{pmatrix},\;T^{\pm}=\begin{pmatrix}w^{\prime}&we^{\mp i2\pi/3}\\ we^{\pm i2\pi/3}&w^{\prime}\end{pmatrix}, (3)

where w=0.0939w^{\prime}=0.0939 eV and w=0.12w=0.12 eV Chebrolu2019 . For the full numerical calculations, we include the lattice corrugation (ww)(w\neq w^{\prime}), whereas we assume the rigid model of equal hopping terms (w=w=0.12w=w^{\prime}=0.12 eV) for simplicity when we study the Hamiltonian analytically. We choose representative twisted angles 33^{\circ}-55^{\circ} for the numerical calculations above the typical magic angle values that lie between 11^{\circ}-22^{\circ}. For these large angles the interlayer coupling substantially reduces the Fermi velocity of the dispersive Dirac cones near the two moiré Brillouin zone (mBZ) corners K¯\bar{K} and K¯\bar{K}^{\prime} but have not completely flattened them.

In the absence of the interlayer potential difference, the effective Hamiltonian of the ATMG at K¯\bar{K} and K¯\bar{K}^{\prime} can be described as a set of TBG models at different angles with an additional decoupled monolayer graphene model at K¯\bar{K} (or at K¯\bar{K}^{\prime} depending on the continuum model we start with) for an odd number of layers Khalaf2019 . The electronic structure of ATMG has a close analogy with Bernal-stacked multilayer graphene where the effective Hamiltonian is described by a set of bilayer graphene models with different effective masses with an additional decoupled monolayer graphene model for an odd number of layers min2008a ; min2008b . This analogy between Bernal stacked multilayer graphene and ATMG can be expanded to their wave functions.

We now construct the wave function of ATMG at K¯\bar{K} or K¯\bar{K}^{\prime} using the first shell model, in which the momentum-space lattice is truncated at the nearest-neighbor shell of the moiré reciprocal lattice 𝑮\bm{G} vectors, assuming the rigid model (w=ww=w^{\prime}). For the bilayer case (N=2N=2), the zero-energy eigenstates near K¯\bar{K} and K¯\bar{K}^{\prime} consist of four two-component spinors as follows:

ψλ,K¯(orK¯)TBG=11+6α2(aλbλ)or(bλaλ),\psi^{\rm{TBG}}_{\lambda,\bar{K}(\rm{or}\;\it{\bar{K}}^{\prime})}=\frac{1}{\sqrt{1+6\alpha^{2}}}\begin{pmatrix}a_{\lambda}\\ b_{\lambda}\end{pmatrix}\;\rm{or}\;\begin{pmatrix}b_{\lambda}\\ a_{\lambda}\end{pmatrix}, (4)

where α=w/[2v0kDsin(θ/2)]\alpha=w/\left[2v_{0}k_{\rm D}\sin(\theta/2)\right] is a dimensionless parameter. Following Bistritzer and MacDonald Bistritzer2011 , we define aλa_{\lambda} as a normalized eigenstate of 𝒌^𝝈θ\hat{\boldsymbol{k}}\cdot\boldsymbol{\sigma}_{\theta_{\ell}} corresponding to the eigenvalue λ=±1\lambda=\pm 1, and bλ=(b𝒒0,λ,b𝒒+,λ,b𝒒,λ)Tb_{\lambda}=(b_{\boldsymbol{q}_{0},\lambda},b_{\boldsymbol{q}_{+},\lambda},b_{\boldsymbol{q}_{-},\lambda})^{\rm T} determined by the equation b𝒒j,λ=hj1Tjaλb_{\boldsymbol{q}_{j},\lambda}=-h_{j}^{-1}T_{j}^{\dagger}a_{\lambda} with hj=v0(𝒌+𝒒j)𝝈θh_{j}=\hbar v_{0}(\boldsymbol{k}+\boldsymbol{q}_{j})\cdot\boldsymbol{\sigma}_{\theta_{\ell}}. In a similar way to Bernal stacked multilayer graphene, the eigenfunctions of ATMG have the form of the solution of a one-dimensional chain problem, thus we can construct the eigenfunctions of ATMG in the following manner Khalaf2019 :

Ψr,λ()=2τN+1sin(θr)ψr,λTBG,\Psi_{r,\lambda}^{(\ell)}=\sqrt{\frac{2\tau}{N+1}}\;\sin(\ell\theta_{r})\;\psi^{\rm{TBG}}_{r,\lambda}, (5)

where τ=2δr,n+1\tau=2-\delta_{r,n+1}, θr=rπ/(N+1)\theta_{r}=r\pi/(N+1) with r=1,2,,nr=1,2,\ldots,n for even N=2nN=2n or for odd N=2n+1N=2n+1 with additional r=(n+1)r=(n+1)th mode near K¯\bar{K}, and ψr,λTBG\psi^{\rm{TBG}}_{r,\lambda} can be obtained from ψλTBG\psi^{\rm{TBG}}_{\lambda} in Eq. (4) by letting αtrα\alpha\rightarrow t_{r}\alpha and bλtrbλb_{\lambda}\rightarrow t_{r}b_{\lambda} with tr=2cosθrt_{r}=2\,\cos{\theta_{r}}. Here, Ψr,λ=(Ψr,λ(1),Ψr,λ(2),,Ψr,λ(N))T\Psi_{r,\lambda}=(\Psi_{r,\lambda}^{(1)},\Psi_{r,\lambda}^{(2)},\ldots,\Psi_{r,\lambda}^{(N)})^{\rm T} is a normalized eigenstate of the effective Hamiltonian Heff=vr(𝒌𝝈)H_{\rm eff}=\hbar v_{r}(\boldsymbol{k}\cdot\boldsymbol{\sigma}) of ATMG with |Ψr,λ|2=1\lvert\Psi_{r,\lambda}\rvert^{2}=1, where vrv_{r} is a Fermi velocity of the Dirac cone Ψr,λ\Psi_{r,\lambda} given by

vrv0=13tr2α21+6tr2α2.\frac{v_{r}}{v_{0}}=\frac{1-3t_{r}^{2}\alpha^{2}}{1+6t_{r}^{2}\alpha^{2}}. (6)

Inserting r=n+1r=n+1 in Eq. (6), one finds that the (n+1)(n+1)th mode for the odd number of layers corresponds to an eigenstate of the decoupled monolayer Hamiltonian.

In the presence of the interlayer potential difference UU, we obtain analytically the low-energy effective Hamiltonian using first-order degenerate-state perturbation theory based on the minimal size Hamiltonian including only the first shell of the moiré 𝑮\bm{G} vectors. Due to the electric field, the Dirac cones near K¯\bar{K} or K¯\bar{K}^{\prime} are hybridized and split from one another, so the effective Hamiltonian of each Dirac node would be altered in the following form:

Heff=Δ(α,U)+v(𝒌𝝈),H_{\rm eff}=\Delta(\alpha,U)+\hbar v^{*}(\boldsymbol{k}\cdot\boldsymbol{\sigma}), (7)

where Δ(α,U)\Delta(\alpha,U) and vv^{*} are the energy shift and modified Fermi velocity of the effective Hamiltonian, respectively. The energy shift due to the interlayer potential UU can be expressed as Δ(α,U)=C(α)U\Delta(\alpha,U)=C(\alpha)U, and the Fermi velocity vv^{*} can be expressed as a linear combination of vrv_{r}. In the following Secs. II.B and II.C, we illustrate the effective Hamiltonian of N=3N=3 and 44 ATMG as examples. We leave the discussions of the analytical results for the N=5N=5 case to Appendix A.

II.2 N=3N=3

Here we derive the effective Hamiltonian of alternating twist trilayer graphene (AT3G) at the K¯\bar{K} and K¯\bar{K}^{\prime} points of the mBZ. There are two Dirac cones centered at K¯\bar{K} with v0v_{0} and v1v_{1} Fermi velocities, as shown in Fig. 2(a), thus the size of the perturbation matrix VK¯V_{\bar{K}} would be 2×22\times 2. Note that v0v_{0} represents the Fermi velocity of monolayer graphene. Using Eq. (4), we obtain the following normalized wave functions Ψr,λ\Psi_{r,\lambda} near K¯\bar{K} in our first shell model:

Refer to caption
Figure 2: Band structure of N=3N=3 ATMG at θ=3\theta=3^{\circ} for (a) U=0U=0 and (b) U=0.1U=0.1 eV. The left and right insets to (a) and (b) represent the schematic band structure near K¯\bar{K} and K¯\bar{K}^{\prime}. (c) C(α)C(\alpha) and (d) v/v0v^{\ast}/v_{0} as a function of twist angle θ\theta for the full numerical calculations (solid line) and the analytical result from the rigid (ω\omega = ω\omega^{\prime}) first shell model (dashed line).
Ψ1,λ=12+24α2(aλ2bλaλ),Ψ2,λ=12(aλ0aλ).\Psi_{1,\lambda}=\frac{1}{\sqrt{2+24\alpha^{2}}}\begin{pmatrix}a_{\lambda}\\ 2b_{\lambda}\\ a_{\lambda}\end{pmatrix},\;\Psi_{2,\lambda}=\frac{1}{\sqrt{2}}\begin{pmatrix}a_{\lambda}\\ 0\\ -a_{\lambda}\end{pmatrix}.\; (8)

At K¯\bar{K} in AT3G, the perturbation V^\hat{V} in the first shell model is given by V^=diag\hat{V}=\rm{diag}(U𝕀2,𝟎6,U𝕀2-U\mathbb{I}_{2},\boldsymbol{0}_{6},U\mathbb{I}_{2}). Then, in the basis of the wave functions in Eq. (8), we obtain the perturbation matrix VK¯V_{\bar{K}} using V11=V22=0V_{11}=V_{22}=0 and V12=V21=U/1+12α2V_{12}=V_{21}=-U/\sqrt{1+12\alpha^{2}}. By diagonalizing VK¯V_{\bar{K}}, we obtain the effective Hamiltonian of biased AT3G near K¯\bar{K} as

Heff,K¯(±)=±C(α)U+v(𝒌𝝈),H_{{\rm eff},\bar{K}}^{(\pm)}=\pm\;C(\alpha)U+\hbar v^{*}(\boldsymbol{k}\cdot\boldsymbol{\sigma}), (9)

where C(α)=1/1+12α2C(\alpha)=1/\sqrt{1+12\alpha^{2}} and v=(v0+v1)/2v^{*}=\left(v_{0}+v_{1}\right)/2. Comparing the left inset of Figs. 2(a) and 2(b), we can deduce that the two Dirac bands near K¯\bar{K} are hybridized equally and split by 2C(α)U2C(\alpha)U acquiring the average Fermi velocity vv^{*} from the unbiased values.

On the other hand, near K¯\bar{K^{\prime}}, only one Dirac cone with v1v_{1} exists, whose wave function is

Ψ1,λ=11+12α2(bλaλbλ).\Psi_{1,\lambda}=\frac{1}{\sqrt{1+12\alpha^{2}}}\begin{pmatrix}b_{\lambda}\\ a_{\lambda}\\ b_{\lambda}\end{pmatrix}. (10)

Then, the perturbation matrix VK¯V_{\bar{K}^{\prime}} vanishes and the Dirac cone at K¯\bar{K}^{\prime} remains unaltered to leading order in UU, resulting in the effective Hamiltonian

Heff,K¯=v1(𝒌𝝈).H_{{\rm eff},\bar{K}^{\prime}}=\hbar v_{1}(\boldsymbol{k}\cdot\boldsymbol{\sigma}). (11)

In Figs. 2(c) and 2(d), we illustrate the result of the leading-order energy splitting coefficient C(α)C(\alpha) and the modified Fermi velocity vv^{*} obtained from the analytic model and numerical method as a function of twist angle.

II.3 N=4N=4

Refer to caption
Figure 3: Similar to panels (a)-(c) in Fig. 2, but for N=4N=4 ATMG. If U<0U<0, the energy shifts are reversed. In panel (d), we show the two Fermi velocities v±v_{\pm}^{*} as given in Eq. (16) of the positively and negatively shifted Dirac cones illustrated in the inset to (b).

In the following, we derive the effective Hamiltonian of alternating twist tetralayer graphene (AT4G) at K¯\bar{K} and K¯\bar{K}^{\prime}. At K¯\bar{K}, there are two Dirac cones with the velocities v1v_{1} and v2v_{2} as shown in Fig. 3(a), and the corresponding wave functions are given by

Ψr,λ=25(1+6tr2α2)(sinθraλsin2θrtrbλsin3θraλsin4θrtrbλ)\Psi_{r,\lambda}=\frac{2}{\sqrt{5(1+6t_{r}^{2}\alpha^{2})}}\begin{pmatrix}\sin{\theta_{r}}\cdot a_{\lambda}\\ \sin{2\theta_{r}}\cdot t_{r}b_{\lambda}\\ \sin{3\theta_{r}}\cdot a_{\lambda}\\ \sin{4\theta_{r}}\cdot t_{r}b_{\lambda}\\ \end{pmatrix} (12)

with r=1,2r=1,2. The perturbation V^\hat{V} at K¯\bar{K} in this case is given by V^=diag\hat{V}=\rm{diag}(3U2𝕀2,U2𝕀6,U2𝕀2,3U2𝕀6-\frac{3U}{2}\mathbb{I}_{2},-\frac{U}{2}\mathbb{I}_{6},\frac{U}{2}\mathbb{I}_{2},\frac{3U}{2}\mathbb{I}_{6}). Since there are two Dirac cones at K¯\bar{K}, the size of the perturbation matrix VK¯V_{\bar{K}} would be 2×22\times 2, and its elements VrrV_{rr^{\prime}} are expressed as

Vrr=4UN+116trtrα2(1+6tr2α2)(1+6tr2α2)×l=0N/2(2N12)sin(2+1)θrsin(2+1)θr=2U516trtrα2(1+6tr2α2)(1+6tr2α2)(3sinθr×sinθr+sin3θrsin3θr+5sin5θrsin5θr).V_{rr^{\prime}}=\frac{4U}{N+1}\frac{1-6t_{r}t_{r^{\prime}}\alpha^{2}}{\sqrt{(1+6t_{r}^{2}\alpha^{2})(1+6t_{r^{\prime}}^{2}\alpha^{2})}}\\ \times\sum_{l=0}^{N/2}\left(2\ell-\frac{N-1}{2}\right)\sin{(2\ell+1)\theta_{r}}\sin{(2\ell+1)\theta_{r^{\prime}}}\\ =\frac{2U}{5}\frac{1-6t_{r}t_{r^{\prime}}\alpha^{2}}{\sqrt{(1+6t_{r}^{2}\alpha^{2})(1+6t_{r^{\prime}}^{2}\alpha^{2})}}(-3\sin{\theta_{r}}\\ \times\sin{\theta_{r^{\prime}}}+\sin{3\theta_{r}}\sin{3\theta_{r^{\prime}}}+5\sin{5\theta_{r}}\sin{5\theta_{r^{\prime}}}). (13)

By diagonalizing the matrix VK¯V_{\bar{K}}, we obtain the effective Hamiltonian of biased AT4G near K¯\bar{K} as

Heff,K¯=C±(α)U+v±(𝒌𝝈),H_{{\rm eff},\bar{K}}=C_{\pm}(\alpha)U+\hbar v_{\pm}^{*}(\boldsymbol{k}\cdot\boldsymbol{\sigma}), (14)

where C±(α)C_{\pm}(\alpha) and v±v_{\pm}^{*} corresponding to upward and downward shifted Dirac cones are given by

C±(α)=12(1+18α2+36α4)[(1+12α236α4)±(1+12α2)(1+18α2+45α4+108α6)]C_{\pm}(\alpha)=\frac{1}{2(1+18\alpha^{2}+36\alpha^{4})}[-(1+12\alpha^{2}-36\alpha^{4})\\ \pm\sqrt{(1+12\alpha^{2})(1+18\alpha^{2}+45\alpha^{4}+108\alpha^{6})}] (15)

and

v±=A±2v1+B2v2A±2+B2.v^{*}_{\pm}=\frac{A_{\pm}^{2}v_{1}+B^{2}v_{2}}{A_{\pm}^{2}+B^{2}}. (16)

Here, A±A_{\pm} and BB are unnormalized mixing coefficients of the two Dirac cones given, respectively, by

A±\displaystyle A_{\pm} =\displaystyle= 1+15α236α4\displaystyle 1+15\alpha^{2}-36\alpha^{4}
±\displaystyle\pm 5(1+12α2)(1+18α2+45α4+108α6),\displaystyle\sqrt{5(1+12\alpha^{2})(1+18\alpha^{2}+45\alpha^{4}+108\alpha^{6})},
B\displaystyle B =\displaystyle= 2(1+6α2)1+18α2+36α4.\displaystyle-2(1+6\alpha^{2})\sqrt{1+18\alpha^{2}+36\alpha^{4}}. (17b)

From the above result, we find that the two Dirac cones with the velocities v1v_{1} and v2v_{2} are hybridized with the ratio of A±A_{\pm} and BB, and shifted by C±(α)UC_{\pm}(\alpha)U, as schematically illustrated in Fig. 3(b).

On the other hand, near K¯\bar{K}^{\prime}, the wave functions for two Dirac cones with the velocity vrv_{r} (r=1,2)(r=1,2) are given by

Ψr,λ=25(1+6tr2α2)(sinθrtrbλsin2θraλsin3θrtrbλsin4θraλ).\Psi_{r,\lambda}=\frac{2}{\sqrt{5(1+6t_{r}^{2}\alpha^{2})}}\begin{pmatrix}\sin{\theta_{r}}\cdot t_{r}b_{\lambda}\\ \sin{2\theta_{r}}\cdot a_{\lambda}\\ \sin{3\theta_{r}}\cdot t_{r}b_{\lambda}\\ \sin{4\theta_{r}}\cdot a_{\lambda}\\ \end{pmatrix}. (18)

Since sinθr=(1)rsin(N+1)θr\sin{\ell\theta_{r}}=(-1)^{r}\sin{(N+1-\ell)\theta_{r}}, the wave function at K¯\bar{K}^{\prime} can be obtained by reversing the components of the wave function at K¯\bar{K}. Therefore, the effective Hamiltonian of biased AT4G near K¯\bar{K}^{\prime} can be obtained by reversing the sign of the interlayer potential difference UU in Eq. (14) as

Heff,K¯=C±(α)U+v±(𝒌𝝈),H_{{\rm eff},\bar{K}^{\prime}}=-\;C_{\pm}(\alpha)U+\hbar v_{\pm}^{*}(\boldsymbol{k}\cdot\boldsymbol{\sigma}), (19)

where C±(α)C_{\pm}(\alpha) and v±v_{\pm}^{*} are the same as those at K¯\bar{K}. In detail, our model Hamiltonian [Eq. (1)] for N=4N=4 has a combined symmetry expressed as

(Σ^𝒯^)H(𝒌)(Σ^𝒯^)1=H(𝒌),(\hat{\Sigma}\hat{\mathcal{T}})H(\boldsymbol{k})(\hat{\Sigma}\hat{\mathcal{T}})^{-1}=-H(-\boldsymbol{k}), (20)

where

Σ^=(000σx00σx00σx00σx000).\hat{\Sigma}=\begin{pmatrix}0&0&0&\sigma_{x}\\ 0&0&-\sigma_{x}&0\\ 0&\sigma_{x}&0&0\\ -\sigma_{x}&0&0&0\end{pmatrix}. (21)

We note that Σ^\hat{\Sigma} changes only the valley index (KKK\leftrightarrow K^{\prime}) keeping the same mBZ corner points (K¯K¯\bar{K}\rightarrow\bar{K}, K¯K¯\bar{K}^{\prime}\rightarrow\bar{K}^{\prime}Moon2013 , whereas the time-reversal operator 𝒯^\hat{\mathcal{T}} changes both the valley index (KKK\leftrightarrow K^{\prime}) and the mBZ corner points (K¯K¯\bar{K}\leftrightarrow\bar{K}^{\prime}). This combined symmetry is preserved even in the presence of an interlayer potential difference, thus the effective Hamiltonians between K¯\bar{K} and K¯\bar{K}^{\prime}, which are respectively described in Eqs. (14) and (19), are also related as Eq. (20).

II.4 Arbitrary NN

As the layer number NN is increased, the size of the perturbation matrix is also proportionally increased and it becomes progressively cumbersome to obtain analytically the effective Hamiltonian of ATMG for a large number of layers in the presence of an applied field even if we use the first shell model. Instead, here want to provide the general behavior patterns of the effective Hamiltonian of biased ATMG for arbitrary NN. Tables 1 and 2 show the summary of the effective Hamiltonian for N=2N=2-88 ATMG in the presence of the interlayer potential difference.

Table 1: Summary of the effective Hamiltonian of ATMG for odd numbers of layers N=3,5,7N=3,5,7.
[Uncaptioned image]

Firstly, for ATMG with an odd number of layers, there are (N1)/2(N-1)/2 TBG Dirac cones labeled by vrv_{r} [r=1,2,,(N1)/2]\left[r=1,2,\ldots,(N-1)/2\right] near the two mBZ corners K¯\bar{K} and K¯\bar{K}^{\prime}, and one decomposed monolayer Dirac cone with v0v_{0} at K¯\bar{K}. Regardless of the layer number NN and mBZ symmetry points, the form of the perturbation matrix VV is solely determined by the number of Dirac cones at K¯\bar{K} or K¯\bar{K}^{\prime}, even though its elements depend on NN. One can thus find the same form of the effective Hamiltonian when the number of Dirac cones is the same, as seen in Table 1. In detail, if there are mm Dirac cones at K¯\bar{K} or K¯\bar{K}^{\prime}, we have m/2m/2 pairs of Dirac-cones shifted by ±Δi\pm\Delta_{i} (i=1,2,,m/2i=1,2,\ldots,m/2) for even mm, whereas we have (m1)/2(m-1)/2 pairs of Dirac cones plus one Dirac cone without energy shift for odd mm. Each pair of Dirac cones has the same effective Fermi velocity.

Table 2: Summary of the effective Hamiltonian of ATMG for even numbers of layers N=2,4,6,8N=2,4,6,8. Here, we assume U>0U>0. For U<0U<0, the energy shifts are reversed.
[Uncaptioned image]

For ATMG with an even number of layers, there are N/2N/2 TBG Dirac cones labeled by vrv_{r} (r=1,2,,N/2)\left(r=1,2,\ldots,N/2\right) near K¯\bar{K} and K¯\bar{K}^{\prime} when U=0U=0. If an external field is applied (U0U\neq 0), the Dirac cones are split with different energy shift Δ\Delta and Fermi velocity vv^{*}. As already mentioned in Sec. II.C, the effect of an applied electric field at K¯\bar{K}^{\prime} (K¯\bar{K}) can be effectively described by flipping its direction (z^z^\hat{z}\rightarrow-\hat{z}) at K¯\bar{K} (K¯\bar{K}^{\prime}). Moreover, we can generalize Eq. (20) by expanding Σ^\hat{\Sigma} symmetry in Eq. (21) by conveniently alternating σx\sigma_{x} and σx-\sigma_{x}. Similarly to the N=4N=4 case, the combined Σ^𝒯^\hat{\Sigma}\hat{\mathcal{T}} symmetry is still preserved for ATMG with an even number of layers in the presence of an interlayer potential difference, relating the effective Hamiltonians between K¯\bar{K} and K¯\bar{K}^{\prime} with flipped energy shifts, as seen in Table 2.

Lastly, let us consider the effective Hamiltonian of biased ATMG in the asymptotic limit (α0\alpha\rightarrow 0) where the twist angle θ\theta becomes much larger than the first magic angle of ATMG. For the first shell model, |bλ|\lvert b_{\lambda}\rvert becomes proportional to α\alpha, so only monolayer terms aλa_{\lambda} of Ψr,λ\Psi_{r,\lambda} survive in this limit. Thus, the energy splitting coefficient C(α)C(\alpha) of ATMG with arbitrary NN at K¯\bar{K} (K¯\bar{K}^{\prime}) can be obtained as odd (even) layer components of V^\hat{V}, as schematically shown in Fig. 4.

Refer to caption
Figure 4: Schematic picture of the energy shifts of Dirac cones at K¯\bar{K} and K¯\bar{K}^{\prime} in biased ATMG in the asymptotic limit (α0\alpha\rightarrow 0).

On the other hand, the modified Fermi velocity vv^{*} converges to v0v_{0} as α0\alpha\rightarrow 0, since all eigenstates of biased ATMG in this limit have just a single monolayer term, giving the monolayer graphene Dirac cone with the velocity v0v_{0}. Therefore, the effective Hamiltonian would be described by a set of monolayer graphene Hamiltonian with the energy shift described by C(α0)UC(\alpha\rightarrow 0)U, which can be obtained by the pattern presented in Fig. 4. Figure 5 shows the band structure of N=58N=5-8 ATMG in the presence of the interlayer potential difference U=0.1U=0.1 eV at θ=5\theta=5^{\circ} along with the analytical result obtained in the asymptotic limit, which agrees closely with the full numerical result except for small deviations in the Fermi velocity of the Dirac cones.

III Optical conductivity

The Kubo formula for the optical conductivity in the non-interacting and clean limit is given by Mahan2000

σij(ω)\displaystyle\sigma_{ij}(\omega) =\displaystyle= ie2s,sd2k(2π)2fs,𝒌fs,𝒌εs,𝒌εs,𝒌\displaystyle-\frac{ie^{2}}{\hbar}\sum_{s,s^{\prime}}\int\frac{d^{2}k}{(2\pi)^{2}}\frac{f_{s,\bm{k}}-f_{s^{\prime},\bm{k}}}{\varepsilon_{s,\bm{k}}-\varepsilon_{s^{\prime},\bm{k}}} (22)
×\displaystyle\times Miss(𝒌)Mjss(𝒌)ω+εs,𝒌εs,𝒌+i0+,\displaystyle\frac{M^{ss^{\prime}}_{i}(\bm{k})M^{s^{\prime}s}_{j}(\bm{k})}{\hbar\omega+\varepsilon_{s,\bm{k}}-\varepsilon_{s^{\prime},\bm{k}}+i0^{+}},

where i,j=x,yi,j=x,y, fs,𝒌=1/[1+e(εs,𝒌μ)/kBT]f_{s,\bm{k}}=1/[1+e^{(\varepsilon_{s,\bm{k}}-\mu)/k_{\rm B}T}] is the Fermi distribution function for the band index ss and wave vector 𝒌\bm{k}, μ\mu is the chemical potential and Miss(𝒌)=s,𝒌|v^i|s,𝒌M^{ss^{\prime}}_{i}(\bm{k})=\langle{s,\bm{k}}|\hbar\hat{v}_{i}|{s^{\prime},\bm{k}}\rangle with the velocity operator v^i\hat{v}_{i} obtained from the relation v^i=1H^ki\hat{v}_{i}=\frac{1}{\hbar}\frac{\partial\hat{H}}{\partial k_{i}}.

Refer to caption
Figure 5: Band structure of N=58N=5-8 ATMG at θ=5\theta=5^{\circ} with U=0.1U=0.1 eV. Solid and dashed lines represent the numerical calculations and the analytical result obtained in the α0\alpha\rightarrow 0 limit, respectively.

In the following, we consider the real part of the longitudinal optical conductivity for μ=0\mu=0 at zero temperature with a finite broadening term η=5\eta=5 meV replacing the 0+0^{+} term in Eq. (22) for numerical calculations. We plot the optical conductivities of AT3G and AT4G for the continuum model [Eq. (1)] with and without the interlayer potential difference UU at the twist angle θ=5\theta=5^{\circ} in Figs. 6 and 7, respectively.

Refer to caption
Figure 6: Band structure and the longitudinal conductivity of N=3N=3 ATMG at θ=5\theta=5^{\circ} for (a), (c) U=0U=0 and (b), (d) U=0.1U=0.1 eV. The insets to (b) show an enlarged view of the band structure near K¯\bar{K} and K¯\bar{K}^{\prime}. The arrows in the band structure indicate interband transitions corresponding to peaks in the conductivity. In (d), a Drude peak appears at low frequencies due to intraband contributions.
Refer to caption
Figure 7: Same as Fig. 6 for N=4N=4 ATMG.

In the absence of the interlayer potential difference, the longitudinal conductivity converges to Nσ0N\sigma_{0} for both low- and high-frequency limits, as shown in Figs. 6(c) and 7(c). Here, σ0=gsve2/16\sigma_{0}=g_{\rm sv}e^{2}/16\hbar is the optical conductivity of charge-neutral monolayer graphene with the spin-valley degeneracy factor gsv=4g_{\rm sv}=4. The low-frequency conductivity originates from transitions within NN hybridized Dirac nodes located at K¯\bar{K} and K¯\bar{K}^{\prime}, whereas at high frequencies the interlayer coupling becomes negligible thus the conductivity approaches that of NN decoupled monolayer graphene sheets. At intermediate frequencies, a dominant peak appears around ω0.9\hbar\omega\sim 0.9 eV for θ=5\theta=5^{\circ} arising from interband transitions between states near the saddle point M¯\bar{M}, as indicated by the red arrows. The frequency where the dominant peak appears depends on the twist angle θ\theta but weakly depends on NN or UU.

In the presence of the interlayer potential difference, the conductivity shows a step-like feature at low frequencies, as shown in Figs. 6(d) and 7(d). For biased AT3G, two Dirac nodes at K¯\bar{K} are split by 2Δ2\Delta, thus, interband transitions are forbidden in the low-frequency limit, while interband transitions are allowed in the unaltered Dirac cone at K¯\bar{K}^{\prime}, giving σ0\sigma_{0}. For biased AT4G, two Dirac nodes at K¯\bar{K} and another two Dirac nodes at K¯\bar{K}^{\prime} are shifted by Δ±\Delta_{\pm}, thus, interband transitions are forbidden in the low-frequency limit, and the optical conductivity vanishes. As the frequency increases, the optical conductivity increases by 2σ02\sigma_{0} when ω2|Δ|\hbar\omega\sim 2\lvert\Delta\rvert and 2|Δ±|2\lvert\Delta_{\pm}\rvert for biased AT3G and AT4G, respectively, eventually approaching Nσ0N\sigma_{0}. This feature is very analogous to the optical conductivity of AA-stacked multilayer graphene, where the optical conductivity increases in steps of 2σ02\sigma_{0} toward Nσ0N\sigma_{0} when interband transitions occur within the same Dirac cones min2009 ; Tabert2012 . Unlike AA-stacked multilayer graphene, however, the velocity changes away from K¯\bar{K} or K¯\bar{K}^{\prime}, so the transition energy deviates from 2|Δ|2\lvert\Delta\rvert and 2|Δ±|2\lvert\Delta_{\pm}\rvert, especially at small twist angles. As the twist angle decreases, interband transitions arising from the M¯\bar{M} and Γ¯\bar{\Gamma} points occur at lower energies and eventually mix with the interband transitions arising from the K¯\bar{K} and K¯\bar{K}^{\prime} points, blurring the step-like features explained above. The evolution of the optical conductivity with decreasing twist angle will be discussed in Appendix C.

IV Discussion

The analytical forms of the effective Hamiltonian for biased ATMG near K¯\bar{K} or K¯\bar{K}^{\prime} were obtained using the first shell model of the moiré 𝑮\bm{G} vectors, which is valid within the radius about kcU/v0k_{c}\sim U/\hbar v_{0} in the 𝒌\bm{k} space beyond which two shifted Dirac cones by UU cross each other. When the twist angle becomes smaller than the typical values of the first magic angle θM(N)θM()2.2\theta_{\rm M}^{(N)}\leq\theta_{\rm M}^{(\infty)}\approx 2.2^{\circ} Khalaf2019 , our first shell model, which employs the nearest-neighbor truncation, is generally insufficient for accurately capturing the bands of an enlarged moiré superlattice of ATMG, resulting in the discrepancy between analytical and numerical results. Nevertheless, the analytical results obtained from our perturbation approach agree well with the full numerical calculations for twist angles θ2.2\theta\gtrsim 2.2^{\circ} where the interlayer coupling is weaker.

In summary, we have studied the effect of a perpendicular electric field on ATMG, focusing on the effects of an interlayer potential difference in altering the low-energy band structure and therefore the optical absorption spectrum, which can be used as a distinguishing experimental signature. Firstly, we analytically derived the low-energy effective Hamiltonian and its energy spectrum near the two moiré Dirac points K¯\bar{K} and K¯\bar{K}^{\prime} up to pentalayer by using first-order degenerate-state perturbation theory, treating the asymmetric interlayer potential difference as a perturbation. Then, we presented general rules for constructing the effective Hamiltonian of biased ATMG with an arbitrary number of layers. Lastly, we investigated the optical absorption spectrum of ATMG with and without an interlayer potential difference. We found that the longitudinal conductivity of biased ATMG showed a step-like feature arising from the splitting of Dirac nodes by the applied electric field, which is reminiscent of the optical conductivity features of AA-stacked multilayer graphene.

Acknowledgements.
This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (Grant No. 2018R1A2B6007837 and No. 2023R1A2C1005996), the Creative-Pioneering Researchers Program through Seoul National University (SNU), and the Center for Theoretical Physics. J. S. was supported by Korea NRF (Grant No. 2021R1A6A3A01087281), and J. J. acknowledges support from Korea NRF (Grant No. 2020R1A2C3009142).

Appendix A Derivation of the effective Hamiltonian for N=5N=5 ATMG

In this Appendix, we derive the effective Hamiltonian of alternating twist pentalayer graphene (AT5G) at the two moiré Dirac points K¯\bar{K} and K¯\bar{K}^{\prime}. Three Dirac cones labeled by v0v_{0}, v1v_{1}, and v2v_{2} exist near K¯\bar{K} as shown in Fig. 8(a), thus the size of the perturbation matrix VK¯V_{\bar{K}} is 3×33\times 3. Using Eq. (4), we obtain the following normalized wave functions Ψr,λ\Psi_{r,\lambda} near K¯\bar{K} in our first shell model:

Ψr,λ=26(1+6tr2α2)(sinθraλsin2θrtrbλsin3θraλsin4θrtrbλsin5θraλ),\Psi_{r,\lambda}=\frac{2}{\sqrt{6(1+6t_{r}^{2}\alpha^{2})}}\begin{pmatrix}\sin{\theta_{r}}\cdot a_{\lambda}\\ \sin{2\theta_{r}}\cdot t_{r}b_{\lambda}\\ \sin{3\theta_{r}}\cdot a_{\lambda}\\ \sin{4\theta_{r}}\cdot t_{r}b_{\lambda}\\ \sin{5\theta_{r}}\cdot a_{\lambda}\end{pmatrix}, (23a)
Ψ3,λ=13(aλ0aλ0aλ)\Psi_{3,\lambda}=\frac{1}{\sqrt{3}}\begin{pmatrix}a_{\lambda}\\ 0\\ -a_{\lambda}\\ 0\\ a_{\lambda}\\ \end{pmatrix} (23b)

with r=1,2r=1,2. At K¯\bar{K} in AT5G, the perturbation V^\hat{V} is given by V^=diag\hat{V}=\rm{diag}(2U𝕀2×2-2U\mathbb{I}_{2\times 2}, U𝕀6×6-U\mathbb{I}_{6\times 6}, 𝟎2×2\boldsymbol{0}_{2\times 2}, U𝕀6×6U\mathbb{I}_{6\times 6}, 2U𝕀2×22U\mathbb{I}_{2\times 2}). Following the same procedure described in Secs. II.B and II.C, we obtain the perturbation matrix VK¯V_{\bar{K}} with the elements of V11=V22=V33=V13=V31=0V_{11}=V_{22}=V_{33}=V_{13}=V_{31}=0, V12=V21=2U(1+9α2)/3(1+6α2)(1+12α2)V_{12}=V_{21}=-2U(1+9\alpha^{2})/\sqrt{3(1+6\alpha^{2})(1+12\alpha^{2})}, and V23=V32=4U/6(1+6α2)V_{23}=V_{32}=-4U/\sqrt{6(1+6\alpha^{2})} in the basis of the wave functions in Eq. (23). Therefore, we obtain the effective Hamiltonian of biased AT5G near K¯\bar{K} by diagonalizing VK¯V_{\bar{K}} as

Heff,K¯(0)\displaystyle H^{(0)}_{{\rm eff},\bar{K}} =\displaystyle= v0(𝒌𝝈),\displaystyle\hbar v_{0}^{*}(\boldsymbol{k}\cdot\boldsymbol{\sigma}), (24a)
Heff,K¯(±1)\displaystyle H^{(\pm 1)}_{{\rm eff},\bar{K}} =\displaystyle= ±CK¯(α)U+v1(𝒌𝝈),\displaystyle\pm C_{\bar{K}}(\alpha)U+\hbar v_{1}^{*}(\boldsymbol{k}\cdot\boldsymbol{\sigma}), (24b)

where

CK¯(α)=21+18α2+27α4(1+6α2)(1+18α2),C_{\bar{K}}(\alpha)=2\sqrt{\frac{1+18\alpha^{2}+27\alpha^{4}}{(1+6\alpha^{2})(1+18\alpha^{2})}},\vspace{-0.5mm} (25)

and

v0=A2v0+B2v1A2+B2,v1=B2v0+A2v12(A2+B2)+v22.v_{0}^{*}=\frac{A^{2}v_{0}+B^{2}v_{1}}{A^{2}+B^{2}},\;\;v_{1}^{*}=\frac{B^{2}v_{0}+A^{2}v_{1}}{2(A^{2}+B^{2})}+\frac{v_{2}}{2}.\vspace{-1mm} (26)

Here, A=1+9α2A=1+9\alpha^{2} and B=2(1+18α2)B=\sqrt{2(1+18\alpha^{2})} are unnormalized mixing coefficients of the Dirac cones.

Refer to caption
Figure 8: Similar to panels (a)-(c) in Fig. 2, but for N=5N=5 ATMG. In panel (d), we show the modified Fermi velocities v0v_{0}^{*} and v1v_{1}^{*} given in Eq. (26) at K¯\bar{K} and v=(v1+v2)/2v^{*}=(v_{1}+v_{2})/2 at K¯\bar{K}^{\prime} as illustrated in the inset to (b).

On the other hand, near K¯\bar{K}^{\prime}, there are two Dirac cones with the velocities v1v_{1} and v2v_{2} as shown in Fig. 8(a), and the corresponding wave functions are given by

Ψr,λ=26(1+6tr2α2)(sinθrtrbλsin2θraλsin3θrtrbλsin4θraλsin5θrtrbλ)\Psi_{r,\lambda}=\frac{2}{\sqrt{6(1+6t_{r}^{2}\alpha^{2})}}\begin{pmatrix}\sin{\theta_{r}}\cdot t_{r}b_{\lambda}\\ \sin{2\theta_{r}}\cdot a_{\lambda}\\ \sin{3\theta_{r}}\cdot t_{r}b_{\lambda}\\ \sin{4\theta_{r}}\cdot a_{\lambda}\\ \sin{5\theta_{r}}\cdot t_{r}b_{\lambda}\end{pmatrix} (27)

with r=1,2r=1,2. Then, the size of perturbation matrix VK¯V_{\bar{K}^{\prime}} would be 2×22\times 2 with the elements of V11=V22=0V_{11}=V_{22}=0 and V12=V21=U(1+12α2)/(1+6α2)(1+18α2)V_{12}=V_{21}=-U(1+12\alpha^{2})/\sqrt{(1+6\alpha^{2})(1+18\alpha^{2})}. By diagonalizing VK¯V_{\bar{K}^{\prime}}, we obtain the effective Hamiltonian of biased AT5G near K¯\bar{K}^{\prime} as

Heff,K¯(±)=±CK¯(α)U+v(𝒌𝝈),H^{(\pm)}_{{\rm eff},\bar{K}^{\prime}}=\pm\;C_{\bar{K}^{\prime}}(\alpha)U+\hbar v^{*}(\boldsymbol{k}\cdot\boldsymbol{\sigma}), (28)

where

CK¯=1+12α2(1+6α2)(1+18α2)C_{\bar{K}^{\prime}}=\frac{1+12\alpha^{2}}{\sqrt{(1+6\alpha^{2})(1+18\alpha^{2})}} (29)
Refer to caption
Figure 9: Same as Fig. 6 for N=5N=5 ATMG.

and v=(v1+v2)/2v^{*}=(v_{1}+v_{2})/2. We here notice that the effective Hamiltonian of biased AT5G near K¯\bar{K}^{\prime} has a similar form of one at biased AT3G near K¯\bar{K} since the number of Dirac cones are the same. In both cases, two Dirac cones are shifted by ±C(α)U\pm C(\alpha)U and hybridized equally, so that the equal Fermi velocity vv^{*} is assigned to be an average of unbiased ones. The only difference between them is the value of off-diagonal matrix elements, determining the energy shift of Dirac cones.

Appendix B Optical conductivity of N=5N=5 ATMG

In this Appendix, we present the real part of the longitudinal conductivity of AT5G with and without the interlayer potential difference UU. Figure 9 illustrates the longitudinal conductivity of AT5G for U=0U=0 and U=0.1U=0.1 eV at the twist angle θ=5\theta=5^{\circ}, respectively.

Refer to caption
Figure 10: Band structure of N=3N=3 ATMG at θ=2\theta=2^{\circ} and θ=3\theta=3^{\circ} for (a), (c) U=0U=0 and (b), (d) U=0.1U=0.1 eV, respectively, and the evolution of the longitudinal optical conductivity with decreasing twist angle for (e) U=0U=0 and (f) U=0.1U=0.1 eV. The arrows in the band structure indicate interband transitions corresponding to peaks in the conductivity.

In the absence of the interlayer potential difference, the longitudinal conductivity converges to 5σ05\sigma_{0} for both low- and high-frequency regions, as shown in Fig. 9(c). On the other hand, in the presence of the interlayer potential difference, the conductivity shows a step-like feature at low frequencies, as shown in Fig. 9(d). Specifically, the conductivity starts with σ0\sigma_{0} from the unshifted Dirac cone at K¯\bar{K} then increases toward 5σ05\sigma_{0} in steps of 2σ02\sigma_{0} when ω2|ΔK¯|\hbar\omega\sim 2\lvert\Delta_{\bar{K}}\rvert and 2|ΔK¯|2\lvert\Delta_{\bar{K}^{\prime}}\rvert, respectively, at which the forbidden interband transitions due to the splitting of Dirac nodes by the applied electric field can occur. The conductivity jump at ω2|ΔK¯|\hbar\omega\sim 2\lvert\Delta_{\bar{K}}\rvert, however, is only 1.5σ01.5\sigma_{0} less than 2σ02\sigma_{0}, approaching the conductivity value for U=0U=0 at that frequency. This mismatch is due to the continuous decrease of the optical conductivity as the frequency increases because interband transitions are no longer described by those between the Dirac nodes near K¯\bar{K} and K¯\bar{K}^{\prime}. When smaller UU or larger θ\theta is used, one may see more clearly the conductivity increase in steps of 2σ02\sigma_{0}. At intermediate frequencies, a dominant peak arises from interband transitions near M¯\bar{M}, as indicated by the red arrows.

Appendix C Evolution of the optical conductivity with decreasing-twist angle

In the following, we consider the evolution of the longitudinal optical conductivities with decreasing twist angle for μ=0\mu=0 at zero temperature with a smaller broadening term η=3\eta=3 meV compared to that used in Figs. 6 and 7 to capture the low-frequency behavior more accurately.

Refer to caption
Figure 11: Same as Fig. 10 for N=4N=4 ATMG.

We plot the optical conductivities of AT3G and AT4G with and without the interlayer potential difference UU at

various twist angles θ=2\theta=2^{\circ}-55^{\circ} in Figs. 10 and 11, respectively. Notice that in this section we only consider interband transitions, ignoring the Drude peak arising from intraband transitions.

When the interlayer potential difference is absent, the longitudinal conductivities converge to Nσ0N\sigma_{0} in the low-frequency limit but drop more quickly as the twist angle decreases due to the decrease in the bandwidth, as shown in Figs. 10(e) and 11(e). Furthermore, interband transitions arising from the M¯\bar{M} and Γ¯\bar{\Gamma} points [see Figs. 10(a), 10(c), 11(a), and 11(c)], which were regarded as high-energy transitions in Sec. III, occur at lower energies, and the corresponding peaks move toward the low-frequency region as the twist angle decreases.

When the interlayer potential difference is present, the step-like feature discussed in Sec. III can still be observed due to the interband transitions within the same Dirac cones, as shown in Figs. 10(f) and  11(f). However, as the twist angle decreases, interband transitions arising from the M¯\bar{M} and Γ¯\bar{\Gamma} points occur at lower energies and eventually mix with the interband transitions arising from the K¯\bar{K} and K¯\bar{K}^{\prime} points, blurring the step-like features. Furthermore, unlike AA-stacked multilayer graphene, the velocity changes away from K¯\bar{K} or K¯\bar{K}^{\prime} [see Figs. 10(b), 10(d), 11(b), and 11(d)], and additional peaks occur due to interband transitions from or to the ring of the crossed Dirac cones [marked as K1K_{1}^{\prime} in Fig. 11(d)] and due to interband transitions between other Dirac cones [marked as K2K_{2} in Fig. 10(b) or K3K_{3} and K4K_{4} in Fig. 11(b)], which become significant for smaller θ\theta or larger UU.

References

  • (1) Y. Cao, V. Fatemi, A. Demir, S. Fang, S. L. Tomarken, J. Y. Luo, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi, E. Kaxiras, R. C. Ashoori, and P. Jarillo-Herrero, Correlated Insulator Behaviour at Half-Filling in Magic-Angle Graphene Superlattices, Nature (London) 556, 80 (2018).
  • (2) Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Unconventional Superconductivity in Magic-Angle Graphene Superlattices, Nature (London) 556, 43 (2018).
  • (3) M. Yankowitz, S. Chen, H. Polshyn, Y. Zhang, K. Watanabe, T. Taniguchi, D. Graf, A. F. Young, and C. R. Dean, Tuning superconductivity in twisted bilayer graphene, Science 363, 1059-1064 (2019).
  • (4) X. Lu, P. Stepanov, W. Yang, M. Xie, M. A. Aamir, I. Das, C. Urgell, K. Watanabe, T. Taniguchi, G. Zhang, A. Bachtold, A. H. MacDonald, and D. K. Efetov, Superconductors, orbital magnets and correlated states in magic-angle bilayer graphene, Nature (London) 574, 653-657 (2019).
  • (5) G. Trambly de Laissardière, D. Mayou, and L. Magaud, Localization of Dirac Electrons in Rotated Graphene Bilayers, Nano Lett. 10, 804 (2010).
  • (6) E. Suárez Morell, J. D. Correa, P. Vargas, M. Pacheco, and Z. Barticevic, Flat bands in slightly twisted bilayer graphene: Tight-binding calculations, Phys. Rev. B 82, 121407(R) (2010).
  • (7) R. Bistritzer and A. H. MacDonald, Moiré bands in twisted double-layer graphene, Proc. Natl. Acad. Sci. U.S.A. 108, 12233-12237 (2011).
  • (8) G. Tarnopolsky, A. J. Kruchkov, and A. Vishwanath, Origin of Magic Angles in Twisted Bilayer Graphene, Phys. Rev. Lett. 122, 106405 (2019).
  • (9) M. Koshino, Band Structure and Topological Properties of Twisted Double Bilayer Graphene, Phys. Rev. B 99, 235406 (2019).
  • (10) N. R. Chebrolu, B. L. Chittari, and J. Jung, Flat bands in twisted double bilayer graphene, Phys. Rev. B 99, 235417 (2019).
  • (11) J. Y. Lee, E. Khalaf, S. Liu, X. Liu, Z. Hao, P. Kim, and A. Vishwanath, Theory of correlated insulating behaviour and spin-triplet superconductivity in twisted double bilayer graphene, Nature Communications 10, 5333 (2019).
  • (12) C. Shen, Y. Chu, Q. Wu, N. Li, S. Wang, Y. Zhao, J. Tang, J. Liu, J. Tian, K. Watanabe et al., Correlated states in twisted double bilayer graphene, Nature Physics 16, 520-525 (2020).
  • (13) X. Liu, Z. Hao, E. Khalaf, J. Y. Lee, Y. Ronen, H. Yoo, D. H. Najafabadi, K. Watanabe, T. Taniguchi, A. Vishwanath, and P. Kim, Tunable spin-polarized correlated states in twisted double bilayer graphene, Nature (London) 583, 221 (2020).
  • (14) M. He, Y. Li, J. Cai, Y. Liu, K. Watanabe, T. Taniguchi, X. Xu, and M. Yankowitz, Symmetry breaking in twisted double bilayer graphene, Nature Physics 17, 26 (2021).
  • (15) J. Shin, B. L. Chittari, Y. Jang, H. Min, and J. Jung, Nearly flat bands in twisted triple bilayer graphene, Phys. Rev. B 105, 245124 (2022).
  • (16) M. H. Naik and M. Jain, Ultraflatbands and Shear Solitons in Moiré Patterns of Twisted Bilayer Transition Metal Dichalcogenides, Phys. Rev. Lett. 121, 266401 (2018).
  • (17) T. Kariyado and A. Vishwanath, Flat band in twisted bilayer Bravais lattices, Phys. Rev. Research 1, 033076 (2019).
  • (18) L. Xian, D. M. Kennes, N. Tancogne-Dejean, M. Altarelli, and A. Rubio, Multiflat Bands and Strong Correlations in Twisted Bilayer Boron Nitride: Doping-Induced Correlated Insulator and Superconductor, Nano Lett. 19, 4934 (2019).
  • (19) G. Chen, A. L. Sharpe, P. Gallagher, I. T. Rosen, E. J. Fox, L. Jiang, B. Lyu, H. Li, K. Watanabe, T. Taniguchi et al., Signatures of tunable superconductivity in a trilayer graphene moiré superlattice, Nature (London) 572, 215–219 (2019).
  • (20) H. S. Arora, R. Polski, Y. Zhang, A. Thomson, Y. Choi, H. Kim, Z. Lin, I. Z. Wilson, X. Xu, J.-H. Chu et al., Superconductivity in metallic twisted bilayer graphene stabilized by WSe2, Nature (London) 583, 379-384 (2020).
  • (21) L. Wang, E.-M. Shih, A. Ghiotto, L. Xian, D. A. Rhodes, C. Tan, M. Claassen, D. M. Kennes, Y. Bai, B. Kim et al., Correlated electronic phases in twisted bilayer transition metal dichalcogenides, Nat. Mater. 19, 861–866 (2020).
  • (22) L. An, X. Cai, D. Pei, M. Huang, Z. Wu, Z. Zhou, J. Lin, Z. Ying, Z. Ye, X. Feng, R. Gao, C. Cacho, M. Watson, Y. Chen, and N. Wang, Interaction effects and superconductivity signatures in twisted double-bilayer WSe2, Nanoscale Horiz. 5, 1309 (2020).
  • (23) E. Khalaf, A. J. Kruchkov, G. Tarnopolsky, and A. Vishwanath, Magic angle hierarchy in twisted graphene multilayers, Phys. Rev. B 100, 085109 (2019).
  • (24) S. Carr, C. Li, Z. Zhu, E. Kaxiras, S. Sachdev, and A. Kruchkov, Ultraheavy and Ultrarelativistic Dirac Quasiparticles in Sandwiched Graphenes, Nano Lett. 20, 3030-3038 (2020).
  • (25) A. Lopez-Bezanilla and J. L. Lado, Electrical band flattening, valley flux, and superconductivity in twisted trilayer graphene, Phys. Rev. Research 2, 033357 (2020).
  • (26) C. Lei, L. Linhart, W. Qin, F. Libisch, and A. H. MacDonald, Mirror symmetry breaking and lateral stacking shifts in twisted trilayer graphene, Phys. Rev. B 104, 035139 (2021).
  • (27) V. T. Phong, P. A. Pantaleón, T. Cea, and F. Guinea, Band structure and superconductivity in twisted trilayer graphene, Phys. Rev. B 104, L121116 (2021).
  • (28) E. Lake and T. Senthil, Reentrant superconductivity through a quantum Lifshitz transition in twisted trilayer graphene, Phys. Rev. B 104, 174505 (2021).
  • (29) Y. W. Choi and H. J. Choi, Dichotomy of Electron-Phonon Coupling in Graphene Moiré Flat Bands, Phys. Rev. Lett. 127, 167001 (2021).
  • (30) V Hung Nguyen, Trinh X Hoang, and J-C Charlier, Electronic properties of twisted multilayer graphene, J. Phys. Mater. 5, 034003 (2022).
  • (31) N. Leconte, Y. Park, J. An, A. Samudrala, and J. Jung, Electronic structure of lattice relaxed alternating twist tNG-multilayer graphene: from few layers to bulk AT-graphite, 2D Mater. 9, 044002 (2022).
  • (32) J. M. Park, Y. Cao, K. Watanabe, T. Taniguchi, and P. Jarillo-Herrero, Tunable strongly coupled superconductivity in magic-angle twisted trilayer graphene, Nature (London) 590, 249 (2021).
  • (33) Z. Hao, A. M. Zimmerman, P. Ledwith, E. Khalaf, D. H. Najafabadi, K. Watanabe, T. Taniguchi, A. Vishwanath, and P. Kim, Electric field–tunable superconductivity in alternating-twist magic-angle trilayer graphene, Science 371, 6534 (2021).
  • (34) Y. Cao, J. M. Park, K. Watanabe, T. Taniguchi, and P. Jarillo-Herrero, Pauli-limit violation and re-entrant superconductivity in moiré graphene, Nature (London) 595, 526-531 (2021).
  • (35) H. Kim, Y. Choi, C. Lewandowski, A. Thomson, Y. Zhang, R. Polski, K. Watanabe, T. Taniguchi, J. Alicea, and S. Nadj-Perge, Evidence for unconventional superconductivity in twisted trilayer graphene, Nature (London) 606, 494 (2022).
  • (36) J. M. Park, Y. Cao, L.-Q. Xia, S. Sun, K. Watanabe, T. Taniguchi, and P. Jarillo-Herrero, Robust superconductivity in magic-angle multilayer graphene family, Nature Materials 21, 877-883 (2022).
  • (37) Y. Zhang, R. Polski, C. Lewandowski, A. Thomson, Y. Peng, Y. Choi, H. Kim, K. Watanabe, T. Taniguchi, J. Alicea, F. von Oppen, G. Refael, and S. Nadj-Perge, Promotion of superconductivity in magic-angle graphene multilayers, Science 377, 1538 (2022).
  • (38) C. Shen, P. J. Ledwith, K. Watanabe, T. Taniguchi, E. Khalaf, A. Vishwanath, and D. K. Efetov, Dirac spectroscopy of strongly correlated phases in twisted trilayer graphene, Nature Materials 22, 316-321 (2023).
  • (39) L. Balents, C. R. Dean, D. K. Efetov, and A. F. Young, Superconductivity and strong correlations in moiré flat bands, Nature Physics 16, 725-733 (2020).
  • (40) K. Uchida, S. Furuya, J.-I. Iwata, and A. Oshiyama, Atomic corrugation and electron localization due to Moiré patterns in twisted bilayer graphenes, Phys. Rev. B 90, 155451 (2014).
  • (41) M. M. van Wijk, A. Schuring, M. I. Katsnelson, and A. Fasolino, Relaxation of moiré patterns for slightly misaligned identical lattices: graphene on graphite, 2D Mater. 2, 034010 (2015).
  • (42) J. Jung, A. Raoux, Z. Qiao, and A. H. MacDonald, Ab initio theory of moiré superlattice bands in layered two-dimensional materials, Phys. Rev. B 89, 205414 (2014).
  • (43) H. Min and A. H. MacDonald, Chiral decomposition in the electronic structure of graphene multilayers, Phys. Rev. B 77, 155416 (2008).
  • (44) H. Min and A. H. MacDonald, Electronic structure of multilayer graphene, Prog. Theor. Phys. Suppl. 176, 227 (2008).
  • (45) P. Moon and M. Koshino, Optical absorption in twisted bilayer graphene, Phys. Rev. B 87, 205404 (2013).
  • (46) Gerald D. Mahan, Many-particle physics (3rd ed.), Springer (2000).
  • (47) H. Min and A. H. MacDonald, Origin of Universal Optical Conductivity and Optical Stacking Sequence Identification in Multilayer Graphene, Phys. Rev. Lett. 103, 067402 (2009).
  • (48) C. J. Tabert and E. J. Nicol, Dynamical conductivity of AA-stacked bilayer graphene, Phys. Rev. B 86, 075439 (2012).