Low-temperature thermal transport in moiré superlattices

Lukas P. A. Krisna [email protected]    Takuto Kawakami    Mikito Koshino Department of Physics, Osaka University, Toyonaka, Osaka 560-0043, Japan
(May 1, 2024)
Abstract

We calculate the thermal conductivity of various moiré bilayer systems using a continuum approach and the semiclassical transport theory. When the twist angle is close to 0, we observe a significant reduction of thermal conductivity in a particular low-temperature regime. This reduction is attributed to a moiré-induced reconstruction of acoustic phonon bands and associated decrease of the group velocity. Conversely, in the zero temperature limit, the thermal conductivity is enhanced by moiré effect, surpassing the original values in non-moiré counterparts. These changes result in a characteristic temperature dependence which deviates from the quadratic behavior in intrinsic two-dimensional systems.

I Introduction

In recent years, significant advancements have emerged in the field of two-dimensional (2D) moiré materials, following the discovery of various exotic phenomena in twisted bilayer systems [1, 2, 3, 4, 5, 6, 7]. The electronic structure of these materials are strongly modulated by the emerged superlattice of the long-range periodic pattern [8, 9, 10, 11, 12, 13, 14, 15, 16]. In a similar manner, moiré patterns also have notable influence on the lattice vibration. In twisted systems, the interlayer moiré potential folds the original monolayer phonon bands into the superlattice Brillouin zone [17, 18, 19, 20, 21, 22]. When the moiré superlattice period is much greater than the atomic scale, spontaneous lattice relaxation renormalizes the phonon modes [23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33], and the acoustic bands are reconstructed with notable band flattening and the appearance of spectral gaps [23, 24, 25]. These changes are expected to have significant influence in the thermal transport phenomena where acoustic phonons are the dominant heat carrier.

Heat transport in 2D materials has attracted notable interest because of its superiority and its sensitivity to heterostructuring. For example, the high thermal conductivity of suspended graphene [34, 35, 36, 37] is significantly decreased in the presence of substrate or other layer due to strong scattering of flexural phonons [38, 39, 40, 41, 42, 43, 44]. Meanwhile, experimental measurements demonstrated further suppression when rotationally stacked graphene layers [45, 46]. Theoretical simulations showed that such a thermal conductivity reduction by a twist can be caused by the enhancement of anharmonic phonon scattering and the redistribution of phonons to higher frequencies [47, 48, 49, 50, 51]. While all these simulations assumed the room temperature and above, the effect of moiré-induced phonon band reconstruction on the thermal properties is expected to be relevant in the low temperature regime, and it remains unknown.

Refer to caption
Figure 1: Top view of atomic structures of (a) twisted bilayer graphene (TBG), (b) twisted graphene/hBN (t-G/hBN), (c) parallel-stacked twisted bilayer MoS2 (t-MoS2 (P)), and (d) antiparallel-stacked twisted bilayer MoS2 (t-MoS2 (AP)). Each yellow circle in (c) and (d) represent overlapped double S atoms at different perpendicular positions.

In this paper, we calculate the thermal conductivity of various representative twisted bilayer systems, including twisted bilayer graphene (TBG) and twisted graphene/hexagonal boron nitride (t-G/hBN), and also twisted bilayer molybdenum disulfide (t-MoS2) in both parallel (P) and antiparallel (AP) stacking arrangement [Fig. 1]. Here, we obtain the phonon band structure by the continuum approach, which has been extensively utilized to describe lattice relaxation and the formation of superlattice mini-bands in the phonon dispersion relation [23, 24, 52, 25]. Then we calculate the thermal conductivity using a semiclassical transport theory in the low temperature regime assuming a constant mean free path. We demonstrate that the flattening of the low-energy phonons bands leads to a significant suppression of thermal conductivity up to 40%, resulting in a characteristic deviation from the generic quadratic temperature dependence of thermal conductivity in two-dimensional system.

This paper is organized as follows. In Sec. II, we introduce the continuum method to describe long-wavelength phonons, and the formulation of semiclassical transport to calculate the thermal conductivity in the low-temperature regime. In Sec. III, we present and discuss the calculated phonon dispersion of each system and the corresponding thermal conductivity. Finally, a brief conclusion is given in Sec. IV.

II Methods

II.1 Lattice geometry

We consider a twisted bilayer system composed of two honeycomb lattice layers with generally different lattice constants, a𝑎aitalic_a and asuperscript𝑎a^{\prime}italic_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT, as illustrated in Fig. 2(a). The layer 2 is stacked on top of the layer 1 with relative rotation angle θ𝜃\thetaitalic_θ around a common honeycomb center. We label the sublattices of layer 1 by A and B, and that of layer 2 by A and B as in Fig. 2(a). The primitive lattice vectors of layer 1 are defined as 𝐚1=a(1,0)subscript𝐚1𝑎10\mathbf{a}_{1}=a(1,0)bold_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_a ( 1 , 0 ) and 𝐚2=a(1/2,3/2)subscript𝐚2𝑎1232\mathbf{a}_{2}=a(1/2,\sqrt{3}/2)bold_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_a ( 1 / 2 , square-root start_ARG 3 end_ARG / 2 ), and those of layer 2 are given by 𝐚i=M^R^𝐚i(i=1,2)subscriptsuperscript𝐚𝑖^𝑀^𝑅subscript𝐚𝑖𝑖12\mathbf{a}^{\prime}_{i}=\hat{M}\hat{R}\,\mathbf{a}_{i}\,(i=1,2)bold_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = over^ start_ARG italic_M end_ARG over^ start_ARG italic_R end_ARG bold_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_i = 1 , 2 ), with rotation matrix R^(θ)^𝑅𝜃\hat{R}(\theta)over^ start_ARG italic_R end_ARG ( italic_θ ) and isotropic expansion matrix M^=(1+ε)I^^𝑀1𝜀^𝐼\hat{M}=(1+\varepsilon)\hat{I}over^ start_ARG italic_M end_ARG = ( 1 + italic_ε ) over^ start_ARG italic_I end_ARG where ε=(aa)/a𝜀superscript𝑎𝑎𝑎\varepsilon=(a^{\prime}-a)/aitalic_ε = ( italic_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_a ) / italic_a. The corresponding reciprocal lattice vectors for layer 1 and 2 are given by 𝐛isubscript𝐛𝑖\mathbf{b}_{i}bold_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and 𝐛isubscriptsuperscript𝐛𝑖\mathbf{b}^{\prime}_{i}bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, respectively, which satisfy 𝐚i𝐛j=𝐚i𝐛j=2πδijsubscript𝐚𝑖subscript𝐛𝑗subscriptsuperscript𝐚𝑖subscriptsuperscript𝐛𝑗2𝜋subscript𝛿𝑖𝑗\mathbf{a}_{i}\cdot\mathbf{b}_{j}=\mathbf{a}^{\prime}_{i}\cdot\mathbf{b}^{% \prime}_{j}=2\pi\delta_{ij}bold_a start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ bold_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = bold_a start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 2 italic_π italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. A long-range moiré interference pattern appears due to a slight mismatch from a small difference in lattice constant or small twist angle. The reciprocal lattice vectors of the pattern are given by 𝐆iM=𝐛i𝐛isubscriptsuperscript𝐆M𝑖subscript𝐛𝑖subscriptsuperscript𝐛𝑖\mathbf{G}^{\rm M}_{i}=\mathbf{b}_{i}-\mathbf{b}^{\prime}_{i}bold_G start_POSTSUPERSCRIPT roman_M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = bold_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, while the corresponding real-space lattice vectors are obtained from the condition 𝐋iM𝐆jM=2πδijsubscriptsuperscript𝐋M𝑖subscriptsuperscript𝐆M𝑗2𝜋subscript𝛿𝑖𝑗\mathbf{L}^{\rm M}_{i}\cdot\mathbf{G}^{\rm M}_{j}=2\pi\delta_{ij}bold_L start_POSTSUPERSCRIPT roman_M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ bold_G start_POSTSUPERSCRIPT roman_M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = 2 italic_π italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT. The moiré period LM=|𝐋iM|subscript𝐿Msubscriptsuperscript𝐋M𝑖L_{\rm M}=|\mathbf{L}^{\rm M}_{i}|italic_L start_POSTSUBSCRIPT roman_M end_POSTSUBSCRIPT = | bold_L start_POSTSUPERSCRIPT roman_M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | can be expressed as

LM=a1+εε2+2(1+ε)(1cosθ).subscript𝐿M𝑎1𝜀superscript𝜀221𝜀1𝜃L_{\rm M}=a\frac{1+\varepsilon}{\sqrt{\varepsilon^{2}+2(1+\varepsilon)(1-\cos{% \theta})}}.italic_L start_POSTSUBSCRIPT roman_M end_POSTSUBSCRIPT = italic_a divide start_ARG 1 + italic_ε end_ARG start_ARG square-root start_ARG italic_ε start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 ( 1 + italic_ε ) ( 1 - roman_cos italic_θ ) end_ARG end_ARG . (1)
Refer to caption
Figure 2: (a) Schematic diagram of a twisted bilayer system. (b) Three local stacking structures AA, AB, and BA. (c) Non-relaxed atomic structure of t-G/hBN with θ=1.25𝜃superscript1.25\theta=1.25^{\circ}italic_θ = 1.25 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT. The inset shows the first moiré Brillouin zone.

For honeycomb lattice components, we take graphene, hexagonal boron nitride, and molybdenum disulfide with lattice constant of a𝑎absenta\approxitalic_a ≈ 0.246 nm, 0.2504 nm, and 0.317 nm, respectively. In this paper, we consider twisted bilayer graphene (TBG) and twisted bilayer molybdenum disulfide (t-MoS2) as examples of homo-bilayer (ε=1𝜀1\varepsilon=1italic_ε = 1) and also twisted graphene/hexagonal boron nitride (t-G/hBN) as a heterobilayer (ε1𝜀1\varepsilon\neq 1italic_ε ≠ 1). In Fig. 2(c), we illustrate the formation of moiré superlattice in t-G/hBN with θ=1.25𝜃superscript1.25\theta=1.25^{\circ}italic_θ = 1.25 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT where a lattice constant difference ε1.8%𝜀percent1.8\varepsilon\approx 1.8\%italic_ε ≈ 1.8 % and the twist angle produce a moiré pattern of LM=|𝐋iM|8.8subscript𝐿Msubscriptsuperscript𝐋M𝑖8.8L_{\rm M}=|\mathbf{L}^{\rm M}_{i}|\approx 8.8italic_L start_POSTSUBSCRIPT roman_M end_POSTSUBSCRIPT = | bold_L start_POSTSUPERSCRIPT roman_M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | ≈ 8.8 nm.

Across the moiré pattern, the local stacking structure changes smoothly at the atomic scale. At a given position 𝐫𝐫\mathbf{r}bold_r, it is characterized by the phase difference (φ1,φ2)subscript𝜑1subscript𝜑2(\varphi_{1},\varphi_{2})( italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) defined as

φj(𝐫)=(𝐛j𝐛j)𝐫=𝐆jM𝐫.subscript𝜑𝑗𝐫subscript𝐛𝑗subscriptsuperscript𝐛𝑗𝐫subscriptsuperscript𝐆M𝑗𝐫\varphi_{j}(\mathbf{r})=(\mathbf{b}_{j}-\mathbf{b}^{\prime}_{j})\cdot\mathbf{r% }=\mathbf{G}^{\text{M}}_{j}\cdot\mathbf{r}.italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_r ) = ( bold_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ⋅ bold_r = bold_G start_POSTSUPERSCRIPT M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⋅ bold_r . (2)

For example, (φ1,φ2)=(0,0)subscript𝜑1subscript𝜑200(\varphi_{1},\varphi_{2})=(0,0)( italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = ( 0 , 0 ), (2π/3,2π/3)2𝜋32𝜋3(2\pi/3,2\pi/3)( 2 italic_π / 3 , 2 italic_π / 3 ) and (4π/3,4π/3)4𝜋34𝜋3(4\pi/3,4\pi/3)( 4 italic_π / 3 , 4 italic_π / 3 ) correspond to AA (complete alignment of the honeycomb lattices), AB (B-site of layer 2 on top of A-site of layer 1), and BA (A-site of layer 2 on top of B-site of layer 1), respectively [Fig. 2(b)]. We note that, when stacking two monolayers with different sublattice atoms, e.g., t-MoS2, there are two possible stacking configuration, parallel (P) and antiparallel (AP), which are related by 180 rotation of layer 1. In the parallel stacking, the two layers have identical atoms at A and A (B and B), whereas in the antiparallel stacking, the two layers have different atoms on the same sublattice sites.

II.2 Continuum model

We describe the in-plane lattice vibration in the twisted bilayer systems using a continuum approach [23, 24, 25]. In this framework, we consider a smoothly-varying displacement field 𝐮(l)(𝐫,t)superscript𝐮𝑙𝐫𝑡\mathbf{u}^{(l)}(\mathbf{r},t)bold_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT ( bold_r , italic_t ) at layer l(=1,2)l(=1,2)italic_l ( = 1 , 2 ), which represents the shift of the atom at 𝐫𝐫\mathbf{r}bold_r in the original rigid lattice, and express the Lagrangian as a functional of 𝐮(l)(𝐫,t)superscript𝐮𝑙𝐫𝑡\mathbf{u}^{(l)}(\mathbf{r},t)bold_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT ( bold_r , italic_t ). Here we only consider the in-plane atomic displacement 𝐮(l)=(ux(l),uy(l))superscript𝐮𝑙superscriptsubscript𝑢𝑥𝑙superscriptsubscript𝑢𝑦𝑙\mathbf{u}^{(l)}=(u_{x}^{(l)},u_{y}^{(l)})bold_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT = ( italic_u start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT , italic_u start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT ) neglecting the out-of-plane component. In the presence of the lattice distortion, the phase difference [Eq. (2)] is modified as

φj(𝐫,t)=𝐆jM[𝐫𝐮+(𝐫,t)/2]+𝐛¯j𝐮(𝐫,t),subscript𝜑𝑗𝐫𝑡subscriptsuperscript𝐆M𝑗delimited-[]𝐫superscript𝐮𝐫𝑡2subscript¯𝐛𝑗superscript𝐮𝐫𝑡\varphi_{j}(\mathbf{r},t)=\mathbf{G}^{\text{M}}_{j}\cdot\left[\mathbf{r}-% \mathbf{u}^{+}(\mathbf{r},t)/2\right]+\bar{\mathbf{b}}_{j}\cdot\mathbf{u}^{-}(% \mathbf{r},t),italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_r , italic_t ) = bold_G start_POSTSUPERSCRIPT M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⋅ [ bold_r - bold_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( bold_r , italic_t ) / 2 ] + over¯ start_ARG bold_b end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⋅ bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( bold_r , italic_t ) , (3)

where we have defined the (interlayer) symmetric and antisymmetric displacements 𝐮±=𝐮(2)±𝐮(1)superscript𝐮plus-or-minusplus-or-minussuperscript𝐮2superscript𝐮1\mathbf{u}^{\pm}=\mathbf{u}^{(2)}\pm\mathbf{u}^{(1)}bold_u start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT = bold_u start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ± bold_u start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT, and 𝐛¯j=(𝐛j+𝐛j)/2subscript¯𝐛𝑗subscript𝐛𝑗subscriptsuperscript𝐛𝑗2\bar{\mathbf{b}}_{j}=(\mathbf{b}_{j}+\mathbf{b}^{\prime}_{j})/2over¯ start_ARG bold_b end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = ( bold_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + bold_b start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) / 2.

The Lagrangian is explicitly written as L=T(UE+UB)𝐿𝑇subscript𝑈𝐸subscript𝑈𝐵L=T-(U_{E}+U_{B})italic_L = italic_T - ( italic_U start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT + italic_U start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ), where T𝑇Titalic_T is the kinetic energy, UEsubscript𝑈𝐸U_{E}italic_U start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT is the elastic energy, and UBsubscript𝑈𝐵U_{B}italic_U start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the interlayer binding energy. The UBsubscript𝑈𝐵U_{B}italic_U start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is defined by

UB=V[φ1(𝐫,t),φ2(𝐫,t)]d2𝐫,subscript𝑈𝐵𝑉subscript𝜑1𝐫𝑡subscript𝜑2𝐫𝑡superscript𝑑2𝐫U_{B}=\int V[\varphi_{1}(\mathbf{r},t),\varphi_{2}(\mathbf{r},t)]\,d^{2}% \mathbf{r},italic_U start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = ∫ italic_V [ italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_r , italic_t ) , italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( bold_r , italic_t ) ] italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r , (4)

where V[φ1,φ2]𝑉subscript𝜑1subscript𝜑2V[\varphi_{1},\varphi_{2}]italic_V [ italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] is the interlayer binding energy per area of the non-rotated bilayer with an interlayer registry specified by φisubscript𝜑𝑖\varphi_{i}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. By 120 symmetry of the system, it should be written within the lowest harmonics as

V[φ1,φ2]=j=132V0cos(φj+φ0),𝑉subscript𝜑1subscript𝜑2superscriptsubscript𝑗132subscript𝑉0subscript𝜑𝑗subscript𝜑0V[\varphi_{1},\varphi_{2}]=\sum_{j=1}^{3}2V_{0}\cos\left(\varphi_{j}+\varphi_{% 0}\right),italic_V [ italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT 2 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos ( italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , (5)

where φ3=φ1φ2subscript𝜑3subscript𝜑1subscript𝜑2\varphi_{3}=-\varphi_{1}-\varphi_{2}italic_φ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = - italic_φ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_φ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and the constant energy offset is neglected. The energy of AA, AB, and BA bilayer stacking is then given by VAA=6V0cos(φ0)subscript𝑉𝐴𝐴6subscript𝑉0subscript𝜑0V_{AA}=6V_{0}\cos(\varphi_{0})italic_V start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT = 6 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos ( italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), VAB=6V0cos(φ0+23π)subscript𝑉𝐴𝐵6subscript𝑉0subscript𝜑023𝜋V_{AB}=6V_{0}\cos(\varphi_{0}+\tfrac{2}{3}\pi)italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT = 6 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos ( italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG 2 end_ARG start_ARG 3 end_ARG italic_π ), and VBA=6V0cos(φ023π)subscript𝑉𝐵𝐴6subscript𝑉0subscript𝜑023𝜋V_{BA}=6V_{0}\cos(\varphi_{0}-\tfrac{2}{3}\pi)italic_V start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT = 6 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos ( italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - divide start_ARG 2 end_ARG start_ARG 3 end_ARG italic_π ). Here we obtain the parameters (V0,φ0)subscript𝑉0subscript𝜑0(V_{0},\varphi_{0})( italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) from the relative values of VAAsubscript𝑉𝐴𝐴V_{AA}italic_V start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT, VABsubscript𝑉𝐴𝐵V_{AB}italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT, and VBAsubscript𝑉𝐵𝐴V_{BA}italic_V start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT found in the literature. Table 1 lists the values of (V0,φ0)subscript𝑉0subscript𝜑0(V_{0},\varphi_{0})( italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) and the corresponding (VAA,VAB,VBA)subscript𝑉𝐴𝐴subscript𝑉𝐴𝐵subscript𝑉𝐵𝐴(V_{AA},V_{AB},V_{BA})( italic_V start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT ) for the considered systems in this paper.

Table 1: Parameters for interlayer binding energy, (V0,φ0)subscript𝑉0subscript𝜑0(V_{0},\varphi_{0})( italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), and the corresponding (VAA,VAB,VBA)subscript𝑉𝐴𝐴subscript𝑉𝐴𝐵subscript𝑉𝐵𝐴(V_{AA},V_{AB},V_{BA})( italic_V start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT ) for the considered systems in this paper. Note that the origin of interlayer binding energy is arbitrary.
V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (eV/nm2) φ0subscript𝜑0\varphi_{0}italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT VAAsubscript𝑉𝐴𝐴V_{AA}italic_V start_POSTSUBSCRIPT italic_A italic_A end_POSTSUBSCRIPT (eV/nm2) VABsubscript𝑉𝐴𝐵V_{AB}italic_V start_POSTSUBSCRIPT italic_A italic_B end_POSTSUBSCRIPT (eV/nm2) VBAsubscript𝑉𝐵𝐴V_{BA}italic_V start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT (eV/nm2)
TBG111Ref. [53, 54] 0.1600.1600.1600.160 00 0.9610.9610.9610.961 0.4810.481-0.481- 0.481 0.4810.481-0.481- 0.481
t-G/hBN222Ref. [55] 0.2020.2020.2020.202 0.9560.9560.9560.956 0.7000.7000.7000.700 1.2081.208-1.208- 1.208 0.5090.5090.5090.509
t-MoS2 (P)333Ref. [56, 57] 0.08890.08890.08890.0889 00 0.5330.5330.5330.533 0.2670.267-0.267- 0.267 0.2670.267-0.267- 0.267
t-MoS2 (AP)333Ref. [56, 57] 0.08010.0801-0.0801- 0.0801 0.8050.805-0.805- 0.805 0.3330.333-0.333- 0.333 0.1330.133-0.133- 0.133 0.4670.4670.4670.467

The energy cost for intralayer distortion, UEsubscript𝑈𝐸U_{E}italic_U start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT is given by standard expression of elastic theory [58, 59],

UE=l=1212subscript𝑈𝐸superscriptsubscript𝑙1212\displaystyle U_{E}=\sum_{l=1}^{2}\frac{1}{2}\intitalic_U start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ (λ(l)+μ(l))(uxx(l)+uyy(l))2superscript𝜆𝑙superscript𝜇𝑙superscriptsubscriptsuperscript𝑢𝑙𝑥𝑥subscriptsuperscript𝑢𝑙𝑦𝑦2\displaystyle(\lambda^{(l)}+\mu^{(l)})\left(u^{(l)}_{xx}+u^{(l)}_{yy}\right)^{2}( italic_λ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT + italic_μ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT ) ( italic_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT + italic_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
+μ(l)[(uxx(l)uyy(l))2+4(uxy(l))2]d2𝐫,superscript𝜇𝑙delimited-[]superscriptsubscriptsuperscript𝑢𝑙𝑥𝑥subscriptsuperscript𝑢𝑙𝑦𝑦24superscriptsubscriptsuperscript𝑢𝑙𝑥𝑦2superscript𝑑2𝐫\displaystyle+\mu^{(l)}\left[\left(u^{(l)}_{xx}-u^{(l)}_{yy}\right)^{2}+4\left% (u^{(l)}_{xy}\right)^{2}\right]d^{2}\mathbf{r},+ italic_μ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT [ ( italic_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT - italic_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 ( italic_u start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r , (6)

where uij(l)=(iuj(l)+jui(l))/2superscriptsubscript𝑢𝑖𝑗𝑙subscript𝑖superscriptsubscript𝑢𝑗𝑙subscript𝑗superscriptsubscript𝑢𝑖𝑙2u_{ij}^{(l)}=(\partial_{i}u_{j}^{(l)}+\partial_{j}u_{i}^{(l)})/2italic_u start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT = ( ∂ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT + ∂ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT ) / 2 is the strain tensor, and λ(l)superscript𝜆𝑙\lambda^{(l)}italic_λ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT and μ(l)superscript𝜇𝑙\mu^{(l)}italic_μ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT are the Lamé parameters for layer l𝑙litalic_l which are given in Table 2.

Table 2: Lamé parameter and mass density used in the calculation.
λ𝜆\lambdaitalic_λ (eV/Å2) μ𝜇\muitalic_μ (eV/Å2) ρ𝜌\rhoitalic_ρ (108superscript10810^{-8}10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT g/cm2)
Graphene111Ref. [60, 16] 3.253.253.253.25 9.579.579.579.57 7.617.617.617.61
hBN222Ref. [61, 16] 3.53.53.53.5 7.87.87.87.8 7.597.597.597.59
MoS2333Ref. [56, 57] 4.234.234.234.23 4.234.234.234.23 30.530.530.530.5

The kinetic energy is expressed as

T=l=1212ρ(l)(u˙x(l)2+u˙y(l)2)d2𝐫,𝑇superscriptsubscript𝑙1212superscript𝜌𝑙superscriptsubscript˙𝑢𝑥𝑙2superscriptsubscript˙𝑢𝑦𝑙2superscript𝑑2𝐫T=\sum_{l=1}^{2}\int\frac{1}{2}\rho^{(l)}\left(\dot{u}_{x}^{(l)2}+\dot{u}_{y}^% {(l)2}\right)d^{2}\mathbf{r},italic_T = ∑ start_POSTSUBSCRIPT italic_l = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∫ divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ρ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT ( over˙ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) 2 end_POSTSUPERSCRIPT + over˙ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_l ) 2 end_POSTSUPERSCRIPT ) italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_r , (7)

where ρ(l)superscript𝜌𝑙\rho^{(l)}italic_ρ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT is the mass density of layer l𝑙litalic_l. The mass density for graphene, hBN, and MoS2 are given in Table 2. For the layer convenience, we define averaged materials parameters

λ=12(λ(1)+λ(2)),μ=12(μ(1)+μ(2)),formulae-sequence𝜆12superscript𝜆1superscript𝜆2𝜇12superscript𝜇1superscript𝜇2\displaystyle\lambda=\frac{1}{2}(\lambda^{(1)}+\lambda^{(2)}),\quad\mu=\frac{1% }{2}(\mu^{(1)}+\mu^{(2)}),italic_λ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_λ start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT + italic_λ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ) , italic_μ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_μ start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT + italic_μ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ) ,
ρ=12(ρ(1)+ρ(2)).𝜌12superscript𝜌1superscript𝜌2\displaystyle\rho=\frac{1}{2}(\rho^{(1)}+\rho^{(2)}).italic_ρ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_ρ start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT + italic_ρ start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ) . (8)

The equation of motion is given by the Euler-Lagrange equation for the Lagrangian L𝐿Litalic_L. It can be effectively decoupled in terms of the symmetric and the antisymmetric components 𝐮±superscript𝐮plus-or-minus\mathbf{u}^{\pm}bold_u start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT [23, 25]. The hybridization between 𝐮±superscript𝐮plus-or-minus\mathbf{u}^{\pm}bold_u start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT is caused by the difference of the parameters ρ(l),λ(l),μ(l)superscript𝜌𝑙superscript𝜆𝑙superscript𝜇𝑙\rho^{(l)},\lambda^{(l)},\mu^{(l)}italic_ρ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT , italic_λ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT , italic_μ start_POSTSUPERSCRIPT ( italic_l ) end_POSTSUPERSCRIPT between two layers, and therefore it is present only in heterobilayers [25]. In bilayers composed of materials with closely matched physical parameters, such as t-G/hBN, these effects are minor and will be neglected in the following argument.

The effect of the moiré interlayer coupling is much more significant in the antisymmetric part 𝐮superscript𝐮\mathbf{u}^{-}bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT than in the symmetric part 𝐮+superscript𝐮\mathbf{u}^{+}bold_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, as the phase difference φisubscript𝜑𝑖\varphi_{i}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT in Eq. (3) is more sensitive to 𝐮superscript𝐮\mathbf{u}^{-}bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT because of |𝐛j¯||𝐆jM|much-greater-than¯subscript𝐛𝑗subscriptsuperscript𝐆M𝑗|\bar{\mathbf{b}_{j}}|\gg|\mathbf{G}^{\rm M}_{j}|| over¯ start_ARG bold_b start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG | ≫ | bold_G start_POSTSUPERSCRIPT roman_M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT | [25]. By neglecting 𝐮+superscript𝐮\mathbf{u}^{+}bold_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT in φisubscript𝜑𝑖\varphi_{i}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, the equation of motion for 𝐮+superscript𝐮\mathbf{u}^{+}bold_u start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT becomes equivalent to that for the single-layer honeycomb lattice with the averaged parameters of Eq. (II.2). As a result, the symmetric phonon modes are simply represented by the longitudinal (LA) and transverse (TA) acoustic modes with phonon velocity

vL=λ+2μρandvT=μρ,formulae-sequencesubscript𝑣L𝜆2𝜇𝜌andsubscript𝑣T𝜇𝜌v_{\rm L}=\sqrt{\frac{\lambda+2\mu}{\rho}}\quad\text{and}\quad v_{\rm T}=\sqrt% {\frac{\mu}{\rho}},italic_v start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_λ + 2 italic_μ end_ARG start_ARG italic_ρ end_ARG end_ARG and italic_v start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG italic_μ end_ARG start_ARG italic_ρ end_ARG end_ARG , (9)

respectively.

For the antisymmetric part 𝐮superscript𝐮\mathbf{u}^{-}bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, we consider the solution for in the form of

𝐮(𝐫,t)=𝐮0(𝐫)+δ𝐮(𝐫,t),superscript𝐮𝐫𝑡subscriptsuperscript𝐮0𝐫𝛿superscript𝐮𝐫𝑡\mathbf{u}^{-}(\mathbf{r},t)=\mathbf{u}^{-}_{0}(\mathbf{r})+\delta\mathbf{u}^{% -}(\mathbf{r},t),bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( bold_r , italic_t ) = bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) + italic_δ bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( bold_r , italic_t ) , (10)

where 𝐮0(𝐫)subscriptsuperscript𝐮0𝐫\mathbf{u}^{-}_{0}(\mathbf{r})bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) gives the relaxed state to minimize UB+UEsubscript𝑈𝐵subscript𝑈𝐸U_{B}+U_{E}italic_U start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT + italic_U start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and δ𝐮(𝐫,t)𝛿superscript𝐮𝐫𝑡\delta\mathbf{u}^{-}(\mathbf{r},t)italic_δ bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( bold_r , italic_t ) is a small vibration around 𝐮0subscriptsuperscript𝐮0\mathbf{u}^{-}_{0}bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The relaxed displacement field 𝐮0(𝐫)subscriptsuperscript𝐮0𝐫\mathbf{u}^{-}_{0}(\mathbf{r})bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) is obtained by self-consistent iteration [59]. Figure 3 presents contour plots of the interlayer binding energy in the relaxed state for the considered moiré systems. Each system exhibits a characteristic domain pattern, which is determined by relative stabilities among AA, AB and BA stacking configurations (Table 1). In TBG [Fig. 3(a)] and parallel t-MoS2 [Fig. 3(c)], the relaxed structure reveals a triangular pattern comprising AB and BA stacking regions, indicative of the energetic equivalence between these configurations [59, 62]. In t-G/hBN [Fig. 3(b)], on the other hand, only the AB stacking dominates in the relaxed structure as it is the only structure with the lowest energy, resulting in a honeycomb domain pattern [55, 63]. The antiparallel t-MoS2 [Fig. 3(d)] exhibits a honeycomb structure reminiscent of t-G/hBN, albeit with a distortion of the hexagonal domains into triangles. This distortion arises from the relatively small energy difference between the most stable AA stacking and the second stable AB stacking (see, Table 1), resulting in a significant broadening of domain walls around the AB position [56, 57].

The vibration modes around the relaxed structure can be obtained as follows. We define the Fourier transform of 𝐮0(𝐫)subscriptsuperscript𝐮0𝐫\mathbf{u}^{-}_{0}(\mathbf{r})bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) and δ𝐮(𝐫,t)𝛿superscript𝐮𝐫𝑡\delta\mathbf{u}^{-}(\mathbf{r},t)italic_δ bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( bold_r , italic_t ) as,

𝐮0(𝐫)subscriptsuperscript𝐮0𝐫\displaystyle\mathbf{u}^{-}_{0}(\mathbf{r})bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) =𝐆𝐮0,𝐆ei𝐆𝐫,absentsubscript𝐆subscriptsuperscript𝐮0𝐆superscript𝑒𝑖𝐆𝐫\displaystyle=\sum_{\mathbf{G}}\mathbf{u}^{-}_{0,\mathbf{G}}e^{i\mathbf{G}% \cdot\mathbf{r}},= ∑ start_POSTSUBSCRIPT bold_G end_POSTSUBSCRIPT bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 , bold_G end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i bold_G ⋅ bold_r end_POSTSUPERSCRIPT , (11)
δ𝐮(𝐫,t)𝛿superscript𝐮𝐫𝑡\displaystyle\delta\mathbf{u}^{-}(\mathbf{r},t)italic_δ bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( bold_r , italic_t ) =1S𝐆𝐪δ𝐮𝐪+𝐆(t)ei(𝐪+𝐆)𝐫,absent1𝑆subscript𝐆subscript𝐪𝛿superscriptsubscript𝐮𝐪𝐆𝑡superscript𝑒𝑖𝐪𝐆𝐫\displaystyle=\frac{1}{\sqrt{S}}\sum_{\mathbf{G}}\sum_{\mathbf{q}}\delta% \mathbf{u}_{\mathbf{q}+\mathbf{G}}^{-}(t)e^{i(\mathbf{q}+\mathbf{G})\cdot% \mathbf{r}},= divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_S end_ARG end_ARG ∑ start_POSTSUBSCRIPT bold_G end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT italic_δ bold_u start_POSTSUBSCRIPT bold_q + bold_G end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_t ) italic_e start_POSTSUPERSCRIPT italic_i ( bold_q + bold_G ) ⋅ bold_r end_POSTSUPERSCRIPT , (12)

where 𝐆=m𝐆1M+n𝐆2M𝐆𝑚subscriptsuperscript𝐆M1𝑛subscriptsuperscript𝐆M2\mathbf{G}=m\mathbf{G}^{\text{M}}_{1}+n\mathbf{G}^{\text{M}}_{2}bold_G = italic_m bold_G start_POSTSUPERSCRIPT M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_n bold_G start_POSTSUPERSCRIPT M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the moiré reciprocal lattice vectors, 𝐪𝐪\mathbf{q}bold_q is the wave vector within MBZ, and S𝑆Sitalic_S is the system’s total area as normalization factor. The equation of motion in the momentum-space is then given as

ρ2d2dt2δ𝐮𝐪+𝐆=𝐆D^𝐪(𝐆,𝐆)δ𝐮𝐪+𝐆,𝜌2superscript𝑑2𝑑superscript𝑡2𝛿superscriptsubscript𝐮𝐪𝐆subscriptsuperscript𝐆subscript^𝐷𝐪𝐆superscript𝐆𝛿subscriptsuperscript𝐮𝐪superscript𝐆\frac{\rho}{2}\frac{d^{2}}{dt^{2}}\delta\mathbf{u}_{\mathbf{q}+\mathbf{G}}^{-}% =-\sum_{\mathbf{G}^{\prime}}\hat{D}_{\mathbf{q}}(\mathbf{G},\mathbf{G}^{\prime% })\delta\mathbf{u}^{-}_{\mathbf{q}+\mathbf{G}^{\prime}},divide start_ARG italic_ρ end_ARG start_ARG 2 end_ARG divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_δ bold_u start_POSTSUBSCRIPT bold_q + bold_G end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT = - ∑ start_POSTSUBSCRIPT bold_G start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT ( bold_G , bold_G start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_δ bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_q + bold_G start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (13)

where

D^𝐪(𝐆,𝐆)=(1/2)K^𝐪+𝐆δ𝐆,𝐆+V^𝐆𝐆subscript^𝐷𝐪𝐆superscript𝐆12subscript^𝐾𝐪𝐆subscript𝛿𝐆superscript𝐆subscript^𝑉superscript𝐆𝐆\hat{D}_{\mathbf{q}}(\mathbf{G},\mathbf{G}^{\prime})=(1/2)\hat{K}_{\mathbf{q}+% \mathbf{G}}\delta_{\mathbf{G},\mathbf{G}^{\prime}}+\hat{V}_{\mathbf{G}^{\prime% }-\mathbf{G}}over^ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT ( bold_G , bold_G start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ( 1 / 2 ) over^ start_ARG italic_K end_ARG start_POSTSUBSCRIPT bold_q + bold_G end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT bold_G , bold_G start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT + over^ start_ARG italic_V end_ARG start_POSTSUBSCRIPT bold_G start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - bold_G end_POSTSUBSCRIPT (14)

is the dynamical matrix. The matrix K^^𝐾\hat{K}over^ start_ARG italic_K end_ARG is defined as

K^𝐪=((λ+2μ)qx2+μqy2(λ+μ)qxqy(λ+μ)qxqy(λ+2μ)qy2+μqx2),subscript^𝐾𝐪matrix𝜆2𝜇superscriptsubscript𝑞𝑥2𝜇superscriptsubscript𝑞𝑦2𝜆𝜇subscript𝑞𝑥subscript𝑞𝑦𝜆𝜇subscript𝑞𝑥subscript𝑞𝑦𝜆2𝜇superscriptsubscript𝑞𝑦2𝜇superscriptsubscript𝑞𝑥2\displaystyle\hat{K}_{\mathbf{q}}=\begin{pmatrix}(\lambda+2\mu)q_{x}^{2}+\mu q% _{y}^{2}&(\lambda+\mu)q_{x}q_{y}\\ (\lambda+\mu)q_{x}q_{y}&(\lambda+2\mu)q_{y}^{2}+\mu q_{x}^{2}\end{pmatrix},over^ start_ARG italic_K end_ARG start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT = ( start_ARG start_ROW start_CELL ( italic_λ + 2 italic_μ ) italic_q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_μ italic_q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL ( italic_λ + italic_μ ) italic_q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL ( italic_λ + italic_μ ) italic_q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT end_CELL start_CELL ( italic_λ + 2 italic_μ ) italic_q start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_μ italic_q start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ) , (15)

and the matrix V^^𝑉\hat{V}over^ start_ARG italic_V end_ARG is defined as

V^𝐆=2V0j=13h𝐆j(b¯j,xb¯j,xb¯j,xb¯j,yb¯j,yb¯j,xb¯j,yb¯j,y),subscript^𝑉𝐆2subscript𝑉0superscriptsubscript𝑗13superscriptsubscript𝐆𝑗matrixsubscript¯𝑏𝑗𝑥subscript¯𝑏𝑗𝑥subscript¯𝑏𝑗𝑥subscript¯𝑏𝑗𝑦subscript¯𝑏𝑗𝑦subscript¯𝑏𝑗𝑥subscript¯𝑏𝑗𝑦subscript¯𝑏𝑗𝑦\hat{V}_{\mathbf{G}}=-2V_{0}\sum_{j=1}^{3}h_{\mathbf{G}}^{j}\begin{pmatrix}% \bar{b}_{j,x}\bar{b}_{j,x}&\bar{b}_{j,x}\bar{b}_{j,y}\\ \bar{b}_{j,y}\bar{b}_{j,x}&\bar{b}_{j,y}\bar{b}_{j,y}\end{pmatrix},over^ start_ARG italic_V end_ARG start_POSTSUBSCRIPT bold_G end_POSTSUBSCRIPT = - 2 italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_h start_POSTSUBSCRIPT bold_G end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT ( start_ARG start_ROW start_CELL over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_x end_POSTSUBSCRIPT over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_x end_POSTSUBSCRIPT end_CELL start_CELL over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_x end_POSTSUBSCRIPT over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_y end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_y end_POSTSUBSCRIPT over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_x end_POSTSUBSCRIPT end_CELL start_CELL over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_y end_POSTSUBSCRIPT over¯ start_ARG italic_b end_ARG start_POSTSUBSCRIPT italic_j , italic_y end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) , (16)

with

cos[𝐆jM𝐫+𝐛¯j𝐮0(𝐫)+φ0]=𝐆h𝐆jei𝐆𝐫.subscriptsuperscript𝐆M𝑗𝐫subscript¯𝐛𝑗subscriptsuperscript𝐮0𝐫subscript𝜑0subscript𝐆superscriptsubscript𝐆𝑗superscript𝑒𝑖𝐆𝐫\cos\left[\mathbf{G}^{\text{M}}_{j}\cdot\mathbf{r}+\bar{\mathbf{b}}_{j}\cdot% \mathbf{u}^{-}_{0}(\mathbf{r})+\varphi_{0}\right]=\sum_{\mathbf{G}}h_{\mathbf{% G}}^{j}e^{i\mathbf{G}\cdot\mathbf{r}}.roman_cos [ bold_G start_POSTSUPERSCRIPT M end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⋅ bold_r + over¯ start_ARG bold_b end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⋅ bold_u start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) + italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ] = ∑ start_POSTSUBSCRIPT bold_G end_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT bold_G end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_i bold_G ⋅ bold_r end_POSTSUPERSCRIPT . (17)

Equation (13) is solved by diagonalizing the dynamical matrix D^𝐪subscript^𝐷𝐪\hat{D}_{\mathbf{q}}over^ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT to obtain the phonon frequency ωn,𝐪subscript𝜔𝑛𝐪\omega_{n,\mathbf{q}}italic_ω start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT of mode n𝑛nitalic_n at 𝐪𝐪\mathbf{q}bold_q. In solving this, we take sufficient cut-off in the reciprocal space, so that the result converges.

Refer to caption
Figure 3: Contour plot of the interlayer binding energy V(𝐫)𝑉𝐫V(\mathbf{r})italic_V ( bold_r ) in the relaxed structure of (a) 0.817 TBG, (b) 0 t-G/hBN, (c) 0.8 t-MoS2 (P), and (d) 0.8 t-MoS2 (AP).

II.3 Thermal transport theory

Based on linearized Boltzmann transport theory with relaxation time approximation [64], thermal conductivity of 2D isotropic material can be expressed as

κ=1Sn,q12vn,q2τn,qωn,qf0(ωn,q)T,𝜅1𝑆subscript𝑛q12subscriptsuperscript𝑣2𝑛qsubscript𝜏𝑛qPlanck-constant-over-2-pisubscript𝜔𝑛qsubscript𝑓0subscript𝜔𝑛q𝑇\kappa=\frac{1}{S}\sum_{n,\textbf{q}}\frac{1}{2}v^{2}_{n,\textbf{q}}\tau_{n,% \textbf{q}}\hbar\omega_{n,\textbf{q}}\frac{\partial f_{0}(\omega_{n,\textbf{q}% })}{\partial T},italic_κ = divide start_ARG 1 end_ARG start_ARG italic_S end_ARG ∑ start_POSTSUBSCRIPT italic_n , q end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n , q end_POSTSUBSCRIPT italic_τ start_POSTSUBSCRIPT italic_n , q end_POSTSUBSCRIPT roman_ℏ italic_ω start_POSTSUBSCRIPT italic_n , q end_POSTSUBSCRIPT divide start_ARG ∂ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_n , q end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_T end_ARG , (18)

where S𝑆Sitalic_S is the total area of the system, f0(ω)=1/[exp(ω/kBT)1]subscript𝑓0𝜔1delimited-[]Planck-constant-over-2-pi𝜔subscript𝑘𝐵𝑇1f_{0}(\omega)=1/[\exp(\hbar\omega/k_{B}T)-1]italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω ) = 1 / [ roman_exp ( roman_ℏ italic_ω / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ) - 1 ] is the Bose-Einstein distribution function, vn,𝐪=|𝐪ωn,𝐪|subscript𝑣𝑛𝐪subscriptbold-∇𝐪subscript𝜔𝑛𝐪v_{n,\mathbf{q}}=|\bm{\nabla}_{\mathbf{q}}\omega_{n,\mathbf{q}}|italic_v start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT = | bold_∇ start_POSTSUBSCRIPT bold_q end_POSTSUBSCRIPT italic_ω start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT | is the phonon velocity, τn,𝐪subscript𝜏𝑛𝐪\tau_{n,\mathbf{q}}italic_τ start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT is the relaxation time, and the summation is taken over all of the mode index n𝑛nitalic_n and wave vector 𝐪𝐪\mathbf{q}bold_q. Here, we note that the isotropic behavior is a consequence of the three-fold rotation symmetry.

The relaxation time, τn,𝐪subscript𝜏𝑛𝐪\tau_{n,\mathbf{q}}italic_τ start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT, describes various scattering mechanisms that limit the mean free path of the phonon which is defined as Λn,𝐪τn,𝐪vn,𝐪subscriptΛ𝑛𝐪subscript𝜏𝑛𝐪subscript𝑣𝑛𝐪\Lambda_{n,\mathbf{q}}\equiv\tau_{n,\mathbf{q}}v_{n,\mathbf{q}}roman_Λ start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT ≡ italic_τ start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT. At low temperature, scattering due to geometric boundary is the dominant scattering mechanism and the mean free path no longer depends on the phonons frequency and wavelength, i.e., Λn,𝐪=ΛsubscriptΛ𝑛𝐪Λ\Lambda_{n,\mathbf{q}}=\Lambdaroman_Λ start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT = roman_Λ (constant) [64]. For example, in graphitic systems, ΛΛ\Lambdaroman_Λ is determined from the size and shape of the sample or the grain boundaries, and it well describes the thermal conductivity for up to 100 K [65, 66, 67]. In moiré systems, the superlattice period is generally observed to be varying across a single sample [68, 69] and this would disrupt the propagation of phonon modes [70]. In such a case, ΛΛ\Lambdaroman_Λ can be regarded as a typical length where the moiré pattern remains uniform. Henceforth, we treat ΛΛ\Lambdaroman_Λ as a phenomenological constant to be determined from direct measurements. The thermal conductivity is then entirely governed by the harmonic properties, and it can be written as

κ=Λ20n~(ω)C(ω,T)𝑑ω,𝜅Λ2superscriptsubscript0~𝑛𝜔𝐶𝜔𝑇differential-d𝜔\kappa=\frac{\Lambda}{2}\int_{0}^{\infty}\tilde{n}(\omega)C(\omega,T)d\omega,italic_κ = divide start_ARG roman_Λ end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT over~ start_ARG italic_n end_ARG ( italic_ω ) italic_C ( italic_ω , italic_T ) italic_d italic_ω , (19)

where velocity-weighted density of states (VDOS) n~~𝑛\tilde{n}over~ start_ARG italic_n end_ARG is defined as

n~(ω)=1Sn,𝐪δ(ωωn,𝐪)vn,𝐪~𝑛𝜔1𝑆subscript𝑛𝐪𝛿𝜔subscript𝜔𝑛𝐪subscript𝑣𝑛𝐪\tilde{n}(\omega)=\frac{1}{S}\sum_{n,\mathbf{q}}\delta(\omega-\omega_{n,% \mathbf{q}})v_{n,\mathbf{q}}over~ start_ARG italic_n end_ARG ( italic_ω ) = divide start_ARG 1 end_ARG start_ARG italic_S end_ARG ∑ start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT italic_δ ( italic_ω - italic_ω start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT ) italic_v start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT (20)

and spectral heat capacity C𝐶Citalic_C is defined as

C(ω,T)𝐶𝜔𝑇\displaystyle C(\omega,T)italic_C ( italic_ω , italic_T ) =ωf0(ω)T=kB[βω/2sinh(βω/2)]2,absentPlanck-constant-over-2-pi𝜔subscript𝑓0𝜔𝑇subscript𝑘𝐵superscriptdelimited-[]𝛽Planck-constant-over-2-pi𝜔2𝛽Planck-constant-over-2-pi𝜔22\displaystyle=\hbar\omega\frac{\partial f_{0}(\omega)}{\partial T}=k_{B}\left[% \frac{\beta\hbar\omega/2}{\sinh(\beta\hbar\omega/2)}\right]^{2},= roman_ℏ italic_ω divide start_ARG ∂ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ω ) end_ARG start_ARG ∂ italic_T end_ARG = italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT [ divide start_ARG italic_β roman_ℏ italic_ω / 2 end_ARG start_ARG roman_sinh ( italic_β roman_ℏ italic_ω / 2 ) end_ARG ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (21)

where β=1/(kBT)𝛽1subscript𝑘𝐵𝑇\beta=1/(k_{B}T)italic_β = 1 / ( italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T ). In Eq. (19), n~(ω)~𝑛𝜔\tilde{n}(\omega)over~ start_ARG italic_n end_ARG ( italic_ω ) contains all information regarding the phonon dispersion while C(ω,T)𝐶𝜔𝑇C(\omega,T)italic_C ( italic_ω , italic_T ) acts as a weight function. The function C(ω,T)𝐶𝜔𝑇C(\omega,T)italic_C ( italic_ω , italic_T ) is plotted in Fig. 4; it equals to kBsubscript𝑘𝐵k_{B}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT when ω=0𝜔0\omega=0italic_ω = 0, and decays exponentially for ω2kBTgreater-than-or-equivalent-toPlanck-constant-over-2-pi𝜔2subscript𝑘𝐵𝑇\hbar\omega\gtrsim 2k_{B}Troman_ℏ italic_ω ≳ 2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T. In general, the summation over mode index n𝑛nitalic_n would include both in-plane and out-of-plane (flexural) phonon modes. However, scattering due to substrate or other layers significantly reduces the mean free path of the flexural mode [39, 40, 41, 42, 43, 44], hence, its contribution to thermal conductivity especially at low temperatures [38]. Hereafter, we will focus only on the in-plane modes contribution in the thermal conductivity.

Refer to caption
Figure 4: Plot of the spectral heat capacity function C(ω,T)𝐶𝜔𝑇C(\omega,T)italic_C ( italic_ω , italic_T ) [Eq. (21)] as a function of phonon energy at different temperatures.

Since only the antisymmetric phonon modes are strongly affected by the interlayer moiré potential, it is useful to write κ𝜅\kappaitalic_κ and n~~𝑛\tilde{n}over~ start_ARG italic_n end_ARG in terms of the symmetric and antisymmetric components

κ=κ++κ,n~=n~++n~,formulae-sequence𝜅superscript𝜅superscript𝜅~𝑛superscript~𝑛superscript~𝑛\displaystyle\kappa=\kappa^{+}+\kappa^{-},\quad\tilde{n}=\tilde{n}^{+}+\tilde{% n}^{-},italic_κ = italic_κ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_κ start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT , over~ start_ARG italic_n end_ARG = over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT , (22)

which takes the contribution from the symmetric (+++) and antisymmetric (--) phonon modes separately. In the absence of the interlayer moiré coupling (i.e., two independent monolayers), they are given as

n~NC+subscriptsuperscript~𝑛NC\displaystyle\tilde{n}^{+}_{\rm NC}over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT =n~NC=12n~NC=ω2πv¯,absentsubscriptsuperscript~𝑛NC12subscript~𝑛NC𝜔2𝜋¯𝑣\displaystyle=\tilde{n}^{-}_{\rm NC}=\tfrac{1}{2}\tilde{n}_{\rm NC}=\frac{% \omega}{2\pi{\bar{v}}},= over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG over~ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT = divide start_ARG italic_ω end_ARG start_ARG 2 italic_π over¯ start_ARG italic_v end_ARG end_ARG , (23)
κNC+subscriptsuperscript𝜅NC\displaystyle\kappa^{+}_{\rm NC}italic_κ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT =κNC=12κNC=3Λζ(3)2π2v¯kB3T2,absentsubscriptsuperscript𝜅NC12subscript𝜅NC3Λ𝜁32𝜋superscriptPlanck-constant-over-2-pi2¯𝑣superscriptsubscript𝑘𝐵3superscript𝑇2\displaystyle=\kappa^{-}_{\rm NC}=\tfrac{1}{2}\kappa_{\rm NC}=\frac{3\Lambda% \zeta(3)}{2\pi\hbar^{2}{\bar{v}}}k_{B}^{3}T^{2},= italic_κ start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_κ start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT = divide start_ARG 3 roman_Λ italic_ζ ( 3 ) end_ARG start_ARG 2 italic_π roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over¯ start_ARG italic_v end_ARG end_ARG italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (24)

where v¯1=vL1+vT1superscript¯𝑣1superscriptsubscript𝑣L1superscriptsubscript𝑣T1{\bar{v}}^{-1}=v_{\rm L}^{-1}+v_{\rm T}^{-1}over¯ start_ARG italic_v end_ARG start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = italic_v start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + italic_v start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, ζ(n)𝜁𝑛\zeta(n)italic_ζ ( italic_n ) is the Riemann zeta function, and NC stands for ‘non-coupled’. Here, we note that we have neglected the out-of-plane (flexural) phonon modes in the calculation of thermal conductivity. The quadratic temperature dependence of the thermal conductivity is a characteristic of linear acoustic phonon-dominated thermal transport of 2D systems in the low-temperature regime [38].

III Results and discussion

III.1 TBG and t-G/hBN

Figure 5 shows the calculated band dispersion and density of states for antisymmetric phonon modes in TBG with θ=𝜃absent\theta=italic_θ = 5, 2.65 and 0.817. Figure 6 presents similar plots for t-G/hBN with θ=𝜃absent\theta=italic_θ = 5, 3, and 0. The phonon dispersion is plotted along the high-symmetry line in the moiré Brillouin zone (MBZ) as illustrated in Fig. 2(c). In each panel, the red-dashed line indicate the dispersion (only the lowest two branches shown) and the density of states in the non-coupled case. Here we can see that the interlayer coupling significantly modifies the band structure of the antisymmetric modes in the energy range below similar-to\sim20 meV. This reconstruction is characterized by sharp peaks in the density of states which results from the flattening of superlattice phonon bands [23, 24, 25, 32]. As the twist angle is reduced towards 0, the number of sharp peaks increases dramatically. On the other hand, the phonon bands of the symmetric phonon modes are not affected by the moiré interlayer coupling, hence they are equivalent to those of the non-coupled case (red-dashed lines).

Refer to caption
Figure 5: Band structure of the interlayer antisymmetric phonon modes (black line) and the symmetric modes (red-dashed line) for (a) 5, (b) 2.65, (c) 0.817 TBG. The right panel plots the density of states. (d)-(f) Corresponding plots for D(ω,v)𝐷𝜔𝑣D(\omega,v)italic_D ( italic_ω , italic_v ), the density of states on the frequency-velocity space [see, Eq. (25)].
Refer to caption
Figure 6: Plots of the band structure and D(ω,v)𝐷𝜔𝑣D(\omega,v)italic_D ( italic_ω , italic_v ) similar to Fig. 5, for t-G/hBN with θ=5𝜃superscript5\theta=5^{\circ}italic_θ = 5 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT, 3, 0.

To provide a more comprehensive understanding of the band flattening effects, we present two-dimensional density maps in Fig. 5(d)-(f) and Fig. 6(d)-(f), which plot the distribution of antisymmetric phonon modes on a space of frequency and velocity, or

D(ω,v)=n,𝐪δ(ωωn,𝐪)δ(vvn,𝐪).𝐷𝜔𝑣subscript𝑛𝐪𝛿𝜔subscript𝜔𝑛𝐪𝛿𝑣subscript𝑣𝑛𝐪D(\omega,v)=\sum_{n,\mathbf{q}}\delta(\omega-\omega_{n,\mathbf{q}})\delta(v-v_% {n,\mathbf{q}}).italic_D ( italic_ω , italic_v ) = ∑ start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT italic_δ ( italic_ω - italic_ω start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT ) italic_δ ( italic_v - italic_v start_POSTSUBSCRIPT italic_n , bold_q end_POSTSUBSCRIPT ) . (25)

In each panel, the vertical axis is scaled by a factor of

v0=λ/ρsubscript𝑣0𝜆𝜌v_{0}=\sqrt{\lambda/\rho}italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG italic_λ / italic_ρ end_ARG (26)

and the color brightness is a linear scale of D(ω,v)𝐷𝜔𝑣D(\omega,v)italic_D ( italic_ω , italic_v ). The horizontal red-dashed lines correspond to the velocity of the TA (vTsubscript𝑣Tv_{\rm T}italic_v start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT) and LA (vLsubscript𝑣Lv_{\rm L}italic_v start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT) phonons of a single layer [Eq. (9)]. Here, the flattening of phonon bands can be immediately seen as a distribution of phonon modes below vTsubscript𝑣Tv_{\rm T}italic_v start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT and vLsubscript𝑣Lv_{\rm L}italic_v start_POSTSUBSCRIPT roman_L end_POSTSUBSCRIPT. In 0.817 TBG [Fig. 5(f)] and 0 t-G/hBN [Fig. 6(f)], particularly, the signals in ω<15Planck-constant-over-2-pi𝜔15\hbar\omega<15roman_ℏ italic_ω < 15 meV predominantly falls below the line of vTsubscript𝑣Tv_{\rm T}italic_v start_POSTSUBSCRIPT roman_T end_POSTSUBSCRIPT, i.e., nearly all of the low-energy antisymmetric phonons become slower than the original acoustic phonons in its non-moiré counterpart. Note that the distribution of symmetric phonon states (not shown) sticks to the vLsubscript𝑣𝐿v_{L}italic_v start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and vTsubscript𝑣𝑇v_{T}italic_v start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT lines, where the intensity increases linearly with energy.

Refer to caption
Figure 7: (a) Velocity-weighted density of states (VDOS) of the antisymmetric modes, n~(ω)superscript~𝑛𝜔\tilde{n}^{-}(\omega)over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ω ), in TBGs with various twist angles. Black dashed lines represents n~NC+=n~NCsubscriptsuperscript~𝑛NCsubscriptsuperscript~𝑛NC\tilde{n}^{+}_{\rm NC}=\tilde{n}^{-}_{\rm NC}over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT = over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT for a non-coupled bilayer [Eq. (23)]. The VDOS for symmetric phonons n~+superscript~𝑛\tilde{n}^{+}over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT is equal to n~NC±subscriptsuperscript~𝑛plus-or-minusNC\tilde{n}^{\pm}_{\rm NC}over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT. (b) Similar plots for t-G/hBN.

Figure 7(a) and (b) show n~(ω)superscript~𝑛𝜔\tilde{n}^{-}(\omega)over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_ω ) (the VDOS of the antisymmetric phonon modes) of TBG and t-G/hBN, respectively, with various small twist angles. The effect of the moiré coupling is observed as a difference from a black-dashed line, which represents n~NC+=n~NCωsubscriptsuperscript~𝑛NCsubscriptsuperscript~𝑛NCproportional-to𝜔\tilde{n}^{+}_{\rm NC}=\tilde{n}^{-}_{\rm NC}\propto\omegaover~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT = over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT ∝ italic_ω [Eq. (23)]. While VDOS is proportional to both phonon density of states and velocity, we find that the reduction of phonon velocity is more significant than the sharpening of density of states, leading to a suppression of n~superscript~𝑛\tilde{n}^{-}over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT over a wide range of phonon frequency. We also find that the linear behavior of n~superscript~𝑛\tilde{n}^{-}over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT remains in the low frequency region, which corresponds to the linear dispersion in the the lowest moiré phonon band in ω<ωedge2πv0/LM𝜔subscript𝜔edge2𝜋subscript𝑣0subscript𝐿M\omega<\omega_{\rm edge}\approx 2\pi v_{0}/L_{\rm M}italic_ω < italic_ω start_POSTSUBSCRIPT roman_edge end_POSTSUBSCRIPT ≈ 2 italic_π italic_v start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUBSCRIPT roman_M end_POSTSUBSCRIPT. For the symmetric phonon modes (not shown), we have n~+(ω)=n~NC±(ω)superscript~𝑛𝜔subscriptsuperscript~𝑛plus-or-minusNC𝜔\tilde{n}^{+}(\omega)=\tilde{n}^{\pm}_{\rm NC}(\omega)over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_ω ) = over~ start_ARG italic_n end_ARG start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT ( italic_ω ), because they are not influenced by the moiré coupling.

Figure 8(a) and (b) summarise the calculated thermal conductivity κ(T)(=κ++κ)annotated𝜅𝑇absentsuperscript𝜅superscript𝜅\kappa(T)(=\kappa^{+}+\kappa^{-})italic_κ ( italic_T ) ( = italic_κ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_κ start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) of TBG and t-G/hBN, respectively. In each figure, the left panel shows a log-log plot of κ(T)𝜅𝑇\kappa(T)italic_κ ( italic_T ) for 2<T<1002𝑇1002<T<1002 < italic_T < 100 K. Here the vertical axis is scaled by the constant mean-free-path length ΛΛ\Lambdaroman_Λ, which is assumed to be independent of temperature. The colored lines correspond to different twist angles. The black-dashed line represents the non-coupled bilayer case, κNCsubscript𝜅NC\kappa_{\rm NC}italic_κ start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT, which is proportional to T2superscript𝑇2T^{2}italic_T start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT as given in Eq. (24). Here, we find that thermal conductivity is suppressed from the non-coupled case, notably around 20 K, while it converges towards the non-coupled value as increasing temperatures.

Refer to caption
Figure 8: (a) Thermal conductivity in TBGs with various twist angles. The left panel shows the thermal conductivity κ(T)𝜅𝑇\kappa(T)italic_κ ( italic_T ) scaled by the mean free path ΛΛ\Lambdaroman_Λ. The right panel shows the relative thermal conductivity to the non-coupled bilayer case, κNC(T)subscript𝜅NC𝑇\kappa_{\rm NC}(T)italic_κ start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT ( italic_T ) [Eq. (24)]. (b) Similar plots for t-G/hBN.

To better understand the change from the intrinsic graphene, we plot the relative thermal conductivity, κ(T)/κNC(T)𝜅𝑇subscript𝜅NC𝑇\kappa(T)/\kappa_{\rm NC}(T)italic_κ ( italic_T ) / italic_κ start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT ( italic_T ), in the right panel of Fig. 8(a) and (b). Here, suppression of thermal conductivity is represented by a value below unity. We can see that the suppression occurs over a wide temperature range except near the low-temperature limit. This suppression becomes more pronounced as the twist angle is smaller, where the largest reduction of up to similar-to\sim35% for 0.817 TBG and up to similar-to\sim40% for 0 t-G/hBN takes place at around 20 K. This can be understood since at around 20 K, heat transport is carried by phonons with energy below similar-to\sim20 meV [see Fig. 4], where the moiré effect on the asymmetric modes are the most notable, as shown by the changes in VDOS [Fig. 7]. There, thermal conductivity is almost entirely came from the symmetric phonon modes, κ+=κNC/2superscript𝜅subscript𝜅NC2\kappa^{+}=\kappa_{\rm NC}/2italic_κ start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT = italic_κ start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT / 2 (red-dashed line).

In the zero-temperature limit, we observe that the relative thermal conductivity κ/κNC𝜅subscript𝜅NC\kappa/\kappa_{\rm NC}italic_κ / italic_κ start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT rapidly rises and even exceeds 1, indicating that the thermal transport is enhanced by the moiré effect. The reason for this phenomenon is explained as follows. Within this temperature range, only phonons in the two lowest moiré phonon bands in ω<ωedge𝜔subscript𝜔edge\omega<\omega_{\rm edge}italic_ω < italic_ω start_POSTSUBSCRIPT roman_edge end_POSTSUBSCRIPT become relevant in the thermal transport equation. These phonon modes have a linear dispersion, where the VDOS is given by ω/(2πv)𝜔2𝜋𝑣\omega/(2\pi v)italic_ω / ( 2 italic_π italic_v ) with the corresponding group velocity v𝑣vitalic_v [Eq. (23) for the noncoupled case]. Since the velocities of the antisymmetric phonon modes are significantly reduced by the moiré effects [see Fig. 5 and Fig. 6], the inverse relation leads to an enhancement of the VDOS, and hence of the thermal conductivity. In decreasing the twist angle, these phonon velocities are monotonically decreased, and eventually converges towards a finite value in the small angle limit [23, 71]. This sets the upper bound of the relative thermal conductivity in the T0𝑇0T\to 0italic_T → 0 limit.

We also find that the overall modifications of the thermal conductivity by the moiré effect results in a change of the power coefficient α𝛼\alphaitalic_α in κ(T)Tαproportional-to𝜅𝑇superscript𝑇𝛼\kappa(T)\propto T^{\alpha}italic_κ ( italic_T ) ∝ italic_T start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT, as seen from the logarithmic plot in the left panel of Fig. 8(a) and (b). In the absence of moiré interlayer coupling, thermal conductivity has quadratic temperature dependence (α=2𝛼2\alpha=2italic_α = 2), which comes from the linear dispersion of the original acoustic phonons. However, the enhancement in the low T𝑇Titalic_T limit and the subsequent suppression in higher T𝑇Titalic_T decrease the power coefficient from 2. For example, it is given by α1.6𝛼1.6\alpha\approx 1.6italic_α ≈ 1.6 in 0.817 TBG and α1.4𝛼1.4\alpha\approx 1.4italic_α ≈ 1.4 in 0 t-G/hBN within the temperature range of 4 K to 8 K. As temperature increases further, moiré effect starts to fade out and thermal conductivity returns towards the original value. This requires the power coefficient to be larger than 2. For example, it is given by α2.25𝛼2.25\alpha\approx 2.25italic_α ≈ 2.25 for 0.817 TBG and α2.4𝛼2.4\alpha\approx 2.4italic_α ≈ 2.4 for 0 for t-G/hBN which occurs in the range from 35 K to 80 K.

III.2 t-MoS2

Refer to caption
Figure 9: Phonon dispersion and density of states of the antisymmetric modes (black line) and the symmetric modes (red-dashed line) for (a) parallel-stacked (P) and (b) antiparallel-stacked (AP) 0.8 t-MoS2.

The calculated phonon dispersion for t-MoS2 is shown in Fig. 9 for both parallel and antiparallel stacking case with twist angle 0.8. We find that the antisymmetric phonon modes of the parallel-stacked case closely resembles that of TBG [Fig. 5(c)]. This similarity occurs because TBG and t-MoS2 share triangular domain wall structures [see Fig. 3(a) and (c)]. In an analogous way, the phonon band structure of the antiparallel t-MoS2 resembles those of t-G/hBN, reflecting a common honeycomb domain-wall structure [Fig. 3(b) and (d)]. This is a natural result because the moiré phonon band structure is qualitatively reproduced by an effective mass-spring model for domain-wall motion [25], and hence it is primarily determined by the geometical structure of domain walls (triangular or honeycomb). We also observe that the characteristic energy scale of the moiré phonon bands in t-MoS2 is much smaller than in TBG and t-G/hBN, because the original acoustic phonon velocity in MoS2 is much lower than that of graphene and hBN.

In Fig. 10, we plot the calculated thermal conductivity of t-MoS2 (P and AP) in a parallel manner to the TBG and t-G/hBN case. Here we find an overall reduction of thermal conductivity and the enhancement near T0𝑇0T\to 0italic_T → 0 just as in TBG and t-G/hBN. However, the characteristic temperature range is much lower than TBG and t-G/hBN because of the smaller energy scale of the corresponding moiré phonons. At Tsimilar-to𝑇absentT\simitalic_T ∼ 4 K, the thermal conductivity is reduced up to around 35% and 40% in the P and AP cases, respectively. In the limit of T0𝑇0T\to 0italic_T → 0, we find that the AP case exhibits greater enhancement of κ𝜅\kappaitalic_κ than in the P case. This is attributed to the smaller phonon velocity v𝑣vitalic_v in the lowest branch in the AP case [Fig. 9] and the fact that κ𝜅\kappaitalic_κ for the linear band regime is proportional to 1/v1𝑣1/v1 / italic_v as argued in the previous section. Accordingly, we have the corresponding change of the power coefficient (α)𝛼(\alpha)( italic_α ) in κ(T)Tαproportional-to𝜅𝑇superscript𝑇𝛼\kappa(T)\propto T^{\alpha}italic_κ ( italic_T ) ∝ italic_T start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT in a similar manner to TBG and t-G/hBN.

Refer to caption
Figure 10: (a) Thermal conductivity of t-MoS2 with θ=0.8𝜃superscript0.8\theta=0.8^{\circ}italic_θ = 0.8 start_POSTSUPERSCRIPT ∘ end_POSTSUPERSCRIPT for the parallel case (red line) and antiparallel case (blue line). (b) Relative thermal conductivity to the noncoupling case, κNC(T)subscript𝜅NC𝑇\kappa_{\rm NC}(T)italic_κ start_POSTSUBSCRIPT roman_NC end_POSTSUBSCRIPT ( italic_T ).

IV Conclusion

We have studied the thermal transport of various moiré superlattice systems in the low-temperature regime. In general, we found a reduction of thermal conductivity due to an overall flattening of the phonon bands. The largest reduction is around 35-40% which occurs at around 20 K for 0.817 TBG and 0 t-G/hBN and around 4 K for 0.8 t-MoS2/MoS2. At higher temperature, the thermal conductivity returns towards the original intrinsic value, as the moiré superlattice effect for the acoustic phonons is significant only in the low energy region. These changes result in a characteristic deviation to the original quadratic temperature dependence of two-dimensional systems where linear acoustic phonons dominate the thermal transport properties. The results could provide a pathway for engineering thermal conductivity and observation of the moiré effect in the phonon properties of twisted bilayer system. While the present study is limited in the harmonic approximation, we expect that anharmonic scattering between moiré phonons and the presence of disorder [70, 72] could further decrease the thermal conductivity. We leave this imposition for future study.

Acknowledgements.
This work was supported by JST SPRING, Grant Number JPMJSP2138, JSPS KAKENHI Grants No. JP21H05236, and No. JP21H05232, No. JP20H01840, and by JST CREST Grant No. JPMJCR20T3, Japan.

References

  • Dean et al. [2013] C. R. Dean, L. Wang, P. Maher, C. Forsythe, F. Ghahari, Y. Gao, J. Katoch, M. Ishigami, P. Moon, M. Koshino, et al., Hofstadter’s butterfly and the fractal quantum hall effect in moiré superlattices, Nature 497, 598 (2013).
  • Hunt et al. [2013] B. Hunt, J. D. Sanchez-Yamagishi, A. F. Young, M. Yankowitz, B. J. LeRoy, K. Watanabe, T. Taniguchi, P. Moon, M. Koshino, P. Jarillo-Herrero, et al., Massive dirac fermions and hofstadter butterfly in a van der waals heterostructure, Science 340, 1427 (2013).
  • Cao et al. [2018a] Y. Cao, V. Fatemi, A. Demir, S. Fang, S. L. Tomarken, J. Y. Luo, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi, E. Kaxiras, et al., Correlated insulator behaviour at half-filling in magic-angle graphene superlattices, Nature 556, 80 (2018a).
  • Cao et al. [2018b] Y. Cao, V. Fatemi, S. Fang, K. Watanabe, T. Taniguchi, E. Kaxiras, and P. Jarillo-Herrero, Unconventional superconductivity in magic-angle graphene superlattices, Nature 556, 43 (2018b).
  • Sharpe et al. [2019] A. L. Sharpe, E. J. Fox, A. W. Barnard, J. Finney, K. Watanabe, T. Taniguchi, M. Kastner, and D. Goldhaber-Gordon, Emergent ferromagnetism near three-quarters filling in twisted bilayer graphene, Science 365, 605 (2019).
  • Polshyn et al. [2019] H. Polshyn, M. Yankowitz, S. Chen, Y. Zhang, K. Watanabe, T. Taniguchi, C. R. Dean, and A. F. Young, Large linear-in-temperature resistivity in twisted bilayer graphene, Nature Physics 15, 1011 (2019).
  • Cao et al. [2020] Y. Cao, D. Chowdhury, D. Rodan-Legrain, O. Rubies-Bigorda, K. Watanabe, T. Taniguchi, T. Senthil, and P. Jarillo-Herrero, Strange metal in magic-angle graphene with near planckian dissipation, Physical review letters 124, 076801 (2020).
  • Dos Santos et al. [2007] J. L. Dos Santos, N. Peres, and A. C. Neto, Graphene bilayer with a twist: Electronic structure, Physical review letters 99, 256802 (2007).
  • Li et al. [2010] G. Li, A. Luican, J. Lopes dos Santos, A. Castro Neto, A. Reina, J. Kong, and E. Andrei, Observation of van hove singularities in twisted graphene layers, Nature physics 6, 109 (2010).
  • Trambly de Laissardière et al. [2010] G. Trambly de Laissardière, D. Mayou, and L. Magaud, Localization of dirac electrons in rotated graphene bilayers, Nano letters 10, 804 (2010).
  • Morell et al. [2010] E. S. Morell, J. Correa, P. Vargas, M. Pacheco, and Z. Barticevic, Flat bands in slightly twisted bilayer graphene: Tight-binding calculations, Physical Review B 82, 121407 (2010).
  • Luican et al. [2011] A. Luican, G. Li, A. Reina, J. Kong, R. Nair, K. S. Novoselov, A. K. Geim, and E. Andrei, Single-layer behavior and its breakdown in twisted graphene layers, Physical review letters 106, 126802 (2011).
  • Bistritzer and MacDonald [2011] R. Bistritzer and A. H. MacDonald, Moiré bands in twisted double-layer graphene, Proceedings of the National Academy of Sciences 108, 12233 (2011).
  • Moon and Koshino [2012] P. Moon and M. Koshino, Energy spectrum and quantum hall effect in twisted bilayer graphene, Physical Review B 85, 195458 (2012).
  • Yankowitz et al. [2012] M. Yankowitz, J. Xue, D. Cormode, J. D. Sanchez-Yamagishi, K. Watanabe, T. Taniguchi, P. Jarillo-Herrero, P. Jacquod, and B. J. LeRoy, Emergence of superlattice dirac points in graphene on hexagonal boron nitride, Nature physics 8, 382 (2012).
  • Jung et al. [2015] J. Jung, A. M. DaSilva, A. H. MacDonald, and S. Adam, Origin of band gaps in graphene on hexagonal boron nitride, Nature communications 6, 6308 (2015).
  • Jiang et al. [2012] J.-W. Jiang, B.-S. Wang, and T. Rabczuk, Acoustic and breathing phonon modes in bilayer graphene with moiré patterns, Applied Physics Letters 101, 023113 (2012).
  • Campos-Delgado et al. [2013] J. Campos-Delgado, L. G. Cançado, C. A. Achete, A. Jorio, and J.-P. Raskin, Raman scattering study of the phonon dispersion in twisted bilayer graphene, Nano Research 6, 269 (2013).
  • Cocemasov et al. [2013] A. I. Cocemasov, D. L. Nika, and A. A. Balandin, Phonons in twisted bilayer graphene, Physical Review B 88, 035428 (2013).
  • Choi and Choi [2018] Y. W. Choi and H. J. Choi, Strong electron-phonon coupling, electron-hole asymmetry, and nonadiabaticity in magic-angle twisted bilayer graphene, Physical Review B 98, 241412 (2018).
  • Lin et al. [2018] M.-L. Lin, Q.-H. Tan, J.-B. Wu, X.-S. Chen, J.-H. Wang, Y.-H. Pan, X. Zhang, X. Cong, J. Zhang, W. Ji, et al., Moiré phonons in twisted bilayer mos2, Acs Nano 12, 8770 (2018).
  • Parzefall et al. [2021] P. Parzefall, J. Holler, M. Scheuck, A. Beer, K.-Q. Lin, B. Peng, B. Monserrat, P. Nagler, M. Kempf, T. Korn, and C. Schüller, Moiré phonons in twisted mose2–wse2 heterobilayers and their correlation with interlayer excitons, 2D Materials 8, 035030 (2021).
  • Koshino and Son [2019] M. Koshino and Y.-W. Son, Moiré phonons in twisted bilayer graphene, Physical Review B 100, 075416 (2019).
  • Suri et al. [2021] N. Suri, C. Wang, Y. Zhang, and D. Xiao, Chiral phonons in moiré superlattices, Nano Letters 21, 10026 (2021), 2108.13000 .
  • Krisna and Koshino [2023] L. P. Krisna and M. Koshino, Moiré phonons in graphene/hexagonal boron nitride moiré superlattice, Physical Review B 107, 115301 (2023).
  • Lamparski et al. [2020] M. Lamparski, B. V. Troeye, and V. Meunier, Soliton signature in the phonon spectrum of twisted bilayer graphene, 2D Materials 7, 025050 (2020).
  • Maity et al. [2020] I. Maity, M. H. Naik, P. K. Maiti, H. R. Krishnamurthy, and M. Jain, Phonons in twisted transition-metal dichalcogenide bilayers: Ultrasoft phasons and a transition from a superlubric to a pinned phase, Physical Review Research 2, 013335 (2020).
  • Gadelha et al. [2021] A. C. Gadelha, D. A. A. Ohlberg, C. Rabelo, E. G. S. Neto, T. L. Vasconcelos, J. L. Campos, J. S. Lemos, V. Ornelas, D. Miranda, R. Nadas, F. C. Santana, K. Watanabe, T. Taniguchi, B. van Troeye, M. Lamparski, V. Meunier, V.-H. Nguyen, D. Paszko, J.-C. Charlier, L. C. Campos, L. G. Cançado, G. Medeiros-Ribeiro, and A. Jorio, Localization of lattice dynamics in low-angle twisted bilayer graphene, Nature 590, 405 (2021).
  • Quan et al. [2021] J. Quan, L. Linhart, M.-L. Lin, D. Lee, J. Zhu, C.-Y. Wang, W.-T. Hsu, J. Choi, J. Embley, C. Young, T. Taniguchi, K. Watanabe, C.-K. Shih, K. Lai, A. H. MacDonald, P.-H. Tan, F. Libisch, and X. Li, Phonon renormalization in reconstructed MoS2 moiré superlattices, Nature Materials 20, 1100 (2021).
  • Maity et al. [2022] I. Maity, A. A. Mostofi, and J. Lischner, Chiral valley phonons and flat phonon bands in moiré materials, Physical Review B 105, l041408 (2022), 2108.03965 .
  • Lu et al. [2022] J. Z. Lu, Z. Zhu, M. Angeli, D. T. Larson, and E. Kaxiras, Low-energy moiré phonons in twisted bilayer van der waals heterostructures, Physical Review B 106, 144305 (2022).
  • Xie and Liu [2023] B. Xie and J. Liu, Lattice distortions, moiré phonons, and relaxed electronic band structures in magic-angle twisted bilayer graphene, Physical Review B 108, 094115 (2023).
  • Girotto et al. [2023] N. Girotto, L. Linhart, and F. Libisch, Coupled phonons in twisted bilayer graphene, Physical Review B 108, 155415 (2023).
  • Balandin et al. [2008] A. A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao, and C. N. Lau, Superior thermal conductivity of single-layer graphene, Nano letters 8, 902 (2008).
  • Ghosh et al. [2008] D. Ghosh, I. Calizo, D. Teweldebrhan, E. P. Pokatilov, D. L. Nika, A. A. Balandin, W. Bao, F. Miao, and C. N. Lau, Extremely high thermal conductivity of graphene: Prospects for thermal management applications in nanoelectronic circuits, Applied Physics Letters 92, 151911 (2008).
  • Cai et al. [2010] W. Cai, A. L. Moore, Y. Zhu, X. Li, S. Chen, L. Shi, and R. S. Ruoff, Thermal transport in suspended and supported monolayer graphene grown by chemical vapor deposition, Nano letters 10, 1645 (2010).
  • Chen et al. [2012] S. Chen, Q. Wu, C. Mishra, J. Kang, H. Zhang, K. Cho, W. Cai, A. A. Balandin, and R. S. Ruoff, Thermal conductivity of isotopically modified graphene, Nature materials 11, 203 (2012).
  • Seol et al. [2010] J. H. Seol, I. Jo, A. L. Moore, L. Lindsay, Z. H. Aitken, M. T. Pettes, X. Li, Z. Yao, R. Huang, D. Broido, et al., Two-dimensional phonon transport in supported graphene, Science 328, 213 (2010).
  • Ghosh et al. [2010] S. Ghosh, W. Bao, D. L. Nika, S. Subrina, E. P. Pokatilov, C. N. Lau, and A. A. Balandin, Dimensional crossover of thermal transport in few-layer graphene, Nature Materials 9, 555 (2010).
  • Guo et al. [2011] Z.-X. Guo, D. Zhang, and X.-G. Gong, Manipulating thermal conductivity through substrate coupling, Physical Review B 84, 075470 (2011).
  • Lindsay et al. [2011] L. Lindsay, D. A. Broido, and N. Mingo, Flexural phonons and thermal transport in multilayer graphene and graphite, Physical Review B 83, 235428 (2011).
  • Pak and Hwang [2016] A. J. Pak and G. S. Hwang, Theoretical analysis of thermal transport in graphene supported on hexagonal boron nitride: the importance of strong adhesion due to π𝜋\piitalic_π-bond polarization, Physical Review Applied 6, 034015 (2016).
  • Zou and Cao [2017] J.-H. Zou and B.-Y. Cao, Phonon thermal properties of graphene on h-bn from molecular dynamics simulations, Applied Physics Letters 110, 103106 (2017).
  • Zou et al. [2019] J.-H. Zou, X.-T. Xu, and B.-Y. Cao, Size-dependent mode contributions to the thermal transport of suspended and supported graphene, Applied Physics Letters 115, 123105 (2019).
  • Li et al. [2014] H. Li, H. Ying, X. Chen, D. L. Nika, A. I. Cocemasov, W. Cai, A. A. Balandin, and S. Chen, Thermal conductivity of twisted bilayer graphene, Nanoscale 6, 13402 (2014).
  • Han et al. [2021] S. Han, X. Nie, S. Gu, W. Liu, L. Chen, H. Ying, L. Wang, Z. Cheng, L. Zhao, and S. Chen, Twist-angle-dependent thermal conduction in single-crystalline bilayer graphene, Applied Physics Letters 118, 193104 (2021).
  • Li et al. [2018] C. Li, B. Debnath, X. Tan, S. Su, K. Xu, S. Ge, M. R. Neupane, and R. K. Lake, Commensurate lattice constant dependent thermal conductivity of misoriented bilayer graphene, Carbon 138, 451 (2018).
  • Nie et al. [2019] X. Nie, L. Zhao, S. Deng, Y. Zhang, and Z. Du, How interlayer twist angles affect in-plane and cross-plane thermal conduction of multilayer graphene: A non-equilibrium molecular dynamics study, International Journal of Heat and Mass Transfer 137, 161 (2019).
  • Mandal et al. [2022] S. Mandal, I. Maity, A. Das, M. Jain, and P. K. Maiti, Tunable lattice thermal conductivity of twisted bilayer mos2, Physical Chemistry Chemical Physics 24, 13860 (2022).
  • Cheng et al. [2023] Y. Cheng, Z. Fan, T. Zhang, M. Nomura, S. Volz, G. Zhu, B. Li, and S. Xiong, Magic angle in thermal conductivity of twisted bilayer graphene, Materials Today Physics , 101093 (2023).
  • Ahmed et al. [2023] S. Ahmed, S. Alam, and A. Jain, Understanding phonon thermal transport in twisted bilayer graphene, Physical Review B 108, 235202 (2023).
  • Samajdar et al. [2022] R. Samajdar, Y. Teng, and M. S. Scheurer, Moiré phonons and impact of electronic symmetry breaking in twisted trilayer graphene, Physical Review B 106, L201403 (2022).
  • Popov et al. [2011] A. M. Popov, I. V. Lebedeva, A. A. Knizhnik, Y. E. Lozovik, and B. V. Potapkin, Commensurate-incommensurate phase transition in bilayer graphene, Physical Review B 84, 045404 (2011), 1108.2254 .
  • Lebedeva et al. [2011] I. V. Lebedeva, A. A. Knizhnik, A. M. Popov, Y. E. Lozovik, and B. V. Potapkin, Interlayer interaction and relative vibrations of bilayer graphene, Physical Chemistry Chemical Physics 13, 5687 (2011).
  • San-Jose et al. [2014] P. San-Jose, A. Gutiérrez-Rubio, M. Sturla, and F. Guinea, Spontaneous strains and gap in graphene on boron nitride, Physical Review B 90, 075428 (2014).
  • Enaldiev et al. [2020] V. Enaldiev, V. Zolyomi, C. Yelgel, S. Magorrian, and V. Fal’Ko, Stacking domains and dislocation networks in marginally twisted bilayers of transition metal dichalcogenides, Physical review letters 124, 206101 (2020).
  • Weston et al. [2020] A. Weston, Y. Zou, V. Enaldiev, A. Summerfield, N. Clark, V. Zólyomi, A. Graham, C. Yelgel, S. Magorrian, M. Zhou, et al., Atomic reconstruction in twisted bilayers of transition metal dichalcogenides, Nature Nanotechnology 15, 592 (2020).
  • Suzuura and Ando [2002] H. Suzuura and T. Ando, Phonons and electron-phonon scattering in carbon nanotubes, Physical review B 65, 235412 (2002).
  • Nam and Koshino [2017] N. N. Nam and M. Koshino, Lattice relaxation and energy band modulation in twisted bilayer graphene, Physical Review B 96, 075311 (2017).
  • Zakharchenko et al. [2009] K. Zakharchenko, M. Katsnelson, and A. Fasolino, Finite temperature lattice properties of graphene beyond the quasiharmonic approximation, Physical review letters 102, 046808 (2009).
  • Sachs et al. [2011] B. Sachs, T. Wehling, M. Katsnelson, and A. Lichtenstein, Adhesion and electronic structure of graphene on hexagonal boron nitride substrates, Physical Review B 84, 195414 (2011), 1105.2379 .
  • Carr et al. [2018] S. Carr, D. Massatt, S. B. Torrisi, P. Cazeaux, M. Luskin, and E. Kaxiras, Relaxation and domain formation in incommensurate two-dimensional heterostructures, Physical Review B 98, 224102 (2018).
  • Woods et al. [2014] C. Woods, L. Britnell, A. Eckmann, R. Ma, J. Lu, H. Guo, X. Lin, G. Yu, Y. Cao, R. V. Gorbachev, et al., Commensurate–incommensurate transition in graphene on hexagonal boron nitride, Nature physics 10, 451 (2014).
  • Ziman [2001] J. Ziman, Electrons and Phonons (Oxford University Press, 2001).
  • Prasher [2008] R. Prasher, Thermal boundary resistance and thermal conductivity of multiwalled carbon nanotubes, Physical Review B 77, 075424 (2008).
  • Heremans and Beetz Jr [1985] J. Heremans and C. Beetz Jr, Thermal conductivity and thermopower of vapor-grown graphite fibers, Physical Review B 32, 1981 (1985).
  • Klemens and Pedraza [1994] P. Klemens and D. Pedraza, Thermal conductivity of graphite in the basal plane, Carbon 32, 735 (1994).
  • Uri et al. [2020] A. Uri, S. Grover, Y. Cao, J. A. Crosse, K. Bagani, D. Rodan-Legrain, Y. Myasoedov, K. Watanabe, T. Taniguchi, P. Moon, et al., Map** the twist-angle disorder and landau levels in magic-angle graphene, Nature 581, 47 (2020).
  • Benschop et al. [2021] T. Benschop, T. A. de Jong, P. Stepanov, X. Lu, V. Stalman, S. J. van der Molen, D. K. Efetov, and M. P. Allan, Measuring local moiré lattice heterogeneity of twisted bilayer graphene, Physical Review Research 3, 013153 (2021).
  • Ochoa and Fernandes [2022] H. Ochoa and R. M. Fernandes, Degradation of phonons in disordered moiré superlattices, Physical review letters 128, 065901 (2022).
  • Gao and Khalaf [2022] Q. Gao and E. Khalaf, Symmetry origin of lattice vibration modes in twisted multilayer graphene: Phasons versus moiré phonons, Physical Review B 106, 075420 (2022).
  • Nakatsuji and Koshino [2022] N. Nakatsuji and M. Koshino, Moiré disorder effect in twisted bilayer graphene, Physical Review B 105, 245408 (2022).