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

thanks: C. L. and P. T. M. contributed equally to this work.

Moiré band theory for M-valley twisted transition metal dichalcogenides

Chao Lei [email protected]    Perry T. Mahon [email protected]    A. H. MacDonald Department of Physics, University of Texas at Austin, Austin, Texas 78712, USA
Abstract

We examine the viability of twisted bilayers of group IV and IVB trigonal transition metal dichalcogenides (TMDs) MX2 (M==Ti, Zr, Hf, Sn and X==S, Se, Te) as a moiré material platform. We show that the low energy conduction band states in these systems are accurately described by a set of three emergent periodic Hamiltonians which we derive for a representative homobilayer from small unit cell density functional theory calculations using a local displacement approximation. In monolayer form, these TMDs have conduction band minima with anisotropic effective masses near the three inequivalent Brillouin zone MM points, implying a moiré band model that has six flavors when spin is included. Because each MM point is time-reversal invariant, spontaneous valley polarization is signaled in transport by anisotropy instead of by the anomalous Hall and magnetic circular dichroism effects commonly seen in graphene and KK-valley TMD based moiré multilayers.

Introduction— Moiré materials, artificial two-dimensional (2D) crystals with large lattice constants created by forming moiré patterns in layered 2D materials, have emerged in recent years [1] as a successful platform for experimental and theoretical discovery. Moiré materials are strictly speaking quasicrystals, but when the constituent layers are semiconductors or semimetals the electronic states near the Fermi energy are accurately described by emergent Hamiltonians that have only the moiré periodicity [2, 3]. Because their effective Hamiltonians are periodic, moiré materials support low-energy Bloch bands, which can have a small bandwidth and be topologically nontrivial. Moreover, the large lattice constants allow the number of electrons per moiré unit cell to be tuned over large ranges with electrical gates - effectively moving through an artificial periodic table without chemical doping. Materials of this type therefore serve as flexible tunable quantum systems in which strong-correlation and topological electron physics can be studied, often together.

The moiré materials that have been studied experimentally to date are almost exclusively formed from either graphene or group VI transition metal dichalcogenide (TMD) 2D layers [1, 4]. Accordingly, there has been a focus on developing theoretical models to describe the low-energy electronic states in these specific materials [5, 2, 6, 7, 8, 9, 10, 11]. In generic weakly coupled multilayer materials, the electronic states near the Fermi energy are formed from the Bloch states near the Fermi energy of the constituent crystalline layers, which in graphene and group VI TMDs usually have crystal momenta near KK or Γ\Gamma points in their respective triangular lattice Brillouin zones (BZ). In moiré materials, Bloch states with crystal momenta that are near distinct low-energy momenta (the set of which we denote 𝒦BZ\mathcal{K}\subset\text{BZ}) decouple and separate periodic models emerge, one associated with each low-energy point (valley) 𝒌𝒦\bm{k}_{*}\in\mathcal{K} [3]. Most attention has been given to developing models for KK-valley materials. Since there are two inequivalent KK-points related by time-reversal symmetry in triangular lattices, the valley-projected moiré band Hamiltonians break time-reversal symmetry. This property is the root cause [12, 13, 14, 9, 15, 16] of the anomalous Hall effect and magnetic circular dichroism, and of the integer and fractional quantum anomalous Hall effects seen [17, 18, 19, 20, 21, 22, 23, 24] in these materials.

In this Letter, we propose that bilayers of group IV and IVB TMDs with small twist angles or differences in lattice constant are realistic candidate moiré materials and derive moiré band models for their low-energy conduction states using Hafnium diselenide (HfS2) twisted homobilayers as a representative example. In HfS2, the valence band maximum (VBM) is at the Γ\Gamma point and the conduction band minima (CBM) are at the three inequivalent MM points [25, 26] in both bulk and single-layer limits. We focus here on the CBM states in homobilayers and use Wannier function (WF) [27] processed density function theory (DFT) [28] to construct the moiré band model. The current model applies to HfS2 twisted homobilayers with nn-type electrostatic doping. Generalizations to other materials, to heterobilayers, and to multilayers are straightforward although the symmetry considerations discussed below are altered. We find that the valley-projected moiré bands are time-reversal invariant, implying that they cannot support an anomalous Hall effect or magnetic circular dichroism, and that they are strongly anisotropic. Thus, interaction-induced spontaneous valley polarization will manifest experimentally by anisotropy in transport.

Moiré band models via the local displacement scheme— The local displacement approach [3] is an approximation scheme that can be used to construct accurate periodic effective models of the low-energy electronic states of semiconductor bilayers that have small relative twist angles (θ\theta) and/or lattice constant ratios (λ\lambda) close to one, and weak coupling between layers. This approach employs a family of Bloch Hamiltonians H(𝗱)H(\bm{\mathsf{d}}) of (fictitious) crystalline insulator bilayers – constructed as aligned stacks of the quasicrystal’s constituent layers, but with lattice constants (of the top layer, for instance) that are artificially adjusted (if necessary) so that the top and bottom layers are characterized by the same Bravais lattice Λ\Lambda 111For every 𝗱2\bm{\mathsf{d}}\in\mathbb{R}^{2} the Bloch Hamiltonian H(𝒓,𝒑(𝒓);𝗱)H(\bm{r},\bm{p}(\bm{r});\bm{\mathsf{d}}) of the bilayer is Λ\Lambda-periodic in 𝒓\bm{r} where Λ\Lambda is the Bravais lattice of the (arbitrarily chosen) bottom layer. The vertical separation between layers is fixed at each 𝗱\bm{\mathsf{d}} by minimizing energy. Therefore, we can (and will) choose the real-space Bravais lattice to be independent of 𝗱\bm{\mathsf{d}}. The same is true of the unit cell Ωuc\Omega_{uc} (BZ) of Λ\Lambda (Λ\Lambda^{*}). – parameterized by an in-plane relative displacement 𝗱2\bm{\mathsf{d}}\in\mathbb{R}^{2} between layers. If H(𝗱)H(\bm{\mathsf{d}}) has time-reversal symmetry, then a set of WFs, denoted by |WI,𝑹(𝗱)\ket{W_{I,\bm{R}}(\bm{\mathsf{d}})}, can be constructed [30] for the set of isolated energy bands just above (below) the band gap at charge neutrality 222That H(𝗱)H(\bm{\mathsf{d}}) has time-reversal symmetry is not a necessary criterion, but is valid in HfS2 bilayers and simplifies our discussion. We also assume that at each 𝒌BZ\bm{k}\in\text{BZ} the set of isolated conduction (valence) bands remains energetically isolated from all other bands as 𝗱\bm{\mathsf{d}} is varied. Again, this is not necessary, but is convenient and valid in HfS2 bilayers.. We will assume that these WFs can be chosen such that the label II decomposes into layer and orbital labels I=(l,αl)I=(l,\alpha_{l}), where l=𝔟l=\mathfrak{b} and l=𝔱l=\mathfrak{t} identify bottom and top layers, respectively; indeed, the identification of layer as a degree of freedom is essential and the main advantage of a Wannier-based approach. This type of Wannier construction is natural in weakly coupled bilayers and succeeds in HfS2 bilayers. We denote the 𝗱\bm{\mathsf{d}}-dependent WF matrix elements by W(l,αl),𝑹(𝗱)|H(𝗱)|W(l,βl),𝑹(𝗱)\bra{W_{(l,\alpha_{l}),\bm{R}}(\bm{\mathsf{d}})}H(\bm{\mathsf{d}})\ket{W_{(l^{\prime},\beta_{l^{\prime}}),\bm{R}^{\prime}}(\bm{\mathsf{d}})} and their Bloch Fourier transforms by

Hl,βll,αl(𝒌;𝗱)ψ~(l,αl),𝒌(𝗱)|H(𝗱)|ψ~(l,βl),𝒌(𝗱).\displaystyle H^{l,\alpha_{l}}_{l^{\prime},\beta_{l^{\prime}}}(\bm{k};\bm{\mathsf{d}})\equiv\bra{\tilde{\psi}_{(l,\alpha_{l}),\bm{k}}(\bm{\mathsf{d}})}H(\bm{\mathsf{d}})\ket{\tilde{\psi}_{(l^{\prime},\beta_{l^{\prime}}),\bm{k}}(\bm{\mathsf{d}})}. (1)

Here |ψ~(l,αl),𝒌(𝗱)\ket{\tilde{\psi}_{(l,\alpha_{l}),\bm{k}}(\bm{\mathsf{d}})} are Bloch-type vectors defined by the WFs and are smooth over the BZ and Λ\Lambda^{*}-periodic in 𝒌\bm{k}. These momentum-space Wannier matrix elements evaluated at valley-center momenta 𝒌𝒦\bm{k}_{*}\in\mathcal{K} provide the microscopic data needed to construct the moiré band models.

We assume that the WFs track smoothly with 𝗱\bm{\mathsf{d}} so that W(l,α),𝑹(𝒓;𝗱+𝗮)=W(l,α),𝑹(𝒓δl,𝔱𝗮;𝗱)W_{(l,\alpha),\bm{R}}(\bm{r};\bm{\mathsf{d}}+\bm{\mathsf{a}})=W_{(l,\alpha),\bm{R}}(\bm{r}-\delta_{l,\mathfrak{t}}\bm{\mathsf{a}};\bm{\mathsf{d}}) after relative translation of layers by 𝗮2\bm{\mathsf{a}}\in\mathbb{R}^{2}. It follows that

|ψ~(l,α),𝒌(𝗱+𝗮)eiδl,𝔱𝒌𝗮|ψ~(l,α),𝒌(𝗱),\displaystyle\ket{\tilde{\psi}_{(l,\alpha),\bm{k}}(\bm{\mathsf{d}}+\bm{\mathsf{a}})}\approx e^{-i\delta_{l,\mathfrak{t}}\bm{k}\cdot\bm{\mathsf{a}}}\ket{\tilde{\psi}_{(l,\alpha),\bm{k}}(\bm{\mathsf{d}})}, (2)

and therefore that

H¯l,βll,αl(𝒌;𝗱)ei𝒌(δl,𝔱δl,𝔱)𝗱Hl,βll,αl(𝒌;𝗱)\displaystyle\bar{H}^{l,\alpha_{l}}_{l^{\prime},\beta_{l^{\prime}}}(\bm{k};\bm{\mathsf{d}})\equiv e^{i\bm{k}\cdot(\delta_{l^{\prime},\mathfrak{t}}-\delta_{l,\mathfrak{t}})\bm{\mathsf{d}}}H^{l,\alpha_{l}}_{l^{\prime},\beta_{l^{\prime}}}(\bm{k};\bm{\mathsf{d}}) (3)

is Λ\Lambda-periodic in 𝗱\bm{\mathsf{d}} and admits a Fourier series expansion with coefficients given by

H¯l,βll,αl(𝒌;𝑮)=1ΩucΩucei𝑮𝗱H¯l,βll,αl(𝒌;𝗱)𝑑𝗱\displaystyle\bar{H}^{l,\alpha_{l}}_{l^{\prime},\beta_{l^{\prime}}}(\bm{k};\bm{G})=\frac{1}{\Omega_{uc}}\int_{\Omega_{uc}}e^{i\bm{G}\cdot\bm{\mathsf{d}}}\bar{H}^{l,\alpha_{l}}_{l^{\prime},\beta_{l^{\prime}}}(\bm{k};\bm{\mathsf{d}})\;d\bm{\mathsf{d}} (4)

for 𝑮Λ\bm{G}\in\Lambda^{*}.

There are two elements to the local displacement scheme. First we replace 𝗱\bm{\mathsf{d}} in each WF matrix element by 𝒅(𝑹)λRθz(𝑹)𝑹\bm{d}(\bm{R})\equiv\lambda R^{z}_{\theta}(\bm{R})-\bm{R}, where 𝑹\bm{R} is the lattice site of a WF in that matrix element and 𝒅(𝑹)\bm{d}(\bm{R}) is its local displacement 333The quasicrystal has top (bottom) layer Bravais lattice Λt\Lambda_{t} (Λb\Lambda_{b}) and we arbitrarily choose ΛΛb\Lambda\equiv\Lambda_{b}. By definition, Λt=λRθz(Λb)\Lambda_{t}=\lambda R_{\theta}^{z}(\Lambda_{b}) and the operator 𝒅:22\bm{d}:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} encodes the local displacement of bottom layer lattice site 𝑹Λb\bm{R}\in\Lambda_{b} via λRθz(𝑹)(𝑹+𝒅(𝑹))Λt\lambda R^{z}_{\theta}(\bm{R})\equiv\big{(}\bm{R}+\bm{d}(\bm{R})\big{)}\in\Lambda_{t}. for twist angle θ\theta and lattice scaling factor λ\lambda [3]. Rθz:22R^{z}_{\theta}:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} is a linear function of position that encodes counter-clockwise rotation by θ\theta about the surface-normal direction 𝒆^z\hat{\bm{e}}_{z}. Note that this replacement coarse grains position from the lattice points Λ\Lambda to continuous space. At this stage we are neglecting in-plane strain relaxation relative to rigid rotation, which can be important, but this correction can be conveniently added, as we discuss below. Secondly we replace 𝒌\bm{k} in the l=ll=l^{\prime}, 𝑮𝟎\bm{G}\neq\bm{0} and the lll\neq l^{\prime} terms by their value at the nearby 𝒌𝒦\bm{k}_{*}\in\mathcal{K} – the wavevector of the CBM at the center of the valley of interest. Both elements are justified [3] for sufficiently long moiré periods because the 𝒌\bm{k} dependence of the Hamiltonian at each 𝗱\bm{\mathsf{d}} occurs on the scale of Λ\Lambda.

Refer to caption
Figure 1: (a) Side view of AA-stacked crystalline bilayer HfS2 with relative in-plane layer displacement 𝗱=𝟎\bm{\mathsf{d}}=\bm{0}. (b) Top view of a bilayer with nonzero displacement 𝗱\bm{\mathsf{d}}. Red (blue) dots identify the positions of Hf (S) ions. (c) Bandstructure for 𝗱=𝟎\bm{\mathsf{d}}=\bm{0} obtained from DFT (black). For all 𝗱\bm{\mathsf{d}} values the six doubly-degenerate low-energy conduction bands are energetically isolated from all others at each 𝒌BZ\bm{k}\in\text{BZ} and are indistinguishable on the scale of this figure from those obtained using a six orbital Wannier90 tight-binding model. (d) and (e) Low-energy conduction bands from Wannier90 for 𝗱=𝒂1/3+𝒂2/3\bm{\mathsf{d}}=\bm{a}_{1}/3+\bm{a}_{2}/3 and 𝗱=2𝒂1/3+2𝒂2/3\bm{\mathsf{d}}=2\bm{a}_{1}/3+2\bm{a}_{2}/3.

With these approximations, the local displacement (LD) Hamiltonian HLD(𝒌)H_{\text{LD}}(\bm{k}_{*}) for valley 𝒌𝒦\bm{k}_{*}\in\mathcal{K} acts in an envelope function space [33] that is the direct product of layer, Wannier orbital (within each layer), and continuous position subspaces, and its eigenstates are spinors with layer and Wannier orbital indices. Using a momentum-space representation for the position degree of freedom one obtains

l,αl,𝒌|HLD(𝒌)|l,βl,𝒌\displaystyle\bra{l,\alpha_{l},\bm{k}}H_{\text{LD}}(\bm{k}_{*})\ket{l^{\prime},\beta_{l^{\prime}},\bm{k}^{\prime}}
=δl,l(δ(𝒌𝒌)H¯l,βll,αl(𝒌;𝑮=𝟎)\displaystyle=\delta_{l,l^{\prime}}\Big{(}\delta\big{(}\bm{k}^{\prime}-\bm{k}\big{)}\bar{H}^{l,\alpha_{l}}_{l,\beta_{l}}(\bm{k};\bm{G}=\bm{0})
+𝑮Λ{𝟎}δ(𝒌𝒌𝑮~)H¯l,βll,αl(𝒌;𝑮))\displaystyle\qquad\qquad+\sum_{\bm{G}\in\Lambda^{*}\setminus\{\bm{0}\}}\delta\big{(}\bm{k}^{\prime}-\bm{k}-\tilde{\bm{G}}\big{)}\bar{H}^{l,\alpha_{l}}_{l,\beta_{l}}(\bm{k}_{*};\bm{G})\Big{)}
+(1δl,l)𝑮Λδ(𝒌𝒌+(1)l,𝔱𝒌~𝑮~)H¯l,βll,αl(𝒌;𝑮).\displaystyle+(1-\delta_{l,l^{\prime}})\sum_{\bm{G}\in\Lambda^{*}}\delta\big{(}\bm{k}^{\prime}-\bm{k}+(-1)^{l,\mathfrak{t}}\tilde{\bm{k}}_{*}-\tilde{\bm{G}}\big{)}\bar{H}^{l,\alpha_{l}}_{l^{\prime},\beta_{l^{\prime}}}(\bm{k}_{*};\bm{G}). (5)

Here 𝑮Λ\bm{G}\in\Lambda^{*}, 𝑮~λRθz(𝑮)𝑮\tilde{\bm{G}}\equiv\lambda R^{z}_{-\theta}(\bm{G})-\bm{G} is a vector in the moiré reciprocal lattice ΛM\Lambda_{\text{M}}^{*} 444The set ΛM{λRθ(𝑮)𝑮:𝑮Λ}2\Lambda_{\text{M}}^{*}\equiv\{\lambda R_{-\theta}(\bm{G})-\bm{G}:\bm{G}\in\Lambda^{*}\}\subset\mathbb{R}^{2} equipped with addition is a Bravais lattice. This is called the moiré reciprocal lattice and its dual, the moiré lattice, is given by ΛM={𝒓2:𝒅(𝖗)Λ}\Lambda_{\text{M}}=\{\bm{r}\in\mathbb{R}^{2}:\bm{d}(\bm{\mathfrak{r}})\in\Lambda\}., 𝒌~λRθz(𝒌)𝒌\tilde{\bm{k}}_{*}\equiv\lambda R^{z}_{-\theta}(\bm{k}_{*})-\bm{k}_{*}, and 𝒌,𝒌\bm{k},\bm{k}^{\prime} are extended from BZ to 2\mathbb{R}^{2}. Because the layer separations in the van der Waals crystals from which moiré materials can be successfully fabricated are always much larger than the lattice constants within layers, sizable values for H¯l,βll,αl(𝒌;𝑮)\bar{H}^{l,\alpha_{l}}_{l^{\prime},\beta_{l^{\prime}}}(\bm{k};\bm{G}) are expected for 𝑮\bm{G} only in the first few shells of Λ\Lambda^{*}, implying that the lowest energy moiré bands involve pockets of momentum space that are small compared to the Brillouin zone size. This property justifies the valley-decoupling that occurs in the local displacement approximation and also justifies an expansion of H¯l,βll,αl(𝒌;𝑮=𝟎)\bar{H}^{l,\alpha_{l}}_{l,\beta_{l}}(\bm{k};\bm{G}=\bm{0}) around the CBM 𝒌\bm{k}_{*} that is parameterized by an effective mass tensor. A key property of MM-valley TMDs is that their effective mass tensors are anisotropic.

𝗱\bm{\mathsf{d}}-dependence of HfS2 crystalline bilayer bands— As illustrated in Fig. 1, monolayer HfS2 has a triangular lattice Λ\Lambda with Hf ions located at Bravais lattice sites around which S ions form octahedral cages [25]. Trigonal monolayers and AA-stacked 555By AA-stacked we mean that top and bottom layer have identical orientation. bilayers (with any 𝗱\bm{\mathsf{d}}) have a center of inversion symmetry that leads to a Kramers spin-degeneracy at each 𝒌BZ\bm{k}\in\text{BZ}. For HfS2 bilayers, we find that spin-orbit interactions have negligible impact on the six spin-degenerate low-energy conduction bands 666DFT bandstructure calculations for HfS2 monolayers that employ different pseudo-potentials can result in more significant, albeit still small, SOI in the low-energy conduction bands (see, e.g., Fig. 9 of Ref. [44]). Ultimately, what is most relevant for our study is the energetics of the doubly-degenerate lowest energy conduction bands near M, and for these states SOI has almost no impact., which remain energetically isolated from all other bands throughout the BZ and vary weakly with 𝗱\bm{\mathsf{d}} (see Supplementary Material and Fig. 1), demonstrating the weak interlayer coupling needed for the accuracy of local displacement moiré band models. Within each szs_{z} spin sector, the low-energy conduction bands are accurately described throughout the BZ by a six-orbital tight-binding Hamiltonian HTB(𝗱)H_{\text{TB}}(\bm{\mathsf{d}}). The conduction band minima occur near the three MM points (𝒦{𝑴1,𝑴2,𝑴3}\mathcal{K}\equiv\{\bm{M}_{1},\bm{M}_{2},\bm{M}_{3}\}).

In principle, we could construct a moiré band model with three orbitals plus spin in each layer by retaining all orbitals needed for the Wanner processing. We find however that near each 𝑴ν𝒦\bm{M}_{\nu}\in\mathcal{K}, the two lowest energy bands are separated from the other four bands retained in the tight-binding model by 2\approx 2 eV. This motivates the construction of a simpler and more physically transparent model with only layer and spin degrees of freedom that still accounts for the higher energy bands perturbatively. To do so we re-express the 6×66\times 6 matrix TB(𝒌;𝗱)\mathcal{H}_{\text{TB}}(\bm{k};\bm{\mathsf{d}}), which has elements Hl,βl,szl,αl,sz(𝒌;𝗱)H^{l,\alpha_{l},s_{z}}_{l^{\prime},\beta_{l^{\prime}},s_{z}}(\bm{k};\bm{\mathsf{d}}), in the basis of the Bloch energy eigenvectors at the relevant valley center 𝑴ν𝒦\bm{M}_{\nu}\in\mathcal{K} and at an arbitrary reference displacement 𝗱0=𝟎\bm{\mathsf{d}}_{0}=\bm{0}. Rewritten in this basis, TB(𝒌;𝗱)\mathcal{H}_{\text{TB}}(\bm{k};\bm{\mathsf{d}}) takes the form

𝑴ν(𝒌;𝗱)=(lowE(𝒌;𝗱)T(𝒌;𝗱)T(𝒌;𝗱)highE(𝒌;𝗱)),\displaystyle\mathcal{H}_{\bm{M}_{\nu}}(\bm{k};\bm{\mathsf{d}})=\begin{pmatrix}\mathcal{H}_{\text{lowE}}(\bm{k};\bm{\mathsf{d}})&T(\bm{k};\bm{\mathsf{d}})\\ T^{\dagger}(\bm{k};\bm{\mathsf{d}})&\mathcal{H}_{\text{highE}}(\bm{k};\bm{\mathsf{d}})\end{pmatrix}, (6)

where lowE(𝒌;𝗱)\mathcal{H}_{\text{lowE}}(\bm{k};\bm{\mathsf{d}}) is a 2×22\times 2 matrix, highE(𝒌;𝗱)\mathcal{H}_{\text{highE}}(\bm{k};\bm{\mathsf{d}}) is a 4×44\times 4 matrix and T(𝒌;𝗱)T(\bm{k};\bm{\mathsf{d}}) is a 2×42\times 4 matrix. lowE(𝒌;𝗱)\mathcal{H}_{\text{lowE}}(\bm{k};\bm{\mathsf{d}}) is the low-energy sector (of the isolated low-energy conduction bands near 𝑴ν\bm{M}_{\nu}) and highE(𝒌;𝗱)\mathcal{H}_{\text{highE}}(\bm{k};\bm{\mathsf{d}}) is the high-energy sector. Defining the zero of energy at the low-energy sector eigenvalues (which are approximately equal), to second-order in perturbation theory 777See, e.g., Appendix B of Ref. [45] and Ref. [46]. we obtain

eff𝑴ν(𝒌;𝗱)\displaystyle\mathcal{H}^{\bm{M}_{\nu}}_{\text{eff}}(\bm{k};\bm{\mathsf{d}}) =lowE(𝒌;𝗱)T(𝒌;𝗱)highE1(𝒌;𝗱)T(𝒌;𝗱).\displaystyle=\mathcal{H}_{\text{lowE}}(\bm{k};\bm{\mathsf{d}})-T^{\dagger}(\bm{k};\bm{\mathsf{d}})\mathcal{H}_{\text{highE}}^{-1}(\bm{k};\bm{\mathsf{d}})T(\bm{k};\bm{\mathsf{d}}). (7)
Refer to caption
Figure 2: (a) Intralayer potential Δl(𝑴1;𝗱)\Delta_{l}(\bm{M}_{1};\bm{\mathsf{d}}) and (b) interlayer tunneling amplitude ΔT(𝑴1;𝗱)\Delta_{T}(\bm{M}_{1};\bm{\mathsf{d}}) vs. stacking 𝗱\bm{\mathsf{d}}. Black parallelograms identify the crystal unit cell of HfS2. Note that the interlayer tunneling is periodic in a doubled unit cell (along 𝒂1\bm{a}_{1} for valley 𝑴1=𝒃1/2\bm{M}_{1}=\bm{b}_{1}/2) due to the 𝗱\bm{\mathsf{d}}-dependence of the WFs discussed in the main text. The 𝗱\bm{\mathsf{d}}-dependence of these terms is scaled to position-dependence on the moiré scale to model twisted bilayers.
𝑮=n1𝒃1+n2𝒃2\bm{G}=n_{1}\bm{b}_{1}+n_{2}\bm{b}_{2} l=ll=l^{\prime} (10310^{-3} eV) lll\neq l^{\prime} (10310^{-3} eV)
n1=0,n2=0n_{1}=0,n_{2}=0 arb. 42.0+8.0𝒊\bm{-42.0+8.0i}
n1=1,n2=0n_{1}=1,n_{2}=0 3.5+3.5i3.5+3.5i 0.30+1.59i0.30+1.59i
n1=1,n2=0n_{1}=-1,n_{2}=0 3.53.5i3.5-3.5i 42.08.0𝒊\bm{-42.0-8.0i}
n1=0,n2=1n_{1}=0,n_{2}=1 3.2+0.0i-3.2+0.0i 29.00.0𝒊\bm{29.0-0.0i}
n1=0,n2=1n_{1}=0,n_{2}=-1 3.20.0i-3.2-0.0i 0.083.21i-0.08-3.21i
n1=1,n2=1n_{1}=1,n_{2}=1 3.20.0i-3.2-0.0i 0.213.2i-0.21-3.2i
n1=1,n2=1n_{1}=-1,n_{2}=-1 3.2+0.0i-3.2+0.0i 29.0+0.0𝒊\bm{29.0+0.0i}
n1=1,n2=1n_{1}=-1,n_{2}=1 0.41.0i-0.4-1.0i 0.08+3.2i-0.08+3.2i
n1=1,n2=1n_{1}=1,n_{2}=-1 0.4+1.0i-0.4+1.0i 0.130.14i-0.13-0.14i
Table 1: Zeroth, first, and two second shell Fourier components of the intralayer potential (l=ll=l^{\prime}) and phase-modified interlayer tunneling (lll\neq l^{\prime}) for 𝑴1=𝒃1/2\bm{M}_{1}=\bm{b}_{1}/2 rounded to 0.1 meV. Symmetry of the lll\neq l^{\prime} elements is most apparent when explicitly accounting for the top-to-bottom layer momentum boost; i.e., considering 𝑸i𝑴1+𝑮i\bm{Q}_{i}\equiv\bm{M}_{1}+\bm{G}_{i}. The first shell of 𝑸i\bm{Q}_{i} vectors is given by (n1,n2)=(0,0),(1,0)(n_{1},n_{2})=(0,0),\,(-1,0), the second shell by (n1,n2)=(0,1),(1,1)(n_{1},n_{2})=(0,1),\,(-1,-1), and the third shell by (n1,n2)=(0,1),(1,1),(1,1),(2,1)(n_{1},n_{2})=(0,-1),\,(1,1),\,(-1,1),\,(-2,-1). Together with the effective masses m,l𝑴1=0.27mem_{\perp,l}^{\bm{M}_{1}}=0.27m_{e} and m,l𝑴1=2.41mem_{\parallel,l}^{\bm{M}_{1}}=2.41m_{e} for mem_{e} the bare electron mass, these values define the moiré model at valley 𝑴1\bm{M}_{1}.

When we diagonalize the matrix (7), we find that its eigenvectors are even and odd combinations of Bloch-type states in top and bottom layers at all values of 𝗱\bm{\mathsf{d}}: |ϕ±,𝒌𝑴ν(𝗱)=|ϕ~b,𝒌𝑴ν(𝗱)|ϕ~t,𝒌𝑴ν(𝗱)\ket{\phi^{\bm{M}_{\nu}}_{\pm,\bm{k}}(\bm{\mathsf{d}})}=\ket{\tilde{\phi}^{\bm{M}_{\nu}}_{b,\bm{k}}(\bm{\mathsf{d}})}\mp\ket{\tilde{\phi}^{\bm{M}_{\nu}}_{t,\bm{k}}(\bm{\mathsf{d}})}, where |ϕ~l,𝒌𝑴ν(𝗱)=α=13Cα𝑴ν(𝒌,𝗱)|ψ~(l,α,sz),𝑴ν(𝗱0)\ket{\tilde{\phi}^{\bm{M}_{\nu}}_{l,\bm{k}}(\bm{\mathsf{d}})}=\sum_{\alpha=1}^{3}C^{\bm{M}_{\nu}}_{\alpha}(\bm{k},\bm{\mathsf{d}})\ket{\tilde{\psi}_{(l,\alpha,s_{z}),\bm{M}_{\nu}}(\bm{\mathsf{d}}_{0})}. When transformed back to a layer representation, the average of the eigenvalues can be interpreted as potentials that are identical in each layer and half the difference as a real tunneling amplitude between layers. These properties are expected since each 𝑴ν𝒦\bm{M}_{\nu}\in\mathcal{K} is time-reversal invariant and the system has inversion symmetry which implies that top and bottom intralayer potentials are identical and interlayer tunneling amplitude is real-valued. The 𝗱\bm{\mathsf{d}}-dependence of the intralayer potential Δl(𝑴ν;𝗱)\Delta_{l}(\bm{M}_{\nu};\bm{\mathsf{d}}) and of the interlayer tunneling ΔT(𝑴ν;𝗱)\Delta_{T}(\bm{M}_{\nu};\bm{\mathsf{d}}) for valley 𝑴1=𝒃1/2\bm{M}_{1}=\bm{b}_{1}/2 is plotted in Fig. 2. In Table 1 we list all sizable Fourier components of Δl(𝑴1;𝗱)\Delta_{l}(\bm{M}_{1};\bm{\mathsf{d}}) and of Δ¯𝔱𝔟(𝑴1;𝗱)ei𝑴1𝗱ΔT(𝑴1;𝗱)\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};\bm{\mathsf{d}})\equiv e^{i\bm{M}_{1}\cdot\bm{\mathsf{d}}}\Delta_{T}(\bm{M}_{1};\bm{\mathsf{d}}) for top-to-bottom tunneling (recall Eq. (3)). As anticipated only a small number of parameters are needed to define the moiré band model for valley 𝑴1\bm{M}_{1}. The valley-projected Hamiltonians for other valleys are related to that for valley 𝑴1\bm{M}_{1} by the C3C_{3} rotational symmetry of HfS2, as we now describe.

The spatial symmetries of HfS2 bilayers constrain the form of the Hamiltonian matrix elements Hj,βj,szi,αi,sz(𝒌;𝗱)H^{i,\alpha_{i},s_{z}}_{j,\beta_{j},s_{z}}(\bm{k};\bm{\mathsf{d}}) and therefore also of the effective masses ma,l𝑴νm_{a,l}^{\bm{M}_{\nu}} (defined precisely below), Δl(𝑴ν;𝗱)\Delta_{l}(\bm{M}_{\nu};\bm{\mathsf{d}}), and ΔT(𝑴ν;𝗱)\Delta_{T}(\bm{M}_{\nu};\bm{\mathsf{d}}). In bulk, each Hf ion location in 1T-HfS2 has D3dD_{3d} point group symmetry. This point group symmetry is inherited in the bilayer with 𝗱=𝟎\bm{\mathsf{d}}=\bm{0} (albeit no longer about Hf sites) and, in fact, its elements have counterparts in bilayers with general 𝗱\bm{\mathsf{d}}, namely: (a) a spatial inversion symmetry, which implies that the effective masses are layer independent (ma,𝔟𝑴ν=ma,𝔱𝑴νm_{a,\mathfrak{b}}^{\bm{M}_{\nu}}=m_{a,\mathfrak{t}}^{\bm{M}_{\nu}}), the potentials are layer independent (Δ𝔟(𝑴ν;𝗱)=Δ𝔱(𝑴ν;𝗱)\Delta_{\mathfrak{b}}(\bm{M}_{\nu};\bm{\mathsf{d}})=\Delta_{\mathfrak{t}}(\bm{M}_{\nu};\bm{\mathsf{d}})), and the tunneling is real (ΔT(𝑴ν;𝗱)=ΔT(𝑴ν;𝗱)\Delta_{T}(\bm{M}_{\nu};\bm{\mathsf{d}})=\Delta_{T}^{*}(\bm{M}_{\nu};\bm{\mathsf{d}})). (Here and below we use that 𝑴ν\bm{M}_{\nu} is equivalent to 𝑴ν-\bm{M}_{\nu}); (b) a symmetry that involves the rotation of the top layer by Θ=2πn/3\Theta=2\pi n/3 for nn\in\mathbb{Z} about the surface-normal 𝒆^z\hat{\bm{e}}_{z} relative to the bottom layer, which implies ma,l𝑴ν=ma,lRΘz(𝑴ν)m_{a,l}^{\bm{M}_{\nu}}=m_{a,l}^{R^{z}_{\Theta}(\bm{M}_{\nu})} 888\parallel” and “\perp” on LHS and RHS are with respect to 𝑴ν\bm{M}_{\nu} and RΘz(𝑴ν)R^{z}_{\Theta}(\bm{M}_{\nu}), respectively. and Δμ(𝑴ν;𝗱)=Δμ(RΘz(𝑴ν);RΘz(𝗱))\Delta_{\mu}(\bm{M}_{\nu};\bm{\mathsf{d}})=\Delta_{\mu}(R^{z}_{\Theta}(\bm{M}_{\nu});R^{z}_{\Theta}(\bm{\mathsf{d}})) for μ{b,t,T}\mu\in\{b,t,T\}; and (c) three symmetries that involve a total rotation by π\pi about in-plane axes 𝒆^i\hat{\bm{e}}_{i} (i{A,B,C}i\in\{\text{A,B,C}\}, which are parallel to one of the three lines connecting nearest-neighbor Hf ions and positioned at midway between the layers) followed by 𝗱Rπi(𝗱)\bm{\mathsf{d}}\rightarrow-R^{i}_{\pi}(\bm{\mathsf{d}}), which for the 𝑴1=𝒃1/2\bm{M}_{1}=\bm{b}_{1}/2 valley implies Δ𝔟(𝑴1;𝒃2)=Δ𝔱(𝑴1;𝒃1𝒃2)\Delta_{\mathfrak{b}}(\bm{M}_{1};\bm{b}_{2})=\Delta_{\mathfrak{t}}(\bm{M}_{1};-\bm{b}_{1}-\bm{b}_{2}), Δ¯𝔱𝔟(𝑴1;𝟎)=Δ¯𝔱𝔟(𝑴1;𝒃1)\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};\bm{0})=\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};-\bm{b}_{1})^{*}, Δ¯𝔱𝔟(𝑴1;𝒃2)\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};\bm{b}_{2})\in\mathbb{R}, Δ¯𝔱𝔟(𝑴1;𝒃1𝒃2)\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};-\bm{b}_{1}-\bm{b}_{2})\in\mathbb{R}, Δ¯𝔱𝔟(𝑴1;𝒃1+𝒃2)=Δ¯𝔱𝔟(𝑴1;𝒃1+𝒃2)\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};\bm{b}_{1}+\bm{b}_{2})=\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};-\bm{b}_{1}+\bm{b}_{2})^{*}. Since szs_{z} spin sectors are identical, time-reversal symmetry implies Δ¯𝔱𝔟(𝑴1;𝑮)=Δ¯𝔱𝔟(𝑴1;𝑮𝒃1)\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};\bm{G})=\bar{\Delta}_{\mathfrak{t}}^{\mathfrak{b}}(\bm{M}_{1};-\bm{G}-\bm{b}_{1})^{*}. This is a minimal set of symmetry relations which explains the most prominent structure of Fig. 2 via Table 1.

Two-orbital anisotropic moiré band model— We now address the position-independent (intralayer 𝑮=𝟎\bm{G}=\bm{0}) term in the moiré band Hamiltonian of Eq. (5), which is obtained by averaging the intralayer Hamiltonian matrix elements over 𝗱\bm{\mathsf{d}}. When described at parabolic order, the effective mass tensor has eigenvectors in the 𝒌\bm{k}-space directions parallel (a=a=\,\parallel) and perpendicular (a=a=\,\perp) to 𝑴ν\bm{M}_{\nu}, which we have denoted above by ma,l𝑴νm_{a,l}^{\bm{M}_{\nu}}. The effective mass eigenvalues are weakly dependent on 𝗱\bm{\mathsf{d}} and approximated here by their spatial averages 999Position-dependent effective masses can be included in the moiré band Hamiltonian when needed.. The explicit form of the plane-wave representation two-orbital continuum model Hamiltonian is

l,𝖖+𝑮~|HLD|l,𝖖+𝑮~=\displaystyle\bra{l,\bm{\mathfrak{q}}+\tilde{\bm{G}}}H_{\text{LD}}\ket{l^{\prime},\bm{\mathfrak{q}}^{\prime}+\tilde{\bm{G}^{\prime}}}=
δl,lδ(𝖖𝖖)(δ𝑮~,𝑮~a{,}22ma,l𝑴ν(𝔮a+G~a)2\displaystyle\delta_{l,l^{\prime}}\delta\big{(}\bm{\mathfrak{q}}^{\prime}-\bm{\mathfrak{q}}\big{)}\Big{(}\delta_{\tilde{\bm{G}}^{\prime},\tilde{\bm{G}}}\sum_{a\in\{\parallel,\perp\}}\frac{\hbar^{2}}{2m_{a,l}^{\bm{M}_{\nu}}}\big{(}\mathfrak{q}^{a}+\tilde{G}^{a}\big{)}^{2}
+Δl(𝑴ν;𝑮~𝑮~))\displaystyle\qquad\qquad\qquad\quad+\Delta_{l}(\bm{M}_{\nu};\tilde{\bm{G}}^{\prime}-\tilde{\bm{G}})\Big{)}
+(1δl,l)δ(𝖖𝖖+(1)l,𝔱𝒌~)Δ¯ll(𝑴ν;𝑮~𝑮~),\displaystyle+(1-\delta_{l,l^{\prime}})\delta\big{(}\bm{\mathfrak{q}}^{\prime}-\bm{\mathfrak{q}}+(-1)^{l,\mathfrak{t}}\tilde{\bm{k}}_{*}\big{)}\bar{\Delta}_{l^{\prime}}^{l}(\bm{M}_{\nu};\tilde{\bm{G}}^{\prime}-\tilde{\bm{G}}), (8)

where k𝒌𝑴^νk^{\parallel}\equiv\bm{k}\cdot\hat{\bm{M}}_{\nu}, k𝒌(𝒆^z×𝑴^ν)k^{\perp}\equiv\bm{k}\cdot(\hat{\bm{e}}_{z}\crossproduct\hat{\bm{M}}_{\nu}), and 𝑴^ν𝑴ν/|𝑴ν|\hat{\bm{M}}_{\nu}\equiv\bm{M}_{\nu}/|\bm{M}_{\nu}|.

Refer to caption
Figure 3: (a) Schematic illustration of the moiré Brillouin zone (black hexagon), where the red (green) hexagon illustrates the bottom (top) monolayer BZ. (b) Moiré (𝑮~i\tilde{\bm{G}}_{i}) and crystalline HfS2 bilayer (𝑮i\bm{G}_{i}) reciprocal lattice vectors. Points in mBZ defined relative to 𝑴1\bm{M}_{1}; m=(0,0)m^{-}=(0,0), m+=𝑴~1m^{+}=\tilde{\bm{M}}_{1}, γ=𝑮~1/2\gamma=\tilde{\bm{G}}_{1}/2, κ=γ(𝑮~1+𝑮~2)/3\kappa=\gamma-(\tilde{\bm{G}}_{1}+\tilde{\bm{G}}_{2})/3. (c) 𝑴1\bm{M}_{1}-valley moiré conduction bands for twist angle θ=5\theta=5^{\circ}. (d) Contour plot of the lowest energy moiré band for θ=5\theta=5^{\circ} and (e) its bandwidth for various values of θ\theta (blue) compared to those in a hypothetical twisted bilayer with isotropic mass equal to the light m,l𝑴1m^{\bm{M}_{1}}_{\perp,l}.
Refer to caption
Figure 4: Density-distribution vs. position 𝒓\bm{r} for the lowest band’s moiré Bloch eigenfunction at different moiré BZ momenta and twist angles.

Our findings for the 𝑴1\bm{M}_{1} valley are summarized in Fig. 3 and additional plots are included in the Supplementary Materials. The effective masses in HfS2 crystal bilayers are highly anisotropic as evidenced in Fig. 1(c)–(e); m,l𝑴ν/m,l𝑴ν9m_{\parallel,l}^{\bm{M}_{\nu}}/m_{\perp,l}^{\bm{M}_{\nu}}\approx 9. As shown in Figs. 3(c) and (d), this leads in the 𝑴1\bm{M}_{1} valley to moiré bands that are significantly much more dispersive along the 𝑴ν\perp\bm{M}_{\nu} direction which is along the 𝒃~1\tilde{\bm{b}}_{1} direction in the moiré BZ (the mm^{-} to m+m^{+} line), and less dispersive along the direction parallel to 𝑴1\bm{M}_{1} (the κ\kappa to γ\gamma line). In real space (see Fig. 4), the eigenfunctions in the lowest energy moiré conduction band are more localized close to favored positions along the large mass directions. Up to an overall scaling factor related to variation in bandwidth, the shape of that band is qualitatively independent of θ\theta for small twist angles (see Supplementary Material). Fig. 4 shows that the spatial distribution of flat-band charge is more momentum dependent within a band than in the case of atomic crystals, which have strong attractive Coulombic potentials centered on nuclear positions. This property is shared with other moiré material systems and implies that interaction effects can shape not only the widths of bands but also their shapes as band filling factors change.

Valleys at time-reversal invariant momenta and inversion symmetry together imply ΔT(𝑴ν;𝗱)\Delta_{T}(\bm{M}_{\nu};\bm{\mathsf{d}}) is real-valued and top and bottom intralayer potentials are identical. Therefore, the previously defined [40] layer pseudo-spin map Δ\Delta is characterized by degree 0. Layer textures can be induced in real and momentum space by applying electrostatic gate fields between layers, but these can never yield Chern bands because of time-reversal symmetry. This implies that the Hall conductivity within each valley vanishes and therefore cannot be used as a harbinger of interaction-induced spontaneous valley polarization, as commonly seen in graphene multilayers and KK-valley group VI TMDs. However, unlike in those materials, the valley-projected bands in t(HfS2)/HfS2 are highly anisotropic and a rotational symmetry implies that the high and low mass directions (in the crystal BZ) are rotated between valleys.

Discussion— In this Letter we employ the WF-based local displacement scheme [3] to construct a moiré band model of the low-energy electronic conduction states in t(HfS2)/HfS2, a prototypical group IV TMD homobilayer. This approach provides a universal moiré band Hamiltonian that applies at all small twist angles by evaluating displacement-dependent corrections to the Hamiltonian that are small on the atomic scale but significant on the moiré scale. We have neglected in-plane strain relaxation, but this can be included by modifying the mapping 𝒅:22\bm{d}:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} between lattice sites and position from the linear form that applies for rigid rotations [41]. Although strain effects can be important for a quantitatively accurate moiré bandstructure, the novel qualitative aspects of these moiré materials is due to time-reversal invariant valley momenta and effective mass anisotropy, which will persist in the presence of strain relaxation. We find that the low-energy conduction bands are determined mainly by interlayer tunneling terms in each valley-projected moiré band Hamiltonian that are peaked near metal on chalcogen layer configurations and have a typical size 100\sim 100 meV, and secondarily by intralayer potential variations that are identical in the two layers and vary by 30\sim 30 meV over the moiré unit cell. The tunneling amplitudes are real because of time-reversal invariance in each valley and this implies that the moiré band Hamiltonian separates into even and odd layer-parity sectors.

Monolayer HfS2 has anisotropic dispersion near its conduction band minimum and this leads to valley-projected moiré bands that are also anisotropic. We find that the lowest energy moiré conduction bands are narrow and isolated, the established recipe for interesting strong correlation physics, with bandwidths below 11 meV for θ4\theta\lesssim 4^{\circ}. At much smaller twist angles we expect classical lattice-gas physics in which hopping between sites plays no role. It appears therefore that the interesting range of twist angles is larger than in the well-studied KK-valley moiré systems. Spin-valley flavor magnetism appears likely to be as common in MM-valley TMDs as it is in KK-valley TMDs, and this can lead to insulating states at integer band fillings ν\nu, smaller than the filling ν=6\nu=6 at which the lowest band is filled. The mismatch in shape between Fermi surface contours in different valleys favors valley polarization over inter-valley coherence [42]. Anisotropy in transport is therefore a likely signature of strong correlation physics in MM-valley TMDs. It seems likely that the strongly anisotropic bands may bring the tendency toward density-wave states and other aspects of quasi-one-dimensional material physics to the moiré platform.

Acknowledgments— We are grateful to Andrei Bernevig, Dumitru Călugăru, Yi Jiang, Haoyu Hu, Hanqi Pi and collaborators for insightful discussions near the conclusion of this work. We acknowledge HPC resources provided by the Texas Advanced Computing Center at The University of Texas at Austin. This work was supported by a Simons Foundation Collaborative Research Grant and by Robert A. Welch Foundation Grant F-2112.

Note Added— As this manuscript was being prepared for publication we learned of a closely related study by Călugăru et al. [43] which reaches similar conclusions.

References

  • Andrei et al. [2021] E. Y. Andrei, D. K. Efetov, P. Jarillo-Herrero, A. H. MacDonald, K. F. Mak, T. Senthil, E. Tutuc, A. Yazdani, and A. F. Young, The marvels of moiré materials, Nature Reviews Materials 6, 201 (2021).
  • 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).
  • Jung et al. [2014] 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).
  • Mak and Shan [2022] K. F. Mak and J. Shan, Semiconductor moiré materials, Nature Nanotechnology 17, 686 (2022).
  • Lopes dos Santos et al. [2007] J. M. B. Lopes dos Santos, N. M. R. Peres, and A. H. Castro Neto, Graphene bilayer with a twist: Electronic structure, Phys. Rev. Lett. 99, 256802 (2007).
  • Bernevig et al. [2021] B. A. Bernevig, Z.-D. Song, N. Regnault, and B. Lian, Twisted bilayer graphene. i. matrix elements, approximations, perturbation theory, and a kpk\cdot{}p two-band model, Phys. Rev. B 103, 205411 (2021).
  • Zou et al. [2018] L. Zou, H. C. Po, A. Vishwanath, and T. Senthil, Band structure of twisted bilayer graphene: Emergent symmetries, commensurate approximants, and wannier obstructions, Phys. Rev. B 98, 085435 (2018).
  • Wu et al. [2018] F. Wu, T. Lovorn, E. Tutuc, and A. H. MacDonald, Hubbard model physics in transition metal dichalcogenide moiré bands, Physical review letters 121, 026402 (2018).
  • Wu et al. [2019a] F. Wu, T. Lovorn, E. Tutuc, I. Martin, and A. MacDonald, Topological insulators in twisted transition metal dichalcogenide homobilayers, Physical review letters 122, 086402 (2019a).
  • Devakul et al. [2021] T. Devakul, V. Crépel, Y. Zhang, and L. Fu, Magic in twisted transition metal dichalcogenide bilayers, Nature communications 12, 6730 (2021).
  • Angeli and MacDonald [2021] M. Angeli and A. H. MacDonald, γ\gamma valley transition metal dichalcogenide moiré bands, Proceedings of the National Academy of Sciences 118, e2021826118 (2021).
  • Randeria et al. [2019] M. T. Randeria, K. Agarwal, B. E. Feldman, H. Ding, H. Ji, R. J. Cava, S. L. Sondhi, S. A. Parameswaran, and A. Yazdani, Interacting multi-channel topological boundary modes in a quantum hall valley system, Nature 566, 363–367 (2019).
  • Liu et al. [2019] J. Liu, Z. Ma, J. Gao, and X. Dai, Quantum valley hall effect, orbital magnetism, and anomalous hall effect in twisted multilayer graphene systems, Phys. Rev. X 9, 031021 (2019).
  • Xie and MacDonald [2020] M. Xie and A. H. MacDonald, Nature of the correlated insulator states in twisted bilayer graphene, Physical Review Letters 12410.1103/physrevlett.124.097601 (2020).
  • Reddy et al. [2023] A. P. Reddy, F. Alsallom, Y. Zhang, T. Devakul, and L. Fu, Fractional quantum anomalous hall states in twisted bilayer mote2{\mathrm{mote}}_{2} and wse2{\mathrm{wse}}_{2}Phys. Rev. B 108, 085117 (2023).
  • Tao et al. [2024] Z. Tao, B. Shen, S. Jiang, T. Li, L. Li, L. Ma, W. Zhao, J. Hu, K. Pistunova, K. Watanabe, T. Taniguchi, T. F. Heinz, K. F. Mak, and J. Shan, Valley-coherent quantum anomalous hall state in ab-stacked mote2/wse2{\mathrm{mote}}_{2}/{\mathrm{w}\mathrm{s}\mathrm{e}}_{2} bilayers, Phys. Rev. X 14, 011004 (2024).
  • Sharpe et al. [2019] A. L. Sharpe, E. J. Fox, A. W. Barnard, J. Finney, K. Watanabe, T. Taniguchi, M. A. Kastner, and D. Goldhaber-Gordon, Emergent ferromagnetism near three-quarters filling in twisted bilayer graphene, Science 365, 605–608 (2019).
  • Serlin et al. [2020] M. Serlin, C. L. Tschirhart, H. Polshyn, Y. Zhang, J. Zhu, K. Watanabe, T. Taniguchi, L. Balents, and A. F. Young, Intrinsic quantized anomalous hall effect in a moiré heterostructure, Science 367, 900–903 (2020).
  • Li et al. [2021] T. Li, S. Jiang, B. Shen, Y. Zhang, L. Li, Z. Tao, T. Devakul, K. Watanabe, T. Taniguchi, L. Fu, J. Shan, and K. F. Mak, Quantum anomalous hall effect from intertwined moiré bands, Nature 600, 641–646 (2021).
  • Zeng et al. [2023] Y. Zeng, Z. Xia, K. Kang, J. Zhu, P. Knüppel, C. Vaswani, K. Watanabe, T. Taniguchi, K. F. Mak, and J. Shan, Thermodynamic evidence of fractional Chern insulator in moiré MoTe2, Nature 622, 69 (2023).
  • Cai et al. [2023] J. Cai, E. Anderson, C. Wang, X. Zhang, X. Liu, W. Holtzmann, Y. Zhang, F. Fan, T. Taniguchi, K. Watanabe, Y. Ran, T. Cao, L. Fu, D. Xiao, W. Yao, and X. Xu, Signatures of fractional quantum anomalous Hall states in twisted MoTe2, Nature 622, 63 (2023).
  • Park et al. [2023] H. Park, J. Cai, E. Anderson, Y. Zhang, J. Zhu, X. Liu, C. Wang, W. Holtzmann, C. Hu, Z. Liu, T. Taniguchi, K. Watanabe, J.-H. Chu, T. Cao, L. Fu, W. Yao, C.-Z. Chang, D. Cobden, D. Xiao, and X. Xu, Observation of fractionally quantized anomalous Hall effect, Nature 622, 74 (2023).
  • Xu et al. [2023] F. Xu, Z. Sun, T. Jia, C. Liu, C. Xu, C. Li, Y. Gu, K. Watanabe, T. Taniguchi, B. Tong, J. Jia, Z. Shi, S. Jiang, Y. Zhang, X. Liu, and T. Li, Observation of Integer and Fractional Quantum Anomalous Hall Effects in Twisted Bilayer MoTe 2, Phys. Rev. X 13, 031037 (2023).
  • Lu et al. [2024] Z. Lu, T. Han, Y. Yao, A. P. Reddy, J. Yang, J. Seo, K. Watanabe, T. Taniguchi, L. Fu, and L. Ju, Fractional quantum anomalous Hall effect in multilayer graphene, Nature 626, 759 (2024).
  • Yan et al. [2018] C. Yan, C. Gong, P. Wangyang, J. Chu, K. Hu, C. Li, X. Wang, X. Du, T. Zhai, Y. Li, et al., 2d group ivb transition metal dichalcogenides, Advanced Functional Materials 28, 1803305 (2018).
  • Zhao et al. [2017] Q. Zhao, Y. Guo, K. Si, Z. Ren, J. Bai, and X. Xu, Elastic, electronic, and dielectric properties of bulk and monolayer zrs2, zrse2, hfs2, hfse2 from van der waals density-functional theory, Physica Status Solidi (b) 254, 1700033 (2017).
  • Marzari et al. [2012] N. Marzari, A. A. Mostofi, J. R. Yates, I. Souza, and D. Vanderbilt, Maximally localized wannier functions: Theory and applications, Rev. Mod. Phys. 84, 1419 (2012).
  • Giannozzi et al. [2009] P. Giannozzi, S. Baroni, N. Bonini, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli, G. L. Chiarotti, M. Cococcioni, I. Dabo, A. Dal Corso, S. de Gironcoli, S. Fabris, G. Fratesi, R. Gebauer, U. Gerstmann, C. Gougoussis, A. Kokalj, M. Lazzeri, L. Martin-Samos, N. Marzari, F. Mauri, R. Mazzarello, S. Paolini, A. Pasquarello, L. Paulatto, C. Sbraccia, S. Scandolo, G. Sclauzero, A. P. Seitsonen, A. Smogunov, P. Umari, and R. M. Wentzcovitch, Quantum espresso: a modular and open-source software project for quantum simulations of materials, Journal of Physics: Condensed Matter 21, 395502 (19pp) (2009).
  • Note [1] For every 𝗱2\bm{\mathsf{d}}\in\mathbb{R}^{2} the Bloch Hamiltonian H(𝒓,𝒑(𝒓);𝗱)H(\bm{r},\bm{p}(\bm{r});\bm{\mathsf{d}}) of the bilayer is Λ\Lambda-periodic in 𝒓\bm{r} where Λ\Lambda is the Bravais lattice of the (arbitrarily chosen) bottom layer. The vertical separation between layers is fixed at each 𝗱\bm{\mathsf{d}} by minimizing energy. Therefore, we can (and will) choose the real-space Bravais lattice to be independent of 𝗱\bm{\mathsf{d}}. The same is true of the unit cell Ωuc\Omega_{uc} (BZ) of Λ\Lambda (Λ\Lambda^{*}).
  • Brouder et al. [2007] C. Brouder, G. Panati, M. Calandra, C. Mourougane, and N. Marzari, Exponential localization of wannier functions in insulators, Phys. Rev. Lett. 98, 046402 (2007).
  • Note [2] That H(𝗱)H(\bm{\mathsf{d}}) has time-reversal symmetry is not a necessary criterion, but is valid in HfS2 bilayers and simplifies our discussion. We also assume that at each 𝒌BZ\bm{k}\in\text{BZ} the set of isolated conduction (valence) bands remains energetically isolated from all other bands as 𝗱\bm{\mathsf{d}} is varied. Again, this is not necessary, but is convenient and valid in HfS2 bilayers.
  • Note [3] The quasicrystal has top (bottom) layer Bravais lattice Λt\Lambda_{t} (Λb\Lambda_{b}) and we arbitrarily choose ΛΛb\Lambda\equiv\Lambda_{b}. By definition, Λt=λRθz(Λb)\Lambda_{t}=\lambda R_{\theta}^{z}(\Lambda_{b}) and the operator 𝒅:22\bm{d}:\mathbb{R}^{2}\rightarrow\mathbb{R}^{2} encodes the local displacement of bottom layer lattice site 𝑹Λb\bm{R}\in\Lambda_{b} via λRθz(𝑹)(𝑹+𝒅(𝑹))Λt\lambda R^{z}_{\theta}(\bm{R})\equiv\big{(}\bm{R}+\bm{d}(\bm{R})\big{)}\in\Lambda_{t}.
  • Yu and Cardona [2007] P. Y. Yu and M. Cardona, Fundamentals of Semiconductors, 3rd ed. (Springer Berlin, Heidelberg, 2007).
  • Note [4] The set ΛM{λRθ(𝑮)𝑮:𝑮Λ}2\Lambda_{\text{M}}^{*}\equiv\{\lambda R_{-\theta}(\bm{G})-\bm{G}:\bm{G}\in\Lambda^{*}\}\subset\mathbb{R}^{2} equipped with addition is a Bravais lattice. This is called the moiré reciprocal lattice and its dual, the moiré lattice, is given by ΛM={𝒓2:𝒅(𝖗)Λ}\Lambda_{\text{M}}=\{\bm{r}\in\mathbb{R}^{2}:\bm{d}(\bm{\mathfrak{r}})\in\Lambda\}.
  • Note [5] By AA-stacked we mean that top and bottom layer have identical orientation.
  • Note [6] DFT bandstructure calculations for HfS2 monolayers that employ different pseudo-potentials can result in more significant, albeit still small, SOI in the low-energy conduction bands (see, e.g., Fig. 9 of Ref. [44]). Ultimately, what is most relevant for our study is the energetics of the doubly-degenerate lowest energy conduction bands near M, and for these states SOI has almost no impact.
  • Note [7] See, e.g., Appendix B of Ref. [45] and Ref. [46].
  • Note [8] \parallel” and “\perp” on LHS and RHS are with respect to 𝑴ν\bm{M}_{\nu} and RΘz(𝑴ν)R^{z}_{\Theta}(\bm{M}_{\nu}), respectively.
  • Note [9] Position-dependent effective masses can be included in the moiré band Hamiltonian when needed.
  • Wu et al. [2019b] F. Wu, T. Lovorn, E. Tutuc, I. Martin, and A. H. MacDonald, Topological insulators in twisted transition metal dichalcogenide homobilayers, Phys. Rev. Lett. 122, 086402 (2019b).
  • 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).
  • Li et al. [2016] X. Li, F. Zhang, and A. MacDonald, Su (3) quantum hall ferromagnetism in snte, Physical Review Letters 116, 026803 (2016).
  • Calugaru et al. [2024] D. Calugaru, Y. Jiang, H. Hu, H. Pi, J. Yu, M. G. Vergniory, J. Shan, C. Felser, L. M. Schoop, D. K. Efetov, K. F. Mak, and B. A. Bernevig, Twist to the m-ax(is): A new moiré platform based on m-point twisting,   (2024), to appear online.
  • Habenicht et al. [2018] C. Habenicht, L. Sponza, R. Schuster, M. Knupfer, and B. Büchner, Mapping of the energetically lowest exciton in bulk 1thfs21t-{\mathrm{hfs}}_{2}Phys. Rev. B 98, 155204 (2018).
  • Winkler [2003] R. Winkler, Spin-orbit Coupling Effects in Two-Dimensional Electron and Hole Systems, 1st ed. (Springer Berlin, Heidelberg, 2003).
  • Löwdin [1951] P.-O. Löwdin, A note on the quantum-mechanical perturbation theory, The Journal of Chemical Physics 19, 1396 (1951).