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

Fate of Fermi-arc States in Gapped Weyl Semimetals under Long-ranged Interactions

Yu-Chin Tzeng Department of Physics, National Chung Hsing University, Taichung 40227, Taiwan    Min-Fong Yang [email protected] Department of Applied Physics, Tunghai University, Taichung 40704, Taiwan
Abstract

For noninteracting Weyl semimetals (WSMs), Fermi-arc surface states can arise when there exist gapless Weyl nodes in the bulk single-particle spectrum. However, in the presence of electronic correlations, it is not clear whether this bulk-boundary correspondence still holds or not. Recently, novel correlated phases are predicted to appear in WSMs with long-ranged interactions, in which the bulk Weyl nodes can be gapped but without destroying their topological properties. Here, we explore the fate of the Fermi-arc states under the influence of such long-ranged interactions. After mapping the system onto a one-dimensional interacting Su-Schrieffer-Heeger (SSH) model with two open ends, we employ numerical exact diagonalizations to address the issue whether the Fermi-arc states will be modified. By extrapolating our data to the thermodynamic limit, we find that the zero-energy edge states of the corresponding SSH model still exist for those momenta giving the noninteracting Fermi-arc states. Moreover, this observation applies to both the single-particle and the collective edge excitations. Since the locus of these edge states constitutes the Fermi arcs, the robustness of the Fermi-arc states against long-ranged interactions is thus demonstrated. In particular, the Fermi arcs of single-particle nature can survive even when the single-particle gaps at the Weyl nodes are opened by interactions. Our results illustrate the subtlety in identifying the topological phases of interacting WSMs and show the limitation of the approaches simply by examining the nodal structure of the single-particle spectrum.

I introduction

There is ongoing interest in understanding the effects of electronic correlations on topological quantum states owing to the promise of uncovering new phenomena in condensed matter Budich and Trauzettel (2013); Hohenadler and Assaad (2013); Meng et al. (2014); Rachel (2018). Among such topological phases of matter, Weyl semimetals (WSMs) represent an nontrivial class as they broaden the topological classification of matter beyond the insulators Hasan et al. (2017); Yan and Felser (2017); Burkov (2018); Armitage et al. (2018). In the absence of interaction, the bulk system of WSMs hosts gapless band-touching points, called Weyl nodes. Each of these nodal points acts as a monopole of the Berry curvature in the momentum space with a definite Chern number, which is proportional to the quantized Berry flux around the node. Another key signature of WSMs is the presence of Fermi arcs that connect pairs of projected Weyl nodes with opposite Chern numbers in the surface Brillouin zone Wan et al. (2011). Actually, as a consequence of the bulk-boundary correspondence, the existence of the surface Fermi arcs is guaranteed by the nontrivial topology of the Berry curvature carried by the bulk Weyl nodes. The fascinating physics in WSMs has been understood well at the level of noninteracting description. Thereafter, one of the subjects that need to be further explored is the correlation effects on WSMs.

The robustness of the above features of WSMs against electronic correlations is being actively pursued Witczak-Krempa et al. (2014); Roy et al. (2017); Carlström and Bergholtz (2018a, b); Acheche et al. (2019); Crippa et al. (2020); Laubach et al. (2016); Wang et al. (2020); Morimoto and Nagaosa (2016); Meng and Budich (2019); Zhang and Zubkov (2020); Xu et al. (2019). Based on perturbative approaches, weak short-range interactions was shown to provide merely band renormalization and WSMs are thus stable Witczak-Krempa et al. (2014); Roy et al. (2017); Carlström and Bergholtz (2018a, b). Nevertheless, transitions to insulating states can be induced by sufficiently strong interactions Roy et al. (2017); Acheche et al. (2019); Crippa et al. (2020). Within the nonperturbative mean-field theories, it was found that, in the presence of either the spin density wave Laubach et al. (2016) or the Fulde-Ferrell-Larkin-Ovchinnikov superconducting Wang et al. (2020) instability, the bulk Weyl nodes can be gapped out, while the Fermi arcs keep unchanged. This is achieved by pair annihilation of Weyl nodes coming from the backfolding of the Brillouin zone caused by ordering. Therefore, contrary to the noninteracting case, the emergence of the surface Fermi arcs needs not be associated with the existence of the gapless Weyl nodes.

Nevertheless, the validity of usual approximations for exploring the strongly correlated phases is not so clear. Thus it is worthwhile to study exactly solvable models for searching possible novel physics in the strongly correlated regime. Recently, exactly solvable models of interacting WSMs have been investigated in Refs. Morimoto and Nagaosa (2016); Meng and Budich (2019), where the two-body interactions are chosen to be local in momentum space and then of infinite range in real space. The advantage of such models is that different momentum states are decoupled and they thus can be solved exactly through diagonalization for each momentum state. Distinct from the mechanism discussed in Refs. Laubach et al. (2016); Wang et al. (2020), it was shown that the Weyl nodes in these models can be gapped out by interactions but without the pair annihilation. Notably, the topological properties of noninteracting WSMs are kept unchanged: these gapped Weyl nodes still carry nonzero Chern numbers with their noninteracting values, and the Fermi arcs on the surfaces persist against the correlation effects Morimoto and Nagaosa (2016). This novel phase behaves as a topological Mott insulator and is dubbed as the Weyl Mott insulator.

We note that, contrary to their bulk behaviors, the surface properties (say, the Fermi arcs) of these solvable models can not be derived analytically. It is because the translational symmetry is broken in the presence of boundaries and the momenta are not good numbers any more. Usually, the Fermi arc states of the interacting WSMs are calculated by considering an equivalent single-particle Hamiltonian HtH_{t} (i.e., the so-called “topological Hamiltonian” Wang and Zhang (2012); Wang and Yan (2013)) under open boundary conditions Morimoto and Nagaosa (2016); Xu et al. (2019). However, as pointed out in the context of topological insulators, the eigenstates of HtH_{t} have no direct physical meaning Gurarie (2011); Essin and Gurarie (2011). Therefore, the Fermi-arc states derived from HtH_{t} may not be real. In order to provide soild supports on the existence of the Fermi arcs in interacting WSMs, approaches other than that using HtH_{t} are thus necessary.

Moreover, as mentioned in Ref. Yang (2019), the surface Fermi-arc states directly implied by quantized Hall conductances (i.e., quantized Chern numbers) consist of bosonic charge-neutral particle-hole excitations, rather than those of single-particle nature. This conclusion is drawn from the theory of quantum Hall edge states Wen (1991); Renn (1995); Yoshioka , since each two-dimensional (2D) momentum sector between a pair of Weyl nodes can be viewed as a quantum anomalous Hall insulator. That is, the topologically induced Fermi-arc states are the collective edge excitations, whose appearance cannot in general be implied by solving the effective single-particle Hamiltonian HtH_{t}.

In this paper, we revisit the issue of the correlation effects on surface Fermi-arc states of WSMs with long-ranged two-body interactions introduced in Refs. Morimoto and Nagaosa (2016); Meng and Budich (2019). In the noninteracting setting, a single pair of gapless Weyl nodes separated by a distance along the kzk_{z} direction in the momentum space is assumed. In order to construct the Fermi-arc states, we consider a finite-slab geometry with open boundary conditions in the xx direction, but keeping translational symmetry along the other two transverse directions. The transverse momenta 𝐤(ky,kz)\mathbf{k}_{\bot}\equiv(k_{y},\,k_{z}) thus remain good quantum numbers. Utilizing the fact that different momentum states are decoupled in the present case, we can solve our system once for a given 𝐤\mathbf{k}_{\bot}. As a consequence, our problem reduces to a one-dimensional (1D) interacting Su-Schrieffer-Heeger model Su et al. (1979) in the xx direction with two open ends.

By means of exact diagonalization, we search for the edge modes within both the single-particle and the particle-hole subspaces of the Hilbert space. The single-particle edge modes are identified by calculating the energy and the density distribution of the quasi-particle added or removed from the corresponding 1D system at half-filling. Interestingly, we find that the zero-energy edge modes of single-particle excitations still appear in the spectrum, even though the long-ranged interactions induce single-particle gaps at the nodes. Besides, these edge modes occur at the same positions on the surface Brillouin zone as those for the noninteracting Fermi arcs. Actually, there exist tiny single-particle excitation gaps for finite systems. However, their values reduce to zero after extrapolating to the thermodynamic limit. Because the single-particle Fermi arcs are constituted by the collection of such zero-energy edge states, our findings provide direct evidences on their robustness against long-ranged interactions.

On the other hand, the collective edge modes can be recognized by evaluating the energy and the density distribution of the particle-hole excitations with Δ𝐤=0\Delta\mathbf{k}_{\bot}=0. As mentioned above, the zero-energy edge modes among these particle-hole excitations represent the quantum Hall edge states implied by the quantized Hall conductances. Our data show that, the same as those single-particle surface states, the collective Fermi-arc states of particle-hole nature persist to exist as well even in the presence of long-ranged interactions. The extent of these zero-energy collective modes on the surface Brillouin zone is in agreement with the calculated results of the bulk Hall conductances Morimoto and Nagaosa (2016). That is, collective edge states will show up for those kzk_{z}’s such that the bulk Hall conductances (i.e., Chern numbers) keep quantized. Our findings thus indicate that the stability of the collective Fermi-arc surface states originates from the nontrivial topology encoded in the particle-hole subspace, no matter whether the single-particle gaps at the nodes are opened or not.

This paper is organized as follows. Our model Hamiltonian is introduced in Sec. II and some of its bulk properties are summarized. In the second part of this section, the 1D interacting Su-Schrieffer-Heeger model for the study of surface Fermi arcs is derived. Our numerical results are presented in Sec. III. They provide clear evidences for the robustness of the Fermi arcs against long-ranged interactions. We summarize our work in Sec. IV.

II model Hamiltonian

For simplicity, we consider a typical two-band lattice model of WSMs [see Eq. (1)], which host a single pair of Weyl nodes located at momenta 𝐤=(0, 0,±kWeyl)\mathbf{k}=(0,\,0,\,\pm k_{\mathrm{Weyl}}) in the noninteracting setting. For a fixed kzk_{z} with |kz|<kWeyl|k_{z}|<k_{\mathrm{Weyl}}, the 2D slice of the Brillouin zone carries a nonzero Chern number C=1C=1 and shows the quantized anomalous Hall effect, while C=0C=0 otherwise. The region between nodes can thus be viewed as a stack of 2D Chern insulators. If we consider a finite geometry along the direction perpendicular to the zz axis, there appear chiral edge states near the Fermi energy on the boundaries of these 2D Chern insulators. The collection or locus of such edge states constitutes the Fermi arcs, which terminate at the projection of the bulk Weyl nodes onto the surface Brillouin zone Wan et al. (2011).

Following Refs. Morimoto and Nagaosa (2016); Meng and Budich (2019), we explore the influence of the momentum-local two-body interactions on the surface Fermi arcs of such a two-band model of WSMs. Let’s begin with the description of our system and a summary on some of its bulk properties.

II.1 Bulk excitation energies

Our model Hamiltonian under periodic boundary conditions is given by

H\displaystyle H =\displaystyle= H0+HU,\displaystyle H_{0}+H_{U}\;, (1)
H0\displaystyle H_{0} =\displaystyle= 𝐤α,βc𝐤α[𝐡(𝐤)𝝈]αβc𝐤β,\displaystyle\sum_{\mathbf{k}}\sum_{\alpha,\beta}c^{\dagger}_{\mathbf{k}\alpha}[\mathbf{h}(\mathbf{k})\cdot\bm{\sigma}]_{\alpha\beta}c_{\mathbf{k}\beta}\;, (2)
HU\displaystyle H_{U} =\displaystyle= U2𝐤(n𝐤+n𝐤1)2\displaystyle\frac{U}{2}\sum_{\mathbf{k}}\left(n_{\mathbf{k}\uparrow}+n_{\mathbf{k}\downarrow}-1\right)^{2} (3)
=\displaystyle= U𝐤n𝐤n𝐤U2𝐤(n𝐤+n𝐤1).\displaystyle U\sum_{\mathbf{k}}n_{\mathbf{k}\uparrow}n_{\mathbf{k}\downarrow}-\frac{U}{2}\sum_{\mathbf{k}}(n_{\mathbf{k}\uparrow}+n_{\mathbf{k}\downarrow}-1)\;.

Here 𝝈\bm{\sigma} is the vector of three Pauli matrices denoting a (pseudo-)spin-1/21/2 operator, c𝐤αc^{\dagger}_{\mathbf{k}\alpha} represents the creation operator for electrons of (pseudo-)spin α=\alpha=\;\uparrow or \downarrow, and n𝐤α=c𝐤αc𝐤αn_{\mathbf{k}\alpha}=c^{\dagger}_{\mathbf{k}\alpha}c_{\mathbf{k}\alpha}. The first part H0H_{0} describes a noninteracting two-band model of a WSM. To be specific, we take 𝐡(𝐤)=(Mcoskxcoskycoskz,sinkx,sinky)\mathbf{h}(\mathbf{k})=(M-\cos k_{x}-\cos k_{y}-\cos k_{z},\;\sin k_{x},\;\sin k_{y}) with 1<M<31<M<3 not (a). In such a case, only a single pair of Weyl nodes appears at momenta 𝐤=(0, 0,±kWeyl)\mathbf{k}=(0,\,0,\,\pm k_{\mathrm{Weyl}}) with kWeyl=arccos(M2)k_{\mathrm{Weyl}}=\arccos(M-2). The second term HUH_{U} gives a Hubbard-like interaction but being local in momentum space, rather than in real space.

Because different momentum states are decoupled, the bulk properties of this interacting model can be derived exactly through diagonalizing the Hamiltonian in the number representation for each momentum 𝐤\mathbf{k}. The four eigenstates for each 𝐤\mathbf{k} are {|0,b𝐤|0,b𝐤+|0,b𝐤b𝐤+|0}\{|0\rangle,\;b_{\mathbf{k}-}^{\dagger}|0\rangle,\;b_{\mathbf{k}+}^{\dagger}|0\rangle,\;b_{\mathbf{k}-}^{\dagger}b_{\mathbf{k}+}^{\dagger}|0\rangle\} with the corresponding eigenvalues {U/2,h(𝐤),h(𝐤),U/2}\{U/2,\;-h(\mathbf{k}),\;h(\mathbf{k}),\;U/2\}. Here b𝐤+b^{\dagger}_{\mathbf{k}+} and b𝐤b^{\dagger}_{\mathbf{k}-} denote the eigenmodes of the noninteracting Hamiltonian H0H_{0}, |0|0\rangle is the vacuum state, and h(𝐤)=|𝐡(𝐤)|h(\mathbf{k})=|\mathbf{h}(\mathbf{k})|. Therefore, the many-body ground state for positive UU at zero temperature and half-filling becomes the state with a completely-filled lower band, |Ψ0=𝐤b𝐤|0|\Psi_{0}\rangle=\prod_{\mathbf{k}}b_{\mathbf{k}-}^{\dagger}|0\rangle. The single-particle/hole excitations, b𝐤+|Ψ0b_{\mathbf{k}+}^{\dagger}|\Psi_{0}\rangle and b𝐤|Ψ0b_{\mathbf{k}-}|\Psi_{0}\rangle, have nonzero excitation energies ΔE1p/1h=h(𝐤)+U/2\Delta E_{1p/1h}=h(\mathbf{k})+U/2 even at the Weyl nodes such that h(𝐤)=0h(\mathbf{k})=0. The system thus behaves as a Mott insulator with a full gap. However, we note that the excitation energies of the bosonic charge-neutral particle-hole excitations, b𝐤+b𝐤|Ψ0b_{\mathbf{k}+}^{\dagger}b_{\mathbf{k}-}|\Psi_{0}\rangle, keep their noninteracting values ΔEph=2h(𝐤)\Delta E_{\mathrm{ph}}=2h(\mathbf{k}), which become zero at the Weyl nodes. That is, while the single-particle/hole excitations are always gapful, this interacting WSM retains its gapless Weyl nodes from the viewpoint of particle-hole excitations.

II.2 One-dimensional model for the study of surface Fermi arcs

In order to construct the Fermi arc surface states, we need to consider a finite-slab geometry with open boundaries. To this end, we first transform Eq. (1) into its real-space form and then impose the open boundary conditions. Here we take a finite size of LL sites in the xx direction. By employing Fourier transform in the xx direction, c~j𝐤α=(1/L)kxeikxjc𝐤α\tilde{c}_{j\mathbf{k}_{\bot}\alpha}=(1/\sqrt{L})\sum_{k_{x}}e^{ik_{x}j}c_{\mathbf{k}\alpha}\,, the Hamiltonian for a given transverse momentum 𝐤(ky,kz)\mathbf{k}_{\bot}\equiv(k_{y},\,k_{z}) reduces to a 1D tight-binding form, where j=1,,Lj=1,\dots,L labels the sites in the xx direction. After imposing the the open boundary conditions by dropping off the boundary hopping terms, the resulting 1D model with two open ends becomes

(𝐤)=0(𝐤)+U(𝐤)\mathcal{H}(\mathbf{k}_{\bot})=\mathcal{H}_{0}(\mathbf{k}_{\bot})+\mathcal{H}_{U}(\mathbf{k}_{\bot}) (4)

with

0(𝐤)=\displaystyle\mathcal{H}_{0}(\mathbf{k}_{\bot})= i=1L1(c~ic~i+1+h.c.)+t𝐤i=1L(c~ic~i+h.c.)\displaystyle-\sum_{i=1}^{L-1}\left(\tilde{c}_{i\uparrow}^{\dagger}\tilde{c}_{i+1\downarrow}+\mathrm{h.c.}\right)+t_{\mathbf{k}_{\bot}}\sum_{i=1}^{L}\left(\tilde{c}_{i\uparrow}^{\dagger}\tilde{c}_{i\downarrow}+\mathrm{h.c.}\right)
+h𝐤i=1L(c~ic~ic~ic~i)\displaystyle+h_{\mathbf{k}_{\bot}}\sum_{i=1}^{L}\left(\tilde{c}_{i\uparrow}^{\dagger}\tilde{c}_{i\uparrow}-\tilde{c}_{i\downarrow}^{\dagger}\tilde{c}_{i\downarrow}\right) (5)

and

U(𝐤)=\displaystyle\mathcal{H}_{U}(\mathbf{k}_{\bot})= ULd=1L1i=1Ldj=d+1L[c~i+dc~ic~jdc~j+()]\displaystyle\frac{U}{L}\sum_{d=1}^{L-1}\sum_{i=1}^{L-d}\sum_{j=d+1}^{L}\left[\tilde{c}_{i+d\uparrow}^{\dagger}\tilde{c}_{i\uparrow}\tilde{c}_{j-d\downarrow}^{\dagger}\tilde{c}_{j\downarrow}+(\uparrow\leftrightarrow\downarrow)\right]
+ULi,j=1Ln~in~jU2i=1L(n~i+n~i1).\displaystyle+\frac{U}{L}\sum_{i,j=1}^{L}\tilde{n}_{i\uparrow}\tilde{n}_{j\downarrow}-\frac{U}{2}\sum_{i=1}^{L}(\tilde{n}_{i\uparrow}+\tilde{n}_{i\downarrow}-1)\;. (6)

Here the labels of the transverse momentum 𝐤\mathbf{k}_{\bot} in the fermion operators is suppressed for simplification and n~iα=c~iαc~iα\tilde{n}_{i\alpha}=\tilde{c}_{i\alpha}^{\dagger}\tilde{c}_{i\alpha}. The two parameters in Eq. (II.2) are given by t𝐤=Mcoskycoskzt_{\mathbf{k}_{\bot}}=M-\cos k_{y}-\cos k_{z} and h𝐤=sinkyh_{\mathbf{k}_{\bot}}=\sin k_{y}. We note that the non-interacting part 0(𝐤)\mathcal{H}_{0}(\mathbf{k}_{\bot}) is equivalent to the Su-Schrieffer-Heeger model Su et al. (1979) with an alternating on-site potential. The first term of U(𝐤)\mathcal{H}_{U}(\mathbf{k}_{\bot}) describes the correlated hopping for up and down (pseudo-)spins with the same distances dd but in opposite directions. The second term in Eq. (II.2) comes from the zero-distance d=0d=0 hopping, which reduces to the density-density interactions between different spin components.

While the model in Eq. (1) under periodic boundary conditions can be solved analytically, the corresponding one under open boundary conditions cannot. In this paper, numerically exact diagonalization is employed to search possible edge states at zero energy for the corresponding 1D models in Eq. (4). The collection of such zero-energy edge states constitutes the surface Fermi arcs of our interacting WSM.

III results

In the present calculations, we set M=2M=2 such that kWeyl=π/2k_{\mathrm{Weyl}}=\pi/2. The system sizes under consideration are up to L=16L=16. In the following, we focus on the U=2U=2 case. The results for other UU’s are qualitatively the same.

Due to the conservation of total particle number, subspaces of the Hilbert space with different particle numbers are decoupled. One can thus diagonalize (𝐤)\mathcal{H}(\mathbf{k}_{\bot}) for each subspace with a fixed number NN of electrons. The largest subspace takes place at half-filling N=LN=L and its dimension is about 6×1086\times 10^{8} for L=16L=16.

III.1 Single-particle Fermi arcs

To show the existence of the single-particle edge states lying between two bulk bands, the quasi-particle and the quasi-hole energies, defined by ϵ+0(N=L+1)0(N=L)\epsilon_{+}\equiv\mathcal{E}_{0}(N=L+1)-\mathcal{E}_{0}(N=L) and ϵ0(N=L)0(N=L1)\epsilon_{-}\equiv\mathcal{E}_{0}(N=L)-\mathcal{E}_{0}(N=L-1), are calculated. Here 0(N)\mathcal{E}_{0}(N) denotes the ground energy of the 1D model in Eq. (4) with NN electrons. The appearance of the in-gap edge states can be recognized when ϵ+\epsilon_{+} and ϵ\epsilon_{-} deviate from their bulk values,

ϵ±bulk\displaystyle\epsilon^{\mathrm{bulk}}_{\pm} =±minkx{ΔE1p/1h}\displaystyle=\pm\min_{k_{x}}\{\Delta E_{1p/1h}\}
=±[sin2ky+(1coskycoskz)2+U/2],\displaystyle=\pm\left[\sqrt{\sin^{2}{k_{y}}+(1-\cos{k_{y}}-\cos{k_{z}})^{2}}+U/2\right]\;,

for the present case of M=2M=2.

Before presenting our numerical results, some symmetry properties of our 1D lattice model (𝐤)\mathcal{H}(\mathbf{k}_{\bot}) in Eq. (4) are discussed in order. First, (𝐤)\mathcal{H}(\mathbf{k}_{\bot}) is invariant under the parity transformation 𝒫\mathcal{P}: c~iαc~Li+1,α\tilde{c}_{i\alpha}\rightarrow\tilde{c}_{L-i+1,-\alpha} in combination with kykyk_{y}\rightarrow-k_{y}. The energy spectrum will thus be symmetric with respect to the ky=0k_{y}=0 axis. Second, (𝐤)\mathcal{H}(\mathbf{k}_{\bot}) is invariant under the operation combining the parity transformation 𝒫\mathcal{P} with the particle-hole transformation: c~ic~i\tilde{c}^{\dagger}_{i\uparrow}\leftrightarrow\tilde{c}_{i\uparrow} and c~ic~i\tilde{c}^{\dagger}_{i\downarrow}\leftrightarrow-\tilde{c}_{i\downarrow}. Therefore, the quasi-particle and the quasi-hole excitation energies, ϵ+\epsilon_{+} and |ϵ||\epsilon_{-}|, above the ground-state energy will be the same. That is, the energy spectrum will become symmetric with respect to ϵ=0\epsilon=0.

Refer to caption
Figure 1: Quasi-particle and quasi-hole energies, ϵ+\epsilon_{+} and ϵ\epsilon_{-}, at two different UU’s as functions of kyk_{y} for kz=π/4k_{z}=\pi/4(<kWeyl<k_{\mathrm{Weyl}}) and 3π/43\pi/4(>kWeyl>k_{\mathrm{Weyl}}). Here the system size L=16L=16 and kWeyl=π/2k_{\mathrm{Weyl}}=\pi/2 for M=2M=2. The upper panels show the U=0U=0 cases, while the lower ones give those for U=2U=2. Black open triangles represent ϵ+\epsilon_{+} and black open circles show ϵ\epsilon_{-}. The green lines denote the corresponding bulk values ϵ±bulk\epsilon^{\mathrm{bulk}}_{\pm}. The shadow regions indicate the quasi-particle/hole bands bounded by ϵ±bulk\epsilon^{\mathrm{bulk}}_{\pm}. For U=0U=0 case with kz=π/4k_{z}=\pi/4, the red (blue) line shows the dispersion relation of the noninteracting edge states, ϵ=sinky\epsilon=\sin{k_{y}} (ϵ=sinky\epsilon=-\sin{k_{y}}), for |ky|0.4π|k_{y}|\leq 0.4\pi.

In Fig. 1, ϵ+\epsilon_{+} and ϵ\epsilon_{-} as functions of kyk_{y} for kz=π/4k_{z}=\pi/4 and 3π/43\pi/4 are displayed for the case of U=2U=2. For comparison, the results for the noninteracting (U=0U=0) case are shown as well. It is clear that our findings do respect the symmetries mentioned above. The U=0U=0 results can be derived analytically. When |kz|<kWeyl|k_{z}|<k_{\mathrm{Weyl}}, the effective 2D system at a fixed kzk_{z} acts as a Chern insulator with a Chern number C=1C=1 and thus supports a pair of chiral edge modes. The edge modes emerge around ky=0k_{y}=0 with the dispersions ϵ=±sinky\epsilon=\pm\sin{k_{y}}. On the other hand, when kWeyl<|kz|<πk_{\mathrm{Weyl}}<|k_{z}|<\pi, the effective 2D system behaves as a C=0C=0 trivial insulator, in which there exists no edge state and the quasi-particle/quasi-hole energies must equal to their bulk values. As seen from Fig. 1, our numerical data for the U=0U=0 case fairly reproduce the analytic predictions.

After turning on the interaction UU, the conclusions are qualitatively unchanged. The calculated quasi-particle/quasi-hole energies for kz=3π/4>kWeylk_{z}=3\pi/4>k_{\mathrm{Weyl}} agree well with their bulk values ϵ±bulk\epsilon^{\mathrm{bulk}}_{\pm} and thus no edge state appears. Interestingly, when kz=π/4<kWeylk_{z}=\pi/4<k_{\mathrm{Weyl}}, the edge modes around ky=0k_{y}=0 are not destroyed by interactions. Nevertheless, the correlation effect makes the slopes of their dispersions around ky=0k_{y}=0 being steeper. This is consistent with the enhanced bulk gap due to electronic correlations. We note that the calculated single-particle gap Δ1\Delta_{1} defined by the minimal value of the energy difference ϵ+ϵ\epsilon_{+}-\epsilon_{-} can have a tiny nonzero value due to finite-size effects. As shown below (Fig. 3), this value approaches zero in the thermodynamic limit, and then the zero-energy edge modes indeed show up in the spectrum. Therefore, our findings clearly demonstrate the robustness of zero-energy single-particle edge modes against the long-ranged interaction UU.

Refer to caption
Figure 2: Density profiles δni±\delta n_{i}^{\pm} at two different UU’s of an extra particle or hole for the states with ky=0.05πk_{y}=0.05\pi for both the cases of kz=π/4k_{z}=\pi/4 and 3π/43\pi/4. Here L=16L=16. The upper panels show the U=0U=0 cases, while the lower ones gives those for U=2U=2.

To further identify the in-gap states around ky=0k_{y}=0 for kz=π/4k_{z}=\pi/4 as edge modes, we need to show that the densities of these quasi-particle/quasi-hole states do distribute near the edges only. The density profile of one electron added is defined as δni+=n~iN=L+1n~iN=L\delta n_{i}^{+}=\langle\tilde{n}_{i}\rangle_{N=L+1}-\langle\tilde{n}_{i}\rangle_{N=L} and that of an extra hole is δni=n~iN=Ln~iN=L1\delta n_{i}^{-}=\langle\tilde{n}_{i}\rangle_{N=L}-\langle\tilde{n}_{i}\rangle_{N=L-1}. Here n~i=αc~iαc~iα\tilde{n}_{i}=\sum_{\alpha}\tilde{c}_{i\alpha}^{\dagger}\tilde{c}_{i\alpha} is the total electron number operator on site ii and n~iN\langle\tilde{n}_{i}\rangle_{N} denotes its expectation value with respect to the ground state with NN electrons. The values of δni+\delta n_{i}^{+} and δni\delta n_{i}^{-} describe how an added particle or hole distributes itself in the system.

Our findings of δni+\delta n_{i}^{+} and δni\delta n_{i}^{-} at U=2U=2 for the states with a typical value of kyk_{y} around ky=0k_{y}=0 are shown in Fig. 2 for both the cases of kz=π/4k_{z}=\pi/4 and 3π/43\pi/4. Again, the data for the U=0U=0 case are displayed as well for comparison. When kz=3π/4>kWeylk_{z}=3\pi/4>k_{\mathrm{Weyl}}, no matter whether the interaction UU is present or not, the extra particle/hole always distributes itself in the center of the lattice system. It shows that these quasi-particle/quasi-hole states indeed describe the bulk states, in support of the energy perspective discussed above. For kz=π4<kWeylk_{z}=\frac{\pi}{4}<k_{\mathrm{Weyl}}, we find that the added particle/hole locates on one or the other edge even in the presence of the long-ranged interaction UU. Moreover, the density profiles for the U=2U=2 case modify little as compared to the noninteracting case. Therefore, the in-gap states for kz=π/4k_{z}=\pi/4 do behave as edge modes and they are stable against electronic correlations.

Refer to caption
Figure 3: Size dependence of the single-particle gap Δ1=ϵ+ϵ\Delta_{1}=\epsilon_{+}-\epsilon_{-} as functions of kzk_{z} for (a) U=0U=0 and (c) U=2U=2. The extrapolation of Δ1\Delta_{1} to the thermodynamic limit for kz=0.6πk_{z}=0.6\pi, 0.55π0.55\pi, 0.5π0.5\pi, 0.45π0.45\pi, and 0.4π0.4\pi from top to bottom are shown in (b) and (d), respectively. Here we take M=2M=2 (thus kWeyl=π/2k_{\mathrm{Weyl}}=\pi/2) and ky=0k_{y}=0. We fit the data in (b) and (d) with a quadratic function f(L)=a0+a1(1/L)+a2(1/L)2f(L)=a_{0}+a_{1}(1/L)+a_{2}(1/L)^{2}, except those of |kz|<π/2|k_{z}|<\pi/2 in (b), in which the exponential function f(L)=aexp(bL)f(L)=a\;\mathrm{exp}(-bL) is employed.

The Fermi arc is constituted by the collection or locus of the zero-energy edge states. In the noninteracting limit, this line segment ends at the projection of the bulk Weyl nodes onto the surface Brillouin zone. It has been shown in Fig. 1 that, for a specific kz=π/4k_{z}=\pi/4, the zero-energy edge modes, indicated by vanishing single-particle gap Δ1\Delta_{1}, persist upon the interaction effect. We now extend such discussions to other kzk_{z}’s and determine the whole Fermi arc for the present interacting model.

In Fig. 3, the single-particle gaps Δ1\Delta_{1} for all kzk_{z}’s within the Brillouin zone are displayed for the case of U=2U=2. For comparison, the results for the noninteracting (U=0U=0) case are shown as well. As illustrated in Fig. 1, the possible zero-energy modes always occur at ky=0k_{y}=0. We thus fix ky=0k_{y}=0 in our calculations of Δ1\Delta_{1}. Besides, M=2M=2 is assumed such that kWeyl=π/2k_{\mathrm{Weyl}}=\pi/2. The U=0U=0 results can be derived analytically. When |kz|<π/2|k_{z}|<\pi/2, Δ1=0\Delta_{1}=0 because of the appearance of a pair of degenerate chiral edge mode at ky=0k_{y}=0 [cf. Fig. 1(a)]. On the other hand, when π/2<|kz|π\pi/2<|k_{z}|\leq\pi, there exists no edge mode and the quasi-particle/quasi-hole energies agree well with their bulk values ϵ±bulk\epsilon^{\mathrm{bulk}}_{\pm} such that Δ1=2[ϵ+bulk]ky=0\Delta_{1}=2\left[\epsilon^{\mathrm{bulk}}_{+}\right]_{k_{y}=0}, which reduces to Δ1=2|coskz|\Delta_{1}=2|\cos{k_{z}}| for the present case of M=2M=2. That is, the Fermi arc at U=0U=0 extends from kz=π/2k_{z}=-\pi/2 to π/2\pi/2. As seen from Fig. 3, our data agree with the analytic predictions, while visible deviation due to finite-size effects is observed around kz=±π/2k_{z}=\pm\pi/2. Nevertheless, after extrapolating to the thermodynamic limit with LL\rightarrow\infty, the agreement becomes excellent.

When the long-ranged interaction UU is turned on, for π/2<|kz|π\pi/2<|k_{z}|\leq\pi, the calculated Δ1\Delta_{1} again agrees well with the analytic prediction Δ1=2[ϵ+bulk]ky=0=2|coskz|+U\Delta_{1}=2\left[\epsilon^{\mathrm{bulk}}_{+}\right]_{k_{y}=0}=2|\cos{k_{z}}|+U. However, for |kz|π/2|k_{z}|\leq\pi/2, enhanced finite-size effects are observed in our calculations not (b). While Δ1\Delta_{1} is found to be always nonzero up to our maximum size L=16L=16, its extrapolated values for |kz|0.4π|k_{z}|\leq 0.4\pi does reduce to zero. We note that the edge states are expected to be more and more delocalized when kzk_{z} approaches the Weyl nodes at kz=±π/2k_{z}=\pm\pi/2. Therefore, severe finite-size effects exist around the Weyl nodes and calculations for much larger sizes are necessary to reduce such effects. Nevertheless, our findings strongly suggest that the Fermi arcs for nonzero UU have the same extent as those in the noninteracting case. We thus conclude that, except the renormalized dispersions, the single-particle Fermi arcs receive no modification by the long-ranged interaction.

It appears that our conclusion disagrees with that obtained in Ref. Meng and Budich (2019), where a gap opening in the single-particle Fermi-arc states is observed for a slightly different model. Actually, after a necessary correction to their model Hamiltonian, gapless single-particle Fermi arcs will be recovered. That is, the same conclusion can be reached even when other long-ranged interactions different from the present model are taken. Instead of Eq. (II.2), the authors in Ref. Meng and Budich (2019) consider the following interaction term to simplify their calculations,

~U(𝐤)=U(kz)2[i=1L(n~i+n~i1)]2.\widetilde{\mathcal{H}}_{U}(\mathbf{k}_{\bot})=\frac{U(k_{z})}{2}\left[\sum_{i=1}^{L}(\tilde{n}_{i\uparrow}+\tilde{n}_{i\downarrow}-1)\right]^{2}\;. (7)

Because this long-ranged density-density interaction depends only on the total particle number N~=i=1L(n~i+n~i)\tilde{N}=\sum_{i=1}^{L}(\tilde{n}_{i\uparrow}+\tilde{n}_{i\downarrow}), it is clear that both the quasi-particle and the quasi-hole energies, ϵ+\epsilon_{+} and ϵ\epsilon_{-}, will be pushed away from the Fermi level by U(kz)/2U(k_{z})/2 for those kzk_{z}’s with nonzero U(kz)U(k_{z})’s. This gives a gap opening of U(kz)U(k_{z}) for the noninteracting Fermi-arc states. However, the form in Eq. (7) is improper physically. For many-body states with a fixed particle density N~/L=1±δρ\tilde{N}/L=1\pm\delta\rho, where δρ\delta\rho denotes the density deviation from half-filling, this interaction term gives a contribution in energy density, U(kz)(δρ)2L/2U(k_{z})(\delta\rho)^{2}L/2, which becomes infinite in the thermodynamic limit, LL\rightarrow\infty. In order to keep the energy density being an intensive quantity, a factor 1/L1/L needs to be added in Eq. (7), i.e., ~U~U/L\widetilde{\mathcal{H}}_{U}\rightarrow\widetilde{\mathcal{H}}_{U}/L. The single-particle gap thus becomes U(kz)/LU(k_{z})/L, which vanishes in the thermodynamic limit. Therefore, the corrected results are actually in agreement with ours.

III.2 Collective Fermi arcs of particle-hole nature

While the Weyl nodes are gapped out by the long-ranged interactions, the Hall conductance (and thus the Chern number) for each 2D momentum sector between the nodes remains its quantized value Morimoto and Nagaosa (2016). According to the theory of quantum Hall edge states Wen (1991); Renn (1995); Yoshioka , there must exist chiral edge modes on the boundaries of these 2D sectors. Notice that, instead of quasiparticle excitations, those edge modes are the capillary waves, whose dynamics is described by a 1D chiral boson theory. That is, the Fermi arc states implied by nonzero Chern number are constituted by the collective edge excitations, rather than the single-particle ones. Since the topological Chern number of the bulk system is not affected by the long-ranged interactions, the collective Fermi arcs of particle-hole nature are expected to maintain their noninteracting structures.

To verify this conclusion, we evaluate the particle-hole excitation energies Δph=10\Delta\mathcal{E}_{\mathrm{ph}}=\mathcal{E}_{1}-\mathcal{E}_{0} for the cases of nonzero UU’s. Here 0\mathcal{E}_{0} (1\mathcal{E}_{1}) denotes the energy of the ground state (first-excited state) for the 1D model in Eq. (4) at half-filling. Because the particle number is fixed, the excited states should represent the particle-hole excitations. The appearance of the edge modes can be recognized once Δph\Delta\mathcal{E}_{\mathrm{ph}} deviates the corresponding bulk value. From the discussions at the end of Sec. II.1, the lowest excitation energy of the bulk particle-hole excitation at zero momentum transfer for a given set of kyk_{y} and kzk_{z} can be obtained by

ΔEphbulk\displaystyle\Delta E_{\mathrm{ph}}^{\mathrm{bulk}} =minkx{2h(𝐤)}\displaystyle=\min_{k_{x}}\{2h(\mathbf{k})\}
=2sin2ky+(1coskycoskz)2\displaystyle=2\sqrt{\sin^{2}{k_{y}}+(1-\cos{k_{y}}-\cos{k_{z}})^{2}}

for the present case of M=2M=2. This expression is valid no matter whether UU is zero or not. We note that, different from the cases of ϵ±bulk\epsilon^{\mathrm{bulk}}_{\pm}, the bulk value of the particle-hole excitation energy ΔEphbulk\Delta E_{\mathrm{ph}}^{\mathrm{bulk}} can approach zero at the Weyl nodes with ky=0k_{y}=0 and kz=±π/2k_{z}=\pm\pi/2, even when UU is nonzero.

Refer to caption
Figure 4: Size dependence of particle-hole excitation energies Δph\Delta\mathcal{E}_{\mathrm{ph}} at U=2U=2 as functions of kyk_{y} for (a) kz=π/4k_{z}=\pi/4(<kWeyl<k_{\mathrm{Weyl}}) and (b) 3π/43\pi/4(>kWeyl>k_{\mathrm{Weyl}}). Here we take M=2M=2 such that kWeyl=π/2k_{\mathrm{Weyl}}=\pi/2. The green lines denote the corresponding bulk values ΔEphbulk\Delta E_{\mathrm{ph}}^{\mathrm{bulk}}. Panels (c) and (d) show the density profiles δni\delta n_{i} of the particle-hole excitations at U=2U=2 with ky=0.05πk_{y}=0.05\pi for the cases of kz=π/4k_{z}=\pi/4 and 3π/43\pi/4, respectively. Here the data for L=16L=16 are presented. Panel (e) displays the size dependence of the particle-hole gap Δph\Delta_{\mathrm{ph}} as functions of kzk_{z} for U=2U=2. The extrapolation of Δph\Delta_{\mathrm{ph}} to the thermodynamic limit for kz=0.6πk_{z}=0.6\pi, 0.55π0.55\pi, 0.5π0.5\pi, 0.45π0.45\pi, and 0.4π0.4\pi from top to bottom is shown in (f). The data in (f) are fitted with a quadratic function f(L)=a0+a1(1/L)+a2(1/L)2f(L)=a_{0}+a_{1}(1/L)+a_{2}(1/L)^{2}, except those of |kz|<π/2|k_{z}|<\pi/2, in which the exponential function f(L)=aexp(bL)f(L)=a\;\mathrm{exp}(-bL) is employed.

Our results of Δph\Delta\mathcal{E}_{\mathrm{ph}} as functions of kyk_{y} for kz=π/4k_{z}=\pi/4 and 3π/43\pi/4 are shown in Figs. 4 (a) and (b) for the U=2U=2 case. Because of the symmetry in the energy spectrum mentioned in the last subsection, only the data for 0ky<π0\leq k_{y}<\pi are plotted. As illustrated in Fig. 4 (a), we find that the edge modes do appear around ky=0k_{y}=0 for kz<kWeylk_{z}<k_{\mathrm{Weyl}} even in the presence of long-ranged interactions. Notably, the calculated energies of bulk excitations for either kz<kWeylk_{z}<k_{\mathrm{Weyl}} or kz>kWeylk_{z}>k_{\mathrm{Weyl}} experience considerable finite-size effects. Nevertheless, their values agree well with the analytic predictions after extrapolating to the thermodynamic limit.

To show the in-gap excited states around ky=0k_{y}=0 for kz=π/4k_{z}=\pi/4 being the edge modes, their density profiles δni=n~i1n~i0\delta n_{i}=\langle\tilde{n}_{i}\rangle_{1}-\langle\tilde{n}_{i}\rangle_{0} are calculated as well. Here 0\langle\cdots\rangle_{0} and 1\langle\cdots\rangle_{1} denote the expectation values with respect to the ground state and the first-excited state at half-filling, respectively. Our result at U=2U=2 for the state with kz=π/4k_{z}=\pi/4 and a typical value of kyk_{y} around ky=0k_{y}=0 is displayed in Fig. 4 (c). For comparison, the counterpart for the case of kz=3π/4k_{z}=3\pi/4 is plotted in Fig. 4 (d). As illustrated in the figure, the in-gap states around ky=0k_{y}=0 for kz=π/4k_{z}=\pi/4 do behave as edge states, in agreement with the energy perspective. Besides, the disturbed density shows a nonlocal charge transfer from one end to the other. That is, this particle-hole excitation implies the quantized charge pumping via the twisted boundary condition in the yy direction with a twist angle ϕ=2π\phi=2\pi. Such a charge pumping process is expected, since the 2D sector at kz=π/4k_{z}=\pi/4 behaves as a quantum Hall system with quantized Hall conductance.

Finally, we determine the extent of the collective Fermi arcs constituted by the collection of the zero-energy edge states of particle-hole excitations. Similar to the discussions in Fig. 3, this can be achieved by evaluating the particle-hole gap Δph\Delta_{\mathrm{ph}} defined by the minimal value of Δph\Delta\mathcal{E}_{\mathrm{ph}}. Therefore, the locus of the states with Δph=0\Delta_{\mathrm{ph}}=0 in the surface Brillouin zone gives the collective Fermi arc. As observed in Figs. 4 (a) and (b), the states with the smallest Δph\Delta\mathcal{E}_{\mathrm{ph}} always occur at ky=0k_{y}=0. We thus fix ky=0k_{y}=0 in our calculations of Δph\Delta_{\mathrm{ph}} not (c). Here kWeyl=π/2k_{\mathrm{Weyl}}=\pi/2 because M=2M=2 is assumed. Our findings of Δph\Delta_{\mathrm{ph}} as functions of kzk_{z} for the case of U=2U=2 are displayed in Fig. 4 (e). Because Δph=Δ1\Delta_{\mathrm{ph}}=\Delta_{1} for U=0U=0, the results of Δph\Delta_{\mathrm{ph}} for the noninteracting case can directly refer to Fig. 3 (a). We find that the zero-energy edge states of particle-hole excitations appear when kzk_{z} lies between two Weyl nodes even in the presence of long-ranged interactions. As compared to the single-particle counterparts in Fig. 3 (c), the particle-hole gaps within |kz|<π/2|k_{z}|<\pi/2 are found to receive less finite-size effects not (b). As mentioned before, the occurrence of such collective zero-energy edge states is actually guaranteed by the nonzero Hall conductances. Our findings provide a numerical support on this bulk-boundary correspondence in terms of collective excitations.

IV conclusions and discussions

In summary, by means of numerical exact diagonalizations, we provide clear evidences on the existence of the Fermi-arc surface states even when the bulk Weyl nodes are gapped out by the long-ranged interactions. That is, the relationship between the existence of the gapless Weyl nodes and the surface Fermi arcs needs not hold in the correlated systems, while it is always true in the noninteracting cases. The possibility of the appearance of Fermi arcs in the absence of gapless nodes exhibit the subtlety in identifying the topological phases of interacting WSMs. For example, the conclusions drawn merely from the nodal structure of the single-particle spectrum should be considered with caution.

We note that the Fermi-arc surface states can be defined both in the ΔN=±1\Delta N=\pm 1 and the ΔN=0\Delta N=0 subspaces of the Hilbert space, where ΔN\Delta N denotes the change in total particle number. The former is of quasiparticle nature, while the latter describes the collective particle-hole excitation. For the model under consideration, both kinds of the Fermi-arc surface states are found to preserve their noninteracting structures even in the presence of the long-ranged interactions.

Due to their topological origin, the robustness of the collective Fermi-arc surface states against electronic correlations is actually expected. It is shown that the bulk Hall conductance (and the Chern number) of each 2D momentum sector still keep its noninteracting value even though the long-ranged interactions produce single-particle gaps at the Weyl nodes Morimoto and Nagaosa (2016). According to the bulk-boundary correspondence given by the theory of quantum Hall edge states Wen (1991); Renn (1995); Yoshioka , the collective chiral edge states and then the constituted Fermi arcs are guaranteed to appear on the boundaries.

On the other hand, the existence of the single-particle Fermi-arc states established by our numerical calculations may be somehow surprising. The appearance of such single-particle states cannot be directly inferred from the results of Hall conductance. This statement is based on the fact that the Hall conductance describes the current-current correlator and thus contains only the information of the particle-hole subspace, rather than the quasiparticle one. While the presence of single-particle Fermi arcs was suggested by solving the effective single-particle Hamiltonian HtH_{t} Morimoto and Nagaosa (2016), such a conclusion can be misleading because the eigenstates of HtH_{t} have no direct physical meaning Gurarie (2011); Essin and Gurarie (2011). Therefore, to prove their existence, one has to diagonalize the many-body Hamiltonian within the single-particle/hole subspaces, as done in the present work.

Our results are achieved for the WSMs with long-ranged interactions. Some comments on their validity in the cases of generic interactions are made as follows. While the single-particle Fermi arcs are shown to be robust against the long-ranged interactions irrespective of the single-particle gap opening, their stability may not hold for generic interactions. Recall that the subspaces of different momenta are decoupled in the present model with momentum-local interactions. Such decoupling will not happen when the momentum-local condition is relaxed. Let’s consider a more realistic model such that the interactions in the two directions perpendicular to the zz axis, along which a pair of Weyl nodes is separated, become short-ranged. Therefore, states with different (kx,ky)(k_{x},\,k_{y}) get coupled. For such a model, the region between nodes can still be viewed as a stack of 2D interacting Chern insulators, if the Hall conductances of these 2D sectors remain quantized. After imposing the open boundary conditions, because of the couplings among states with different momenta along the boundaries, the effective model on the edges of these 2D systems will be described by the Luttinger-liquid theory Wen (1991); Renn (1995); Yoshioka . This theory shows that the quasiparticle edge modes become gapped and the gapless edge modes must consist of the particle-hole collective excitations. As a consequence, we conclude that, in the generic cases, the single-particle Fermi arcs can disappear, even though the collective ones remain topologically protected by nonzero Chern numbers. In addition, we stress that the stability of the collective Fermi arcs relies only on the topology in the particle-hole subspace and is irrelevant to the fact whether the single-particle gaps at the nodes are opened or not. That is, it is possible to have interacting WSMs hosting gapped nodes and with collective Fermi arcs on their surfaces.

Due to distinct characters in the single-particle and the collective Fermi arcs, one needs to measure different physical quantities to determine their existence. For example, the angle-resolved photoemission spectroscopy, which measures the single-particle spectrum functions, can be employed to observe only the single-particle Fermi arcs. To detect directly the collective Fermi arcs consisting of particle-hole excitations, more delicate arrangements are necessary. Recently, experimental techniques to reveal the momentum and energy dependent two-particle correlations are proposed Su and Zhang (2020). Due to the advances in experiments, it may be possible to observe collective Fermi arcs in the near future and then to verify our conclusions.

Acknowledgements.
We are grateful to Tobias Meng for useful discussions. Y.C.T. and M.F.Y. acknowledge support from the Ministry of Science and Technology of Taiwan under Grant No. MOST 108-2112-M-029-005 and 108-2811-M-005-522, respectively.

References