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

Spinon quantum spin Hall state in the kagome antiferromagnet with a Dzyaloshinskii-Moriya interaction

Li-Wei He National Laboratory of Solid State Microstructures and School of Physics, Nanjing University, Nanjing 210093, China    Jian-Xin Li [email protected] National Laboratory of Solid State Microstructures and School of Physics, Nanjing University, Nanjing 210093, China Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210093, China
Abstract

We investigate the spin-12\frac{1}{2} antiferromagnetic Heisenberg model with a Dzyaloshinskii-Moriya interaction on kagome lattice, making use of the variational Monte Carlo technique. An exotic quantum spin state is found to arise from a melting of the 𝑸=0\bm{Q}=0 long-range magnetic order by a topological transition, when a small anisotropic third nearest-neighbor antiferromagnetic Heisenberg interaction is turned on. This novel state is a gapped quantum spin liquid, characterized by a topological order with ground-state degeneracy ng=4n_{g}=4 and topological entanglement entropy γ=ln2\gamma=\ln 2, suggesting it is an Abelian topological phase. Furthermore, the Chern numbers of the spin-up (-down) spinon occupied bands of this state are C=±1C_{\uparrow\downarrow}=\pm 1, respectively. From this perspective, this state is also a time-reversal symmetric (total Chern number Ctotal=0C_{total}=0) topological insulator with spinons as the chiral edge states, which carry opposite spin and move in the opposite direction. It is analogous to the quantum spin Hall state but the spin current is carried by deconfined spinons in a quantum spin liquid, so is dubbed as the spinon quantum spin Hall state.

preprint: APS/123-QED

I introduction

One of the exotic and intriguing phase receiving extensive attention and research in condensed matter physics is the quantum spin liquid (QSL) [1], which is magnetic disordered and has the fractionalized elementary excitation called spinon with spin-12\frac{1}{2}. One class of this phase is the gapped QSL, which is a crucial representative of the topological orders with the long-range many-body quantum entanglement. Among the gapped QSLs, the chiral spin liquid (CSL) [2] breaks the time-reversal (TR) symmetry and usually has nontrivial Chern number to characterize itself. In a CSL, the spinons with up and down spins usually couple the same gauge field and have the same Chern number. For those gapped QSLs with TR symmetry, the ground-state degeneracy (GSD) [3, 4] other than the Chern number is a good quantity to describe the global topological property. Furthermore, another important quantity, the topological entanglement entropy (TEE) [5, 6] related to the quantum dimension of topological excitaitons, is very useful for both chiral and achiral topological orders.

Another exotic and extensively studied phase is the quantum spin Hall (QSH) state [7, 8], which is realized in the topological insulator [9] characterized by an insulating bulk gap and gapless edge or surface states (the bulk-edge correspondence of topological system) protected by TR symmetry. In the QSH state, the electrons with opposite spins move along the opposite direction on a given edge [7, 10, 11]. Consequently, the two states on the edge possess the spin Chern number with opposite signs so that the QSH state is characterized by a Z2 topological number.

Pictorially, both the gapped QSL with TR symmetry and the QSH state are insulators and topological nontrivial phases. However, the conventional topological insulator is a non (or weak)-interacting single particle state whose quasiparticles are electrons in the framework of Landau Fermi liquid. While, the QSL is a strong interacting Mott insulator without conventional Landau quasiparticles but the fractionalized excitations such as spinons. Hence, thus far, these two states are studied separatively from each other. Thus, an issue arises as to if there exists a possible exotic quantum state which is a gapped QSL with TR symmetry meanwhile the quasiparticles (spinons) exhibit similar QSH effect.

To search for QSLs, the kagome antiferromagnet is an appealing platform [12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26], in view of its strong geometric frustration and the resulting strong quantum fluctuations. Many novel many-body quantum states have been explored in this lattice recently  [27, 28, 29, 30, 31, 32, 33, 34, 35, 36, 37, 38]. It is generally believed that the Heisenberg model with the nearest-neighbor (NN) antiferromagnetic (AFM) interaction J1J_{1} on a kagome lattice hosts a QSL ground state. However, it is controversial that this QSL is gapped [12] or not [13, 14, 15]. Further, the long-range AFM Heisenberg spin interactions beyond the NN term are always possible and will affect the properties of ground states. It is reported that the second NN AFM interaction J2J_{2} favors a 𝑸=0\bm{Q}=0 magnetic order [16, 14, 17, 18]. However, the interplay between the J2J_{2} term and one of the third NN AFM couplings which is across the diagonals of hexagons JdJ_{d} can induce a CSL [19, 20, 17] with spontaneous TR symmetry breaking. And even, the CSL can arise in the XXZ model with anisotropic J2J_{2} and JdJ_{d} terms [21, 22]. These results point to the dominant role that a relatively large JdJ_{d} might play to stabilize the CSL though it alone does not. Moreover, the scalar three-spin interaction χ=𝑺i𝑺j×𝑺k\chi=\bm{S}_{i}\cdot\bm{S}_{j}\times\bm{S}_{k} (subscripts mean vertexes of each triangle), which breaks TR symmetry, can naturally stabilize the CSL [20, 23]. Therefore, it suggests that the novel gapped QSL with TR symmetry we look for does not exist in these models.

Another choice is to consider the Dzyaloshinskii-Moriya interaction (DM) interaction, whose effects in a kagome lattice have recently been investigated theoretically and experimentally [39, 40, 18, 41, 42, 43, 44, 45, 46]. However, it has been shown that this small interaction in fact favors the in-plane 𝑸=0\bm{Q}=0 magnetic order, when the direction of DM vector is perpendicular to the xyxy plane and leads to the local vector chirality 𝝌v=𝑺1×𝑺2+𝑺2×𝑺3+𝑺3×𝑺1\bm{\chi}_{\mathrm{v}}=\bm{S}_{1}\times\bm{S}_{2}+\bm{S}_{2}\times\bm{S}_{3}+\bm{S}_{3}\times\bm{S}_{1}. Hence, if proceeding along this direction, we need to consider additional exchange interactions to melt this magnetic order. We notice that in this magnetic order spins aline ferromagnetically along the diagonal direction of the hexagon in the kagome lattice. So, it arises a possibility to consider the interplay between the DM interaction and the third NN AFM interaction along the diagonal direction.

Based on these considerations, in this work, we study the nature of the quantum spin states in the J1J_{1}-J2J_{2}-J3J_{3} kagome antiferromagnet with additional DM interaction and the third NN AFM interaction JdJ_{d} along the diagonal direction. We do not intend to obtain the global phase diagram with so many spin interaction parameters, instead mainly focus on the study of quantum spin states emerging out of the 𝑸=0\bm{Q}=0 magnetic ordered state. When only J1J_{1} term exists, our numerical simulation suggests a Dirac spin liquid (DSL) as the ground state, and the introduction of a weak DM interaction transits the system into the long-range 𝑸=0\bm{Q}=0 magnetic order. These results are consistent with previous researches [13, 14, 15, 18, 41]. Therefore, without loss of generality, we will fix the magnitude of DM interaction DD as 0D0.20\lesssim D\lesssim 0.2 in our main calculations to investigate the effects of other longer-range AFM Heisenberg interactions. When a weak diagonal third NN interaction JdJ_{d} turns on, the magnitude MM of the magnetic moment for the 𝑸=0\bm{Q}=0 state decreases and eventually drops to zero when Jd0.21J_{d}\simeq 0.21 (setting J1=1J_{1}=1). Our calculation suggests there is a continuous phase transition into a disordered phase, a gapped QSL with TR symmetry. We also show that when a small additional J2J_{2} is present, this transition still occurs, for example a slightly larger Jd0.27J_{d}\simeq 0.27 is required with J2=0.05J_{2}=0.05 because the J2J_{2} term favors the 𝑸=0\bm{Q}=0 order, as mentioned above. Obviously, this state can survive in a broad range of JdJ_{d}. It is verified that the CSL and the so-called cuboc1cuboc1 magnetic state breaking TR symmetry are not found in the range of JdJ_{d} we considered, while the cuboc1cuboc1 order is indeed found in the large JdJ_{d} range, such as D=0.1,J2,3=0,Jd=0.3D=0.1,J_{2,3}=0,J_{d}=0.3. Beyond the diagonal third NN term JdJ_{d}, we have also checked the effects of the usual NN term J3J_{3} and find that it further enhances the effect of suppressing the 𝑸=0\bm{Q}=0 order. Therefore, this phase transition is general in the sense of the reasonable DM interaction and diagonal third NN term, and the resulting gapped QSL with TR symmetry is robust.

To describe the intrinsic property of the gapped QSL, we calculate the TEE and find γ0.748\gamma\simeq 0.748 on a torus, which agrees numerically with the exact value γideal=ln20.693\gamma_{\mathrm{ideal}}=\ln 2\simeq 0.693. So, the quantum dimension Dq=2D_{q}=2 is obtained via γ=lnDq\gamma=\ln D_{q}. In the meantime, we find that its GSD is ng=4n_{g}=4. These results suggest that the gapped QSL holds an Abelian topological order with ng=Dq2n_{g}=D_{q}^{2}. It is expected that the total Chern number of this QSL is zero because of TR symmetry. Interestingly, we find that the spin-up and -down spinons see opposite gauge fluxes and there is no coupling between them. As a result, we can independently define the Chern numbers for the spin-up and -down spinons, respectively. It turns out that the Chern number of the spin-up (-down) spinons is C()=+()1C_{\uparrow(\downarrow)}=+(-)1, which is right the Z2 index [47]. So, the spinons with different spins move along opposite directions on a given edge with the opposite chiral central charges c±=±1/2c_{\pm}=\pm 1/2. That shows that this disordered state is an exotic gapped QSL with TR symmetry and shares the same properties as those of a QSH state at the mean time. Thus, we name it the spinon quantum spin Hall (SQSH) state. According to the topological long-range entanglement, ground-state degeneracy and the Chern number of this SQSH state, we suggest that it is a double-semion topological order (doubled Chern-Simons state), which is described by a sum of two topological quantum field theories with opposite chiralities and is suggested in the string-net model [48]. Generalizing to spin-kk (kk is integer) antiferromagnets, the spin liquids with parity and time-reversal symmetry have been proposed, such as the doubled CSL termed as CSL+CSL- with k=1k=1 [49], which holds the same topological properties with the SQSH state. In addition, this state has also been studied by the wire deconstructionism of the long-range entanglement topological phase [50].

II model and method

We start with the following model,

H=\displaystyle H= (i,j)Jij𝑺i𝑺j+Di,j𝑫ij𝑺i×𝑺j.\displaystyle\ \sum_{\left(i,j\right)}J_{ij}\bm{S}_{i}\cdot\bm{S}_{j}+D\sum_{\langle i,j\rangle}\bm{D}_{ij}\cdot\bm{S}_{i}\times\bm{S}_{j}. (1)

The first term in model (1) is the AFM Heisenberg spin interaction, including the first, second and two anisotropic third NN spin couplings, expressed respectively as the J1J_{1}, J2J_{2}, J3J_{3} and JdJ_{d} terms. The second term denotes the DM interaction connecting the first NN bond spins, and the direction of the DM vector 𝑫ij\bm{D}_{ij} is perpendicular to the NN bond i,j\langle i,j\rangle with DD its magnitude. All these terms are illustrated in Fig. 1. In this paper, we only consider the case that the vector 𝑫ij\bm{D}_{ij} is perpendicular to the lattice plane, namely only DijzD^{z}_{ij} is finite, so that total SzS_{z} is conserved. We set J1=1J_{1}=1 as the energy unit for convenience.

The model (1) will be investigated by the variational Monte Carlo (VMC) method. First, we use the fermionic doublet representation to rewrite the spin interactions as following,

𝑺i𝑺j=14(TijTij+PijPij)+const,\displaystyle\bm{S}_{i}\cdot\bm{S}_{j}=-\frac{1}{4}(T_{ij}T_{ij}^{\dagger}+P_{ij}P_{ij}^{\dagger})+\mathrm{const}, (2)
𝑺i×𝑺j=i4(Tij𝑻ij+Pij𝑷ijh.c.)+const,\displaystyle\bm{S}_{i}\times\bm{S}_{j}=-\frac{i}{4}(T_{ij}\bm{T}_{ij}^{\dagger}+P_{ij}\bm{P}_{ij}^{\dagger}-\mathrm{h.c.})+\mathrm{const},

where Tij=ψiψjT_{ij}=\psi^{\dagger}_{i}\psi_{j} (Pij=ψiψ¯jP_{ij}=\psi^{\dagger}_{i}\bar{\psi}_{j}) is the singlet hopping (pairing) term, while 𝑻ij=ψi𝝈ψj\bm{T}_{ij}=\psi^{\dagger}_{i}\bm{\sigma}\psi_{j} (𝑷ij=ψi𝝈ψ¯j\bm{P}_{ij}=\psi^{\dagger}_{i}\bm{\sigma}\bar{\psi}_{j}) the triplet hopping (pairing) term. The fermionic doublet field and its particle-hole partner are given by ψ=(c,c)\psi=(c_{\uparrow},c_{\downarrow}){{}^{\top}} and ψ¯=(c,c)\bar{\psi}=(c^{\dagger}_{\downarrow},-c^{\dagger}_{\uparrow}){{}^{\top}}, respectvely. Considering the SU(2) gauge structure of this fermionic representation [51], it is necessary to implement Lagrangian multipliers 𝝀\bm{\lambda} to enforce the generators of the SU(2) gauge group 𝚲i=0\bm{\Lambda}_{i}=0 to return the subspace of real physical states. Their expressions with fermionic doublet representation are as following,

Λix=14(ψiψ¯i+ψ¯iψi),\displaystyle\Lambda_{i}^{x}=-\frac{1}{4}(\psi_{i}^{\dagger}\bar{\psi}_{i}+\bar{\psi}_{i}^{\dagger}\psi_{i}), (3)
Λiy=i4(ψiψ¯iψ¯iψi),\displaystyle\Lambda_{i}^{y}=-\frac{i}{4}(\psi_{i}^{\dagger}\bar{\psi}_{i}-\bar{\psi}_{i}^{\dagger}\psi_{i}),
Λiz=12(1ψiψi).\displaystyle\Lambda_{i}^{z}=\frac{1}{2}(1-\psi_{i}^{\dagger}\psi_{i}).

Then, we use the fermionic parton approximation to decouple the spin interactions into noninteracting quadratic structure, and obtain the mean-filed Hamiltonian (irrelevant constants are omitted),

Hmfall=\displaystyle H_{\mathrm{mf}}^{all}= i,j(tijsψiψj+𝒕ijtψi𝝈ψj+Δijsψiψ¯j\displaystyle\sum_{i,j}(t_{ij}^{s}\psi_{i}^{\dagger}\psi_{j}+\bm{t}^{t}_{ij}\cdot\psi_{i}^{\dagger}\bm{\sigma}\psi_{j}+\Delta_{ij}^{s}\psi_{i}^{\dagger}\bar{\psi}_{j} (4)
+𝚫ijtψi𝝈ψ¯j+H.c.)\displaystyle+\bm{\Delta}^{t}_{ij}\cdot\psi_{i}^{\dagger}\bm{\sigma}\bar{\psi}_{j}+\mathrm{H.c.})
+i𝝀𝚲i𝑴iψi𝝈ψi/2,\displaystyle+\sum_{i}\bm{\lambda}\cdot\bm{\Lambda}_{i}-\bm{M}_{i}\cdot\psi^{\dagger}_{i}\bm{\sigma}\psi_{i}/2,

where tijst_{ij}^{s}, 𝒕ijt𝑫ij\bm{t}^{t}_{ij}\parallel\bm{D}_{ij} are spinon hopping parameters and Δijs\Delta_{ij}^{s}, 𝚫ijt𝑫ij\bm{\Delta}^{t}_{ij}\parallel\bm{D}_{ij} are pairing ones, and a background field 𝑴i\bm{M}_{i} is applied to induce a static magnetic order. Therefore, all the variational parameters are αall=(tijs,𝒕ijt,Δijs,𝚫ijt,𝝀,𝑴i)\alpha^{all}=(t_{ij}^{s},\bm{t}^{t}_{ij},\Delta_{ij}^{s},\bm{\Delta}^{t}_{ij},\bm{\lambda},\bm{M}_{i}).

Obviously, there must be various different ansatzes from the mean-field Hamiltonian Eq. (4) with a plenty of variational parameters. We selectively consider various singlet and triplet hopping terms and several pairing terms combined with the projective symmetry group[52, 53, 54] (PSG).

Actually, we have found that the spinon-pairing instability (the allowed first NN triplet pairing term with complex numbers) is vanishingly small (|𝚫ijt|/tijs<103|\bm{\Delta}_{\langle ij\rangle}^{t}|/t_{\langle ij\rangle}^{s}<10^{-3}) in our calculation. So, we will ignore all pairing terms reasonably. Moreover, the 𝝀\bm{\lambda} term can also be ignored because of vanishing pairing terms in the actual VMC procedure. In this way, we can obtain the reduced mean-field Hamiltonian, which is written as,

Hmf=\displaystyle H_{\mathrm{mf}}= i,j(tijsψiψj+H.c.)+ij(𝒕ijtψi𝝈ψj+H.c.)\displaystyle\sum_{i,j}(t_{ij}^{s}\psi_{i}^{\dagger}\psi_{j}+\mathrm{H.c.})+\sum_{\langle ij\rangle}(\bm{t}^{t}_{ij}\cdot\psi_{i}^{\dagger}\bm{\sigma}\psi_{j}+\mathrm{H.c.}) (5)
i𝑴iψi𝝈ψi/2,\displaystyle-\sum_{i}\bm{M}_{i}\cdot\psi^{\dagger}_{i}\bm{\sigma}\psi_{i}/2,

where tijst_{ij}^{s}, 𝒕ijt𝑫ij\bm{t}^{t}_{ij}\parallel\bm{D}_{ij} are the singlet and triplet spinon hopping parameters, respectively. The former includes the first and second NN hopping terms and the later only include the first NN one.

With the mean-field ground-state wave function |GS(α)mf|\mathrm{GS(\alpha)}\rangle_{\mathrm{mf}}, we can obtain a trial wave function |Φ(α)=PG|GS(α)mf|\Phi(\alpha)\rangle=P_{G}|\mathrm{GS(\alpha)}\rangle_{\mathrm{mf}}, where PGP_{G} is the Gutzwiller projection to guarantee the single occupancy condition, and α=(tijs,𝒕ijt,𝑴i\alpha=(t_{ij}^{s},\bm{t}^{t}_{ij},\bm{M}_{i}) are variational parameters. We emphasize that these parameters could be complex numbers in principle, so the whole parameter range is in fact large. In the variational process, we employ the stochastic reconfiguration scheme [55] to optimize so many parameters. We have considered various Z2 QSLs, U(1) QSLs, 𝑸=0\bm{Q}=0 and cuboc1cuboc1 magnetic orders as initial trial states selectively (see Appendix A for details). We adopt the torus geometry with L1=L2=12L_{1}=L_{2}=12 for main results, where L1,2L_{1,2} are the lengths along the two Bravais kagome-lattice vectors 𝒂1=(1,0)\bm{a}_{1}=(1,0) and 𝒂2=(1/2,3/2)\bm{a}_{2}=(-1/2,\sqrt{3}/2).

III result

In this paper, we do not intend to obtain the comprehensive phase diagram of the spin model in the kagome lattice defined by model (1). Instead, we will focus on the study of the possible states stabilized or induced by the additional DM interaction DD and the AFM interactions JdJ_{d} across the diagonals of the hexagons of the kagome lattice, in the presence of the first, second and usual third NN spin couplings.

Refer to caption

(a)

Refer to caption

(b)

Refer to caption

(c)

Refer to caption

(d)

Figure 1: (a) Illustration of the spin interactions included in model (1) on kagome lattice. The dashed lines with different colors indicate different Heisenberg terms. The arrows on the nearest-neighbor bonds indicate the direction of the DM interaction and the DM vector 𝑫ij\bm{D}_{ij} is oriented parallel (antiparallel) to zz axis in up (down) triangles indicated by \bigodot (\bigotimes). (b) denotes the ansatz for the exotic SQSH state. The +()θ+(-)\theta is the flux for spin-up (-down) spinon in each triangle, respectively, and the π(+)2θ\pi-(+)2\theta is the same way in each hexagon. The marked red and blue dashed bonds denote that the hopping terms along these bonds in HmfH_{\mathrm{mf}} have the opposite sign compared with those unmarked ones so that the unit cell is doubled. (c) is the schematic diagram indicating the in-plane configuration of classical spins in the long-range magnetic order with 𝑸=0\bm{Q}=0, three spins in all triangles form a 120120^{\circ} distribution, which preserves the original translational symmetry. (d) illustrates the energy curves versus DD when J2,d,3=0J_{2,d,3}=0 (see Appendix A for the details of the four states). We note that all of the standard errors in this paper are considered as confidence intervals.

We start with the result in the presence of only the first NN J1J_{1} AF couplings. Our VMC results suggest that a Dirac spin liquid is the most energetically favored state in this case, which is consistent with previous results[13, 14, 15]. In this U(1) QSL, all triplet terms vanish and only the singlet hopping terms survive. In the case of only the first NN hopping term is included, there are zero (π\pi) fluxes through triangles (hexagons) in the kagome lattice, corresponding to the pattern for θ=0\theta=0 in Fig. 1. We also consider the Jastrow-type wave functions for the magnetic ordered state, which is widely used in VMC studies. This state is constructed based on the solution with only a finite 𝑴i\bm{M}_{i} term in HmfH_{\mathrm{mf}} (see Appendix A). It has the magnetic order 𝑸=0\bm{Q}=0 in the lattice (XY) plane as illustrated in Fig. 1 , and is called as Jastrow + (𝑸=0)(\bm{Q}=0) state. In the case of only J1J_{1} term, the energy of this magnetic order is obviously higher than that of the Dirac spin liquid, as shown in Fig. 1.

When turning on the DM interaction DD, we find that the energy of the Dirac spin liquid does not depend on DD, while the energy of the Jastrow + (𝑸=0)(\bm{Q}=0) state decreases nearly linearly with DD (see Fig. 1). However, before the Jastrow + (𝑸=0)(\bm{Q}=0) state surpasses energetically the Dirac spin liquid, a novel competing state emerges. This state has the same forms of the first and second NN singlet hopping terms as the Dirac spin liquid, but has a finite and pure imaginary first NN triplet hopping term. The selection of this state is done via a comprehensive comparison with other possible states. We have considered the ansatzes with complex first and second NN singlet hopping terms, but find that their imaginary parts are almost zero (Im(tijs)/tijs<102\mathrm{Im}(t_{\langle ij\rangle}^{s})/t_{\langle ij\rangle}^{s}<10^{-2}, Im(tijs)/tijs<102\mathrm{Im}(t_{\langle\langle ij\rangle\rangle}^{s})/t_{\langle ij\rangle}^{s}<10^{-2}) in our variational process. We have also checked the existence of the first NN triplet pairing term (complex number), and it turns out that this term vanishes (|𝚫ijt|/tijs<103|\bm{\Delta}_{\langle ij\rangle}^{t}|/t_{\langle ij\rangle}^{s}<10^{-3}) in our calculation. In particular, we have considered two candidate states, which are regarded as the derivative states of the uniform resonating valence bond state (see Appendix A). One candidate carries the same fluxes in all triangles and the other carries the opposite fluxes in the up and down triangles. Combining the triplet hopping terms (complex number) and finite 𝑴i\bm{M}_{i}, we find that both these states are energetically unfavored.

In this novel state, the complex triplet hopping term will lead to the consequence that free spinons at the VMC mean-field level will carry nonzero flux when hops along closed loops, as the direction of the DM vector 𝑫ij\bm{D}_{ij} considered in this paper is perpendicular to the lattice plane. Specifically, the spinon with up (down) spin carries a +()θ+(-)\theta flux in all triangles and π(+)2θ\pi-(+)2\theta flux in all hexagons. This shares the same physics as a quantum spin Hall state or topological insulator. So, we dub it as the spinon quantum spin Hall (SQSH) state and will discuss its properties in detail later. When J2=Jd=J3=0J_{2}=J_{d}=J_{3}=0, this state emerges firstly at D0.01D\sim 0.01 in the sense that our numerical calculations can determine it. However, the mix of this state with the 𝑸=0\bm{Q}=0 magnetic order state, i.e. SQSH + (𝑸=0\bm{Q}=0) state, is always lower in energy in the range 0.01D0.20.01\lesssim D\lesssim 0.2. Moreover, this magnetic order state is energetically more favored than the Jastrow + (𝑸=0\bm{Q}=0) state, as shown in Fig. 1.

The SQSH + (𝑸=0\bm{Q}=0) state is induced and stabilized by the DM interaction. It is also analogous to the most possible ground state of the AFM J1J_{1} Heisenberg model on the triangular lattice [56, 57, 58]. As the DM vector 𝑫ij\bm{D}_{ij} is considered to be along the perpendicular direction of the lattice plane, it confines the spins to lie in the lattice plane and effectively plays a role of the easy-plane anisotropy. In the meantime, it is able to induce the local vector chirality 𝝌v\bm{\chi}_{\mathrm{v}}. These two properties make the DM interaction to favor the in-plane 𝑸=0\bm{Q}=0 magnetic order. In detail, this state requires a finite 𝑴i\bm{M}_{i} and additionally the same hopping term as the SQSH state. However, due to the existence of the magnetic order, the elementary excitation of this SQSH + (𝑸=0\bm{Q}=0) state is the magnon which is the confined state of two spinons. Therefore, there are no free spinons existing in the sense of the VMC mean-field level and no SQSH effects.

Refer to caption

(a)

Refer to caption

(b)

Refer to caption

(c)

Figure 2: Magnitude of magnetic moment (MM) in the SQSH + (𝑸=0\bm{Q}=0) state versus JdJ_{d} term for different J2,J3J_{2},J_{3} interactions. (a) and (b) is obtained with the DM interaction D=0.1D=0.1, (c) with D=0.2D=0.2. We note that the legends of all these curves are labeled by (J2,J3)(J_{2},J_{3}) and the lattice size we employ in (b) and (c) is 12×6×312\times 6\times 3.

It turns out that we need to melt the 𝑸=0\bm{Q}=0 magnetic order in order to realize SQSH state. We find that the AFM interaction JdJ_{d} across the diagonals of hexagons can effectively weaken the magnitude of the magnetic moment MM. As can be seen in Fig. 2, MM decreases nearly linearly with JdJ_{d} and vanishes at Jd0.21J_{d}\simeq 0.21, when only the DM interaction besides the first NN term is considered, namely D=0.1D=0.1 and J2=J3=0J_{2}=J_{3}=0 [Fig. 2)]. The introduction of the J2J_{2} term enhances effectively the moment of the 𝑸=0\bm{Q}=0 magnetic order, so that a larger JdJ_{d} is needed to eliminate the magnetic order. As shown in Fig. 2, when J2J_{2} starts from zero to 0.05, the critical JdcJ^{c}_{d} term leading to M0M\simeq 0 increases to be 0.27. On the other hand, the effect of the J3J_{3} term behaves conversely with J2J_{2}, but cooperatively with JdJ_{d} [Fig. 2]. For example, if the J2J_{2} term is fixed to be 0.10.1, we can find that the moment MM with J3=0.1J_{3}=0.1 is significantly smaller than that with J3=0.05J_{3}=0.05 for the same JdJ_{d}, and JdcJ^{c}_{d} drops to be near 0.16 for a fixed J3=J2=0.1J_{3}=J_{2}=0.1. To show more clearly the effect of the J3J_{3} term, we present the results for the moment MM in the case of D=0.1D=0.1, J2,d=0J_{2,d}=0 in Table 1. To check the effects of the DM interaction, we show results obtained with D=0.2D=0.2 in Fig. 2, one can see that JdcJ^{c}_{d} increases noticeably compared to the case of D=0.1D=0.1 for various J2J_{2} and J3J_{3}. We note that those spin exchange parameters to melt the 𝑸=0\bm{Q}=0 magnetic order can be nearly one magnitude smaller than the dominant nearest-neighbor exchange interaction J1J_{1}, which are considered to be acceptable physically and realizable experimentally.

   J3J_{3}   0.05   0.08   0.1    0.12   0.15
  MM   0.2074   0.1239   0.0728   0.0279   0.0062
Table 1: Magnetic moment MM in the SQSH + (𝑸=0\bm{Q}=0) magnetic order state calculated with lattice size 12×12×312\times 12\times 3 in the case of D=0.1D=0.1, J2,d=0J_{2,d}=0.
   state    DD    J2,3J_{2,3}    JdJ_{d}           EE
SQSH 0.1 0.0 0.29 -0.43371
0.1 0.0 0.3 -0.43345
cuboc1cuboc1 0.1 0.0 0.29 -0.43371
0.1 0.0 0.3 -0.43379
Table 2: Variational energy (per site) for the SQSH state and cuboc1cuboc1 magnetically ordered state with lattice size 12×12×312\times 12\times 3. When D=0.1D=0.1, J2,3=0.0J_{2,3}=0.0 and Jd0.3J_{d}\gtrsim 0.3, the cuboc1cuboc1 order state with TR symmetry breaking will be dominant. All the errors of the energies are 105\sim 10^{-5}.
   state    DD    J2J_{2}    JdJ_{d}    J3J_{3}       EE
SQSH 0.1 0.05 0.3 0 -0.43444
0.1 0.05 0.32 0 -0.43382
cuboc1cuboc1 0.1 0.05 0.3 0 -0.43443
0.1 0.05 0.32 0 -0.43387
Table 3: Variational energy (per site) for the SQSH state and cuboc1cuboc1 magnetically ordered state with lattice size 12×12×312\times 12\times 3. When D=0.1D=0.1, J2=0.05J_{2}=0.05, Jd0.32J_{d}\gtrsim 0.32 and J3=0.0J_{3}=0.0, the cuboc1cuboc1 ordered state with TR symmetry breaking will be dominant. All the errors of the energies are 105\sim 10^{-5}.

After the 𝑸=0\bm{Q}=0 magnetic order is melted, we further study the robustness of this disordered SQSH state against the cuboc1cuboc1 order and present the results in Table 2 and 3 . It shows that the SQSH state can survive in a relatively broad range of JdJ_{d} and eventually gives way to the cuboc1cuboc1 order, such as for Jd0.3J_{d}\gtrsim 0.3. In addition, we find that the transition from the finite MM magnetic order state to the M=0M=0 spin liquid state with SQSH is a continuous phase transition. Now, let us discuss the detail properties of the pure SQSH state. As stated above, this state is a quantum spin liquid and inherits the same singlet hopping pattern as the Dirac quantum spin liquid [See Fig. 1]. The key difference between them is that the SQSH state includes a pure imaginary zz component in the first NN triplet hoppings induced by the DM interaction. As a result, the spin-up and -down spinons see the opposite flux (±θ\pm\theta) in all triangles. Because of the absence of the coupling between spin-up and -down spinons, we can rewrite the mean-field Hamiltonian,

Hmf=(h00h),H_{\mathrm{mf}}=\begin{pmatrix}h&0\\ 0&h^{*}\end{pmatrix}, (6)

where hh denotes the Hamiltonian for the spin-up spinons while hh^{*} that for the spin-down spinons. When we go into the 𝒌\bm{k} space, at an arbitrary 𝒌\bm{k}, there is always a couple of degenerate states, which are conjugate to each other. So, the spin-up and -down spinons have opposite Berry phase in any plaqutte of 𝒌\bm{k} space, and consequently the opposite Chern number. We have calculated the Chern number of the filled three bands for the spin-up (-down) spinons as shown in Fig. 3, and find that C()=+()1C_{\uparrow(\downarrow)}=+(-)1. We then calculate the energy dispersions of the spin-up (-down) spinons in the SQSH state with the period-open boundary condition, and the results are shown in Fig. 4.

Refer to caption

(a)

Refer to caption

(b)

Figure 3: (a) Mean-field band structure of the SQSH state, where each band is doubly degenerate so that there is a significant energy gap because of the one spinon per site. (b) The chirality-chirality correlation of the SQSH state |χiχj||\langle\chi_{i}\chi_{j}\rangle| (iji\neq j) with respect to the distance along 𝒂1\bm{a}_{1} for different lattice sizes. And the |𝒓i𝒓j||\bm{r}_{i}-\bm{r}_{j}| means the distance between two up-pointing triangles. We only need to consider the distance up to 6, 9 and 12 because of the period boundary condition for a system of L1,2=12L_{1,2}=12, L1=18,L2=8L_{1}=18,L_{2}=8 and L1=24,L2=6L_{1}=24,L_{2}=6, respectively. These results are obtained in the system with D=0.1D=0.1, Jd=0.21J_{d}=0.21 and J2=J3=0J_{2}=J_{3}=0.

It shows clearly that two edge states emerge in the gap of the energy bands for each spin species spinon. In particular, the energy bands of the spin-up and -down spinons are antisymmetric with respect to momentum kk, indicating that the spin-up state is just the time-reversal copy of the spin-down one. So, the spin-up and -down spinons move along opposite direction on the edge. We have also checked the chirality-chirality correlation in the SQSH state, the results are shown in Fig. 3. It shows |χiχj|0|\langle\chi_{i}\chi_{j}\rangle|\sim 0 within numerical error, so the SQSH indeed has TR symmetry. It suggests that the two edge states of this topological SQSH are protected by TR symmetry. Thus, we believe that the SQSH state is the spinon version of the quantum spin Hall state.

Refer to caption
Figure 4: Left (right) panel is the spin-up (-down) energy dispersion of the SQSH state with the same optimal parameters as that of Fig. 3, respectively. In the calculation, the period (open) boundary condition is used in the direction of basis vectors 𝒏1=(2,0)\bm{n}_{1}=(2,0) (𝒏2=(1/2,3/2)\bm{n}_{2}=(-1/2,\sqrt{3}/2)) defined in the doubled unit cell and the reciprocal vector 𝐛1=(π,π/3)\mathbf{b}_{1}=(\pi,\pi/\sqrt{3}).

On the other hand, the SQSH state is essentially the strongly correlated state, which is different from the conventional quantum spin Hall state based on the single-particle picture. Its elementary excitations are fractionalized spinons out of the QSL ground state and could embrace more intrinsic topological properties. To further explore its property, we calculate the topological entanglement entropy and ground-state degeneracy. The TEE is obtained numerically by partitioning the system into two subsystems and calculating the second order Renyi entropy (see Appendix C for details). Then, the Renyi entropy is expected to follow S()=αγS(\mathcal{L})=\alpha\mathcal{L}-\gamma, where α\alpha depends on the details of the state, \mathcal{L} represents the boundary length of a contractible patch with codimension-1 boundary in the system and γ\gamma is the universal TEE. In order to eliminate the area-law contribution α\alpha\mathcal{L}, we calculate the entanglement entropy of plaquette P1\mathrm{P}_{1} (the shaded region in the inset of Fig. 5) with different sizes, and the result is presented in Fig. 5. It shows that S()S(\mathcal{L}) increases linearly with \mathcal{L}. Then, we apply a linear extrapolation to 0\mathcal{L}\rightarrow 0, and obtain the TEE for the SQSH state as γ0.748\gamma\approx 0.748, which is very close to ln2=0.693\ln 2=0.693. It shows that the SQSH state has intrinsic topological order. This is different from the quantum spin Hall state which is a symmetry-protected topological state.

Refer to caption
Figure 5: Entanglement entropy as a function of the area for the SQSH state with lattice size 12×12×312\times 12\times 3. The corresponding optimal parameters are obtained when D=0.1D=0.1, J2=0.05J_{2}=0.05, J3=0J_{3}=0 and Jd=0.27J_{d}=0.27. The xx-axis labeled as \mathcal{L} means that the area is 2\mathcal{L}^{2} times of the primitive cell, i.e., the shaded rhombus as shown in the insert, where different colored sites label the different sublattices of kagome lattice.
  state   ε1\varepsilon_{1}   ε2\varepsilon_{2}   ε3\varepsilon_{3}   ε4\varepsilon_{4}   ngn_{g}
𝑸=0\bm{Q}=0 3.997 1.3×103\times 10^{-3} 9.72×104\times 10^{-4} 9.27×104\times 10^{-4} 1
SQSH 1.985 0.6774 0.673 0.6646 4
Table 4: Eigenvalues and GSDs of the 𝑸=0\bm{Q}=0 and SQSH states, calculated in a kagome lattice of 12×\times12×\times3. And εi\varepsilon_{i} denotes the eigenvalue of the overlap matrix while ngn_{g} denotes GSD.

The nontrivial topological properties of this state can be further characterized by its GSD. The calculation of GSD is carried out by constructing four states |ϕ±,±mf|\phi_{\pm,\pm}\rangle_{\mathrm{mf}} with a Gutzwiller projection, where the + (-) subscript denotes the period (anti-period) boundary condition along the directions of two lattice basis vectors, respectively. After diagonalizing the overlap matrix on the basis vector of these four states, we can obtain their eigenvalues, and the number of the significant finite eigenvalues is just GSD. In Table. 4, the results of four eigenvalues and the corresponding GSD are presented for the 𝑸=0\bm{Q}=0 magnetic order state and SQSH state. It shows that there is only one significant finite eigenvalue for the 𝑸=0\bm{Q}=0 state, so its GSD is ng=1n_{g}=1. In the category of topological order [59, 60], this magnetic order only holds a trivial (identity) topological excitation 𝕀\mathbb{I} with its quantum dimension Dq=1D_{q}=1 due to ng=1n_{g}=1. According to the relation γ=lnDq\gamma=\ln D_{q}, we find its topological entanglement entropy is γ=0\gamma=0. Besides, its Chern number is found to be zero. So, the 𝑸=0\bm{Q}=0 magnetic order state is a topological trivial phase without long-range entanglement with γ=0\gamma=0. For the SQSH state, there are four finite eigenvalues and its GSD is ng=4n_{g}=4. From the above result of the TEE for the SQSH state γln2\gamma\approx\ln 2, we obtain its quantum dimension Dq=2D_{q}=2. For an Abelian topological phase, GSD = Dq2D_{q}^{2}. Therefore, we suggest that this SQSH state is an Abelian topological phase.

IV conclusions

In summary, we have studied the interplay between the Dzyaloshinskii-Moriya interaction and the long-range AFM Heisenberg interactions in the kagome antiferromagnet by using the variational Monte Carlo method. We find that the Dzyaloshinskii-Moriya interaction alone favors the 𝑸=0\bm{Q}=0 long-range magnetic order, and an additional antiferromagnetic interaction across the diagonals of the hexagons of the kagome lattice can suppress and eliminate eventually this order phase. This topological phase transition leads to an exotic quantum spin state with fruitful topological properties. We elaborate that it is a topological gapped quantum spin liquid with the ground-state degeneracy ng=4n_{g}=4 and the topological entanglement entropy γ=ln2\gamma=\ln 2. As the fractionalized excitations in a quantum spin liquid, the spinons constitute the two chiral edge states protected by the time-reversal symmetry. So, the spin-up and -down spinons move along opposite directions on a given edge, and it gives rise to the quantum spin Hall effect existing in a topological insulator.

We suggest that, by doping magnetic impurities into this spinon quantum spin Hall state to suppress one of the so-call helical states and retain the topological properties of the system as has been done for the celebrated quantum anomalous Hall effect [61], it is possible to detect the fractionalized spinons in this exotic quantum spin liquid.

Acknowledgements.
We would like to thank Q.-H. Wang, Z.-X. Liu, Z.-L. Gu, X.-M. Cui and J.-B. Liao for many helpful and valuable discussions. This work was supported by National Key Projects for Research and Development of China (Grant No. 2021YFA1400400) and the National Natural Science Foundation of China (No. 92165205).

Appendix A MEAN-FIELD ANSATZES

1. Q=0\bm{Q}=0 magnetic order

Firstly, we consider a simple classical magnetic order 𝑸=0\bm{Q}=0 in the lattice (XY) plane. This state is constructed based on the solution with only a finite 𝑴i\bm{M}_{i} term in HmfH_{\mathrm{mf}}, and is restricted to the Stotz=0S_{tot}^{z}=0 subspace [14] by the application of the projector PStotz=0P_{S_{tot}^{z}=0}. Then, quantum fluctuations are included by the long-range Jastrow projector:

PJz=exp(1/2ijμijSizSjz),P_{J}^{z}=\mathrm{exp}\left(1/2\sum_{ij}\mu_{ij}S_{i}^{z}S_{j}^{z}\right), (7)

where μij\mu_{ij} is the pseudopotential that only depends on the absolute distance |𝑹i𝑹j||\bm{R}_{i}-\bm{R}_{j}| of two sites. It decays exponentially with the distance so that we can just consider the first three without loss of generality. Finally, we get the state as |Ψmag=PStotz=0PJz|GSmf|\mathrm{\Psi_{\mathrm{mag}}}\rangle=P_{S_{tot}^{z}=0}P_{J}^{z}|\mathrm{GS}\rangle_{\mathrm{mf}}, and call it the Jastrow + (𝑸=0)(\bm{Q}=0) state.

2. DSL

The Dirac quantum spin liquid (DSL) is one of the most competitive candidate ground states in the Heisenberg model with only J1J_{1} term on kagome lattice. In this U(1) QSL, all triplet terms vanish and only singlet hopping terms survive. If we only consider the first NN hopping ones, there are zero (π\pi) fluxes through triangles (hexagons), respectively. Based on PSG, the second singlet hopping terms can also exist. All bond patterns are shown in Fig. 1 in the main text. We also emphasize that the third NN terms in this state are forbidden by PSG, so that we rationally abandon them. In fact, the second NN singlet pairing terms are allowed by PSG, namely, the so-called Z[0,π]2β{}_{2}[0,\pi]\beta, a gapped QSL [53]. But, it’s not energy favorable in our model by our calculation. In detail, our numerical results suggest that the J1J_{1}, JdJ_{d} and J3J_{3} interactions suppress this pairing term, which is consist with Ref. [14]. For this reason, we also throw away the second NN singlet pairing terms all the time.

3. SQSH

The SQSH state is a novel state we found in this paper. In this state, the first and second NN singlet hopping terms have the same forms as those of the DSL discussed above, but the first NN triplet hopping term is finite and pure imaginary number. According to the direction of the DM interaction vector 𝑫ij\bm{D}_{ij} we chosen in this paper, this triplet hopping term will lead to that the spin-up (-down) spinon sees +()θ+(-)\theta flux in all triangles and π(+)2θ\pi-(+)2\theta flux in all hexagons. Intrinsically, the spin-up and -down spinons have opposite Chern number C,=±1C_{\uparrow,\downarrow}=\pm 1.

As shown in Fig. 1, when 0.01D<0.20.01\lesssim D<0.2 and other interactions beyond J1J_{1} term are absent, another magnetic order state as a magnetic instability of SQSH, we call it the SQSH + (𝑸=0\bm{Q}=0) state, is energetically favored. For convenience, we stipulate the unique 𝑸=0\bm{Q}=0 magnetic ordered state mentioned in the main text is just the SQSH + (𝑸=0\bm{Q}=0) state, because the Jastrow + (𝑸=0\bm{Q}=0) state is not energy favorable. Henceforth, we also call it the 𝑸=0\bm{Q}=0 state.

4. CSL

We have considered the chiral quantum spin liquid (CSL), in which the first and second NN singlet hopping terms are complex numbers as discussed in Ref. [20]. But, we find that their imaginary parts are almost zero (Im(tijs)/tijs<102\mathrm{Im}(t_{\langle ij\rangle}^{s})/t_{\langle ij\rangle}^{s}<10^{-2}, Im(tijs)/tijs<102\mathrm{Im}(t_{\langle\langle ij\rangle\rangle}^{s})/t_{\langle ij\rangle}^{s}<10^{-2}) in our variational process. At least, in the range of interactions we considered in this work, this result excludes the chiral QSL stabilized by strong enough JdJ_{d} [20, 23].

5. cuboc1cuboc1 magnetic order

We also study the so-called cuboc1cuboc1 magnetic order as the instability of the TR symmetry breaking chiral spin liquid, i.e., CSL + cuboc1cuboc1 [17]. Without loss of generality, we additionally include the first NN triple hopping term t<ij>,ztt_{<ij>,z}^{t} into this ansatz. The detail of the classical cuboc1cuboc1 order is shown in Fig. 6. For the sake of simplicity, we call it cuboc1cuboc1 state. We indeed find that the variational energy of this magnetic order is lower than that of the SQSH state, when the JdJ_{d} term is relatively large, as listed in Table 2 and 3. Before the phase transition, we find that the cuboc1cuboc1 order almost reduces to SQSH state by VMC calculation because the optimal parameters t<ij>st_{<ij>}^{s} and t<<ij>>st_{<<ij>>}^{s} are real number, the optimal parameter t<ij>,ztt_{<ij>,z}^{t} is pure imaginary and the magnetic moment MM is almost vanishing. As a result, the energies of both cuboc1cuboc1 order and SQSH state look like degeneracy within numerical error. Therefore, we believe this exotic SQSH state can survive in a relatively broad range of JdJ_{d}.

Refer to caption
Figure 6: 12 sublattices of the magnetic unit cell for classical cuboc1cuboc1 order. 𝑺1=(1,0,0)\bm{S}_{1}=(1,0,0), 𝑺2=(12,32,0)\bm{S}_{2}=(-\frac{1}{2},\frac{\sqrt{3}}{2},0), 𝑺3=(12,36,63)\bm{S}_{3}=(-\frac{1}{2},\frac{\sqrt{3}}{6},\frac{\sqrt{6}}{3}), 𝑺4=(0,33,63)\bm{S}_{4}=(0,\frac{\sqrt{3}}{3},-\frac{\sqrt{6}}{3}), 𝑺5=(12,32,0)\bm{S}_{5}=(\frac{1}{2},-\frac{\sqrt{3}}{2},0), 𝑺6=(12,32,0)\bm{S}_{6}=(-\frac{1}{2},-\frac{\sqrt{3}}{2},0), 𝑺7=(1,33,63)\bm{S}_{7}=(1,-\frac{\sqrt{3}}{3},\frac{\sqrt{6}}{3}), 𝑺8=(12,36,63)\bm{S}_{8}=(-\frac{1}{2},-\frac{\sqrt{3}}{6},-\frac{\sqrt{6}}{3}), 𝑺9=(12,36,63)\bm{S}_{9}=(\frac{1}{2},-\frac{\sqrt{3}}{6},-\frac{\sqrt{6}}{3}), 𝑺10=(1,0,0)\bm{S}_{10}=(-1,0,0), 𝑺11=(12,36,63)\bm{S}_{11}=(\frac{1}{2},\frac{\sqrt{3}}{6},\frac{\sqrt{6}}{3}) and 𝑺12=(12,32,0)\bm{S}_{12}=(\frac{1}{2},\frac{\sqrt{3}}{2},0).

6. uRVB state

In addition, we have considered another two candidate states. Both of them are regarded as the derivative states of uRVB states. Without loss in generality, we allow the first NN singlet tijst_{ij}^{s} terms are complex number so that there are gauge fluxes though triangular and hexagonal plaquette. And we consider two flux patterns, one where fluxes through all triangles are the same and another one where fluxes through up and down triangles are the opposite. Combining the triplet hopping tij,ztt_{ij,z}^{t} terms (complex number) and finite 𝑴i\bm{M}_{i} of the 𝑸=0\bm{Q}=0 magnetic order, we find that the trial energies of both states are much higher than aforementioned states. So, we discard them and do not draw up their energy curves in Fig. 1 in main text.

Appendix B CHERN NUMBER

Nonzero Chern number is one of the fundamental topological numbers to characterize topological phase of matter. Here we won’t go into details about the concepts of Berry connection, Berry phase and Chern number with formal analytical expression. We just introduce the numerical calculation of Chern number in the filled bands [62].

Firstly, we obtain the corresponding mean-field Hamiltonian with the optimized variational parameters by VMC and then transform it into 𝒌\bm{k}-space, H(𝒌)H(\bm{k}). And we emphasize that the H(𝒌)H(\bm{k}) is periodic along the directions of reciprocal basis vectors 𝒃1,2\bm{b}_{1,2}, H(𝒌)=H(𝒌+n1𝒃1+n2𝒃2)H(\bm{k})=H(\bm{k}+n_{1}\bm{b}_{1}+n_{2}\bm{b}_{2}), where n1,2n_{1,2} are any integer numbers. In other word, H(𝒌)H(\bm{k}) is in Bloch form. Therefore, this Fourier transformation must be handled with care. To be specific, it usually needs a gauge transformation, c𝒌c𝒌ei𝒌𝜹c_{\bm{k}}\longrightarrow c_{\bm{k}}e^{i\bm{k}\cdot\bm{\delta}}. In general, the 𝜹\bm{\delta} is different for different c𝒌c_{\bm{k}} and not unique.

Then, for a lattice with finite size, the Brillouin zone is filled with discrete 𝒌\bm{k} points. And we define intervals of 𝒌\bm{k} points in two directions of reciprocal basis vectors,

𝒖i=li𝒃i2Niπ,(i=1,2;Ni/li𝐍).\bm{u}_{i}=\frac{l_{i}\bm{b}_{i}}{2N_{i}\pi},\quad(i=1,2;\quad N_{i}/l_{i}\in\bm{\mathrm{N}}^{*}). (8)

In our calculation, we take li=1l_{i}=1 to guarantee the highest numerical precision. We also note larger intervals are also allowed as long as the result is convergent. And then, we require that the eigenstate |n(𝒌)|n(\bm{k})\rangle of H(𝒌)H(\bm{k}) is also periodic in the Brillouin zone to eliminate the effect of any U(1) gauge of eigenstate. Now we can define the U(1) quantity for a certain 𝒌\bm{k} as following,

η(𝒌)𝒖in(𝒌)|n(𝒌+𝒖i)|n(𝒌)|n(𝒌+𝒖i)|.\eta(\bm{k})_{\bm{u}_{i}}\equiv\frac{\langle n(\bm{k})|n(\bm{k}+\bm{u}_{i})\rangle}{|\langle n(\bm{k})|n(\bm{k}+\bm{u}_{i})\rangle|}. (9)

η(𝒌)𝒖i\eta(\bm{k})_{\bm{u}_{i}} is well defined as long as the denominator of Eq. 9 is nonzero. And then, we can define another variable about phase in a loop with η(𝒌)𝒖i\eta(\bm{k})_{\bm{u}_{i}},

θ(𝒌)=1iln(η(𝒌)𝒖1η(𝒌+𝒖1)𝒖2η(𝒌+𝒖2)𝒖1η(𝒌)𝒖2),\displaystyle\theta(\bm{k})=\frac{1}{i}\ln\left(\eta(\bm{k})_{\bm{u}_{1}}\eta(\bm{k}+\bm{u}_{1})_{\bm{u}_{2}}{\eta(\bm{k}+\bm{u}_{2})^{\dagger}_{\bm{u}_{1}}}{\eta(\bm{k})^{\dagger}_{\bm{u}_{2}}}\right), (10)
π<θ(𝒌)π.\displaystyle-\pi<\theta(\bm{k})\leq\pi.

Finally, the Chern number of the nnth filled band is obtained by,

Cn12π𝒌BZθ(𝒌).C_{n}\equiv\frac{1}{2\pi}\sum_{\bm{k}\in\mathrm{BZ}}\theta(\bm{k}). (11)

Appendix C GROUND-STATE DEGENERACY AND TOPOLOGICAL ENTANGLEMENT ENTROPY

The low-energy gauge fluctuations of gapped QSLs (topological orders) are characterized by ground-state degeneracy (GSD). When one compacts the lattice to a torus, in the thermodynamics limit, there is no energy cost when a Z2 π\pi flux is inserted in any hole of the torus. In the mean-field theory, this procedure is equivalent to changing the boundary condition of HmfH_{\mathrm{mf}} from period to anti-period. For a two-dimensional system, in general, we can always construct four states |ϕ±,±mf|\phi_{\pm,\pm}\rangle_{\mathrm{mf}}, where the + (-) subscript denotes the periodic (anti-periodic) boundary condition along the directions of two lattice basis vectors, respectively. Then, we enforce a Gutzwiller projection to these four ground states of HmfH_{\mathrm{mf}} to recover physical Hilbert space. Thus, we rewrite the symbols of the four states for convenience,

|1=PG|ϕ+,+mf,|2=PG|ϕ+,mf,\displaystyle|1\rangle=P_{G}|\phi_{+,+}\rangle_{\mathrm{mf}},\quad|2\rangle=P_{G}|\phi_{+,-}\rangle_{\mathrm{mf}}, (12)
|3=PG|ϕ,+mf,|4=PG|ϕ,mf.\displaystyle|3\rangle=P_{G}|\phi_{-,+}\rangle_{\mathrm{mf}},\quad|4\rangle=P_{G}|\phi_{-,-}\rangle_{\mathrm{mf}}.

We can calculate the 4 by 4 overlap matrix 𝒪\mathcal{O} based on these four states. In detail, the matrix element 𝒪ij=i|j/i|ij|j\mathcal{O}_{ij}=\langle i|j\rangle/\sqrt{\langle i|i\rangle\langle j|j\rangle}, where i,j=1,2,3,4i,j=1,2,3,4. After diagonalizing this overlap matrix, we can obtain its eigenvalues. The number of the significantly finite eigenvalues is just GSD. We calculate the overlap matrices of the 𝑸=0\bm{Q}=0 magnetically ordered state and SQSH states as following,

𝒪𝑸=0(1ei0.29ei0.79ei0.76ei0.291ei0.5ei0.46ei0.79ei0.51ei0.03ei0.76ei0.46ei0.031),\displaystyle\mathcal{O}_{\bm{Q}=0}\simeq\begin{pmatrix}1&e^{-i0.29}&e^{-i0.79}&e^{-i0.76}\\ e^{i0.29}&1&e^{-i0.5}&e^{-i0.46}\\ e^{i0.79}&e^{i0.5}&1&e^{i0.03}\\ e^{i0.76}&e^{i0.46}&e^{-i0.03}&1\end{pmatrix}, (13)
𝒪SQSH\displaystyle\mathcal{O}_{\mathrm{SQSH}}\simeq
(10.32ei2.640.33ei2.130.33ei3.050.32ei2.6410.32ei1.510.33ei0.420.33ei2.130.32ei1.5110.33ei1.10.33ei3.050.33ei0.420.33ei1.11).\displaystyle\begin{pmatrix}1&0.32e^{i2.64}&0.33e^{-i2.13}&0.33e^{i3.05}\\ 0.32e^{-i2.64}&1&0.32e^{i1.51}&0.33e^{i0.42}\\ 0.33e^{i2.13}&0.32e^{-i1.51}&1&0.33e^{-i1.1}\\ 0.33e^{-i3.05}&0.33e^{-i0.42}&0.33e^{i1.1}&1\end{pmatrix}.

And now, we can obtain their eigenvalues and GSDs, as shown in Table 4.

Another important quantity to characterize topological order is topological entanglement entropy (TEE) [5, 6]. Firstly, we divide a system into two parts, as shown in Fig. 7.

Refer to caption

(a)

Refer to caption

(b)

Figure 7: (a) shows a schematic diagrm of bipartition on an arbitrary system. (b) indicates the system is compacted to a torus. And this bipartition is trivial so that the boundaries are contractile.

And then, the von Neumann entanglement entropy of P1 for a system can be represented as follows,

S(P1)=αγ,S({\mathrm{P_{1}}})=\alpha\mathcal{L}-\gamma, (14)

where the coefficient α\alpha is not universal and depends on the details of the state, \mathcal{L} is the codimensional-1 boundary of P1, and γ\gamma is just the universal TEE. As we know, S(P1)0S({\mathrm{P_{1}}})\geq 0 for an arbitrary gapped system [63]. For those systems with γ=0\gamma=0, it is possible to deform the ground state to obtain α=0\alpha=0, namely one can remove the leading area-low contribution (α=0\alpha\mathcal{L}=0). Thus, the state with α=0\alpha=0 is the pure direct product one without long-range entanglement, i.e. non-topological gapped phase. On the contrary, the deformation could never occur for the state with γ>0\gamma>0, which is the case for a topological gapped phase. Therefore, the nonzero γ\gamma is a direct sign of long-range entanglement.

Numerical calculation of TEE from the von Neumann entropy is difficult, so we focus on the Renyi entropy here. The Renyi entropy for the gapped state with bipartition is defined as [64],

Sn=11nln[Tr(ρ1n)],S_{n}=\frac{1}{1-n}\ln\left[\mathrm{Tr}\left(\rho_{1}^{n}\right)\right], (15)

where ρ1\rho_{1} is the reduced density matrix obtained by tracing out the subsystem P2, ρ1=Tr2|ΨΨ|\rho_{1}=\mathrm{Tr}_{2}|\Psi\rangle\langle\Psi|, where |Ψ|\Psi\rangle is a normalized wave function of the system. In this paper, we just focus on the Renyi entropy with index n=2n=2, S2=ln[Tr(ρ12)]S_{2}=-\ln\left[\mathrm{Tr}\left(\rho_{1}^{2}\right)\right]. We define a swap operator XX[65] with the purpose as follow,

X|α1|α2=|β1|β2,X|\alpha_{1}\rangle\otimes|\alpha_{2}\rangle=|\beta_{1}\rangle\otimes|\beta_{2}\rangle, (16)

where |α1=|a|b|\alpha_{1}\rangle=|a\rangle|b\rangle and |α2=|m|n|\alpha_{2}\rangle=|m\rangle|n\rangle are two configurations, the |a|a\rangle and |m|m\rangle are in P1 while |b|b\rangle and |n|n\rangle are in P2, and then |β1=|m|b|\beta_{1}\rangle=|m\rangle|b\rangle and |β2=|a|n|\beta_{2}\rangle=|a\rangle|n\rangle.

Finally, we can rewrite S2S_{2} in terms of the expectation of XX with respect to the wave function |Ψ|Ψ|\Psi\rangle\otimes|\Psi\rangle, S2=lnXS_{2}=-\ln\langle X\rangle. Empirically, X\langle X\rangle is predicted to be a complex number in actual calculation if |Ψ|\Psi\rangle is complex. So we can divide this expectation into two parts, X=XmodXphase\langle X\rangle=\langle X_{mod}\rangle\langle X_{phase}\rangle, which can be individually calculated by Monte Carlo (MC) method as shown in Eq. 17. It is worth mentioning that ρ~α1,α2\tilde{\rho}_{\alpha_{1},\alpha_{2}} is a joint probability distribution. Besides, we note that for large size \mathcal{L}, the computational cost is relatively high because we have to taken more samples in the MC process to reduce numerical error. Therefore, we calculate the S()S(\mathcal{L}) with =1\mathcal{L}=1 up to 44 to eliminate the area contribution and obtain the TEE.

Xmod=α1,α2ρα1ρα2|f(α1,α2)|,\displaystyle\langle X_{mod}\rangle=\sum_{\alpha_{1},\alpha_{2}}\rho_{\alpha_{1}}\rho_{\alpha_{2}}\left|f(\alpha_{1},\alpha_{2})\right|, (17)
Xphase=α1,α2ρ~α1,α2eiθ(α1,α2),\displaystyle\langle X_{phase}\rangle=\sum_{\alpha_{1},\alpha_{2}}\tilde{\rho}_{\alpha_{1},\alpha_{2}}e^{i\theta(\alpha_{1},\alpha_{2})},
ραi=|αi|Ψ|2Ψ|Ψ,f(α1,α2)=β1|Ψβ2|Ψα1|Ψα2|Ψ,\displaystyle\rho_{\alpha_{i}}=\frac{\left|\langle\alpha_{i}|\Psi\rangle\right|^{2}}{\langle\Psi|\Psi\rangle},f(\alpha_{1},\alpha_{2})=\frac{\langle\beta_{1}|\Psi\rangle\langle\beta_{2}|\Psi\rangle}{\langle\alpha_{1}|\Psi\rangle\langle\alpha_{2}|\Psi\rangle},
ρ~α1,α2=|α1|Ψα2|Ψ|2|f(α1,α2)|α1,α2|α1|Ψα2|Ψ|2|f(α1,α2)|,\displaystyle\tilde{\rho}_{\alpha_{1},\alpha_{2}}=\frac{\left|\langle\alpha_{1}|\Psi\rangle\langle\alpha_{2}|\Psi\rangle\right|^{2}\left|f(\alpha_{1},\alpha_{2})\right|}{\sum_{\alpha_{1},\alpha_{2}}\left|\langle\alpha_{1}|\Psi\rangle\langle\alpha_{2}|\Psi\rangle\right|^{2}\left|f(\alpha_{1},\alpha_{2})\right|},
eiθ(α1,α2)=f(α1,α2)|f(α1,α2)|.\displaystyle e^{i\theta(\alpha_{1},\alpha_{2})}=\frac{f(\alpha_{1},\alpha_{2})}{\left|f(\alpha_{1},\alpha_{2})\right|}.

References