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

Nematic chiral spin liquid in a Kitaev magnet under external magnetic field

Shun-Yao Yu Department of Physics, Beihang University, Beijing 100191, China    Han Li Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China    Qi-Rong Zhao Department of Physics, Renmin University of China, Beijing 100872, China    Yuan Gao Department of Physics, Beihang University, Beijing 100191, China CAS Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China    Xiao-Yu Dong Hefei National Laboratory, Hefei 230088, China    Zheng-Xin Liu liuzxphys@ruc.edu.cn Department of Physics, Renmin University of China, Beijing 100872, China    Wei Li w.li@itp.ac.cn CAS Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China CAS Center of Excellence in Topological Quantum Computation, University of Chinese Academy of Sciences, Beijing 100190, China Peng Huanwu Collaborative Center for Research and Education, Beihang University, Beijing 100191, China    Shou-Shu Gong shoushu.gong@buaa.edu.cn Department of Physics, Beihang University, Beijing 100191, China School of Physical Sciences, Great Bay University, Dongguan 523000, China
(July 28, 2025)
Abstract

The possible existence of a quantum spin liquid (QSL) phase, an exotic state of matter with long-range quantum entanglement and fractionalized excitations, in α\alpha-RuCl3 has sparked widespread interests in exploring QSLs in various Kitaev models under magnetic fields. Recently, a KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model has been proposed to accurately describe the compound by fitting the measured thermodynamic data [Nat. Commun. 12, 4007 (2021)], where JJ is Heisenberg interaction, and Γ\Gamma, Γ\Gamma^{\prime} are off-diagonal exchanges on top of the dominant Kitaev coupling KK. Based on this effective model, an intermediate QSL phase in presence of an out-of-plane fields along the [111][111] direction, between the low-field zigzag order and high-field polarized phase, has been predicted. By combining density matrix renormalization group (DMRG), exponential tensor renormalization group (XTRG), and variational Monte Carlo (VMC) calculations, we address the nature of this QSL phase in the honeycomb KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model under the [111][111]-direction field. Our DMRG calculations find the algebraic-like decay of spin correlation function, the finite spin scalar chiral order, and lattice nematic order. Together with the XTRG results of power-law specific heat at low temperature, our findings naturally suggest a gapless nematic chiral spin liquid. On the other hand, our VMC study finds a gapped nematic chiral spin liquid with the variational energy very close to that obtained in DMRG. As VMC finds a very small gap that is beyond the current resolution of both ground-state DMRG and finite-temperature XTRG calculations on finite-width cylinders, we resort the full clarification for the nature of the intermediate QSL to future studies. Lastly, we discuss the implications of our results to the recent experiment on α\alpha-RuCl3 and the QSL-like phase in a similar KK-Γ\Gamma-Γ\Gamma^{\prime} model.

I Introduction

Quantum spin liquid (QSL) is an exotic quantum state of matter where the spins of electrons remain disordered even at absolute zero temperature [1, 2]. Unlike conventional phases of matter, QSL states do not fit the Landau-Ginzburg-Wilson framework, and exhibit long-range quantum entanglement and fractionalized excitations [3, 4, 5]. The exactly soluble Kitaev honeycomb model plays the milestone role for understanding the physics in QSL. In particular, a non-Abelian chiral spin liquid (CSL) can be induced by breaking time-reversal symmetry in the Kitaev model, which possesses the non-Abelian anyon that has potential application in topological quantum computing [6, 7]. Experimentalists have been searching for the Kitaev QSL candidate materials in various compounds with 4d4d and 5d5d transition metal-based honeycomb lattice. These compounds have strong spin-orbit coupling that may induce the bond-dependent Kitaev interaction [8]. Some examples of these compounds are A2IrO3 with A = Li, Na [9], H3LiIr2O6, which is a hydrogen intercalated modification of A2IrO3 [10], and α\alpha-RuCl3 [11, 12, 13, 14]. Recently, the cobalt-based 3d73d^{7} honeycomb magnets such as A2Co2TeO6 [15, 16], A3Co2SbO6 [17, 18], and BaCo2(AsO4)2 [19, 20] are also thought to be candidates of Kitaev materials.

Among these compounds, α\alpha-RuCl3 has been widely recognized as a prominent Kitaev material. Due to the non-Kitaev interactions in the material, the compound exhibits a zigzag magnetic order at low temperature [11, 12]. By suppressing the zigzag order with external magnetic fields [21, 22, 23], in the intermediate fields experimental evidences for a QSL have been reported, e.g., the excitation continuum in the inelastic neutron scattering spectrum [13, 14, 24], the half-quantized thermal Hall conductivity [25, 26, 27] suggesting the presence of gapless Majorana edge modes [28, 29], and the quantum oscillations in the low-temperature longitudinal heat conductivity possibly due to a spinon Fermi surface [30]. Despite these intriguing progresses, nature of this state are still under debate [31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46].

Motivated by the exciting experimental progress, there have been a plenty of theoretical studies in an effort to find QSL in related spin models. Amongst others, the KK-Γ\Gamma-Γ\Gamma^{\prime} model with the Kitaev interaction KK and off-diagonal couplings Γ,Γ\Gamma,\Gamma^{\prime} has been considered as a minimal model to describe a broad class of Kitaev materials including α\alpha-RuCl3 [47, 48, 49, 50, 51, 52, 53]. Besides the well known Kitaev spin liquids, other promising QSL candidates have also been proposed in extended Kitaev models, including the QSL in the antiferromagnetic (AFM) Kitaev model under an intermediate magnetic field [54, 55, 56, 57, 58, 59, 60], different QSL phases in the KK-Γ\Gamma model with or without a magnetic field [61, 62, 63], the multinode QSL [64] and nematic QSL in the KK-Γ\Gamma-Γ\Gamma^{\prime} model [65, 66, 67]. However, it is hard to accurately estimate the microscopic parameters that can match the experimental data and thus it is unclear how to connect the found QSLs in model studies to the compounds in experiment.

Recently, a KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model has been proposed for α\alpha-RuCl3. Multiple accurate many-body simulations have been combined to examine this model, which reproduces the major experimental features in equilibrium and dynamical measurements, giving the dominant ferromagnetic Kitaev interaction, additional Heisenberg coupling JJ, and other parameters [68]. Interestingly, this model under the [111][111]-direction magnetic field shows an emergent QSL phase between the low-field zigzag order and high-field polarized phase [68]. While the variational Monte Calro (VMC) study proposes a gapped CSL, the exact diagonalization results seem to suggest a gapless state [68]. Remarkably, the theoretically predicted two-transition scenario in Ref. [68] has been witnessed by recent high magnetic field experiment on α\alpha-RuCl3 at 35\sim 35 T and 83\sim 83 T, respectively, with the field along the out-of-plane cc^{*} axis [69]. A field-induced intermediate phase is found experimentally in the field-angle phase diagram [69]. The experimental findings and the good agreements with model study open a new route to search for QSL in α\alpha-RuCl3 and call for further understanding on the QSL phase in the KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model.

In this work, we study the nature of the intermediate QSL phase in the KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model using three methods: density matrix renormalization group (DMRG), VMC, and exponential tensor renormalization group (XTRG) for finite-temperature calculations. The DMRG calculations on different cylinder geometries up to the circumference of 66 unit cells identify the existence of an intermediate phase, with a robust nonzero scalar spin chiral order 𝐒i(𝐒j×𝐒k)\langle\mathbf{S}_{i}\cdot(\mathbf{S}_{j}\times\mathbf{S}_{k})\rangle and a lattice nematic order characterizing a C3C_{3} symmetry breaking. In addition, the algebraic-like spin correlation decay observed in the ground state and the power-law behavior of the specific heat CmT2C_{m}\sim T^{2} obtained at low temperature strongly suggest a gapless QSL state.

We further investigate the intermediate phase using the VMC with extended variational parameters, which unveils that allowing the C3C_{3} symmetry breaking can further lower the variational energy of the previously identified gapped CSL [68] and enlarge the spinon gap, giving a variational energy very close to that obtained in DMRG. Using quantitative analysis, we examine how the mean field dispersion varies and find that the spinon gap is very small compared to the total band width (0.02\sim 0.02), which may explain why the spin correlation decay looks like algebraic on our studied cylinders. The spinon density of states which is approximately proportional to energy above the spinon gap can agree with the CmT2C_{m}\sim T^{2} behavior at low temperature. The Van Hove singularity at the higher energy E/|K|0.4E/|K|\sim 0.4 also supports the double peak structure found in the specific heat. The gapless or gapped nature of the QSL may require further simulations on larger sizes, but the predicted low-temperature properties above T2T\approx 2K, which are less affected by finite-size effects, could be tested in high-field experiments on α\alpha-RuCl3.

This paper is organized as follows. In Sec. II, we introduce the KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model and numerical methods employed in this study. The DMRG simulations are demonstrated in Sec. III, and the XTRG thermodynamics results are presented in Sec. IV. In Sec. V, we show the nematic order obtained in the VMC calculations and compare the results with the DMRG and XTRG results. We present our summary and discussion in Sec. VI.

Refer to caption
Refer to caption
Figure 1: Lattice geometries and the 1D phase diagram obtained by DMRG. (a) and (b) show the XC and YC cylinders used in our calculations. The cylinders have the open boundary condition along the XX direction and periodic boundary condition along the YY direction. The three different colors in the solid bonds denote the different dominant Kitaev couplings. The dashed bonds mark the periodic connection. These two cylinders have the size Lx=3,Ly=2L_{x}=3,L_{y}=2 and are denoted as XC2-3 and YC2-3. The external magnetic field is applied perpendicular to the honeycomb plane. (c) Quantum phase diagram of the model Eq. (1). As the magnetic field increases, the system shows a zigzag (ZZ) magnetic order phase for h0.1h\lesssim 0.1, a polarized (PL) phase for h0.4h\gtrsim 0.4, and an intermediate quantum spin liquid (QSL) phase with spin scalar chiral order and lattice nematic order. (d)-(f) show the characteristic intra-sublattice spin structure factor S~(𝐤)\tilde{S}(\bf k) defined in Eq. (2) for the three phases, which are obtained on the YC4-18 cylinder for h=0.02,0.24,0.46h=0.02,0.24,0.46, respectively. The dashed white hexagon denotes the Brillouin zone of the sublattice, and the 𝚪{\bf\Gamma} and 𝐌{\bf M} points are marked in (f).

II Model and Methods

We study the honeycomb-lattice KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model in an out-of-plane magnetic field along the [111][111] direction, which has the Hamiltonian defined as

H=i,jγ[KSiγSjγ+J𝐒i𝐒j+Γ(SiαSjβ+SiβSjα)\displaystyle H=\sum_{\langle i,j\rangle_{\gamma}}[KS_{i}^{\gamma}S_{j}^{\gamma}+J\mathbf{S}_{i}\cdot\mathbf{S}_{j}+\Gamma(S_{i}^{\alpha}S_{j}^{\beta}+S_{i}^{\beta}S_{j}^{\alpha}) (1)
+Γ(SiγSjα+SiγSjβ+SiαSjγ+SiβSjγ)]+hi,αSαi,\displaystyle+\Gamma^{\prime}(S_{i}^{\gamma}S_{j}^{\alpha}+S_{i}^{\gamma}S_{j}^{\beta}+S_{i}^{\alpha}S_{j}^{\gamma}+S_{i}^{\beta}S_{j}^{\gamma})]+h\sum_{i,\alpha}S^{\alpha}_{i},

where i,jγ\langle i,j\rangle_{\gamma} denotes the nearest-neighbor (NN) γ\gamma bond with {α,β,γ}\{\alpha,\beta,\gamma\} being {x,y,z}\{x,y,z\} under a cyclic permutation, 𝐒i={Six,Siy,Siz}\mathbf{S}_{i}=\{S_{i}^{x},S_{i}^{y},S_{i}^{z}\} are the pseudo spin-1/21/2 operators at the site ii. KK denotes the FM Kitaev interaction, JJ is the FM Heisenberg interaction, Γ\Gamma and Γ\Gamma^{\prime} are the off-diagonal spin couplings. Following the microscopic model determined in Ref. [68], we choose the Kitaev coupling as the energy unit K=1.0K=-1.0, J=0.1J=-0.1, Γ=0.3\Gamma=0.3 and Γ=0.02\Gamma^{\prime}=-0.02. Without external magnetic field, this system has the finite magnetic point group D3d×Z2TD_{3d}\times Z_{2}^{T}, where Z2T={E,T}Z_{2}^{T}=\{E,T\} is the time-reversal symmetry group, and each element in D3dD_{3d} stands for a combination of spin rotation and lattice rotation owing to the spin-orbit coupling. In this paper, we focus on the system under an external magnetic field along the [111][111] direction, which gives the Zeeman coupling hi,αSiαh\sum_{i,\alpha}S^{\alpha}_{i} in Eq. (1).

In DMRG calculation of the ground state [70], we study the system on a cylinder geometry with the periodic boundary condition along the circumference direction (YY direction) and the open boundary condition along the axis direction (XX direction). To avoid possible bias from the lattice geometry, we consider two geometries, namely the XC and YC cylinders shown in Figs. 1(a) and 1(b). For the XC cylinder, there is a π/3\pi/3 angle between the YY and XX direction. For the YC cylinder, the YY direction is perpendicular to the XX direction. We denote the system size as XCLyL_{y}-LxL_{x} and YCLyL_{y}-LxL_{x}, where LyL_{y} and LxL_{x} denote the numbers of unit cells along the two directions, respectively. When we show the results on both cylinders together, we use the length along the circumference direction to demonstrate different system sizes, which is denoted as WyW_{y} with the lattice constant as the unit. In our DMRG simulation, we keep up to D=2000D=2000 bond states, which allow us to obtain accurate results on the Ly=6L_{y}=6 cylinder with the truncation error at about 10610^{-6}. To reduce the finite-size effect, we also implement the infinite-DMRG (iDMRG) [71] by using the TeNPy library [72] to check our results, where the system size along the XX direction can be taken as long enough to eliminate the size effect in this direction. In the iDMRG simulations, we denote the lattice as XCLyL_{y} and YCLyL_{y}.

To compute the thermodynamic properties, we employ the exponential tensor renormalization group (XTRG) method [73, 74] to simulate the model on the YC cylinder with Ly=4L_{y}=4 and LxL_{x} up to 1212. The retained bond dimension DD is up to 10001000, which ensures the truncation error smaller than 10410^{-4} down to the lowest temperature T/|K|=0.008T/|K|=0.008.

III DMRG results for the ground state

In this section, we demonstrate our DMRG simulation results, focusing on the QSL phase. Since the phase boundaries have been investigated in a recent study [68], we leave our results on the determination of phase transition points in Appendix A, which identify the transitions at h0.1h\simeq 0.1 and h0.4h\simeq 0.4 as shown in Fig. 1(c), in consistent with the previous finding [68]. With increased magnetic field, an intermediate QSL emerges between a zigzag magnetic order at low field and a polarized phase at high field.

The three phases have different features of spin structure factor. In Figs. 1(d-f), we show the intra-sublattice structure factor defined as

S~(𝐤)=1Ni,jei𝐤𝐫ij[𝐒i𝐒j𝐒i𝐒j],\tilde{S}({\bf k})=\frac{1}{N}\sum_{i,j}e^{-i{\bf k}\cdot{\bf r}_{ij}}[\langle{\bf S}_{i}\cdot{\bf S}_{j}\rangle-\langle{\bf S}_{i}\rangle\cdot\langle{\bf S}_{j}\rangle], (2)

where the summation is for the sites in the same sublattice. The obtained results of S~(𝐤)\tilde{S}({\bf k}) are the same for the two sublattices of the honeycomb lattice. In the zigzag phase [Fig. 1(d)], the prominent peaks at the 𝐌{\bf M} points characterize the zigzag order. In the QSL phase [Fig. 1(e)], a dominant peak appears at the 𝚪{\bf\Gamma} point with a broad peak at a 𝐌{\bf M} point. In the polarized phase [Fig. 1(f)], the featureless structure factor reflects the very short-range spin correlation. In the following, we focus on the characterization of the QSL phase.

Refer to caption
Figure 2: Absolute spin correlation function C(r)C(r) along the XX direction in the QSL phase. rr is the distance of the sites to the reference site in the same row. (a) and (b) are the double-logarithmic plots of the spin correlations for h=0.24h=0.24 on the XC and YC cylinders with Lx=24L_{x}=24. KsK_{s} is the fitted power exponent of algebraic decay. The insets are the semi-logarithmic plots for the same data. (c) and (d) are the similar figures for h=0.34h=0.34. The reference site is chosen on the fifth column of the cylinder.

III.1 Algebraic-like spin correlation function and static spin structure factor

Refer to caption
Figure 3: Static intra-sublattice spin structure factor at h=0.24h=0.24 in the QSL phase. (a) shows the typical results of S~(𝐤)\tilde{S}({\bf k}) obtained on the XC4-24 cylinder, with the white dashed hexagon indicating the Brillouin zone. The red dashed lines denote the two momentum cuts crossing the 𝚪{\bf\Gamma} and 𝐌{\bf M} points, which represent the momentum used in (b) and (c). (b) and (c) show the S~(𝐤)\tilde{S}({\bf k}) along the momentum cuts in (a), in which S~(𝐤)\tilde{S}({\bf k}) are obtained from the Fourier transform of the middle Ly×3LyL_{y}\times 3L_{y} sites.

We first study the spin-spin correlation function C(r)C(r) defined as

C(r)=|𝐒i𝐒j𝐒i𝐒j|,C(r)=|\langle\mathbf{S}_{i}\cdot\mathbf{S}_{j}\rangle-\langle\mathbf{S}_{i}\rangle\cdot\langle\mathbf{S}_{j}\rangle|, (3)

where rr is the distance between the two sites i,ji,j in the same row along the XX direction. In Fig. 2, we show the double-logarithmic plots of the spin correlations on different geometries and system sizes. In all the cases, the spin correlations drop markedly with about an order of magnitude in the distance of r1r\simeq 1, which is similar to the observation in the spin systems with the dominant bond-directional Kitaev interaction [7]. Remarkably, the spin correlations decay slowly at long distance, which can be fitted quite well using the algebraic behavior C(r)1/rKsC(r)\sim 1/r^{K_{s}} with small power exponents Ks1.5K_{s}\lesssim 1.5. Although the spin correlations decay slowly, one can find in Fig. 2 that with growing LyL_{y}, the spin correlations only change slightly, which agrees with the vanished magnetic order in two-dimensional limit (also see Fig. 3 below). With increasing the field, the power exponent KsK_{s} remains small in the QSL phase until the transition to the polarized phase with spin correlations decaying exponentially.

As a comparison, we also demonstrate the semi-logarithmic plots of the same data in the insets of Fig. 2. We fit the most data as shown by the dashed lines, showing the deviation of the longer-distance data from the exponential decay. These analyses strongly suggest the algebraic-like decay of spin correlation, which supports a gapless QSL.

We further discuss the static spin structure factor. For the sake of convenience, we stick to the intra-sublattice structure factor defined in Eq. (2). As shown in Fig. 3 (for XC) and Fig. 1(e) (for YC), S~(𝐤)\tilde{S}({\bf k}) exhibits a dominant peak at the 𝚪{\bf\Gamma} point, which appears to be singular [Fig. 3(b)] and is consistent with the small power exponent (Ks<2K_{s}<2) of spin correlation. In Figs. 3(b) and 3(c), we plot S~(𝐤)\tilde{S}({\bf k}) versus kxk_{x} following the two cuts in Fig. 3(a), crossing the 𝚪{\bf\Gamma} and 𝐌{\bf M} points, respectively. For a direct comparison, we plot the S~(𝐤)\tilde{S}({\bf k}) obtained from the middle Ly×3LyL_{y}\times 3L_{y} sites. With increased LyL_{y}, the peak at S~(𝚪)\tilde{S}({\bf\Gamma}) becomes more diffusive, which agrees with the absent magnetic order. The round peak at the 𝐌{\bf M} point may characterize the residual fluctuation of the neighboring zigzag order.

III.2 Spin scalar chiral order

While the magnetic order is absent in the QSL phase, the CSL found in the previous VMC study [68] motivates us to examine the spin chiral order. We calculate the spin scalar chiral order χijk\chi_{ijk} defined as

χijk=𝐒i(𝐒j×𝐒k),\chi_{ijk}=\langle\mathbf{S}_{i}\cdot(\mathbf{S}_{j}\times\mathbf{S}_{k})\rangle, (4)

where the sites i,j,ki,j,k from the same sublattice form a triangle in the clockwise direction as shown in the inset of Fig. 4(a). For a time-reversal symmetric system, the spontaneously emergent chiral order could either be left-chiral or right-chiral, which is chosen randomly in DMRG simulations [75]. For our studied system with an external magnetic field, we always obtain a certain chirality. We have checked the chiral orders of different types of triangles allowed on the honeycomb lattice, and we find that the chiral orders of the equilateral triangles shown in the inset of Fig. 4(a) have the maximal magnitudes (see the results for other types of triangles in Appendix B). Thus, we focus on the equilateral triangles inside the hexagon. In each hexagon, there are two types of equilateral triangles, and both of them have the same chirality denoted as χn\chi_{n}.

We first check the spatial dependence of the chiral order χn\chi_{n}. Two examples at h=0.24,0.34h=0.24,0.34 on the XC6-18 cylinder are demonstrated in Fig. 4(a). Indeed, the two types of triangles have the same chirality and both of them are quite uniform in the bulk of system, revealing a small finite-size effect along the XX direction. We have also checked the chiral orders using iDMRG simulations, which obtain the results in agreement with the calculations on the finite-length cylinders.

Furthermore, we examine the LyL_{y} dependence of the chiral orders on both the XC and YC cylinders [Fig. 4(b)]. Clearly, the chiral orders only change slightly with growing LyL_{y}, strongly indicating the robust scalar chiral order in the thermodynamic limit. We remark that this finite chiral order is not from a trivial product of local magnetic moments. We have checked the result of 𝐒i(𝐒j×𝐒k)\langle\mathbf{S}_{i}\rangle\cdot(\langle\mathbf{S}_{j}\rangle\times\langle\mathbf{S}_{k}\rangle), which turns out to be vanishingly small in the QSL phase.

The nonzero scalar chiral order has also been found in some other Kitaev systems. One model is a similar KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model either without or with a small magnetic field [76, 77]. Near the dominant Γ\Gamma interaction region, a QSL-like phase with chiral order has been found. Different from the studied model, in this Γ\Gamma-dominant phase the two equilateral triangles inside the hexagon have the opposite chiralities [76, 77]. The other system is the KK-Γ\Gamma model in the [111][111]-direction magnetic field, which has both the uniform chiral and staggered chiral phases on two-leg ladder, with tuning coupling ratio and field strength [78]. The spin chiral order can emerge in magnetically disordered phases of various Kitaev models due to the rich interplay among different interactions and magnetic field.

Refer to caption
Figure 4: Scalar spin chiral order in the QSL phase. (a) The spatial dependence of the chiral orders computed for the equilateral triangles defined following the clockwise direction. ii denotes the column number and the data for h=0.24,0.34h=0.24,0.34 are obtained on the XC6-18 cylinder. (b) System width dependence of the chiral order obtained on various cylinders. WyW_{y} is the width of the cylinder along the circumference direction, with the lattice constant as the unit. The dashed lines are linear extrapolations of the data for the guide to the eye.

III.3 Lattice nematic order

Refer to caption
Figure 5: Size dependence of the lattice nematic order in the QSL phase. (a) and (b) show the definitions of the lattice nematic order parameters ONO_{N} on the XC and YC cylinders, respectively. The nematic order is the bond-energy difference between the bond along the axis (X)(X) and the circumference (Y)(Y) direction. (c) Size dependence of the nematic order on the XC and YC cylinders with different widths WyW_{y} at the field h=0.24h=0.24 and 0.340.34. For a comparison, the nematic orders in the zigzag phase (h=0.02h=0.02) and the polarized phase (h=0.46h=0.46) are also shown.

To rule out the valence bond order in the QSL phase, we have measured the bond energy to detect the translation symmetry breaking. In the bulk of the cylinder, we find a uniform bond energy distribution that agrees with the absent valence bond order. Furthermore, we find the bond energies along the axis (XX) direction and the circumference (YY) direction show a clear difference, indicating a possible lattice nematic order associated with C3C_{3} lattice rotational symmetry breaking. To detect such a nematic order, we can measure the difference of the bond energies as the nematic order parameter ONO_{N}. As shown in Figs. 5(a) and 5(b), we define the bond energies along the three bond directions as EγE_{\gamma} (including all the energy contributions to the NN γ\gamma bond), where γ{x,y,z}\gamma\in\{x,y,z\}. On the XC cylinder, Ex=EzE_{x}=E_{z} is guaranteed due to the symmetry along the YY direction, and the nematic order can be described by ON|ExEy|O_{N}\equiv|E_{x}-E_{y}|. Similarly, on the YC cylinder ONO_{N} can be defined as ON|ExEz|O_{N}\equiv|E_{x}-E_{z}|.

Refer to caption
Figure 6: Entanglement entropy versus the conformal distance l~x\tilde{l}_{x} in the QSL phase. The conformal distance is defined as l~x=Lxπsin(πlxLx)\tilde{l}_{x}=\frac{L_{x}}{\pi}\sin(\frac{\pi l_{x}}{L_{x}}). (a) and (b) are the entropy for h=0.24h=0.24 obtained on the Lx=24L_{x}=24 XC and YC cylinders, respectively. The dashed lines denote the fitting of entropy using the formula S(l~x)=c6ln(l~x)+gS(\tilde{l}_{x})=\frac{c}{6}\ln(\tilde{l}_{x})+g, giving the fitted central charge cc. The insets show the entropy versus the subsystem length lxl_{x}. (c) and (d) are the similar plots for h=0.34h=0.34.

We emphasize that for our studied systems, a nonzero ONO_{N} always exists in the bulk of the system due to the quasi-one-dimensional cylindrical geometry that has already broken the lattice C3C_{3} rotational symmetry, but the LyL_{y} dependence of ONO_{N} can determine whether the ground state has an intrinsic nematic order in two-dimensional limit. Such a scheme for determining nematic order has been adopted in previous DMRG studies on various spin systems [79, 80, 81]: for the quantum states with lattice rotational symmetry breaking, ONO_{N} will remain finite after the extrapolation to LyL_{y}\to\infty; otherwise ONO_{N} decreases rapidly with LyL_{y} and appears to approach zero in the thermodynamic limit.

Following this strategy, we first check the nematic order ONO_{N} in the bulk of the long cylinders for different LxL_{x} to ensure the converged ONO_{N} with system length (not shown here). Then, we examine the LyL_{y} dependence of ONO_{N} as shown in Fig. 5(c). For a comparison, one can see that in the polarized phase (h=0.46h=0.46) ONO_{N} is vanishing small and decreases rapidly to zero, which agrees with the preserved lattice rotational symmetry in this state. On the other hand, in the zigzag phase (h=0.02h=0.02), ONO_{N} increases with LyL_{y} and approaches a finite value, which is consistent with the spacial anisotropy associated with the zigzag spin order. Interestingly, in the QSL phase, ONO_{N} shows relatively large values and the results slightly increase with growing LyL_{y}, strongly indicating a finite nematic order in the thermodynamic limit. We remark that although this method can identify an intrinsic nematic order, the finite-size cylinder calculation will not allow the symmetries along the YY direction to break. Thus, our results in Fig. 5 cannot determine whether the C3C_{3} symmetry breaking state has ExEyEzE_{x}\neq E_{y}\neq E_{z} or two of the three bonds have the same energy.

In the recent studies on a similar KK-Γ\Gamma-Γ\Gamma^{\prime} model in the [111][111]-direction magnetic field, a possible nematic QSL phase has also been found [65, 66, 67]. The iDMRG simulation can obtain the ground state with ExEyEzE_{x}\neq E_{y}\neq E_{z} in some field regime of the nematic phase [67], which may be owing to the infinite system length that allows a complete C3C_{3} symmetry breaking. We have also tested the nematic order of our model using iDMRG (see the details in Appendix C), which can obtain the complete C3C_{3} symmetry breaking in some range of the QSL phase as well, indicating that such a nematic order with ExEyEzE_{x}\neq E_{y}\neq E_{z} is highly possible in the thermodynamic limit. While our results can identify the nematic order, the explicit nature of the nematicity still needs further studies on larger system sizes, e.g., by tensor network simulations [66].

III.4 Entanglement entropy

In this subsection, we calculate and analyze the entanglement entropy versus subsystem length lxl_{x} by cutting the cylinder into two parts. For gapless quasi-one-dimensional systems, the entropy will have a logarithmic correction to the area law and follow the formula [82]

S(lx)=c6ln[(Lxπ)sin(πlxLx)]+g,S(l_{x})=\frac{c}{6}\ln{[(\frac{L_{x}}{\pi})\sin(\frac{\pi l_{x}}{L_{x}})]}+g, (5)

where S(lx)S(l_{x}) is the bipartite entanglement entropy, LxL_{x} is the length of system, gg is a non-universal constant, and cc is the central charge characterizing the number of gapless modes. In principle, the fitting of the central charge from Eq. (5) should be applied for a large LxL_{x}. In our study, we study the entanglement entropy behavior with the system length up to Lx=24L_{x}=24, and we find the obtained central charge results are consistent at different system lengths (see Appendix D).

In Fig. 6, we show the entanglement entropy for h=0.24,0.34h=0.24,0.34 on both the XC and YC cylinders with Lx=24L_{x}=24, including the entanglement entropy versus the conformal distance l~x=(Lx/π)sin(πlx/Lx)\tilde{l}_{x}=(L_{x}/\pi)\sin(\pi l_{x}/L_{x}) and the lxl_{x} dependence of the entanglement entropy in the insets. On both geometries, the entanglement entropy can be fitted well using Eq. (5), giving nonzero central charge. Although the obtained central charge dependents on lattice geometry and magnetic field, the nonzero values support the gapless nature of the phase within present resolution. One inspiring observation is that with growing LyL_{y} for each case, the obtained central charge seems to approach either c0.5c\sim 0.5 or c1.0c\sim 1.0, which may suggest emergent gapless Majorana excitations.

IV Gapless nature identified from thermodynamics

To further explore the nature of the intermediate QSL phase, we turn to study the specific heat CmC_{\rm m} and thermal entropy SS. In a previous work by some of the authors [68], a prominent double-peak/hump feature has been revealed in the curve of the specific heat CmC_{\rm m} versus temperature, for the intermediate field regime, which characterizes two temperature scales THT_{\rm H} (the higher-temperature scale) and TL′′T^{\prime\prime}_{\rm L} (the lower-temperature scale). Below TL′′T^{\prime\prime}_{\rm L}, the calculated specific heat data seem to fall into a power-law decay [68].

To clarify this possible power-law behavior, here we improve the XTRG calculation to larger system sizes, namely YC4-LxL_{x} cylinders (Lx=6,8,10,12L_{x}=6,8,10,12) and down to lower temperature. As shown in Fig. 7, we find that both the position and height of the CmC_{\rm m} hump at TL′′0.07T^{\prime\prime}_{\rm L}\simeq 0.07 (under the magnetic field h=0.34h=0.34) are quite robust with growing system size. Below TL′′T^{\prime\prime}_{L}, the CmC_{\rm m} curves exhibit a power-law decay CmTαC_{\rm m}\sim T^{\alpha} with the power exponent approaching α2\alpha\simeq 2, reflecting the gapless excitations (or with a very small excitation gap) in this low-energy scale.

Besides, we also show the thermal entropy SS in the inset of Fig. 7, where the low-temperature data has well converged versus bond dimension DD and also follows a power-law decay close to ST2S\sim T^{2}. It is noteworthy that there are considerable amount of entropies below TL′′T^{\prime\prime}_{L}, indicating that the low-temperature states are strongly fluctuating and have large density of states in low-energy excitations.

Refer to caption
Figure 7: The low-temperature specific heat CmC_{\rm m} curves under the out-of-plane field h=0.34h=0.34 on the YC4-LL cylinders with LL up to 1212. The corresponding thermal entropy SS are shown in the inset. Below the lower-temperature scale TL′′0.07T^{\prime\prime}_{\rm L}\simeq 0.07 as marked by the black arrow in both the main plot and the inset, the dashed line indicates a power-law decay (T2\sim T^{2}) and serves as a guide to the eye.

V Variational Monte Carlo results

Refer to caption
Figure 8: Mean-field band dispersion of the CSL state either without or with the nematic order. k1k_{1} and k2k_{2} are the two reciprocal lattice basis vectors. The bands are shown as 3D surface plots and viewed from the 45 angle. (a) The band dispersion of the CSL with the C3C_{3} symmetry. (b) Zoom in of the lowest-energy band in (a). (c) The band dispersion of the CSL breaking the C3C_{3} symmetry. (d) Zoom in of the lowest-energy band in (c). These band dispersions are obtained at h=0.3h=0.3.

Finally, we investigate the intermediate phase of the system using VMC, and try to provide a physical picture to help understand part of the above numerical results.

We introduce the fermionic slave-particle representation for spin operator as Siα=12CiσαCiS_{i}^{\alpha}=\frac{1}{2}C_{i}^{\dagger}\sigma^{\alpha}C_{i} with Ci=(ci,ci)C_{i}^{\dagger}=(c_{i\uparrow}^{\dagger},c_{i\downarrow}^{\dagger}), αx,y,z,\alpha\equiv x,y,z, and σα\sigma^{\alpha} the Pauli matrices. To ensure that the Hilbert space of the fermions faithfully represents that of the physical spin, a constraint for particle number N^icici+cici=1\hat{N}_{i}\equiv c_{i\uparrow}^{\dagger}c_{i\uparrow}+c_{i\downarrow}^{\dagger}c_{i\downarrow}=1 is imposed for every site. This complex fermion representation is equivalent to the Majorana representation using bx,by,bz,cb_{x},b_{y},b_{z},c fermions [7]. For the model described by Eq. (1), the most general expression of the mean-field Hamiltonian reads [83]

HmfSL\displaystyle H^{\rm SL}_{\rm mf} =\displaystyle= i,jαβ(γ)(Tr[Uji(0)ΨiΨj]+Tr[Uji(1)Ψi(iRαβγ)Ψj]\displaystyle\sum_{\langle i,j\rangle\in\alpha\beta(\gamma)}\Big{(}\operatorname{Tr}[U_{ji}^{(0)}\Psi_{i}^{\dagger}\Psi_{j}]+\operatorname{Tr}[U_{ji}^{(1)}\Psi_{i}^{\dagger}(iR^{\gamma}_{\alpha\beta})\Psi_{j}] (6)
+\displaystyle+ Tr[Uji(2)ΨiσγΨj]+Tr[Uji(3)ΨiσγRαβγΨj]+H.c.)\displaystyle\operatorname{Tr}[U_{ji}^{(2)}\Psi_{i}^{\dagger}\sigma^{\gamma}\Psi_{j}]+\operatorname{Tr}[U_{ji}^{(3)}\Psi_{i}^{\dagger}\sigma^{\gamma}R^{\gamma}_{\alpha\beta}\Psi_{j}]+{\rm H.c.}\Big{)}
+\displaystyle+ i[Tr(𝝀iΨi𝝉Ψi)+h4αTr(ΨiσαΨi)],\displaystyle\sum_{i}\Big{[}\operatorname{Tr}({\boldsymbol{\lambda}_{i}}\cdot\Psi_{i}{\boldsymbol{\tau}}\Psi^{\dagger}_{i})+{h\over 4}\sum_{\alpha}{\rm Tr}(\Psi_{i}^{\dagger}\sigma_{\alpha}\Psi_{i})\Big{]},

where Ψi=(Ci,C¯i)\Psi_{i}=(C_{i},\bar{C}_{i}), C¯i=(ci,ci)T\bar{C}_{i}=(c_{i\downarrow}^{\dagger},-c_{i\uparrow}^{\dagger})^{T}, and Rαβγ=i(σα+σβ)/2R^{\gamma}_{\alpha\beta}=-i(\sigma^{\alpha}+\sigma^{\beta})/\sqrt{2} is a rotation matrix. Considering the SU(2) gauge symmetry in this fermionic representation, the particle-number constraint N^i=1\hat{N}_{i}=1 is extended to the SU(2) gauge-invariant three-component form Tr(ΨiτΨi)=0\operatorname{Tr}(\Psi_{i}\tau\Psi^{\dagger}_{i})=0, which is implemented by the Lagrangian multipliers λx,y,z\lambda^{x,y,z} with τx,y,z\tau^{x,y,z} being the generators of the SU(2) gauge group. The mean-field matrices Uji0,1,2,3U^{0,1,2,3}_{ji} can be expanded with the identity matrix and τx,y,z\tau^{x,y,z}. For the studied model, the SU(2) gauge symmetry breaks down to its Z2Z_{2} subgroup named the invariant gauge group (IGG) [84].

While a QSL state preserves the space group symmetry, the symmetry group of its mean-field Hamiltonian is an enlarged one called the projective symmetry group (PSG), whose group elements are the space group operations followed by certain SU(2) gauge transformation. We adopt the Kitaev PSG [85, 86], namely the PSG of the mean-field description of the pure Kitaev model, then the general form of the coefficients Uji(m)U_{ji}^{(m)} can be expressed as

Uji(0)z=iθ(η0+ρa+ρc)τ0,\displaystyle U^{(0)^{z}}_{ji}\ =i\theta(\eta_{0}+\rho_{a}+\rho_{c})\tau^{0},
Uji(0)x,y=i(η0+ρa+ρc)τ0,\displaystyle U^{(0)^{x,y}}_{ji}=i(\eta_{0}+\rho_{a}+\rho_{c})\tau^{0},
Uji(1)z=iθη3τc+iθ(ρaρc+ρd+2ρf)(τx+τy),\displaystyle U^{(1)^{z}}_{ji}\ =i\theta\eta_{3}\tau^{c}+i\theta(\rho_{a}-\rho_{c}+\rho_{d}+2\rho_{f})(\tau^{x}+\tau^{y}),
Uji(1)x,y=iη3τc+i(ρaρc+ρd+2ρf)(τy,x+τz),\displaystyle U^{(1)^{x,y}}_{ji}=i\eta_{3}\tau^{c}+i(\rho_{a}-\rho_{c}+\rho_{d}+2\rho_{f})(\tau^{y,x}+\tau^{z}),
Uji(2)z=iθη5τc+iθ(ρa+ρc)τz+iθρf(τx+τy),\displaystyle U^{(2)^{z}}_{ji}\ =i\theta\eta_{5}\tau^{c}+i\theta(\rho_{a}+\rho_{c})\tau^{z}+i\theta\rho_{f}(\tau^{x}+\tau^{y}),
Uji(2)x,y=iη5τc+i(ρa+ρc)τx,y+iρf(τy,x+τz),\displaystyle U^{(2)^{x,y}}_{ji}=i\eta_{5}\tau^{c}+i(\rho_{a}+\rho_{c})\tau^{x,y}+i\rho_{f}(\tau^{y,x}+\tau^{z}),
Uji(3)z=i(η7z+θρcθρaθρd)(τxτy),\displaystyle U^{(3)^{z}}_{ji}\ =i(\eta^{z}_{7}+\theta\rho_{c}-\theta\rho_{a}-\theta\rho_{d})(\tau^{x}-\tau^{y}),
Uji(3)x,y=±iη7xτc±i(ρcρaρd)(τy,xτz).\displaystyle U^{(3)^{x,y}}_{ji}=\pm i\eta^{x}_{7}\tau^{c}\pm i(\rho_{c}-\rho_{a}-\rho_{d})(\tau^{y,x}-\tau^{z}).

with τc=13(τx+τy+τz)\tau^{c}={1\over\sqrt{3}}(\tau^{x}+\tau^{y}+\tau^{z}). Hence we have the complete set of variational parameters {ρa,ρc,ρd,ρf,η0,η3,η5,η7x,η7z,θ,λ}\{\rho_{a},\rho_{c},\rho_{d},\rho_{f},\eta_{0},\eta_{3},\eta_{5},\eta^{x}_{7},\eta^{z}_{7},\theta,\lambda\}, in which the subset {ρa,ρc,ρd,ρf,η0,η3,η5}\{\rho_{a},\rho_{c},\rho_{d},\rho_{f},\eta_{0},\eta_{3},\eta_{5}\} preserves all the symmetries of the original Hamiltonian Eq. (1). To make the VMC compatible with the spacial nematic order, we introduce three additional parameters {η7x,η7z,θ}\{\eta^{x}_{7},\eta^{z}_{7},\theta\} in the mean-field Hamiltonian to break the C3C_{3} symmetry but preserve a residual mirror symmetry. The mirror plane is parallel to the zz bond and perpendicular to the lattice plane. Then Gutzwiller projection is performed to the mean-field ground state to enforce the particle number constraint. Thus we obtain the trial wave function |Ψ({x})=PG|Ψmf({x})=αf(α)|α|\Psi(\{x\})\rangle=P_{G}|\Psi_{\rm mf}(\{x\})\rangle=\sum_{\alpha}f(\alpha)|\alpha\rangle, where {x}\{x\} denote the variational parameters and α\alpha stands for the Ising bases in the many-body Hilbert space. The energy of the trial state E({x})=Ψ({x})|H|Ψ({x})/Ψ({x})|Ψ({x})=α|f(α)|2γ|f(γ)|2(ββ|H|αf(β)f(α))E(\{x\})=\langle\Psi(\{x\})|H|\Psi(\{x\})\rangle/\langle\Psi(\{x\})|\Psi(\{x\})\rangle=\sum_{\alpha}\frac{|f(\alpha)|^{2}}{\sum_{\gamma}|f(\gamma)|^{2}}(\sum_{\beta}\langle\beta|H|\alpha\rangle\frac{f(\beta)*}{f(\alpha)*}) is computed using Monte Carlo sampling. Then the optimal parameters {x}\{x\} are determined by minimizing the energy E({x})E(\{x\}). Our VMC calculations are performed on the torus geometry up to the size 9×9×29\times 9\times 2 with 162162 sites.

In the absence of the variational parameters {η7x,η7z,θ}\{\eta^{x}_{7},\eta^{z}_{7},\theta\}, an intermediate CSL phase between the zigzag phase and the polarized trivial phase is identified with Chern number ν=2\nu=2 [68]. This CSL is an Abelian QSL with the quasiparticle excitations aa, a¯\bar{a}, ϵ\epsilon, 11, where 11 denotes the vacuum, ϵ\epsilon is the fermion, aa and a¯\bar{a} are two types of vortices with topological spin eiπ/4e^{i\pi/4} [7]. These excitations follow the fusion rules ϵ×ϵ=1\epsilon\times\epsilon=1, a×a¯=1a\times\bar{a}=1, a×ϵ=a¯a\times\epsilon=\bar{a}, a¯×ϵ=a\bar{a}\times\epsilon=a, and a¯×a¯=ϵ\bar{a}\times\bar{a}=\epsilon [7]. The gapless edge of this CSL contains 22 branches of chiral Majorana excitations with total chiral central charge c=1c_{-}=1.

By considering the additional parameters {η7x,η7z,θ}\{\eta^{x}_{7},\eta^{z}_{7},\theta\}, our VMC calculations still support a CSL state, but with a nematic order which further lowers the variational energy. For example, on the 8×8×28\times 8\times 2 size at the magnetic field h=0.3h=0.3, the variational energy of the CSL state with the C3C_{3} symmetry is 0.3691-0.3691. By allowing the C3C_{3} symmetry breaking, the energy is lowered to 0.3746-0.3746 with an emergent nematic order. This optimized variational energy is very close to the energies found in the DMRG simulation, e.g., 0.3790-0.3790 in the bulk of the XC6 cylinder. The details of the VMC identification of the nematic order can be found in Appendix E. Besides the spontaneous nematic order in the new ground state, the obtained low-lying energies also have a drastic change. In Fig. 8, we show the dispersion of the fermion bands in the optimized mean-field Hamiltonian Eq. (6) with h=0.3h=0.3, which indicates that with the C3C_{3} symmetry breaking a small excitation gap is opened.

The mean field dispersion is plotted for a system with 120×120×2120\times 120\times 2 sites. There are totally eight bands due to the two sublattices, two spin components and the particle-hole redundancy. A remarkable feature of the spinon dispersion is that six of the eight bands are quit flat and are close to zero energy, and the rest two bands are very dispersive. The existence of the nearly flat bands is very similar to the mean field spectrum of the pure Kitaev model, where the dispersive band mainly come from the cc fermions and the flat bands result from the bx,y,zb_{x,y,z} fermions. Our variational calculation indicates that four parameters are dominating in the intermediate QSL phase, namely ρa,ρc,η0\rho_{a},\rho_{c},\eta_{0} and η3\eta_{3}, where ρa,ρc\rho_{a},\rho_{c} come from the Kitaev interactions, η0\eta_{0} comes from the Heisenberg interaction and η3\eta_{3} comes from the Γ\Gamma interaction.

It will be illustrative to compare the VMC results with those from DMRG and XTRG. The size of the spinon gap with respect to the total band width is 0.02|K|0.02|K| (we have set the band width of the spinons to be 2|K|2|K|) which is very small. Practically, it is not easy to detect such a small gap using correlation functions. This qualitatively interprets the algebraic-like spin correlation decay observed in DMRG results. Furthermore, above the energy 0.02|K|0.02|K|, the fermionic spinons start to show density of states which is proximately proportional to energy in a small energy window. This agrees with the CmT2C_{m}\propto T^{2} law of the specific heat at the temperature region 0.02|K|0.04|K|0.02|K|\sim 0.04|K|. With further increasing energy, a Van Hove singularity appears in the dispersive band at E0.5|K|E\sim 0.5|K| which yields a large density of states. Hence a peak (or shoulder) is expected to appear in the specific heat CmC_{m} at around 0.5|K|0.5|K|. We notice that the existence of two-temperature scale structure in the specific heat is indeed observed in the XTRG results shown in Fig. 7.

VI Summary and discussion

By combining the DMRG, XTRG, and VMC calculations, we investigate the nature of the intermediate QSL phase of a honeycomb KK-JJ-Γ\Gamma-Γ\Gamma^{\prime} model in an out-of-plane magnetic field along the [111][111] direction, which is sandwiched between a low-field zigzag order phase and a high-field polarized phase.

By using DMRG simulations on different cylinder geometries, we confirm the absent magnetic order and unveil the algebraic-like spin correlation function decay C(r)1/rKsC(r)\sim 1/r^{K_{s}} with small power exponent Ks1.5K_{s}\lesssim 1.5. Besides, we identify a robust spin scalar chiral order 𝐒i(𝐒j×𝐒k)\langle\mathbf{S}_{i}\cdot(\mathbf{S}_{j}\times\mathbf{S}_{k})\rangle and a lattice nematic order characterizing the C3C_{3} symmetry breaking. We also extend the XTRG simulation to longer 4-leg cylinders and down to lower temperature. Below the lower-temperature scale TL′′0.07T^{\prime\prime}_{\rm L}\simeq 0.07, we observe a robust power-law behavior of the specific heat CmT2C_{\rm m}\sim T^{2}, as well as considerable amount of thermal entropy. By allowing C3C_{3} symmetry breaking in the VMC calculation, we obtain the same gapped CSL reported in Ref. [68] but with an additional nematic order that can further lower the variational energy.

While our DMRG results also find the chiral order and nematic order that can agree with the nematic CSL found by VMC, the ground-state and thermal measurements suggest a nearly gapless QSL. Yet, the VMC results provide another picture to understand the finite-size simulation results. The small ratio of the gap with respect to the spinon band width (0.02\sim 0.02) indicates a large correlation length that may go beyond the resolution of present DMRG calculation. The finite density of states for spinons, which are nearly proportional to energy in a small energy window above the gap, also agrees with the behavior of the specific heat CmT2C_{m}\sim T^{2}. Identifying the nature of gapped or gapless may need further studies on larger system size.

It is noteworthy that the QSL predicted here for α\alpha-RuCl3 with relatively large low-energy excitation could be further studied experimentally in high-field NMR [87], ESR spectroscopy [88], and specific heat measurements. Besides, this intermediate QSL may also relate to other recently proposed field-induced states in the Kitaev candidates, e.g., the broad magnetic continuum observed by time-domain terahertz spectroscopy measurements on BaCo2(AsO4)2 [20].

We also notice that this QSL phase might be related to the nematic QSL-like phase in the KK-Γ\Gamma-Γ\Gamma^{\prime} model with dominant FM Kitaev interaction and off-diagonal exchanges Γ>0\Gamma>0, Γ<0\Gamma^{\prime}<0 [65, 66, 67]. While this phase was found not connected to the Kitaev spin liquid, many-body simulations found the C3C_{3} symmetry breaking [66, 67] and gapless-like excitations [67]. It would be interesting to explore the potential connection of the two phases and whether the phase in the KK-Γ\Gamma-Γ\Gamma^{\prime} model may also be a nematic CSL.

Acknowledgements.
We acknowledge stimulating discussions with Lei Wang, Xing-Yu Zhang, and Qiang Luo. This work was supported by the National Natural Science Foundation of China Grants No. 11874078, No. 11834014, No. 12222412, No. 11974036, N0. 11974421, No. 12047503 and No.12134020; the Innovation Program for Quantum Science and Technology (2021ZD0301900); CAS Project for Young Scientists in Basic Research (Grant No. YSBR-057); and China National Postdoctoral Program for Innovative Talents (Grant No. BX20220291).

Appendix A Quantum phase transitions

In Ref. [68], the two phase transitions have been identified from the results of magnetization and susceptibility. Here we calculate entanglement entropy to characterize the phase transitions. As shown in Fig. 9 on the XC cylinders, the entropy exhibits quick drops near the transition points. Although the entropy on the XC4 cylinder behaves differently from the wider sizes near the lower transition point, which may be owing to the finite-size effect, the estimated transition points are very close on different system sizes. Our determined transition points also agree with those found in Ref. [68].

Refer to caption
Figure 9: Field dependence of entanglement entropy and its derivative. (a) shows the entropy SES_{E} versus the field hh on the XC4-18, XC5-18, and XC6-18 cylinders. (b) shows the derivative of SES_{E} to the field dSE/dhdS_{E}/dh.

Appendix B Spin scalar chiral orders in different triangles

In the main text, we have shown the spin scalar chiral order of the equilateral triangles inside the hexagon. Here we demonstrate the chiral orders for other types of triangles. Four typical triangles are shown in Fig. 10, where the chiral order of the triangle Δ4\Delta_{4} has the maximum magnitude and has the opposite chirality to the chiral order shown in the main text (inside the hexagon).

Refer to caption
Figure 10: The scalar spin chiral orders in different types of triangles. The data are obtained for h=0.24h=0.24 on a XC6-18 cylinder. The inset shows the definitions of the four kinds of triangles. ii denotes the column number.

Appendix C Nematic order from iDMRG simulation

In the matin text, we have shown the nematic order obtained by finite DMRG calculation. Here, we show more results from iDMRG calculation. We define Oγ=Eγ(Ex+Ey+Ez)/3O_{\gamma}=E_{\gamma}-(E_{x}+E_{y}+E_{z})/3, where γ=x,y,z\gamma=x,y,z denote the three bond directions. In Fig. 11, we show OγO_{\gamma} obtained by DMRG and iDMRG. In DMRG results, two bond energies are always the same, which is required by the symmetries along the circumference direction. In iDMRG simulation, probably due to the infinite system length allows the complete C3C_{3} symmetry breaking, one can obtain the ground state with ExEyEzE_{x}\neq E_{y}\neq E_{z} in a regime of the intermediate phase, indicating that the complete C3C_{3} symmetry breaking is possible in two-dimensional limit.

Refer to caption
Figure 11: Lattice nematic orders versus the magnetic field. (a) and (b) are the results obtained from DMRG on the XC4 and XC6 cylinders. (c) and (d) are the results obtained from iDMRG on the XC4 and XC6 cylinders. Oγ=Eγ(Ex+Ey+Ez)/3O_{\gamma}=E_{\gamma}-(E_{x}+E_{y}+E_{z})/3, where γ=x,y,z\gamma=x,y,z denote the three bond directions.

Appendix D Fitting of the central charge

In the main text, we have shown the entanglement entropy versus the conformal distance l~x\tilde{l}_{x} on the Lx=24L_{x}=24 cylinders. Here, we show more results of the entropy on different system sizes. In Figs. 12 and 13, we present the entropy obtained on the XC and YC cylinders with different LxL_{x}. All the entropy data can be well fitted using Eq. (5), giving the close central charge on different system lengths.

Refer to caption
Figure 12: Entanglement entropy versus the subsystem length lxl_{x} in the intermediate phase on the XC cylinders. The central charge cc is fitted using the formula Eq. (5), S(lx)=c6ln[(Lxπ)sin(πlxLx)]+gS(l_{x})=\frac{c}{6}\ln{[(\frac{L_{x}}{\pi})\sin(\frac{\pi l_{x}}{L_{x}})]}+g. (a), (c), (e) are the entropy for h=0.24h=0.24 obtained on the Ly=4,5,6L_{y}=4,5,6 XC cylinders, respectively. (b), (d), (f) are the similar plots for h=0.34h=0.34.
Refer to caption
Figure 13: Entanglement entropy versus the subsystem length lxl_{x} in the intermediate phase on the YC cylinders. The central charge cc is fitted using the formula Eq. (5), S(lx)=c6ln[(Lxπ)sin(πlxLx)]+gS(l_{x})=\frac{c}{6}\ln{[(\frac{L_{x}}{\pi})\sin(\frac{\pi l_{x}}{L_{x}})]}+g. (a) and (c) are the entropy for h=0.24h=0.24 obtained on the Ly=4,6L_{y}=4,6 YC cylinders, respectively. (b) and (d) are the similar plots for h=0.34h=0.34.
Table 1: Different variational states and the corresponding energies. The system sizes are chosen from 3×3×23\times 3\times 2 to 9×9×29\times 9\times 2, and the magnetic field is h=0.3h=0.3. Here we use η=(1θ)\eta=(1-\theta). Bdr1 = Bdr2 = 11 denotes the periodic boundary conditions along both the 𝐚1{\bf a}_{1} and 𝐚2{\bf a}_{2} directions. EE is the variational energy per site. While the best variational state on 3×3×23\times 3\times 2 preserves the C3C_{3} symmetry with the vanished parameters {η7x,η7z,η}\{\eta^{x}_{7},\eta^{z}_{7},\eta\}, the best variational states on larger sizes break the C3C_{3} symmetry.
size Bdr1 Bdr2 EE ρa\rho_{a} ρc\rho_{c} ρd\rho_{d} ρf\rho_{f} η0\eta_{0} η3\eta_{3} η5\eta_{5} η7x\eta_{7}^{x} η7z\eta_{7}^{z} η\eta λ0\lambda_{0}
3×3×23\times 3\times 2 1 1 -0.3725 0.9397 0.1847 -0.1887 -0.0079 -0.1778 0.3167 -0.0847 0 0 0 0.0257
3×3×23\times 3\times 2 1 1 -0.3721 0.6129 0.1742 -0.1291 -0.0095 -0.2113 0.3062 -0.0586 -0.0974 0.0418 0 0.0311
3×3×23\times 3\times 2 1 1 -0.3720 0.5928 0.1928 -0.1276 -0.0085 -0.2180 0.3063 -0.0653 -0.1005 0.0406 0.0356 0.0258
4×4×24\times 4\times 2 1 1 -0.3720 1.1047 0.1519 -0.1041 -0.0094 -0.1865 0.2858 -0.1019 0 0 0 0.0280
4×4×24\times 4\times 2 1 1 -0.3743 1.2157 0.1769 -0.1264 -0.0115 -0.1918 0.3187 -0.0938 -0.0494 0.0327 0 0.0365
4×4×24\times 4\times 2 1 1 -0.3742 1.1839 0.1698 -0.1280 -0.0118 -0.1870 0.3179 -0.0887 -0.0554 0.0348 0.0181 0.0394
6×6×26\times 6\times 2 1 1 -0.3715 0.5972 0.1329 -0.1953 -0.0153 -0.1585 0.3386 -0.0715 0 0 0 0.0407
6×6×26\times 6\times 2 1 1 -0.3731 1.3208 0.1746 -0.1183 -0.0078 -0.1770 0.3039 -0.0945 -0.0721 0.0604 0 0.0235
6×6×26\times 6\times 2 1 1 -0.3731 1.3363 0.2109 -0.1219 -0.0167 -0.1882 0.3225 -0.0966 -0.0959 0.0573 0.0008 0.0234
8×8×28\times 8\times 2 1 1 -0.3691 0.8860 0.1738 -0.1822 -0.0072 -0.1793 0.3173 -0.1013 0 0 0 0.0267
8×8×28\times 8\times 2 1 1 -0.3746 0.6559 0.1854 -0.1317 -0.0092 -0.1877 0.3065 -0.0740 -0.0689 0.0232 0 0.0310
8×8×28\times 8\times 2 1 1 -0.3743 0.7866 0.1881 -0.1335 -0.0094 -0.1771 0.3152 -0.0821 -0.0804 0.0322 0.0351 0.0238
9×9×29\times 9\times 2 1 1 -0.3684 1.0719 0.1778 -0.1533 -0.0083 -0.1836 0.3028 -0.0991 0 0 0 0.0276
9×9×29\times 9\times 2 1 1 -0.3740 0.8519 0.1835 -0.1335 -0.0080 -0.1812 0.3172 -0.0834 -0.0754 0.0439 0 0.0291
9×9×29\times 9\times 2 1 1 -0.3739 0.9048 0.1904 -0.1323 -0.0084 -0.1829 0.3183 -0.0816 -0.0784 0.0439 0.0294 0.0264

Appendix E Nematic order in Variational Monte Carlo results

Using VMC calculation, we examine the mean-field Hamiltonian Eq. (6) in the intermediate phase. Here we show the results at the magnetic field h=0.3h=0.3 as an example. On different system sizes with La1×La2×2L_{\rm a_{1}}\times L_{\rm a_{2}}\times 2 (La1L_{\rm a_{1}} and La2L_{\rm a_{2}} denote the numbers of unit cell along the two lattice spacing directions 𝐚1{\bf a}_{1} and 𝐚2{\bf a}_{2}) under the periodic boundary conditions (PBC), which preserve the C3C_{3} symmetry, we calculate the variational energies up to the system size 9×9×29\times 9\times 2, as shown in Table 1. On the smaller 3×3×23\times 3\times 2 system sizes, the variational ground states preserve the C3C_{3} symmetry and have a large overlap with an low-lying exact state obtained by exact diagonalization. Remarkably, on the larger sizes (4×4×2\geq 4\times 4\times 2), the optimized ground state always breaks the C3C_{3} symmetry (see Table 1).

Next, to interpret the difference results for systems with different size, we perform the ‘symmetrization’ procedure among the projected states under different boundary conditions to obtain a C3C_{3} symmetric wave functions. We diagonalize the original Hamiltonian Eq. (1) in the Hilbert space with anti-periodic boundary conditions (APBC), and the obtained results are shown in Table 2, where the parameters {η7x,η7z,η}=0\{\eta^{x}_{7},\eta^{z}_{7},\eta\}=0 (with η=(1θ)\eta=(1-\theta)) for the three projected states in the APBC. For the 3×3×23\times 3\times 2 system size, the energy of the symmetrized ground state is significantly lowered. Actually we have confirmed that the overlap of the symmetrized state has and the exact ground state is close to 1, and its energy also very close to that of the exact state. However, for systems with 6×6×26\times 6\times 2 and 8×8×28\times 8\times 2 sizes, the symmetrized states are not significantly proved (as shown in Table 2 the energy correction is very small) and their energy are still higher than those of the corresponding nematic states in Table 1. This result indicate that with increasing system size, within the Hilbert space of projected ‘non-nematic’ states there does not exist a C3C_{3} symmetrized state whose energy is lower than the ones with nematic order. This result supports the C3C_{3} symmetry breaking in the thermodynamic limit.

For the C3C_{3} symmetry breaking nematic states, we can also perform similar symmetrization procedure in the space spaned by the three nematic states related by the C3C_{3} operation. The results are shown in Table 3. We find that the symmetrization cannot efficiently lower the energy. Furthermore eigenvalues of the fidelity matrix are all close to 11. The reason is that the overlaps between the different nematic states as well as the hopping elements of the Hamiltonian among the three states are all vanishing small, i.e. ψinematic|ψjnematicψinematic|H^|ψjnematic0\langle\psi^{nematic}_{i}|\psi^{nematic}_{j}\rangle\approx\langle\psi^{nematic}_{i}|\hat{H}|\psi^{nematic}_{j}\rangle\approx 0, which indicates a three-fold degeneracy that agrees with a C3C_{3} symmetry breaking.

Table 2: Symmetrization of the variational wavefunctions under the three anti-periodic boundary conditions. The system sizes include 3×3×23\times 3\times 2, 6×6×26\times 6\times 2, and 8×8×28\times 8\times 2. EE is the ground-state energy of the symmetrized state. E+,E_{+,-}, E,+E_{-,+}, and E,E_{-,-} are the ground state energies under the three anti-periodic boundary conditions, before the symmetrization.
size EE |O0|2\sqrt{|O_{0}|^{2}} E+,E_{+,-} E,+E_{-,+} E,E_{-,-}
3×3×23\times 3\times 2 -0.3781 0.9846 -0.3709 -0.3706 -0.3708
6×6×26\times 6\times 2 -0.3701 / -0.3690 -0.3690 -0.3689
8×8×28\times 8\times 2 -0.3701 / -0.3696 -0.3695 -0.3696
Table 3: Symmetrization of the three nematic states. The system sizes include 6×6×26\times 6\times 2 and 8×8×28\times 8\times 2. EE is the ground-state energy of the renormalized state. EnematicE_{nematic} are the ground state energies of the nematic states before the symmetrization. ρ1,ρ2,ρ3\rho_{1},\rho_{2},\rho_{3} are the eigenvalues of the density matrix.
size EE EnematicE_{nematic} ρ1\rho_{1} ρ2\rho_{2} ρ3\rho_{3}
6×6×26\times 6\times 2 -0.3732 -0.3729 0.9558 1.0028 1.0414
8×8×28\times 8\times 2 -0.3757 -0.3757 0.9870 1.0000 1.0130

References

  • Savary and Balents [2016] L. Savary and L. Balents, Quantum spin liquids: a review, Reports on Progress in Physics 80, 016502 (2016).
  • Zhou et al. [2017] Y. Zhou, K. Kanoda, and T.-K. Ng, Quantum spin liquid states, Rev. Mod. Phys. 89, 025003 (2017).
  • Wen [1991] X. G. Wen, Mean-field theory of spin-liquid states with finite energy gap and topological orders, Phys. Rev. B 44, 2664 (1991).
  • Senthil and Fisher [2000] T. Senthil and M. P. A. Fisher, Z2{Z}_{2} gauge theory of electron fractionalization in strongly correlated systems, Phys. Rev. B 62, 7850 (2000).
  • Senthil and Fisher [2001] T. Senthil and M. P. A. Fisher, Fractionalization in the cuprates: Detecting the topological order, Phys. Rev. Lett. 86, 292 (2001).
  • Kitaev [2003] A. Kitaev, Fault-tolerant quantum computation by anyons, Annals of Physics 303, 2 (2003).
  • Kitaev [2006] A. Kitaev, Anyons in an exactly solved model and beyond, Annals of Physics 321, 2 (2006).
  • Jackeli and Khaliullin [2009] G. Jackeli and G. Khaliullin, Mott insulators in the strong spin-orbit coupling limit: From Heisenberg to a quantum compass and Kitaev models, Phys. Rev. Lett. 102, 017205 (2009).
  • Singh et al. [2012] Y. Singh, S. Manni, J. Reuther, T. Berlijn, R. Thomale, W. Ku, S. Trebst, and P. Gegenwart, Relevance of the Heisenberg-Kitaev model for the honeycomb lattice iridates A2IrO3{A}_{2}{I}r{O}_{3}Phys. Rev. Lett. 108, 127203 (2012).
  • Kitagawa et al. [2018] K. Kitagawa, T. Takayama, Y. Matsumoto, A. Kato, R. Takano, Y. Kishimoto, S. Bette, R. Dinnebier, G. Jackeli, and H. Takagi, A spin-orbital-entangled quantum liquid on a honeycomb lattice, Nature 554, 341 (2018).
  • Sears et al. [2015] J. A. Sears, M. Songvilay, K. W. Plumb, J. P. Clancy, Y. Qiu, Y. Zhao, D. Parshall, and Y.-J. Kim, Magnetic order in α\alpha-RuCl3: A honeycomb-lattice quantum magnet with strong spin-orbit coupling, Phys. Rev. B 91, 144420 (2015).
  • Johnson et al. [2015] R. D. Johnson, S. C. Williams, A. A. Haghighirad, J. Singleton, V. Zapf, P. Manuel, I. I. Mazin, Y. Li, H. O. Jeschke, R. Valentí, and R. Coldea, Monoclinic crystal structure of α\alpha-RuCl3 and the zigzag antiferromagnetic ground state, Phys. Rev. B 92, 235119 (2015).
  • Banerjee et al. [2017] A. Banerjee, J. Yan, J. Knolle, C. A. Bridges, M. B. Stone, M. D. Lumsden, D. G. Mandrus, D. A. Tennant, R. Moessner, and S. E. Nagler, Neutron scattering in the proximate quantum spin liquid α\alpha-RuCl3Science 356, 1055 (2017).
  • Do et al. [2017] S.-H. Do, S.-Y. Park, J. Yoshitake, J. Nasu, Y. Motome, Y. Kwon, D. T. Adroja, D. J. Voneshen, K. Kim, T.-H. Jang, J.-H. Park, K.-Y. Choi, and S. Ji, Majorana fermions in the Kitaev quantum spin system α\alpha-RuCl3Nature Physics 13, 1079 (2017).
  • Lin et al. [2021] G. Lin, J. Jeong, C. Kim, Y. Wang, Q. Huang, T. Masuda, S. Asai, S. Itoh, G. Günther, M. Russina, Z. Lu, J. Sheng, L. Wang, J. Wang, G. Wang, Q. Ren, C. Xi, W. Tong, L. Ling, Z. Liu, L. Wu, J. Mei, Z. Qu, H. Zhou, X. Wang, J.-G. Park, Y. Wan, and J. Ma, Field-induced quantum spin disordered state in spin-1/2 honeycomb magnet Na2Co2TeO6Nature Communications 12, 5559 (2021).
  • Yao et al. [2022] W. Yao, K. Iida, K. Kamazawa, and Y. Li, Excitations in the ordered and paramagnetic states of honeycomb magnet Na2Co2TeO6Phys. Rev. Lett. 129, 147202 (2022).
  • Sano et al. [2018] R. Sano, Y. Kato, and Y. Motome, Kitaev-Heisenberg Hamiltonian for high-spin d7{d}^{7} Mott insulators, Phys. Rev. B 97, 014408 (2018).
  • Songvilay et al. [2020] M. Songvilay, J. Robert, S. Petit, J. A. Rodriguez-Rivera, W. D. Ratcliff, F. Damay, V. Balédent, M. Jiménez-Ruiz, P. Lejay, E. Pachoud, A. Hadj-Azzem, V. Simonet, and C. Stock, Kitaev interactions in the Co honeycomb antiferromagnets Na3Co2SbO6 and Na2Co2TeO6Phys. Rev. B 102, 224429 (2020).
  • Zhong et al. [2020] R. Zhong, T. Gao, N. P. Ong, and R. J. Cava, Weak-field induced nonmagnetic state in a Co-based honeycomb, Science Advances 6, eaay6953 (2020).
  • Zhang et al. [2023] X. Zhang, Y. Xu, T. Halloran, R. Zhong, C. Broholm, R. J. Cava, N. Drichko, and N. P. Armitage, A magnetic continuum in the cobalt-based honeycomb magnet BaCo2(AsO4)2Nature Materials 22, 58 (2023).
  • Sears et al. [2017] J. A. Sears, Y. Zhao, Z. Xu, J. W. Lynn, and Y.-J. Kim, Phase diagram of α\alpha-RuCl3 in an in-plane magnetic field, Phys. Rev. B 95, 180411 (2017).
  • Zheng et al. [2017] J. Zheng, K. Ran, T. Li, J. Wang, P. Wang, B. Liu, Z.-X. Liu, B. Normand, J. Wen, and W. Yu, Gapless spin excitations in the field-induced quantum spin liquid phase of α\alpha-RuCl3Phys. Rev. Lett. 119, 227208 (2017).
  • Banerjee et al. [2018] A. Banerjee, P. Lampen-Kelley, J. Knolle, C. Balz, A. A. Aczel, B. Winn, Y. Liu, D. Pajerowski, J. Yan, C. A. Bridges, A. T. Savici, B. C. Chakoumakos, M. D. Lumsden, D. A. Tennant, R. Moessner, D. G. Mandrus, and S. E. Nagler, Excitations in the field-induced quantum spin liquid state of α\alpha-RuCl3npj Quantum Materials 3, 8 (2018).
  • Banerjee et al. [2016] A. Banerjee, C. A. Bridges, J.-Q. Yan, A. A. Aczel, L. Li, M. B. Stone, G. E. Granroth, M. D. Lumsden, Y. Yiu, J. Knolle, S. Bhattacharjee, D. L. Kovrizhin, R. Moessner, D. A. Tennant, D. G. Mandrus, and S. E. Nagler, Proximate Kitaev quantum spin liquid behaviour in a honeycomb magnet, Nature Materials 15, 733 (2016).
  • Kasahara et al. [2018] Y. Kasahara, T. Ohnishi, Y. Mizukami, O. Tanaka, S. Ma, K. Sugii, N. Kurita, H. Tanaka, J. Nasu, Y. Motome, T. Shibauchi, and Y. Matsuda, Majorana quantization and half-integer thermal quantum Hall effect in a Kitaev spin liquid, Nature 559, 227 (2018).
  • Yokoi et al. [2021] T. Yokoi, S. Ma, Y. Kasahara, S. Kasahara, T. Shibauchi, N. Kurita, H. Tanaka, J. Nasu, Y. Motome, C. Hickey, S. Trebst, and Y. Matsuda, Half-integer quantized anomalous thermal Hall effect in the Kitaev material candidate α\alpha-RuCl3Science 373, 568 (2021).
  • Bruin et al. [2022] J. A. N. Bruin, R. R. Claus, Y. Matsumoto, N. Kurita, H. Tanaka, and H. Takagi, Robustness of the thermal Hall effect close to half-quantization in α\alpha-RuCl3Nature Physics 18, 401 (2022).
  • Vinkler-Aviv and Rosch [2018] Y. Vinkler-Aviv and A. Rosch, Approximately quantized thermal Hall effect of chiral liquids coupled to phonons, Phys. Rev. X 8, 031032 (2018).
  • Ye et al. [2018] M. Ye, G. B. Halász, L. Savary, and L. Balents, Quantization of the thermal Hall conductivity at small Hall angles, Phys. Rev. Lett. 121, 147201 (2018).
  • Czajka et al. [2021] P. Czajka, T. Gao, M. Hirschberger, P. Lampen-Kelley, A. Banerjee, J. Yan, D. G. Mandrus, S. E. Nagler, and N. P. Ong, Oscillations of the thermal conductivity in the spin-liquid state of α\alpha-RuCl3Nature Physics 17, 915 (2021).
  • Janssen and Vojta [2019] L. Janssen and M. Vojta, Heisenberg–Kitaev physics in magnetic fields, Journal of Physics: Condensed Matter 31, 423002 (2019).
  • Baek et al. [2017] S.-H. Baek, S.-H. Do, K.-Y. Choi, Y. S. Kwon, A. U. B. Wolter, S. Nishimoto, J. van den Brink, and B. Büchner, Evidence for a field-induced quantum spin liquid in α\alpha-RuCl3Phys. Rev. Lett. 119, 037201 (2017).
  • Wolter et al. [2017] A. U. B. Wolter, L. T. Corredor, L. Janssen, K. Nenkov, S. Schönecker, S.-H. Do, K.-Y. Choi, R. Albrecht, J. Hunger, T. Doert, M. Vojta, and B. Büchner, Field-induced quantum criticality in the Kitaev system α\alpha-RuCl3Phys. Rev. B 96, 041405 (2017).
  • Gass et al. [2020] S. Gass, P. M. Cônsoli, V. Kocsis, L. T. Corredor, P. Lampen-Kelley, D. G. Mandrus, S. E. Nagler, L. Janssen, M. Vojta, B. Büchner, and A. U. B. Wolter, Field-induced transitions in the Kitaev material α\alpha-RuCl3 probed by thermal expansion and magnetostriction, Phys. Rev. B 101, 245158 (2020).
  • Hentrich et al. [2018] R. Hentrich, A. U. B. Wolter, X. Zotos, W. Brenig, D. Nowak, A. Isaeva, T. Doert, A. Banerjee, P. Lampen-Kelley, D. G. Mandrus, S. E. Nagler, J. Sears, Y.-J. Kim, B. Büchner, and C. Hess, Unusual phonon heat transport in α\alpha-RuCl3: Strong spin-phonon scattering and field-induced spin gap, Phys. Rev. Lett. 120, 117204 (2018).
  • Hentrich et al. [2019] R. Hentrich, M. Roslova, A. Isaeva, T. Doert, W. Brenig, B. Büchner, and C. Hess, Large thermal Hall effect in α\alpha-RuCl3: Evidence for heat transport by Kitaev-Heisenberg paramagnons, Phys. Rev. B 99, 085136 (2019).
  • Hentrich et al. [2020] R. Hentrich, X. Hong, M. Gillig, F. Caglieris, M. Čulo, M. Shahrokhvand, U. Zeitler, M. Roslova, A. Isaeva, T. Doert, L. Janssen, M. Vojta, B. Büchner, and C. Hess, High-field thermal transport properties of the Kitaev quantum magnet α\alpha-RuCl3: Evidence for low-energy excitations beyond the critical field, Phys. Rev. B 102, 235155 (2020).
  • Balz et al. [2019] C. Balz, P. Lampen-Kelley, A. Banerjee, J. Yan, Z. Lu, X. Hu, S. M. Yadav, Y. Takano, Y. Liu, D. A. Tennant, M. D. Lumsden, D. Mandrus, and S. E. Nagler, Finite field regime for a quantum spin liquid in α\alpha-RuCl3Phys. Rev. B 100, 060405 (2019).
  • Balz et al. [2021] C. Balz, L. Janssen, P. Lampen-Kelley, A. Banerjee, Y. H. Liu, J.-Q. Yan, D. G. Mandrus, M. Vojta, and S. E. Nagler, Field-induced intermediate ordered phase and anisotropic interlayer interactions in α\alpha-RuCl3Phys. Rev. B 103, 174417 (2021).
  • Bachus et al. [2020] S. Bachus, D. A. S. Kaib, Y. Tokiwa, A. Jesche, V. Tsurkan, A. Loidl, S. M. Winter, A. A. Tsirlin, R. Valentí, and P. Gegenwart, Thermodynamic perspective on field-induced behavior of α\alpha-RuCl3Phys. Rev. Lett. 125, 097203 (2020).
  • Bachus et al. [2021] S. Bachus, D. A. S. Kaib, A. Jesche, V. Tsurkan, A. Loidl, S. M. Winter, A. A. Tsirlin, R. Valentí, and P. Gegenwart, Angle-dependent thermodynamics of α\alpha-RuCl3Phys. Rev. B 103, 054440 (2021).
  • Leahy et al. [2017] I. A. Leahy, C. A. Pocs, P. E. Siegfried, D. Graf, S.-H. Do, K.-Y. Choi, B. Normand, and M. Lee, Anomalous thermal conductivity and magnetic torque response in the honeycomb magnet α\alpha-RuCl3Phys. Rev. Lett. 118, 187203 (2017).
  • Janša et al. [2018] N. Janša, A. Zorko, M. Gomilšek, M. Pregelj, K. W. Krämer, D. Biner, A. Biffin, C. Rüegg, and M. Klanjšek, Observation of two types of fractional excitation in the Kitaev honeycomb magnet, Nature Physics 14, 786 (2018).
  • Winter et al. [2018] S. M. Winter, K. Riedl, D. Kaib, R. Coldea, and R. Valentí, Probing α\alpha-RuCl3 beyond magnetic order: Effects of temperature and magnetic field, Phys. Rev. Lett. 120, 077203 (2018).
  • Kaib et al. [2019] D. A. S. Kaib, S. M. Winter, and R. Valentí, Kitaev honeycomb models in magnetic fields: Dynamical response and dual models, Phys. Rev. B 100, 144445 (2019).
  • Chern et al. [2021] L. E. Chern, E. Z. Zhang, and Y. B. Kim, Sign structure of thermal Hall conductivity and topological magnons for in-plane field polarized Kitaev magnets, Phys. Rev. Lett. 126, 147201 (2021).
  • Winter et al. [2017] S. M. Winter, K. Riedl, P. A. Maksimov, A. L. Chernyshev, A. Honecker, and R. Valentí, Breakdown of magnons in a strongly spin-orbital coupled magnet, Nature Communications 8, 1152 (2017).
  • Cookmeyer and Moore [2018] T. Cookmeyer and J. E. Moore, Spin-wave analysis of the low-temperature thermal Hall effect in the candidate Kitaev spin liquid α\alpha-RuCl3Phys. Rev. B 98, 060412 (2018).
  • Kim and Kee [2016] H.-S. Kim and H.-Y. Kee, Crystal structure and magnetism in α\alpha-RuCl3: An ab initio study, Phys. Rev. B 93, 155143 (2016).
  • Suzuki and Suga [2018] T. Suzuki and S.-i. Suga, Effective model with strong Kitaev interactions for α\alpha-RuCl3Phys. Rev. B 97, 134424 (2018).
  • Ran et al. [2017] K. Ran, J. Wang, W. Wang, Z.-Y. Dong, X. Ren, S. Bao, S. Li, Z. Ma, Y. Gan, Y. Zhang, J. T. Park, G. Deng, S. Danilkin, S.-L. Yu, J.-X. Li, and J. Wen, Spin-wave excitations evidencing the Kitaev interaction in single crystalline α\alpha-RuCl3Phys. Rev. Lett. 118, 107203 (2017).
  • Ozel et al. [2019] I. O. Ozel, C. A. Belvin, E. Baldini, I. Kimchi, S. Do, K.-Y. Choi, and N. Gedik, Magnetic field-dependent low-energy magnon dynamics in α\alpha-RuCl3Phys. Rev. B 100, 085108 (2019).
  • Kim et al. [2015] H.-S. Kim, V. S. V., A. Catuneanu, and H.-Y. Kee, Kitaev magnetism in honeycomb RuCl3 with intermediate spin-orbit coupling, Phys. Rev. B 91, 241110 (2015).
  • Zhu et al. [2018] Z. Zhu, I. Kimchi, D. N. Sheng, and L. Fu, Robust non-abelian spin liquid and a possible intermediate phase in the antiferromagnetic Kitaev model with magnetic field, Phys. Rev. B 97, 241110 (2018).
  • Gohlke et al. [2018a] M. Gohlke, R. Moessner, and F. Pollmann, Dynamical and topological properties of the Kitaev model in a [111] magnetic field, Phys. Rev. B 98, 014418 (2018a).
  • Jiang et al. [2018] H.-C. Jiang, C.-Y. Wang, B. Huang, and Y.-M. Lu, Field induced quantum spin liquid with spinon Fermi surfaces in the Kitaev model, arXiv:1809.08247  (2018).
  • Patel and Trivedi [2019] N. D. Patel and N. Trivedi, Magnetic field-induced intermediate quantum spin liquid with a spinon Fermi surface, Proceedings of the National Academy of Sciences 116, 12199 (2019).
  • Hickey and Trebst [2019] C. Hickey and S. Trebst, Emergence of a field-driven U(1) spin liquid in the Kitaev honeycomb model, Nature Communications 10, 530 (2019).
  • Ronquillo et al. [2019] D. C. Ronquillo, A. Vengal, and N. Trivedi, Signatures of magnetic-field-driven quantum phase transitions in the entanglement entropy and spin dynamics of the Kitaev honeycomb model, Phys. Rev. B 99, 140413 (2019).
  • Zou and He [2020] L. Zou and Y.-C. He, Field-induced QCD3-Chern-Simons quantum criticalities in Kitaev materials, Phys. Rev. Res. 2, 013072 (2020).
  • Gohlke et al. [2018b] M. Gohlke, G. Wachtel, Y. Yamaji, F. Pollmann, and Y. B. Kim, Quantum spin liquid signatures in Kitaev-like frustrated magnets, Phys. Rev. B 97, 075126 (2018b).
  • Liu and Normand [2018] Z.-X. Liu and B. Normand, Dirac and chiral quantum spin liquids on the honeycomb lattice in a magnetic field, Phys. Rev. Lett. 120, 187201 (2018).
  • Yamada and Fujimoto [2021] M. G. Yamada and S. Fujimoto, Quantum liquid crystals in the finite-field K-Γ\Gamma model for α\alpha-RuCl3arXiv:2107.03045  (2021).
  • Wang et al. [2020] J. Wang, Q. Zhao, X. Wang, and Z.-X. Liu, Multinode quantum spin liquids on the honeycomb lattice, Phys. Rev. B 102, 144427 (2020).
  • Gordon et al. [2019] J. S. Gordon, A. Catuneanu, E. S. Sørensen, and H.-Y. Kee, Theory of the field-revealed Kitaev spin liquid, Nature Communications 10, 2470 (2019).
  • Lee et al. [2020] H.-Y. Lee, R. Kaneko, L. E. Chern, T. Okubo, Y. Yamaji, N. Kawashima, and Y. B. Kim, Magnetic field induced quantum phases in a tensor network study of Kitaev magnets, Nature Communications 11, 1639 (2020).
  • Gohlke et al. [2020] M. Gohlke, L. E. Chern, H.-Y. Kee, and Y. B. Kim, Emergence of nematic paramagnet via quantum order-by-disorder and pseudo-Goldstone modes in Kitaev magnets, Phys. Rev. Research 2, 043023 (2020).
  • Li et al. [2021] H. Li, H.-K. Zhang, J. Wang, H.-Q. Wu, Y. Gao, D.-W. Qu, Z.-X. Liu, S.-S. Gong, and W. Li, Identification of magnetic interactions and high-field quantum spin liquid in α\alpha-RuCl3Nature Communications 12, 4007 (2021).
  • [69] X.-G. Zhou, H. Li, Y. H. Matsuda, A. Matsuo, W. Li, N. Kurita, K. Kindo, and H. Tanaka, Intermediate Quantum Spin Liquid Phase in the Kitaev Material α\alpha-RuCl3 under High Magnetic Fields up to 100 T,  arXiv:2201.04597 (2022) .
  • White [1992] S. R. White, Density matrix formulation for quantum renormalization groups, Phys. Rev. Lett. 69, 2863 (1992).
  • McCulloch [2008] I. P. McCulloch, Infinite size density matrix renormalization group, revisited (2008).
  • Hauschild and Pollmann [2018] J. Hauschild and F. Pollmann, Efficient numerical simulations with tensor networks: Tensor network python (tenpy), SciPost Phys. Lect. Notes , 5 (2018).
  • Chen et al. [2018] B.-B. Chen, L. Chen, Z. Chen, W. Li, and A. Weichselbaum, Exponential thermal tensor network approach for quantum lattice models, Phys. Rev. X 8, 031082 (2018).
  • [74] Q. Li, Y. Gao, Y.-Y. He, Y. Qi, B.-B. Chen, and W. Li, Tangent space approach for thermal tensor network simulations of 2D Hubbard model, arXiv: 2212.11973 (2022) .
  • Gong et al. [2014] S.-S. Gong, W. Zhu, and D. N. Sheng, Emergent chiral spin liquid: Fractional quantum Hall effect in a kagome Heisenberg model, Scientific Reports 4, 6317 (2014).
  • Luo et al. [2022] Q. Luo, P. P. Stavropoulos, J. S. Gordon, and H.-Y. Kee, Spontaneous chiral-spin ordering in spin-orbit coupled honeycomb magnets, Phys. Rev. Research 4, 013062 (2022).
  • Luo and Kee [2022] Q. Luo and H.-Y. Kee, Interplay of magnetic field and trigonal distortion in the honeycomb Γ\mathrm{\Gamma} model: Occurrence of a spin-flop phase, Phys. Rev. B 105, 174435 (2022).
  • Sørensen et al. [2021] E. S. Sørensen, A. Catuneanu, J. S. Gordon, and H.-Y. Kee, Heart of entanglement: Chiral, nematic, and incommensurate phases in the Kitaev-Γ\Gamma ladder in a field, Phys. Rev. X 11Phys. Rev. X 11.011013 (2021).
  • Hu et al. [2019] W.-J. Hu, S.-S. Gong, H.-H. Lai, H. Hu, Q. Si, and A. H. Nevidomskyy, Nematic spin liquid phase in a frustrated spin-1 system on the square lattice, Phys. Rev. B 100, 165142 (2019).
  • Hu et al. [2020a] W.-J. Hu, S.-S. Gong, H.-H. Lai, Q. Si, and E. Dagotto, Density matrix renormalization group study of nematicity in two dimensions: Application to a spin-1 bilinear-biquadratic model on the square lattice, Phys. Rev. B 101, 014421 (2020a).
  • Hu et al. [2020b] W.-J. Hu, H.-H. Lai, S.-S. Gong, R. Yu, E. Dagotto, and Q. Si, Quantum transitions of nematic phases in a spin-1 bilinear-biquadratic model and their implications for FeSe, Phys. Rev. Research 2, 023359 (2020b).
  • Calabrese and Cardy [2004] P. Calabrese and J. Cardy, Entanglement entropy and quantum field theory: A non-technical introduction, Journal of Statistical Mechanics: Theory and Experiment 2004, P06002 (2004).
  • Wang et al. [2019] J. Wang, B. Normand, and Z.-X. Liu, One proximate Kitaev spin liquid in the KK-JJ-Γ\Gamma model on the honeycomb lattice, Phys. Rev. Lett. 123, 197201 (2019).
  • Wen [2002] X.-G. Wen, Quantum orders and symmetric spin liquids, Phys. Rev. B 65, 165113 (2002).
  • Song et al. [2016] X.-Y. Song, Y.-Z. You, and L. Balents, Low-energy spin dynamics of the honeycomb spin liquid beyond the Kitaev limit, Physical Review Letters 117Phys. Rev. Lett. 117.037209 (2016).
  • You et al. [2012] Y.-Z. You, I. Kimchi, and A. Vishwanath, Doping a spin-orbit Mott insulator: Topological superconductivity from the Kitaev-Heisenberg model and possible application to (Na2/Li2)IrO3Phys. Rev. B 86, 085145 (2012).
  • Meier et al. [2011] B. Meier, S. Greiser, J. Haase, T. Herrmannsdörfer, F. Wolff-Fabris, and J. Wosnitza, NMR signal averaging in 62 T pulsed fields, Journal of Magnetic Resonance 210, 1 (2011).
  • Akaki et al. [2018] M. Akaki, K. Kanaida, and M. Hagiwara, Magnetoelectric properties of Ca2CoSi2O7 studied by high-field electron spin resonance, Journal of Physics: Conference Series 969, 012102 (2018).
  • Yao and Li [2020] W. Yao and Y. Li, Ferrimagnetism and anisotropic phase tunability by magnetic fields in na2co2teo6{\mathrm{na}}_{2}{\mathrm{co}}_{2}{\mathrm{teo}}_{6}Phys. Rev. B 101, 085120 (2020).