Global phase diagram of a spin–orbit-coupled Kondo lattice model on the honeycomb lattice
Li Xin1, 2, Yu Rong3, †, Si Qimiao4
Beijing National Laboratory for Condensed Matter Physics and Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
Department of Physics and Beijing Key Laboratory of Opto-electronic Functional Materials and Micro-nano Devices, Renmin University of China, Beijing 100872, China
Department of Physics & Astronomy, Rice Center for Quantum Materials, Rice University, Houston, Texas 77005, USA

 

† Corresponding author. E-mail: rong.yu@ruc.edu.cn

Abstract

Motivated by the growing interest in the novel quantum phases in materials with strong electron correlations and spin–orbit coupling, we study the interplay among the spin–orbit coupling, Kondo interaction, and magnetic frustration of a Kondo lattice model on a two-dimensional honeycomb lattice. We calculate the renormalized electronic structure and correlation functions at the saddle point based on a fermionic representation of the spin operators. We find a global phase diagram of the model at half-filling, which contains a variety of phases due to the competing interactions. In addition to a Kondo insulator, there is a topological insulator with valence bond solid correlations in the spin sector, and two antiferromagnetic phases. Due to the competition between the spin–orbit coupling and Kondo interaction, the direction of the magnetic moments in the antiferromagnetic phases can be either within or perpendicular to the lattice plane. The latter antiferromagnetic state is topologically nontrivial for moderate and strong spin–orbit couplings.

1. Introduction

Exploring novel quantum phases and the associated phase transitions in systems with strong electron correlations is a major subject of contemporary condensed matter physics.[13] In this context, heavy fermion (HF) compounds play a crucial role.[37] In these materials, the coexisted itinerant electrons and local magnetic moments (from localized f electrons) interact via the antiferromagnetic exchange coupling, resulting in the famous Kondo effect.[8] Meanwhile, the Ruderman–Kittel–Kasuya–Yosida (RKKY) interaction, namely the exchange coupling among the local moments mediated by the itinerant electrons, competes with the Kondo effect.[9] This competition gives rise to a rich phase diagram with an antiferromagnetic (AFM) quantum critical point (QCP) and various emergent phases nearby.[3,10]

In the HF metals, experiments[11,12] have provided strong evidence for local quantum criticality,[13,14] which is characterized by the beyond-Landau physics of Kondo destruction at the AFM QCP. Across this local QCP, the Fermi surface jumps from large in the paramagnetic HF liquid phase to small in the AFM phase with Kondo destruction. A natural question arises: how does this local QCP connect to the conventional spin density wave (SDW) QCP described by the Hertz–Millis theory?[15,16] A proposed global phase diagram[1720] makes this connection via the tuning of the quantum fluctuations in the local-moment magnetism. Besides the HF metals, it is also interesting to know whether a similar global phase diagram can be realized in Kondo insulators (KIs), where the chemical potential is inside the Kondo hybridization gap when the electron filling is commensurate. The KIs are band insulators where the band gap originates from nontrivial strong electron-correlation effects. A Kondo-destruction transition is expected to accompany the closure of the band gap. The question that remains open is: do the local moments immediately order or do they form a different type of magnetic state, such as spin liquid or valence bond solid (VBS), when the Kondo-destruction takes place?

Recent years have seen extensive studies about the effect of a fine spin–orbit coupling (SOC) on the electronic bands. In topological insulators (TIs), the bulk band gap opens due to a non-zero SOC, and there exist gapless surface states. The nontrivial topology of the band structure is protected by the time reversal symmetry (TRS). Even for a system with broken TRS, the conservation of combination of TRS and translational symmetry can give rise to a topological antiferromagnetic insulator (T-AFMI).[21] In general, these TIs and TAFIs can be tuned to topologically trivial insulators via topological quantum phase transitions. But how the strong electron correlations influence the properties of these symmetry dictated topological phases and the related phase transitions is still under active discussion.

The SOC also has important effects in HF materials.[20] For example, the SOC can produce a topologically nontrivial band structure and induce exotic Kondo physics.[22,23] It may give rise to a topological Kondo insulator (TKI),[24] which has been invoked to understand the resistivity plateau of the heavy-fermion SmB6 at low temperatures.[25]

From a more general perspective, the SOC provides an additional tuning parameter enriching the global phase diagram of the HF systems.[20,26] Whether and how the topological nontrivial quantum phases can emerge in this phase diagram is a timely issue. Recent studies have advanced a Weyl–Kondo semimetal phase.[27] The new heavy fermion compound Ce3Bi4Pd3 displays thermodynamic[28] and zero-field Hall transport[29] properties that provide evidence for the salient features of the Weyl–Kondo semimetal. These measurements probe the linearly dispersing electronic excitations with a velocity that is renormalized by several orders of magnitude and singularities in the Berry-curvature distribution.

Theoretical studies are also of interest for a Kondo lattice model defined on a honeycomb lattice,[30] which readily accommodates the SOC.[31] In the diluted carrier limit, this model supports a nontrivial Dirac–Kondo semimetal (DKSM) phase, which can be tuned to a TKI by increasing the SOC.[32] In Ref. [30], it was shown that, at half-filling, increasing the Kondo coupling induces a direct transition from a TI to a KI. A related model, with the conduction-electron part of the Hamiltonian described by a Haldane model[33] on the honeycomb lattice, was subsequently studied.[34]

Here we investigate the global phase diagram of a spin–orbit-coupled Kondo lattice model on the honeycomb lattice at half-filling. We show that the competing interactions in this model give rise to a very rich phase diagram containing a TI, a KI, and two AFM phases. We focus on discussing the influence of magnetic frustration on the phase diagram. In the TI, the local moments develop a VBS order. In the two AFM phases, the moments are ordered, respectively, in the plane of the honeycomb lattice (denoted as AFMxy) and perpendicular to the plane (AFMz). Particularly in the AFMz phase, the conduction electrons may have a topologically nontrivial band structure, although the TRS is explicitly broken. This T-AFMz state connects to the trivial AFMz phase via a topological phase transition as the SOC is reduced.

The remainder of the paper is organized as follows. We start by introducing the model and our theoretical procedure in Section 2. In Section 3, we discuss the magnetic phase diagram of the Heisenberg model for the local moments. Next, we obtain the global phase diagram of the full model in Section 4. In Section 5, we examine the nature of the conduction-electron band structures in the AFM states, with a focus on their topological characters. We discuss the implications of our results in Section 6.

2. Model and method

The model that we consider here is defined on an effective double-layer honeycomb lattice. The top layer contains conduction electrons realizing the Kane–Mele Hamiltonian.[31] The conduction electrons are Kondo coupled (i.e., via AF exchange coupling JK) to the localized magnetic moments in the bottom layer. The local moments interact among themselves through direct exchange interaction, as well as the conduction electron mediated RKKY interaction; the interactions are described by a simple J1J2 model. Both the conduction bands and the localized bands are half-filled. This Kondo-lattice Hamiltonian takes the following form on the honeycomb lattice:

where creates a conduction electron at site i with spin index σ. t is the hopping parameter between the nearest neighboring (NN) sites, and is the strength of the SOC between the next-nearest neighboring (NNN) sites. , depending on the direction of the NNN hopping. is the spin operator of the conduction electrons at site i with being the Pauli matrices. refers to the spin operator of the local moments with spin S = 1/2. In the model we considered here, JK, J1, and J2 are all AFM. By incorporating the Heisenberg interactions, the studied Kondo-lattice model readily captures the effect of geometrical frustration. In addition, instead of treating the Kondo screening and magnetic order in terms of the longitudinal and transverse components of the Kondo-exchange interactions,[3436] we will treat both effects in terms of interactions that are spin-rotationally invariant; this will turn out to be important in mapping out the global phase diagram.

We use the spinon representation for , i.e., rewriting along with the constraint , where is the spinon operator. The constraint is enforced by introducing the Lagrange multiplier term in the Hamiltonian. To study both the non-magnetic and magnetic phases, We decouple the Heisenberg Hamiltonian into two channels

Here x is a parameter that is introduced in keeping with the generalized procedure of Hubbard–Stratonovich decouplings and will be fixed to conveniently describe the effect of quantum fluctuations. The corresponding valence bond (VB) parameter Qij and sublattice magnetization are and , respectively. Throughout this paper, we consider the two-site unit cell, thus excluding any states that break the lattice translation symmetry. Under this construction, there are three independent VB mean fields Qi, i = 1,2,3 for the NN bonds and six independent VB mean fields Qi, for the NNN bonds, which are illustrated in Fig. 1. We consider only AFM exchange interactions, and , and will thus only take into account AFM order with .

Fig. 1. (a), (b) Definition of nearest neighboring and next nearest neighboring valence bond mean fields Qij. Filled and empty circles denote the two sublattices A and B, respectively. Different bond directions are labeled by different colors. (c) First Brilluion zone corresponds to the two-sublattice unit cell.

To take into account the Kondo hybridization and the possible magnetic order on an equal footing, we follow the treatment of the Heisenberg interaction as outlined in Eq. (2) and decouple the Kondo interaction as follows:

Here we have introduced the mean-field parameter for the Kondo hybridization, , and the conduction electron magnetization . For nonzero b, the conduction electrons will Kondo hybridize with the local moments and the system at half-filling is a KI. On the other hand, when b is zero and is nonzero, magnetization ( ) on the conduction electron band will be induced by the Kondo coupling, and various AFM orders can be stabilized depending on the strength of the SOC. Just like the parameter x in Eq. (2) is chosen so that a saddle-point treatment captures the quantum fluctuations in the form of spin-singlet bond parameters,[19] the parameter y will be specified according to the criterion that the treatment at the same level describes the quantum fluctuations in the form of Kondo-insulator state.

3. Phase diagram of the Heisenberg model for the local moments

Because of the complexity of the full Hamiltonian, we start by setting JK = 0 and discuss the possible ground-state phases of the J1J2 Heisenberg model for the local moments. By treating the problem at the saddle-point level in Eq. (2), we obtain the phase diagram in the x plane shown in Fig. 2. Here, the x-dependence is studied in the same spirit as that for the Shastry–Sutherland lattice in Ref. [19]. In the parameter regime explored, an AFM ordered phase (labeled as “AFM” in the figure) and a valence bond solid (VBS) phase are stabilized. The AF order stabilized is the two-sublattice Néel order on the honeycomb lattice, and the VBS order refers to covering of dimer singlets with for one out of the three NN bonds (e.g., ) and for all the NNN bonds. This VBS state spontaneously breaks the C3 rotational symmetry of the lattice. We thus define the order parameter for the VBS state to be .

Fig. 2. Ground-state phase diagram of the J1J2 Hamiltonian for the local moments in the x plane. A NN VBS and an AFM state are stabilized in the parameter regime shown.

In Fig. 3, we plot the evolution of VBS and AF order parameters Q and M as a function of . A direct first-order transition (signaled by the mid-point of the jump of the order parameters) between these two phases is observed for . For the sake of understanding the global phase diagram of the full Kondo–Heisenberg model, we limit our discussion to , where only the NN VBS is relevant. A different decoupling scheme approach has been used to study this model,[37] and the obtained results are consistent with ours in the parameter regime of overlap. To fix the parameter x, we compare our results with those about the J1J2 model derived from previous numerical studies. DMRG studies[38] found that the AFM state is stabilized for , and VBS exists for , while in between the nature of the ground states is still under debate. In this parameter regime, the DMRG calculations suggest a plaquette resonating valence bond (RVB) state,[38] while other methods implicate possibly spin liquids.[39] In light of these numerical results, we take x = 0.4 in our calculations. This leads to a direct transition from AFM to VBS at , close to the values of phase boundaries of these two phases determined by other numerical methods.

Fig. 3. Evolution of (a) the VBS order parameter Q and (b) the AFM order parameter M as a function of for x = 0.3, 0.4, 0.5.
4. Global phase diagram of the Kondo-lattice model

We now turn to the global phase diagram of the full model by turning on the Kondo coupling. For definiteness, we set J1 = 1 and consider t = 1 and . As prescribed in the previous section, we take x = 0.4. Similar considerations for y require that its value allows for quantum fluctuations in the form of Kondo-singlet formation. This has guided us to take y = 0.7 (see below). The corresponding phase diagram as a function of JK and the frustration parameter is shown in Fig. 4.

Fig. 4. Global phase diagram at T = 0 from the saddle-point calculations with x = 0.4 and y = 0.7. The ground states include the valence-bond solid (VBS) and Kondo insulator (KI), as well as two antiferromagnetic orders, T-AFMz and AFMxy, as described in Section 5.

In our calculation, the phase boundaries are determined by sweeping JK while along multiple horizontal cuts for several fixed J2/J1 values, as shown in Fig. 5. For small JK and large J2/J1, the local moments and the conduction electrons are still effectively decoupled. The conduction electrons form a TI for finite SOC, and the local moments are in the VBS ground state as discussed in the previous section. When both JK and J2/J1 are small, the ground state is AFM. Due to the Kondo coupling, finite magnetization is induced for the conduction electrons. This opens an SDW gap in the conduction band, and therefore the ground state of the system is an AFM insulator. The SOC couples the rotational symmetry in the spin space to the one in the real space. As a consequence, the ordered moments in the AFM phase can be either along the z direction (AFMz) or in the xy plane (AFMxy). For finite SOC, these two AFM states with different energies, which can be tuned by JK. As shown in the phase diagram, the AFM phase contains two ordered states, the AFMz and AFMxy. They are separated by a spin reorientation transition at . For the value of SOC taken, the AFM state is topologically nontrivial, and is hence denoted as T-AFMz state. The nature of this state and the associated topological phase transition is discussed in detail in the next section.

Fig. 5. Evolution of parameters (a) b, (b) Q, (c) Mx, and (d) Mz as a function of JK for different ratios of J2/J1.

For sufficiently large JK, the Kondo hybridization b is nonzero (see Fig. 5(a)), and the ground state is a KI. Note that for finite SOC, this KI does not have a topological nontrivial edge state, as a consequence of the topological no-go theorem.[30,40,41] In our calculation at the saddle-point level, the KI exists for ; this provides the basis for taking y = 0.7, as noted earlier. Going beyond the saddle-point level, the dynamical effects of the Kondo coupling will appear, and we will expect the KI phase to arise for other choices of y.

Several remarks are in order. The phase diagram, Fig. 4, has a similar profile of the global phase diagram for the Kondo insulating systems.[26,42] However, the presence of SOC has enriched the phase diagram. In the AF state, the ordered moment may lie either within the plane or be perpendicular to it. These two states have very different topological properties. We now turn to a detailed discussion of this last point.

5. Topological properties of the AFM states

In this section, we discuss the properties of the AFMxy and AFMz states, in particular to address their topological nature. For a clear discussion, we fix t = 1, J1 = 1, and J2 = 0. Since the Kondo hybridization is not essential to the nature of the AFM states, in this section we simply the discussion by setting y = 0.

We start by defining the order parameters of the two states

Note that for the AFMxy state, we set without losing generality. In Fig. 6, we plot the evolution of these AFM order parameters with JK for a representative value of SOC . Due to the large J1 value we take, the sublattice magnetizations of the local moments are already saturated to 0.5. Therefore, at the saddle-point level, they serve as effective (staggered) magnetic fields to the conduction electrons. The Kondo coupling then induces finite sublattice magnetizations for the conduction electrons, and they increase linearly with JK for small JK values. But mx is generically different from mz, which is important for the stabilization of the states.

Fig. 6. The conduction electron magnetization for the AFMxy and AFMz states at .

We then discuss the energy competition between the AFMxy and AFMz states. The conduction electron part of the mean-field Hamiltonian reads

with
for the AFMxy state and
for the AFMz state. Here
, is the complex conjugate of , and , are the primitive vectors. For both states, the eigenvalues are doubly degenerate

The eigenenergies of the spinon band can be obtained in a similar way

The expression of the total energy for either state is then
The first line of the above expression comes from filling the bands up to the Fermi energy (which is fixed to be zero here). The second line is the constant term in the mean-field decomposition. The factor 2 in the summation is to take into account the double degeneracy of the energies. Nk refers to the number of points in the first Brillouin zone.

By comparing the expressions of in Eqs. (11) and (12), we find that adding a small Mx is to increase the size of the gap at both of the two (inequivalent) Dirac points, thereby pushing the states further away from the Fermi-energy. While adding a small Mz is to enlarge the gap at one Dirac point but reduce the gap size at the other one. Therefore, an AFMxy state is more favorable than the AFMz state in lowering the energy of the conduction electrons .

Meanwhile, from Eqs. (13)–(15), we see that the overall effect of adding a magnetization of the conduction band, , is to increase the total energy (the main energy increase comes from the term). Because from the self consistent solution, as shown in Fig. 6, the energy increase of the AFMz state is smaller than that in the AFMxy state.

With increasing JK, the two effects from the magnetic orders compete, resulting in different magnetic ground states as shown in Fig. 4. This analysis is further supported by our self-consistent mean-field calculation. In Fig. 7, we plot the energy difference between these two states as a function of JK at several values. In the absence of SOC, the model has the spin SU(2) symmetry, and the AFMz and AFMxy states are degenerate with . For finite , at small JK values, the energy gain from the term dominates, , and the ground state is an AFMz state. With increasing JK, the contribution from the term is more important. crosses zero to be negative, and the AFMxy state is eventually energetically favorable for large JK.

Fig. 7. Energy difference between AFMz and AFMxy states as a function of JK for various values of spin–orbital coupling .

Next we discuss the topological nature of the AFMz and AFMxy states. In the absence of Kondo coupling JK, the conduction electrons form a TI, which is protected by the TRS. There, the left- and right-moving edge states connecting the conduction and valence bands are respectively coupled to up and down spin flavors (eigenstates of the Sz operator) as the consequence of SOC, and these two spin polarized edge states do not mix.

Once the TRS is broken by the AFM order, generically, topologically nontrivial edge states are no longer guaranteed. However, in the AFMz state, the structure of the Hamiltonian for the conduction electrons is the same as that in a TI. This is clearly shown in Eq. (10) the effect of magnetic order is only to shift to . In particular, the spin-up and spin-down sectors still do not mix with each other. Therefore, the two spin polarized edge states are still well defined as in the TI, and the system is topologically nontrivial, though without the protection of TRS. Note that the above analysis is based on assuming , where the bulk gap between the conduction and valence bands is finite. For , the bulk gap closes at one of the inequivalent Dirac points and the system is driven to a topologically trivial phase via a topological phase transition.[30] We also note that a similar AFMz state arises in a Kondo lattice model without SOC but with a Haldane coupling, as analyzed in Ref. [34].

For the AFMxy state, we can examine the Hamiltonian for the conduction electrons in a similar way. As shown in Eq. (9), the transverse magnetic order Mx mixes the spin-up and spin-down sectors. As a result, a finite hybridization gap opens between the two edge states, making the system topologically trivial.

To support this analysis, we perform calculations of the energy spectra of the conduction electrons in the AFMz and AFMxy states, as shown in Eqs. (9) and (10), on a finite slab of size , with and . The boundary condition is chosen to be periodic along the x direction and open and zig-zag-type along the y direction. In Fig. 8, we show the energy spectra with three different sets of parameters: (a) , JK = 0.4, , (b) , JK = 0.4, , and (c) , JK = 0.8, , which respectively correspond to the topologically trivial AFMz state, topological AFMz insulator, and AFMxy state. As clearly seen, the gapless edge states only exist for parameter set (b), where the system is in the topological AFMz state. Note that in this state, the spectrum is asymmetric with respect to the Brilluion zone boundary ( ), reflecting the explicit breaking of TRS. Based on our analysis and numerical calculations, we construct a phase diagram (as shown in Fig. 9) to illustrate the competition of these AFM states. As expected, the AFMz state is stabilized for , and is topological for (above the red line).

Fig. 8. Energy spectra of (a) the trivial AFMz state, (b) the topological AFMz insulator, and (c) the AFMxy state from finite slab calculations. Black lines denote the bulk states and red lines denote the edge states. The topological AFMz state is characterized by the gapless edge states. See text for detailed information on the parameters.
Fig. 9. Phase diagram in the JK plane showing the competition of various AFM states. The red line denotes a topological phase transition between the topological trivial and topological nontrivial AFMz states, and the black curve gives the boundary between the AFMz and AFMxy states. These two states become equivalent in the limit of .
6. Discussion

We have discussed the properties of various phases in the ground-state phase diagram of the spin–orbit-coupled Kondo lattice model on the honeycomb lattice at half filling. We have shown how the competition of SOC, Kondo interaction, and magnetic frustration stabilizes these phases. For example, in the AFM phase the moments can order either along the z-direction or within the xy plane. In our model, the AFM order is driven by the RKKY interaction, and the competition of SOC and Kondo interaction dictates the direction of the ordered magnetic moments.

Throughout this work, we have discussed the phase diagram of the model at half filling. The phase diagram away from half-filling is also an interesting problem. We expect that the competition between the AFMz and AFMxy states persist at generic fillings, but the topological feature will not. Another interesting filling would be the dilute-carrier limit, where a DKSM exists, and can be tuned to a TKI by increasing the SOC.[32]

In this work, we have considered a particular type of SOC, which is inherent in the band structure of the itinerant electrons. In real materials, there are also SOC terms that involve the magnetic ions. Such couplings will lead to models beyond the current work, and may further enrich the global phase diagram.

Although the model in this work is defined on the honeycomb lattice, our conclusion on the global phase diagram is quite general, and will be important in understanding the nature of the transition between the Kondo insulating phase and the antiferromagnetic phase in real materials. For example, the Kondo insulator compound SmB6 undergoes a magnetic transition under pressure.[44] The Kondo-insulator-to-antiferromagnet transition may also be realized by doping CeNiSn with Pt or Pd ions since both CePtSn and CePdSn are antiferromagnetic at low temperatures.[4446] Nontrivial topological properties in the antiferromagnetic phase are expected given that the 5d electrons of CePtSn may contain a large SOC. Moreover, because the Kane–Mele model describes the electron states in graphene, our model may also shed light on the properties of graphene with 5d adatoms.[47]

7. Conclusion

We have investigated the ground state phase diagram of a spin–orbit coupled Kondo lattice model at half-filling. The combination of SOC, Kondo and RKKY interactions produces various quantum phases, including a Kondo insulator, a topological insulator with VBS spin correlations, and two AFM phases. Depending on the strength of SOC, the magnetic moments in the AFM phase can be either ordered perpendicular to or in the xy plane. We further show that the z-AFM state is topologically nontrivial for strong and moderate SOC, and can be tuned to a topologically trivial one via a topological phase transition by varying either the SOC or the Kondo coupling. Our results shed new light on the global phase diagram of heavy fermion materials.

Reference
[1] Löhneysen H 2010 J. Low Temp. Phys. 161 1 https://doi.org/10.1007/s10909-010-0219-y
[2] Sachdev S 1999 Quantum Phase Transitions New York Cambridge University Press
[3] Si Q Steglich F 2010 Science 329 1161 http://dx.doi.org/10.1126/science.1191195
[4] Gegenwart P Si Q Steglich F 2008 Nat. Phys. 4 186 http://dx.doi.org/10.1038/nphys892
[5] Löhneysen H von Rosch A Vojta M Wolfle P 2007 Rev. Mod. Phys. 79 1015 http://dx.doi.org/10.1103/RevModPhys.79.1015
[6] Tsunetsugu H Sigrist M Ueda K 1997 Rev. Mod. Phys. 69 809 http://dx.doi.org/10.1103/RevModPhys.69.809
[7] Yang Y F Yu L 2015 Acta Phys. Sin. 64 217401 (in Chinese) http://dx.doi.org/10.7498/aps.64.217401
[8] Hewson A C 1993 The Kondo Problem to Heavy Fermions Cambridge Cambridge University Press
[9] Doniach S 1977 Physica B+C 91 231 http://dx.doi.org/10.1016/0378-4363(77)90190-5
[10] Custers J Gegenwart P Wilhelm H Neumaier K Tokiwa Y Trovarelli O Geibel C Steglich F Pépin C Coleman P 2003 Nature 424 524 http://dx.doi.org/10.1038/nature01774
[11] Schröder A Aeppli G Coldea R Adams M Stockert O L?hneysen H v Bucher E Ramazashvili R Coleman P 2000 Nature 407 351 http://dx.doi.org/10.1038/35030039
[12] Paschen S Luhmann T Wirth S Gegenwart P Trovarelli O Geibel C Steglich F Coleman P Si Q 2004 Nature 432 881 http://dx.doi.org/10.1038/nature03129
[13] Si Q Rabello S Ingersent K Smith J L 2001 Nature 413 804 http://dx.doi.org/10.1038/35101507
[14] Coleman P Pépin C Si Q Ramazashvili R 2001 J. Phys.: Condens. Matter 13 R723 http://dx.doi.org/10.1088/0953-8984/13/35/202
[15] Hertz J A 1976 Phys. Rev. B 14 1165 http://dx.doi.org/10.1103/PhysRevB.14.1165
[16] Millis A J 1993 Phys. Rev. B 48 7183 http://dx.doi.org/10.1103/PhysRevB.48.7183
[17] Si Q 2006 Physica B 378 23
[18] Si Q 2010 Phys. Stat. Solid B 247 476 http://dx.doi.org/10.1002/pssb.v247:3
[19] Pixley J H Yu R Si Q 2014 Phys. Rev. Lett. 113 176402 http://dx.doi.org/10.1103/PhysRevLett.113.176402
[20] Si Q Paschen S 2013 Phys. Stat. Solid (b) 250 425 http://dx.doi.org/10.1002/pssb.v250.3
[21] Mong R S K Essin A M Moore J E 2010 Phys. Rev. B 81 245209 http://dx.doi.org/10.1103/PhysRevB.81.245209
[22] Nakatsuji S Machida Y Maeno Y Tayama T Sakakibara T Duijn J v Balicas L Millican J N Macaluso R T Chan J Y 2006 Phys. Rev. Lett. 96 087204 http://dx.doi.org/10.1103/PhysRevLett.96.087204
[23] Chen G 2017 Phys. Rev. B 94 205107 https://doi.org/10.1103/PhysRevB.94.205107
[24] Dzero M Sun K Galitski V Coleman P 2010 Phys. Rev. Lett. 104 106408 http://dx.doi.org/10.1103/PhysRevLett.104.106408
[25] Barla A Derr J Sanchez J P Salce B Lapertot G Doyle B P Rüffer R Lengsdorf R Abd-Elmeguid M M Flouquet J 2005 Phys. Rev. Lett. 94 166401 http://dx.doi.org/10.1103/PhysRevLett.94.166401
[26] Yamamoto S J Si Q 2010 J. Low Temp. Phys. 161 233 http://dx.doi.org/10.1007/s10909-010-0221-4
[27] Lai H H Grefe S E Paschen S Si Q 2018 Pro. Natl. Acad Sci. USA 115 93 http://dx.doi.org/10.1073/pnas.1715851115
[28] Dzsaber S Prochaska L Sidorenko A Eguchi G Svagera R Waas M Prokofiev A Si Q Paschen S 2017 Phys. Rev. Lett. 118 246601 http://dx.doi.org/10.1103/PhysRevLett.118.246601
[29] Dzsaber S Yan X Eguchi G Prokofiev A Shiroka T Blaha P Rubel O Grefe S E Lai H H Si Q Paschen S 2018 arXiv:1811.02819
[30] Feng X Y Chung C H Dai J Si Q 2013 Phys. Rev. Lett. 111 016402 http://dx.doi.org/10.1103/PhysRevLett.111.016402
[31] Kane C L Mele E J 2005 Phys. Rev. Lett. 95 226801 http://dx.doi.org/10.1103/PhysRevLett.95.226801
[32] Feng X Y Zhong H Dai J Si Q 2016 arXiv:1605.02380
[33] Haldane F D M 1988 Phys. Rev. Lett. 61 2015 http://dx.doi.org/10.1103/PhysRevLett.61.2015
[34] Zhong Y Wang Y F Wang Y Q Luo H G 2013 Phys. Rev. B 87 035128 http://dx.doi.org/10.1103/PhysRevB.87.035128
[35] Lacroix C Cyrot M 1979 Phys. Rev. B 20 1969 http://dx.doi.org/10.1103/PhysRevB.20.1969
[36] Li H Liu Y Zhang G M Yu L 2015 J. Phys. Condens. Matter 27 425601 http://dx.doi.org/10.1088/0953-8984/27/42/425601
[37] Li H Song H F Liu Y 2016 Euro. Phys. Lett. 116 37005 http://dx.doi.org/10.1209/0295-5075/116/37005
[38] Ganesh R Brink J van den Nishimoto S 2013 Phys. Rev. Lett. 110 127203 http://dx.doi.org/10.1103/PhysRevLett.110.127203
[39] Clark B K Abanin D A Sondhi S L 2011 Phys. Rev. Lett. 107 087204 http://dx.doi.org/10.1103/PhysRevLett.107.087204
[40] Hasan M Z Kane C L 2010 Rev. Mod. Phys. 82 3045 http://dx.doi.org/10.1103/RevModPhys.82.3045
[41] Qi X L Zhang S C 2011 Rev. Mod. Phys. 83 1057 http://dx.doi.org/10.1103/RevModPhys.83.1057
[42] Pixley J H Yu R Paschen S Si Q 2018 Phy. Rev. B 98 085110 http://dx.doi.org/10.1103/PhysRevB.98.085110
[43] Zhou Y Wu Q Rosa P F S Yu R Guo J Yi W Zhang S Wang Z Wang H Cai S Yang K Li A Jiang Z Zhang S Wei X Huang Y Yang Y F Fisk Z Si Q Sun L Zhao Z 2017 Sci. Bull. 62 1439 http://dx.doi.org/10.1016/j.scib.2017.10.008
[44] Kasaya M Tani T Iga F Kasuya T 1988 J. Magn. Magn. Mater. 76&77 278 https://doi.org/10.1016/0304-8853(88)90395-2
[45] Malik S K Adroja D T Dhar S K Vijayaraghavan R Padalia B D 1989 Phys. Rev. B 40 2414 http://dx.doi.org/10.1103/PhysRevB.40.2414
[46] Adroja D T Rainford B D 1994 Physica B 194–196 363 https://doi.org/10.1016/0921-4526(94)90511-8
[47] Hu J Alicea J Wu R Franz M 2012 Phys. Rev. Lett. 109 266801 http://dx.doi.org/10.1103/PhysRevLett.109.266801