Brought to you by:
Paper The following article is Open access

Exact chiral spin liquids and mean-field perturbations of gamma matrix models on the ruby lattice

, and

Published 28 November 2012 © IOP Publishing and Deutsche Physikalische Gesellschaft
, , Focus on Quantum Spin Liquids Citation Seth Whitsitt et al 2012 New J. Phys. 14 115029 DOI 10.1088/1367-2630/14/11/115029

1367-2630/14/11/115029

Abstract

We theoretically studied an exactly solvable gamma matrix generalization of the Kitaev spin model on the ruby lattice, which is a honeycomb lattice with 'expanded' vertices and links. We find that this model displays an exceptionally rich phase diagram that includes (i) gapless phases with stable spin Fermi surfaces, (ii) gapless phases with low-energy Dirac cones and quadratic band touching points and (iii) gapped phases with finite Chern numbers possessing the values ±4,±3,±2 and ±1. The model is then generalized to include Ising-like interactions that break the exact solvability of the model in a controlled manner. When these terms are dominant, they lead to a trivial Ising ordered phase which is shown to be adiabatically connected to a large coupling limit of the exactly solvable phase. In the limit where these interactions are weak, we treat them within mean-field theory and present the resulting phase diagrams. We discuss the nature of the transitions between various phases. Our results show the richness of possible ground states in closely related magnetic systems.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The study of zero-temperature properties of 'frustrated' magnetic systems has been reinvigorated in recent years with the development of powerful numerical methods [13], the discovery of new classes of exactly solvable models [4, 5] and the realization that intriguing topological ground states could occur [6, 7]. In particular, there is a large class of exactly solvable quantum spin models known as Kitaev models [5] which have revitalized the study of exactly solvable spin-liquid systems. The appeal of this class of models lies in the relative ease with which Kitaev's original version can be generalized, spawning many variants [819], and its possession of many nontrivial properties [2029] even in the presence of disorder [3032]. Moreover, the exact solution of Kitaev models does not require any more effort than solving the problem of noninteracting particles moving in a static background magnetic field. Nonetheless, the exact eigenstates of all Kitaev models are nontrivial entangled many-body wavefunctions [22, 23]. More specifically, their respective ground states are examples of quantum spin liquids, which are insulating quantum states of matter exhibiting no conventional long-range magnetic order at zero temperature. Although no physical example of these models has been established in nature, there have been several experimental proposals for realizing them [3339]. Nevertheless, the importance of this entire class of models lies in providing a strong case—by proof of principle—for the existence of exotic emergent phenomena in many-body quantum phases of matter such as quantum spin liquids, as well as their utility as model systems for the study of nontrivial, nonperturbative emergent quantum phenomena in general.

In this paper, we study a new exactly solvable Kitaev model and examine its response to the inclusion of interactions that destroy the exact solvability. This will motivate us to analyze the order that is favored by these new interactions and their effects on the exactly solvable ground state. The model is a gamma matrix model (GMM) extension [11, 12, 16] on the two-dimensional ruby lattice. Previous studies of the ruby lattice have yielded a topological insulator [40], fractional quantum anomalous Hall states [41] and topological anyons [42]. In this paper, we show that a GMM on the ruby lattice realizes a quantum spin liquid ground state with an unusually rich phase diagram. The model is then generalized by the inclusion of Ising-like interactions that spoil the exact solvability of the model. We explore the interplay between these interactions and the many ground state spin-liquid phases.

This paper is organized as follows. In section 2, we introduce the Hamiltonian and discuss its solution by mapping it onto a system of noninteracting Majoranas moving in an emergent $\mathbb {Z}_2$ gauge field. In section 3, we discuss the various ground state phases that the model realizes and present a phase diagram. In section 4, we examine the Ising order in one of the gamma matrix operators which may be induced by including Ising-like interactions. In section 5, we treat these extra interactions within the mean-field approximation and discuss their effects on the phase diagram. We then present the conclusions in section 6.

2. The gamma matrix model and Hamiltonian

We consider spin-3/2 moments located at the sites of a ruby lattice, shown in figure 1(a). Alternatively, one can think of each site as being occupied by a single spin-1/2 which has an additional orbital degree of freedom. At each site, we introduce five 4 × 4 Hermitian gamma matrices which satisfy a Clifford algebra locally {Γaibi} = 2δab, with $a,b=1,\dots ,5$ [11, 12, 16]. These matrices also have a representation in terms of bilinears of SU(2) spin-3/2 operators on each site i as follows:

Equation (1)

The gamma matrices may also be expressed as tensor products of Pauli spin-1/2 matrices {σμ,τν} on the site i,

Equation (2)

where σ0i = τ0i = 1, the 2 × 2 identity matrix. In the spin–orbit interpretation (2), σ acts on the real spin degree of freedom and τ on the orbital degree of freedom. Thus, one can view the GMM we study below as applying to both a spin-3/2 model and a two-orbital spin-1/2 model. In either case, there are four physical states per lattice site.

Figure 1.

Figure 1. (a) The ruby lattice. The ruby lattice can be viewed as an 'expanded' honeycomb lattice with triangles replacing the vertices and squares replacing the bonds. (b) The couplings in our Hamiltonian equation (3) and the ground state flux configuration of the $\mathbb {Z}_2$ gauge field. Note that the J' couplings (not shown) are also defined on the same links as the Js. The arrows indicate the choice of gauge used for the uij fields in (5).

Standard image

The Hamiltonian we study is given by

Equation (3)

where we define the operators Γabi = [Γaibi]/(2i) in terms of the Γai given in either equation (1) or equation (2). The couplings $J_{\triangle },J'_{\triangle },J_{\unicode{x2394} }$ and $J'_{\unicode{x2394} }$ are taken to be positive and are link dependent, as shown in figure 1(b). For generic couplings, the Hamiltonian has translational and sixfold rotational lattice symmetry. With the interpretation in terms of spin-3/2 moments, it has global Ising spin symmetry under 180° rotations about the z-axis, and time-reversal symmetry (TRS), although TRS will be spontaneously broken in the ground state as we will be show below.

While the model (3) has unusual spin symmetries in terms of the underling spin and orbital degrees of freedom, its structure allows an exact solution. A key ingredient for the exact solvability of the model is the existence of conserved operators on every plaquette of the lattice which commute with the Hamiltonian [5]. These operators are the three ($\hat {W}_{\triangle }$ ), four ($\hat {W}_\square $ ) and six point ($\hat {W}_{\unicode{x2394} }$ ) operators defined on the triangles, squares and hexagons of the lattice (as their subscripts suggest). Explicitly, they are given by $\hat {W}_{\triangle } = \Gamma ^{12}_i \Gamma ^{12}_j \Gamma ^{12}_k,$ $\hat {W}_{\unicode{x2394} } = \Gamma ^{34}_i \Gamma ^{34}_j \Gamma ^{34}_k \Gamma ^{34}_l \Gamma ^{34}_m \Gamma ^{34}_n$ and $\hat {W}_{\square } = \Gamma ^{23}_i \Gamma ^{14}_j \Gamma ^{23}_k \Gamma ^{14}_l,$ where the sites are labeled counter-clockwise and for $\hat {W}_{\square }$ the first link 〈ij〉 lies on a triangle. The model is then solved by introducing a Majorana representation of the Γ-matrices using six flavors of Majorana operators, {ξ1i,ξ2i,ξ3i,ξ4i,ci,di}, at each site i:

Equation (4)

where the Majorana fermions have the important property that (ξai) = ξai, ci = ci and di = di. The Hamiltonian written in terms of these operators then reduces to that of two species of Majorana fermions, c and d, which are noninteracting and move in a background $\mathbb {Z}_2$ gauge field uij,

Equation (5)

Here the fields uij are defined by uij = −i ξ1iξ2j if $ij \;\in \triangle $ and uij = −i ξ3iξ4j if $ij\in \unicode{x2394} $ . The uij are, in general, quantum fields with eigenvalues ±1 since u2ij = 1, hence their identification with a $\mathbb {Z}_2$ gauge theory. Moreover, we can simultaneously diagonalize ${\tilde {\mathcal {H}}}$ and {uij} since they commute. We emphasize that the interacting Hamiltonian (3) with the representation (4) is reduced to the effectively noninteracting Hamiltonian (5) because the uij behave as constants in each flux sector of the $\hat {W}_{\triangle }$ , $\hat {W}_\square $ and $\hat {W}_{\unicode{x2394} }$ .

However, the full Hilbert space spanned by the Majorana fermions is overcomplete, and a constraint must be enforced to ensure that the Clifford algebra of Γ operators is satisfied. This constraint is expressed by the operator equation Di = − Γ1iΓ2iΓ3iΓ4iΓ5i = − i ξ1iξ2iξ3iξ4icidi = 1 and is enforced by the projector $P=\prod _{i} [\frac {1+D_{i}}{2} ]$ , where the product is over all sites in the lattice. The original Hamiltonian is obtained through ${\cal H}=P \tilde{{\cal H}}P$ .

The Z2 gauge fields define fluxes ϕp via $\exp (\mathrm {i}\phi _p)\equiv \prod _{jk\in p} \mathrm {i}u_{jk}$ , where jk is taken counterclockwise on each elementary plaquette p. These fluxes are then related to the conserved $\hat {W}$ s by $W_{p} \propto \prod _{ij\in p}u_{ij}$ , where ij is also taken counterclockwise. Since the eigenstates of the Hamiltonian are also eigenstates of the fluxes, once the fluxes have been specified, the uij = ± 1s are uniquely specified up to $\mathbb {Z}_2$ gauge transformations, and the Hamiltonian ${\skew3\tilde{\mathcal {H}}}$ is then diagonalized and projected to yield to eigenstates of $\mathcal {H}$ . The ground state is then determined by a many-body Majorana wavefunction minimizing the total energy. Due to translational symmetry, ${\skew3\tilde{\mathcal{H}}}$ describes a band structure for the c and d fermions. It can be shown that the minimal energy configuration is the one where all the negative 'eigenstates' of the effective band Hamiltonian (5) are occupied [5, 11, 12].

One finds that ϕp = ± π/2 in the triangular plaquettes, and ϕp = 0,π in the hexagonal and square plaquettes. Under TRS, Wp → ± Wp where −(+) is for triangle (hexagon and square) plaquettes; it follows that ϕp → − ϕp for triangle plaquettes, while ϕp remains unchanged for hexagon and square plaquettes. Consequently, a ground state with a certain flux pattern {ϕp} spontaneously breaks TRS. The ground state energy of a flux configuration must be degenerate with the flux pattern obtained from {ϕp} by changing ϕp → − ϕp on all triangular plaquettes [5, 11].

To determine the physical ground state, the flux configuration that minimizes the ground state energy needs to be determined. This was accomplished by first numerically determining the ground state energy for each flux configuration for the symmetric couplings $J_{\triangle }=J'_{\triangle }=J_{\unicode{x2394} }=J'_{\unicode{x2394} }=J$ and J5 = 0. The ground state flux with the least energy is depicted by figure 1(b). Generally, the ground state flux will be a function of J5, but for simplicity we will not consider additional flux configurations in this work because an additional term (depending on $\hat {W}_{\triangle },\hat {W}_{\square },\hat {W}_{\unicode{x2394} }$ ) can always be included to favor a particular configuration without destroying the exact solvability [20, 28]. By general arguments regarding the Majorana representation of the spin operators in the 3/2 representation [5, 25], the spin–spin correlation functions of the ground state are identically 0 beyond nearest-neighbor sites. Hence, the ground state is also a quantum spin liquid in these observables.

With the gauge sector of $\mathcal {H}$ specified, we can solve our system by mapping the Majorana fermions to complex fermions ai, defined by ci = ai + ai and di = −i(ai − ai). This puts our Hamiltonian into the form

Equation (6)

where we have defined the quantities $\mathrm {i} u^c_{ij} = \mathrm {i} u_{ij} J_{\triangle (\unicode{x2394} )}$ and $\mathrm {i} u^d_{ij} = \mathrm {i} u_{ij} J'_{\triangle (\unicode{x2394} )}$ for $ij \in \triangle (\unicode{x2394} )$ . This type of Hamiltonian is solved by taking a Bogoliubov transformation in momentum space. Taking the Fourier transform of (6) and introducing the Nambu spinor as $\Psi^{\rm T}(k) = ( a_1 (k) a_2 (k) \cdots a_6 (k) a^\dagger _1 (-k) \cdots a^\dagger _6 (-k))$ , where ai(k) is the Fourier mode of the fermion creation operator on the ith site in the unit cell, we obtain

Equation (7)

with the identifications

Equation (8)

where the $\tilde {u}$ s are the Fourier modes of the corresponding us defined above, and where C+ is half of the Brillouin zone (BZ). We only sum over half of the BZ to avoid the double counting introduced by the Nambu spinor. Because H(k) = −H(−k), the eigenvalues of H(k) appear in pairs {Ej(k),−Ej(−k)}, j = 1,2,...,6, and our model is particle–hole symmetric. This is a consequence of our Hamiltonian being written in terms of the Nambu spinor, which satisfies Ψi(−k) = Ψi(k); when we diagonalize H(k), half of the states are redundant.

3. Band structure and phase diagram

For general couplings, the ground state may be gapped or gapless. For the gapped phases, the band structure may possess a nontrivial (nonzero) Chern number. We will refer to these nontrivial gapped phases as Chern phases. The physical consequence of a nonzero Chern number is the appearance of chiral gapless edge modes in a system with a boundary. If the system has a finite Chern number, it will exhibit a quantized thermal Hall conductance. We find these signatures of a nontrivial topological phase by calculating the Chern number numerically [29, 43] and diagonalizing a system with boundaries (a 'strip' geometry) to determine the existence of gapless edge modes. This is most conveniently done by taking a Fourier transform to momentum space along the 'length' of the strip geometry [5, 11, 12].

An example band structure is shown in figure 2, which illustrates the dispersion of Majorana excitations $E(\vec k)$ as a function of the two planar momentum components kx,ky. The state in figure 2 is gapped about zero energy (below which all states are occupied and above which all states are empty) and has a nontrivial Chern number. There are six bands above and six bands below zero energy, corresponding to the six sites in the unit cell of the ruby lattice. Shown in figure 3 is the spectrum of a strip geometry in the nontrivial phase with Chern number −1.

Figure 2.

Figure 2. Band structure for the coupling parameters $\{J_{\triangle },J'_{\triangle },J_{\unicode{x2394} },J'_{\unicode{x2394} },J_5\} =\{1.0,0.75,1.0,1.33,1.0 \} $ , where $q_1 = (2\pi /3)(\sqrt {3}-1)$ and $q_2 =\pi (1-1/\sqrt {3})$ . There are 12 bands that exhibit a redundant particle–hole symmetry and the energy spectrum is gapped with Chern number 2.

Standard image
Figure 3.

Figure 3. Spectrum of the strip geometry of the system as a function of the momentum ky, parallel to the edge. The system has Chern number −1 and couplings $\{J_{\triangle },J'_{\triangle },J_{\unicode{x2394} },J'_{\unicode{x2394} },J_5\}=\{1.0,0.6,1.0,1.67,0.2\}$ . Two chiral edge modes that traverse the band gap at E = 0 are clearly visible and each corresponds to a mode on a single edge.

Standard image

When the ground state is gapless, it generically has a Fermi surface (FS). Shown in figures 4 and 5 are band structures with an FS realized in the ground state. In figure 4 the FS is in the center of the first BZ, while in figure 5 the FS is near the high-symmetry K and K' points of the hexagonal BZ. Note that the FS is 'connected' once the appropriate reciprocal lattice vectors are used to shift them together. At a phase boundary between two gapped phases with different Chern numbers, the ground state is gapless with Dirac nodes at a discrete set of points in the BZ. An example of such a band structure is shown in figure 6, which has four nodes in the first BZ (one at the zone center and three others at the inversion symmetric M points).

Figure 4.

Figure 4. Band structure and shape of the FS (blue) at $\{J_{\triangle },J'_{\triangle },J_{\unicode{x2394} },J'_{\unicode{x2394} },J_5\}=\{1.0,1.0,1.0,1.0,1.4\}$ and where $q_1 = (2\pi /3)(\sqrt {3}-1)$ and $q_2 =\pi (1-1/\sqrt {3})$ . The FS pocket is located around the BZ center. The dots are indicative of the zone boundary.

Standard image
Figure 5.

Figure 5. Band structure and shape of the FS (blue) at $\{J_{\triangle },J'_{\triangle },J_{\unicode{x2394} },J'_{\unicode{x2394} },J_5\}=\{1.0,1.0,1.0,1.0,0.55\}$ and where $q_1 = (2\pi /3)(\sqrt {3}-1)$ and $q_2 =\pi (1-1/\sqrt {3})$ . The FS pockets are located around the K and K' symmetry points of the BZ. The dots are indicative of the zone boundary.

Standard image
Figure 6.

Figure 6. Band structure of a nodal phase with four nodes and at parameters $\{J_{\triangle },J'_{\triangle },J_{\unicode{x2394} },J'_{\unicode{x2394} },J_5\}=\{1.0,0.58,1.0,1.73,2.62\}$ . The nodes are located at the zone center and the three M points of the BZ.

Standard image

In general, the phase diagram is complicated due to the large number of tunable parameters, even when one insists on translational, sixfold rotation and inversion symmetry. We report on a phase diagram resulting from an interesting selection of parameters which reveals the general richness of the model. We explored the phase diagram for the specific case when the couplings $J_{\triangle }=J_{\unicode{x2394} }$ are set equal to a fixed constant J, and the J' couplings are allowed to vary in such a way that their product remains constant; explicitly $J'_{\unicode{x2394} } \times J'_{\triangle }=J^2$ . This phase diagram is shown in figure 7 and is exceptionally rich with many gapped and gapless phases.

Figure 7.

Figure 7. Phase diagram of the spin liquid at couplings $J_{\triangle }=J_{\unicode{x2394} }=J$ and $ J'_{\triangle }\times J'_{\unicode{x2394} }=J^2$ . The different colored regions correspond to either a gapped phase with Chern number as given in the legend or if gapless correspond to an FS phase. Lines that are boundaries between gapped phases are gapless phases with nodes or Fermi points at specific points in the BZ. The single-node phase is gapless at the zone center k = (0,0) and can be either a Dirac cone or a quadratic band touching point; the double-node phase is gapless at the K and K' points with $\mathbf{k} = (\pm \pi (\sqrt {3}-1)/3,\pm \pi (1-1/\sqrt {3}))$ and the triple-node phase is gapless at the three inversion symmetric M points, $ {\bf k} = (0,\pm \pi (1-1/\sqrt{3}))$ and $ {\bf k} = (\pm \pi (\sqrt{3}-1)/2,\;\pm \pi (1-1/\sqrt{3})/2)$ . These lines are labeled according to the number of such nodes.

Standard image

The regions of the phase diagram are classified by whether or not they are gapped or gapless and whether or not the gapless phases possess a nontrivial Chern number (Chern phase). If gapless, the phases are classified according to whether the gap closes at discrete Fermi points in the BZ or along a line or FS. Often the former contains Dirac nodes on the BZ boundary or the zone center where the Fermi point may also be a quadratic band touching point. The Fermi points only occur on critical lines separating two Chern phases or a Chern phase and an FS. We will refer to these critical phases as nodal lines (NL).

The phase diagram in figure 7 contains NLs with up to three Fermi points. Since the NLs are obtained by solving a secular equation of a Bloch Hamiltonian which is analytic, the NLs trace out smooth trajectories in the phase diagram. Moreover, the complicated manner in which they cross yields the many Chern phases in figure 7. Most interestingly, we see a three-node NL which closes in on itself to form an ellipse.

Another interesting region is the high-symmetry line at constant $J'_{\triangle }/J=1$ where $J_{\triangle }=J'_{\triangle }=J_{\unicode{x2394} }=J'_{\unicode{x2394} }=J$ . This line contains a gapless FS phase in a line between 0.7 ≲ J5 ≲ 2.0, and there are four NLs which cross this FS line at its endpoints. Along this line there are no nearby Chern phases in the phase diagram. This FS phase is noncritical with regard to tuning $J'_{\triangle }$ and $J'_{\unicode{x2394} }$ , appears to be particularly stable and is the FS phase with the largest volume in the phase diagram.

Finally, we note that in the phase diagram, transitions between Chern phases and FS phases involve critical NLs. This transition will lead to FS pockets forming around the nodes located at the points of high symmetry. Figures 4 and 5 show examples of such pockets. There are also transitions between Chern phases and FS phases which do not involve an NL. In this case, the FS becomes gapped by a pairing (Cooper) instability as the system is driven into the Chern phase.

In section 4, we will study the effects of the inclusion of new interaction terms that spoil the exact solvability of the model. These new interactions also motivate the study of the Ising order in one of the gamma matrices, namely Γ5i.

4. The Ising order of Γ5

In this section, we consider the following extension to $\mathcal {H}$ :

Equation (9)

We limit ourselves to the case where λ > 0. For our analysis it is convenient to take the spin–orbit interpretation of the model where we have four local degrees of freedom with basis states $\{ \left | \uparrow \uparrow \right \rangle , \left | \uparrow \downarrow \right \rangle , \left | \downarrow \uparrow \right \rangle , \left | \downarrow \downarrow \right \rangle \}$ . On this basis, Γ5i is trivially diagonal with eigenvalue +1 if the 1/2-spins are aligned and eigenvalue −1 if they are anti-aligned. Thus, we can regard Γ5 as an Ising spin, albeit one that is doubly degenerate for each Ising polarization. Then, $\mathcal {H}_{\mathrm {Ising}}$ , as its name suggests, describes ferromagnetic nearest-neighbor Ising–Ising interactions in the case where λ > 0. It is also trivially true that $[\mathcal {H}_{\mathrm {Ising}} , \Gamma _i^5 ]=0$ for all sites i, which implies that both operators are simultaneously diagonal in the local spin–orbit basis. Thus in the $\mathcal {H}_0 = 0$ limit, the model is trivially exactly solvable with eigenstates given by tensor products of local Γ5i eigenstates. A typical eigenstate would have the form

Equation (10)

where the N sites of the lattice are partitioned into disjoint sets I and J with NI and NJ sites each, respectively. The energy of such a state is then given by

Equation (11)

where Nlink is the total number of nearest-neighbor links and Ldom. is the total length of all the boundaries between different 'Ising domains'. The arbitrariness in the parameters α,β,γ,δ on every site leads to a macroscopic degeneracy of the state $\left |\psi \right \rangle $ ; this stems from the local degeneracy of Γ5i. Also, an exact eigenstate of Γ5i will exhibit no fluctuations in the Γ5i observables. Hence, the ground states of $\mathcal {H}_{\mathrm {Ising}} + \mathcal {H}_5$ are trivial Ising ferromagnets with order in 〈Γ5i〉 = ± 1. The sign of the order parameter will depend on the sign of J5 if it is nonzero. Otherwise, if J5 = 0, the ground state will spontaneously break the Ising symmetry in Γ5i.

However, when $\mathcal {H}_0 \neq 0 $ the situation is no longer trivial and no longer exactly solvable. In particular, $[\mathcal {H}_0 , \Gamma ^5_i ] \neq 0$ suggests that it would be difficult to diagonalize the Hamiltonian in the spin–orbit basis or any basis where Γ5i is diagonal. Nevertheless, one still has $[W_p,\mathcal {H}_{\mathrm {Ising}}]=0$ for all plaquettes p. This implies that the flux invariants are still good quantum numbers for both the GMM and the $\mathcal {H}_{\mathrm {Ising}}$ , and it is sensible to solve for the eigenstates within a given flux sector {Wp}. Writing $\mathcal {H}_{\mathrm {Ising}}$ after performing Majorana transformation makes this explicit,

Equation (12)

Hence, $\tilde {\mathcal {H}}_{\mathrm {Ising}}$ does not couple to the $\mathbb {Z}_2$ gauge field degrees of freedom and the fluxes are conserved. If we now consider again the limit where $\mathcal {H}_0=0$ , then $\tilde {\mathcal {H}}$ describes a trivial insulating Hamiltonian with no kinetic terms but with local a-number conservation. One can then write down the exact eigenstates in this representation which is just any occupation state of the complex fermion ai for all sites i where the flux configuration can remain arbitrary. This gives another explanation for the macroscopic degeneracy of |ψ〉, which in the gauge field picture is due to the many degenerate {Wp} sectors. Equivalently, the flux invariant operators are insensitive to whether or not the local 1/2-spins are aligned or anti-aligned; rather, they operate on degrees of freedom that are orthogonal. We can now give also a physical interpretation of the complex fermion a. Namely, its occupation parity Pi = (2aiai − 1) is the Γ5i observable whose eigenvalue determines whether or not the local 1/2-spins are aligned or anti-aligned. Also, the a occupation numbers do not fluctuate in an eigenstate, which is consistent with the nonfluctuating behavior of the Ising order parameter.

Now we return to considering general $\mathcal {H}$ . Note that $\mathcal {H}_0$ commutes neither with Γ5i nor with its sum total $\sum _i\Gamma _i^5$ . This leads to non-conservation of the total a-particle number in the ground state, and we can associate this nonconservation with the pairing terms in the a-representation of the problem. Nevertheless, the product of parities as defined by $\mathcal {P}=\prod _i P_i = \prod _i \Gamma _i^5$ is conserved since $\mathcal {H}_0,\mathcal {H}_5$ and $\mathcal {H}_{\mathrm {Ising}}$ all individually commute with it. In fact $\mathcal {P}$ implements a global $\mathbb {Z}_2$ gauge transform, ai← − ai, ai← − ai on every site i. This is merely another manifestation of the conservation of the total a-number modulo 2 by $\mathcal {H}$ .

Next, we will consider the opposite limit where $\mathcal {H}_{\mathrm {Ising}} + \mathcal {H}_5 = 0$ and argue that the exact ground states of $\mathcal {H}_0$ will not possess any Ising order in Γ5 by themselves, but that the exact ground states of the more general GMM with J5 ≠ 0 will. Moreover, in the limit of small J5, these ground states are adiabatically connected. Given a fixed flux sector {Wp} and given the existence of a nondegenerate ground state of $\tilde {\mathcal {H}}_0$ which we denote by |Ω〉, we will show that 〈Ω|Γ5i|Ω〉 = 0 for all i.

First consider equation (6) in the limit J5 = 0 where the gauge fields uij have been fixed. Define a linear unitary particle–hole conjugation operator $\mathcal {C}$ by its conjugation on ai and ai and its action on the a-number vacuum |0〉 as follows:

Equation (13)

θ is a phase which we can fix by demanding that $\mathcal {C}^2|0\rangle =|0\rangle $ . For N even, this leads to $\theta =\frac {N\pi }{4}$ . These relations then totally specify $\mathcal {C}$ and also imply the relations $\mathcal {C}^\dagger =\mathcal {C}$ and $\mathcal {C}^2=1$ . Note that $[\tilde {\mathcal {H}}_0,\mathcal {C}] =0$ , implying that if |Ω〉 is nondegenerate, then $\mathcal {C}|\Omega \rangle $ can only differ from |Ω〉 by a phase. Thus, $\langle \Omega |a_i^\dagger a_i | \Omega \rangle = \langle \Omega |\mathcal {C}^\dagger a_i a_i^\dagger \mathcal {C}|\Omega \rangle = \langle \Omega |a_ia_i^\dagger |\Omega \rangle $ . Using {ai,ai} = 1, we then conclude that 〈Ω|aiai|Ω〉 = 1/2 and 〈Ω|Γ5i|Ω〉 = 0. Since aiai is $\mathbb {Z}_2$ gauge invariant, we expect this to hold true even after projection. One could also heuristically argue that because J5 couples to aiai like a chemical potential, the ground state will have a particle–hole symmetry and is the half-filled Fermi sea. However, the presence of pairing terms aiaj + aiaj complicates this argument.

For the model that was solved in section 2 where J5 ≠ 0 generally, 〈Γ5i〉 may take nonzero values in (0,1] when in the ground state. In this more general situation, the presence of $\mathcal {H}_5$ breaks the symmetry under $\mathcal {C}$ conjugation. In addition, since the ground states of the GMM were shown to not undergo a phase transition as J5 is tuned from 0, we conclude that the exactly solvable ground states of the GMM may exhibit Ising order in the Γ5i spins. That is, the exactly solvable phase and Ising order are not mutually exclusive. Hence the ground state is not a spin liquid in these observables. But fluctuations in Γ5i remain nontrivial in the exactly solvable phase. This poses the interesting question of to what extent an exactly solvable GMM ground state is adiabatically connected to the trivial Ising states such as that described by |ψ〉 above.

In the limit where $\mathcal {H}_0$ is treated as a perturbation to $\mathcal {H}_5+\mathcal {H}_{\mathrm {Ising}}$ , the weakly fluctuating Ising order Γ5i will strongly renormalize the J5 coupling as seen by the a fermions, which can be seen from equation (12). Thus, we can expect this limit to be equivalent to a large J5 limit and to be described by a low-energy effective Hamiltonian derived from perturbation theory. Schematically, the Hamiltonian can be written in terms of Wp operators [5, 11, 44],

Equation (14)

where {αp} are coupling constants and the dots represent higher order terms. Such an effective theory will break the degeneracy among the different flux sectors and favor certain flux sectors over others. However, TRS is still preserved and must be spontaneously broken by the ground state. If J5 = 0, then the ground state must also spontaneously break the Ising symmetry of Γ5.

In the opposite limit where $\mathcal {H}_{\mathrm {Ising}}$ is the perturbation to $\mathcal {H}_0+\mathcal {H}_5$ , the situation is more complicated as the ground state of the exactly solvable phase is, in general, nontrivial. We try to address this question in the next section of the paper where $\mathcal {H}_{\mathrm {Ising}}$ is treated as a perturbation to the exactly solvable GMM within the mean-field approximation.

5. Mean-field approximation of the Ising perturbation

In this section, we study the effects of $\mathcal {H}_{\mathrm {Ising}}$ as a perturbation to the exactly solvable $\mathcal {H}_0+\mathcal {H}_5$ within the mean-field approximation. Performing a mean-field decomposition with respect to the order parameter 〈Γ5i〉 results in the following mean-field Hamiltonian:

Equation (15)

where the mean field 〈Γ5i〉 has to be determined self-consistently. Motivated by the fact that the Ising couplings are ferromagnetic, we make the ansatz that 〈Γ5i〉 is uniform across the entire lattice. At the mean-field level, the effect of the interaction $\mathcal {H}_{\mathrm {Ising}}$ can already be seen to renormalize J5 with Jeff5 = J5 + 4λ〈Γ5〉.

Within this approximation, we determined the phase diagrams for varying J5 and λ with fixed J and J' couplings and fixed flux sector. Two phase diagrams are shown in figures 8 and 9. We note that generically the topological phases are stable against weak $\mathcal {H}_{\mathrm {Ising}}$ perturbations. At larger λ, however, the energetics always prefer the trivial gapped phase which has zero Chern number and large 〈Γ5〉. Hence, large λ eventually renormalizes J5 to larger effective values. This is consistent with the large J5 and large λ trivial phases being adiabatically connected to each other.

Figure 8.

Figure 8. Phase diagram with the Ising perturbation treated within the mean-field approximation. The couplings are $\{J_{\triangle },J'_{\triangle },J_{\unicode{x2394} },J'_{\unicode{x2394} }\}=\{1.67,1.0,0.3,1.0\}$ .

Standard image
Figure 9.

Figure 9. Phase diagram with the Ising perturbation treated within the mean-field approximation. The couplings are $\{J_{\triangle },J'_{\triangle },J_{\unicode{x2394} },J'_{\unicode{x2394} }\}=\{1.0,1.0,0.2,0.2\}$ . The phase boundary separating the two trivial phases is a first-order transition line where 〈Γ5i〉 changes discontinuously.

Standard image

Focusing along the line at constant λ = 0 and increasing J5, we see a series of gapped and gapless phases, which eventually ends with a trivial phase at large J5. This is as expected from the results of the previous section, but since J5 acts as an external field, the Ising order is 〈Γ5i〉 ≠ 0 on this line. When λ is tuned into greater values, the volume of these intermediate phases diminishes continuously, eventually giving way to the trivial phase. This confirms that the trivial Ising phase is adiabatically connected to the trivial large J5 phase.

In figure 8, the gapped phase at J5 = 0 appears to be the most stable with respect to increasing λ. However, for the couplings shown in figure 9, the J5 = 0 phase gives way to a trivial gapped phase (Chern number zero) which is not adiabatically connected to the trivial phase at large J5. This phase undergoes a first-order phase transition to the larger J5 trivial phase as shown in figure 9. Thus, the Ising interaction may choose to stabilize certain gapped phases over the default J5 = λ = 0 phase. The stabilized phases are highly dependent on the particular values of the other coupling constants.

At constant J5 > 0, increasing λ may stabilize certain phases that would typically require larger J5 values. For example, in figure 8 along the line J5 = 0.75, the system is driven to an FS phase with increasing λ. Other examples of such transitions driven by the Ising interaction may also been seen in figures 8 and 9. Hence, the Ising perturbation treated at mean field may stabilize certain exotic phases which are otherwise only accessible by intermediate values J5.

In summary, these results show that with increasing λ, the ground state is eventually adiabatically connected to the trivial Ising ordered state of section 4. However, in general, various phase transitions involving other exotic phases must occur before this happens. Moreover, this depends crucially on the initial (λ = 0) exactly solvable phase.

6. Conclusions

In this work, we studied an exactly solvable Γ-matrix generalization of Kitaev's original spin-1/2 model on the ruby lattice shown in figure 1. Our model can be interpreted as describing either a spin-3/2 system or a double-orbital spin-1/2 model. We have shown that the ruby lattice, with its many possible tunable parameters, exhibits a rich phase diagram with many exotic phases realized by the same Hamiltonian. We have found gapless phases with FSs and Fermi points, as well as gapped topologically nontrivial phases. Next, we studied the effects of including Ising-like interaction terms that destroy the exact solvability of the model. We analyzed the ordered phase that these terms lead to and argued that the ground state in the large J5 limit of the exactly solvable model is adiabatically connected to it. We then analyzed in detail the general model near the limit of an exactly solvable phase and treated the Ising terms as perturbations within the mean-field approximation. We derived phase diagrams that show that the Ising interactions will eventually favor the trivial Ising ordered phase of the model over the other more exotic phases. We also confirm that the large J5 exactly solvable phase is adiabatically connected to the trivial Ising ordered phase.

Although this study has focused on a particular exactly solvable spin model and special perturbations on it, the connections previously drawn to other topological phases indicate that our results are rather broadly applicable [45]. Moreover, many directions for further study in the future suggest themselves. These include the following. (i) Considering anti-ferromagnetic Ising interactions (λ < 0) in order to determine whether the interplay between the exactly solvable GMM and the Ising interactions might be different. (ii) The effect of disorder in interacting many-body quantum systems remains a key open problem. The model we study here with its unusually rich phase diagram provides an opportunity to study disorder effects within a well-controlled model [32]. (iii) Doping magnetic systems is one route to high-temperature superconductivity. To date, no work has been reported on the doping of gamma matrix generalizations of the Kitaev model, and only a few published studies exist on the original model [46]. The richness of the model we study here could shed light on doping magnetic systems in much a more general context and could ultimately help guide the discovery of strongly correlated materials exhibiting high-temperature or unconventional superconductivity.

Acknowledgment

We gratefully acknowledge funding from ARO grant W911NF-09-1-0527 and NSF grant DMR-0955778.

Please wait… references are loading.
10.1088/1367-2630/14/11/115029