+ All Categories
Home > Documents > arXiv:1706.09908v1 [cond-mat.str-el] 29 Jun 2017 · PDF file3Department of Applied Physics,...

arXiv:1706.09908v1 [cond-mat.str-el] 29 Jun 2017 · PDF file3Department of Applied Physics,...

Date post: 16-Mar-2018
Category:
Upload: phungtuong
View: 214 times
Download: 1 times
Share this document with a friend
17
Signatures of Quantum Spin Liquid in Kitaev-like Frustrated Magnets Matthias Gohlke, 1 Gideon Wachtel, 2 Youhei Yamaji, 3, 4 Frank Pollmann, 5 and Yong Baek Kim 2, 6, 7 1 Max-Planck-Institut f¨ ur Physik komplexer Systeme, 01187 Dresden, Germany 2 Department of Physics, University of Toronto, Toronto, Ontario M5S 1A7, Canada 3 Department of Applied Physics, The University of Tokyo, Hongo, Bunkyo-ku, Tokyo, 113-8656, Japan 4 JST, PRESTO, Hongo, Bunkyo-ku, Tokyo, 113-8656, Japan 5 Technische Universit¨ at M¨ unchen, 85747 Garching, Germany 6 Canadian Institute for Advanced Research/Quantum Materials Program, Toronto, Ontario MSG 1Z8, Canada 7 School of Physics, Korea Institute for Advanced Study, Seoul 130-722, Korea Motivated by recent experiments on α-RuCl3, we investigate a possible quantum spin liquid ground state of the honeycomb-lattice spin model with bond-dependent interactions. We consider the K - Γ model, where K and Γ represent the Kitaev and symmetric-anisotropic interactions between spin-1/2 moments on the honeycomb lattice. Using the infinite density matrix renormal- ization group (iDMRG), we provide compelling evidence for the existence of quantum spin liquid phases in an extended region of the phase diagram. In particular, we use transfer matrix spectra to show the evolution of two-particle excitations with well-defined two-dimensional dispersion, which is a strong signature of quantum spin liquid. These results are compared with predictions from Majorana mean-field theory and used to infer the quasiparticle excitation spectra. Further, we com- pute the dynamical structure factor using finite size cluster computations and show that the results resemble the scattering continuum seen in neutron scattering experiments on α-RuCl3. We discuss these results in light of recent and future experiments. I. INTRODUCTION One of the hallmarks of quantum spin liquid is the existence of fractionalized excitations 1 . While it is gen- erally difficult to detect the dispersion of single-particle excitations in fractionalized systems, the information about two-particle and other multi-particle excitations is contained in the dynamical spin structure factor mea- sured in inelastic neutron scattering. If the ground state is a quantum spin liquid, the lower boundary of multi- particle continuum should have a well-defined disper- sion. Many candidate materials for quantum spin liq- uid, however, do not allow such scattering experiment due to unavailability of large single crystal. Possible ex- periments are, therefore, quite often limited to thermo- dynamic measurements such as specific heat and sus- ceptibility, as well as thermal transport. Hence it is difficult to identify smoking-gun evidence for quantum spin liquid. In this context, recent inelastic neutron scattering experiments 2–4 on α-RuCl 3 provide valuable informa- tion about multi-particle continuum in a putative quan- tum spin liquid material 5 . α-RuCl 3 is one of the can- didate materials that support the Kitaev interaction 6 between j eff =1/2 pseudo-spin moments, which is a bond-dependent frustrated Ising interaction and arises from the combination of correlation effects and spin- orbit coupling 7–11 . When only the Kitaev interaction is present, such a model can be exactly solved 6 and the ground state is a quantum spin liquid. On the other hand, other interactions are generally present and they may drive a transition to a magnetically ordered state 12–15 . The compound α-RuCl 3 orders magnetically at low temperatures 16–18 , possibly due to the existence of other interactions mentioned above. In spite of this, the dy- namical structure factor shows a continuum of excita- tions at high energies both below and above the ordering temperature, which may be related to a nearby quantum spin liquid 19 . Elaborate ab initio computations indicate that the dominant exchange interactions are the Kitaev K and symmetric-anisotropic Γ interactions with small third neighbor J 3 Heisenberg interaction 13 . The dom- inance of K and Γ interactions is also pointed out in studies of multi-orbital Hubbard model 20 . Furthermore, it is interesting to notice that the K - Γ model is pro- posed for metal-organic frameworks with Ru 3+ or Os 3+ ions 21 . A previous theoretical work 22 of exact diago- nalization (ED) on finite size clusters suggest that the K - Γ model may host quantum spin liquid phases in an extended region of the phase diagram and small J 3 would drive a transition to the zig-zag order that is seen in the experiment. Such a study is naturally subject to finite size effect and it is still difficult to nail down the precise nature of the ground state. In fact, the ED study mentioned above introduce a spatial anisotropy in the Kitaev interaction to avoid strong finite size effect. In this work, we investigate the K - Γ model using then infinite density matrix renormalization group 23–25 (iDMRG), where the system is placed on an infinite cylinder. We first study the ground state energy, entanglement entropy, magnetization, and static struc- arXiv:1706.09908v2 [cond-mat.str-el] 14 Feb 2018
Transcript

Signatures of Quantum Spin Liquid in Kitaev-like Frustrated Magnets

Matthias Gohlke,1 Gideon Wachtel,2 Youhei Yamaji,3, 4 Frank Pollmann,5 and Yong Baek Kim2, 6, 7

1Max-Planck-Institut fur Physik komplexer Systeme, 01187 Dresden, Germany2Department of Physics, University of Toronto, Toronto, Ontario M5S 1A7, Canada

3Department of Applied Physics, The University of Tokyo, Hongo, Bunkyo-ku, Tokyo, 113-8656, Japan4JST, PRESTO, Hongo, Bunkyo-ku, Tokyo, 113-8656, Japan5Technische Universitat Munchen, 85747 Garching, Germany

6Canadian Institute for Advanced Research/Quantum Materials Program, Toronto, Ontario MSG 1Z8, Canada7School of Physics, Korea Institute for Advanced Study, Seoul 130-722, Korea

Motivated by recent experiments on α-RuCl3, we investigate a possible quantum spin liquidground state of the honeycomb-lattice spin model with bond-dependent interactions. We considerthe K − Γ model, where K and Γ represent the Kitaev and symmetric-anisotropic interactionsbetween spin-1/2 moments on the honeycomb lattice. Using the infinite density matrix renormal-ization group (iDMRG), we provide compelling evidence for the existence of quantum spin liquidphases in an extended region of the phase diagram. In particular, we use transfer matrix spectra toshow the evolution of two-particle excitations with well-defined two-dimensional dispersion, whichis a strong signature of quantum spin liquid. These results are compared with predictions fromMajorana mean-field theory and used to infer the quasiparticle excitation spectra. Further, we com-pute the dynamical structure factor using finite size cluster computations and show that the resultsresemble the scattering continuum seen in neutron scattering experiments on α-RuCl3. We discussthese results in light of recent and future experiments.

I. INTRODUCTION

One of the hallmarks of quantum spin liquid is theexistence of fractionalized excitations1. While it is gen-erally difficult to detect the dispersion of single-particleexcitations in fractionalized systems, the informationabout two-particle and other multi-particle excitationsis contained in the dynamical spin structure factor mea-sured in inelastic neutron scattering. If the ground stateis a quantum spin liquid, the lower boundary of multi-particle continuum should have a well-defined disper-sion. Many candidate materials for quantum spin liq-uid, however, do not allow such scattering experimentdue to unavailability of large single crystal. Possible ex-periments are, therefore, quite often limited to thermo-dynamic measurements such as specific heat and sus-ceptibility, as well as thermal transport. Hence it isdifficult to identify smoking-gun evidence for quantumspin liquid.

In this context, recent inelastic neutron scatteringexperiments2–4 on α-RuCl3 provide valuable informa-tion about multi-particle continuum in a putative quan-tum spin liquid material5. α-RuCl3 is one of the can-didate materials that support the Kitaev interaction6

between jeff = 1/2 pseudo-spin moments, which is abond-dependent frustrated Ising interaction and arisesfrom the combination of correlation effects and spin-orbit coupling7–11. When only the Kitaev interactionis present, such a model can be exactly solved6 andthe ground state is a quantum spin liquid. On theother hand, other interactions are generally present and

they may drive a transition to a magnetically orderedstate12–15.

The compound α-RuCl3 orders magnetically at lowtemperatures16–18, possibly due to the existence of otherinteractions mentioned above. In spite of this, the dy-namical structure factor shows a continuum of excita-tions at high energies both below and above the orderingtemperature, which may be related to a nearby quantumspin liquid19. Elaborate ab initio computations indicatethat the dominant exchange interactions are the KitaevK and symmetric-anisotropic Γ interactions with smallthird neighbor J3 Heisenberg interaction13. The dom-inance of K and Γ interactions is also pointed out instudies of multi-orbital Hubbard model20. Furthermore,it is interesting to notice that the K − Γ model is pro-posed for metal-organic frameworks with Ru3+ or Os3+

ions21. A previous theoretical work22 of exact diago-nalization (ED) on finite size clusters suggest that theK − Γ model may host quantum spin liquid phases inan extended region of the phase diagram and small J3

would drive a transition to the zig-zag order that is seenin the experiment. Such a study is naturally subject tofinite size effect and it is still difficult to nail down theprecise nature of the ground state. In fact, the ED studymentioned above introduce a spatial anisotropy in theKitaev interaction to avoid strong finite size effect.

In this work, we investigate the K − Γ modelusing then infinite density matrix renormalizationgroup23–25(iDMRG), where the system is placed on aninfinite cylinder. We first study the ground state energy,entanglement entropy, magnetization, and static struc-

arX

iv:1

706.

0990

8v2

[co

nd-m

at.s

tr-e

l] 1

4 Fe

b 20

18

2

ture factor for the K−Γ model. Similar to the previousED study, we find quantum paramagnetic ground statesfor ferro-like Kitaev interaction for an extended regionin the phase diagram. It is also shown that the entan-glement entropy in this region remains relatively high.In comparison, the entanglement entropy of a magneti-cally ordered state for anti-ferro-like Kitaev interactionis quite small.

In order to test the existence of coherent excita-tions, we compute the eigenvalues of the transfer matrix(TM) in the matrix product wavefunction. Using themapping26 between the complex eigenvalues and low-energy excitation spectra as a function of momentum,we identify the lower edge of the multi-particle excita-tion continuum as a function of momentum27. Fromthis, we find that the lower edge of the excitation spec-tra, which also corresponds to the longest correlationlength in the system, moves coherently in momentumspace in a well-defined fashion as a function of Γ/K.This alone tells us that these states are non-trivial andlikely correspond to quantum spin liquid phases. Fur-thermore, the TM spectrum exhibits an emergent sym-metry in its momentum dependence. Assuming that thelowest momentum-dependent eigenvalues correspond tothe lower boundaries of the single- and multi-particlecontinuum in a quantum spin liquid, we compare the re-sults of the transfer matrix computations and the two-particle spectra obtained in the Majorana mean-fieldtheory for the K − Γ model. It is found that the samesymmetry which emerges in the TM spectrum is alsoa property of the Majorana spectrum in the mean-fieldtheory. This suggests that the mean-field picture cancapture some essential features of the two-particle exci-tation spectrum seen in the transfer matrix eigenvalues.

As pointed out in earlier works28–30, the dynamicalspin structure factor in the Kitaev model involves notjust two-particle spectra, but also the flux degrees offreedom. In fact, this makes it difficult to extract the in-formation about the single-particle dispersion from thedynamical spin structure factor as the flux degrees offreedom may still play important roles even in the pres-ence of the additional Γ interaction. In contrast, thetransfer matrix eigenvalues computed here are directlyrelated to the convolution of the single-particle spectra.

In order to make a more direct contact with scat-tering experiment, we compute the dynamical structurefactor on a 24-site cluster. It is found that the low en-ergy dynamical structure factor exhibit a star-like fea-ture in momentum space, just like what is seen in recentneutron scattering experiments3 on α-RuCl3. We alsodemonstrate that these results are consistent with theequal-time spin structure factor computed in the Majo-rana mean-field theory. We have further confirmed thatthe magnetic-field dependence of the magnetization in

0.0 0.2 0.4 0.6 0.8 1.0

φ/π

−0.2

−0.3

−0.4

−0.5

EGS

0

1

2

SEz

xy

FIG. 1. Ground state energy density, EGS , of the K − Γmodel, and the corresponding entanglement entropy, SE ,for a bipartition of the cylinder into two infinite cylinders,as determined using iDMRG for cylinders with circumfer-ence L = 6. Kinks in EGS are indications of first orderphase transitions. Inset shows the cylinder’s three-unit-cellcircumference.

the 24-site cluster is consistent with the experimentalresults15,31–37 on α-RuCl3. Taken together, all of theresults above suggest that α-RuCl3 may be close to aquantum spin liquid described by the K−Γ model withferro-like Kitaev interaction.

The remainder of the paper is organized as follows.We present the model and iDMRG results in section II.In section III we present and discuss the transfer matrix(TM) spectrum. A Majorana based MFT is presentedin section IV, while results of the ED calculation of thedynamic structure factor appear in section V. Details ofthe iDMRG and ED calculations, as well as additionalresults are given in the appendices.

II. iDMRG STUDY OF THE K − Γ MODEL

The K − Γ Hamiltonian is given by

H =∑〈ij〉∈x

KxSxi S

xj + Γx(Syi S

zj + Szi S

yj )

+∑〈ij〉∈y

KySyi S

yj + Γy(Sxi S

zj + Szi S

xj )

+∑〈ij〉∈z

KzSzi S

zj + Γz(S

xi S

yj + Syi S

xj ), (1)

where Sαi are spin-1/2 operators at site i of a honey-comb lattice. Unless otherwise noted, we consider herethe isotropic case, Kα = K,Γα = Γ. Throughout thefollowing, K and Γ are parameterized using φ such thatK = − cosφ and Γ = sinφ. We use the iDMRG methodwith bond dimensions of up to χ = 400 to obtain the

3

0.0 0.2 0.4 0.6 0.8 1.0

φ/π

0.0

2.5

5.0

7.5

10.0

12.5

S(q)

Γ

Γ′

Γ′2Γ′3K

K2

K3

M

0.0 0.2 0.4 0.6 0.8 1.0

φ/π

0

1

〈|~ M|〉

Γ

K3 K2

K

M2 MM3

Γ′

Γ′2

Γ′3

FIG. 2. Staggered magnetization (top) and static spin struc-ture factor (bottom), calculated using iDMRG. Inset: Bril-louin zone with labeled positions of symmetry points.

0 1/3 2/3 1

kx/π, η/π

0.0

0.5

1.0

1.5

2.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

FIG. 3. Transfer matrix spectrum λi in the Kitaev limit(φ→ 0) for L = 6. Plotted is the spectrum E(k) = − log |λ|with each point corresponding to a single eigenvalue λ, whereλ = |λ|eiη and η is identified with the momentum kx alongthe cylinder. ky denotes the transverse momentum obtainedas a quantum number with respect to translation along thecylinder. See App. A for more details.

ground state of this model on a narrow infinite cylin-der with a three unit cell circumference (L = 6). InFig. 1 we plot the ground state energy density, EGS ,as a function of the parameter 0 < φ < π. Startingfrom the ferromagnetic Kitaev limit, φ = 0, EGS evolvessmoothly through the Γ limit, φ = π/2. A discontinu-ity appears at φ ≈ 0.6π and again slightly before theanti-ferromagnetic Kitaev limit, φ = π. The two dis-continuities are associated with a transition into, andout of, a magnetically ordered vortex state, which be-comes an exact product state for φ = 3π/4. This isevident in a plot of the entanglement entropy, SE , alsoin Fig. 1, showing a vanishing SE at this point. No-

0

0.4

0.8

0 1/3 2/3 1

ε c(k)

0

0.4

0.8

0 1/3 2/3 1

Ωmin.(q)

kx/π

(a)

qx/π

(b)

FIG. 4. (a) Single Majorana fermion spectrum εc(k) for theisotropic Kitaev model, Kx = Ky = Kz = −1, on a cylin-der with a three unit cell circumference. (b) Correspondingminimum energy for two-particle excitations, Ωmin.(q).

tice that the entanglement entropy remains as high asthat of the ferro-like Kitaev limit in the entire regionof 0 < φ < 0.6π. Furthermore iDMRG shows a finitestaggered magnetization in 0.6π < φ < 0.96π, as wellas an enhanced spin structure factor, Fig. 2, all con-sistent with a magnetically ordered phase. The mainquestion we want to address here is: what is the natureof the ground state outside of the magnetically orderedstate? The large entanglement and lack of magneticorder suggest that the ground state in this region is oc-cupied by quantum spin liquid phases. However, a smalldiscontinuity at φ ≈ 0.025π in both the entanglemententropy and the spin structure factor, raises the questionwhether there exists a subtle transition between differ-ent kinds of spin liquid phases. To address this issue,and to gain insight into the low energy physics of theK − Γ model as φ is tuned from the Kitaev to Γ limits,we turn to a detailed examination of the transfer ma-trix spectrum, obtained from the ground state matrixproduct state (MPS).

III. TRANSFER MATRIX SPECTRUM

A. The Kitaev limit

We begin by analyzing the transfer matrix spectrum,E(kx, ky), of the pure Kitaev model, shown in Fig. 3,as a function of the momentum along the cylinder, kx.We were also able to resolve the transverse momentumky = 0,±2π/3, which are depicted in the figure by differ-ent colors (see Fig. 7). Here we use the demonstratedcorrespondence26 between the complex eigenvalues ofthe transfer matrix and the excitation spectrum, E(k).Namely, given a TM eigenvalue λi = e−εi+iηi , the cor-responding momentum (along the infinite dimension) isgiven by ki ∼ ηi = argλi, while the corresponding en-ergy is given by Ei ∼ εi = − ln |λi| (see appendix Afor details). The Kitaev model is exactly solvable in

4

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

φ/π = 0.02 φ/π = 0.025

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

φ/π = 0.03 φ/π = 0.05

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

φ/π = 0.10 φ/π = 0.20

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

0 1/3 2/3 1

kx/π, φ/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky = 0

ky = 2/3π

ky = −2/3π

kx/π, η/π kx/π, η/π

φ/π = 0.35 φ/π = 0.50

FIG. 5. Same as Fig. 3 for different φ.

terms of Majorana fermions, and therefore it is possi-ble to readily identify the features in Fig. 3 with theknown Majorana excitations. The most prominent fea-ture of the Majorana spectrum, εc(k), in the Kitaev

model is the existence of two gapless Dirac nodes at thecorners of the Brillouin zone, K = (2π/3,−2π/3) andK ′ = (−2π/3, 2π/3) (see Fig. 4a). A continuum ofexcitations may thus be obtained if multiple Majorana

5

0.0 0.1 0.2 0.3 0.4 0.5

φ/π

−0.30

−0.25

−0.20

−0.15

−0.10

Lcirc = 6

〈Ex〉〈Ey〉〈Ez〉

Lcirc = 12

〈Ex〉〈Ey〉〈Ez〉

FIG. 6. Energy density per bond, as obtained using iDMRG,for systems with a three (L = 6) and six (L = 12) unit cellcircumference.

fermions are excited. Fig. 4b shows the minimum exci-tation energies for the two particle excitation spectrum,as defined by

Ωmin.(q) = mink

(|εc(q− k)|+ |εc(k)|) . (2)

Minima in Ωmin.(qx, qy), as a function of qx, in thetwo particle spectrum, appear at (0, 0), (π/3, 2π/3) and(2π/3,−2π/3), which are consistent with the blue, red,and green pillars shown in Fig. 3a at these momenta.We note, however, that, at least in the pure Kitaevmodel, single-particle excitations seem to appear in theTM spectrum as well (see Appendix A for details).

B. The K − Γ model

Moving away from the exactly solvable Kitaev limit,we now turn to analyze the TM spectrum of the K − Γmodel, Fig. 5, which shows the transfer matrix spec-trum, E(kx, ky), for various values of φ. Similarly to theKitaev limit, minima in the continuum of excitations areclearly identified at (0, 0), (π/3, 2π/3), (2π/3,−2π/3),and (π, 0). All, however, are gapped. This can be un-derstood in the context of Majorana fermions by notingthat the cylindrical geometry breaks the symmetry be-tween x bonds and y, z bonds, which in turn can lead,for Γ > 0, to anisotropic hopping amplitudes, and thegapping out of the fermions. To corroborate this point,Fig. 6 depicts the energy density per bond as a func-tion of φ, displaying that indeed the symmetry betweenbonds is broken for φ > 0.

Several additional minima appear for φ > 0, withtheir momentum position moving as φ is increased.Strikingly, these additional minima seem to obey an un-derlying symmetry, i.e., a considerable number of eigen-values obey E(kx, ky) = E(kx+2π/3, ky−2π/3). For in-

kxky

K

K ′

M

FIG. 7. Allowed momentum cuts for the cylindrical geome-try. Symbols indicate the positions of the soft two-particleexcitations, as deduced from the TM spectrum. • indicatethe leading soft modes at K,K′ and M points. Additionalsoft modes are plotted for φ = 0.03π (N) and φ = 0.2π ().As φ is tuned between these values, we expect that theseminima smoothly move from the squares () to the trian-gles (N).

stance, the φ = 0.1π panel in Fig. 5 has a minimum near(π/6,−2π/3) (green +), which has a symmetric counter-part near (5π/6, 2π/3) (red x), i.e., shifted in momen-tum by (2π/3,−2π/3). An additional counterpart islocated near (π/2, 0) (blue circle), which can be reachedby inversion k→ −k, followed by the same shift in mo-mentum. Interpreting the TM spectrum as being asso-ciated with two-particle excitations, the above symme-try suggests the existence of single-particle excitationswhich, in addition to inversion symmetry ε(−k) = ε(k),obey also ε(k) = ε(k±K), where ±K are the momentaat the Brillouin zone corners, K and K ′, respectively.Figure 7 shows the positions of the soft two-particle ex-citations, for φ = 0.03π and φ = 0.2π, further demon-strating the above symmetry.

In summary, the features of the TM spectrumstrongly indicate that the paramagnetic phase of theK − Γ model harbours coherent excitations commonlyassociated with quantum spin liquids. However, it is dif-ficult to determine the nature of this spin liquid phase,based on the iDMRG data alone. On the one hand, theTM features suggest that in the region 0 < φ < 0.6π theK − Γ model harbours Majorana fermion excitations,sharing basic properties with the ground state of theferromagnetic Kitaev model. On the other hand, theapparent transition at φ = 0.025π may indicate thatthere are two distinct spin liquid phases with a sharptransition between them. In the next section we intro-duce a mean-field approximation, which can be used toelucidate the above results.

6

IV. MAJORANA MEAN-FIELD THEORY

A. Majorana spectrum

Motivated by the iDMRG results of the previous sec-tion, we would like to formulate a Fermionic mean-fieldtheory which closely resembles the exact solution of theKitaev model. Therefore, following Kitaev6, we replacethe spin operators in the Hamiltonian with products ofMajorana fermion operators, 2Sαi → ibαi ci,

H = −∑〈ij〉αβ

Kαβij ibαi b

βj icicj , (3)

where for 〈ij〉 a z-type bond,

Kαβij =

1

4

K α = β = zΓ α 6= β 6= z0 otherwise

. (4)

Similar definitions follow for x and y-type bonds. Here,the Majorana fermion operators are normalized such

that bαi , bβj = 2δijδαβ and ci, cj = 2δij . The phys-ical Hilbert space of the spin Hamiltonian H is thenobtained by projecting the Majorana Hamiltonian Honto the subspace of states |Ψ〉 which obey Di |Ψ〉 ≡bxi b

yi bzi ci |Ψ〉 = |Ψ〉. Within a mean-field approach, we

can approximate H with

HMF =−∑〈ij〉αβ

Bij Kαβij ibαi b

βj−∑〈ij〉

Aij icicj+∑〈ij〉

AijBij ,

(5)where the fields Aij and Bij obey the mean-field self-consistency equations on each bond,

Aij =∑αβ

Kαβij 〈ibαi bβj 〉B , (6)

Bij = 〈icicj〉A . (7)

Given the ground state |Ψ0〉MF of HMF , it is possi-ble to construct an approximate ground state for Hby projection onto the physical Hilbert space, |Ψ0〉 ≈∏i(1 +Di)/2 |Ψ0〉MF .It is straightforward to obtain a uniform, Z2-flux-

free, solution of Equations (6) and (7), in the two-dimensional thermodynamic limit. In the following weuse the convention that in Aij , Bij etc., the subscript iindicates a site on the odd sublattice and j a site on theeven sublattice. Assuming that Aij ≡ A on all bonds,we obtain Bij ≡ B = 0.5248, which is independent ofA. Similarly, A is independent of B, but it does de-pend on the ratio K/Γ. The mean-field ground stateenergy per bond is given by EMF = −AB. By solv-ing HMF , it is possible to obtain the Majorana fermion

-1

0

1

Γ M K Γ-1

0

1

Γ M K Γ

-1

0

1

Γ M K Γ-1

0

1

Γ M K Γ

φ/π = 0

ω

φ/π = 0.2

ω

φ/π = 0.4

ω

φ/π = 0.5

ω

FIG. 8. Band structure of c (red) and b (blue) Majoranafermions, plotted along high symmetry lines of the Brillouinzone, for several values of φ in the two dimensional ther-modynamic limit. The flat bands for φ = 0 are three-folddegenerate; for φ/π = 0.5, the lower energy flat bands aretwo-fold degenerate. When φ/π ∼ 0.2, there is a finite den-sity of zero energy b fermion states, around the K (and Γ)points in the Brillouin zone. At φ/π = 0.4 the b fermionbands are still dispersive, but gapped.

spectrum, shown in Fig. 8 along high symmetry linesin the Brillouin zone, for several values of φ. In the Ki-taev limit, φ = 0, one finds a single dispersing c-fermionband, with Dirac nodes at the K points of the Brillouinzone, and whose band width is set by A = K. Threeflat bands describe the b-fermions, which are localizedon the bonds. For φ = π/2, one obtains a similarlydispersing c-fermion band, with band width A = 4Γ/3,and three flat bands which describe the b-fermions local-ized on the hexagons. When both K and Γ are nonzero,the b-fermions are dispersive, and become gapless in theparameter range 0.15 < φ/π < 0.25.

B. Two-Majorana spectrum

We have suggested that the TM spectrum can beassociated with two-particle continua of fractionalizedexcitations, some of which obey a symmetry relationE(kx, ky) = E(kx + 2π/3, ky − 2π/3). Interestingly, theMajorana mean-field Hamiltonian, Eq. (5), exhibits thissymmetry for the b fermion spectrum, εb(k) = εb(k±K)(see Appendix B for details), inducing the same symme-try in the two-particle spectrum, Ωb+c(q,k) = |ε(q −k)| + |ε(k)|, as well. We note that this symmetry,Ωb+c(q±K,k) = Ωb+c(q,k), holds irrespectively of the

7

0

0.1

0.2

-1 -2/3 -1/3 0 1/3 2/3 1

Ωmin.(q)

qx/π

qy/π = 02/3

−2/3

φ/π = 0.275

FIG. 9. b+c (solid) and c+c (dashed) two-Majorana fermionspectrum plotted along the three momentum cuts allowed ona cylinder with a three unit cell circumference (see fig. 7)

properties of the c-fermion spectrum. In Appendix A,we show that this is consistent with the iDMRG results,which exhibit this symmetry in the TM spectrum evenwhen the K − Γ coupling are anisotropic such that theminima in the c spectrum move away from the K,K ′

points.Thus, the form of HMF may give a good description

of the fractionalized excitations, as probed by iDMRG.However, due to the strongly interacting nature of theK − Γ model, the actual amplitudes Aij and Bij whichshould be used for such a description, as well as thevalue of φ itself, will most likely be very different fromtheir values as determined by MFT. Nevertheless, wemay still compare the MF spectra with the TM spec-tra, demonstrating the usefulness of HMF . To do so westudy the same cylindrical geometry considered aboveusing iDMRG. As in the iDMRG calculation where thecylindrical geometry breaks the symmetry between xand y, z bonds, also here we choose different amplitudesAij , Bij for different bonds. In Fig. 9 we plot the mini-mal energies required to excite two Majorana fermions,as given by

Ωmin.(q) = mink

(|εb,c(q− k)|+ |εc(k)|) , (8)

where εb,c(k) is the Majorana spectrum of HMF . Forfinite anisotropy, εb,c(k) opens a gap at all allowed mo-menta, and consequently, also in Ωmin.(q). Neverthe-less, the K and K ′ points remain soft, as in the TMspectrum. Furthermore, additional soft modes appearat finite Γ in the b+c spectrum, which obeys the symme-try Ωmin.(q±K) = Ωmin.(q). For example, shifting thesolid green curve (qy = −2π/3) in Fig. 9 by qx = 2π/3,yields the solid red curve (qy = 2π/3). By demonstrat-ing this symmetry, we conclude that HMF may give agood description of the low energy excitations of theK − Γ model, as seen in the TM spectrum.

ω ω

ω ω

0

0.2

0.4

0.6

0.8

1

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

φ/π = 0

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

φ/π = 0.2

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

φ/π = 0.3

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

φ/π = 0.5

XKΓ

YΓ∗M

FIG. 10. Intensity plot of the dynamic structure factor,S(q, ω), presented along high symmetry lines of the Bril-louin zone for several values of φ, as obtained using theED method. Pseudo-color indicates relative intensity. In-set: Brillouin zone with labels of symmetry points used inthis figure.

V. DYNAMIC STRUCTURE FACTOR

A. Zero magnetic field

Next, we turn to spectral signatures of the K − Γspin liquid, which can be observed in experiments. Thedynamic structure factor, which is probed in inelasticneutron scattering experiments, is defined as

S(q, ω) =∑j,α

∫dt 〈Sαj (t)Sα0 (t = 0)〉 e−iq·(rj−r0)+iωt.

(9)We calculated S(q, ω) using an ED method for a 24-site cluster, see appendix C for details. Fig. 10 showsS(q, ω) for several values of φ. The first evident fea-ture is the existence of a broad excitation continuumat high frequencies. In the Kitaev limit, Γ = 0, mostof the spectral weight is found at relatively low ener-gies, which is in agreement with literature29,30. As Γ isincreased, the low frequency spectral weight at the Bril-louin zone center (Γ point) is pushed towards higherfrequencies, while a low ω signal remains at the M, Yand K/2 (midway between K and Γ) points. In contrastto the spectra of the Kitaev-Heisenberg model38,39, thespin gap at the Kitaev limit seems to remain finite evenin the presence of large Γ. The difference in momen-tum dependence between high and low frequencies isclearly evident by integrating over different ranges ofω. In Fig. 11 we show the relative intensity of S(q, ω),integrated over low and high frequency ranges. In theKitaev limit, φ = 0, S(q, ω) is rather featureless. As

80<

ω<

0.6

φ/π = 0

0<

ω<

0.6

φ/π = 0

0.6<

ω<

20.6<

ω<

2

φ/π = 0.2φ/π = 0.2 φ/π = 0.3φ/π = 0.3 φ/π = 0.5φ/π = 0.5

FIG. 11. Dynamical structure factor, S(q, ω), integratedover low (top) and high (bottom) frequencies, as obtainedfrom exact diagonalization. The discrete set of peaks in mo-mentum space, resulting from the finite size of the studiedcluster, was broadened in order to improve the visualizationof the obtained pattern. The black lines depict the bound-aries of the first and second Brillouin zones.

φ/π = 0φ/π = 0 φ/π = 0.3φ/π = 0.3 φ/π = 0.4φ/π = 0.4 φ/π = 0.5φ/π = 0.5

FIG. 12. Equal-time structure factor S(q), obtained usingthe Majorana mean-field theory.

Γ is increased, S(q, ω), integrated over a low frequencyrange, shows a star shaped pattern, similar to the pat-tern seen in the α-RuCl3 neutron scattering experimentsat low energies. In contrast, integrating over a range ofhigher frequencies, shows an almost featureless momen-tum dependence even for finite Γ, again, in qualitativeagreement with the experiments.

To calculate the dynamic spin structure factor in thecontext of the Majorana MFT, one must consider Z2

flux excitations with respect to the ground state, sinceeach spin operator inserts a flux28,29. Technically, thisrequires solutions to Eqns. (6) and (7) which go beyondthe uniform ansatz considered here. It is however stillpossible to approximately calculate the equal-time spinstructure factor,

S(q) =∑i,α

〈Sαi Sα0 〉 e−iq·(ri−r0)

≈ 1

4

∑i,α

〈bαi bα0 〉B 〈cic0〉A e−iq·(ri−r0). (10)

Since S(q) =∫

(dω/2π)S(q, ω), and noting that accord-ing to the ED results, most of the dynamic structurefactor signal is concentrated at low frequencies, we ex-

ω ω

ω ω

0

0.2

0.4

0.6

0.8

1

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

h = 0

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

h = 0.1

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

h = 0.2

0

0.5

1

1.5

2

2.5

X K Γ Y Γ∗M Γ

h = 0.4

FIG. 13. Dynamic structure factor for φ/π = 0.2, J3 = 0.05,and several values of the magnetic field h⊥c∗ . Symmetrypoints labeled as in Fig. 10.

0

10

20

30

40

50

0 0.2 0.4 0.6 0.8 1 0

0.2

0.4

0.6

0.8

1

FIG. 14. Magnetization and second derivative of the energyas a function of magnetic field, parallel , h||c∗ , and perpen-dicular, h⊥c∗ , to c∗ = (1, 1, 1).

pect that S(q) resemble the integrated low frequencypatterns of S(q, ω). Indeed, as seen in Fig. 12, themean-field theory reproduces the star shaped patternseen both in experiments and in the ED calculation.

B. J3 terms and finite magnetic field

Zig-zag magnetic ordering, similar to the magneticallyordered state observed in α-RuCl3 at low temperatures,can be stabilized by adding a third neighbor Heisenbergterm, J3

∑α,〈ij〉∈3rdn.n. S

αi S

αj to the Hamiltonian, Eq.

(1). At this level, nearest neighbor Heisenberg terms areless important since microscopic calculations13 suggestthat they are smaller, and J3 is enough to stabilize thezig-zag magnetic order. Furthermore, it is possible to

9

suppress this ordering tendency by applying a magneticfield, −∑iα h

αSαi . This is evident in Fig 13, whichdisplays S(q, ω) for φ/π = 0.2, J3 = 0.05 and severalvalues of the in-plane magnetic field h⊥c∗ ∝ (−1, 1, 0),perpendicular to c∗ = (1, 1, 1), which corresponds tothe out-of-plane direction. With h⊥c∗ = 0 the low ωspectral weight is increased at the M and Y points, butnot at K/2. Notice, however, that most of the zonecenter (Γ point) spectral weight is found at relativelyhigh energies15. This aspect of the spectra is similarto the case with J3 = 0, h = 0, shown in Fig. 10.When h⊥c∗ is increased beyond h⊥c∗ ∼ 0.1, the zonecenter spectral weight shifts towards lower energies, anda continuum of excitations emerges.

Fig. 14 shows magnetization curves as obtained withED, for magnetic fields pointing parallel, h||c∗ , and per-pendicular, h⊥c∗ , to c∗. A peak at h⊥c∗ ∼ 0.1 inthe second derivative of the energy −d2E/dh2

⊥c∗ , in-dicates an apparent transition away from zig-zag or-der. In addition, the magnetization curves displayan easy axis anisotropy, also consistent with α-RuCl3experiments16,17,40. A simple mean-field analysis qual-itatively explains the easy-plane anisotropy as follows:If a ferromagnetically ordered moment ~m is assumedas a classical Weiss field, the mean-field energy Emf isobtained as

Emf/N = −K − 3J3

2(m2

x +m2y +m2

z)

+Γ(mymz +mzmx +mxmy)

= −K + Γ− 3J3

2(m2

x +m2y +m2

z)

2(mx +my +mz)

2, (11)

which is minimized, for finite Γ > 0, when mx + my +mz = 0 is satisfied, i.e., when the magnetic momentsare in-plane.

VI. CONCLUSIONS

In this work, we investigated a spin model with boththe Kitaev (K) and symmetric-anisotropic (Γ) interac-tions on the honeycomb lattice using iDMRG, exact di-agonalization, and Majorana mean-field theory. Thismodel is strongly motivated by recent experiments onα-RuCl3, where K and Γ are likely to be the dominantexchange interactions.

We found strong numerical evidence for the existenceof a quantum spin liquid for arbitrary ratio of Γ/K forferro-like Kitaev interactions in iDMRG. In particular,the entanglement entropy remains very high in this en-tire region while we do not see any sign of magneticorder in iDMRG computations. In contrast, we found

a magnetically ordered state with very small entangle-ment entropy on the antiferro-like Kitaev side. More-over, we demonstrated the existence of coherent two-dimensional multi-particle excitations using the corre-spondence between transfer-matrix eigenvalues and thelower boundary of multi-particle excitation spectrum.The cylinder geometry in iDMRG induces an anisotropyin bond-dependent energy, which is expected to movethe locations in momentum space of low energy excita-tions. We show that this can indeed be seen in thetransfer-matrix spectra. The existence of such two-dimensional coherence excitations without magnetic or-der is a very strong evidence of quantum spin liquid.

In order to make direct connection to neutron scatter-ing experiments, we computed the dynamical structurefactor for the K − Γ model without and with a smallthird neighbor Heisenberg interaction J3 using exact di-agonalization of 24-site cluster. The J3 on top of theK − Γ interactions is shown to drive a transition tothe zig-zag order in agreement with previous numericalstudies13,15,22. Upon introduction of Γ starting from theferro-like Kitaev interaction, the dynamical structurefactor develops the scattering continuum with star-likeintensity profile at low energies, just like what is seen inrecent neutron scattering experiment3. The magneticfield dependence of the dynamical structure factor isalso investigated when J3 is finite such that the groundstate is the zig-zag magnetic order in zero field. Thereis a transition to a paramagnetic state with dominantscattering intensity at the zone center when the exter-nal magnetic field along the honeycomb plane reachesabout 1/10 of the largest exchange interactions, namelyK or Γ. This is seen in the magnetization profile anddynamical structure factor computed in exact diagonal-ization. All of these features are consistent with recentexperimental data.

Further we used the Majorana mean-field theory togain analytical insight in these numerical results. Forexample, we computed single and two-particle excita-tion spectra for the K − Γ model showed they exhibitan emergent symmetry in their momentum dependence,which is also found in the transfer matrix spectrum. Theequal-time structure factor and real-space spin correla-tions computed in the Majorana mean-field theory arealso consistent with main features in the exact diago-nalization results.

Combining all these results together, we conclude thatthe mysterious scattering continuum seen in the neu-tron scattering experiment on α-RuCl3 may come froma nearby quantum spin liquid supported by K−Γ inter-actions. In our numerical computations, the spin liquidphases at finite Γ/K show qualitatively the same behav-ior as the Kitaev spin liquid. We do see, however, a jumpin the bond-dependent energy at some value of Γ/K in

10

iDMRG on cylinder geometry, which causes a small kinkin the entanglement entropy. This can be interpreted asa meta-nematic transition, where the bond-anisotropy(or broken 3-fold rotation symmetry) increases abruptly.Whether such a transition would survive in the 2D limitis not clear to us at present. If it does survive, we wouldneed to consider two possible scenarios. (i) The trans-fer matrix spectra on both sides of the transition sharesome qualitative features, suggesting that they are ac-tually the same spin liquid phase, while the apparenttransition may be interpreted as a Lifshitz transition onthe Fermi surface topology of the underlying quasipar-ticles. (ii) Although the transfer matrix spectra maybe described using Majorana fermions both before andafter the transition, the underlying spin liquid groundstates may be distinct. These questions will have to beaddressed in future theoretical investigations. Furtherexperimental data in external magnetic field would pro-vide additional clues for the validity of the assumptionthat the K − Γ or K − Γ− J3 is a good minimal modelfor α-RuCl3.

ACKNOWLEDGEMENTS

We would like to thank Andrei Catuneanu, Yin-Chen He, Masatoshi Imada, Hae-Young Kee, Young-June Kim, Stephen Nagler, Tsuyoshi Okubo, NataliaPerkins, Roser Valenti and Ruben Verresen for usefuldiscussions. YY further thanks Mitsuaki Kawamura forhis technical support. YY was supported by PRESTO,JST (JPMJPR15NF). The ED computation has partlybeen done using the facilities of the Supercomputer Cen-ter, the Institute for Solid State Physics, the Universityof Tokyo. MG and FP acknowledge support from theGerman Research Foundation (DFG) via SFB 1143 andResearch Unit FOR 1807. GW and YBK are supportedby the NSERC of Canada, Canadian Institute for Ad-vanced Research, and the Center for Quantum Materialsat the University of Toronto.

Appendix A: iDMRG and Transfer Matrix

This appendix is devoted to the infinite Density Ma-trix Renormalization (iDMRG) method and the transfermatrix spectrum. The first part exposes the geometryused in order to apply iDMRG onto a 2D lattice model.We present more in-depth discussion of possible finitesize effects due to the finite circumference. The secondpart introduces the transfer matrix spectrum and itsconnection to the excitation spectrum.

Infinite Density Matrix Renormalization Group. Weuse iDMRG23–25 to study ground state properties of the

K-Γ model. Initially developed for 1D systems, it hasbeen successfully applied on 2D systems by wrappingthe lattice on a cylinder and mapping the cylinder to achain. Furthermore employing translational invarianceenables to study infinite cylinders24,25. Using the cylin-der geometry, one dimension of the lattice is finite andleads to a discretization of the related reciprocal vec-tor. Thus, accessible momenta lie on lines in reciprocalspace. We chose cylinder geometries, i.e. circumferenceLcirc and unit cell, such that the accessible momentumlines go through the gapless nodes of the isotropic Ki-taev spin liquid, that are located at the K-points of thefirst Brillouin zone. The results presented here and inthe main text are obtained using a rhombic unit cell anda narrow cylinder with Lcirc = 6 sites circumference.

We extend the discussion of the main text by consid-ering the average energy 〈Ex,y,z〉 of the x, y and z bondsfor 0 ≤ φ ≤ 0.5π, which is presented in Fig. 6. In thelimit of small Γ almost no anisotropy exists, which indi-cates negligible finite size effects caused by the cylindergeometry. Once Γ increases, the anisotropy raises andreaches max(〈Ey,z〉/〈Ex〉) ≈ 2 near φ = 0.3π. Usingwider cylinders with Lcirc = 12 reduces the anisotropyonly slightly. This suggests, that a Γ-like interactionis highly sensitive to finite size and a cylinder geometryand presumably causing the gapless Dirac nodes to shiftaway from the K-points in reciprocal space.

Transfer Matrix Spectrum. The transfer matrix (TM)of a wave function encoded as an infinite matrix productstate (iMPS) contains full information about the staticcorrelations26. Intuitively, the TM translates the iMPSby a lattice vector along the chain in 1D or the cylinderin 2D. For Hamiltonians with only local interactions, thestatic correlations are related to the spectral gap41, e.g.ξ ∼ 1/∆ for z = 1. This statement has been extendedby Zauner et al.26 to also include momentum, such thatthe length scale of the decay of static correlations with amomentum k gives an upper bound on the spectral gapε(k) at k close to k. Hence, the TM spectrum, a groundstate property, provides information about the positionof the minimal energy excitation within the reciprocalspace. A connection between the TM eigenvalues andthe exact excitation energies can only be made know-ing the Lieb-Robinson velocity42, e.g., from dynamics.Here, the Lieb-Robinson velocity is not known and thusthe quasi-energies Ei = − log λi, where λi are the eigen-values of the TM, are only given up to an overall energyscale of the Hamiltonian.

On the cylinder geometry and If the symmetry upontranslation along the cylinder’s circumference is not bro-ken, the transverse momentum ky is a good quantumnumber. Then, for each ky independently a set of λiexists with a longitudinal momentum kx = arg λi cor-responding to the momentum of minimal energy excita-

11

0 1/3 2/3 1

kx/π, η/π

0.00

0.25

0.50

0.75

1.00

1.25

1.50

E(k

),−

log|λ|

TM

single particle

two particles

four particles

six particles

FIG. 15. Quasi-energies E(k) = − log λi of the regular trans-fer matrix spectrum λi compared to analytical excitation en-ergies for single- (thin dotted), two- (solid), four- (dashed),and six-particle excitations (dash-dotted) of the anisotropicXY-Heisenberg chain with transverse field. The parametersare: γ = 0.1, h = 0.75. The analytical results are scaledby a factor a = 1.73 in order to match the min(Ei) withmin(E2p(k)), where E2p(k) is the lower edge of the two-particle excitation band.

tion.In Fig. 15 we illustrate an exemplary TM spectrum

for the anisotropic XY-Heisenberg chain with transversefield h and anisotropy γ:

H = −J∑i

[(1 + γ)Sxi S

xi+1 + (1− γ)Syi S

yi+1 + hSzi

],

(A1)This model can be solved exactly43–47 and its energyspectrum is known47. The plot compares the TM spec-trum Ei(ki) = − log λi (blue dots) with the analyticalsingle- and multi-particle excitations (lines). The posi-tion ki = arg(λi) of the minimal eigenvalues Ei coincidenicely with the minimum in the excitation bands withan even number of particles. Single particle excitationsare not present in the regular TM.

We provide a second example related to the modelinvestigated in the main text. Taking the limit φ → 0of eq. 1 one obtains the exactly solvable Kitaev modelon a honeycomb lattice6. If the Kitaev couplings areisotropic Kx = Ky = Kz, the model exhibits gaplessDirac nodes at the K-points of the Brillouin zone. Aslong as |Kα| < |Kβ |+ |Kγ | with α, β, γ = x, y, z, theexcitation remain gapless. Once the K’s are tuned awayfrom isotropy, the Dirac node moves and odd-numberedparticle excitations get separated in reciprocal-spacefrom even-numbered. Furthermore, the nodes may leavethe allowed momenta cuts of the cylinder geometry in-troducing an effective gap. In Fig. 16 we present acomparison of the TM spectrum with analytical resultsfor the single- and two-particle spectrum clearly illus-

0 1/3 2/3 1

kx/π, φ/π

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

E(k

),−

log|λ|

ky = 0

ky = 2π/3

ky = −2π/3

FIG. 16. Quasi-energies E(k) of the anisotropic Kitaevmodel with Kx = −1, Ky = −1.2, and Kz = −0.9. Linesdisplay the analytic results of the single-particle (solid) andtwo-particles (dashed) lowest energy excitation6 on the cor-responding momentum cuts: ky = 0 (blue), ky = 2π/3 (red),and ky = −2π/3 (green). The analytic results are scaled bya factor a = 0.38.

trating a correspondence between (kx, ky) of the TMeigenvalues and the minimum of the excitations bands.Remarkably, the TM spectrum recovers single-, three-, etc. particle excitations as well as two-, four-, etc.particle excitations.

The reader will find a more rigorous and detailed ex-planation as well as more examples in Zauner et al.26.

Implementation. We turn now to the technical real-ization of obtaining the momentum resolved TM spec-trum. Let λi be the eigenvalues of the transfer matrixwith the ordering |λ0| > |λ1| ≥ |λ2| ≥ ... . By defi-nition the dominant eigenvalue is |λ0| = 1. Generally,λi are complex and can be decomposed as λi = |λi|eiη.The angle η is connected to the momentum kx along thechain or cylinder. Exploiting the rotational invarianceof the Hamiltonian on the cylinder geometry yields thetransverse momentum ky as will be explained now. Inthe following, we require the iMPS to be in canonicalform48. A translation with a lattice vector along thecircumference keeps the Hamiltonian invariant and assuch ky can be treated as a regular quantum number.We extract ky by computing the dominant eigenvector

Λ0 of the mixed transfer matrix constructed out of theground state iMPS and a iMPS with the translation ap-plied, see also Fig. 17b). We like to remark, that thetranslation along the circumference is simply given by apermutation of sites within a ring. If the iMPS is suffi-ciently converged and the applied translation is indeeda symmetry, then the dominant eigenvalue λ0 of themixed TM is 1. Its eigenvector Λ0 has a diagonal formwith eigenvalues |λi|eiq and q being discretized in steps2π/n, where n is the number of unit cells around the

12

a)T : Λi = λi Λi

b)T T : T Λi = λi Λi

c)Λ0 = δi,je

iqi

qi

qj

FIG. 17. Schematic representation of a) the regular andb) the mixed transfer matrix T with translation T applied

along the circumference. c) Dominant eigenvector Λ0 of T Tdetermines the q quantum numbers associated with eachbond leg.

cylinder. If Schmidt values are degenerate, the diagonalform becomes block diagonal with blocks for each set ofdegenerate Schmidt values. Each block can be diago-nalized separately by a unitary transformation which isthen applied to the non-translated iMPS. The momen-tum quantum number qi are associated with the entriesi along a bond leg in the same way as Schmidt valuesare. The TM connects states i and j with correspond-ing qi and qj , hence the transverse momentum is givenby ky,(i,j) = qj − qi. The ky label of λi can be read offfrom its eigenvector Λi due to the fact, that Λi has onlynon-zero entries with the same change of the quantumnumber qi − qj .

Anisotropic K−Γ model. Here, we provide TM spec-tra in the case of anisotropic K − Γ couplings in orderto strengthen the discussion in the main text, sectionIII B and IV B, about the symmetry of the b-fermions.

We introduce an anisotropy by scaling the couplingparameter Kα,Γα by a factor a

(Kx,Γx)→ (aKx, aΓx)

(Ky,Γy)→ ((3/2− a/2)Ky, (3/2− a/2)Γy)

(Kz,Γz)→ ((3/2− a/2)Kz, (3/2− a/2)Γz) ,

(A2)

with respect to the x-bond. The anisotropy leads to ashift of the minimum in the c-fermion spectrum awayfrom the K-point as is apparent in the MFT spectrumFig. 20. The TM spectra, Fig. 18, do not display such ashift of the eigenvalues at the K-point (green +), neitherfor φ/π = 0.1 nor for 0.35. Furthermore, the symmetryΩmin.(k±K) = Ωmin.(k) is unaffected by the anisotropy,which is consistent with the MFT prediction, App. B,for the b and the b+ c fermion spectrum.

0 1/3 2/3 1

kx/π, η/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky/π = 0.000

ky/π = −0.666

ky/π = 0.667

φ/π = 0.10, a = 1.1

0 1/3 2/3 1

kx/π, η/π

0.0

0.5

1.0

1.5

2.0

2.5

3.0

E(k

),−

log|λ|

ky/π = 0.000

ky/π = −0.666

ky/π = 0.667

kx/π, η/π

φ/π = 0.35, a = 1.1

FIG. 18. Quasi-energies E(k) of the anisotropic KΓ modelfor φ = 0.1 (top) and φ = 0.35 (bottom) with an anisotropya = 1.1 according to eq. A2.

Appendix B: Majorana MFT details

Even before solving equations (6) and (7), it is im-portant to characterize the behavior of the Majoranafermions under any configuration of Aij and Bij . The cioperators describe fermions which move about the wholelattice with hopping amplitudes given by Aij . Simi-larly, the bαi operators describe fermion hopping with

amplitudes BijKαβij . Most importantly, the structure

of Kαβij , given in Eq. (4), separates the bαi fermions

into three independent, uncorrelated sectors. Each sec-tor is associated with one of the three sublattices ofhexagons. Within each sector, the hopping amplitudeis BijΓ around a hexagon of the corresponding sublat-tice, and BijK between neighboring hexagons. WhenΓ = 0, bαi fermions are bound to a bond, while theyare bound to a hexagon when K = 0. This behaviorechoes the macroscopic degeneracies of the parent clas-sical models: in the classical Kitaev model there are amacroscopic number of degenerate ground states whichare related to each other by reversing the sign of a singlespin component for two spins on the same bond. Theclassical Γ-model has a similar macroscopic degeneracy,obtained by reversing the signs of one component of each

13

(a)

-0.1

-0.05

0

0.05

0.1(b)

-0.3-0.2-0.100.10.20.3

FIG. 19. Spin-spin correlations in real space. The open blackcircles indicate site i = 0 and the filled circles indicate i. Thecircle size and color represent the magnitude and value of thecorresponding correlation 〈Sxi Sx0 〉. (a) Mean-field results forφ/π = 0.4. (b) ED results for φ/π = 0.5. In both cases onefinds same spin component correlations only among a subsetof all sites.

of the six spins on one hexagon49. Furthermore, return-ing to the uniform mean-field solution, the b-fermionband structure is similar to the band structure obtainedin a Luttinger-Tisza study of the corresponding classi-cal models, since both are determined by the exchange

matrix Kαβij . In the classical case, the flat bands for

Γ = 0 or K = 0 indicate the existence of a macroscopicnumber of ground states, mentioned above.

Spin correlations in real space, shown in Fig. 19a,reveal a pattern which manifests the separation of bfermions into independent sectors: Szi on site i is cor-related only with a certain subset of Szj ’s on other sitesj. Similar patterns are also obtained with ED, see Fig19b although some differences should be pointed out.According to the MFT, at φ = 0.5 (K = 0), thereare non-zero spin-spin correlation only within the samehexagons since the b-fermions are localized. Thus, bothfinite K and Γ are requried to get longer range corre-lations. Furthermore, the correlation in Fig. 19b don’tdecay very fast since they were obtained using ED on asmall cluster with periodic boundary condition, whereasthe correlations in 19a where calculating assuming aninfinite system. Finally, even among the the subsetof sites which have same spin component correlations,MFT gives nonzero static correlations only between op-posite sublattice sites.

Next we show how the symmetry εb(k±K) emergesfrom the Majorana mean-field Hamiltonian, Eq. (5).We begin by assuming that the mean-field amplitudesBij preserve translational invariance, and inversionsymmetry, but are not necessarily isotropic. We cantherefore write the b-fermion Hamiltonian as

Hb =∑k

∑L,L′=A,B

∑αβ

KαβLL′(k)ibαL(k)bβL′(−k), (B1)

where the Fourier transform of the bMajorana operatorsis

bi =1√N

∑k

bL(k)eik·ri , (B2)

0

0.1

0.2

0 1/3 2/3 1

ε b,c(k)

0

0.1

0.2

0 1/3 2/3 1

Ωmin.(q)

kx/π

qy/π = 02/3

−2/3

φ/π = 0.2625

qx/π

FIG. 20. Left: εc(k) (dashed) and εb(k) (solid) foranisotropic mean-field amplitudes Ax = 1, Ay = 1.1, Az =1.2 and Bx = 0.45, By = 0.5, Bz = 0.55. Right: minimumenergy for b + c excitations. The spectra are plotted alongthe allowed momentum cuts, as in Fig. 7.

and where L = A,B denotes the sublattice of site i.

The matrixKαβLL′(k) is the Fourier transform of BijK

αβij ,

with Kαβij defined in Eq. (4). Thus, the sublattice off-

diagonal block take the form

KAB(k) =

KBxT1 ΓBz ΓByT2

ΓBz KByT2 ΓBxT1

ΓByT2 ΓBxT1 KBz

, (B3)

where Bx,y,z are the anisotropic mean-field amplitudes,and Ti = eik·ai , with a1,2 the primitive lattice vectorsfor the honeycomb lattice. K itself is given by

K(k) =

(KAB(k)

K†AB(k)

). (B4)

Under a shift in momentum k→ k±K, T1 → T1η andT2 → T3η

∗, where η = e2πi/3. Using the phases η wenext observe that

KAB(k) =

1η∗

η

KAB(k±K)

η1η∗

.

(B5)The above relation shows that there exists a unitarytransformation which takes K(k ± K) → K(k), andtherefore, the spectrum at the shifted momentum mustbe the same. As noted in the main text, this sym-metry holds irrespective of the details of the c fermionspectrum, and is therefore always induced in the b + cspectrum as well. Fig. 20 shows the single and twoparticle spectrum with fully anisotropic mean-field cou-plings. The minimum in the c fermions spectrum isclearly shifted away from the K point (2π/3,−2π/3),while the b + c spectrum still exhibits the symmetryΩmin.(k ± K) = Ωmin.(k). Finally, in Fig. 21 we use

14

0

0.1

0.2

-1 -2/3 -1/3 1/3 2/3 1

Ωmin.(q)

0

0.1

0.2

-1 -2/3 -1/3 1/3 2/3 1

Ωmin.(q)

0

0.1

0.2

-1 -2/3 -1/3 1/3 2/3 1

Ωmin.(q)

φ/π = 0.1

qx/π

φ/π = 0.225

qx/π

qy/π = 02/3

−2/3

φ/π = 0.4

qx/π

FIG. 21. Ωmin.(q) for c + c (dashed) and b + c (solid) ex-citations, plotted along the momentum cuts in Fig. 7 fordifferent values of φ.

the mean-field Hamiltonian, Eq. (5), to demonstrate,that the soft modes in the two-particle spectrum movein momentum space, as φ is increased.

Appendix C: Dynamical structure factors and theKrylov subspace method

The dynamical structure factor may be written as

S( ~Q, ω) =∑

α=x,y,z

Sαα( ~Q, ω), (C1)

where

Sαα( ~Q, ω) = − 1

πIm 〈0| Sα−~Q

1

ω + iδ − H + E0

Sα+~Q|0〉 ,(C2)

and~S+~Q = (Sx

+~Q, Sy

+~Q, Sz

+~Q) is the Fourier transform of

the spin operators~Si defined as,

~S~Q = N−1/2

N∑i

~Sie

+i ~Q·~ri . (C3)

The dynamical structure factors are conventionally cal-culated by using the Lanczos algorithm initialized withan excited state Sα

+~Q|0〉50,51. However, naive imple-

mentations of the Lanczos algorithm and continuedfraction50 requires careful examination of the conver-gence of excited states relevant to the spectra to controlthe truncation errors. In contrast, modern Krylov sub-space methods, which extract essence from the Lanc-zos algorithm, offer controlled convergence without ad-ditional numerical costs.

Here, we solve a linear equation by employing a conju-gate gradient (CG) method, instead of explicitly calcu-

lating the resolvent of H in Eq.(C2). The CG methodsfind the solution in a Krylov subspace, as follows. First,by introducing the following two vectors,

|χ(ζ)〉 = (ζ − H)−1Sα+~Q|0〉 , (C4)

|φ〉 = Sα+~Q|0〉 (C5)

we rewrite Sαα( ~Q, ω) as

Sαα( ~Q, ω) = − 1

πIm 〈φ|χ(ω + iδ)〉 . (C6)

To obtain the unknown vector |χ(ζ)〉, we solve the fol-lowing linear equation,

(ζ − H) |χ(ζ)〉 = |φ〉 . (C7)

When the linear dimension of the matrix H, L, is toolarge to store the whole matrix in the memory, the lin-ear equation is solved iteratively, for example, by us-ing the CG methods. At nth iteration, the conjugategradient algorithm initialized with |χ0(ζ)〉 = |φ〉 findsan approximate solution |χn(ζ)〉 within a n-dimensional

Krylov subspace Kn(ζ − H, |φ〉) = span|φ〉 , (ζ −H) |φ〉 , . . . , (ζ−H)n−1 |φ〉. At each steps, the CG-typealgorithms search the approximate solution |χn(ζ)〉 tominimize the 2-norm of the residual vector,

|ρn(ζ)〉 = (ζ − H) |χn(ζ)〉 − |φ〉 . (C8)

We note that one needs to solve Eq.(C7) essentiallyonce at a fixed complex number ζ = ω + iδ to obtain

whole spectrum Sαα( ~Q, ω). Due to the shift invariance

of the Krylov subspace52, namely, Kn(ζ − H, |φ〉) =

Kn(ζ ′ − H, |φ〉) for any complex number ζ ′ 6= ζ, we can

15

obtain |χ(ζ ′)〉 from |χ(ζ)〉 with a numerical complexityof O(L0)52. The Krylov subspace methods utilizing theshift invariance are called the shifted Krylov subspacemethods.

For the calculations of S( ~Q, ω), we employ the shiftedbiconjugate gradient (BiCG) method implemented in anumerical library Kω for the shifted Krylov subspacemethod53. The condition for truncating the shifted

BiCG iteration is set maxω‖ |ρn(ω + iδ)〉 ‖ < 10−4 for

the following calculations with δ = 0.02. The number ofthe iteration steps for satisfying the condition dependson the parameters (φ, J3, and B), and is typically ofthe order of one thousand and at most of the order often thousand.

1 Leon Balents, “Spin liquids in frustrated magnets,” Na-ture 464, 199–208 (2010).

2 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. Moess-ner, D. A. Tennant, D. G. Mandrus, and S. E. Nagler,“Proximate Kitaev quantum spin liquid behaviour in ahoneycomb magnet,” Nature Materials 15, pages 733740(2016).

3 Arnab Banerjee, Jiaqiang Yan, Johannes Knolle, Craig A.Bridges, Matthew B. Stone, Mark D. Lumsden, David G.Mandrus, David A. Tennant, Roderich Moessner, andStephen E. Nagler, “Neutron scattering in the proximatequantum spin liquid α-RuCl3,” Science 356, 1055–1059(2017).

4 Kejing Ran, Jinghui Wang, Wei Wang, Zhao-Yang Dong,Xiao Ren, Song Bao, Shichao Li, Zhen Ma, Yuan Gan,Youtian Zhang, J. T. Park, Guochu Deng, S. Danilkin,Shun-Li Yu, Jian-Xin Li, and Jinsheng Wen, “Spin-WaveExcitations Evidencing the Kitaev Interaction in Sin-gle Crystalline α−RuCl3,” Phys. Rev. Lett. 118, 107203(2017).

5 K. W. Plumb, J. P. Clancy, L. J. Sandilands, V. V.Shankar, Y. F. Hu, K. S. Burch, Hae-Young Kee, andYoung-June Kim, “α−RuCl3: A spin-orbit assisted Mottinsulator on a honeycomb lattice,” Phys. Rev. B 90,041112 (2014).

6 Alexei Kitaev, “Anyons in an exactly solved model andbeyond,” Annals of Physics 321, 2–111 (2006).

7 G. Jackeli and G. Khaliullin, “Mott Insulators in theStrong Spin-Orbit Coupling Limit: From Heisenberg to aQuantum Compass and Kitaev Models,” Phys. Rev. Lett.102, 017205 (2009).

8 William Witczak-Krempa, Gang Chen, Yong Baek Kim,and Leon Balents, “Correlated Quantum Phenomena inthe Strong Spin-Orbit Regime,” Annual Review of Con-densed Matter Physics 5, 57–82 (2014).

9 Jeffrey G. Rau, Eric Kin-Ho Lee, and Hae-Young Kee,“Spin-orbit physics giving rise to novel phases in corre-lated systems: Iridates and related materials,” AnnualReview of Condensed Matter Physics 7, 195–221 (2016).

10 Stephen M. Winter, Alexander A. Tsirlin, MariaDaghofer, Jeroen van den Brink, Yogesh Singh, PhilippGegenwart, and Roser Valenti, “Models and Materialsfor Generalized Kitaev Magnetism,” Journal of Physics:Condensed Matter 29, 493002 (2017).

11 S. Trebst, “Kitaev Materials,” ArXiv e-prints (2017),arXiv:1701.07056 [cond-mat.str-el].

12 Jeffrey G. Rau, Eric Kin-Ho Lee, and Hae-Young Kee,“Generic Spin Model for the Honeycomb Iridates beyondthe Kitaev Limit,” Phys. Rev. Lett. 112, 077204 (2014).

13 Stephen M. Winter, Ying Li, Harald O. Jeschke, andRoser Valentı, “Challenges in design of Kitaev materi-als: Magnetic interactions from competing energy scales,”Phys. Rev. B 93, 214431 (2016).

14 Heung-Sik Kim and Hae-Young Kee, “Crystal structureand magnetism in α-RuCl3: An ab initio study,” Phys.Rev. B 93, 155143 (2016).

15 Stephen M. Winter, Kira Riedl, Andreas Honecker, andRoser Valenti, “Breakdown of magnons in a strongly spin-orbital coupled magnet,” Nature Communications 8, 1152(2017).

16 R. D. Johnson, S. C. Williams, A. A. Haghighirad, J. Sin-gleton, V. Zapf, P. Manuel, I. I. Mazin, Y. Li, H. O.Jeschke, R. Valentı, and R. Coldea, “Monoclinic crystalstructure of α-RuCl3 and the zigzag antiferromagneticground state,” Phys. Rev. B 92, 235119 (2015).

17 J. A. Sears, M. Songvilay, K. W. Plumb, J. P. Clancy,Y. Qiu, Y. Zhao, D. Parshall, and Young-June Kim,“Magnetic order in α-RuCl3: A honeycomb-lattice quan-tum magnet with strong spin-orbit coupling,” Phys. Rev.B 91, 144420 (2015).

18 A. Little, Liang Wu, P. Lampen-Kelley, A. Banerjee,S. Patankar, D. Rees, C. A. Bridges, J.-Q. Yan, D. Man-drus, S. E. Nagler, and J. Orenstein, “AntiferromagneticResonance and Terahertz Continuum in α−RuCl3,” Phys.Rev. Lett. 119, 227201 (2017).

19 M. Gohlke, R. Verresen, R. Moessner, and F. Pollmann,“Dynamics of the Kitaev-Heisenberg Model,” Phys. Rev.Lett. 119, 157203 (2017).

20 Wei Wang, Zhao-Yang Dong, Shun-Li Yu, and Jian-XinLi, “Theoretical investigation of magnetic dynamics inα−RuCl3,” Phys. Rev. B 96, 115103 (2017).

21 Masahiko G. Yamada, Hiroyuki Fujita, and MasakiOshikawa, “Designing Kitaev Spin Liquids in Metal-Organic Frameworks,” Physical Review Letters 119,057202 (2017).

22 Andrei Catuneanu, Youhei Yamaji, Gideon Wachtel, Hae-Young Kee, and Yong Baek Kim, “Realizing quantumspin liquid phases in spin-orbit driven correlated materi-als,” arXiv:1701.07837 [cond-mat] (2017).

23 Steven R. White, “Density matrix formulation for quan-

16

tum renormalization groups,” Phys. Rev. Lett. 69, 2863–2866 (1992).

24 I. P. McCulloch, “Infinite size density matrix renor-malization group, revisited,” arXiv:0804.2509 [cond-mat](2008).

25 Jonas A. Kjall, Michael P. Zaletel, Roger S. K. Mong,Jens H. Bardarson, and Frank Pollmann, “Phase diagramof the anisotropic spin-2 XXZ model: Infinite-system den-sity matrix renormalization group study,” Phys. Rev. B87, 235106 (2013).

26 V. Zauner, D. Draxler, L. Vanderstraeten, M. Degroote,J. Haegeman, M. M. Rams, V. Stojevic, N. Schuch, andF. Verstraete, “Transfer matrices and excitations withmatrix product states,” New J. Phys. 17, 053002 (2015).

27 Yin-Chen He, Michael P. Zaletel, Masaki Oshikawa, andFrank Pollmann, “Signatures of Dirac Cones in a DMRGStudy of the Kagome Heisenberg Model,” Physical Re-view X 7, 031020 (2017).

28 G. Baskaran, Saptarshi Mandal, and R. Shankar, “ExactResults for Spin Dynamics and Fractionalization in theKitaev Model,” Phys. Rev. Lett. 98, 247201 (2007).

29 J. Knolle, D.L. Kovrizhin, J.T. Chalker, and R. Moess-ner, “Dynamics of a Two-Dimensional Quantum SpinLiquid: Signatures of Emergent Majorana Fermions andFluxes,” Phys. Rev. Lett. 112, 207203 (2014).

30 J. Knolle, D. L. Kovrizhin, J. T. Chalker, and R. Moess-ner, “Dynamics of fractionalization in quantum spin liq-uids,” Phys. Rev. B 92, 115127 (2015).

31 Ravi Yadav, Nikolay A. Bogdanov, Vamshi M. Katukuri,Satoshi Nishimoto, Jeroen van den Brink, and Liviu Ho-zoi, “Kitaev exchange and field-induced quantum spin-liquid states in honeycomb α-RuCl3,” Scientific Reports6, 37925 (2016).

32 S.-H. Baek, S.-H. Do, K.-Y. Choi, Y. S. Kwon, A. U. B.Wolter, S. Nishimoto, Jeroen van den Brink, andB. Buchner, “Evidence for a Field-Induced Quantum SpinLiquid in α-RuCl3,” Physical Review Letters 119, 037201(2017).

33 Richard Hentrich, Anja U. B. Wolter, Xenophon Zotos,Wolfram Brenig, Domenic Nowak, Anna Isaeva, ThomasDoert, Arnab Banerjee, Paula Lampen-Kelley, David G.Mandrus, Stephen E. Nagler, Jennifer Sears, Young-JuneKim, Bernd Buchner, and Christian Hess, “Large field-induced gap of Kitaev-Heisenberg paramagnons in α-RuCl3,” arXiv:1703.08623 [cond-mat] (2017).

34 Jiacheng Zheng, Kejing Ran, Tianrun Li, Jinghui Wang,Pengshuai Wang, Bin Liu, Zheng-Xin Liu, B. Normand,Jinsheng Wen, and Weiqiang Yu, “Gapless Spin Excita-tions in the Field-Induced Quantum Spin Liquid Phaseof α−RuCl3,” Phys. Rev. Lett. 119, 227208 (2017).

35 J. A. Sears, Y. Zhao, Z. Xu, J. W. Lynn, and Young-JuneKim, “Phase diagram of α-RuCl3 in an in-plane magneticfield,” Physical Review B 95, 180411 (2017).

36 A. U. B. Wolter, L. T. Corredor, L. Janssen, K. Nenkov,S. Schonecker, S.-H. Do, K.-Y. Choi, R. Albrecht,J. Hunger, T. Doert, M. Vojta, and B. Buchner, “Field-induced quantum criticality in the Kitaev system α-RuCl3,” Physical Review B 96, 041405 (2017).

37 Zhe Wang, S. Reschke, D. Huvonen, S.-H. Do, K.-Y. Choi,

M. Gensch, U. Nagel, T. Room, and A. Loidl, “MagneticExcitations and Continuum of a Possibly Field-InducedQuantum Spin Liquid in α−RuCl3,” Phys. Rev. Lett.119, 227202 (2017).

38 Youhei Yamaji, Takafumi Suzuki, Takuto Yamada, Sei-ichiro Suga, Naoki Kawashima, and Masatoshi Imada,“Clues and criteria for designing a Kitaev spin liquid re-vealed by thermal and spin excitations of the honeycombiridate Na2IrO3,” Phys. Rev. B 93, 174425 (2016).

39 Dorota Gotfryd, Juraj Rusnacko, Krzysztof Wohlfeld,George Jackeli, Ji rı Chaloupka, and Andrzej M. Oles,“Phase diagram and spin correlations of the Kitaev-Heisenberg model: Importance of quantum effects,” Phys.Rev. B 95, 024426 (2017).

40 Lukas Janssen, Eric C. Andrade, and Matthias Vojta,“Magnetization processes of zigzag states on the honey-comb lattice: Identifying spin models for α-RuCl3 andNa2IrO3,” Physical Review B 96, 064430 (2017).

41 M. B. Hastings, “Locality in Quantum and Markov Dy-namics on Lattices and Networks,” Phys. Rev. Lett. 93,140402 (2004).

42 E. Lieb and D. W. Robinson, “The finite group velocity ofquantum spin systems,” Commun.Math. Phys. 28, 251–257 (1972).

43 E. Lieb, T. Schultz, and D. Mattis, “Two soluble mod-els of an antiferromagnetic chain,” Annals of Physics 16,407–466 (1961).

44 S. Katsura, “Statistical Mechanics of the AnisotropicLinear Heisenberg Model,” Phys. Rev. 127, 1508–1518(1962).

45 Eytan Barouch, Barry M. McCoy, and Max Dresden,“Statistical Mechanics of the XY Model. I,” Phys. Rev.A 2, 1075–1092 (1970).

46 Eytan Barouch and Barry M. McCoy, “Statistical Me-chanics of the XY Model. II. Spin-Correlation Functions,”Phys. Rev. A 3, 786–804 (1971).

47 U. Schmidt, “The Excitation Spectrum of the One-dimensional XY-Model in a Magnetic Field,” Z. Physik267, 271–275 (1974).

48 Guifre Vidal, “Efficient Classical Simulation of SlightlyEntangled Quantum Computations,” Phys. Rev. Lett.91, 147902 (2003).

49 Ioannis Rousochatzakis and Natalia B. Perkins, “ClassicalSpin Liquid Instability Driven By Off-Diagonal Exchangein Strong Spin-Orbit Magnets,” Physical Review Letters118, 147204 (2017).

50 E. R. Gagliano and C. A. Balseiro, “Dynamical Propertiesof Quantum Many-Body Systems at Zero Temperature,”Phys. Rev. Lett. 59, 2999–3002 (1987).

51 Elbio Dagotto, “Correlated electrons in high-temperaturesuperconductors,” Rev. Mod. Phys. 66, 763–840 (1994).

52 Andreas Frommer, “Bicgstab (`) for families of shiftedlinear systems,” Computing 70, 87–109 (2003).

53 A numerical library for the shifted Krylov methods,Kω, developed by M. Kawamura based on Refs.54 and55, is available through https://github.com/issp-center-dev/Komega.

54 Tomohiro Sogabe, Takeo Hoshi, Shao-Liang Zhang, andTakeo Fujiwara, “A numerical method for calculating the

17

Green’s function arising from electronic structure theory,”in Frontiers of Computational Science (Springer, 2007)pp. 189–195.

55 Susumu Yamamoto, Tomohiro Sogabe, Takeo Hoshi,

Shao-Liang Zhang, and Takeo Fujiwara, “Shiftedconjugate-orthogonalconjugate-gradient method and itsapplication to double orbital extended Hubbard model,”Journal of the Physical Society of Japan 77, 114713(2008).


Recommended