Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

PHYSICAL REVIEW B 97, 161108(R) (2018)

Rapid Communications

Mechanism for subgap optical conductivity in honeycomb Kitaev materials

Adrien Bolens,* Hosho Katsura, Masao Ogata, and Seiji Miyashita


Department of Physics, University of Tokyo, Hongo, Bunkyo-ku, Tokyo 113-0033, Japan

(Received 13 November 2017; published 13 April 2018)

Motivated by recent terahertz absorption measurements in α-RuCl3 , we develop a theory for the electromagnetic
absorption of materials described by the Kitaev model on the honeycomb lattice. We derive a mechanism for the
polarization operator at second order in the nearest-neighbor hopping Hamiltonian. Using the exact results of
the Kitaev honeycomb model, we then calculate the polarization dynamical correlation function corresponding
to electric dipole transitions in addition to the spin dynamical correlation function corresponding to magnetic
dipole transitions.

DOI: 10.1103/PhysRevB.97.161108

Introduction. In Mott insulators, the electronic charge is attributed to direct coupling to magnetic dipole (MD) moments
localized at each site due to Coulomb repulsion, and the so that there must be a contribution from electric dipole (ED)
low-energy properties are described by the remaining spin transitions. We study the response of low-energy excitations of
and orbital degrees of freedom. Small charge fluctuations Kitaev materials to an electromagnetic field by deriving a new
subsist due to virtual hopping of the electrons and generate the microscopic mechanism. We show that the interplay of Hund’s
effective interaction. The same fluctuations can be responsible coupling, spin-orbit coupling (SOC), and a trigonal crystal field
for a finite effective polarization operator [1–5]. As a result, (CF) distortion results in a finite polarization operator up to
some magnetic systems can respond to an external ac electric second order in the nearest-neighbor hopping term, which is
field in a nontrivial way [6–9]. In the case of the simple different from previous results obtained only at the third order.
single-band Hubbard model, such an effect has been calculated This is an important result as it sheds light on a new way to
at third order in the virtual hopping and is generally predicted derive an electric polarization of pure electronic origin which
only in frustrated lattices [1,2,5]. is potentially relevant for various multiorbital Mott insulators.
In this Rapid Communication, motivated by recent experi- We then calculate the optical conductivity at T = 0 in the
ments of terahertz spectroscopy of α-RuCl3 [10,11], we con- ideal case of a pure Kitaev model by combining analytical
sider the case of Kitaev materials: multiorbital Mott insulators and numerical methods. We thus show that the fractionalized
which are in close proximity to the Kitaev honeycomb model low-energy excitations, although emerging from an effective
[12]. We show that by introducing additional on-site degrees spin Hamiltonian, respond to an external electric field.
of freedom, the restriction to frustrated lattices can be lifted, Model. The Hamiltonian of Kitaev materials has been
and we derive an effective polarization operator on each bond discussed extensively in the literature [36–44]. The nearly
of the lattice. octahedral ligand field strongly splits the eg orbitals from the
The Kitaev honeycomb model is exactly solvable and pos- t2g ones. For d 5 filling, one hole occupies the three t2g orbitals
sesses a quantum spin liquid (QSL) ground state. Its potential per site with an effective angular momentum L = 1. We thus
realization in real materials through the Jackeli-Khaliullin study a tight-binding model for the holes,
mechanism [13] attracted much attention in recent years. Exact
analytical results for spin correlations have been derived for H = Hhop + HSOC + HCF + Hint , (1)
the Kitaev model [14] and used to predict the signatures of
Majorana quasiparticles in inelastic neutron [15–18], Raman which is the sum of the kinetic hopping term, SOC, CF splitting
[18–20], and resonant x-ray scatterings [21]. As yet, potential among the t2g orbitals, and the Coulomb and Hund interactions,
Kitaev materials, such as Na2 IrO3 [13,22–25] and α-RuCl3 respectively.
[26–31], all eventually reach a magnetically ordered state at In the present Rapid Communication, we consider Hamil-
sufficiently low temperatures [23–25,29–31], indicating sig- tonians with the full C3 symmetry (which may be appropriate
nificant deviations from the Kitaev model [32]. Nevertheless, for α-RuCl3 [39,42]). In addition, we only consider nearest-
in the case of α-RuCl3 , experimental observations of a residual neighbor hopping processes.
continuum of excitations have been interpreted as remnants of The Hamiltonians are concisely expressed by using the hole
the Kitaev physics [10,33–35]. operators,
In the terahertz absorption measurement [10], it is argued
† † † † † † †
that the absorption continuum of α-RuCl3 is too strong to be ci = (ci,yz,↑ ,ci,yz,↓ ,ci,xz,↑ ,ci,xz,↓ ,ci,xy,↑ ,ci,xy,↓ ). (2)
 †
The kinetic term is Hhop = − ij  ci (Tij ⊗ I2×2 )cj , where
*
bolens@spin.phys.s.u-tokyo.ac.jp I2×2 is the 2 × 2 identity matrix and Tij ’s are the hopping

2469-9950/2018/97(16)/161108(5) 161108-1 ©2018 American Physical Society


BOLENS, KATSURA, OGATA, AND MIYASHITA PHYSICAL REVIEW B 97, 161108(R) (2018)

êy êx at second order in Hhop is not forbidden, even though the lattice
z is bipartite.
B
y Ax In the atomic limit (t1–4 = 0), the system has exactly one
b 
a c hole per site, i.e., ni = 1 for all sites i, where ni = 6α=1 niα (α
êz
labels the six t2g states). The polarization operator measures
x
f
e
d y deviation from this configuration and is defined as P =
the
e i ri δni , where δni = ni − 1 and ri is the position of site i.
z Conservation of charge entails i δni = 0. In the following,
we set e = 1.
We find the existence of a finite effective polarization
FIG. 1. One hexagon of the honeycomb lattice. The different bond
operator at the second order in perturbation theory in Hhop
types (x,y,z) are indicated along with their respective unit vectors
êx , êy , and êz . The A and B sublattices are colored in red and blue,
if  and JH are finite. The effective polarization can be
respectively. written as

Peff = (Pij − Pj i )êγ ≡ Pij  êγ , (6)
matrices among the dyz , dxz , and dxy orbitals, ij γ ij γ
⎛ ⎞ ⎛ ⎞
t3 t4 t4 t1 t4 t2 where êγ is the unit vector along the γ bond connecting the
⎜ ⎟ ⎜ ⎟ sites i ∈ A sublattice and j ∈ B sublattice [see Fig. 1] and
T =⎝4
x
t t 1 t 2 ⎠, T = ⎝ 4
y
t t3 t4 ⎠ ,
Pij is given by perturbation theory. We can furthermore use
t4 t2 t1 t2 t4 t1 the symmetry group of the bond ij  to narrow down the
⎛ ⎞ possible terms in Pij  [48]. Due to the octahedral CF, the
t1 t2 t4
⎜ ⎟ symmetry group of a γ bond of the honeycomb lattice is
T = ⎝ 2 t1 t4 ⎠,
z
t (3) {e,i,C2 (γ ),m⊥ (γ )} whose elements are the identity element,
t4 t4 t3 the inversion transformation, the C2 rotation around êγ , and the
where x, y, and z refer the type of the bond considered [see reflection relative to the plane perpendicular to êγ , respectively.
Fig. 1] and t1–4 are the different hopping integrals. The SOC Note that the trigonal CF distortion does not affect the structure
 † of the group. Considering how Sia Sjb êγ (a,b ∈ {x,y,z}) trans-
Hamiltonian is given by HSOC = λ/2 i,a ci (La ⊗ σ a )ci with
forms under the different group elements, Eq. (6) reduces to
λ > 0, where (La )bc = −iabc and σ a are the Pauli matrices.
The C3 symmetric CF splitting of the t2g orbitals corresponds to √
a trigonal distortion along the axis perpendicular to the plane of
 † Peff = 2A [êγ · (Si × Sj )]êγ . (7)
the honeycomb lattices HCF =  i ci [(L · n̂CF )2 ⊗ I2×2 ]ci ij γ
with n̂CF = [111]. The interaction Hamiltonian Hint is the
Kanamori Hamiltonian [45–47] with intraorbital Coulomb In the basis fixed by the octahedral CF (in which √ the
repulsion U , interorbital repulsion U  = U − 2JH , and Hund’s Kitaev Hamiltonian
√ is written), √ê x = (0,1,−1)/ 2, ê y =
coupling JH , (−1,0,1)/ 2, and êz = (1,−1,0)/ 2. Equation (7) is valid for
any general real symmetric hopping matrices which preserve
Hint = U ni,a,↑ ni,a,↓ + (U  − JH ) ni,a,σ ni,b,σ the C3 symmetry. The unitless constant A is calculated at
i,a i,a<b,σ second order in Hhop by using the eigenstates of the 15 × 15
† † two-hole on-site Hamiltonian. In order to obtain an analytical

+ Ui,a =b ni,a,↑ ni,b,↓ − JH ci,a,↑ ci,a,↓ ci,b,↓ ci,b,↑ result, we only keep terms linear in . For t1 = t3 = t4 = 0,
i,a=b we find
† †

+ JH ci,a,↑ ci,a,↓ ci,b,↓ ci,b,↑ . (4) 128(21λ + 8U ) 2
A = t22  J H + O J , (8)
i,a=b 81λ(3λ + 2U )4 H

In the limit λ t 2 /U , it is well known (see, e.g., which scales as t22 JH /(U 3 λ) for U JH , λ . The full
Refs. [39,40]) that at second-order perturbation theory in Hhop , expression (exact in JH ) and the expression including all the
the effective Hamiltonian is the KH  model which includes hopping integrals t1–4 = 0 are included in the Supplemental
the Heisenberg (H) and anisotropic ( and   ) interactions (not Material [49] together with details about the perturbation
shown here) in addition to the Kitaev model (K) described by theory and numerical calculations of A exact in .
γ γ Only a few hopping processes are possible at second order
HK = −4JK Si Sj , (5)
in Hhop (see the Supplemental Material [49]). Even when
ij γ
 = 0 or JH = 0, different allowed processes contribute to
for the effective spins 1/2. The trigonal distortion is usually Pij  , but they interfere destructively, and their contributions
small, and we treat it as a perturbation (||
λ) unless stated overall vanish. The interference is not completely destructive
otherwise. only when both  = 0 and JH = 0.
Polarization. The on-site Hamiltonian breaks the particle- Optical conductivity. The spin dynamics of the pure Ki-
hole symmetry as HSOC and HCF are both antisymmetric under taev model have been investigated thoroughly. However, for
the particle-hole transformation. Therefore, in contrast to the Kitaev materials, additional integrability breaking terms are
single-band Hubbard model [1], a finite polarization operator indispensable. In particular, they explain the magnetic ordering

161108-2
MECHANISM FOR SUBGAP OPTICAL CONDUCTIVITY IN … PHYSICAL REVIEW B 97, 161108(R) (2018)

at low temperatures. In this case, even the calculation of the


spin structure factor becomes a challenge. Very recent works π 4 4
1 5 1 5 5
indicate that the spin dynamics evolve smoothly from the
2 π 6 2 6
results of the pure Kitaev model using numerical [50–52] and 3
parton mean-field methods [53] close to the QSL regime. In
the following, we limit ourselves to the pure Kitaev model to O1 = S2y S3z O2 = S2z S1y O3 = S4z S5y O4 = S6y S5z
calculate the polarization dynamical response of the fraction-
alized excitations, expecting that our results are meaningful FIG. 2. Four different operators Oi appearing in Peff which create
the same given pair of adjacent π fluxes. The shaded hexagons
physically in the putative proximate Kitaev spin liquids. We
represent the π fluxes (plaquettes with Wp = −1).
show that in this limit, the polarization dynamics is remarkably
similar to the spin dynamics. The pure Kitaev limit is obtained
from the electronic Hamiltonian by keeping only the 90◦ metal- In addition to adding a Majorana matter fermion at site i, Ŝia
ligand-metal hoppings (t1 = t3 = t4 = 0). Then, for small trig- changes the bond fermion number of the bond ij a , which
onal CF, the spin Hamiltonian becomes HK + O(). Peff is corresponds to the change ûij a → −ûij a . Therefore, Ŝia adds
itself linear in , therefore we do not need the O() correction one π flux to the two plaquettes sharing the bond ij a , which
in the Hamiltonian to calculate the response at first order in . is also true for Ŝia (t), at all times t since all ûij a ’s are constants
The optical conductivity along the arbitrary in-plane direc- of motion.
tion êα at T = 0 for ω > 0 is We now have a good criterion to identify which
 ∞  Sia (t)Sjb (t)Skc Sld  terms contribute to the optical conductivity.
ω
σ α (ω) = Re dt eiωt P α (t)P α (0) , (9) The expectation value of a product of any operator can be finite
V 0 only if it does not change the flux in any hexagons. This is a
where P(t) = eiHt Pe−iHt , P α = P · êα , and V is the vol- direct consequence of the orthogonality of the subspaces with
ume of the system. In the effective Kitaev model, we can different flux patterns. Therefore, only combinations of four
substitute P α (t)P α (0)Hubbard → Peff α α
(t)Peff (0)Kitaev with spin operators which overall leave the flux in each hexagon
−iHK t unchanged are relevant.
Peff (t) = e iHK t
Peff e where the expectation value is taken
with respect to the ground state of the Kitaev Hamiltonian. For a fixed bond ij , each pair Sia Sjb in Peff changes the
For the calculation of Peff α
(t)Peff α
(0), we need to evaluate fluxes in two adjacent hexagons. There are four different pos-
spin correlations of the form Si (t)Sjb (t)Skc Sld  for pairs of
a sible pairs of hexagons, located around the bond ij . Inversely,
bonds ij  and kl and for a,b,c,d ∈ {x,y,z}. This is rem- a fixed pair of adjacent hexagons is affected by four different
iniscent of Raman scattering in the Kitaev Hamiltonian [19]. operators Sia Sjb . Let us consider the situation depicted in Fig. 2.
However, unlike Raman scattering for which only terms with The four aforementioned operators are labeled O1–4 . The
a = b and c = d are relevant, only terms with a = b and c = d different symmetries of the Kitaev model (mirror symmetries,
terms appear in Peff α α
(t)Peff (0) due to the antisymmetric nature C3 symmetry, and inversion symmetry) leave us with only four
of Peff . Moreover, we must have either a = γij or b = γij and independent correlation functions i (t) = 24 O1 (t)Oi (0) for
similarly either c = γkl or d = γkl . i = 1–4 from which we can calculate the full response,
On each hexagon, a specific product of Pauli matrices Wp = e 2 A2 ω
y y
26 Sa Sbz Scx Sd Sez Sfx (refer to Fig. 1 for site labels) commutes σ (ω) = √ (2˜ 1 − 2˜ 3 − ˜ 2 + ˜ 4 ), (13)
h̄a⊥ 8 3
with the Hamiltonian and has eigenvalues ±1 so that there is ∞
a conserved Z2 flux in each hexagon. Kitaev introduced an where ˜ i (ω) = Re{ 0 dt eiωt i (t)}. The calculated optical
enlarged Hilbert space of Majorana fermions [12] in which conductivity is independent of the direction of êα and therefore
the spin operators read 2Ŝia = i ĉi b̂ia . We use a hat symbol to isotropic on the plane.
indicate that the operators act on the enlarged Hilbert space. Two different matter Hamiltonians are needed to calculate
The Majorana fermions ĉi and b̂ia are the matter and gauge 1–4 (see the Supplemental Material [49]): the flux-free Hamil-
fermions, respectively. The Kitaev Hamiltonian in terms of tonian Ĥ0 (all ûij  = 1’s) and the two-flux Hamiltonian Ĥ ,
Majorana fermions is given by whose π fluxes correspond to those depicted in Fig. 2 (ûij  = 1
expected for û25 = −1). The ground state of the full Kitaev
ĤK = iJK ûij a ĉi ĉj , ûij a = i b̂ia b̂ja , (10) Hamiltonian is in the flux-free sector [54] so that it is given by
ij a the ground-state |M0  of Ĥ0 . Interestingly, the calculation of
where ûij a = ±1 are constants of motions which fix the Z2 the simpler spin-spin dynamical correlation functions requires
flux Wp in each hexagon. For a fixed flux pattern, the remaining the same Hamiltonian Ĥ [14,15], implying that the magnetic
matter Hamiltonian is quadratic and thus solvable. We further dipole and electric dipole transitions take place between the
introduce the bond fermions, which are complex fermions same flux sectors. Using the Lehmann spectral representation,
defined by (i ∈ A, j ∈ B) we generally define the operator,

χ̂ij a = 21 b̂ia + i b̂ja , a = x,y,z. (11) ab (ω) = π M0 |ca |λλ|cb |M0 δ(ω − λ ), (14)
λ
In terms of the bond fermions, the spin operators become
†  where |λ’s are the eigenvectors of Ĥ with energy Eλ and
Ŝia = i ĉi χ̂ij a + χ̂ij a 2 (i ∈ A) λ = Eλ − E0 . We find
† 
= ĉi χ̂ij a − χ̂ij a 2 (i ∈ B). (12) ˜ 1 = 33 , ˜ 2 = − 31 , ˜ 3 = i 34 , ˜ 4 = −i 36 , (15)

161108-3
BOLENS, KATSURA, OGATA, AND MIYASHITA PHYSICAL REVIEW B 97, 161108(R) (2018)

honeycombs and λ is the reduced Compton wavelength. In


the literature, a wide range of values for the different physical
parameters has been reported, resulting in different values for
A. For α-RuCl3 , the ratio R roughly ranges between 0.01
and 10. Figure 3 shows IED in units of e2 A2 /(h̄a⊥ ) and the
corresponding IMD where the ratio R is arbitrarily set to 0.5.
Around ω = 0 , IMD seems to be dominant. However, IMD
is of course independent of A, and its calculated value (with
g = 2) can only account for about 7% of the measured signal
just above the sharp gap in Ref. [10]. With a ratio of R = 10,
the order of magnitude of IED is comparable to the measured
quantity, but the sharp gap at 0 disappears as IMD becomes
negligible.
Discussion. We showed that the complex interplay of the
Hund’s coupling, SOC, and a trigonal CF distortion results
in a nontrivial polarization operator originating from nearest-
neighbor hopping processes, shedding some light on an un-
FIG. 3. Single-particle responses IED 1
= σ (ω) in units of
expected charge fluctuation mechanism in Kitaev materials.
e2 A2 /(h̄a⊥ ) and IMD
1
where we set R = (A/g)2 × (a/λ)2 to 0.5. The
1 3 By calculating the effective polarization operator and its
inset: single- and three-particle responses IED and IED .
dynamical correlation function, we determined the electric
dipole absorption spectrum originating from the pure Kitaev
where the matter fermions are labeled according to Fig. 2. model. This shows that, like other spin liquids with a continuum
The imaginary part of the spin susceptibility can be written of low-energy excitations [6–9], the fractionalized magnetic
as χ  (ω) ∝ ( 22 + 55 − i 25 + i 52 ) [15]. Using complex excitations respond to an external ac electric field. As measured
matter fermions and the appropriate Bogoliubov transforma- in the terahertz absorption measurements of α-RuCl3 [10],
tions, the Hamiltonians become the electric dipole spectral weight is expected to dominate
over the magnetic dipole one. Our results for the optical
Ĥ0 = ωn an† an + E0 , Ĥ = ωn bn† bn + E0 , (16)
conductivity in Fig. 3 are valid for the pure Kitaev model.
n>0 n>0
However, the derived polarization operator (7) is valid even in
1  
where E0 = − 2 n>0 ωn and E0 = − 21 n>0 ωn . Even the effective KH  model (only the coefficient A is affected).
though all the |λ states are defined in the same flux sector, Therefore, we expect the optical response to be modified
we find that the states that contribute to σ (ω) and χ  (ω) are smoothly when introducing integrability breaking terms in
mutually exclusive. This can be explained using symmetries of the proximity of the QSL regime as for the spin structure
the Hamiltonians Ĥ0 and Ĥ (see the Supplemental Material factor [50–53]. Nonetheless, substantial changes in the spin
[49] and Ref. [55]). Hamiltonian, such as a large  term (expected in real materi-
The electric dipole and magnetic dipole response functions als), most probably significantly alter the calculated response
were numerically calculated in systems of sizes up to 82 × 82 [17].
unit cells. As mentioned in Refs. [15,18,19], the leading con- Additionally, other corrections can potentially affect the

tribution comes from the single-particle states |λ = bλ |M  , optical conductivity in real materials, such as longer-range
where |M   is the ground state of H which satisfies bλ |M   = hopping or breaking of the C3 symmetry, which should explain
0. We verified this property by calculating the single- and three- the dependence on the direction of the probing ac field.
particle responses of a 20 × 20 system (see the Supplemental Acknowledgments. A.B. thanks K. Penc and C. Hotta for
Material [49]), shown in the inset of Fig. 3. helpful comments and acknowledges FMSP for the encour-
The electric and magnetic dipole absorption rates scale agement of the present Rapid Communication. H.K. was sup-
as σ (ω) and ω(gμB /c)2 χ  (ω), respectively, where μB is the ported, in part, by JSPS KAKENHI Grants No. JP15K17719
Bohr magneton and g is the effective Landé g factor so and No. JP16H00985. The present Rapid Communication
that we set IED = σ (ω) and IMD = ω(gμB /c)2 χ  (ω). The was supported by the Elements Strategy Initiative Center for
intensities are related by the ratio R = (A/g)2 × (a/λ)2 , where Magnetic Materials (ESICMM) under the outsourcing project
a is the spacing between the transition-metal atoms on the of MEXT.

[1] L. N. Bulaevskii, C. D. Batista, M. V. Mostovoy, and D. I. [4] K. Hwang, S. Bhattacharjee, and Y. B. Kim, New J. Phys. 16,
Khomskii, Phys. Rev. B 78, 024402 (2008). 123009 (2014).
[2] D. Khomskii, J. Phys.: Condens. Matter 22, 164209 (2010). [5] C. D. Batista, S.-Z. Lin, S. Hayami, and Y. Kamiya, Rep. Prog.
[3] Y. Kamiya and C. D. Batista, Phys. Rev. Lett. 108, 097202 Phys. 79, 084504 (2016).
(2012). [6] T.-K. Ng and P. A. Lee, Phys. Rev. Lett. 99, 156402 (2007).

161108-4
MECHANISM FOR SUBGAP OPTICAL CONDUCTIVITY IN … PHYSICAL REVIEW B 97, 161108(R) (2018)

[7] S. Elsässer, D. Wu, M. Dressel, and J. A. Schlueter, Phys. Rev. [30] J. A. Sears, M. Songvilay, K. W. Plumb, J. P. Clancy, Y. Qiu,
B 86, 155150 (2012). Y. Zhao, D. Parshall, and Y.-J. Kim, Phys. Rev. B 91, 144420
[8] D. V. Pilon, C. H. Lui, T.-H. Han, D. Shrekenhamer, A. J. Frenzel, (2015).
W. J. Padilla, Y. S. Lee, and N. Gedik, Phys. Rev. Lett. 111, [31] H. B. Cao, A. Banerjee, J.-Q. Yan, C. A. Bridges, M. D.
127401 (2013). Lumsden, D. G. Mandrus, D. A. Tennant, B. C. Chakoumakos,
[9] A. C. Potter, T. Senthil, and P. A. Lee, Phys. Rev. B 87, 245106 and S. E. Nagler, Phys. Rev. B 93, 134423 (2016).
(2013). [32] H. Matsuura and M. Ogata, J. Phys. Soc. Jpn. 83, 093701
[10] A. Little, L. Wu, P. Lampen-Kelley, A. Banerjee, S. Patankar, D. (2014).
Rees, C. A. Bridges, J.-Q. Yan, D. Mandrus, S. E. Nagler, and [33] L. J. Sandilands, Y. Tian, K. W. Plumb, Y.-J. Kim, and K. S.
J. Orenstein, Phys. Rev. Lett. 119, 227201 (2017). Burch, Phys. Rev. Lett. 114, 147201 (2015).
[11] Z. Wang, S. Reschke, D. Hüvonen, S.-H. Do, K.-Y. Choi, M. [34] A. Banerjee et al., Nature Mater. 15, 733 (2016).
Gensch, U. Nagel, T. Rõõm, and A. Loidl, Phys. Rev. Lett. 119, [35] A. Banerjee, J. Yan, J. Knolle, C. A. Bridges, M. B. Stone, M.
227202 (2017). D. Lumsden, D. G. Mandrus, D. A. Tennant, R. Moessner, and
[12] A. Kitaev, Ann. Phys. (Berlin) 321, 2 (2006). S. E. Nagler, Science 356, 1055 (2017).
[13] G. Jackeli and G. Khaliullin, Phys. Rev. Lett. 102, 017205 [36] J. G. Rau, E. K.-H. Lee, and H.-Y. Kee, Phys. Rev. Lett. 112,
(2009). 077204 (2014).
[14] G. Baskaran, S. Mandal, and R. Shankar, Phys. Rev. Lett. 98, [37] J. G. Rau and H.-Y. Kee, arXiv:1408.4811.
247201 (2007). [38] Y. Sizyuk, C. Price, P. Wölfle, and N. B. Perkins, Phys. Rev. B
[15] J. Knolle, D. L. Kovrizhin, J. T. Chalker, and R. Moessner, Phys. 90, 155126 (2014).
Rev. Lett. 112, 207203 (2014). [39] H.-S. Kim and H.-Y. Kee, Phys. Rev. B 93, 155143 (2016).
[16] J. Knolle, D. L. Kovrizhin, J. T. Chalker, and R. Moessner, Phys. [40] S. M. Winter, Y. Li, H. O. Jeschke, and R. Valenti, Phys. Rev. B
Rev. B 92, 115127 (2015). 93, 214431 (2016).
[17] X.-Y. Song, Y.-Z. You, and L. Balents, Phys. Rev. Lett. 117, [41] R. Yadav, N. A. Bogdanov, V. M. Katukuri, S. Nishimoto, J. van
037209 (2016). den Brink, and L. Hozoi, Sci. Rep. 6, 37925 (2016).
[18] J. Knolle, Dynamics of a Quantum Spin Liquid (Springer, Cham, [42] W. Wang, Z.-Y. Dong, S.-L. Yu, and J.-X. Li, Phys. Rev. B 96,
Switzerland, 2016). 115103 (2017).
[19] J. Knolle, G.-W. Chern, D. L. Kovrizhin, R. Moessner, and N. [43] S. M. Winter, A. A. Tsirlin, M. Daghofer, J. van den Brink, Y.
B. Perkins, Phys. Rev. Lett. 113, 187201 (2014). Singh, P. Gegenwart, and R. Valenti, J. Phys.: Condens. Matter
[20] J. Nasu, J. Knolle, D. L. Kovrizhin, Y. Motome, and R. Moessner, 29, 493002 (2017).
Nat. Phys. 12, 912 (2016). [44] Y. S. Hou, H. J. Xiang, and X. G. Gong, Phys. Rev. B 96, 054410
[21] G. B. Halász, N. B. Perkins, and J. van den Brink, Phys. Rev. (2017).
Lett. 117, 127203 (2016). [45] J. Kanamori, Progr. Theor. Phys. 30, 275 (1963).
[22] J. Chaloupka, G. Jackeli, and G. Khaliullin, Phys. Rev. Lett. 105, [46] A. Georges, L. de’ Medici, and J. Mravlje, Annu. Rev. Condens.
027204 (2010). Matter Phys. 4, 137 (2013).
[23] X. Liu, T. Berlijn, W.-G. Yin, W. Ku, A. Tsvelik, Y.-J. Kim, H. [47] N. B. Perkins, Y. Sizyuk, and P. Wölfle, Phys. Rev. B 89, 035143
Gretarsson, Y. Singh, P. Gegenwart, and J. P. Hill, Phys. Rev. B (2014).
83, 220403 (2011). [48] S. Miyahara and N. Furukawa, Phys. Rev. B 93, 014445
[24] S. K. Choi, R. Coldea, A. N. Kolmogorov, T. Lancaster, I. I. (2016).
Mazin, S. J. Blundell, P. G. Radaelli, Y. Singh, P. Gegenwart, [49] See Supplemental Material at http://link.aps.org/supplemental/
K. R. Choi, S. W. Cheong, P. J. Baker, C. Stock, and J. Taylor, 10.1103/PhysRevB.97.161108 for details of the perturbation
Phys. Rev. Lett. 108, 127204 (2012). theory and for detailed calculations of the dynamical correlation
[25] F. Ye, S. Chi, H. Cao, B. C. Chakoumakos, J. A. Fernandez-Baca, functions with the Kitaev Hamiltonian.
R. Custelcean, T. F. Qi, O. B. Korneta, and G. Cao, Phys. Rev. [50] S. M. Winter, K. Riedl, P. A. Maksimov, A. L. Chernyshev,
B 85, 180403 (2012). A. Honecker, and R. Valenti, Nat. Commun. 8, 1152
[26] I. Pollini, Phys. Rev. B 53, 12769 (1996). (2017).
[27] K. W. Plumb, J. P. Clancy, L. J. Sandilands, V. V. Shankar, Y. [51] M. Gohlke, R. Verresen, R. Moessner, and F. Pollmann, Phys.
F. Hu, K. S. Burch, H.-Y. Kee, and Y.-J. Kim, Phys. Rev. B 90, Rev. Lett. 119, 157203 (2017).
041112 (2014). [52] D. Gotfryd, J. Rusnačko, K. Wohlfeld, G. Jackeli, J. Chaloupka,
[28] H.-S. Kim, Vijay Shankar V., A. Catuneanu, and H.-Y. Kee, Phys. and A. M. Oleś, Phys. Rev. B 95, 024426 (2017).
Rev. B 91, 241110(R) (2015). [53] J. Knolle, S. Bhattacharjee, and R. Moessner, arXiv:1801.03774.
[29] R. D. Johnson, S. C. Williams, A. A. Haghighirad, J. Singleton, [54] E. H. Lieb, Phys. Rev. Lett. 73, 2158 (1994).
V. Zapf, P. Manuel, I. I. Mazin, Y. Li, H. O. Jeschke, R. Valenti, [55] J. Blaizot and G. Ripka, Quantum Theory of Finite Systems (MIT
and R. Coldea, Phys. Rev. B 92, 235119 (2015). Press, Cambridge, MA, 1986).

161108-5

You might also like