Main

The Kitaev model has recently attracted attention as a canonical example of a QSL with emergent fractionalized fermionic excitations2,5. The model is defined for S = 1/2 spins on a honeycomb lattice with anisotropic bond-dependent interactions, as shown in Fig. 1a2. Recent theoretical work—by providing access to properties of excited states—has predicted signs of Kitaev QSLs in the dynamical response at T = 0 (refs 6,7) and in the T dependence of thermodynamic quantities8,9. However, the dynamical properties at finite T have remained a theoretical challenge, as it is necessary to handle quantum and thermal fluctuations simultaneously. Here, by calculating dynamical correlation functions over a wide temperature range, we directly identify signatures of fractionalization in available experimental inelastic light scattering data.

Figure 1: Schematic pictures for the Kitaev model and Raman processes.
figure 1

a, Honeycomb lattice structure. Blue, green and red bonds represent Ising-like interactions between x-, y- and z-components of the S = 1/2 spins, respectively. Incoming and outgoing photons, whose frequencies are ωi and ωf, are also depicted. b,c, Feynman diagrams of the Raman scattering processes that correspond to a creation or annihilation of a pair of matter fermions (process (A)) (b) and a combination of creation and annihilation of the matter fermions (process (B)) (c). In process (A), a photon scattering creates two fermions with energies ɛ1 and ɛ2, hence the Raman shift ω = ωfωi is equal to ɛ1 + ɛ2. In process (B), the scattering creates a fermion with energy ɛ2 and annihilates a fermion with ɛ1 simultaneously, hence, ω is equal to ɛ2ɛ1.

In real materials, Kitaev-type anisotropic interactions may appear through a superexchange process between jeff = 1/2 localized moments in the presence of strong spin–orbit coupling10. Such a situation is believed to be realized in several materials, such as iridates A2IrO3 (A = Li, Na)11,12 and a ruthenium compound α-RuCl3 (refs 4,13,14,15). These materials show magnetic ordering at a low T (10 K), indicating that some exchange interactions coexist with the Kitaev exchange and give rise to the magnetic order instead of the QSL ground state16,17,18,19. Nevertheless, evidence suggests that the Kitaev interaction is predominant (several tens to hundreds of kelvin)15,18,19,20,21,22, which may provide an opportunity to observe the fractional excitations in a quantum paramagnetic state above the transition temperature as a proximity effect of the QSL phase.

In particular, unconventional excitations were observed by polarized Raman scattering in α-RuCl3 (ref. 4). In this material, Néel ordering sets in only at Tc 14 K, whereas there is growing evidence15,22 that the Kitaev interaction is much larger than additional interactions such as the Heisenberg interaction, and hence finite-temperature signatures of the Kitaev QSL are expected to be observed in the paramagnetic state persisting in a broad temperature window above Tc.

The inset of Fig. 2 shows the integrated experimental Raman intensity for α-RuCl3 as a function of temperature4. A background contribution, probably due to phonons, has been identified and subtracted4, as it persists up to very high T much larger than any magnetic scale. In this limit, it can be fitted to standard one-particle scattering which is proportional to n + 1, with n = 1/(eβω − 1) being the Bose distribution function (β = 1/T is the inverse temperature and ω is the energy for bosons). The main panel (red symbols) shows the remaining, presumably dominantly magnetic contribution.

Figure 2: Comparison between the numerical results and the experimental data for α-RuCl3.
figure 2

Main panel: blue circles represent QMC data for a L = 20 cluster for the integrated Raman intensity Imid shown in Fig. 3c. The errors evaluated by the standard deviation of the MC samplings are sufficiently smaller than the symbol size. Red squares are the experimental data in the energy window from 5 to 12.5 meV (ref. 4), from which the non-magnetic background is subtracted (see text). Green dashed lines represent the fitting by aM[1 − f(ɛM)]2 + bM (see caption of Fig. 3). We take J = 10 meV in calculating Imid. Inset: red squares show the experimental raw data and the orange curve indicates the bosonic background. Note that the assignment of the bosonic background is slightly different from that in ref. 4. Details of the fitting procedure are given in Methods.

Most remarkably, the T dependence of the spectral weight up to high temperatures (more than an order of magnitude above Tc), does not follow the bosonic form expected for conventional insulating magnets in which both magnons and phonons obey Bose statistics. It is thus imperative to understand the origin of this anomalous contribution. This will provide a more direct test of the proximity to QSLs than an asymptotic low-T behaviour which is sensitive to the subdominant exchange interactions.

Results

The main panel of Fig. 2 provides a comparison of the T dependence of our theoretical results (blue circles) with the experimental data. The good agreement over a wide temperature range, from just above Tc up to a much higher scale (15Tc), offers compelling evidence that our Kitaev QSL theory correctly identifies the nature of fundamental excitations in the form of fractionalized fermions. This is further reinforced by noticing that the asymptotic two-fermion-scattering form (1 − f)2, with f = 1/(1 + eβɛ) being the Fermi distribution function (ɛ is the energy for fermions), is a good fit of the response. In the following, we outline our calculations and explain how the two-fermion-scattering T-dependence emerges as a result of fractionalization.

We investigate the Raman spectrum at finite T for the Kitaev model using quantum Monte Carlo (QMC) simulations which enable the numerically exact calculation. This directly utilizes the fractionalization of quantum spins into two species of Majorana fermions: itinerant ‘matter’ and localized ‘flux’ fermions (see Methods for details). Crucially, the Raman response is elicited only by the itinerant Majorana fermions7, which allows us to detect their Fermi statistics more directly than in other dynamical responses6. Below we focus on the case of isotropic exchange couplings, Jx = Jy = Jz = J; a small anisotropy plausible in real materials does not alter our main conclusions (see Supplementary Information). The thermodynamic behaviour exhibits two characteristic crossover T-scales originating from fractionalization at T/J 0.012 and T/J 0.38: the former is related to the condensation of flux Majorana fermions, set by the flux gap 0.06J (ref. 2), whereas the latter arises from the formation of matter Majorana fermions at much higher T, set by their bandwidth 1.5J.

Figure 3a shows the QMC data for the Raman spectrum I(ω) at several T. At T = 0, it exhibits ω-linear behaviour in the low-energy region, due to a linear Dirac dispersion of matter Majorana fermions7. With increasing T above T, the low-energy part increases and the ω = 0 contribution becomes nonzero, as shown in the figure for T/J = 0.0375. At higher T, the broad peak in the intermediate energy range at ω/J 1 is suppressed above T T. Indeed, the Raman spectrum at T/J = 0.75 shows no substantial energy dependence for 0 < ω/J 2, as shown in Fig. 3a. For higher T, the intermediate-to-high-energy weight gradually decreases. The T and ω dependence of the Raman spectrum is summarized in Fig. 3b. The result clearly shows that the broad peak structure is slightly shifted to the low-energy side above T and the spectrum becomes featureless above T.

Figure 3: Calculated Raman intensity for finite temperatures.
figure 3

aω dependences of the Raman spectra I(ω) at several T. b, Contour map of the Raman spectrum I(ω) in the Tω plane. c,d, Integrated spectral weights Imid for 0.5 < ω/J < 1.25 (c) and Ilow for 0.0 < ω/J < 0.25 (d), whose energy ranges are indicated by the hatched and shaded areas in Fig. 3a, respectively. Green dashed lines represent fits using aM[1 − f(ɛM)]2 + bM withɛM/J = 0.58 in c, and aLf(ɛL)[1 − f(ɛL)] + bL with ɛL/J = 0.42 in d (see Supplementary Information). Here, f(ɛ) = (1 + eβɛ)−1 is the Fermi distribution function with zero chemical potential. Horizontal dashed-dotted lines represent the values of Ilow and Imid at T = 0 (refs 7,23). The red and blue areas in c and d highlight the contributions to the integrated Raman intensities from the processes (A) and (B), respectively. The errors evaluated by the standard deviation of the MC samplings are sufficiently smaller than the symbol size. Vertical dotted lines indicate two crossover temperatures, T and T.

For further understanding of the T dependence of the Raman spectra, it is helpful to work in a basis of complex matter fermions constructed as a superposition of real Majorana fermions (see Methods). These elementary excitations determine the T-dependence because their occupation (in a fixed background of fluxes) is given by the Fermi distribution function. In detail, one needs to analyse two different processes contributing to Raman scattering23: one consists of creation or annihilation of a pair of fermions (process (A)), with the other a combination of the creation of one fermion and the annihilation of another (process (B)) (see Methods for details). Process (A) is proportional to [1 − f(ɛ1)][1 − f(ɛ2)]δ(ωɛ1ɛ2), where ω is the Raman shift, and ɛ1 and ɛ2 are the energies of fermions (see Fig. 1b). Process (B) is proportional to f(ɛ1)[1 − f(ɛ2)]δ(ω + ɛ1ɛ2) and vanishes at T = 0 due to the absence of matter fermions in the ground state (see Fig. 1c). Because of their different frequency dependence—for example, (A) vanishes for ω → 0 at low T—their distinct T-behaviour can be extracted by looking at different frequency windows.

Figure 3c shows the T dependence of the integrated spectral weight in the intermediate energy window, Imid for 0.5 < ω/J < 1.25 (see the hatched region in Fig. 3a). The same is used in Fig. 2 in accordance with the frequency window for the experimental data with J = 10 meV. We emphasize that the value of J is consistent not only with the spectral width and peak position of the Raman continuum at the lowest T (ref. 4), but also with the inelastic neutron scattering in α-RuCl3 (ref. 15). As shown in Fig. 3c, Imid has a non-monotonic change as a function of T: it grows around T with increasing T, but turns over to decrease above T/J 0.1, yielding the shift of the peak structure in I(ω) to the low-energy side shown in Fig. 3b. Note that the decrease persists up to temperatures much higher than J due to thermal fluctuations of the itinerant Majorana fermions. We also highlight the contributions from the processes (A) and (B) in Fig. 3c. The result clearly indicates that Imid is dominated by the process (A), which supports the scaling with (1 − f)2 (see Supplementary Information).

Meanwhile, the results presented in Fig. 3d covering the low-energy window, Ilow for 0.0 < ω/J < 0.25 (see the shaded region in Fig. 3a), have a different T-dependence. The increase around T is because the Dirac semimetallic dip in the itinerant fermion system is filled in due to thermal fluctuations of the flux fermions9. Moreover, with increasing T, Ilow saturates around the high-T crossover T. As shown in Fig. 3d, above T/J 0.1, Ilow is dominated by the process (B), indicating that the T dependence is well fitted by f(1 − f). However, the intensity Ilow, is one order of magnitude smaller than Imid.

Discussion

The striking T dependence of the Raman intensity observed in experiments can be naturally attributed to the response from fractionalized fermionic Majorana excitations, dominantly from pairs of creation and annihilation of matter fermions. The T dependence is qualitatively different from that of conventional insulating magnets which show bosonic Raman spectra from two-magnon scattering24. It is important to note that here we are dealing with a two-dimensional magnet13,14,15. In one dimension, there is no such crisp distinction between Bose and Fermi statistics, as in the absence of true exchange processes, bosons with hardcore repulsion are rather similar to fermions obeying the Pauli principle; and on the other hand the roles of topology and order in two dimensions are quite distinct from a one-dimensional case25.

We note that the behaviour below and around Tc is non-universal, as indicated by the strong sample dependence of Tc (ref. 13,15). All features connected to the long-range ordering quickly disappear above Tc, and the universal aspect related with the dominant Kitaev interaction appears in a wide T range. In fact, a recent exact diagonalization study presents evidence of fractional excitations above Tc (ref. 26).

The crucial observation here is that the unexpected fermionic contribution is clearly observed over a remarkably wide T range, more than an order of magnitude higher than the transition temperature into the incidental low-temperature Néel order. This approach is distinct from the conventional quest for exotic properties of QSLs, where the experimental hallmark of fermionic excitations has mainly been pursued in asymptotic T behaviour—for example, in the T-linear specific heat for temperatures much lower than the interaction energy. However, the low-T analyses of such thermodynamic quantities are further complicated by the need to distinguish between QSLs, glassy behaviour, spurious order, and other low-energy contributions typified, for example, by nuclear spins. Our finding provides a direct way of identifying QSL behaviour, and in particular, the presence of fermionic excitations. This, we hope, will stimulate further studies of other dynamical quantities in the wide T range15 as well as studies of other candidate materials such as A2IrO3 (A = Li, Na)27.

Methods

Monte Carlo simulation.

The Hamiltonian of the Kitaev model on the honeycomb lattice is given by

where Sj = (Sjx, Sjy, Sjz) represents an S = 1/2 spin on site j, and 〈jkγ stands for a nearest-neighbour γ(=x, y, z) bond shown in Fig. 1a2. By using the Jordan–Wigner transformation and introducing two kinds of Majorana fermions, cj and 28,29, the model is rewritten as

where (jk)γ is the nearest-neighbour pair satisfying j < k on the γ bond, and is a Z2 variable defined on the z bond (r is the label for the bond), which takes ±1. Equation (2) describes free itinerant Majorana fermions coupled to classical Z2 variables {ηr}. Although the configurations of {ηr} are thermally disturbed away from the ground state configuration with all ηr = 1, the thermodynamic behaviour can be obtained by properly sampling {ηr} as follows. As the Hamiltonian for a given configuration of {ηr} is bilinear in terms of c operators, it is easily diagonalized as

Here, we introduce complex matter fermions fλ with the eigenenergies ɛλ(≥0), which are related to c by

where X is introduced so as to diagonalize the Hamiltonian. Then, we evaluate the free energy Ff({ηr}) = −β−1 lnZf({ηr}) for the configuration {ηr}, where ; β = 1/T is the inverse temperature, and we set kB = 1. The thermal average of an operator is given by

where we define and 〈 η = Z−1 Σ { η r } [ ] e - β F f ( { η r } ) , with Z being the partition function of the system. In our calculations, we take the sum over configurations {ηr} in the average 〈 η by performing Monte Carlo (MC) simulations so as to reproduce the distribution e - β F f ( { η r } ) . This admits the quantum MC (QMC) simulation which is free from the sign problem9.

Raman spectrum.

To calculate the Raman spectrum at finite T, we employ the Loudon–Fleury (L–F) approach30,31 by following previous T = 0 studies7,23: the LF operator for the Kitaev model is given by , where εin and εout are the polarization vectors of the incoming and outgoing photons and dγ is the vector connecting sites on a NN γ bond. Using the LF operator, the Raman intensity is given by where and N is the number of sites; l and l′ denote the directions of εin and εout in , respectively. Note that the following relation is satisfied in the isotropic case7:

In terms of the Majorana fermions, the LF operator is described by a bilinear form of c operators as

where B({ηr}) is a Hermitian matrix with pure imaginary elements. Note that is simply given by 1/2∑ jkBjk({ηr})cj(t)ck(t) as all {ηr} commute with the Hamiltonian. It is this property which allows us to evaluate exactly the dynamical correlator of . Using equation (4), we obtain

where Cλλ = ∑ jkBjkXX and Dλλ = ∑ jkBjkXX. By applying Wick’s theorem, we obtain the Raman intensity for a given configuration {ηr} as

where ω > 0. Finally, the thermal average is evaluated as using the QMC simulation.

The terms in equation (9) describe two different Raman processes, which show different T dependences via the Fermi distribution function f(ɛ): the first term corresponds to the process (B) (Fig. 1c) and the second term corresponds to the process (A) (Fig. 1b). Thus, the T dependence of the Raman intensity provides a good indicator of fermionic excitations in Kitaev QSLs.

Following our previous QMC study9, we have performed more than 30,000 MC steps for the measurements after 10,000 MC steps for the thermalization using parallel tempering technique, for N = 2L2 clusters with L = 12 and 20. The Raman intensity Ill(ω) is computed from 3,000 samples during the 30,000 MC steps. The statistical errors are evaluated by the corresponding standard deviation. However, it turns out that the errors are sufficiently smaller than the symbol sizes in the plots in Figs 2, 3c, and 3d, and hence, are not shown.

Details of the fitting to experimental results.

We define the Raman spectrum of the Kitaev model with J = Jx = Jy = Jz as IK(ω, J), which is I(ω) in equation (6) calculated by QMC for the L = 20 cluster. The bosonic background is taken to be proportional to IB(ω) ≡ n(ω) + 1 = eβω/(eβω − 1), as in ref. 4. Then, we assume that the total spectral weight Itotal(ω, J) is given by

where αK and αB are the coefficients to be determined in the fitting procedure below. The integrated intensity for the intermediate energy window 0.5 < ω/J < 1.25 is also introduced as

where IKmid(J) is equivalent to Imid in the main text. On the other hand, the experimental data for the integrated intensity within the same energy window is obtained by

where Iexp(ω) is the Eg component of the Raman spectrum measured for α-RuCl3 (ref. 4). In Fig. 2, we assume J = 10 meV and determine the coefficients αK and αB so as to fit Itotalmid(J, αK, αB) to Iexp(J) in the energy window from 5 to 12.5 meV (see Supplementary Information for the choice of J); we compute Iexp(J) from the experimental data provided by the authors of ref. 4. We denote the optimized coefficients by αK and αB. In the main panel of this figure, the red squares and blue circles represent Iexp(J) − αBIB(J) and αKIK(J), respectively. In the inset, the red squares represent Iexp(J) and the orange line represents αBIB(J).

Data availability.

Experimental Raman spectroscopy data were originally reported in ref. 4. All other data that support the plots within this paper and other findings of this study are available from the corresponding authors upon request.