Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Letter
  • Published:

A spin–orbital-entangled quantum liquid on a honeycomb lattice

Abstract

The honeycomb lattice is one of the simplest lattice structures. Electrons and spins on this simple lattice, however, often form exotic phases with non-trivial excitations. Massless Dirac fermions can emerge out of itinerant electrons, as demonstrated experimentally in graphene1, and a topological quantum spin liquid with exotic quasiparticles can be realized in spin-1/2 magnets, as proposed theoretically in the Kitaev model2. The quantum spin liquid is a long-sought exotic state of matter, in which interacting spins remain quantum-disordered without spontaneous symmetry breaking3. The Kitaev model describes one example of a quantum spin liquid, and can be solved exactly by introducing two types of Majorana fermion2. Realizing a Kitaev model in the laboratory, however, remains a challenge in materials science. Mott insulators with a honeycomb lattice of spin–orbital-entangled pseudospin-1/2 moments have been proposed4, including the 5d-electron systems α-Na2IrO3 (ref. 5) and α-Li2IrO3 (ref. 6) and the 4d-electron system α-RuCl3 (ref. 7). However, these candidates were found to magnetically order rather than form a liquid at sufficiently low temperatures8,9,10, owing to non-Kitaev interactions6,11,12,13. Here we report a quantum-liquid state of pseudospin-1/2 moments in the 5d-electron honeycomb compound H3LiIr2O6. This iridate does not display magnetic ordering down to 0.05 kelvin, despite an interaction energy of about 100 kelvin. We observe signatures of low-energy fermionic excitations that originate from a small number of spin defects in the nuclear-magnetic-resonance relaxation and the specific heat. We therefore conclude that H3LiIr2O6 is a quantum spin liquid. This result opens the door to finding exotic quasiparticles in a strongly spin–orbit-coupled 5d-electron transition-metal oxide.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Crystal structure and basic physical properties of H3LiIr2O6.
Figure 2: Evidence from NMR spectroscopy of a spin-liquid ground state in H3LiIr2O6 down to 1.0 K.
Figure 3: The Knight shift and relaxation rate for H3LiIr2O6 reveal the intrinsic static susceptibility and spin dynamics.
Figure 4: Low-lying excitations are captured by the specific heat.

Similar content being viewed by others

References

  1. Geim, A. K. & Novoselov, K. S. The rise of graphene. Nat. Mater. 6, 183–191 (2007)

    Article  ADS  CAS  Google Scholar 

  2. Kitaev, A. Anyons in an exactly solved model and beyond. Ann. Phys. 321, 2–111 (2006)

    Article  ADS  MathSciNet  CAS  Google Scholar 

  3. Balents, L. Spin liquids in frustrated magnets. Nature 464, 199–208 (2010)

    Article  ADS  CAS  Google Scholar 

  4. Jackeli, G. & Khaliullin, G. Mott insulators in the strong spin-orbit coupling limit: from Heisenberg to a quantum compass and Kitaev models. Phys. Rev. Lett. 102, 017205 (2009)

    Article  ADS  CAS  Google Scholar 

  5. Singh, Y. & Gegenwart, P. Antiferromagnetic Mott insulating state in single crystals of the honeycomb lattice material Na2IrO3 . Phys. Rev. B 82, 064412 (2010)

    Article  ADS  Google Scholar 

  6. Singh, Y. et al. Relevance of the Heisenberg-Kitaev model for the honeycomb lattice iridates A2IrO3 . Phys. Rev. Lett. 108, 127203 (2012)

    Article  ADS  Google Scholar 

  7. Plumb, K. W. et al. α-RuCl3: a spin-orbit assisted Mott insulator on a honeycomb lattice. Phys. Rev. B 90, 041112(R) (2014)

    Article  ADS  Google Scholar 

  8. Ye, F. et al. Direct evidence of a zigzag spin-chain structure in the honeycomb lattice: a neutron and X-ray diffraction investigation of single-crystal Na2IrO3 . Phys. Rev. B 85, 180403(R) (2012)

    Article  ADS  Google Scholar 

  9. Williams, S. C. et al. Incommensurate counterrotating magnetic order stabilized by Kitaev interactions in the layered honeycomb α-Li2IrO3 . Phys. Rev. B 93, 195158 (2016)

    Article  ADS  Google Scholar 

  10. Johnson, R. D. et al. Monoclinic crystal structure of α-RuCl3 and the zigzag antiferromagnetic ground state. Phys. Rev. B 92, 235119 (2015)

    Article  ADS  Google Scholar 

  11. Chaloupka, J., Jackeli, G. & Khaliullin, G. Kitaev-Heisenberg model on a honeycomb lattice: possible exotic phases in iridium oxides A2IrO3 . Phys. Rev. Lett. 105, 027204 (2010)

    Article  ADS  Google Scholar 

  12. Rau, J. G., Lee, E. K. H. & Kee, H. Y. Generic spin model for the honeycomb iridates beyond the Kitaev limit. Phys. Rev. Lett. 112, 077204 (2014)

    Article  ADS  Google Scholar 

  13. Choi, S. K. et al. Spin waves and revised crystal structure of honeycomb iridate Na2IrO3 . Phys. Rev. Lett. 108, 127204 (2012)

    Article  ADS  CAS  Google Scholar 

  14. Anderson, P. W. Resonating valence bonds: a new kind of insulator? Mater. Res. Bull. 8, 153–160 (1973)

    Article  CAS  Google Scholar 

  15. Shimizu, Y., Miyagawa, K., Kanoda, K., Maesato, M. & Saito, G. Spin liquid state in an organic Mott insulator with a triangular lattice. Phys. Rev. Lett. 91, 107001 (2003)

    Article  ADS  CAS  Google Scholar 

  16. Itou, T., Oyamada, A., Maegawa, S., Tamura, M. & Kato, R. Quantum spin liquid in the spin-1/2 triangular antiferromagnet EtMe3Sb[Pd(dmit)2]2 . Phys. Rev. B 77, 104413 (2008)

    Article  ADS  Google Scholar 

  17. Yamashita, M. et al. Highly mobile gapless excitations in a two-dimensional candidate quantum spin liquid. Science 328, 1246–1248 (2010)

    Article  ADS  CAS  Google Scholar 

  18. Olariu, A. et al. 17O NMR study of the intrinsic magnetic susceptibility and spin dynamics of the quantum kagome antiferromagnet ZnCu3(OH)6Cl2 . Phys. Rev. Lett. 100, 087202 (2008)

    Article  ADS  CAS  Google Scholar 

  19. Han, T.-H. et al. Fractionalized excitations in the spin-liquid state of a kagome-lattice antiferromagnet. Nature 492, 406–410 (2012)

    Article  ADS  CAS  Google Scholar 

  20. Fu, M., Imai, T., Han, T.-H. & Lee, Y. S. Evidence for a gapped spin-liquid ground state in a kagome Heisenberg antiferromagnet. Science 350, 655–658 (2015)

    Article  ADS  CAS  Google Scholar 

  21. Kim, B. J. et al. Phase-sensitive observation of a spin-orbital Mott state in Sr2IrO4 . Science 323, 1329–1332 (2009)

    Article  ADS  CAS  Google Scholar 

  22. Chun, S. H. et al. Direct evidence for dominant bond-directional interactions in a honeycomb lattice iridate Na2IrO3 . Nat. Phys. 11, 462–466 (2015)

    Article  CAS  Google Scholar 

  23. Banerjee, A. et al. Proximate Kitaev quantum spin liquid behaviour in a honeycomb magnet. Nat. Mater. 15, 733–740 (2016)

    Article  ADS  CAS  Google Scholar 

  24. O’Malley, M. J., Woodward, P. M. & Verweij, H. Production and isolation of pH sensing materials by carbonate melt oxidation of iridium and platinum. J. Mater. Chem. 22, 7782–7790 (2012)

    Article  Google Scholar 

  25. Bette, S. et al. Solution of the heavily stacking faulted crystal structure of the honeycomb iridate H3LiIr2O6 . Dalton Trans. 46, 15216–15227 (2017)

    Article  CAS  Google Scholar 

  26. Winter, S. M., Li, Y., Jeschke, H. O. & Valentí, R. Challenges in design of Kitaev materials: magnetic interactions from competing energy scales. Phys. Rev. B 93, 214431 (2016)

    Article  ADS  Google Scholar 

  27. Willans, A. J., Chalker, J. T. & Moessner, R. Site dilution in the Kitaev honeycomb model. Phys. Rev. B 84, 115146 (2011)

    Article  ADS  Google Scholar 

  28. Nasu, J., Udagawa, M. & Motome, Y. Thermal fractionalization of quantum spins in a Kitaev model: temperature-linear specific heat and coherent transport of Majorana fermions. Phys. Rev. B 92, 115122 (2015)

    Article  ADS  Google Scholar 

  29. Yoshitake, J., Nasu, J. & Motome, Y. Fractional spin fluctuations as a precursor of quantum spin liquids: Majorana dynamical mean-field study for the Kitaev model. Phys. Rev. Lett. 117, 157203 (2016)

    Article  ADS  Google Scholar 

  30. Momma, K . & Izumi, F. VESTA: a three-dimensional visualization system for electronic and structural analysis. J. Appl. Cryst. 41, 653–658 (2008)

    Article  CAS  Google Scholar 

  31. Bette, S., Dinnebier, R. D. & Freyer, D. Structure solution and refinement of stacking faulted NiCl(OH). J. Appl. Cryst. 48, 1706–1718 (2015)

    Article  CAS  Google Scholar 

  32. Tsujii, H., Andraka, B., Muttalib, K. A. & Takano, Y. Distributed τ2 effect in relaxation calorimetry. Physica B 329–333, 1552–1553 (2003)

    Article  ADS  Google Scholar 

  33. Mendels, P. et al. Ga NMR study of the local susceptibility in kagomé-based SrCr8Ga4O19: pseudogap and paramagnetic defects. Phys. Rev. Lett. 85, 3496–3499 (2000)

    Article  ADS  CAS  Google Scholar 

  34. Olariu, A. et al. Unconventional dynamics in triangular Heisenberg antiferromagnet NaCrO2 . Phys. Rev. Lett. 97, 167203 (2006)

    Article  ADS  CAS  Google Scholar 

  35. Takeya, H. et al. Spin dynamics and spin freezing behavior in the two-dimensional antiferromagnet NiGa2S4 revealed by Ga-NMR, NQR and μSR measurements. Phys. Rev. B 77, 054429 (2008)

    Article  ADS  Google Scholar 

  36. Khuntia, P. et al. Spin liquid state in the 3D frustrated antiferromagnet PbCuTe2O6: NMR and muon spin relaxation studies. Phys. Rev. Lett. 116, 107203 (2016)

    Article  ADS  CAS  Google Scholar 

  37. Harris, R. K. et al. Further conventions for NMR shielding and chemical shifts (IUPAC recommendations 2008). Pure Appl. Chem. 80, 59–84 (2008)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank Y. Motome, M. Udagawa, R. Valentí, A. Gibbs, Y. B. Kim, A. Smerald and N. Shannon for discussions, and U. Wedig, Y. Ishikuro, T. Nishioka and S. Nakatsuji for experimental support and discussions. This work was partly supported by the Japan Society for the Promotion of Science (JSPS) KAKAENHI (numbers 24224010, 26707018, 15K13523, JP15H05852, JP15K21717 and 17H01140) and the Alexander von Humboldt foundation.

Author information

Authors and Affiliations

Authors

Contributions

T.T. and A.K. prepared the sample and performed the bulk experiments. K.K., R.T. and Y.K. carried out the NMR measurements. Y.M. carried out the low-temperature specific heat measurements. S.B. and R.D. performed structural analysis. G.J. gave theoretical inputs. T.T., K.K., Y.M. and H.T. wrote manuscript and all authors commented on it. H.T. designed and supervised the experiments.

Corresponding author

Correspondence to H. Takagi.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Additional information

Reviewer Information Nature thanks M. Mourigal and S. Todadri for their contribution to the peer review of this work.

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Figure 1 Powder X-ray diffraction pattern.

The measured powder X-ray diffraction pattern of H3LiIr2O6 at room temperature was recorded with Ag Kα1 radiation; selected reflection indices based on the C2/m space group symmetry are also shown. The reflections that are vastly broadened owing to stacking faults are indicated in pink.

Extended Data Figure 2 Temperature and magnetic-field dependence of magnetization at low temperatures.

a, Temperature dependence of magnetization M/B under magnetic fields up to B = 14 T for the sample used in the specific heat measurements shown in Fig. 4. The Curie-like contribution that is absent in K(T) is clearly seen, originating from magnetic defects. The dashed line represents a power-law behaviour with T−1/2 dependence. The susceptibility in the low-field limit appears to follow the T−1/2 dependence better than the conventional Curie–Weiss dependence. At very low temperatures, the Curie-like contribution becomes independent of T for high B fields, implying the saturation of moment from magnetic defects. In the inset, M(T) at B = 7.0 T (green line) is compared with that calculated by integrating ∂M(T, B)/∂T, which we obtained from S(T, B) using the Maxwell relation (red pluses). b, M(B)–B curves at low temperatures. From the offset from the linear magnetization at high fields, we estimate the saturation moment that originates from magnetic defects to be 0.022μB (as indicated by the arrow). This corresponds to 2% of the magnetic defects with g = 2 and a S = 1/2 moment. The Zeeman gap in the density-of-states model in Fig. 4 is 2αμBB with α = 2.9, suggesting a g-factor of 5.8 for the moment that comes from the magnetic defects. Incorporating g = 5.8, the best estimate of the number of magnetic defects with S = 1/2 moments is 0.8%. This estimate is consistent with the estimate from the specific heat in Fig. 4, which indicates an entropy S(T) of 1%–2% of Rln(2) at T = 5 K, where R is the gas constant. c, B1/2 × ∂M(T,B)/∂T shows a scaling with T/B similar to that for C and T1, indicating that the three probes capture the same excitations.

Extended Data Figure 3 Resistivity versus temperature.

The dependence of the resistivity ρ on the insulating temperature T is measured on a polycrystalline pellet of H3LiIr2O6. The inset shows the Arrhenius plot of the same data, indicating the transport activation energy of approximately 0.12 eV.

Extended Data Figure 4 NMR spectral parameters and estimate of the hyperfine coupling constant in H3LiIr2O6.

a, HWHM obtained by performing a Gaussian fit near the top of each peak. The open symbols are for B perpendicular to the honeycomb plane (); the filled symbols are for B parallel to the plane (). The length of the arrow corresponds to a hyperfine field at a Li site when a moment of 0.002μB is placed on the Ir atoms. b, The integrated NMR signal intensities after T1 and T2 corrections, which is proportional to the number of nuclei under observation. c, Bulk magnetic susceptibility χ(T) of aligned powder at 7 T after the subtraction of core diamagnetism, with B parallel (purple crosses) and perpendicular (green pluses) to the honeycomb plane. The solid lines represent the intrinsic susceptibility χi calculated from χ by subtracting the Curie contribution that originates from the impurities and/or defects. The dotted lines are high-temperature (above 250 K) Curie–Weiss fits, which yield a Curie–Weiss temperature and effective moment in a parallel field of and , respectively, and in a perpendicular field of and . NMR Knight shifts K for 7Li measured at 2 T (purple triangles and green diamonds) are superposed for comparisons. d, K for 7Li plotted against χi. The data in a temperature region T > 150 K are well described by the linear relation K(T) = [Ahf/(NAμB)]χi (solid line), from which we determine the isotropic hyperfine coupling constant Ahf = 0.44/μB T.

Extended Data Figure 5 T/B scaling of C and (T1T)−1, and their fitting with the model density of excitations.

B1/2(C/T) versus T/B (filled symbols, left axis) and B1/2(T1T)−1/2 versus T/B (open symbols, right axis) under various magnetic fields. All of the C(T, B) data points (closed symbols) and T1(T, B) data points (open symbols) fall onto the respective universal curves at low T/B, indicating a scaling behaviour. Because , the plot for T1 is another way of representing the same (T/B) scaling as the inset to Fig. 3b. The physical meaning of B1/2(T1T)−1/2 is the Fermi average of D(E) with a renormalization factor B1/2, which is closely related to B1/2(C/T). C/T and (T1T)−1 are given as follows using the standard equations, which express a Fermi averaging of D(E) and D(E)2, respectively:

Here, f(T, E) = 1/{exp[(E/(kBT)] + 1} is Fermi distribution function. The solid and dashed lines indicate B1/2(C/T) and B1/2(T1T)−1/2, respectively, calculated using the above equations for the model for D(E) shown in Fig. 4c. With α = 2.9 and Γ = 4.3 × 108 J−1/2 per Ir atom, the two scaling curves observed experimentally for B1/2(C/T) and B1/2(T1T)−1/2 are well reproduced by the calculations. Ahf = 0.44/μB T and γn/ 2π = 16.54680 MHz T−1 were used as known parameters for 7Li. The calculated B1/2(C/T) and B1/2(T1T)-1/2 show different behaviour at high T/B (greater than about 1), which reflects the different methods of thermal averaging. This difference reasonably accounts for the difference between the two universal curves at high T/B that was observed experimentally.

Extended Data Figure 6 NMR relaxation details.

a, Ratio of for 1H and 7Li, and anisotropy of T1 for 7Li with perpendicular and parallel fields. The error bars are calculated using the errors in the estimate of T1 (see Fig. 3b). b, Exponent β in the stretched-exponential function describing the NMR relaxation curve as a function of temperature. c, Comparison between the relaxation curves of 6Li and 7Li at 10 K and 5 T, yielding a ratio of for 6Li and 7Li of 0.14. P1 is the time period between a saturation pulse and the echo sequence in the relaxation measurements. This value agrees well with the squares of the gyromagnetic ratios. β = 0.6 was used in the calculation.

Extended Data Figure 7 Estimate of the intrinsic specific heat Ci by subtracting the specific heat due to magnetic defects.

Fitting the specific heat C in Extended Data Fig. 5 with the model for D(E) in Fig. 4c gives an estimate of the specific heat due to magnetic defects. The best-fitting function, represented dashed line in Extended Data Fig. 5, is subtracted from the total specific heat C(B, T) shown in Fig. 4a. The residual specific heat ΔC is almost independent of B and gives a measure of intrinsic specific heat Ci. ΔC/T appears to extrapolate almost to zero as the temperature approaches 0 K, within the uncertainty of the estimate of the contribution from magnetic defects. We do not attempt to separate the possible contribution from the bulk spin liquid to ΔC because of the difficulty in estimating the lattice contribution. The inset shows the plot of ΔC/T as a function of T2.

Extended Data Figure 8 NMR spectra of H3LiIr2O6 before and after the alignment of the powder sample.

Before the alignment (top), an asymmetric line shape in obtained (reflecting the anisotropy in the Knight shift K versus K; Fig. 2a), which represents a powder pattern. The middle and bottom curves shown the line shapes obtained after the alignment for B along the honeycomb plane (the magnetic easy plane) and perpendicular to the plane, respectively. Because of the easy-plane anisotropy of H3LiIr2O6, the alignment for the direction perpendicular to the plane is not as complete as that for the direction parallel to the plane.

Extended Data Figure 9 Subtraction of the nuclear Schottky contribution from the raw specific heat data.

a, Open circles represent the raw specific heat Ctot/T at B = 0, 1 T and 3 T. Filled circles represent the non-nuclear specific heat C/T obtained after subtracting the nuclear contribution CNS/T T−3 (dashed lines) from Ctot/T. Filled triangles and pluses indicate the non-nuclear C/T obtained directly from a time-domain measurement, either by heating for a short period of 100–200 s and then measuring the short relaxation for a period of 100–200 s (triangles) or by heating for a long period of 1,500–2,500 s and then measuring only the short relaxation for a period of 100–200 s (pluses). The non-nuclear specific heat C/T at B = 3 T obtained by the three different methods agree reasonably well, which strongly supports the validity of our analyses. The CNS/T T−3 term for B = 0 T is slightly larger than that for B = 1 T. The origin of the extra T−3 contribution observed in the B = 0 data remains elusive. b, Example of the relaxation curve at B = 3 T and T = 124 mK, which indicates the clear separation in the time domain between the fast relaxation due to the non-nuclear contribution and the slow relaxation due to the nuclear contributions. The inset shows the temperature dependence of the slow relaxation time constant at B = 3 T.

Extended Data Table 1 Refined structural parameters of H3LiIr2O6 at 300 K

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kitagawa, K., Takayama, T., Matsumoto, Y. et al. A spin–orbital-entangled quantum liquid on a honeycomb lattice. Nature 554, 341–345 (2018). https://doi.org/10.1038/nature25482

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature25482

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing