Skip to main content
NIST Author Manuscripts logoLink to NIST Author Manuscripts
. Author manuscript; available in PMC: 2024 Oct 15.
Published in final edited form as: Nat Phys. 2019;15(10):10.1038/s41567-019-0594-5. doi: 10.1038/s41567-019-0594-5

Field-tunable quantum disordered ground state in the triangular-lattice antiferromagnet NaYbO2

Mitchell M Bordelon 1, Eric Kenney 2, Chunxiao Liu 3, Tom Hogan 4, Lorenzo Posthuma 1, Marzieh Kavand 5, Yuanqi Lyu 5, Mark Sherwin 5, N P Butch 6, Craig Brown 6,7, M J Graf 2, Leon Balents 8, Stephen D Wilson 1,*
PMCID: PMC11474925  NIHMSID: NIHMS1588374  PMID: 39411404

Abstract

Antiferromagnetically coupled S=1/2 spins on an isotropic triangular lattice are the paradigm of frustrated quantum magnetism, but structurally ideal realizations are rare. Here, we investigate NaYbO2, which hosts an ideal triangular lattice of effective Jeff=1/2 moments with no inherent site disorder. No signatures of conventional magnetic order appear down to 50 mK, strongly suggesting a quantum spin liquid ground state. We observe a two-peak specific heat and a nearly quadratic temperature dependence, in agreement with expectations for a two-dimensional Dirac spin liquid. Application of a magnetic field strongly perturbs the quantum disordered ground state and induces a clear transition into a collinear ordered state, consistent with a long-predicted up–up–down structure for a triangular-lattice XXZ Hamiltonian driven by quantum fluctuations. The observation of spin liquid signatures in zero field and quantum-induced ordering in intermediate fields in the same compound demonstrates an intrinsically quantum disordered ground state. We conclude that NaYbO2 is a model, versatile platform for exploring spin liquid physics with full tunability of field and temperature.


Exotic ground states of quantum antiferromagnets are encouraged by the combination of low dimensionality, geometric frustration and inherent anisotropies. Planar triangular lattices are long-sought platforms for stabilizing them17; however, ideal manifestations that do not break crystallographic or exchange symmetries when approaching the quantum regime are rare. The organic compounds κ-(BEDT-TTF)2Cu2(CN)3 (ref. 8) and EtMe3Sb[Pd(dmit)2]2 (ref. 9) are two promising examples of triangular lattices with S=1/2 moments and a dynamically disordered spin ground state. However, S=1/2 inorganic analogues such as Ba3CoSb2O9 (refs. 10,11) and NaTiO2 (refs. 1214) either order magnetically or undergo a lattice deformation when cooling. A roadblock in inorganic systems is the identification of a material with a high crystallographic symmetry, rigid structure and minimal defect mechanisms that also contains magnetic ions possessing strong quantum fluctuations. Ideally, the magnetic ions should reside at high-symmetry positions that preclude antisymmetric Dzyaloshinskii–Moriya exchange from lifting geometric frustration.

As an alternative to S=1/2-based compounds, rare-earth ions with ground state doublets also engender enhanced quantum fluctuations. Recent studies have shown that the spin–orbit entangled Jeff=1/2 moments of Yb3+ ions on a triangular lattice may exhibit a variety of nearly degenerate states1522. Given the appropriate anisotropies and proximity to phase boundaries, spin liquid states are predicted to appear22. Recent experimental studies of one candidate material, YbMgGaO4, report continuum spin dynamics and a spin-liquid-like ground state; however, chemical disorder endemic to the mixed occupancies of magnesium and gallium atoms on the same crystallographic site of this material has clouded interpretation1523. Alternative models propose that the random cation distribution incites exchange disorder, facilitating the formation of a weakly bound spin glass state that experimentally resembles a spin liquid2426. Fully removing this exchange disorder and accessing the physics inherent to an ideal triangular lattice of Jeff=1/2 moments remains a challenge.

The rare-earth moments in compounds of the form NaRO2 (where R are rare-earth ions) are known to form an ideal triangular lattice in the α-NaFeO2 (R-3m) structure. Previous studies suggest that they realize a large degree of magnetic frustration2729, and NaYbO2 in particular stands out as an appealing candidate material. Specifically, the NaYbO2 lattice promotes enhanced exchange through short nearest-neighbour bonds, and the ytterbium (Yb) moments occupy high-symmetry sites that forbid Dzyaloshinskii–Moriya interactions. This, combined with large crystal field splitting between the ground state and first excited doublet20,21, renders this lattice an appealing framework.

Here, we present an investigation of the triangular lattice of Yb moments in NaYbO2. Our data show that the lattice is robust and forms with fully occupied sodium and oxygen sites, minimizing potential chemical/exchange disorder. Furthermore, the structure preserves the ideal equilateral triangular lattice into the quantum regime beyond the point at which the majority of the magnetic entropy has frozen out. Zero-field susceptibility data collected down to 50 mK reveal no signatures of spin freezing or glassiness, and heat capacity data collected over the same temperature range reveal only a broad two-peak structure—a common signature of the onset of short-range correlations in materials thought to host quantum disordered ground states3033. This disordered state is strongly perturbed via the application of a magnetic field that induces antiferromagnetic order consistent with an up–up–down plateau state for the triangular lattice and reflects an underlying XXZ Hamiltonian with enhanced fluctuations due to interlayer frustration. Our data reveal that NaYbO2 hosts an enticing quantum disordered ground state controllable via modest external fields and presents a cleanly tunable platform for exploring order to quantum disorder phase boundaries inherent to the XXZ triangular lattice.

Polycrystalline NaYbO2 was synthesized and characterized via neutron powder diffraction measurements (see Methods for further synthesis and experiment details). Figure 1a shows the structure at 1.6 K, revealing R3m symmetry with fully occupied sodium and oxygen sites. The D3d distorted YbO6 octahedra and bond lengths are illustrated, and a similar YbO6 environment in YbMgGaO4 is known to generate a large 38 meV splitting between the first excited state and the ground state doublet20,21. A comparison of the Yb environments of the two materials is provided in Supplementary Table 2. At low temperatures, the ground state therefore behaves as an isolated Jeff=1/2 Kramers doublet. Nearest-neighbour Yb–Yb distances were refined to 3.3507(1) Å at 300 K, consistent with previous reports2729, and naively support enhanced exchange relative to other frustrated Yb-based compounds.

Fig. 1 |. Crystal structure and magnetic (H, T) phase diagram of NaYbO2.

Fig. 1 |

a, Refined NaYbO2 structure (1.6 K, R3m) contains equilateral triangular layers of D3d YbO6 distorted octahedra separated by 3.346 Å. Sodium cations refine to full occupation, creating a uniform chemical environment surrounding the triangular layers. Purple spheres, Yb atoms; black spheres, Na atoms; brown spheres, O atoms. b, Low-temperature phase boundary between quantum disordered and antiferromagnetic ordered states in NaYbO2, plotted as a function of field and temperature, extracted from a.c. susceptibility and neutron-scattering experiments. The dashed line denotes the boundary of Zeeman-driven quenching of a minority fraction of free Yb moments under field, above which free moments are quenched. These free moments coexist with a quantum disordered ground state. Values in parentheses and error bars indicate one standard deviation. kB is the Boltzmann constant, Hq denotes the onset temperature of the up–up–down ordered state and HZeeman denotes the suppression in magnetic susceptibility observed due to the quenching of a small fraction of free Yb moments.

Characterizing this exchange, magnetic susceptibility (χ) and magnetization data are plotted in Figs. 2 and 3. Below 50 K, the Van Vleck contribution to the susceptibility is negligible. The data from 20 K to 100 K were modelled by Curie–Weiss fits of the form 1χχ0=CTθCW1 (where C is specific heat, T is temperature, χ0 the temperature-independent contribition to the susceptibility and θCW is the Curie–Weiss temperature), shown in Fig. 2a, and yield a local moment of 2.63(8) Bohr magnetons μB with an antiferromagnetic θCW=10.3(8) K. Relative to YbMgGaO4, θCW is substantially enhanced θCW=4K15,16,18,20, consistent with enhanced exchange. The local moment suggests a heightened g-factor, which was validated via electron paramagnetic resonance (EPR) measurements (Fig. 2a inset). A powder-averaged g-factor of gavg=3.03 is implied by the local moment extracted from susceptibility data. Fits to the EPR line shape reveal anisotropic g-factors of gab=3.294(8) and gc=1.726(9) corresponding to the triangular ab plane and c axis, respectively.

Fig. 2 |. Low-field magnetization and magnetic susceptibility data.

Fig. 2 |

a, Low-temperature Curie–Weiss fit to the constant field magnetic susceptibility χd.c. in a temperature range free from Van Vleck contributions from high-energy crystal field doublets (where the majority of trivalent Yb ions are in the Jeff=1/2 ground state). A large mean-field interaction strength of −10.3(8) K with an effective local moment, μeff, of 2.63(19) μB is fit with a temperature-independent χ0=0.0053(3) e.m.u. mol−1 background term. Inset: EPR data collected at 4.2 K fit to anisotropic g-factors of gab=3.294(8) and gc=1.726(9). b, Isothermal magnetization versus field data reaching only 67% of the expected 1.5μB per Yb ion polarized moment under μ0H=9T. c, Temperature and frequency dependence of a.c. magnetic susceptibility χ(T) from 50 mK to 4 K under zero field. d, χ(T) data collected under applied magnetic fields. A minority fraction of free Yb moments are quenched at low temperatures and high fields, resulting in a peak in χ(T), and the downward inflection parameterizing this Zeeman splitting is denoted by orange stars. Inset shows field-subtracted 0T2Tχ(T) data between 1 K and 3 K and a Curie–Weiss fit quantifying the fraction of free Yb moments in the system, as described in the text. Values in parentheses and error bars indicate one standard deviation. a.u., arbitrary units.

Fig. 3 |. High-field magnetic susceptibility and heat capacity data.

Fig. 3 |

a, χ(H) data collected at 330 mK showing the phase boundary (dashed line) between the quantum disordered (QD) ground state and magnetically ordered q=(1/3,1/3,0) state near 3 T. A second transition back into the quantum disordered state or a quantum paramagnetic phase begins at higher fields. b, χ(T) data collected under a series of magnetic fields that traverse the ordered state; 4 T and 5 T χ(T) data illustrate the onset of the ordered phase below 1 K, while 6 T and 7 T data suggest partial re-entry into a disordered magnetic state. c, Specific heat of NaYbO2 measured down to 80 mK under zero field and overplotted with the non-magnetic NaLuO2 analogue. The resulting magnetic entropy ΔSM approaches 95% of Rln(2). d, Values of Cp(T) under varying magnetic fields. The lower peak centred around 1K develops a sharp anomaly at 5 T, indicative of the phase transition into the q=(1/3,1/3,0) state that is suppressed by 9 T. The inset shows the low-temperature portion of the 0TCp(T) data fit to a power law. The resulting fit to T2.04 is shown as a solid yellow line through the data. Error bars denote one standard deviation of the data.

Although the saturated moment for this system is expected to be approximately 1.5μB per Yb ion, M(H) (where M is magnetization and H is magnetic field strength) measurements collected at 2 K up to 9 T (Fig. 2b) were able to polarize Yb moments up to 1μB only, consistent with notable θCW exchange. Figure 2c shows zero-field a.c. susceptibility data collected down to 50 mK. No signatures of freezing, frequency dependence or long-range order are observed. Instead, χ(T) continues to diverge as the sample is cooled, generating an empirical frustration parameter of θCW/0.050 K>200. As we will later argue, this zero-field state is an inherently quantum disordered state dressed by a small fraction of free Yb moments that are quenched in a magnetic field.

Data on χ(T) collected under a variety of H fields are plotted in Fig. 2d. Under small H values, the divergence in χ(T) (the real part of the a.c. susceptibility) is suppressed and a maximum appears. The temperature of this maximum increases with field until μ0H=2T (where μ0 is the permeability of free space) is reached, beyond which χ(T) becomes nearly temperature independent. The inflection in χ(T) increases linearly with H and is plotted in Fig. 1b. This matches the expected Zeeman splitting (ΔE) of isolated Jeff=1/2 moments, ΔE=2μRgavgJeffH, and suggests that χ(T) at μ0H=2T represents the remaining majority of the correlated/bound Yb moments. As an estimate of the fraction of free spins, μ0H=0T data were fit to a Curie–Weiss form after removing the majority response accessed at μ0H=2T. Fits to a Curie–Weiss form between 1 K and 4 K (Fig. 2d) are described by a model of 14.4(6)% free spins with a full moment of 2.63μB and θCW=0.45(4)K. In addition, 2 K M(H) data plotted in Supplementary Fig. 2d were fit to a two-component model of Brillouin-like free spins and exchange-field-bound moments, which yielded a free spin fraction of approximately 7%. These fits roughly parametrize the limits of a free spin fraction in the material and suggest that free spins coexist within a quantum disordered ground state.

At higher fields, the nearly temperature-independent χ(T) at 2 T evolves into an ordered state. Isothermal χ(H) data at 330 mK plotted in Fig. 3a show an increase in the susceptibility as a phase boundary is traversed at 3 T, followed by nearly total suppression of χ(H) at 5 T. For μ0H>5T, χ(H) begins to recover, suggesting a higher-field phase boundary—one marking the quenching of the ordered state as spins are further polarized toward a quantum paramagnetic phase. Data collected across the ordered regime are plotted in Fig. 3b and show a sharp transition below 1 K at μ0H=4T. The probable origin of the enhancement in χ(T) when entering the ordered state at 3 T is the proximity of the quantum critical point associated with the nearby 0 K phase boundary. These quantum fluctuations are suppressed when crossing the finite-temperature phase boundaries away from this point.

To further characterize NaYbO2, heat capacity measurements were performed. Figure 3c shows the zero-field C(T) of both NaYbO2 and a non-magnetic comparator, NaLuO2, plotted from 80 mK to 40 K. Consistent with susceptibility data, no sharp anomaly indicative of the onset of long-range order is observed in NaYbO2. Instead, a broad feature comprising two peaks is apparent—one peak centred near 1 K and the other near 2.5 K. Two peaks in C(T) are predicted in a number of theoretical models for both trian gular34,35 and kagome-based36,37 Heisenberg lattices where a quantum spin liquid state appears. Integrating the magnetic entropy, ΔSM(T), with the lattice contribution subtracted yields an entropy reaching 95% of Rln(2) where R is the ideal gas constant and is consistent with the nominal Jeff=1/2 magnetic doublet of NaYbO2.

When applying a magnetic field, data in Fig. 3d show that the 2.5 K peak in specific heat as a function of temperature, Cp(T), where pressure (p) is constant, shifts upward in temperature, similar to other frustrated magnets; however, under 5 T, a sharp anomaly appears near 1 K and is coincident with the downturn in χ(T) at this field. Under 9 T, this sharp peak broadens and shifts lower in temperature as the system is driven into the disordered state. Integrated under 5 T, the magnetic entropy matches that of 0 T, and the lowest-temperature Cp(T) is strongly suppressed once order is generated. This is consistent with the suppression of low-energy spin fluctuations when entering the ordered state, which return when the high-field quantum paramagnetic phase is approached. Determining the precise form of the zero-field Cp(T) is complicated by a nuclear Schottky feature that dominates below 100 mK (Supplementary Fig. 2c); however, attempts to do so away from this feature yield C(T)T2, as shown in the inset of Fig. 3d.

Low-temperature neutron-scattering measurements were also performed. Figure 4a shows temperature-subtracted (330 mK − 1.5 K) diffraction data and the absence of zero-field magnetic order. Field-subtracted data at 450 mK plotted in Fig. 4b reveal that under 5 T, new superlattice reflections appear at the momentum transfers Q=(1/3,1/3,0),(1/3,1/3,1) and (1/3, 1/3, 3). Given the symmetry constraints of the R3m structure, these either represent a 120° non-collinear spin structure or an up–up–down pattern of spin order. The absence of a reflection at Q=(1/3,1/3,2) suggests that the field-induced order is collinear. Additionally, magnetic intensity appears at the Q=(0,0,3) position, consistent with the two-q or two-propagation vector structure (q=(1/3,1/3,0)+q=(0,0,0)) expected for the equal-moment up–up–down state38. The best fit to this model is shown in Fig. 4c, where spins refine to be oriented nearly parallel to the (1, −1, −1) direction with an ordered moment of 1.36±0.1μB. This value is less than the 1.5μB expected, probably due to the presence of a minority fraction of free moments as well as the influence of remnant fluctuations in the ordered state. Magnetic peaks are resolution-limited with a minimum spin–spin correlation length ξmin=450. Further data collected at 67 mK determine the low-temperature, magnetic field phase boundaries (Supplementary Fig. 4b).

Fig. 4 |. Neutron diffraction and inelastic neutron-scattering data.

Fig. 4 |

a, Temperature-subtracted neutron powder diffraction data (330 mK − 1.5 K) collected under 0 T, showing the absence of low-temperature magnetic order. The red line is a constant fit to the subtracted data. b, Under an applied field of 5 T at 450 mK, new magnetic peaks appear at (1/3, 1/3, z) positions (z=0,1,3), corresponding to an ordering wave vector of q=(1/3,1/3,0). The data were refined by analysing field-subtracted data (5 T − 0 T), which are constrained by the suppressed (1/3, 1/3, 2) reflection. c, The best fit to the 5 T induced magnetic state using the two-q structure q=(1/3,1/3,0)+q=(0,0,0) is generated by a collinear spin structure with Yb moments of 1.36(10)μB. The displayed structure aligns moments approximately along the <1, −1, −1> direction and has six symmetrically equivalent structures generated by three-fold in-plane rotational and mirror symmetries. d, Inelastic neutron-scattering spectrum collected at 67 mK and 0 T. e, Inelastic neutron-scattering spectrum collected at 74 mK and μ0H=5T. f, Linear spin wave calculations showing the powder-averaged S(Q,E) for a two-dimensional triangular lattice of anisotropic Yb3+ moments of NaYbO2 in a 5 T field and three-sublattice ordering. Error bars denote one standard deviation of the data.

Inelastic scattering data plotted in Fig. 4d,e reveal a renormalization of the low-energy spin dynamics when transitioning from the quantum disordered state into the up–up–down phase. The zero-field data show a diffuse spectrum of excitations centred around the (1/3, 1/3, 0) wave vector, and when applying a 5 T field, much of this spectral weight is shifted into the elastic channel and a nearly flat band of excitations centred at 1 meV. Powder-averaged linear spin wave calculations assuming a purely two-dimensional triangular lattice in a 5 T magnetic field reproduce this flat feature, and the simulated dynamic structure factor S(Q,E) is plotted in Fig. 4f. This simulation was generated using nearest-neighbour coupling with a nearly Heisenberg Hamiltonian with a slight easy-plane anisotropy, Jz=0.45 and Jxy=0.51 meV. The subtle downturn at low Q of the 1 meV band requires easy-plane anisotropy, as discussed in the Supplementary Information.

We now discuss the implications of our results. The similar YbO6 octahedra of NaYbO2 and YbMgGaO4 intimate that the local crystal fields and in-plane exchange couplings between Yb ions are comparable; however, the main distinction between the two systems is the much shorter interplane distance in NaYbO2. This suggests that the interlayer coupling is non-negligible, and therefore, a minimal Hamiltonian should include nearest-neighbour bonds within the planes and between neighbouring layers. Based on the structure, a symmetry analysis leads to the following exchange Hamiltonian:

H2d=ijJxySixSjx+SiySjy+JzSizSjz+Jce^ijSie^ijSj+Jczz^e^ij×SiSjz+z^e^ij×SjSizH=ijJxySixSjx+SiySjy+JzSizSjz+Jcf^ijSif^ijSj+Jczf^ijSiSjz+f^ijSjSiz

The first line H2d contains interactions within a triangular layer, and the second between layers. The unit vectors e^ij are oriented along the ij bond, and f^ij is a unit vector along the projection of the ij bond into the ab plane. The in-plane Hamiltonian is identical to that in YbMgGaO4, but is rewritten here (following Iaconis et al.39) in a more physically transparent ‘compass model’ form. The interlayer exchange H in the second line also has a compass-like structure. We expect that this form applies to the full family of delafossite-like antiferromagnets, ARX2 (A = alkali metal, R = rare earth and X = chalcogen ions), with dipolar Kramers doublets on the R site, sharing the space group 166.

Notably, H2d contains a wide range of phase space favouring three types of classical orders: (1) three-sublattice 120° structures, (2) collinear two-sublattice stripe phases and (3) out-of-plane Ising anisotropy with up–up–down structures. Because we do not observe zero-field order, and we expect that interplane exchange is substantial, we infer that the interactions in H should be frustrated by the in-plane order or correlations. Consideration of the coupling between layers J uniquely singles out the three-sublattice 120° structure: to leading order, only this in-plane order allows the staggered magnetization to effectively cancel the exchange field between neighbouring planes (see Supplementary Information)39. This is even true to a large extent also for the anisotropic Jc and Jcz couplings.

Using this deduction that NaYbO2 has 120° correlations, we expect fluctuations among many classically degenerate or nearly degenerate states to strongly suppress order. Furthermore, recent density matrix renormalization group studies of H2d find that for S=1/2 quantum spins, a spin liquid state indeed occurs in a corner of the classically 120° ordered phase space with moderate Jcz coupling40. Consequently, it is plausible that a spin liquid state occurs in NaYbO2, and if so, it is likely to be smoothly connected to the spin liquid of the two-dimensional problem. The optimal spin liquid ground state for the 2d model based on variational parton calculations39 is a U(1) Dirac state with gapless fermionic spinons described theoretically as a 2+1 -dimensional quantum electrodynamics (QED) conformal field theory: QED3. The second implication of our Hamiltonian in this regime is that, on applying a magnetic field, the degeneracy is strongly lifted. This is because a large part of the zero-field cancellation is reliant on the specific 120° structure of the in-plane ground state, which is modified by the application of a magnetic field. Therefore, it is natural to expect ordering to become more robust in an applied magnetic field. The three-sublattice (1/3, 1/3, 0) wave vector is indeed germane to triangular antiferromagnets in a magnetic field, which stabilize a quantized magnetization plateau at 1/3 saturation in XXZ models41,42.

With this in mind, we return to a discussion of the data. Theory predicts that for the 2d U(1) Dirac state, Cp(T) is quadratic in temperature, consistent with measurements43. An alternative explanation of T2 specific heat might come from the degenerate line of spiral states found by Rastelli and Tassi for the zero-field rhombohedral XXZ model44, which has 2d-like spin fluctuations despite 3d coupling. The incommensurate long-range order of the Rastelli–Tassi spiral does not appear in our measurements; however, the field-induced Bragg peaks seen in experiment are consistent with the three-sublattice plateau states that emerge in the XXZ model in a field41,45. Indeed, the magnetization at 5 T, where the ordered phase is maximal, is approximately 1/3 of the expected saturation moment and corresponds to a plateau where χ(T)=M/H reaches zero. The best fit to neutron diffraction data further corresponds to the equal-moment two-q up–up–down structure of the plateau state.

The two peaks observed in the zero-field Cp(T) of NaYbO2 evoke a number of theoretical models of Heisenberg spins on both triangular34,35 and kagome36,37 lattices that predict dual entropy anomalies when cooling into spin liquid ground states. Both peaks are rarely observed experimentally, and interpretations of the nature of each peak vary with the specific model. Exact diagonalization studies of the XXZ Hamiltonian on a triangular lattice predict a high-temperature peak corresponding to the formation of trimers of doublet states (that is, short-range correlations), followed by a lower-temperature peak that marks the onset of a quantum spin singlet state35. Recent work exploring the S=1/2 triangular lattice using tensor renormalization group techniques predicts a dual Cp(T) anomaly with the lower-temperature peak signifying the onset of short-range/incipient order and the upper peak reflective of the onset of gapped low-energy chiral fluctuations31. The ratio of low-temperature to high-temperature peak temperatures predicted in this S=1/2 model, T1/Th0.36, is consistent with those observed in NaYbO2, and the J5 K inferred from the model is reasonably close to the θCW determined from susceptibility data.

Our data demonstrate that the nearly ideal triangular lattice of Yb ions in NaYbO2 realize an unconventional quantum disordered ground state. Unlike the majority of other spin liquid candidates such as herbertsmithite (ZnCu3(OH)6Cl2) (ref. 46), the ground state in NaYbO2 can be driven into an intermediate antiferromagnetic ordered regime in relatively weak magnetic fields. The origin of the small fraction of free spins coexisting with this ground state remains an open question; however, they are not reflective of trivial disorder, which favours the least collinear state47. Additionally, rather than hosting a purely two-dimensional network of spins where the two-dimensionality precludes long-range order such as in Ba8CoNb6O24 (refs. 48,49), interlayer geometric frustration is critical to the exclusion of order in NaYbO2. This results in the strong perturbation that field provides to a complex interplay between interlayer frustration and the nearly degenerate ground states inherent to the XXZ triangular-lattice Hamiltonian. Due to this, NaYbO2 uniquely stands able to provide considerable insight into the critical phase behaviour that manifests at the phase boundaries between the ordered and quantum disordered states in a chemically ideal frustrated triangular lattice.

Methods

Sample preparation.

Polycrystalline NaYbO2 powder was prepared by a solidstate reaction of Yb2O3 (99.99%, Alfa Aesar) with Na2CO3 (99.997%, Alfa Aesar) in a 1/1.25 molar ratio and reacted at 1,000 °C for three days, with subsequent regrinding and reheating to 1,000 °C for another day. A slight excess of Na2CO3 remains in the powder and is required to prevent the formation of magnetic Yb2O3 impurities and stabilize the NaYbO2 phase while reacting. In contrast to previous reports2729, attempts to remove this excess Na2CO3 via washing with common solvents (acidic/basic/neutral water, methanol, ethanol, isopropanol) induced NaYbO2 degradation and Yb2O3 reformation. Therefore, the initial reaction was optimized to include minimal non-magnetic Na2CO3 impurities while maintaining complete reaction of Yb2O3 powder. All subsequent measurements accounted for the known Na2CO3 impurity fractions present in samples. All samples were stored in a dry, inert atmosphere and minimally exposed to atmospheric conditions before measurements. Neutron diffraction data show that the Na content of NaYbO2 is stoichiometric within resolution (~1%). Although managing alkali content in many compounds can be challenging, the strong preference for Yb3+ aids in yielding fully occupied Na sites.

Neutron diffraction.

Powder neutron diffraction data were collected on a high-resolution neutron powder diffractometer (BT-1, National Institute of Standards and Technology (NIST)). For zero-field measurements, the instrument was equipped with a cryostat capable of reaching 1.6–300 K. A 7 T vertical field cryostat with a 3He insert was used to collect magnetic field data at 0 T, 5 T and 7 T between 0.33 K and 1.5 K. Samples were placed in a vanadium canister for high-temperature measurements and a copper cell for low-temperature measurements. The 300 K data were collected with neutrons of wavelengths 1.5399 Å and 2.0774 Å produced by a Cu(311) and Ge(311) monochromator, respectively. All other temperatures used neutrons of wavelength 2.0774 Å for maximum intensity. Rietveld refinement was performed with the FullProf software suite50 and GSAS/EXPGUI programs51,52, and fits to the data at 300 K and 1.6 K are shown in Supplementary Fig. 1. Additional inelastic neutron-scattering data were collected on a disc chopper spectrometer (NIST) in a 10 T magnetic field with a dilution insert. Incident neutrons with 5 Å wavelength were used with the medium-resolution chopper setting.

Magnetic measurements.

Magnetic properties were measured using several different instruments. Isothermal d.c. magnetization up to 9 T was collected on a physical property measurement system (PPMS, Quantum Design) with a vibrating sample magnetometer insert, and low-field d.c. magnetic susceptibility from 2 K to 300 K was obtained on a magnetic property measurement system (MPMS3, Quantum Design). Isothermal a.c. susceptibility in applied fields up to 7 T and temperature-dependent a.c. magnetic susceptibility were measured on an a.c. susceptometer at 711.4 Hz with a 0.1 Oe (7.96 A m−1) drive field and a 3He insert capable of cooling to 330 mK. All fields collected are displayed in Supplementary Fig. 1. Within our a.c. susceptibility measurements at 711.4 Hz, the sample holder’s contribution is nearly linear at this range, and this frequency-dependent background was subtracted. This was confirmed via frequency sweeps of the empty holder loaded with an equivalent amount of silver paint and thermal-link wire used for the experiment. Further zero-field, frequency-dependent, a.c. susceptibility data between 10 Hz and 10 kHz and between 50 mK and 4 K were collected with the a.c. susceptibility for dilution refrigerator option on a PPMS with a 1 Oe drive field. Measurements of the empty sample holder were performed in this a.c. susceptibility for dilution refrigerator set-up at various excitation amplitudes and frequencies to quantify the total background. Trim coils were then used to cancel (null) the measured frequency-dependent background.

EPR spectra were recorded at 4.2 K with an EMXplus (Bruker) EPR spectrometer in the perpendicular operation mode. The observed resonance line shape was modelled with the EasySpin package implemented in MATLAB (ref. 53). Broadening of the EPR line shape was observed and could be fit to a normal distribution of the out-of-plane gc-factor (full-width at half-maximum = 0.40(7)) with no resolvable broadening in the in-plane gab-factor.

Heat capacity measurements.

Specific heat measurements between 80 mK and 300 K were performed for sintered pellets of NaYbO2 under zero-field conditions on a PPMS using the dilution refrigerator insert for temperatures below 1.8 K. Data under 2.5 T, 5 T, and 9 T applied fields were also collected. The magnetic portion of specific heat was determined by subtracting the non-magnetic structural analogue NaLuO2 measured between 2 K and 200 K (this was zero within resolution for temperatures <2 K). Calculated entropy was determined by integrating Cmag/T (where Cmag is the magnetic contribution to the heat capacity) between 80 mK and 40 K.

Supplementary Material

suppl

Acknowledgements

This work was supported by the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering under award DE-SC0017752 (S.D.W. and M.B.). M.B. acknowledges partial support by the National Science Foundation Graduate Research Fellowship Program under grant no. 1650114. Work by L.B. and C.L. was supported by the DOE, Office of Science, Basic Energy Sciences under award no. DE-FG02-08ER46524. Identification of commercial equipment does not imply recommendation or endorsement by NIST.

Footnotes

Online content

Any methods, additional references, Nature Research reporting summaries, source data, statements of code and data availability and associated accession codes are available at https://doi.org/10.1038/s41567-019-0594-5.

Competing interests

The authors declare no competing interests.

Additional information

Supplementary information is available for this paper at https://doi.org/10.1038/s41567-019-0594-5.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request. Neutron data were collected on the BT-1 diffractometer and the Disk Chopper Spectrometer at the NIST Center for Neutron Research.

References

  • 1.Anderson PW Resonating valence bonds: a new kind of insulator? Mater. Res. Bull 8, 153–160 (1973). [Google Scholar]
  • 2.Anderson PW The resonating valence bond state in La2CuO4 and superconductivity. Science 235, 1196–1198 (1987). [DOI] [PubMed] [Google Scholar]
  • 3.Lee PA An end to the drought of quantum spin liquids. Science 321, 1306–1307 (2008). [DOI] [PubMed] [Google Scholar]
  • 4.Balents L. Spin liquids in frustrated magnets. Nature 464, 199–208 (2010). [DOI] [PubMed] [Google Scholar]
  • 5.Savary L. & Balents L. Quantum spin liquids: a review. Rep. Prog. Phys 80, 016502 (2017). [DOI] [PubMed] [Google Scholar]
  • 6.Witczak-Krempa W, Chen G, Kim YB & Balents L. Correlated quantum phenomena in the strong spin–orbit regime. Annu. Rev. Condens. Matter Phys 5, 57–82 (2014). [Google Scholar]
  • 7.Zhou Y, Kanoda K. & Ng T-K Quantum spin liquid states. Rev. Mod. Phys 89, 025003 (2017). [Google Scholar]
  • 8.Lee S-S & Lee PAU (1) gauge theory of the Hubbard model: spin liquid states and possible application to κ-(BEDT-TTF)2Cu2(CN)3. Phys. Rev. Lett 95, 036403 (2005). [DOI] [PubMed] [Google Scholar]
  • 9.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). [Google Scholar]
  • 10.Ma J. et al. Static and dynamical properties of the spin-1/2 equilateral triangular-lattice antiferromagnet Ba3CoSb2O9. Phys. Rev. Lett 116, 087201 (2016). [DOI] [PubMed] [Google Scholar]
  • 11.Shirata Y, Tanaka H, Matsuo A. & Kindo K. Experimental realization of a spin-1/2 triangular-lattice Heisenberg antiferromagnet. Phys. Rev. Lett 108, 057205 (2012). [DOI] [PubMed] [Google Scholar]
  • 12.Jackeli G. & Ivanov DA Dimer phases in quantum antiferromagnets with orbital degeneracy. Phys. Rev. B 76, 132407 (2007). [Google Scholar]
  • 13.Clarke SJ, Fowkes AJ, Harrison A, Ibberson RM & Rosseinsky MJ Synthesis, structure, and magnetic properties of NaTiO2. Chem. Mater 10, 372–384 (1998). [Google Scholar]
  • 14.McQueen TM et al. Successive orbital ordering transitions in NaVO2. Phys. Rev. Lett 101, 166402 (2008). [DOI] [PubMed] [Google Scholar]
  • 15.Li Y. et al. Gapless quantum spin liquid ground state in the two-dimensional spin-1/2 triangular antiferromagnet YbMgGaO4. Sci. Rep 5, 16419 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Li Y. et al. Rare-earth triangular lattice spin liquid: a single-crystal study of YbMgGaO4. Phys. Rev. Lett 115, 167203 (2015). [DOI] [PubMed] [Google Scholar]
  • 17.Li Y. et al. Muon spin relaxation evidence for the U(1) quantum spin-liquid ground state in the triangular antiferromagnet YbMgGaO4. Phys. Rev. Lett 117, 097201 (2016). [DOI] [PubMed] [Google Scholar]
  • 18.Shen Y. et al. Evidence for a spinon Fermi surface in a triangular-lattice quantum-spin-liquid candidate. Nature 540, 559–562 (2016). [DOI] [PubMed] [Google Scholar]
  • 19.Xu Y. et al. Absence of magnetic thermal conductivity in the quantum spin-liquid candidate YbMgGaO4. Phys. Rev. Lett 117, 267202 (2016). [DOI] [PubMed] [Google Scholar]
  • 20.Paddison JAM et al. Continuous excitations of the triangular-lattice quantum spin liquid YbMgGaO4. Nat. Phys 13, 117–122 (2017). [Google Scholar]
  • 21.Li Y. et al. Crystalline electric-field randomness in the triangular lattice spin-liquid YbMgGaO4. Phys. Rev. Lett 118, 107202 (2017). [DOI] [PubMed] [Google Scholar]
  • 22.Li Y-D, Wang X. & Chen G. Anisotropic spin model of strong spin–orbit-coupled triangular antiferromagnets. Phys. Rev. B 94, 035107 (2016). [Google Scholar]
  • 23.Li Y-D, Shen Y, Li Y, Zhao J. & Chen G. Effect of spin–orbit coupling on the effective-spin correlation in YbMgGaO4. Phys. Rev. B 97, 125105 (2018). [Google Scholar]
  • 24.Zhu Z, Maksimov PA, White SR & Cheryshev AL Disorder-induced mimicry of a spin liquid in YbMgGaO4. Phys. Rev. Lett 119, 157201 (2017). [DOI] [PubMed] [Google Scholar]
  • 25.Kimchi I, Nahum A. & Senthil T. Valence bonds in random quantum magnets: theory and application to YbMgGaO4. Phys. Rev. X 8, 031028 (2018). [Google Scholar]
  • 26.Ma Z. et al. Spin-glass ground state in a triangular-lattice compound YbZnGaO4. Phys. Rev. Lett 120, 087201 (2018). [DOI] [PubMed] [Google Scholar]
  • 27.Hashimoto Y, Wakeshima M. & Hinatsu Y. Magnetic properties of ternary sodium oxides NaLnO2 (Ln = rare earths). J. Solid State Chem 176, 266–272 (2003). [Google Scholar]
  • 28.Liu W. et al. Rare-earth chalcogenides: a large family of triangular lattice spin liquid candidates. Chin. Phys. Lett 35, 117501 (2018). [Google Scholar]
  • 29.Baenitz M. et al. NaYbS2: a planar spin-1/2 triangular-lattice magnet and putative spin liquid. Phys. Rev. B 98, 220409(R) (2018). [Google Scholar]
  • 30.Zeng C. & Elser V. Numerical studies of antiferromagnetism on a Kagomé net. Phys. Rev. B 42, 8436 (1990). [DOI] [PubMed] [Google Scholar]
  • 31.Chen L. et al. Two-temperature scales in the triangular-lattice Heisenberg antiferromagnet. Phys. Rev. B 99, 140404(R) (2019). [Google Scholar]
  • 32.Nambu Y, Nakatsuji S. & Maeno Y. Coherent behavior and nonmagnetic impurity effects of spin disordered state in NiGa2S4. J. Phys. Soc. Jpn 75, 043711 (2006). [Google Scholar]
  • 33.Gardner JS, Gingras MJP & Greedan JE Magnetic pyrochlore oxides. Rev. Mod. Phys 82, 53–107 (2010). [Google Scholar]
  • 34.Wang YR Specific heat of a quantum Heisenberg model on a triangular lattice with two exchange parameters and its application to 3He adsorbed on graphite. Phys. Rev. B 45, 12608(R) (1992). [DOI] [PubMed] [Google Scholar]
  • 35.Isoda M, Nakano H. & Sakai T. Specific heat and magnetic susceptibility of Ising-like anisotropic Heisenberg model on kagome lattice. J. Phys. Soc. Jpn 80, 084704 (2011). [Google Scholar]
  • 36.Elstner N. & Young AP Spin-1/2 Heisenberg antiferromagnet on the kagomé lattice: high-temperature expansion and exact-diagonalization studies. Phys. Rev. B 50, 6871–6876 (1994). [DOI] [PubMed] [Google Scholar]
  • 37.Singh RRP & Oitmaa J. High-temperature series expansion study of the Heisenberg antiferromagnet on the hyperkagome lattice: comparison with Na4Ir3O8. Phys. Rev. B 85, 104406 (2012). [Google Scholar]
  • 38.Garlea VO et al. Exotic magnetic field-induced spin-superstructures in a mixed honeycomb-triangular lattice system. Phys. Rev. X 9, 011038 (2019). [Google Scholar]
  • 39.Iaconis J, Liu C, Haláz GB & Balents L. Spin liquid versus spin orbit coupling on the triangular lattice. SciPost Phys. 4, 003 (2018). [Google Scholar]
  • 40.Zhu Z, Maksimov PA, White SR & Chernyshev AL Topography of spin liquids on a triangular lattice. Phys. Rev. Lett 120, 207203 (2018). [DOI] [PubMed] [Google Scholar]
  • 41.Starykh OA Unusual ordered phases of highly frustrated magnets: a review. Rep. Prog. Phys 78, 052502 (2015). [DOI] [PubMed] [Google Scholar]
  • 42.Chubokov AV & Golosov DI Quantum theory of an antiferromagnet on a triangular lattice in a magnetic field. J. Phys.: Condens. Matter 3, 69–82 (1991). [Google Scholar]
  • 43.Ran Y, Hermele M, Lee PA & Wen X-G Projected-wave-function study of the spin-1/2 Heisenberg model on the Kagomé lattice. Phys. Rev. Lett 98, 117205 (2007). [DOI] [PubMed] [Google Scholar]
  • 44.Rastelli E. & Tassi A. The rhombohedral Heisenberg antiferromagnet: infinite degeneracy of the ground state and magnetic properties of solid oxygen. J. Phys. C: Solid State Phys 19, L423–L428 (1986). [Google Scholar]
  • 45.Ono T. et al. Magnetization plateaux of the S=1/2 two-dimensional frustrated antiferromagnet Cs2CuBr4. J. Phys.: Condens. Matter 16, S773–S778 (2004). [Google Scholar]
  • 46.Helton JS et al. Spin dynamics of the spin-1/2 kagome lattice antiferromagnet ZnCu3(OH)6Cl2. Phys. Rev. Lett 98, 107204 (2007). [DOI] [PubMed] [Google Scholar]
  • 47.Maryasin VS & Zhitomirsky ME Triangular antiferromagnet with nonmagnetic impurities. Phys. Rev. Lett 111, 247201 (2013). [DOI] [PubMed] [Google Scholar]
  • 48.Rawl R. et al. Ba8CoNb6O24: a spin-1/2 triangular-lattice Heisenberg antiferromagnet in the two-dimensional limit. Phys. Rev. B 95, 060412(R) (2017). [Google Scholar]
  • 49.Cui Y. et al. Mermin–Wagner physics, (H, T) phase diagram, and candidate quantum spin-liquid phase in the spin-1/2 triangular-lattice antiferromagnet Ba8CoNb6O24. Phys. Rev. Mater 2, 044403 (2018). [Google Scholar]

References

  • 50.Rodríguez-Carvajal J. Recent advances in magnetic structure determination by neutron powder diffraction. Phys. B: Condens. Matter 192, 55–69 (1993). [Google Scholar]
  • 51.Larson AC & Von Dreele RB General Structure Analysis System (GSAS) Report LAUR 86–748 (Los Alamos National Laboratory, 2004). [Google Scholar]
  • 52.Toby BH EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr 34, 210–213 (2001). [Google Scholar]
  • 53.Stoll S. & Schweiger A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson 178, 42–55 (2006). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

suppl

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request. Neutron data were collected on the BT-1 diffractometer and the Disk Chopper Spectrometer at the NIST Center for Neutron Research.

RESOURCES