Skip to main content
Nature Communications logoLink to Nature Communications
. 2024 Nov 15;15:9778. doi: 10.1038/s41467-024-53900-3

Kitaev interactions through extended superexchange pathways in the jeff=1/2 Ru3+ honeycomb magnet RuP3SiO11

Aly H Abdeldaim 1,2,3,, Hlynur Gretarsson 4, Sarah J Day 3, M Duc Le 2, Gavin B G Stenning 2, Pascal Manuel 2, Robin S Perry 2,5, Alexander A Tsirlin 6, Gøran J Nilsen 2,7,, Lucy Clark 1,
PMCID: PMC11568271  PMID: 39548058

Abstract

Magnetic materials are composed of the simple building blocks of magnetic moments on a crystal lattice that interact via magnetic exchange. Yet from this simplicity emerges a remarkable diversity of magnetic states. Some reveal the deep quantum mechanical origins of magnetism, for example, quantum spin liquid (QSL) states in which magnetic moments remain disordered at low temperatures despite being strongly correlated through quantum entanglement. A promising theoretical model of a QSL is the Kitaev model, composed of unusual bond-dependent exchange interactions, but experimentally, this model is challenging to realise. Here we show that the material requirements for the Kitaev QSL survive an extended pseudo-edge-sharing superexchange pathway of Ru3+ octahedra within the honeycomb layers of the inorganic framework solid, RuP3SiO11. We confirm the requisite jeff=12 state of Ru3+ in RuP3SiO11 and resolve the hierarchy of exchange interactions that provide experimental access to an unexplored region of the Kitaev model.

Subject terms: Magnetic properties and materials


Recent theoretical studies indicate that the Kitaev model may be realized in framework materials exhibiting extended superexchange pathways. Here the authors report experimental evidence showing that the material requirements for a Kitaev quantum spin liquid are satisfied in a inorganic framework material.

Introduction

The pursuit of new magnetic materials provides a route towards novel quantum states of matter, and thus serves as an important demonstration of the symbiosis between condensed matter experiment and theory. Traditionally, the impetus to search for novel phenomena in magnetic materials stems directly from advances in theory—a prime example being the landmark developments of the concepts of topology in condensed matter1,2 that are a major motivation for modern experimental materials research3. But increasingly, the design and synthesis of new magnetic materials provide an equally important guide to push the boundaries of state-of-the-art theory. Materials-led discoveries often highlight the critical role played by the complexities of real magnetic systems—such as further near-neighbour exchange interactions, lattice distortions and disorder—that go beyond idealised models of magnetism but ultimately govern the magnetic properties we can observe and exploit4. Quantum spin liquids (QSLs) present particularly compelling challenges for both experiment and theory and are highly sought by both communities for their manifestation of long-range quantum entanglement and topological excitations that may provide alternative routes to quantum computing5. Experimentally, QSLs are widely sought in magnetic materials where the underlying crystal structure introduces a geometric frustration that prevents the simultaneous energy minimisation of every pair-wise magnetic exchange interaction within the system6. However, from a theoretical perspective, this geometric magnetic frustration often compromises the analytical and numerical tractability of the exchange Hamiltonian of model systems that are predicted to host QSL ground states, such that the true nature of the QSL states predicted for many archetypal models of geometrically frustrated magnetism is still outstanding7,8. For instance, the character of the QSL describing the ground state of the S=12 Heisenberg kagome antiferromagnet is still widely debated9,10 and, while several materials candidates of this model have been identified11, the presence of disorder within their crystal structures makes it challenging to determine unambiguously the most relevant model to describe the correct magnetic ground state1214. Accordingly, magnetic models with exactly solvable QSL ground states that can be extended to account for the complexities of candidate materials are highly attractive.

In this vein, the Kitaev model serves as an important and rare example of an exactly solvable model of frustrated magnetism15. It is most widely formulated in terms of a honeycomb lattice of ferromagnetically coupled magnetic moments (see Fig. 1a), whose interactions are frustrated by an easy-axis exchange anisotropy orthogonal to the bond that connects nearest-neighbour moments16. The resulting exchange Hamiltonian is commonly represented in a cubic Cartesian reference frame as8,17

H=i,jγKijγSiγSjγ, 1

where the summation of the spin operators, S, covers the 〈ij〉 pairs of nearest-neighbour spins along some [α, β, γ]  = [(yxz), (zxy), (xyz)] bond direction, and K is the bond-dependent Kitaev exchange interaction (see Fig. 1a). In the quantum limit, the Kitaev model is exactly solvable by fractionalizing the spin degrees of freedom into Majorana fermions and gauge fluxes, yielding a QSL ground state with characteristic topological excitations15.

Fig. 1. Mapping the crystal structure of RuP3SiO11 (RPSO) to the Kitaev model on a honeycomb network.

Fig. 1

a The Kitaev model on a honeycomb network describes a system of magnetic moments (vertices) with anisotropic bond-dependent exchange interactions on cubic (x, y, z) axes resulting in magnetic frustration. The basis vectors of the coordinate system used to define the magnetic Hamiltonian are indicated by arrows. Such a model may be realised in materials in which jeff=12 moments are connected on a honeycomb network via octahedral edge-sharing. b In the R3¯c crystal structure of RPSO, honeycomb layers of Ru3+ ions are stacked by pyrophosphate (P2O74) and pyrosilicate (Si2O76) units in an ABC stacking sequence along the c-axis. c The honeycomb rings of trigonally distorted RuO6 octahedra viewed down the c-axis with a d pseudo-edge-sharing connectivity through two phosphate (PO43) units.

Initially, it was unclear how the bond-dependent exchange interactions of the Kitaev model could be realised experimentally in a magnetic material until the breakthrough development of the Jackeli-Khaliullin mechanism, a superexchange theory that sets out the structural and electronic criteria required for the formation of dominant Kitaev-like interactions in real materials18. The Jackeli-Khaliullin mechanism demonstrates that the Kitaev model can be realised by placing spin-orbit entangled jeff=12 moments on a two-dimensional honeycomb lattice via an edge-sharing octahedral connectivity. The anisotropy of this spin-orbit entangled state ensures that exchange interactions between neighbouring jeff=12 moments are extremely sensitive to their bonding geometry, giving rise to the frustrated bond-dependent exchange interaction through a 90° metal-ligand-metal edge-sharing geometry. This arrangement ensures that the electron hopping processes responsible for the isotropic nearest-neighbour Heisenberg interaction, J, exactly cancel out, leaving only the anisotropic Kitaev interaction, K. While this theory was originally developed for low-spin d5 electronic configurations, more recent work has extended the Jackeli-Khaliullin mechanism to include d719,20 and f-electron21 systems with strong spin-orbit coupling.

The requirements of the Jackeli-Khaliullin mechanism, however, pose considerable synthetic challenges for realising candidate materials for the Kitaev QSL. For the originally proposed d5 electronic configuration, only five transition metal ions with sufficiently strong spin-orbit coupling required for the requisite jeff=12 state are available for study, with, typically, chemically inaccessible oxidation states. This is reflected in the very limited pool of candidate materials studied in the context of the Kitaev model22, which includes only a handful of Ru3+ and Ir4+-containing materials, with the more synthetically challenging Re2+, Os3+, and Rh4+-based materials remaining largely unexplored23,24. Additionally, the ideal octahedral crystal field environment and edge-sharing bonding geometry are often broken by the structural symmetry of real materials16, and the extended orbital wavefunctions of their 4d and 5d transition metal ions lead to appreciable direct nearest-neighbour exchange, as well as further neighbour interactions. Therefore, additional terms in the exchange Hamiltonian are generally inevitable in candidate systems, which in turn are more accurately described by an extended Kitaev model, HJKΓΓ. This includes an isotropic Heisenberg exchange interaction, J, an off-diagonal bond-dependent frustrated interaction, Γ, and an additional interaction that arises from trigonal distortions to the perfect octahedral symmetry in the crystal field of candidate materials, Γ (see Methods). The vital role that these additional exchange interactions play in driving candidate Kitaev materials away from the QSL ground state is exemplified in one of the most-studied candidate systems to date, α-RuCl322,25,26. Despite hosting all prerequisites for dominant Kitaev interactions, the ground state of α-RuCl3 is magnetically ordered27,28. This is likely because in addition to a dominant ferromagnetic Kitaev interaction, significant contributions from all other terms in the extended HJKΓΓ model along with further-neighbour interactions are present in the exchange Hamiltonian of α-RuCl317,29. However, the proximity of the ground state of α-RuCl3 to the Kitaev QSL is still intensely debated17,29, and it and all other candidate materials highlight the experimental challenges in attaining the criteria described by the Jackeli-Khaliullin mechanism as well as the fragility of the Kitaev QSL22.

Given the experimental complexities associated with the existing set of Kitaev materials, recent theoretical studies highlight the importance of extending the pool of candidates to encompass framework materials with more open and extended crystal structures than the dense inorganic materials studied to date22. In particular, ab initio studies have predicted that the Jackeli-Khaliullin mechanism will still apply via an extended pseudo-edge-sharing superexchange pathway in a metal-organic framework solid30,31. At the same time, this approach should also increase the typical nearest-neighbour distances within the honeycomb layers of candidate materials, thus reducing the direct orbital overlap of 4d and 5d transition metal ions that give rise to the non-Kitaev exchange interactions in the extended HJKΓΓ model16,31. In this way, framework materials may provide a route to tuning the ratio of exchange interactions in the extended Kitaev model, providing experimental access to new regions of the magnetic phase diagram, including, possibly, the Kitaev QSL ground state. Motivated by this hypothesis, we have identified the inorganic framework material, RuP3SiO11 (RPSO), as an alternative candidate for exploring the Kitaev QSL beyond α-RuCl3. Through comprehensive structural characterisation and resonant inelastic X-ray scattering, we show that RPSO consists of well-separated honeycomb layers of pseudo-edge-sharing jeff=12 Ru3+ ions. Low-temperature measurements of the magnetic susceptibility, specific heat and neutron diffraction data of RPSO confirm it adopts a magnetic ground state below TN=1.3 K that is distinct from α-RuCl3 due to its unique exchange Hamiltonian, which we show contains a dominant anisotropic Kitaev interaction through analysis of inelastic neutron scattering data. Above a critical field HC=3.55 T, we show that TN is suppressed and RPSO enters a field-polarised phase, highlighting its more readily tuneable exchange interactions in comparison to α-RuCl3 with HC8 T32.

Results

Crystal Structure of RPSO: Two-dimensional honeycomb layers with pseudo-edge-sharing Ru3+ octahedra

To determine the crystal structure of RPSO, we have performed Rietveld analysis of high-resolution synchrotron X-ray and neutron powder diffraction data. This analysis confirms that RPSO adopts a trigonal R3¯c structure33 over the measured temperature range of 0.08 − 300 K (see Methods). The structure of RPSO is composed of quasi-two-dimensional buckled honeycomb layers of octahedrally coordinated Ru3+ ions (see Fig. 1b, c). Each corner of the octahedron is occupied by an O2− anion from a phosphate (PO43) tetrahedron. The latter connect neighbouring Ru3+ ions within the honeycomb layers in a pseudo-edge-sharing fashion (see Fig. 1d). Each phosphate group within the honeycomb layers is part of a larger pyrophosphate (P2O74) linker that connects the honeycomb layers along the c-axis of the crystal structure in an ABC stacking sequence. In the centre of each honeycomb ring is a pyrosilicate (Si2O76) group which stabilises the open inorganic framework structure (see Fig. 1c). The arrangement of the pyrophosphate groups within the crystal structure of RPSO results in a subtle trigonal distortion of the Ru3+ octahedral crystal field, with two distinct Ru-O bond lengths within each octahedron of 2.027(1) Å and 2.049(1) Å at 300 K.

Inspection of the crystal structure of RPSO allows us to infer the possible magnetic exchange pathways between neighbouring Ru3+ ions. The leading nearest-neighbour exchange interaction within the honeycomb layers, J, occurs through the two equivalent Ru-O-P-O-Ru pathways of the two phosphate groups that form the pseudo-edge-sharing connectivity of neighbouring octahedra (see Fig. 1d). The next-nearest-neighbour exchange interaction within the honeycomb layers does not have an obvious superexchange pathway, while the further-neighbour coupling, J3, requires a much longer pathway comprised of one silicate and two phosphate groups. Between the honeycomb layers, an interlayer exchange pathway, J, runs along the pyrophosphate linkers. Compared with α-RuCl3, the more open framework structure extends the intra- and interlayer exchange pathways. In RPSO, the Ru-Ru distance across the nearest-neighbour exchange pathway is 4.800(1) Å compared with  ≈ 3.46 Å in α-RuCl32628. The interlayer Ru-Ru distance is also extended in RPSO to 7.172(1) Å from  ≈ 6.01 Å in α-RuCl32628. At the same time, the nearest-neighbour superexchange angles are comparable in both systems at 94.36(3)° and  ≈ 93.1° in RPSO and α-RuCl32628, respectively. Thus, we hypothesise that the local coordination environment and connectivity of the Ru3+ ions in RPSO should preserve the requirements for anisotropic Kitaev exchange interactions, while the more open framework in comparison to α-RuCl3 should tune the hierarchy of exchange interactions within the extended HJKΓΓ model and enhance the two-dimensionality of the honeycomb layers.

Local and Collective Magnetic Properties of RPSO: jeff=1/2 Ru3+ moment and Néel ground state

An essential ingredient to produce the anisotropic Kitaev exchange coupling on the honeycomb network of RPSO is a jeff=12 state for its Ru3+ ions. In the case of the 4d5 configuration of Ru3+, a jeff=12 state emerges when an octahedral crystal field splitting, ΔO, combined with a low-spin electron configuration creates a single electron hole in the threefold degenerate t2g manifold with electron spin s=12 and an effective orbital angular momentum leff=1. The spin-orbit coupling interaction, λ, mixes these levels, creating the required jeff=12 state (see Fig. 2a). To examine the relevance of this spin-orbit entangled jeff=12 state in RPSO, we measured its resonant inelastic X-ray scattering (RIXS) spectrum at the Ru L3-edge at 25 K (see Methods). The measured spectrum along with the calculated electronic transitions are shown in Fig. 2b. A single sharp transition is observed below 1 eV, which bears a strong resemblance to the jeff=1232 excitation of α-RuCl334, both in terms of its width—with a FWHM ≈ 150 meV—and its energy, E ≈ 250 meV. Moreover, unlike in the related Ir4+ honeycomb compounds, (Na,Li)2IrO335, no trigonal crystal field splitting is observed within the instrumental resolution. Taken together, this provides strong experimental evidence for the existence of a jeff=12 state in RPSO.

Fig. 2. Confirming the jeff=12 state of Ru3+ in RPSO by resonant inelastic X-ray scattering (RIXS).

Fig. 2

a In an octahedral crystal field environment, the d-orbitals of a Ru3+ 4d5 ion are split into a t2g manifold with a single electron hole and an empty eg manifold separated by the octahedral crystal field splitting parameter, ΔO. In the presence of spin-orbit coupling, λ, the t2g manifold can be further split, creating the possibility of a spin-orbit-entangled jeff=12 moment for Ru3+. If such moments are coupled by edge-sharing octahedral connectivity, then anisotropic Kitaev-type exchange interactions may dominate between them. b RIXS data collected for RPSO at 25 K (IRIXS, DESY) confirm the jeff=12 state of the Ru3+ ions in this system through the single sharp transition observed below 1 eV. The RIXS spectrum can be simulated using full atomic multiplet calculations, from which estimates for the octahedral crystal field splitting, ΔO, Hund’s coupling, JH, and spin-orbit coupling, λ, of RPSO can be extracted.

Above 1 eV in the RIXS spectrum of RPSO (see Fig. 2b), a series of peaks are observed originating from excitations into the eg manifold of the octahedral crystal field environment. Although the overall spectral form is similar to α-RuCl3, these excitation bands are shifted by about 250 meV higher in energy and are also noticeably sharper. The energy shift relative to α-RuCl3 is due to the larger octahedral crystal field splitting expected for O2− compared to Cl, while the peak sharpness likely stems from a smaller overlap within the electron-hole continuum, i.e., a larger optical gap. Additional insight can be gained by comparing the measured spectrum with full atomic multiplet calculations (see Methods). In particular, the Hund’s rule coupling, JH=340 meV, and crystal field splitting, ΔO=2.4 eV, determined previously for α-RuCl3 by RIXS34 must be increased to JH=460 meV and ΔO=2.8 eV for RPSO to capture the sharper eg spectral features as well as their shift to higher energies. Such an increase in JH can arise from a weaker screening effect in more insulating compounds, for example, as has been observed in K2RuCl636 and RuX3 (X  = Cl, Br, I)37. Despite the differences in ΔO and JH, the spin-orbit coupling in RPSO, λ = 130 meV, is comparable to that reported for α-RuCl3, λ = 150 meV34, which highlights the significant role of spin-orbit coupling in the physics of RPSO.

Having established the relevance of the jeff=12 state to the local magnetic properties of RPSO, we now turn to the collective magnetic properties and elucidation of its magnetic ground state. Curie-Weiss fitting the inverse magnetic susceptibility of RPSO (see Fig. 3a, Methods) between 100 and 300 K yields a Weiss constant, θCW=5.70(1) K, a Curie constant, C = 0.33(1) emu K mol−1, and a temperature-independent term, χ0 = 2.90(1) × 10−4 emu mol−1. This implies dominant antiferromagnetic interactions between neighbouring Ru3+ moments in RPSO, and an effective moment μeff=1.63(1)μB per Ru3+ ion consistent with the jeff=12 state confirmed by RIXS. The temperature-independent term, χ0, most likely stems from the Pauli paramagnetism of the small Ru and RuO2 phases identified in the sample by powder diffraction (see Methods). Efforts to improve the Curie-Weiss analysis by incorporating temperature-dependent magnetic moments—often a more suitable approach for strongly spin-orbit coupled systems38—proved inconclusive and most likely require single-crystal magnetometry data. Below 15 K, the development of short-range magnetic correlations is evidenced in the magnetic susceptibility data through the deviation from Curie-Weiss behaviour followed by a cusp just above 1 K that indicates the onset of long-range magnetic order. This behaviour is also observed in the low-temperature specific heat data of RPSO (see Fig. 3b, Methods)—where the lattice contribution to the total heat capacity will be minimal—in which a broad peak near 15 K is followed by a sharp λ-type anomaly at TN=1.3 K.

Fig. 3. Elucidating Néel order in RPSO below TN=1.3 K.

Fig. 3

a The magnetic susceptibility data (χm, left axis) measured for RPSO between 0.4 − 300 K in an applied field of 0.1 T after zero-field cooling shows a cusp just above 1 K, indicating the onset of a magnetic phase transition. Curie-Weiss fitting the inverse magnetic susceptibility (χm1, right axis) between 100 − 300 K gives a Weiss constant, θCW=5.70(1) K, a Curie constant, C = 0.33(1) emu K mol−1, and a temperature-independent term, χ0 = 2.90(1) × 10−4 emu mol−1. b The low-temperature zero-field specific heat (Cp) of RPSO confirms the magnetic phase transition at TN=1.3 K with a sharp λ-type anomaly at this temperature. The broader feature at T ≈ 15 K likely reflects the development of short-range magnetic correlations. c Temperature-subtracted powder neutron diffraction data (WISH, ISIS) reveal magnetic Bragg peaks below TN that can be indexed with a k = (0, 0, 0) propagation vector. Of the four magnetic models compatible with this k and the symmetry of the crystal structure, the best magnetic Rietveld fit (shown) is obtained with the mΓ1− irreducible representation (χ2 = 1.07 and Rmag=10.2%). This model corresponds to the magnetic space group R3¯c and a d G-type antiferromagnetic order of the Ru3+ moments in the honeycomb layers of RPSO with an ordered moment, μord=0.35(1)μB.

Two similar features have been observed previously in specific heat measurements of polycrystalline samples of α-RuCl339. These have been attributed to the onset of two magnetically ordered regimes that arise due to the presence of stacking faults between the van der Waals layers of α-RuCl327,39. In the case of RPSO, the pyrophosphate and pyrosilicate linkers that pillar the honeycomb layers will increase the energy barrier to stacking fault formation compared to the weaker van der Waals forces between the layers of α-RuCl3, and indeed, there is no evidence of stacking faults from the peak shape and intensity of high-resolution synchrotron powder X-ray diffraction data (see Supplementary Note 1). Alternatively, Monte-Carlo simulations have indicated that such a two-step magnetic entropy release for the extended HJKΓΓ model is evidence of spin fractionalisation40. The temperatures at which these features appear, however, are inconsistent with the energy scales of the HJKΓΓ model in RPSO obtained through analysis of inelastic neutron scattering data (see Section 2.3). Thus, we hypothesise that the broad feature near 15 K in the specific heat of RPSO simply reflects the onset of the correlated paramagnetic regime, rather than a distinct long-range ordered magnetic phase.

To determine the nature of the magnetic ground state of RPSO below TN=1.3 K, neutron magnetic scattering was isolated by subtracting neutron powder diffraction data collected above and below TNT = 0.08 K  − 2.5 K, see Methods). This temperature subtraction (see Fig. 3c) reveals three magnetic Bragg peaks at d-spacings corresponding to the (012), (104), and (018) reflections of the R3¯c crystal structure, which indicates a ground state magnetic structure described by the commensurate propagation vector, k = (0, 0, 0). Symmetry analysis (see Methods) results in four maximal magnetic space groups with irreducible representations that are compatible with this propagation vector and the underlying crystal structure of RPSO: mΓ1+, mΓ1−, mΓ2+ and mΓ2− in Miller-Love notation41. Here, mΓ2+ corresponds to a ferromagnetically ordered structure, while the mΓ2−, mΓ1+ and mΓ1− modes describe C, A and G-type Néel ordered states, respectively. We also considered the magnetic subgroups defined by the irreducible representations mΓ3+ and mΓ3−. Refining each of these models against the temperature-subtracted data reveals that only the model corresponding to the G-type Néel order (mΓ1− representation) is consistent with the observed magnetic diffraction (see Fig. 3c). Thus, the magnetic ground state of RPSO is composed of antiferromagnetic honeycomb layers of Ru3+ ions that couple antiferromagnetically along the c-axis (G-type Néel order, see Fig. 3d). The ordered magnetic moment obtained from this magnetic structure refinement, μord=0.35(1)μB per Ru3+, is reduced from the full jeff=12 ordered moment of 1μB and the calculated ordered moment for the S=12 honeycomb Heisenberg antiferromagnet with μord0.55μB42. Such a reduced ordered moment is characteristic of low-dimensional magnetic materials and implies the presence of frustrated exchange interactions. It is also consistent with the observed ordered moment in α-RuCl3, which is typically reported in the range of 0.30.7μB per Ru3+ ion27,43.

The significance of the G-type Néel ordered ground state of RPSO is that it provides experimental access to an otherwise unexplored region of the magnetic phase diagram of the extended HJKΓΓ Kitaev model. Indeed, all other experimental realisations of Ru3+ and Ir4+ honeycomb magnets studied in the context of the Kitaev QSL and that undergo a magnetic phase transition to a long-range ordered ground state adopt either zig-zag or incommensurate spin spiral magnetic structures at low temperatures22. This implies that RPSO has a unique hierarchy of exchange interactions within the HJKΓΓ model that gives rise to its distinct magnetic ground state. The delicate balance of competing interactions that yields the rich magnetic phase diagram for this model means that candidate materials may be tuned from one phase to another by an external perturbation, such as applied pressure or magnetic field44. For instance, in the case of α-RuCl3, the zig-zig order within the magnetic ground is suppressed upon application of a critical field HC8 T, which has been attributed to the formation of a field-induced quantum critical phase32,45,46.

To explore the tunability of the magnetic ground state of RPSO, we have measured the field dependence of its magnetic susceptibility χ(TH) and isothermal magnetisation M(HT) below TN (see Methods). Upon increasing applied field strength, the cusp in the magnetic susceptibility of RPSO—that indicates the onset of the long-range Néel order—broadens and shifts lower in temperature up to an applied field of 3.5 T, above which TN is no longer observable down to 0.5 K (see Supplementary Note 2). Below TN, the isothermal magnetisation increases linearly with the applied field up to 2 T, beyond which the rate of increase in magnetisation becomes steeper before reaching a critical field, HC=3.55 T (see Supplementary Fig. 2). Such a field dependence of the magnetisation is typical of a field-induced phase transition, which can be more easily observed in the field derivative of the magnetisation (see Fig. 4a). The resulting field-temperature phase diagram (see Fig. 4b) reveals the reduction in TN upon increasing the applied field until a critical paramagnetic phase is reached above HC=3.55 T. Understanding the nature of this field-induced phase of RPSO—whether it is a simple field-polarised or a quantum paramagnetic state—requires further investigation of single-crystal samples, but the field-temperature phase diagram is again reminiscent of α-RuCl3, albeit with a lower critical field, which stems from the weaker exchange interactions in RPSO.

Fig. 4. Supressing Néel order in RPSO upon application of a magnetic field.

Fig. 4

a The field derivative of the magnetisation isotherms shows the development of the critical field below TN=1.3 K. b The temperature-field phase diagram of RPSO mapped through magnetic susceptibility and isothermal magnetisation measurements reveals the suppression of the zero-field TN=1.3 K upon application of a magnetic field. Above a critical field of HC=3.55 T, TN appears completely suppressed within the temperature range of the measurement.

Exchange Hamiltonian of RPSO: Kitaev interactions from inelastic neutron scattering

To understand the origin of the G-type Néel ordered ground state of RPSO, we need to establish the interactions within the underlying exchange Hamiltonian (see Methods). Experimentally, this can be achieved through the analysis of the dynamical structure factor of RPSO, S(Q,ΔE=ω)exp, measured by inelastic neutron scattering (INS, see Methods). Figure 5a shows the S(Q,ΔE=ω)exp measured for RPSO below TN at T = 0.08 K with an incident neutron energy Ei=1.78 meV. The magnetic INS spectrum is gapped, with a gap Δ ≈ 0.1 meV, and two clear bands of magnetic scattering intensity are observed with bandwidths 0.1 − 0.4 meV and 0.42 − 0.82 meV, respectively. The sharply dispersing feature in Fig. 5a above 0.8 meV is the roton excitation of superfluid 4He arising from the He exchange gas loaded in the sample can to cool the sample (see Methods). The intensity seen at low-∣Q∣ and small−ΔE in Fig. 5a is also spurious, as verified by the observation that its ∣Q∣-dependence varies with Ei. Overall, the form of the magnetic excitation spectrum of RPSO below TN—particularly that the spectrum is gapped—indicates a strongly anisotropic exchange Hamiltonian.

Fig. 5. Determining the exchange Hamiltonian of RPSO through inelastic neutron scattering.

Fig. 5

a Experimental dynamical structure factor, S(Q,ΔE=ω)exp, measured at 0.08 K with neutrons of incident energy Ei=1.78 meV (LET, ISIS). b Powder-averaged and energy-convoluted S(Q, ΔE = ω)calc simulated using linear spin wave theory for the exchange Hamiltonian, HJKΓΓ, with J = 0.32 meV, K = − 0.54 meV, Γ = 0.44 meV, and Γ=0.12 meV (Solution 1, Table 1). c–f ΔE-integrated cuts (blue circles) fitted to S(Q, ΔE = ω)calc (red lines). Error bars in cf represent one standard deviation.

To establish the relevance of HJKΓΓ in RPSO, we performed a grid search using linear spin wave theory (LSWT), in which we compared S(Q,ΔE=ω)exp to solutions of the extended Kitaev HJKΓΓ model within the region of the phase diagram where the experimentally observed Néel ground state is stable (see Methods). The resulting sets of exchange parameters that gave the best fit to the data in the initial grid search were then further optimised using a simulated annealing algorithm (see Methods). This optimisation of the HJKΓΓ model yields two regions of the phase diagram for which the fits to the experimental data are indistinguishable, which have either a dominant ferromagnetic (K < 0) or antiferromagnetic (K > 0) Kitaev exchange coupling. However, these two regions are physically equivalent and are related by the self-duality of the extended Kitaev HJKΓΓ model when defined in the cubic axes of the Kitaev framework and are introduced by a global π rotation about the crystallographic c-axis (see Supplementary Note 4)17,47. Thus, we find five solutions to the HJKΓΓ model that give the lowest χ2 against the experimental data and that are well separated in χ2 from the next set of solutions (see Table 1, Supplementary Fig. 7). Figure 5b shows the S(Q, ΔE = ω)calc simulated by LSWT for the set of exchange parameters in Solution 1, and Fig. 5c–f show representative comparisons of this solution to several Q-integrated cuts to the data. Within the ferromagnetic K region, all five of the lowest χ2 LSWT solutions of the HJKΓΓ model unveil two dominant anisotropic exchange interactions for RPSO: a ferromagnetic Kitaev exchange interaction K, which is approximately equal in strength to an antiferromagnetic Γ. In addition to these two bond-dependent, frustrated exchange interactions, we also find an antiferromagnetic isotropic nearest-neighbour exchange interaction, J ≈ −2/3K, and a ferromagnetic Γ that is consistently smaller than the other three parameters. Notably, the latter aligns with our analysis of the RIXS spectrum and supporting DFT calculations (see Supplementary Note 3), which all show that the trigonal splitting of the octahedral crystal field in RPSO is small.

Table 1.

Exchange parameters (in meV) of the lowest χ2 solutions of the HJKΓΓ model obtained for RPSO in this work by fitting inelastic neutron scattering data to linear spin wave theory

RPSO J K Γ Γ χ2
Solution 1 0.32 −0.54 0.44 −0.12 67.7
Solution 2 0.23 −0.31 0.38 −0.08 68.9
Solution 3 0.16 −0.15 0.33 −0.04 69.7
Solution 4 0.25 −0.39 0.41 −0.08 69.7
Solution 5 0.37 −0.65 0.49 −0.14 70.5
Normalised J K Γ Γ J3
RPSO 0.60 – 1.10 −1 0.75 − 2.20 −0.22 − −0.27
α-RuCl316,17,55 −0.30 −  −0.53 −1 0.41 − 0.85 0.22 − 0.52 0.25 − 0.50

For comparison, the range of  −K normalised exchange parameters reported for α-RuCl316,17,55 have also been included.

Further efforts to distinguish between the optimised parameters of Solutions 1 – 5 using experimental data did not yield a single, unique solution. When comparing the predicted critical field between the Néel ordered ground state and field-induced phase with the experimental HC=3.55 T, all solution sets consistently yield HC,calc=3.84.1 T. Additionally, calculating the mean-field Weiss constant38 results in θCW1 K for all solutions, which makes a direct comparison between the Curie-Weiss analysis of the magnetic susceptibility data and LSWT challenging. However, the calculated HJKΓΓ model successfully captures the key features in S(Q,ΔE=ω)exp (see Fig. 5), in particular, the two clear branches of excitations and the gaps between them, which place strong constraints on the possible values of the exchange parameters. Thus, LSWT provides a well-defined region of exchange parameter space for RPSO and accurately determines the ratio of the exchange interactions in the system. The observed discrepancy with the experimental intensities could imply that further exchange interactions are required for a full microscopic description of RPSO. Indeed, the Ru-Ru bond symmetry of further near-neighbour couplings in RPSO allows for anisotropic exchange tensors, resulting in an exchange Hamiltonian with at least 18 independent parameters that cannot be fit reasonably to INS data. However, this fitting discrepancy most likely stems from the semi-classical formalism of LSWT, which underestimates both quasi-particle interactions48,49 and other quantum effects50, and thus creates a bottleneck for the analysis of INS data of low-dimensional, frustrated quantum magnets. Multi-magnon interactions and the resulting reduction in the single magnon lifetime generally result in a characteristic broadening of higher energy spectral lines51,52. The significance of this broadening effect has been underscored in several other systems with strongly anisotropic exchange interactions, including Kitaev-related materials such as α-RuCl353, CoI249, and BaCo2(AsO4)254. Hence, in the case of RPSO, where the two dominant exchange interactions are strongly anisotropic—combined with the small ordered moment and quantum spin—the current analysis strategy approaches the limits of what can be extracted from INS data with LSWT. Any further insight into the exchange Hamiltonian of RPSO demands future developments in experiment and theory, and will require a combination of computational and experimental tools. These will include new capabilities for the measurement of the field evolution and angular dependence of single-crystal INS spectra with neutron polarisation analysis, along with calculations of multi-magnon dispersion relations.

Discussion

While the energy scales of the exchange parameters determined for RPSO are at least an order of magnitude smaller than those observed in α-RuCl316,17,55, considering their relative magnitudes allows for a useful comparison of the two systems. Table 1 summarises a representative set of HJKΓΓ parameters normalised to the ferromagnetic Kitaev exchange for α-RuCl3 and RPSO. We note that Table 1 is not exhaustive, as the full microscopic description of α-RuCl3 is still heavily debated with at least 20 estimates of the relevant exchange parameters (e.g., see Table 1 in17,29). However, overall, the absolute relative magnitudes of the HJKΓΓ parameters appear consistent across both systems, with dominant K and Γ and smaller J and Γ. Interestingly, the relative magnitudes of the nearest-neighbour Heisenberg exchange, J, are similar in both RPSO and α-RuCl3. This is surprising given that one might expect the strength of this interaction to be relatively reduced by the more open framework structure of RPSO31 and perhaps highlights the more complex influence of the multi-atom ligand facilitating the pseudo-edge-sharing superexchange on the relative magnitude of J. Otherwise, in the analysis of S(Q,ΔE=ω)exp for RPSO, the further-near-neighbour Heisenberg exchange interaction, J3, appears to be negligible, which is consistent with DFT (see Supplementary Note 3). This is in stark contrast to α-RuCl3 and other Kitaev-related materials where J3 is at least ∣K∣/416,17,29,56, and thus appears to be an important consequence of the more open framework structure of RPSO.

Collectively, the exchange parameters determined here for RPSO align with the classically calculated phase diagram of the extended HJKΓΓ Kitaev model29,57, forming a line of solutions that closely border the predicted stripy antiferromagnetic and the experimentally observed incommensurate spin-spiral22 magnetic ground states. To the best of our knowledge, this makes RPSO the first material realisation of the HJKΓΓ model to fall within the Néel ordered region of the phase diagram and the second Ru3+-based material relevant to this model beyond the α-RuX3 family (X = Cl, Br, I)58. Perhaps most importantly, RPSO also serves as an experimental proof-of-concept that anisotropic Kitaev interactions can be transmitted through a complex extended pseudo-edge-sharing superexchange pathway, corroborating the ab initio predictions of Yamada et al.30,31. Consequently, this work significantly broadens the pool of materials suitable for exploring the Kitaev QSL. Coupled with other recent ab initio studies mapping the Kitaev model to more synthetically accessible d7 systems19,20, this opens the door to an extensive experimental exploration of the HJKΓΓ phase diagram in framework materials. Moving forward, it will also be important to investigate why the frustrated bond-dependent XY-type interaction, Γ, also appears to play a dominant role in the exchange Hamiltonian of a multi-atom superexchange pathway. This might be fruitful for future theoretical and experimental investigations as further framework materials are identified as candidate Kitaev systems. Nevertheless, in the case of RPSO, the extended superexchange pathway appears to have significantly diminished the influence of the further-near-neighbour Heisenberg interaction, J3, which is a significant energy scale in all other Kitaev-related materials studied to date22,29. This—in addition to its relatively weaker exchange interactions—makes RPSO an ideal system in which to examine the effect of external perturbations, such as applied strain, pressure, and magnetic field, that may ultimately lead us to the Kitaev QSL in this inorganic framework solid.

Methods

Synthesis of RPSO

Polycrystalline samples of RuP3SiO11 (RPSO) were prepared via a three-step synthesis. In the first step, a precursor of RPSO, H2RuP3O10, was synthesised via a modified procedure reported in the literature33. In a typical reaction, RuCl3 ⋅ xH2O (Sigma-Aldrich, 99.98%) and H3PO4 (Sigma-Aldrich, 85 wt.% in H2O, 99.99% trace metals basis) were combined in a round bottom flask in a 1:4 molar ratio, respectively, and heated in air at 350 K for 6 hours under constant stirring. The resulting viscous brown solution was transported to an alumina crucible and heated at 1 K min−1 to 650 K under flowing Ar gas. After 7 days of heating, the furnace was cooled to room temperature at a rate of 1 K min−1. This resulted in a light brown, water insoluble, and glass-like amorphous product, which has been reported as H2RuP3O1059. In the second step, a water-washed pressed-powder pellet of H2RuP3O10 was heated for 7 days in an evacuated sealed quartz ampoule refilled with 425 mbar of Ar gas to form a pale orange powder of Ru(PO3)3. Temperature control at this stage of the reaction is crucial, as Ru(PO3)3 can adopt three different polymorphs59 of which the monoclinic cyclo-hexaphosphate (Ru2P6O18)—prepared by slow heating (0.5 K min−1) to 823 K—reacts to form RPSO. Finally, in the third step, a stoichiometric mixture of Ru(PO3)3 and SiO2 (Alfa Aesar, 99.99%) was ground for 30 minutes, pressed into a pellet and sealed in an evacuated quartz ampoule refilled with 425 mbar of Ar gas without a crucible. This mixture was heated at 3 K min−1 to 1233 K for 3 days before cooling to room temperature. The resulting final product of RPSO is yellow-coloured, water-insoluble and air-stable for at least a month.

Crystal and magnetic structure determination

High-resolution synchrotron powder X-ray diffraction data were collected on the I11 beamline at the Diamond Light Source using the MAC detector. The sample was densely packed in a 0.5 mm borosilicate capillary by sonication and diffraction data were measured at 300 K with an incident wavelength of λ = 0.826015(5) Å. Neutron powder diffraction data were collected on the time-of-flight diffractometer WISH at the ISIS Neutron and Muon Source. 1 g of RPSO powder was packed in a cylindrical copper can and diffraction data were measured at 0.08 K and 2.5 K using a dilution refrigerator. Crystal structure Rietveld refinements to 300 K (X-ray), 2.5 K and 0.08 K (neutron) diffraction data were performed using the GSASII60 software package (Supplementary Fig.1, Supplementary Note 1). For the Rietveld analysis of data collected at 300 K, lattice parameters, atomic positions and isotropic thermal parameters were varied within the R3¯c structural model, yielding the refined structure summarised in Supplementary Table 1 (Supplementary Note 1). Additional phases of RuO2 and Ru metal were found to account for approximately 4.6% of the sample mass. For the Rietveld fit to 2.5 K data, lattice parameters, atomic positions and isotropic thermal parameters were varied, with additional Al and Cu phases modelled using the Le Bail method to account for the scattering from the dilution fridge insert and sample can used (Supplementary Table 2, Supplementary Note 1). At 0.08 K, all structural and instrumental parameters were fixed to their values obtained from the 2.5 K fit, with the exception of the lattice parameters, which contracted isotropically upon cooling. Representational analysis and magnetic structure refinements performed against temperature-subtracted (ΔT = 0.08 K − 2.5 K) neutron diffraction data utilised the MAXMAGN61 and FullProf suite62 softwares, respectively. The Ru3+ form factor used was interpolated from data presented in63.

Magnetic susceptibility and specific heat measurements

Temperature-dependent magnetic susceptibility data were measured in DC mode for a 28 mg sample of RPSO in a Quantum Design MPMS3 with a low-temperature 3He insert. Data were collected in an applied field of 0.1 T between 0.4 and 300 K in both zero-field cooled and field-cooled protocols. The sample was contained within a gelatine capsule packed tightly with Teflon tape and held in a plastic straw. Additional temperature-dependent magnetic susceptibility measurements were carried out within the temperature range of 0.4 − 1.78 K. These measurements were taken in 1 T steps of applied field between 1 and 5 T and additionally at 3.25 T, 3.5 T, 3.75 T and 4.25 T. Isothermal magnetisation data were measured over the temperature range of 0.40 − 1.8 K, with measurements taken in increments of 0.06 K, over an applied magnetic field range of  −7 − +7 T. Specific heat data were measured on a Quantum Design Physical Property Measurement System (PPMS). 2.6 mg of RPSO was pressed into a pellet, mounted onto a puck using Apiezon N-grease and 300 data points were recorded in zero applied magnetic field between 2 and 300 K using a cryostat and between 0.1 and 4 K using a dilution refrigerator. The measurement at each temperature point was repeated twice to obtain an average.

Resonant inelastic X-ray scattering

Resonant inelastic X-ray scattering (RIXS) data were collected on the IRIXS instrument at beamline P01 of PETRA III synchrotron radiation source at DESY. RPSO powder was pressed into a pellet and mounted on a copper sample holder. The position of the zero-energy-loss spectral line was determined by measuring non-resonant spectra from glue deposited next to the sample. Data were collected at 25 K at the Ru L3-edge, where the overall energy resolution (FWHM) of the IRIXS spectrometer is 75 meV. To model the measured RIXS spectrum of RPSO, full atomic multiplet calculations were performed using the Quanty code64 assuming a 4d5 configuration for Ru3+. Spin-orbit coupling (λ), Hund’s rule coupling (JH), and octahedral crystal field splitting (ΔO) were included in the model Hamiltonian for RPSO, similar to the fitting approach developed for recent measurements of α-RuX3 (X = Cl, Br, I)37,65. These three parameters were then refined to best fit the experimental data.

Defining the exchange Hamiltonian

The extended Kitaev model, HJKΓΓ, used widely to describe the relevant exchange Hamiltonian of candidate Kitaev materials is defined as

HJKΓΓ=ijγJSiSj+KSiγSjγ+ΓSiαSjβ+SiβSjα+ΓSiαSjγ+SiγSjα±ΓijSiβSjγ+SiγSjβ+ijJ3SiSj. 2

Here, J is the isotropic Heisenberg exchange interaction, Γ—which is symmetry allowed in the Kitaev model—is an off-diagonal bond-dependent frustrated XY-type interaction that depends on a complex combination of metal-metal and metal-ligand assisted electron hoppings, Γ is dependent on the local symmetry at the metal ion site and arises from trigonal distortions to the perfect octahedral symmetry in the crystal field of candidate materials66, and J3 is the third-nearest-neighbour Heisenberg exchange interaction. The  ± sign emphasises that the bond-dependent sign structure of Γ varies depending on the frame of reference used to define the model29.

The standard parameterisation of the extended Kitaev model in Eqn. (2) assumes C2 symmetry at the Ru-Ru bond centre. In RPSO, the 1¯ point symmetry at the Ru-Ru bond centre allows for two more parameters in the exchange tensor, ξ and ζ. The resulting nearest-neighbour exchange tensor, Jaγ, of the exchange Hamiltonian is thus populated with six nonzero parameters such that

Jax=J+KΓζΓ+ζΓζJξΓΓ+ζΓJ+ξ,Jay=J+ξΓ+ζΓΓ+ζJ+KΓζΓΓζJξ,
Jaz=JξΓΓζΓJ+ξΓ+ζΓζΓ+ζJ+K,

for spins that are projected on axes with γ defining the local basis vectors, given in the crystallographic coordinates:

xyz=0.2240.1120.0290.1120.2240.0290.1120.1120.029.

The exchange tensors furthermore transform with the threefold rotation axis of the R3¯c space group between bonds in the ab-plane, and the c-glide operation between adjacent planes

Jbx=J+KΓ+ζΓζΓ+ζJ+ξΓΓζΓJξ,Jby=JξΓζΓΓζJ+KΓ+ζΓΓ+ζJ+ξ,
Jbz=J+ξΓΓ+ζΓJξΓζΓ+ζΓζJ+K.

Alternatively, HJKΓΓ can be rewritten in a basis aligned with the crystallographic axes17,44 of the R3¯c space group such that

HXXZ=ijJijxySixSjx+SiySjy+JijzSizSjz2Ji,j±±SixSjx+SiySjyc~αSixSjy+SiySjxs~αJi,jz±SixSjz+SizSjxc~α+SiySjz+SizSjys~α. 3

Here, c~α=cos(ϕα), s~α=sin(ϕα), and the phase ϕα = [2π/3, − 2π/3, 0] for the [XYZ] bonds, respectively. We emphasise that HXXZ and HJKΓΓ are equivalent descriptions of the exchange Hamiltonian that are defined using different Cartesian bases. The HJKΓΓ and HXXZ parameters are related by17

Jxy=J+13(KΓ2Γ),Jz=J+13(K+2Γ+4Γ),2Ji,j±±=13(K+2Γ2Γ),2Ji,jz±=23(KΓ+Γ). 4

Inelastic neutron scattering

Inelastic neutron scattering (INS) data were collected on the direct geometry time-of-flight cold neutron multi-chopper spectrometer, LET, at the ISIS Neutron and Muon Source. 1 g of RPSO was packed and sealed with He exchange gas in an annular copper can mounted in a dilution refrigerator. Data were collected at 0.08 K, 2 K, 3 K, 4 K, 5 K, 10 K, and 15 K with incident neutron energies, Ei= 1.78 meV, 3.14 meV and 7 meV. The choppers were operated on high-flux mode with the resolution disk and pulse-removal disk choppers spinning at frequencies of 200 Hz and 100 Hz, respectively, yielding an elastic line resolution of approximately 0.03 meV at Ei= 1.78 meV.

To analyse the data, linear spin wave theory (LSWT) simulations were performed on the SpinW MATLAB package67. All computations were powder averaged and convoluted with the energy-dependent instrumental resolution. To extract the relevant exchange Hamiltonian of RPSO from the data, eight constant momentum transfer ∣Q∣ cuts of the INS data collected at 0.08 K were used in the exchange parameter grid search, integrated over δQ = 0.1 Å−1, and centred about 0.25 − 0.95 Å−1 in 0.1 Å−1 steps. For each cut, the amplitude was fitted and a constant background term was subtracted. Given the large parameter space of the extended Kitaev model, we made the starting assumption that ζ, ξ, and any further neighbour couplings in RPSO are comparatively smaller than the four leading parameters, J, K, Γ and Γ in the HJKΓΓ model. This is justified by the relative magnitudes of the hopping integrals contributing to these parameters estimated by DFT (see Supplementary Note 3, Supplementary Information) and the open framework structure of RPSO which will likely weaken couplings beyond nearest neighbours. In the initial grid search, the four exchange parameters in the HJKΓΓ model were varied in 14 linearly spaced steps spanning two parameter spaces, one in which  − 0.6 < J < 0.6 meV,  − 1 < K < 0 meV,  − 1 < Γ < 1 meV, and 1<Γ<1 meV, and another with the same J, Γ, and Γ but with 0 < K < 1 meV. Together these represent parameter spaces with ferromagnetic and antiferromagnetic Kitaev exchange coupling, K, respectively. This approach yielded multiple almost equivalent solutions (see Supplementary Note 4, Supplementary Fig. 7) with the lowest χ2, all with either dominant anisotropic Kitaev-type ferromagnetic K and antiferromagnetic Γ interactions (see Table 1), or antiferromagnetic K for the symmetry-related dual set.

From this initial grid search, a series of local optimisations were performed around the lowest χ2 solutions of the HJKΓΓ model in an attempt to identify a unique set of exchange parameters for RPSO. These local optimisations were carried out on grid search solutions with χ2 < 30, and a simulated annealing algorithm was used to fit model exchange Hamiltonians to 60 ΔE cuts spanning the full experimental INS spectrum at 0.08 K. Three models were tested in this local optimisation (1) the HJKΓΓ model, (2) the HJKΓΓ model with the additional ξ and ζ exchange couplings allowed by the symmetry of the Ru-Ru bond centre in RPSO and (3) a 9-parameter model, including  J, K, Γ, Γ, ξ, ζ and the isotropic further-neighbour couplings J2, J3, and J. However, including parameters beyond the HJKΓΓ model did not yield improved fits. While this does not conclusively dismiss the relevance of the additional parameters tested, it underscores the limitation of the powder-averaged datasets currently available, for which there is a risk of over-parameterising when adding further parameters beyond the HJKΓΓ model. Further details and results of the INS fitting procedure can be found in Supplementary Note 4.

Supplementary information

Acknowledgements

This work was supported by the UKRI Science and Technology Facilities Council through the award of an ISIS Facility Development and Utilisation PhD Studentship for AHA as well as access to beamtime at the Diamond Light Source and ISIS Neutron and Muon Source. Work at the University of Birmingham was supported by the UKRI Engineering and Physical Sciences Research Council grant EP/V028774/1 (LC). We also acknowledge DESY—a member of the Helmholtz Association HGF—for access to beamtime. Work at the University of Leipzig was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) grant TRR 360 – 492547816 (AAT). The authors thank Dr Matthew Coak (University of Birmingham) for providing helpful feedback on a manuscript draft.

Author contributions

L.C. and G.J.N. conceived and supervised all aspects of the study. A.H.A. devised and performed the synthesis of RPSO with support from R.S.P. A.H.A. collected and analysed the magnetometry data. A.H.A. collected heat capacity with support from G.B.G.S. A.H.A. collected and analysed synchrotron X-ray diffraction, neutron diffraction and neutron spectroscopy datasets with support from S.J.D., P.M. and G.J.N., respectively. A.H.A., M.D.L. and G.J.N. devised the analysis procedure for the neutron spectroscopy data. A.A.T. performed and interpreted the density-functional theory calculations. H.G. performed and analysed resonant inelastic X-ray scattering measurements. AHA and LC wrote the manuscript with input from all the authors.

Peer review

Peer review information

Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.

Data availability

The data that support the findings of this study are available via the following: https://syncandshare.desy.de/index.php/s/xZ2Jwrzw9mpDkxN (IRIXS, DESY), 10.5286/ISIS.E.RB2210080 (WISH, ISIS), 10.5286/ISIS.E.RB2210081 (LET, ISIS). 10.5281/zenodo.13890183 (I11, Diamond, magnetization, specific heat, LSWT grids). Any requests for further data, analysis or code should be made to the corresponding authors.

Competing interests

The authors declare no competing interests.

Footnotes

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

Contributor Information

Aly H. Abdeldaim, Email: aly.abdeldaim@diamond.ac.uk

Gøran J. Nilsen, Email: goran.nilsen@stfc.ac.uk

Lucy Clark, Email: l.m.clark@bham.ac.uk.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-024-53900-3.

References

  • 1.Haldane, F. D. M. Nonlinear field theory of large-spin Heisenberg antiferromagnets: Semiclassically quantized solitons of the one-dimensional easy-axis Néel state. Phys. Rev. Lett.50, 1153–1156 (1983). [Google Scholar]
  • 2.Kosterlitz, J. M. & Thouless, D. J. Ordering, metastability and phase transitions in two-dimensional systems. J. Phys. C6, 1181–1203 (1973). [DOI] [PubMed] [Google Scholar]
  • 3.Bernevig, B. A., Felser, C. & Beidenkopf, H. Progress and prospects in magnetic topological materials. Nature603, 41–51 (2022). [DOI] [PubMed] [Google Scholar]
  • 4.Balz, C. et al. Physical realization of a quantum spin liquid based on a complex frustration mechanism. Nat. Phys.12, 942–949 (2016). [Google Scholar]
  • 5.Broholm, C. et al. Quantum spin liquids. Science367, 0668 (2020). [DOI] [PubMed] [Google Scholar]
  • 6.Clark, L. & Abdeldaim, A. H. Quantum spin liquids from a materials perspective. Annu. Rev. Mater. Sci.51, 495–519 (2021). [Google Scholar]
  • 7.Savary, L. & Balents, L. Quantum spin liquids: A review. Rep. Prog. Phys.80, 016502 (2016). [DOI] [PubMed] [Google Scholar]
  • 8.Hermanns, M., Kimchi, I. & Knolle, J. Physics of the Kitaev model: Fractionalization, dynamic correlations, and material connections. Annu. Rev. Condens. Matter Phys.9, 17–33 (2018). [Google Scholar]
  • 9.Ran, Y., Hermele, M., Lee, P. A. & 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]
  • 10.Yan, S., Huse, D. A. & White, S. R. Spin-liquid ground state of the S = 1/2 kagome Heisenberg antiferromagnet. Science332, 1173–1176 (2011). [DOI] [PubMed] [Google Scholar]
  • 11.Norman, M. R. Colloquium: Herbertsmithite and the search for the quantum spin liquid. Rev. Mod. Phys.88, 041002 (2016). [Google Scholar]
  • 12.Han, T.-H. et al. Fractionalized excitations in the spin-liquid state of a kagome-lattice antiferromagnet. Nature492, 406–410 (2012). [DOI] [PubMed] [Google Scholar]
  • 13.Fu, M., Imai, T., Han, T.-H. & Lee, Y. S. Evidence for a gapped spin-liquid ground state in a kagome Heisenberg antiferromagnet. Science350, 655–658 (2015). [DOI] [PubMed] [Google Scholar]
  • 14.Khuntia, P. et al. Gapless ground state in the archetypal quantum kagome antiferromagnet ZnCu3(OH)6Cl2. Nat. Phys.16, 469–474 (2020). [Google Scholar]
  • 15.Kitaev, A. Anyons in an exactly solved model and beyond. Ann. Phys.321, 2–111 (2006). [Google Scholar]
  • 16.Winter, S. M. et al. Models and materials for generalized Kitaev magnetism. J. Phys. Condens. Matter.29, 493002 (2017). [DOI] [PubMed] [Google Scholar]
  • 17.Maksimov, P. A. & Chernyshev, A. L. Rethinking α-RuCl3. Phys. Rev. Res.2, 033011 (2020). [Google Scholar]
  • 18.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). [DOI] [PubMed] [Google Scholar]
  • 19.Stavropoulos, P. P., Pereira, D. & Kee, H.-Y. Microscopic mechanism for a higher-spin Kitaev model. Phys. Rev. Lett.123, 037203 (2019). [DOI] [PubMed] [Google Scholar]
  • 20.Liu, H. & Khaliullin, G. Pseudospin exchange interactions in d7 cobalt compounds: Possible realization of the Kitaev model. Phys. Rev. B97, 014407 (2018). [Google Scholar]
  • 21.Motome, Y., Sano, R., Jang, S., Sugita, Y. & Kato, Y. Materials design of Kitaev spin liquids beyond the Jackeli–Khaliullin mechanism. J. Phys. Condens. Matter.32, 404001 (2020). [DOI] [PubMed] [Google Scholar]
  • 22.Takagi, H., Takayama, T., Jackeli, G., Khaliullin, G. & Nagler, S. E. Concept and realization of Kitaev quantum spin liquids. Nat. Rev. Phys.1, 264–280 (2019). [Google Scholar]
  • 23.Kataoka, K. et al. Kitaev spin liquid candidate OsxCl3 comprised of honeycomb nano-domains. J. Phys. Soc. Japan89, 114709 (2020). [Google Scholar]
  • 24.Katukuri, V. M. et al. Strong magnetic frustration and anti-site disorder causing spin-glass behavior in honeycomb Li2RhO3. Sci. Rep.5, 14718 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Plumb, K. W. et al. α-RuCl3: A spin-orbit assisted Mott insulator on a honeycomb lattice. Phys. Rev. B90, 041112 (2014). [Google Scholar]
  • 26.Banerjee, A. et al. Proximate Kitaev quantum spin liquid behaviour in a honeycomb magnet. Nat. Mater.15, 733–740 (2016). [DOI] [PubMed] [Google Scholar]
  • 27.Johnson, R. D. et al. Monoclinic crystal structure of α-RuCl3 and the zigzag antiferromagnetic ground state. Phys. Rev. B92, 235119 (2015). [Google Scholar]
  • 28.Sears, J. A. et al. Magnetic order in α-RuCl3: A honeycomb-lattice quantum magnet with strong spin-orbit coupling. Phys. Rev. B91, 144420 (2015). [Google Scholar]
  • 29.Rousochatzakis, I., Perkins, N., Luo, Q., Kee, H.-Y. Beyond Kitaev physics in strong spin-orbit coupled magnets. Rep. Prog. Phys. (2024). [DOI] [PubMed]
  • 30.Yamada, M. G., Dwivedi, V. & Hermanns, M. Crystalline Kitaev spin liquids. Phys. Rev. B96, 155107 (2017). [Google Scholar]
  • 31.Yamada, M. G., Fujita, H. & Oshikawa, M. Designing Kitaev spin liquids in metal-organic frameworks. Phys. Rev. Lett.119, 057202 (2017). [DOI] [PubMed] [Google Scholar]
  • 32.Banerjee, A. et al. Excitations in the field-induced quantum spin liquid state of α-RuCl3. npj Quantum Mater.3, 8 (2018). [Google Scholar]
  • 33.Fukuoka, H., Imoto, H. & Saito, T. Synthesis and crystal structure of a new ruthenium silicophosphate: RuP3SiO11. J. Solid State Chem.121, 247–250 (1996). [Google Scholar]
  • 34.Suzuki, H. et al. Proximate ferromagnetic state in the Kitaev model material α-RuCl3. Nat. Commun.12, 4512 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Gretarsson, H. et al. Crystal-field splitting and correlation effect on the electronic structure of A2IrO3. Phys. Rev. Lett.110, 076402 (2013). [DOI] [PubMed] [Google Scholar]
  • 36.Takahashi, H. et al. Nonmagnetic J = 0 state and spin-orbit excitations in K2RuCl6. Phys. Rev. Lett.127, 227201 (2021). [DOI] [PubMed] [Google Scholar]
  • 37.Gretarsson, H. et al. J = 1/2 Pseudospins and d-p Hybridization in the Kitaev Spin Liquid Candidates RuX3 (X = Cl, Br, I) (2024).
  • 38.Li, Y., Winter, S. M., Kaib, D. A. S., Riedl, K. & Valentí, R. Modified Curie-Weiss law for jeff magnets. Phys. Rev. B103, 220408 (2021). [Google Scholar]
  • 39.Kim, S., Yuan, B. & Kim, Y.-J. α-RuCl3 and other Kitaev materials. APL Mater.10, 080903 (2022). [Google Scholar]
  • 40.Hickey, C. & Trebst, S. Emergence of a field-driven U(1) spin liquid in the Kitaev honeycomb model. Nat. Commun.10, 530 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Miller, S.C., Love, W.F. Tables of Irreducible Representations of Space Groups and Co-representations of Magnetic Space Groups. Pruett Press, Colorado (1967).
  • 42.Reger, J. D., Riera, J. A. & Young, A. P. Monte carlo simulations of the spin-1/2 Heisenberg antiferromagnet in two dimensions. J. Phys. Condens. Matter.1, 1855–1865 (1989). [Google Scholar]
  • 43.Cao, H. B. et al. Low-temperature crystal and magnetic structure of α-RuCl3. Phys. Rev. B93, 134423 (2016). [Google Scholar]
  • 44.Trebst, S. & Hickey, C. Kitaev materials. Phys. Rep.950, 1–37 (2022). [Google Scholar]
  • 45.Baek, S.-H. et al. Evidence for a field-induced quantum spin liquid in α-RuCl3. Phys. Rev. Lett.119, 037201 (2017). [DOI] [PubMed] [Google Scholar]
  • 46.Hentrich, R. et al. High-field thermal transport properties of the Kitaev quantum magnet α-RuCl3: Evidence for low-energy excitations beyond the critical field. Phys. Rev. B102, 235155 (2020). [Google Scholar]
  • 47.Chaloupka, J. & Khaliullin, G. Hidden symmetries of the extended Kitaev-heisenberg model: Implications for the honeycomb-lattice iridates A2IrO3. Phys. Rev. B92, 024413 (2015). [Google Scholar]
  • 48.Bai, X. et al. Instabilities of heavy magnons in an anisotropic magnet. Nat. Commun.14, 4199 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Kim, C. et al. Bond-dependent anisotropy and magnon decay in cobalt-based Kitaev triangular antiferromagnet. Nat. Phys.19, 1624–1629 (2023). [Google Scholar]
  • 50.Taylor, J. H. & Müller, G. Limitations of spin-wave theory in T = 0 spin dynamics. Phys. Rev. B28, 1529–1533 (1983). [Google Scholar]
  • 51.Zhitomirsky, M. E. & Chernyshev, A. L. Colloquium: Spontaneous magnon decays. Rev. Mod. Phys.85, 219–242 (2013). [Google Scholar]
  • 52.Rau, J. G., Moessner, R. & McClarty, P. A. Magnon interactions in the frustrated pyrochlore ferromagnet yb2ti2o7. Phys. Rev. B100, 104423 (2019). [Google Scholar]
  • 53.Winter, S. M. et al. Breakdown of magnons in a strongly spin-orbital coupled magnet. Nat. Commun.8, 1152 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Halloran, T. et al. Geometrical frustration versus Kitaev interactions in BaCo2(AsO4)2. Proc. Natl. Acad. Sci. USA.120, 2215509119 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Liu, H., Chaloupka, J. & Khaliullin, G. Exchange interactions in d5 Kitaev materials: From Na2IrO3 to α-RuCl3. Phys. Rev. B105, 214411 (2022). [Google Scholar]
  • 56.Kim, J. et al. Dynamic spin correlations in the honeycomb lattice Na2IrO3 measured by resonant inelastic X-ray scattering. Phys. Rev. X10, 021034 (2020). [Google Scholar]
  • 57.Rau, J.G., Kee, H.-Y. Trigonal distortion in the honeycomb iridates: Proximity of zigzag and spiral phases in Na2IrO3 (2014).
  • 58.Kaib, D. A. S. et al. Electronic and magnetic properties of the RuX3 (X = Cl, Br, I) family: two siblings—and a cousin? npj Quantum Materials7, 75 (2022). [Google Scholar]
  • 59.Fukuoka, H., Imoto, H. & Saito, T. New polymorphs of RuIIIP3O9: Cyclo-hexaphosphate Ru2P6O18 and metaphosphate Ru(PO3)3 with a novel structure. J. Solid State Chem.119, 107–114 (1995). [Google Scholar]
  • 60.Toby, B. H. & Von Dreele, R. B. GSAS-II: The Genesis of a Modern Open-source All Purpose Crystallography Software Package. J. Appl. Crystallogr.46, 544–549 (2013). [Google Scholar]
  • 61.Perez-Mato, J. M. et al. Symmetry-based computational tools for magnetic crystallography. Annu. Rev. Mater. Res.45, 217–248 (2015). [Google Scholar]
  • 62.Rodríguez-Carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Phys. B: Condens. Matter.192, 55–69 (1993). [Google Scholar]
  • 63.Cromer, D. T. & Waber, J. T. Scattering factors computed from relativistic Dirac–Slater wave functions. Acta Crystallogr.18, 104–109 (1965). [Google Scholar]
  • 64.Haverkort, M. W. Quanty for core level spectroscopy - excitons, resonances and band excitations in time and frequency domain. J. Phys. Conf. Ser.712, 012001 (2016). [Google Scholar]
  • 65.Yang, Z. et al. Resonant inelastic x-ray scattering from electronic excitations in α-RuCl3 nanolayers. Phys. Rev. B108, 041406 (2023). [Google Scholar]
  • 66.Winter, S. M., Li, Y., Jeschke, H. O. & Valentí, R. Challenges in design of Kitaev materials: Magnetic interactions from competing energy scales. Phys. Rev. B93, 214431 (2016). [Google Scholar]
  • 67.Toth, S. & Lake, B. Linear spin wave theory for single-Q incommensurate magnetic structures. J. Phys. Condens. Matter.27, 166002 (2015). [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Data Availability Statement

The data that support the findings of this study are available via the following: https://syncandshare.desy.de/index.php/s/xZ2Jwrzw9mpDkxN (IRIXS, DESY), 10.5286/ISIS.E.RB2210080 (WISH, ISIS), 10.5286/ISIS.E.RB2210081 (LET, ISIS). 10.5281/zenodo.13890183 (I11, Diamond, magnetization, specific heat, LSWT grids). Any requests for further data, analysis or code should be made to the corresponding authors.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES