This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Evidence of quantum spin liquid state in a Cu2+-based S=1/2S=1/2 triangular lattice antiferromagnet

K. Bhattacharya Department of Physics, Shiv Nadar Institution of Eminence, Gautam Buddha Nagar, UP 201314, India    S. Mohanty School of Physics, Indian Institute of Science Education and Research, Thiruvananthapuram 695551, India    A. D. Hillier ISIS Facility, STFC Rutherford Appleton Laboratory, Chilton, Oxfordshire OX11 0QX, United Kingdom    M. T. F. Telling ISIS Facility, STFC Rutherford Appleton Laboratory, Chilton, Oxfordshire OX11 0QX, United Kingdom    R. Nath School of Physics, Indian Institute of Science Education and Research, Thiruvananthapuram 695551, India    M. Majumder [email protected] Department of Physics, Shiv Nadar Institution of Eminence, Gautam Buddha Nagar, UP 201314, India
Abstract

The layered triangular lattice owing to 1:21:2 order of BB and BB^{\prime} sites in the triple perovskite A3BB2A_{3}BB^{\prime}_{2}O9 family provides an enticing domain for exploring the complex phenomena of quantum spin liquids (QSLs). We report a comprehensive investigation of the ground state properties of Sr3CuTa2O9 that belongs to the above family, by employing magnetization, specific heat, and muon spin relaxation (μ\muSR) experiments down to the lowest temperature of 0.1 K. Analysis of the magnetic susceptibility indicates that the spin-lattice is a nearly isotropic S=1/2S=1/2 triangular lattice. We illustrate the observation of a gapless QSL, in which conventional spin ordering or freezing effects are absent, even at temperatures more than two orders of magnitude smaller than the exchange energy (JCW/kB5.04J_{\rm CW}/k_{\rm B}\simeq-5.04 K). Magnetic specific heat in zero-field follows a power law, CmTηC_{\rm m}\sim T^{\eta}, below 1.2 K with η2/3\eta\approx 2/3, which is consistent with a theoretical proposal of the presence of spinon Fermi surface. Below 1.2 K, the μ\muSR relaxation rate shows no temperature dependence, suggesting persistent spin dynamics as expected for a QSL state. Delving deeper, we also analyze longitudinal field μ\muSR spectra revealing strong dynamical correlations in the spin-disordered ground state. All of these highlight the characteristics of spin entanglement in the QSL state.

Introduction: Quantum spin liquid (QSL) is an exotic and highly entangled quantum state with no spontaneous symmetry breaking down to absolute zero temperature despite strong correlations among spins. Such a quantum phase is characterized by an emergent gauge field and fractional excitations, called spinons [1, 2]. This state was first theoretically predicted by P. W. Anderson as a resonating valance bond (RVB) state for interacting Heisenberg spins in a two-dimensional (2D) triangular lattice antiferromagnet (TLAF) [3]. Subsequently, it was recognized that the true ground state of the isotropic Heisenberg TLAF is a three sub-lattice 120 Ne´\acute{e}el order [4, 5]. Thereafter, considerable effort has been dedicated to stabilizing the QSL state in a Heisenberg TLAF. For instance, (i) in an isotropic TLAF with Heisenberg interactions, the competing nearest-neighbor (NN) (J1J_{1}) and next-nearest-neighbor (NNN) (J2J_{2}) interactions can stabilize a gapless Dirac QSL state for 0.08J2/J10.160.08\lesssim J_{2}/J_{1}\lesssim 0.16 that is sandwiched between the 120 Ne´\acute{e}el order and stripe state in the phase diagram [6, 7]. A tiny J3J_{3} can also destabilize the magnetic long-range-order (LRO) leading to a QSL phase in this model [8]. However, there are theoretical results that illustrate the importance of third nearest-neighbor exchange coupling (J3J_{3}) to achieve QSL state. [9], (ii) Considering spatially anisotropic Heisenberg interactions (JJ and JJ^{\prime}), two different types of QSL states are favored: gapless QSL for J/J0.65J^{\prime}/J\lesssim 0.65 and gapped QSL for 0.65J/J0.80.65\lesssim J^{\prime}/J\lesssim 0.8 [10].

Over the years, the pursuit of new TLAFs with low spin values (e.g. S=1/2S=1/2) has become a focal area of research, as it is believed that QSL is a manifestation of enhanced quantum fluctuations and magnetic frustration. However, the experimental realization of QSL in S=1/2S=1/2 Heisenberg TLAFs is scarce due to the availability of limited model materials. Typically, the majority of the Heisenberg TLAFs show magnetic LRO at finite temperatures due to the inter-layer couplings and exchange anisotropy which are inherently present in real materials [11, 12, 13, 14, 15, 16, 17]. To the best of our knowledge as far as inorganic compounds are concerned, QSL in a Cu2+-based S=1/2S=1/2 Heisenberg TLAF is realized only in Sr3CuSb2O9 [18] reported to date (Note that, Ba3CuSb2O9, likewise exhibits a QSL ground state; however, there is still disagreement over the arrangement of Cu2+ ions, specifically whether it is triangular or Honeycomb [19, 20, 21].). Sr3CuSb2O9 belongs to the triple perovskite family. The family of triple perovskite (A3BB2A_{3}BB^{\prime}_{2}O9 with AA = Sr/Ba, BB = Cu/Ca/Te/Os, BB^{\prime}= Sb/Ru/Ir/Fe) [22, 23, 24, 25, 26] compounds are interesting since some of the family members have 1:21:2 ordering of B/BB/B^{\prime} sites, owing to the site sharing of BB and BB^{\prime} elements [23, 24]. Depending on their crystal structure and space group, the BB site could form a superlattice structure with a specific propagation vector. For Sr3CuSb2O9, Mahajan et. al. have shown that it forms a superlattice structure with a propagation vector (13\frac{1}{3}, 13\frac{1}{3}, 13\frac{1}{3}) and effectively forms an edge-shared triangular lattice on the (111) plane. In this pseudo-cubic system, the Cu2+ and Sb5+ planes are present in 1:21:2 order and the compound evinces QSL state [18]. Thus, this family of compounds holds great potential to showcase quantum magnetism due to frustration. With this motivation, we investigated the ground state of another member of this family Sr3CuTa2O9 (SCTO), where the S=1/2S=1/2 moments are embedded in an edge-shared triangular lattice. From a detailed experimental investigation by employing magnetization, specific heat, and muon spin relaxation/rotation (μ\muSR) techniques, we provide solid evidence of a gapless QSL state in SCTO. This compound serves as a model isotropic edge-shared TLAF that hosts QSL.

Refer to caption
Figure 1: Rietveld refinement of the powder XRD pattern taking the space group P1¯P\bar{1}. Inset: depicts Cu-layers arranged in a triangular lattice. The possible nearest-neighbor (J1J_{1} and J2J_{2}), next-nearest-neighbor (JnnJ_{nn}), and inter-plane (JJ_{\perp}) exchange couplings are illustrated.

Powder x-ray diffraction: Polycrystalline SCTO sample was synthesized following the procedure described in the supplementary material (SM) [27]. Rietveld refinement of the powder x-ray diffraction (XRD) pattern of SCTO at room temperature was carried out assuming a tetragonal space group P4/mmmP4/mmm (No. 123) with lattice parameters a=b=3.93a=b=3.93 and c=4.14c=4.14 Å, which is a pseudocubical structure (details are in SM [27]). The best refinement was achieved with a goodness-of-fit χ25.06\chi^{2}\sim 5.06. Some peaks at lower angles (2θ<302\theta<30^{\circ}) couldn’t be indexed with this space group as shown in SM [27]. These peaks correspond to a kk-value of 13(111)\frac{1}{3}(111). Thus, these peaks are associated with the superlattice structure owing to the inter-site mixing of Cu2+/Ta5+ with occupancies 13/23\frac{1}{3}/\frac{2}{3} and form a 1:21:2 order, which is common in this triple perovskite systems. Note that the presence of superlattice structure is dependent on the synthesis conditions [28]. Superlattice peaks along with Bragg peaks of the P4/mmmP4/mmm space group are well fitted with a lower symmetry triclinic space group P1¯P\bar{1} (yield a goodness-of-fit χ21.49\chi^{2}\sim 1.49) as shown in Fig. 1. The 1:2 site ordering with propagation vector 13(111)\frac{1}{3}(111) in a pseudocubic structure forms a layered edged-shared triangular lattice of Cu2+ in the (111) plane. The two successive triangular layers of Cu2+ are separated by two layers of Ta5+ as shown in the inset of Fig. 1. A similar structure is also found in Sr3CuSb2O9 [18]. Note that the Cu2+ moments form an edge-shared bilateral or quasi-equilateral triangle with a very small difference in the bond lengths (5.56, 5.70, and 5.70 Å) along the edges of the triangle. The smallest bond length between inter-plane Cu2+ ions is about 6.93\sim 6.93 Å.

Refer to caption
Figure 2: Left yy-axis: χ\chi as a function of TT measured at μ0H=1\mu_{0}H=1 T. The solid line represents the fit using an anisotropic (J1J2J_{1}-J_{2}) S=1/2S=1/2 TLAF model [Eq. (1)]. Right yy-axis: inverse-susceptibility (after subtracting χ0\chi_{0}) as a function of TT at μ0H=1\mu_{0}H=1 T and solid line is the Curie-Weiss fit.

Magnetization: Temperature dependent dc magnetic susceptibility χ(T)(M/H\chi(T)(\equiv M/H) measured at μ0H=1\mu_{0}H=1 T is depicted in Fig. 2. χ\chi increases monotonically with lowering temperature and no anomaly associated with magnetic LRO is observed down to 1.85 K. χ(T)\chi(T) measured in zero-field-cooled and field-cooled protocols in a low magnetic field (50 Oe) shows no bifurcation down to 1.85 K [27], ruling out a spin-glass like transition. In Fig. 2, we fitted the χ(T)\chi(T) data in high temperatures (T>100T>100 K) by the modified Curie-Weiss (CW) law [χ(T)=χ0+C/(TθCW)\chi(T)=\chi_{0}+C/(T-\theta_{\rm CW})] that yields temperature independent susceptibility χ02.18×105\chi_{0}\simeq-2.18\times 10^{-5} emu/mol-Cu2+, Curie constant C0.49C\simeq 0.49 emu.K/mol-Cu2+, and characteristic CW temperature θCW7.56\theta_{\rm CW}\simeq-7.56 ±0.11\pm 0.11 K. The negative sign of θCW\theta_{\rm CW} indicates dominant antiferromagnetic (AFM) interaction between Cu2+ spins. From the value of CC, the effective moment is estimated to be μeff=3kBC/NA1.97(8)μB\mu_{\rm eff}=\sqrt{3k_{\rm B}C/N_{\rm A}}\simeq 1.97(8)\mu_{\rm B}/Cu2+ (where, kBk_{\rm B}, NAN_{\rm A}, μB\mu_{\rm B}, and gg are the Boltzmann constant, Avogadro’s number, Bohr magneton, and Landé gg-factor, respectively). This value is indeed comparable with the expected value of μeff1.73μB\mu_{\rm eff}\simeq 1.73~{}\mu_{\rm B}/Cu2+ [=gS(S+1)μB=g\sqrt{S(S+1)}\mu_{\rm B}] for Cu2+ (S=1/2S=1/2). Here, θCW\theta_{\rm CW} represents the overall energy scale of the exchange couplings and one can estimate the average intra-layer coupling (JCW/kBJ_{\rm CW}/k_{\rm B}) as θCW=zJCWS(S+1)/3kB\theta_{\rm CW}=-zJ_{\rm CW}S(S+1)/3k_{\rm B}. Taking the experimental value of θCW\theta_{\rm CW} and the number of NN spins z=6z=6 for a 2D-TLAF, we got JCW/kB5.04J_{\rm CW}/k_{\rm B}\simeq-5.04 K.

As inferred from the crystal structure, one expects a spatial anisotropy in the triangular unit due to a slightly varying bond length. Therefore, to analyze χ(T)\chi(T) we fitted the data in the high-temperature region (T>100T>100 K) by χ(T)=χ0+χTLAF(T)\chi(T)=\chi_{0}+\chi_{\rm TLAF}(T). χTLAF\chi_{\rm TLAF} is the expression of high-temperature series expansion (HTSE) for a S=1/2S=1/2 2D spatially anisotropic TLAF which has the form [29]

χTLAF=NAg2μB2kBT×n=0(J2T)nm=0ncm,nym4n+1n!.\chi_{\rm TLAF}=\frac{N_{\rm A}g^{2}\mu_{\rm B}^{2}}{k_{\rm B}T}\times\sum_{\rm n=0}\left(\frac{J_{2}}{T}\right)^{\rm n}\sum_{\rm m=0}^{\rm n}\frac{c_{\rm m,n}y^{\rm m}}{4^{\rm n+1}{\rm n}!}. (1)

Here, y=J1/J2y=J_{1}/J_{2} with J1J_{1} and J2J_{2} representing the NN and NNN exchange interactions, respectively. The integer coefficients cm,nc_{m,n} are given in Ref. [29]. By fixing the χ0\chi_{0} value (obtained from CW fit), the fit yields J2/kB6.08J_{2}/k_{\rm B}\simeq 6.08 ±0.35\pm 0.35 K, and J2/J11J_{2}/J_{1}\simeq 1. The value J2/J11J_{2}/J_{1}\simeq 1 indicates an effective equilateral triangular lattice as far as the strength of exchange interactions are concerned. Moreover, the obtained value of J2/kBJ_{2}/k_{\rm B} is in good agreement with JCW/kBJ_{\rm CW}/k_{\rm B}, further endorsing a nearly isotropic triangular lattice. We also fitted the χ(T)\chi(T) data by the HTSE of a S=1/2S=1/2 isotropic TLAF that uses Padé approximation [27]. The fit results in nearly identical exchange coupling, Jiso/kB6.09J_{\rm iso}/k_{\rm B}\simeq 6.09 ±0.40\pm 0.40 K.

Refer to caption
Figure 3: (a) Magnetic specific heat (CmC_{\rm m}) as a function of TT for different fields. The solid lines are the fits using CmγTηC_{\rm m}\sim\gamma T^{\eta} and red dashed lines correspond to T0.66T^{0.66} and T1.46T^{1.46}. (b) Variation of η\eta and γ\gamma with the magnetic field. (c) Magnetic entropy (ΔSm\Delta S_{\rm m}) as a function of TT and the dotted lines mark the theoretically expected and experimentally observed values of entropy.

Specific Heat: Specific heat not only provides information about magnetic LRO but also the low-energy excitations of a spin system. Specific heat data in zero-field as well as in applied fields show the absence of a λ\lambda-type anomaly reminiscent of a magnetic LRO down to 0.33 K. To extract the magnetic contribution to the specific heat (CmC_{\rm m}), we have subtracted the lattice contribution (ClatC_{\rm lat}) from the total specific heat (CpC_{\rm p}). The lattice contribution was estimated considering both the Debye and Einstein models in the temperature range of 3019030-190 K with one Debye and two Einstein terms (see SM) [27]. CmC_{\rm m} vs TT presented in Fig. 3(a) features a broad hump at around 4\sim 4 K (ThumpT_{hump}), which indicates crossover from a thermally disordered paramagnet to a quantum paramagnetic QSL state, typically observed in QSL systems [30, 31]. Moreover, for a Heisenberg TLAF, such a hump is expected at Thump/J0.55T_{hump}/J\simeq 0.55 [32], which yields J/kB7.27J/k_{B}\approx 7.27 K consistent with the JJ evaluated from magnetization measurements. For T1.2T\leq 1.2 K, CmC_{\rm m} measured in zero-field follows a power-law behavior CmγTηC_{\rm m}\sim\gamma T^{\eta} with η=2/3\eta=2/3. While for different gapless QSL candidates, CmC_{m} follows either linear [20, 33] or quadratic behavior with the temperature (for Dirac spin liquid [18, 34]), in contrast to that for SCTO η=2/3\eta=2/3 is quite rare, and interestingly, is indeed consistent with a theoretical prediction for an equilateral TLAF with the presence of gapless Fermi surface as CmkBν0tspinon1/3(kBT)2/3C_{\rm m}\sim k_{\rm B}\nu_{0}t_{\rm spinon}^{1/3}(k_{\rm B}T)^{2/3}. Here, ν0\nu_{0} is the density of states at the spinon Fermi surface, and tspinont_{\rm spinon} is the spinon hopping amplitude [35]. The obtained field dependence of η\eta and γ\gamma are shown in Fig. 3(b). With the application of a magnetic field, the exponent η\eta grows slowly, probably due to the gradual suppression of 2D quantum correlations. We also calculated the magnetic entropy change (ΔSm\Delta S_{\rm m}) for different fields as shown in Fig. 3(c). ΔSm\Delta S_{\rm m} recovers only 81\sim 81% of Rln2ln2 above 10 K. A signature of frustrated magnet is that the entropy releases over a broad range of temperatures, thus the remaining 19% of entropy will be released at further low temperatures, which means persistence of strong correlation among the Cu2+ spins are present at further low temperatures. Such a reduction in entropy has been observed in other frustrated magnets [36, 37, 38].

In addition, we have also measured temperature-dependent thermal conductivity κ(T)\kappa(T) down to T=2T=2 K in different magnetic fields (see SM) [27]. The low-temperature data are fitted by κ/T=a+bT2\kappa/T=a+bT^{2} that yields a non-zero intercept a0.0211a\simeq 0.0211 mWK-2cm-1 which further corroborate a gapless QSL state with a finite spinon density of states. However, a detailed quantitative analysis and solid evidence of QSL from thermal conductivity measurements require data at ultra-low temperatures.

Refer to caption
Figure 4: (a) ZF-μ\muSR asymmetry spectra as a function of tt for different temperatures. Solid lines are the fits using Eq. (2). (b) ZF relaxation rate (λZF\lambda_{\rm ZF}) vs TT. (c) βZF\beta_{\rm ZF} as a function of TT. The orange-shaded region represents the crossover region (1.2T101.2\leq T\leq 10 K) from paramagnetic to the QSL state. (d) μ\muSR asymmetry spectra measured in different longitudinal fields at T=0.1T=0.1 K and the solid lines represent the fits using Eq. (2). (e) The corresponding relaxation rate λLF\lambda_{\rm LF} as a function of the longitudinal field. (f) βLF\beta_{\rm LF} as a function of the longitudinal field.

Muon Spin Relaxation (μ\muSR): Being sensitive to a local internal magnetic field as small as 0.1 Oe, μ\muSR is an ideal microscopic tool to probe magnetic LRO. In addition, μ\muSR can distinguish between static and dynamic correlations among the spins making it a powerful tool to uncover the putative QSL behaviour. Hence, to elucidate the magnetic ground state of SCTO, we collected μ\muSR data in zero-field (ZF) as a function of temperature as well as in longitudinal fields (LFs) at the base temperature of 0.1 K. In the following, we delineate our observations from the μ\muSR data.

(i) ZF μ\muSR asymmetries displayed in Fig. 4(a) confirm the absence of magnetic LRO down to 0.1 K, as they decay continuously without any oscillations or initial aymmetry drop [Fig. 4(a)]. The ZF μ\muSR asymmetries are well-fitted by a stretch exponential function with an extra background term BbgB_{\rm bg} (owing to some muons missing the sample and placed at the Ag-sample holder and the cryostat wall.) [39] as

P(t)=P(0)exp(λt)β+Bbg.P(t)=P(0)\exp{(-\lambda t)^{\beta}}+B_{\rm bg}. (2)

Here, P(0)0.28P(0)\sim 0.28 (weakly temperature and field dependent) is the initial asymmetry, λ\lambda is the relaxation rate, and β\beta is the stretching exponent. While fitting the ZF μ\muSR asymmetries, Bbg0.08B_{\rm bg}\approx 0.08 is kept constant for all the temperatures. The obtained fitting parameters (λZF\lambda_{\rm ZF} and βZF\beta_{\rm ZF}) are plotted in Fig. 4(b and c).

(ii) In the high temperature (T>10T>10 K) paramagnetic regime, the uncorrelated Cu2+ spins fluctuate rapidly and randomly. The fluctuation rate can be calculated using the strength of exchange coupling (J/kB5.04J/k_{\rm B}\simeq 5.04 K) as ν=zJs/1.6×1010\nu=\sqrt{z}Js/\hbar\sim 1.6\times 10^{10} Hz [40, 41], where ν\nu is the spin fluctuation rate and z=6z=6 is the coordination number for a 2D TLAF. λZF\lambda_{\rm ZF} (at T>10T>10 K) can provide an idea about the local internal field distribution (Δ\Delta) based on the Redfield formula [42], λ(T10K,H)=2Δ2ν/(ν2+γμ2μ02H2)\lambda(T\geq 10~{}K,H)=2\Delta^{2}\nu/(\nu^{2}+\gamma^{2}_{\mu}\mu^{2}_{0}H^{2}), which yields Δ5.1×107\Delta\sim 5.1\times 10^{7} Hz. This confirms the fast-fluctuation limit (Δν\Delta\ll\nu[43, 41] is responsible for the temperature-independent relaxation rate in the high-temperature range. With lowering the temperature, the relaxation rate λZF\lambda_{\rm ZF} starts increasing below T10T\leq 10 K, which implies slowing down of fluctuating moments due to the growth of correlations among the Cu2+ spins [43, 44]. This temperature regime is marked as the crossover region from paramagnetic to QSL state in Fig. 4(b), consistent with the broad hump observed in Cm(T)C_{\rm m}(T) [see Fig. 3(a)]. With further lowering in temperature, λZF\lambda_{\rm ZF} becomes temperature independent (for T1.2T\leq 1.2 K), indicating the persistence of spin dynamics. This persistent spin dynamics is considered to be a generic feature of QSL, as reported for other celebrated QSL candidates [39, 18, 30, 45]. It is interesting to note that the saturation of λZF\lambda_{\rm ZF} below 1.2\sim 1.2 K exactly coincides with the temperature range where Cm(T)C_{\rm m}(T) follows a power-law behavior. Thus, the behavior of λZF(T)\lambda_{\rm ZF}(T) along with the power-law dependent of Cm(T)C_{\rm m}(T) provides a robust signature of QSL at low temperatures.

(iii) LF μ\muSR experiments are performed to explore the nature of the spin dynamics at low temperatures. LF-μ\muSR asymmetry spectra measured in different fields at T=0.1T=0.1 K are shown in Fig. 4(d). They are fitted well by Eq. (2). The obtained field dependence of λLF\lambda_{\rm LF} and βLF\beta_{\rm LF} are depicted in Fig. 4(e) and (f), respectively. The quick suppression of λLF\lambda_{\rm LF} below 100 Oe reflects the decoupling of the nuclear contribution to the relaxation rate [46, 45]. More interestingly, above 0.01 T, the remaining λLF0.05μs1\lambda_{\rm LF}\sim 0.05~{}\mu s^{-1}, originating from the electronic contribution, is almost field independent up to a LF of 3000 Oe. At low temperatures (T1.2T\leq 1.2 K), if we assume that the relaxation process in the plateau regime of λZF\lambda_{\rm ZF} is due to a static local field then it would correspond to 0.25\sim 0.25 mT (as Bloc=λ/γμB_{\rm loc}=\lambda/\gamma_{\mu}, where γμ=2π×135.5s1μT1\gamma_{\mu}=2\pi\times 135.5~{}s^{-1}\mu T^{-1} is the gyromagnetic ratio of muons). In such a scenario, an LF of 750 Oe, which is 5 times larger than BlocB_{\rm loc} would be sufficient to decouple the relaxation channel. On the contrary, even in an LF of 3000 Oe, the decoupling of the relaxation channel was not achieved. One may require higher LFs to completely decouple the relaxation channel [47]. This observation demonstrates that the correlation among the spins is entirely dynamic (not static) in nature, as expected for a QSL state [18, 48, 39, 49, 50].

(iv) Typically, in a spin-glass state, the value of β\beta is predicted to be about 1/3 [51, 52]. On the contrary, the obtained β\beta value shown in Fig. 4(c) attains a constant value of 1.2\sim 1.2 below the crossover region. Furthermore, the magnitude of λZF\lambda_{\rm ZF} (0.12μs1\sim 0.12~{}\mu s^{-1} at T10T\geq 10 K) becomes double (0.22μs1\sim 0.22~{}\mu s^{-1} at T1T\leq 1 K) upon cooling the system below the crossover regime, which is also in contrast to that expected for a spin-glass type transition, where the relaxation rate should increase by few orders of magnitude [53, 41]. These observations rule out the possibility of a spin-glass state.

It is important to highlight that the system doesn’t order down to Tmin0.1T_{\rm min}\sim 0.1 K that sets the lower limit of the frustration parameter f=θCW/Tmin76f=\theta_{\rm CW}/T_{\rm min}\simeq 76, characterizing SCTO a highly frustrated magnet. Specific heat and μ\muSR results establish a highly dynamic ground state with spinon excitations and a footprint of gapless QSL. The analysis of magnetic susceptibility suggests that the system can be treated as an isotropic TLAF with Heisenberg interactions. In such a scenario, the ground state is expected to be a 120 Ne´\acute{e}el order rather than a QSL state, if only the NN couplings are considered. Further, despite partial site occupancy, the Cu and Ta sites are arranged periodically with 1:2 order, forming separate layers of Cu2+ with a triangular lattice. This ensures the minimal effect of disorder driving QSL state. Note that, systems with disorder may stabilize random-singlet-state, which exhibits scaling behavior in their physical properties as observed in different compounds with site disorder [54, 55, 56]. However, as expected the physical properties of SCTO do not show any such scaling behavior. All these observations reflect the possible role of NNN interaction (JnnJ_{\rm nn}) or inter-layer interaction (JJ_{\perp}) [inset (b) of Fig. 1] in stabilizing the gapless QSL state. Nevertheless, the inelastic neutron scattering experiments on a good-quality single crystal would be essential to shed light on this issue.

Conclusion: In summary, our studies divulge that the Cu2+ and Ta5+ ions feature a precisely calibrated 1:2 site ordering, resulting in seperate planes of Cu2+ moments with propagation vector k=13(111)k=\frac{1}{3}(111) which creates effective equilateral triangular lattice. Magnetic susceptibility data agree well with the isotropic S=1/2S=1/2 TLAF model with a leading exchange coupling of J/kB6.09J/k_{\rm B}\simeq 6.09 K. Specific heat data suggest the absence of magnetic LRO down to 0.33 K and provide evidence for spinon excitations. The absence of magnetic LRO was further corroborated by the μ\muSR data measured down to 0.1 K, setting a very high degree of magnetic frustration (f>76f>76) in SCTO. From the μ\muSR analysis, the ground state is found to be a highly dynamic state with no static order, a hallmark of QSL. Thus, our detailed investigation unambiguously established that SCTO is one of the rare examples of a gapless QSL realized in a S=1/2S=1/2 Cu2+-based Heisenberg TLAF. The onset of such a gapless QSL is anticipated to be due to the complex interplay of different exchange couplings (NN and interactions beyond NN). We believe that our results would instigate further experimental as well as theoretical investigations to settle the origin of the emergence of QSL in SCTO.

Acknowledgment: K.B. and M.M. would like to thank the Department of Science and Technology, India for the access to the experimental facility and financial support for the experiment conducted at ISIS muon source [57] and Jawaharlal Nehru Centre for Advanced Scientific Research (JNCASR) for managing the project. S.M. and R.N. would like to acknowledge SERB, India, bearing sanction Grant No. CRG/2022/000997. K.B. and M.M. also thank I. Ishant for fruitful discussions.

References

  • Balents [2010] L. Balents, Spin liquids in frustrated magnets, Nature 464, 199 (2010).
  • Lancaster [2023] T. Lancaster, Quantum spin liquids, Contemporary Physics 64, 127 (2023).
  • Anderson [1973] P. Anderson, Resonating valence bonds: A new kind of insulator ?, Materials Research Bulletin 8, 153 (1973).
  • Huse and Elser [1988] D. A. Huse and V. Elser, Simple variational wave functions for two-dimensional heisenberg spin-½ antiferromagnets, Phys. Rev. Lett. 60, 2531 (1988).
  • Jolicoeur and Le Guillou [1989] T. Jolicoeur and J. C. Le Guillou, Spin-wave results for the triangular heisenberg antiferromagnet, Phys. Rev. B 40, 2727 (1989).
  • Sherman et al. [2023] N. E. Sherman, M. Dupont, and J. E. Moore, Spectral function of the J1J2{J}_{1}-{J}_{2} H\mathrm{H}eisenberg model on the triangular lattice, Phys. Rev. B 107, 165146 (2023).
  • Drescher et al. [2023] M. Drescher, L. Vanderstraeten, R. Moessner, and F. Pollmann, Dynamical signatures of symmetry-broken and liquid phases in an S=12\mathrm{S}=\frac{1}{2} H\mathrm{H}eisenberg antiferromagnet on the triangular lattice, Phys. Rev. B 108, L220401 (2023).
  • Merino et al. [2014] J. Merino, M. Holt, and B. J. Powell, Spin-liquid phase in a spatially anisotropic frustrated antiferromagnet: A\mathrm{A} S\mathrm{S}chwinger boson mean-field approach, Phys. Rev. B 89, 245112 (2014).
  • Jiang and Jiang [2023] Y.-F. Jiang and H.-C. Jiang, Nature of quantum spin liquids of the S=12\mathrm{S}=\frac{1}{2} H\mathrm{H}eisenberg antiferromagnet on the triangular lattice: A parallel DMRG\mathrm{DMRG} study, Phys. Rev. B 107, L140411 (2023).
  • Yunoki and Sorella [2006] S. Yunoki and S. Sorella, Two spin liquid phases in the spatially anisotropic triangular H\mathrm{H}eisenberg model, Phys. Rev. B 74, 014408 (2006).
  • Lal et al. [2023] S. Lal, S. J. Sebastian, S. S. Islam, M. P. Saravanan, M. Uhlarz, Y. Skourski, and R. Nath, Double magnetic transitions and exotic field-induced phase in the triangular lattice antiferromagnets Sr3Co(Nb,Ta)2O9\mathrm{Sr}_{3}\mathrm{Co(Nb,Ta)}_{2}\mathrm{O}_{9}Phys. Rev. B 108, 014429 (2023).
  • Li et al. [2020] N. Li, Q. Huang, X. Y. Yue, W. J. Chu, Q. Chen, E. S. Choi, X. Zhao, H. D. Zhou, and X. F. Sun, Possible itinerant excitations and quantum spin state transitions in the effective spin-1/2 triangular-lattice antiferromagnet Na2BaCo(PO4)2\mathrm{Na}_{2}\mathrm{BaCo(PO}_{4})_{2}Nat. Commun. 11, 4216 (2020).
  • Kamiya et al. [2018] Y. Kamiya, L. Ge, T. Hong, Y. Qiu, D. L. Quintero-Castro, Z. Lu, H. B. Cao, M. Matsuda, E. S. Choi, C. D. Batista, M. Mourigal, H. D. Zhou, and J. Ma, The nature of spin excitations in the one-third magnetization plateau phase of Ba3CoSb2O9\mathrm{Ba}_{3}\mathrm{CoSb}_{2}\mathrm{O}_{9}Nat. Commun. 9, 2666 (2018).
  • Yokota et al. [2014] K. Yokota, N. Kurita, and H. Tanaka, Magnetic phase diagram of the S=12\mathrm{S}=\frac{1}{2} triangular-lattice hheisenberg antiferromagnet Ba3CoNb2O9\mathrm{Ba}_{3}\mathrm{CoNb}_{2}\mathrm{O}_{9}Phys. Rev. B 90, 014403 (2014).
  • Kojima et al. [2018] Y. Kojima, M. Watanabe, N. Kurita, H. Tanaka, A. Matsuo, K. Kindo, and M. Avdeev, Quantum magnetic properties of the spin-12\frac{1}{2} triangular-lattice antiferromagnet Ba2La2CoTe2O12\mathrm{Ba}_{2}\mathrm{La}_{2}\mathrm{CoTe}_{2}\mathrm{O}_{12}Phys. Rev. B 98, 174406 (2018).
  • Ranjith et al. [2015] K. M. Ranjith, R. Nath, M. Skoulatos, L. Keller, D. Kasinathan, Y. Skourski, and A. A. Tsirlin, Collinear order in the frustrated three-dimensional spin12\text{spin}-\frac{1}{2} antiferromagnet Li2CuW2O8{\mathrm{Li}}_{2}{\mathrm{CuW}}_{2}{\mathrm{O}}_{8}Phys. Rev. B 92, 094426 (2015).
  • Ranjith et al. [2016] K. M. Ranjith, R. Nath, M. Majumder, D. Kasinathan, M. Skoulatos, L. Keller, Y. Skourski, M. Baenitz, and A. A. Tsirlin, Commensurate and incommensurate magnetic order in spin-1 chains stacked on the triangular lattice in Li2NiW2O8{\mathrm{Li}}_{2}{\mathrm{NiW}}_{2}{\mathrm{O}}_{8}Phys. Rev. B 94, 014415 (2016).
  • Kundu et al. [2020] S. Kundu, A. Shahee, A. Chakraborty, K. M. Ranjith, B. Koo, J. Sichelschmidt, M. T. F. Telling, P. K. Biswas, M. Baenitz, I. Dasgupta, S. Pujari, and A. V. Mahajan, Gapless Q\mathrm{Q}uantum S\mathrm{S}pin L\mathrm{L}iquid in the T\mathrm{T}riangular S\mathrm{S}ystem Sr3CuSb2O9\mathrm{Sr}_{3}\mathrm{CuSb}_{2}\mathrm{O}_{9}Phys. Rev. Lett. 125, 267202 (2020).
  • Katayama et al. [2015] N. Katayama, K. Kimura, Y. Han, J. Nasu, N. Drichko, Y. Nakanishi, M. Halim, Y. Ishiguro, R. Satake, E. Nishibori, M. Yoshizawa, T. Nakano, Y. Nozue, Y. Wakabayashi, S. Ishihara, M. Hagiwara, H. Sawa, and S. Nakatsuji, Absence of J\mathrm{J}ahn-T\mathrm{T}eller transition in the hexagonal Ba3CuSb2O9\mathrm{Ba}_{3}\mathrm{CuSb}_{2}\mathrm{O}_{9} single crystal, Proceedings of the National Academy of Sciences 112, 9305 (2015).
  • Zhou et al. [2011a] H. D. Zhou, E. S. Choi, G. Li, L. Balicas, C. R. Wiebe, Y. Qiu, J. R. D. Copley, and J. S. Gardner, Spin liquid state in the S=1/2\mathrm{S}=1/2 triangular lattice Ba3CuSb2O9\mathrm{Ba}_{3}\mathrm{CuSb}_{2}\mathrm{O}_{9}Phys. Rev. Lett. 106, 147204 (2011a).
  • Nakatsuji et al. [2012] S. Nakatsuji, K. Kuga, K. Kimura, R. Satake, N. Katayama, E. Nishibori, H. Sawa, R. Ishii, M. Hagiwara, F. Bridges, T. U. Ito, W. Higemoto, Y. Karaki, M. Halim, A. A. Nugroho, J. A. Rodriguez-Rivera, M. A. Green, and C. Broholm, Spin-orbital short-range order on a honeycomb-based lattice, Science 336, 559 (2012).
  • Zhou et al. [2011b] H. D. Zhou, E. S. Choi, G. Li, L. Balicas, C. R. Wiebe, Y. Qiu, J. R. D. Copley, and J. S. Gardner, Spin liquid state in the S=1/2\mathrm{S}=1/2 triangular lattice Ba3CuSb2O9\mathrm{Ba}_{3}\mathrm{CuSb}_{2}\mathrm{O}_{9}Phys. Rev. Lett. 106, 147204 (2011b).
  • Wallace and McQueen [2015] D. C. Wallace and T. M. McQueen, New honeycomb iridium(v) oxides: NaIrO3\mathrm{NaIrO}_{3} and Sr3CaIr2O9\mathrm{Sr}_{3}\mathrm{CaIr}_{2}\mathrm{O}_{9}Dalton Trans. 44, 20344 (2015).
  • Rijssenbeek et al. [2002] J. T. Rijssenbeek, S. Malo, V. Caignaert, and K. R. Poeppelmeier, Site and O\mathrm{O}xidation-S\mathrm{S}tate S\mathrm{S}pecificity Y\mathrm{Y}ielding D\mathrm{D}imensional C\mathrm{C}ontrol in P\mathrm{P}erovskite R\mathrm{R}uthenates, J. Am. Chem. Soc. 124, 2090 (2002).
  • Tang et al. [2016] Y. Tang, E. C. Hunter, P. D. Battle, R. P. Sena, J. Hadermann, M. Avdeev, and J. Cadogan, Structural chemistry and magnetic properties of the perovskite Sr3Fe2TeO9\mathrm{Sr}_{3}\mathrm{Fe}_{2}\mathrm{Te}\mathrm{O}_{9}J. Solid State Chem. 242, 86 (2016).
  • Thakur et al. [2022] G. S. Thakur, T. C. Hansen, W. Schnelle, S. Guo, O. Janson, J. van den Brink, C. Felser, and M. Jansen, Buckled H\mathrm{H}oneycomb L\mathrm{L}attice C\mathrm{C}ompound Sr3CaOs2O9\mathrm{Sr}_{3}\mathrm{CaOs}_{2}\mathrm{O}_{9} E\mathrm{E}xhibiting A\mathrm{A}ntiferromagnetism above R\mathrm{R}oom T\mathrm{T}emperature, Chem. Mater. 34, 4741 (2022).
  • [27] See the Supplemental Material for further details about sample preparation, experimental conditions, x-ray diffraction, magnetic susceptibility analysis, estimation of lattice heat capacity, and thermal conductivity data analysis .
  • Levin et al. [2000] I. Levin, L. Bendersky, J. Cline, R. Roth, and T. Vanderah, Octahedral T\mathrm{T}ilting and C\mathrm{C}ation O\mathrm{O}rdering in P\mathrm{P}erovskite-L\mathrm{L}ike Ca4Nb2O9=3Ca(Ca1/3Nb2/3)O3\mathrm{Ca}_{4}\mathrm{Nb}_{2}\mathrm{O}_{9}\mathrm{=3\textbullet Ca(Ca}_{1/3}\mathrm{Nb}_{2/3}\mathrm{)O}_{3} P\mathrm{P}olymorphs, J. Solid State Chem. 150, 43 (2000).
  • Zheng et al. [2005] W. Zheng, R. R. P. Singh, R. H. McKenzie, and R. Coldea, Temperature dependence of the magnetic susceptibility for triangular-lattice antiferromagnets with spatially anisotropic exchange constants, Phys. Rev. B 71, 134422 (2005).
  • Dey et al. [2017] T. Dey, M. Majumder, J. C. Orain, A. Senyshyn, M. Prinz-Zwick, S. Bachus, Y. Tokiwa, F. Bert, P. Khuntia, N. Büttgen, A. A. Tsirlin, and P. Gegenwart, Persistent low-temperature spin dynamics in the mixed-valence iridate Ba3InIr2O9\mathrm{Ba}_{3}\mathrm{InIr}_{2}\mathrm{O}_{9}Phys. Rev. B 96, 174411 (2017).
  • Li et al. [2015] Y. Li, H. Liao, Z. Zhang, S. Li, F. Jin, L. Ling, L. Zhang, Y. Zou, L. Pi, Z. Yang, J. Wang, Z. Wu, and Q. Zhang, Gapless quantum spin liquid ground state in the two-dimensional spin-1/2 triangular antiferromagnet YbMgGaO4\mathrm{YbMgGaO}_{4}Sci. Rep. 5, 16419 (2015).
  • Elstner et al. [1993] N. Elstner, R. R. P. Singh, and A. P. Young, Finite temperature properties of the spin-1/2 heisenberg antiferromagnet on the triangular lattice, Phys. Rev. Lett. 71, 1629 (1993).
  • Mustonen et al. [2018] O. Mustonen, S. Vasala, E. Sadrollahi, K. P. Schmidt, C. Baines, H. C. Walker, I. Terasaki, F. J. Litterst, E. Baggio-Saitovitch, and M. Karppinen, Spin-liquid-like state in a spin-1/2 square-lattice antiferromagnet perovskite induced by d10d0d^{10}-d^{0} cation mixing, Nature Communications 9, 1085 (2018).
  • Zeng et al. [2022] Z. Zeng, X. Ma, S. Wu, H.-F. Li, Z. Tao, X. Lu, X.-h. Chen, J.-X. Mi, S.-J. Song, G.-H. Cao, G. Che, K. Li, G. Li, H. Luo, Z. Y. Meng, and S. Li, Possible dirac quantum spin liquid in the kagome quantum antiferromagnet YCu3(OH)6Br2[Brx(OH)1x]\mathrm{YCu}_{3}\mathrm{(OH)}_{6}\mathrm{Br}_{2}[\mathrm{Br}_{x}\mathrm{(OH)}_{1-x}]Phys. Rev. B 105, L121109 (2022).
  • Motrunich [2005] O. I. Motrunich, Variational study of triangular lattice spin1/2-1/2 model with ring exchanges and spin liquid state in κ(ET)2Cu2CN3\kappa-\mathrm{(ET)}_{2}\mathrm{Cu}_{2}\mathrm{CN}_{3}Phys. Rev. B 72, 045105 (2005).
  • Jana et al. [2022] Y. Jana, A. Ghosal, S. Nandi, J. Alam, P. Bag, S. Islam, and R. Nath, Spin-ice behavior of mixed pyrochlore Dy2GaSbO7 exhibiting enhanced Pauling zero-point entropy, J. Magn. Magn. Mater. 562, 169814 (2022).
  • Ranjith et al. [2019] K. M. Ranjith, D. Dmytriieva, S. Khim, J. Sichelschmidt, S. Luther, D. Ehlers, H. Yasuoka, J. Wosnitza, A. A. Tsirlin, H. Kühne, and M. Baenitz, Field-induced instability of the quantum spin liquid ground state in the Jeff=12{J}_{\mathrm{eff}}=\frac{1}{2} triangular-lattice compound NaYbO2{\mathrm{NaYbO}}_{2}Phys. Rev. B 99, 180401 (2019).
  • Singh et al. [2024] V. K. Singh, K. Nam, M. Barik, K. Boya, E. Kermarrec, P. Khuntia, K. H. Kim, S. Bhowal, and B. Koteswararao, Bi2YbO4Cl\mathrm{Bi}_{2}\mathrm{YbO}_{4}\mathrm{Cl}: A two-dimensional square-lattice compound with jeff=12j_{eff}=\frac{1}{2} magnetic moments, Phys. Rev. B 109, 075128 (2024).
  • Li et al. [2016] Y. Li, D. Adroja, P. K. Biswas, P. J. Baker, Q. Zhang, J. Liu, A. A. Tsirlin, P. Gegenwart, and Q. Zhang, Muon spin relaxation evidence for the U(1)\mathrm{U(1)} quantum spin-liquid ground state in the triangular antiferromagnet YbMgGaO4\mathrm{YbMgGaO}_{4}Phys. Rev. Lett. 117, 097201 (2016).
  • Lee et al. [2021] S. Lee, C. H. Lee, A. Berlie, A. D. Hillier, D. T. Adroja, R. Zhong, R. J. Cava, Z. H. Jang, and K.-Y. Choi, Temporal and field evolution of spin excitations in the disorder-free triangular antiferromagnet Na2BaCo(PO4)2\mathrm{Na}_{2}\mathrm{BaCo}{({\mathrm{PO}}_{4})}_{2}Phys. Rev. B 103, 024413 (2021).
  • Uemura et al. [1994] Y. J. Uemura, A. Keren, K. Kojima, L. P. Le, G. M. Luke, W. D. Wu, Y. Ajiro, T. Asano, Y. Kuriyama, M. Mekata, H. Kikuchi, and K. Kakurai, Spin fluctuations in frustrated kagomé lattice system SrCu8Ga4O19\mathrm{SrCu}_{8}\mathrm{Ga}_{4}\mathrm{O}_{19} studied by muon spin relaxation, Phys. Rev. Lett. 73, 3306 (1994).
  • Müller-Warmuch [1991] W. Müller-Warmuch, C. P. Slichter, Principles of Magnetic Resonance, 3rd enlarged and updated edition (Springer, Berlin, 1989), Vol. 95 (Wiley, 1991) p. 449–450.
  • Bono et al. [2004] D. Bono, P. Mendels, G. Collin, N. Blanchard, F. Bert, A. Amato, C. Baines, and A. D. Hillier, μSR\mu\mathrm{S}\mathrm{R} study of the quantum dynamics in the frustrated S=32\mathrm{S}=\frac{3}{2} kagomé bilayers, Phys. Rev. Lett. 93, 187201 (2004).
  • Keren et al. [2004] A. Keren, J. S. Gardner, G. Ehlers, A. Fukaya, E. Segal, and Y. J. Uemura, Dynamic properties of a diluted pyrochlore cooperative paramagnet (TbpY1p)2Ti2O7(\mathrm{Tb}_{p}\mathrm{Y}_{1-p})_{2}\mathrm{Ti}_{2}\mathrm{O}_{7}Phys. Rev. Lett. 92, 107204 (2004).
  • Zhang et al. [2022] Z. Zhang, J. Li, M. Xie, W. Zhuo, D. T. Adroja, P. J. Baker, T. G. Perring, A. Zhang, F. Jin, J. Ji, X. Wang, J. Ma, and Q. Zhang, Low-energy spin dynamics of the quantum spin liquid candidate NaYbSe2\mathrm{NaYbSe}_{2}Phys. Rev. B 106, 085115 (2022).
  • Kermarrec et al. [2011] E. Kermarrec, P. Mendels, F. Bert, R. H. Colman, A. S. Wills, P. Strobel, P. Bonville, A. Hillier, and A. Amato, Spin-liquid ground state in the frustrated kagome antiferromagnet MgCu3(OH)6Cl2Phys. Rev. B 84, 100401 (2011).
  • Ishant et al. [2024] I. Ishant, T. Shiroka, O. Stockert, V. Fritsch, and M. Majumder, Frustration-induced quantum criticality in ni-doped CePdAl\mathrm{CePdAl} as revealed by the µSR\text{\textmu{}}\mathrm{SR} technique, Phys. Rev. Res. 6, 023112 (2024).
  • Sarkar et al. [2019] R. Sarkar, P. Schlender, V. Grinenko, E. Haeussler, P. J. Baker, T. Doert, and H.-H. Klauss, Quantum spin liquid ground state in the disorder free triangular lattice NaYbS2{\mathrm{NaYbS}}_{2}Phys. Rev. B 100, 241116 (2019).
  • Balz et al. [2016] C. Balz, B. Lake, J. Reuther, H. Luetkens, R. Schönemann, T. Herrmannsdörfer, Y. Singh, A. T. M. Nazmul Islam, E. M. Wheeler, J. A. Rodriguez-Rivera, T. Guidi, G. G. Simeoni, C. Baines, and H. Ryll, Physical realization of a quantum spin liquid based on a complex frustration mechanism, Nature Physics 12, 942 (2016).
  • Tripathi et al. [2022] R. Tripathi, D. T. Adroja, C. Ritter, S. Sharma, C. Yang, A. D. Hillier, M. M. Koza, F. Demmel, A. Sundaresan, S. Langridge, W. Higemoto, T. U. Ito, A. M. Strydom, G. B. G. Stenning, A. Bhattacharyya, D. Keen, H. C. Walker, R. S. Perry, F. Pratt, Q. Si, and T. Takabatake, Quantum critical spin-liquid-like behavior in the S=12\mathrm{S}=\frac{1}{2} quasikagome-lattice compound CeRh1xPdxSn\mathrm{CeRh}_{1-x}\mathrm{Pd}_{x}\mathrm{Sn} investigated using muon spin relaxation and neutron scattering, Phys. Rev. B 106, 064436 (2022).
  • Ogielski [1985] A. T. Ogielski, Dynamics of three-dimensional ising spin glasses in thermal equilibrium, Phys. Rev. B 32, 7384 (1985).
  • Campbell et al. [1994] I. A. Campbell, A. Amato, F. N. Gygax, D. Herlach, A. Schenck, R. Cywinski, and S. H. Kilcoyne, Dynamics in canonical spin glasses observed by muon spin depolarization, Phys. Rev. Lett. 72, 1291 (1994).
  • Uemura et al. [1985] Y. J. Uemura, T. Yamazaki, D. R. Harshman, M. Senba, and E. J. Ansaldo, Muon-spin relaxation in AuFe\mathrm{AuFe} and CuMn\mathrm{CuMn} spin glasses, Phys. Rev. B 31, 546 (1985).
  • Murayama et al. [2020] H. Murayama, Y. Sato, T. Taniguchi, R. Kurihara, X. Z. Xing, W. Huang, S. Kasahara, Y. Kasahara, I. Kimchi, M. Yoshida, Y. Iwasa, Y. Mizukami, T. Shibauchi, M. Konczykowski, and Y. Matsuda, Effect of quenched disorder on the quantum spin liquid state of the triangular-lattice antiferromagnet 1TTaS21\mathrm{T}-\mathrm{TaS}_{2}Phys. Rev. Res. 2, 013099 (2020).
  • Murayama et al. [2022] H. Murayama, T. Tominaga, T. Asaba, A. d. O. Silva, Y. Sato, H. Suzuki, Y. Ukai, S. Suetsugu, Y. Kasahara, R. Okuma, I. Kimchi, and Y. Matsuda, Universal scaling of specific heat in the S=12\mathrm{S}=\frac{1}{2} quantum kagome antiferromagnet herbertsmithite, Phys. Rev. B 106, 174406 (2022).
  • Kimchi et al. [2018] I. Kimchi, J. P. Sheckelton, T. M. McQueen, and P. A. Lee, Scaling and data collapse from local moments in frustrated disordered quantum spin systems, Nature Communications 9, 4367 (2018).
  • et al. [2024] M. M. et al., Microscopic investigation of the ground state of quantum spin liquid candidate Sr3CuTa2O9\mathrm{Sr}_{3}\mathrm{CuTa}_{2}\mathrm{O}_{9} (2024).