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

Field evolution of magnetic phases and spin dynamics in the honeycomb lattice magnet Na2Co2TeO6: 23Na NMR study

Jun Kikuchi [email protected]    Takayuki Kamoda    Nobuyoshi Mera    Yodai Takahashi    Kouji Okumura    Yukio Yasui Department of Physics, School of Science and Technology, Meiji University, Kawasaki 214-8571, Japan
Abstract

We report on the results of 23Na NMR in the honeycomb lattice magnet Na2Co2TeO6 which has been nominated as a Kitaev material. Measurements of magnetic shift and width of the NMR line as functions of temperature and magnetic field show that a spin-disordered phase does not appear up to a field of 9 T. In the antiferromagnetic phase just below the Néel temperature TNT_{N}, we find a temperature region extending down to TN/2\sim T_{N}/2 where the nuclear spin-lattice relaxation rate 1/T11/T_{1} remains enhanced and is further increased by magnetic fields. This region crosses over to a low temperature region characterized by the rapidly decreasing 1/T11/T_{1} which is less field-sensitive. These observations suggest incoherent spin excitations with a large spectral weight at low energies in the intermediate temperature region transforming to more conventional spin-wave excitations at low temperatures. The drastic change of the low-energy spin dynamics is likely caused by strong damping of spin waves activated only in the intermediate temperature region, which may be realized for triple-𝐪\mathbf{q} magnetic order possessing partially-disordered moments as scattering centers of spin waves. In the paramagnetic phase near TNT_{N}, dramatic field suppression of 1/T11/T_{1} is observed. From analysis of the temperature dependence of 1/T11/T_{1} based on the renormalized-classical description of a two-dimensional quantum antiferromagnet, we find the field-dependent spin stiffness constant that scales with TNT_{N} as a function of magnetic field. This implies field suppression of the energy scale characterizing both two-dimensional spin correlations and three-dimensional long-range order, which may be associated with an increasing effect of frustration in magnetic fields.

I Introduction

The search for novel quantum phases in frustrated magnets has long been the subject of intense studies since the first proposal of a resonating valence bond state in a triangular lattice antiferromagnet [1]. One of the most intriguing phases to be sought is a quantum spin liquid (QSL) which breaks no spontaneous symmetry and is characterized by topological quantities. Since QSLs are expected to appear under the influence of strong frustration and quantum fluctuations both suppressing magnetic long-range order (LRO), frustrated quantum-spin systems on low-dimensional lattices have been studied extensively from experimental and theoretical viewpoints [2, 3, 4, 5].

A honeycomb lattice has the smallest coordination number among two-dimensional (2D) lattices and hence strong quantum effects are expected. Despite such fundamental importance, honeycomb lattice magnets seem less explored compared with triangular and square lattice magnets. The honeycomb lattice is bipartite and is not frustrated for nearest-neighbor interactions like the square lattice, but the presence of further neighbor interactions introduces frustration, leading to various competing phases in both classical and quantum cases [6, 7, 8, 9, 10, 11]. In the frustrated spin-1/2 Heisenberg model, quantum fluctuations are sufficient to destroy magnetic LRO in some parameter regions and gives rise to disordered ground states such as a gapped QSL and a dimer or plaquette valence-bond solid. In the case of XYXY anisotropy, the honeycomb lattice magnets provide a playground for studying the fascinating Berezinskii-Kosterlitz-Thouless transition driven by topological excitations.

Another direction of research on the honeycomb lattice magnets has been prompted by an exactly solvable quantum spin-1/2 model formulated by Kitaev [12]. The model consists of nearest-neighbor bond-dependent Ising interactions of which easy axes are mutually orthogonal and is strongly frustrated. It has a gapped QSL ground state with fractionalized Majorana fermion excitations coupled to a Z2Z_{2} gauge field and has attracted growing interest in recent years [13, 14, 15, 16, 17].

Materials realizations of the Kitaev model have been proposed for 4d4d and 5d5d transition-metal compounds including ions with a low-spin d5(t2g5)d^{5}\,(t_{2g}^{5}) configuration [18, 19]. These ions can host spin-orbital entangled pseudospin jeff=1/2j_{\mathrm{eff}}=1/2 in an octahedral crystal field [20], which enables bond-dependent coupling via anisotropic electronic wave functions. Extensive researches on the candidate materials such as Na2IrO3 and α\alpha-RuCl3 revealed that most of them display magnetic LRO, indicating the importance of non-Kitaev interactions in real materials. Generalized Kitaev models including the Heisenberg and off-diagonal exchange terms were developed to show that magnetic LRO is largely stabilized by these additional interactions but the QSL survives in small but finite regions of the parameter space [21, 22].

In spite of the absence of a QSL phase, the relevance of Kitaev physics to magnetic properties of the candidate materials is recognized in many respects. One such example is an unconventional continuum of the magnetic excitations in α\alpha-RuCl3 and Na2IrO3 reminiscent of itinerant Majorana-fermion bands expected in the Kitaev QSLs [23, 24, 25]. The two-step release of magnetic entropy is also interpreted as a signature of fractionalization of spin degrees of freedom [26, 27]. The most prominent hallmark would be the field-induced disordered phase of α\alpha-RuCl3 found above the in-plane critical field of 7-8 T [28, 29, 30]. Because of a close connection of this phase to field-driven QSL states in the Kitaev model [12, 31] as well as the robustness of the excitation continuum against a field [24], α\alpha-RuCl3 has been considered proximate to the Kitaev QSL.

Recently, 3d3d transition-metal compounds including Co2+ ions with a high-spin d7(t2g5eg2)d^{7}\,(t_{2g}^{5}e_{g}^{2}) configuration have been suggested to be more suitable for the Kitaev magnet [32, 33, 34, 35]. The compound Na2Co2TeO6 has received special attention among the Co candidates because it shows a field-induced transition from an antiferromagnetic to a putative spin-disordered phase similar to α\alpha-RuCl3. The crystal structure of Na2Co2TeO6 belongs to the hexagonal space group P6322P6_{3}22 (No. 182) [36, 37, 38, 39, 40]. The CoO6 octahedra comprised of two independent Co1 and Co2 sites share their edges to form nearly ideal honeycomb lattices in the abab planes. Stacking of the honeycomb layers in the cc direction is such that the six-fold screw axis going through the Co1 site transforms a honeycomb lattice to the next layer. The Co2 site is hence on top of the Te site which is at the center of the adjacent Co hexagon. Na sites are in between the honeycomb layers and are partially occupied. Structural analysis indicated that a stacking fault and intermixing of cations in the honeycomb layers are not evident [38, 36].

The Co2+ ions host pseudospin 1/21/2 as revealed from the observation of spin-orbit excitations at 22-23 meV via inelastic neutron scattering (INS) [41, 42]. Antiferromagnetic (AFM) LRO characterized by a propagation vector 𝐐=(12,0,0)\mathbf{Q}=(\frac{1}{2},0,0) is observed below the Néel temperature TN27T_{N}\approx 27 K at zero field [37, 38, 40]. The proposed magnetic structures are shown schematically in Fig. 1. The single-𝐪\mathbf{q} collinear structure [Fig. 1(a)] has a zigzag spin arrangement similar to Na2IrO3 [43] and α\alpha-RuCl3 [28]; ferromagnetic zigzag chains running along the bb axis (perpendicular to 𝐐\mathbf{Q}) align alternately in the direction of 𝐐\mathbf{Q} (parallel to aa^{*}) with the ordered moment either pointing along the bb axis [37, 38] or lying in the bcbc plane [40]. The triple-𝐪\mathbf{q} noncollinear structure [Fig. 1(b)] is formed by superposing the zigzag structure with 𝐐=(12,0,0)\mathbf{Q}=(\frac{1}{2},0,0) and the equivalents related by C3C_{3} rotation [44]. The ordered moment on the Co1 site is larger by a factor of 1.1-1.2 than that on the Co2 site, reflecting the fact that the ordered phase is in a strict sense ferrimagnetic [45].

Refer to caption
Figure 1: Magnetic structures proposed for Na2Co2TeO6. The arrows indicate in-plane ordered moments. (a) Zigzag magnetic structure described by a propagation vector 𝐐=(12,0,0)\mathbf{Q}=(\frac{1}{2},0,0). The dashed line represents the magnetic unit cell with a two-fold screw axis normal to the abab plane at the origin. (b) An example of the triple-𝐪\mathbf{q} magnetic structure. The dashed line represents the magnetic unit cell with a six-fold screw axis 636_{3} at the origin. Spinfull (“spinless”) Co atoms are shown by solid (open) circles. Note that only three quarters of Co atoms have an in-plane moment. Each Co atom including spinless one may have a cc-axis Néel component allowed by C3C_{3} symmetry.

The magnetic susceptibility shows easy-plane anisotropy nominally described by a direction dependent Weiss temperature [39, 45] suggesting anisotropic exchange interactions. Standard analysis in evaluating exchange parameters from the Curie-Weiss fit of the susceptibility was unsuccessful owing to the presence of competing interactions and an effect of low-lying spin-orbit excited states [34, 35]. Powder INS experiments have been attempted to extract a set of parameters which consistently describes the magnetic excitation spectrum based on the linear spin-wave theory [41, 46, 40, 42, 47]. The generalized Kitaev model rather than the frustrated Heisenberg model is preferred, but the results are still diverse; even the sign of the leading interaction is unsettled, although theories predict a dominant ferromagnetic Kitaev interaction [32, 33]. A recent INS study on single crystals has revealed unusual features of the magnetic excitations: the existence of an incoherent continuum around 𝐐=(12,0,0)\mathbf{Q}=(\frac{1}{2},0,0) persisting down to TN/2\sim T_{N}/2 and the formation of a spin-wave mode below that temperature [44].

One of the most attracting features of Na2Co2TeO6 is an anisotropic field response of the AFM phase resembling α\alpha-RuCl3. The AFM order of Na2Co2TeO6 is suppressed by a moderate in-plane field, above which a QSL state is expected to emerge [45, 48, 46]. Yao and Li are the first who reported strong suppression of AFM order by an in-plane field and related anomalies in magnetization as well as in magnetic specific heat and suggested the existence of a high-field spin-disordered phase [45]. They also found that the magnetization jump detected in powders around 6 T [37, 38] is observable only for the field 𝐁a\mathbf{B}\parallel a^{*}, i.e., perpendicular to the zigzag-ordered moment, which indicates that the transition is not of spin-flop type. Subsequent magnetization and thermal transport measurements at higher fields demonstrated that the AFM phase closes at the critical field Bc10B_{c}\approx 10 T, above which a phase with gapped spin excitations may appear [48]. On the other hand, Lin et al. claimed from the magnetization and magnetic specific heat data that there is a QSL-like disordered phase in an intermediate field range 7.5T<B<10.5T7.5\mathrm{~{}T}<B<10.5\mathrm{~{}T} before entering a high-field polarized phase [46]. There is also a seemingly important anomaly in the AFM phase not identified as yet, a board hump in magnetic specific heat around 10 K present already at zero field and merging into the anomaly at TNT_{N} with increasing in-plane field [45, 46]. The hump is robust against an out-of-plane field and is pronounced at higher fields, which implies a transition to the phase with a different magnetic structure.

Although a lot of effort has been devoted to elucidate magnetic characteristics of Na2Co2TeO6 and their relevance to Kitaev physics, fundamental aspects regarding the low-energy spin dynamics remain unresolved. A microscopic investigation of the phases and their evolution in magnetic fields has also been lacking. In this paper, we report results of comprehensive 23Na NMR measurements on polycrystalline samples of Na2Co2TeO6, paying special attention to how magnetic phases and low-energy spin dynamics evolve with temperature and magnetic field. We construct the magnetic phase diagram using the microscopic quantities measured by NMR and demonstrate the absence of a spin-disordered phase up to 9 T. The nuclear spin-lattice relaxation rate 1/T11/T_{1} is found to exhibit strong temperature and field variations especially at temperatures T2TNT\lesssim 2T_{N}. A contrasting field response of 1/T11/T_{1} is observed above and below TNT_{N}. The measurements of 1/T11/T_{1} in the AFM phase reveal an unconventionally large spectral weight of low-energy spin excitations which persists down to TN/2\sim T_{N}/2 and dies out below that temperature. As a possible origin for the qualitative change of the low-energy spin dynamics, we discuss spin-wave damping due to partially-disordered moments residing in the triple-𝐪\mathbf{q} magnetic structure. In the paramagnetic (PM) phase, field suppression of 2D spin correlations is inferred from the field dependence of 1/T11/T_{1}. A dominant exchange energy is evaluated from quantitative analysis of 1/T11/T_{1} in the high-temperature limit.

During the course of this study, several groups have reported 23Na NMR in a single crystal of Na2Co2TeO6 [44, 49]. Although some interesting features such as successive anomalies in the nuclear spin-spin relaxation rate 1/T21/T_{2} and a possible signature of slow dynamics have been observed in the AFM phase, the measurements performed at a relatively low field are not intended to trace field variations of the measurable quantities. The present study will give additional information necessary for deeper understanding of field-dependent phenomena in this compound.

II Experiments

Polycrystalline samples of Na2Co2TeO6 were synthesized by a solid state reaction. A stoichiometric mixture of Na2CO3, Co3O4 and TeO2 was pressed into pellets after fine grinding and was sintered in a preheated furnace at 860 C\mathrm{{}^{\circ}C} for 12 h in air. The samples were then cooled slowly in the furnace to room temperature over 50 h. The final product was characterized by x-ray diffraction at room temperature and was confirmed to be a single phase with no trace of impurity. Temperature-dependent magnetization (MM) measurements were carried out in a magnetic field (BB) range of 1-9 T using a vibrating sample magnetometer (Quantum Design, Dynacool) under the zero field cooling (ZFC) and field cooling (FC) conditions.

NMR measurements are performed on the 23Na nucleus (the spin I=3/2I=3/2 and the gyromagnetic ratio γ23/2π=11.2623{}^{23}\gamma/2\pi=11.2623 MHz/T) with a standard phase-coherent pulsed spectrometer. The π/2\pi/2-τ\tau-π\pi two-pulse sequence was used to excite the spin-echo signals. The NMR spectra were taken by recording the spin-echo signal while sweeping the external magnetic field at a fixed frequency. 23Na nuclear spin-lattice relaxation rate 1/T11/T_{1} was measured by an inversion recovery method at the peak position of the NMR spectrum. 1/T11/T_{1} was determined by fitting the recovery of the spin-echo intensity M(t)M(t) as a function of the time tt after an inversion pulse to the following stretched multi-exponential function with the exponent β(1)\beta\;(\leq 1) incorporating the distribution of T1T_{1} and the overlap of central (m=1/21/2m=1/2\leftrightarrow-1/2) and satellite (±3/2±1/2\pm 3/2\leftrightarrow\pm 1/2) transition lines in powder samples [50, 51, 52, 53, 54]:

M(t)=MM0k=13αke(λkt/T1)β.\displaystyle M(t)=M_{\infty}-M_{0}\sum_{k=1}^{3}\alpha_{k}e^{-(\lambda_{k}t/T_{1})^{\beta}}. (1)

Here MM_{\infty} is the intensity at the thermal equilibrium; M0M_{0} is a degree of inversion; {λk}={1,3,6}\{\lambda_{k}\}=\{1,3,6\} are mode eigenvalues, {αk}\{\alpha_{k}\} are amplitudes of the corresponding modes satisfying kαk=1\sum_{k}\alpha_{k}=1. Selective inversion of the populations of m=±1/2m=\pm 1/2 states (central transition) gives {αk}={1/10,0,9/10}\{\alpha_{k}\}=\{1/10,0,9/10\} frequently used in the literature [55, 56, 57, 58, 59, 44, 49]. Deviation of β\beta from unity measures the distribution of 1/T11/T_{1}.

III Results

III.1 Magnetic susceptibility

Figure 2 shows the temperature and field dependences of the magnetic susceptibility M/BM/B under the ZFC condition. No essential difference was observed between the ZFC and FC conditions except a minor difference at 1 T below TNT_{N} (not shown) which may be attributed to compensation behavior of ferrimagnetism [45]. The results are in good agreement with those reported on polycrystalline samples [36, 37, 38] but are slightly different in the absolute magnitude from that of a single crystal because of powder averaging of the anisotropic susceptibility [45, 46]. The Néel temperature and its field variation also agree with those in the previous reports. The susceptibility is essentially field independent in the PM phase, contrasting to the behavior in α\alpha-RuCl3 where it is enhanced by field at temperatures T3TNT\lesssim 3T_{N} [60].

Refer to caption
Figure 2: Temperature dependences of the magnetic susceptibility.

III.2 NMR spectrum

Figure 3(a) shows the temperature (TT) variation of the 23Na NMR spectrum taken at a frequency ν0=33.790\nu_{0}=33.790 MHz. The signal intensity is plotted against the field shift B0BB_{0}-B relative to the reference field B0=ν0/γB_{0}=\nu_{0}/\gamma where BB is the external field. The spectra at high TT in the PM phase exhibit a quadrupolar-split powder pattern characterized by asymmetric electric field gradients (EFGs) at the Na site [61]. This is in clear contrast to the previous reports on single crystals, none of which observed quadrupolar satellites [44, 49]. The TT-dependent magnetic broadening is also obvious (see below). From singular positions of the satellites, the principal values of the EFG tensor are determined as νQ23=1.68{}^{23}\nu_{\mathrm{Q}}=1.68 MHz and η=0.49\eta=0.49 which are almost TT independent. We observed neither line splitting due to the presence of several crystallographic Na sites nor severe broadening or smearing of quardupolar satellites due to disorder [36, 38, 37]. This suggests preferential occupation of some Na site and/or a tendency of atomic ordering rather than strong disorder in Na layers.

Refer to caption
Figure 3: (a) Temperature variation of the 23Na NMR spectrum taken at the reference field B0=ν0/γ3B_{0}=\nu_{0}/\gamma\approx 3 T. The spectra have been normalized to the maximum intensity. The spectra in the paramagnetic (antiferromagnetic) phase are colored yellow (red). (b) Field variation of the 23Na NMR spectrum at 4.2 K. The spectra with a positive peak shift (B07B_{0}\geq 7 T) are colored blue. The spectrum for B0=8.9B_{0}=8.9 T is terminated because of an upper limit of our apparatus. The dashed lines represent a reference field position B=B0B=B_{0}.

We measured the NMR spectrum at various frequencies and temperatures and determined the temperature and field dependences of the hyperfine field at the Na sites and the line width of the spectrum. In the following, we label the spectra and the related quantities by the reference field B0B_{0}. Figure 4(a) shows the TT dependences of the line width (fwhm) ΔB\Delta B taken at various B0B_{0}’s. ΔB\Delta B is also plotted in Fig. 4(b) as a function of uniform magnetization MM per Co atom with TT the implicit parameter. It was found that ΔB\Delta B in the PM phase above 50 K is proportional to MM in a wide range of B0B_{0}. The proportionality constant ΔB/M=108mT/μB\Delta B/M=108\mathrm{~{}mT}/\mu_{B} is comparable to the root mean square of the principal values of the dipolar field tensor of 81mT/μB81\mathrm{~{}mT}/\mu_{B} calculated for the most occupied (\sim70%) 12i12i site for Na. The line width in the PM phase is thus dominated by anisotropic dipolar coupling between Na and Co.

Refer to caption
Figure 4: (a) Temperature dependences of the fwhm ΔB\Delta B of the NMR spectrum. (b) ΔB\Delta B plotted against the uniform magnetization per Co atom. The solid line is a fit of the data above 50 K. The linear slope gives a coupling constant of 108 mT/μB\mathrm{mT}/\mu_{B}. (c) Temperature dependences of the hyperfine field Bhf,pkB_{\mathrm{hf,pk}} at the peak position of the NMR spectrum. (d) Bhf,pkB_{\mathrm{hf,pk}} plotted against the uniform magnetization per Co atom. The solid line is a fit of the data above TNT_{N}. The linear slope gives the hyperfine coupling constant ahf=23.7mT/μBa_{\mathrm{hf}}=23.7\mathrm{~{}mT}/\mu_{B}.

As shown in Figs. 3(a) and 4(a), sudden increase of the line width was observed on entering the AFM phase at relatively low fields B05B_{0}\leq 5 T. The low-field line widths exhibit order-parameter-like TT dependence, saturating at low TT with a slightly field-dependent value. The spectrum is broadened almost symmetrically about the reference field B=B0B=B_{0}, and ΔB\Delta B no longer scales with MM [Fig. 4(b)]. These observations demonstrate the appearance of a staggered hyperfine field at the Na site of which magnitude is much smaller than the external field. The anomaly in ΔB\Delta B at TNT_{N} is obscured at high fields B07B_{0}\geq 7 T owing to magnetic broadening present already in the PM phase. The high-field line widths saturate in the low-TT limit to a value somewhat larger than the low-field value, possibly related to a reorientation of ordered moments with the external field.

It should be emphasized that at low TT the scaling between ΔB\Delta B and MM breaks down even at the highest field B0=8.9B_{0}=8.9 T. This is also true for the hyperfine field Bhf,pkB_{\mathrm{hf,pk}} at the peak position of the spectrum as will be shown below [Fig. 4(d)]. If there appeared a field-induced spin-disordered phase as suggested in Ref. 46, both ΔB\Delta B and Bhf,pkB_{\mathrm{hf,pk}} should scale linearly with MM down to the lowest temperature because of the disappearance of a staggered component of the hyperfine field. The upward deviation of ΔB\Delta B from the ΔB\Delta B-MM scaling means that the spectrum is much broader than is expected from the uniform magnetization, demonstrating a significant contribution of the staggered hyperfine field to ΔB\Delta B. Our results thus point to the absence of a field-induced disordered phase up to B9B\sim 9 T consistent with a report that the in-plane critical field is around 10 T [48].

Figure 3(b) shows the field evolution of the NMR spectrum at 4.2 K. The line shape as well as the line width does not change much up to B0=8.9B_{0}=8.9 T, which confirms the persistence of magnetic LRO. A shift of the peak due to slight asymmetry in the line shape is discernible at B02B_{0}\geq 2 T. The peak shifts to a positive side of B0BB_{0}-B above B0=7B_{0}=7 T, which may be associated with a magnetization jump around 6 T for the field applied perpendicular to the zigzag chains [39, 45].

The magnetic shift KK is an important measure of the local static spin susceptibility of a magnetic ion. For nuclei with I1I\geq 1, care must be taken to correct a contribution of the quadrupole interaction to the shift of the NMR line [61]. We measured the field dependence of the relative line shift (B0B)/B{(B_{0}-B)/B} and confirmed that the second-order quadrupolar shift is negligible compared with the magnetic shift for B03B_{0}\geq 3 T. The field shift B0BB_{0}-B near the peak of the spectrum can then be taken as the hyperfine field BhfB_{\mathrm{hf}} at the Na sites from which the magnetic shift is defined as K=Bhf/BK=B_{\mathrm{hf}}/B.

Shown in Fig. 4(c) is the TT dependence of Bhf,pk=B0BpkB_{\mathrm{hf,pk}}=B_{0}-B_{\mathrm{pk}} for various B0B_{0}’s where BpkB_{\mathrm{pk}} is the external field at the peak position of the spectrum. Figure 4(d) shows a plot of Bhf,pkB_{\mathrm{hf,pk}} against MM alternative to the conventional KK-χ\chi plot 111This plot is useful in determining the hyperfine coupling constant when magnetization shows non-linear increase at high fields so that the susceptibility χ\chi defined by M/BM/B depends on a magnetic field.. The linear relation Bhf,pk=ahfMB_{\mathrm{hf,pk}}=a_{\mathrm{hf}}M holds in the PM phase in a wide range of B0B_{0}. The slope ahf=23.7mT/μBa_{\mathrm{hf}}=23.7\mathrm{~{}mT}/\mu_{B} is by definition the hyperfine coupling constant between Na and Co. The small and positive value of ahfa_{\mathrm{hf}} suggests that the Na-3s3s orbital is polarized a little due to spin transfer from the neighboring Co atoms. Breakdown of the scaling between Bhf,pkB_{\mathrm{hf,pk}} and MM in the AFM phase is signaled by deflection of the plot from the straight line. The deviation of the low-TT data at the highest field B0=8.9B_{0}=8.9 T from the scaling confirms the absence of a spin-disordered phase up to B9B\sim 9 T.

Another interesting behavior in the AFM phase is a field-dependent shift of Bhf,pkB_{\mathrm{hf,pk}} accompanying the sign change [Fig. 4(c)]. As noted above, this is due to asymmetry in the line shape that reflects the distribution of staggered hyperfine fields. The negative shift of the peak at low fields may be associated with domains of net magnetic moment directed opposite to the external field which exist in ferrimagnets showing compensation behavior. The positive shift at high fields is probably caused by a decrease in the number of such domains as well as an increasing contribution of uniform magnetization to the hyperfine field.

Integrated intensity of the NMR spectrum is one of the essential quantities to be measured in frustrated magnets because it is highly sensitive to slow dynamics of which existence is manifested by the loss of signal intensity, or wipeout [63, 64, 65]. Recently, the slow dynamics of Co spins in the AFM phase has been suggested from partial wipeout of the 23Na NMR signal below TNT_{N} [44]. To avoid an artifact of finite separation τ\tau between rf exciting pulses in determining the integrated intensity, we measured spin-echo decay at the peak position of the spectrum and corrected the intensity by extrapolating it to τ=0\tau=0. The results are shown in Fig. 5 where we plotted the corrected intensity multiplied by TT, a factor arising from nuclear paramagnetism, as a function of TT. Unlike the previous report, we observed no anomaly around and below TNT_{N}; the integrated intensity is essentially TT independent down to 4.2 K. This clearly shows the absence of slow dynamics in the measured TT range 222We detected a dip in the uncorrected intensity (typically taken at τ=40μs\tau=40\mathrm{~{}\mu s}) around TNT_{N} as reported in Ref. 44 which is recovered by the extrapolation process. The apparent decrease in intensity around TNT_{N} reported in Ref. 44 could be caused by an insufficient correction of shortening of spin-spin relaxation time T2T_{2} accompanied by critical slowing down. The unrecovered signal loss below TNT_{N} might also be an artifact of insufficient integration range resulting from the large spectral broadening due to magnetic LRO.

Refer to caption
Figure 5: Temperature dependences of the integrated intensity of the NMR spectrum multiplied by temperature. The intensities have been normalized at 30 K. The arrows indicate TNT_{N} at the reference field of the corresponding color.

III.3 Nuclear spin-lattice relaxation rate

Figure 6(d) shows examples of the recovery of 23Na spin-echo intensity M(t)M(t). The data are well fitted by Eq. (1), which demonstrates reliability of the analyses. The TT and BB dependences of 1/T11/T_{1} at the Na sites are shown in Fig. 6(a), and the low-TT close-up in the form of 1/T1T1/T_{1}T in Fig. 6(b). The 23Na 1/T11/T_{1} in Na2Ni2TeO6 with a similar Néel temperature TN26T_{N}\approx 26 K but with a slightly different stacking sequence of honeycomb layers are shown for comparison [67]. The 23Na 1/T11/T_{1} in Na2Co2TeO6 is characterized by a strong variation not only with temperature but with magnetic field. This makes a striking contrast to the macroscopic susceptibility which depends hardly on magnetic field in the PM phase up to 9 T [Fig. 2]. The four characteristic temperature regions are identified: Region I, above 50-60 K (2TN\sim 2T_{N}) where 1/T11/T_{1} is nearly BB independent and the TT dependence is relatively weak; Region II, TN<T2TNT_{N}<T\lesssim 2T_{N} where 1/T11/T_{1} increases toward TNT_{N} and the BB dependence becomes noticeable; Region III, TN/2T<TNT_{N}/2\lesssim T<T_{N} where 1/T11/T_{1} decreases with decreasing TT but remains enhanced, still exhibiting strong field evolution; and Region IV, TTN/2T\lesssim T_{N}/2 where a rapid decrease of 1/T11/T_{1} is observed regardless of the field. On the BB dependence, 1/T11/T_{1} shows a contrasting response to magnetic fields above and below TNT_{N}; 1/T11/T_{1} is suppressed by field above TNT_{N} whereas it is enhanced below TNT_{N}.

Refer to caption
Figure 6: (a) Temperature dependences of 1/T11/T_{1}. The data of Na2Ni2TeO6 were taken from Ref. 67. Four temperature regions are labeled with numbers from I to IV. The dashed line shows a power law 1/T1T51/T_{1}\propto T^{5}. (b) Temperature dependences of 1/T1T1/T_{1}T at low temperatures. (c) Temperature dependences of the stretching exponent β\beta of the magnetization recovery. (d) Examples of the magnetization recovery taken at B=3B=3 T. The solid lines are fits to the stretched multi-exponential function Eq. (1).

Let us inspect first the TT dependence of 1/T11/T_{1} at the lowest field. 1/T11/T_{1} measured at B=1B=1 T depends only weakly on TT above about 60 K, approaching an almost TT- and BB-independent value of 1/T1100s11/T_{1\infty}\sim 100\mathrm{~{}s^{-1}} as usually observed in magnets with exchange-coupled local moments. On decreasing TT across 60 K, 1/T11/T_{1} starts to increase due to the development of short-range spin correlations consistent with the neutron diffuse scattering [38]. This is followed by a divergent increase of 1/T11/T_{1} on approaching TNT_{N} below about 30 K, an indication of three-dimensional (3D) critical slowing down. 1/T11/T_{1} takes a maximum at TN26.5T_{N}\approx 26.5 K and then decreases rapidly on cooling. However, the decrease of 1/T11/T_{1} is rather gradual compared with that in a conventional antiferromagnet in which 1/T11/T_{1} often decreases many orders of magnitude not far below TNT_{N} as observed in Na2Ni2TeO6. 1/T11/T_{1} remains in the same order as in the PM phase, which indicates residual low-energy spin excitations in Region III. This is closely related to the result of single-crystal INS that the low-energy spectral weight survives down to \sim14 K without forming a spin-wave mode and a gap in the excitation spectrum [44]. The absence of an upturn of thermal conductivity on entering the AFM phase [48] may have a close connection to the residual low-energy spin excitations. On further cooling below 13-15 K to Region IV, 1/T11/T_{1} decreases rapidly over two decades. The TT dependence of 1/T11/T_{1} is approximated by a power law 1/T1Tn1/T_{1}\propto T^{n} with n5n\approx 5 predicted for the three-magnon process at TΔT\gg\Delta rather than an activation law expected at TΔT\ll\Delta where Δ\Delta is the gap in the spin-wave spectrum [68].

Our result at B=1B=1 T is in good agreement with the previous report of 1/T11/T_{1} for a single crystal by Chen et al. taken at B=0.75B=0.75 T with 𝐁a\mathbf{B}\parallel a^{*} from 15 to 60 K [44]. Minor differences in the absolute magnitude of 1/T11/T_{1} may be attributed to our use of the stretched exponential function Eq. (1) as well as of polycrystals 333The stretched exponential function used in Ref. 44 is slightly different from conventional ones: the exponential terms are like exp[6(t/T1)β]\exp[-6(t/T_{1})^{\beta}] rather than exp[(6t/T1)β]\exp[-(6t/T_{1})^{\beta}]. If this form of stretched exponentials is applied to our recovery, the resulting 1/T11/T_{1} agrees quantitatively with their 1/T11/T_{1}.. A more extensive single-crystal study has been reported by Lee et al. who measured both 1/T11/T_{1} and 1/T21/T_{2} at B3.1B\sim 3.1 T with 𝐁c\mathbf{B}\parallel c and c\perp c up to room temperature [49]. The TT dependence of their 1/T11/T_{1} above \sim10 K is qualitatively similar to ours at 3 T, but the absolute values are larger by a factor of \sim2. The discrepancy becomes progressively greater below \sim10 K in Region IV, leading to a moderate TT variation of their 1/T11/T_{1} described by a power law with a smaller exponent n3n\sim 3.

In the AFM phase where the powder NMR spectrum is largely broadened due to the staggered hyperfine field, we measured 1/T11/T_{1} at several positions of the spectrum other than the peak position and found that the 1/T11/T_{1} differs by at most 10% in parallel with the result of 1/T11/T_{1} for the split peaks in Ref. 49. We also performed the inverse Laplace transform analysis of magnetization recovery [70, 71] in order to see whether the stretched exponential analysis captures the true distribution of 1/T11/T_{1}. The distribution function of 1/T11/T_{1} similar to that expected for the stretched exponential recovery [72, 73] was obtained as detailed in the Appendix. This justifies our phenomenological analysis of 1/T11/T_{1} using Eq. (1) and rules out a possibility of the inequivalent Na sites showing a distinct TT variation of 1/T11/T_{1} from the one displayed in Fig. 6.

The origin of the large discrepancy between our 1/T11/T_{1} and that reported in Ref. 49 especially in the behavior in Region IV is unclear. This is partly because the authors of Ref. 49 did not give fundamental information on the recovery such as the TT dependence of β\beta to be compared with ours and the recovery itself from which 1/T11/T_{1} is extracted. A possible origin of the discrepancy is a different degree of atomic disorder in Na layers between the samples; their single-crystal sample may have stronger disorder than our polycrystals because quadrupolar satellites are missing in their NMR spectrum 444Vanishingly small quadrupolar splitting or negligible quadrupole coupling seems improbable for their missing of quadrupolar satellites because they can fit the magnetization recovery using the form for a quadrupolar-split center line.. Disorder in Na layers might change low-energy spin excitations by modulating interlayer coupling and/or by introducing bond randomness in Co layers, affecting 1/T11/T_{1} effectively at low TT where the intrinsic 1/T11/T_{1} falls off due to magnetic LRO. We should also point out that in the PM phase the Redfield contribution (T1T_{1} process) to 1/T21/T_{2} for 𝐁c\mathbf{B}\parallel c evaluated from their 1/T11/T_{1} for 𝐁c\mathbf{B}\parallel c and c\perp c exceeds the observed 1/T21/T_{2} for 𝐁c\mathbf{B}\parallel c, which is impossible for the quadrupolar-split center line [75, 76]. To resolve the conflicts, single-crystal NMR measurements are under way and will be reported in a future publication.

Let us go back to our results at higher fields. Increasing magnetic field suppresses 1/T11/T_{1} above TNT_{N}, most drastically around TNT_{N} where the peak gets broadened and is shifted to lower TT. This suggests field-induced suppression of short-range spin correlations. In contrast, 1/T11/T_{1} is enhanced by field below TNT_{N} to exhibit a shoulder-like anomaly around 10 K. The anomaly appears as a broad hump in a plot of 1/T1T1/T_{1}T and is most pronounced at 6-7 T [Fig. 6(b)], which indicates field enhancement of the spectral weight remaining at low energies. The field enhancement of 1/T11/T_{1} cannot be ascribed to defect spin fluctuations because in that case 1/T11/T_{1} would be suppressed by magnetic fields [53, 77]. On the other hand, the magnetic field does not affect much the TT dependence of 1/T11/T_{1} in Region IV; 1/T11/T_{1} shows an approximate T5T^{5} law even at 9 T. There is no indication of strong field enhancement of the low-energy excitations. This corroborates the absence of a spin-disordered phase up to 9 T.

The rounding of the peak of 1/T11/T_{1} around TNT_{N} and the appearance of a broad hump in 1/T1T1/T_{1}T around 10 K in magnetic fields resemble the behavior of magnetic specific heat Cm/TC_{m}/T [45, 46]. This suggests that 1/T11/T_{1} probes excitations governing the magnetic specific heat. The resemblance between 1/T1T1/T_{1}T and Cm/TC_{m}/T also suggests that the rounded peak of 1/T11/T_{1} in magnetic fields is not an artifact due to the use of powder samples but an intrinsic property of this compound 555Since the magnetic response of Na2Co2TeO6 is insensitive to the field applied perpendicular to the honeycomb planes, it seems reasonable to consider that the field dependence measured in powders is governed by in-plane field responses. This is further supported by the fact that in powders a grain with the field lying in the honeycomb plane is found more frequently than a grain with the field normal to the plane, because the probability of a field making an angle θ\theta with respect to the direction normal to the plane is proportional to sinθ\sin\theta..

The stretching exponent β\beta in Eq. (1) also includes valuable information on the spin dynamics. As shown in Fig. 6(c), β\beta is nearly TT- and BB-independent from 30 to 200 K and takes a value close to unity, which indicates a nearly uniform relaxation process. A rapid decrease of β\beta above 200 K implies the appearance of additional relaxation channels possibly related to the spin-orbit excited state lying 22-23 meV above the ground state [41, 42]. This provides an indirect support for the spin-orbital entangled state of Co2+ in Na2Co2TeO6. β\beta decreases steeply below 30 K and takes a local minimum around TNT_{N}. A large distribution of 1/T11/T_{1} close to TNT_{N} may result from strong temperature and field-orientation dependence of 1/T11/T_{1}. While modest inhomogeneity of 1/T11/T_{1} (β0.9\beta\gtrsim 0.9) is found above \sim10 K, inhomogeneous relaxation prevails very rapidly below \sim10 K. This suggests a qualitative change of the spin dynamics on entering Region IV.

III.4 Phase diagram

The magnetic phase diagram is constructed from the TT- and BB-dependences of various quantities measured by NMR. The result is summarized in Fig. 7 together with a contour plot of 1/T1T1/T_{1}T for comparison. TNT_{N} may be determined in several ways: the temperature below which the Bhf,pkB_{\mathrm{hf,pk}}-MM scaling breaks down [Fig. 4(d)], the temperature at which 1/T1T1/T_{1}T takes a maximum, and the temperature at which β\beta takes a local minimum. All of these agree within experimental accuracies. Determination of the boundary between Regions III and IV seems more difficult because of a gradual nature of the transition. We tentatively adopt the temperature at which the derivative of 1/T1T1/T_{1}T takes a local maximum. The field above which Bhf,pkB_{\mathrm{hf,pk}} shifts to a positive side [Fig. 3(b)] constitutes another boundary dividing low- and high-field AFM phases, although the small shift of Bhf,pkB_{\mathrm{hf,pk}} relative to the broad powder spectrum leads to large errors in the boundary field. The phase diagram is in good agreement with that determined from the macroscopic quantities [45, 48, 46].

Our main finding is the existence of two distinct temperature regions in the AFM phase where Co spins exhibit contrasting low-energy dynamics. In Region III, the spin excitation spectrum has a significant low-energy weight that is enhanced strongly with magnetic field. In contrast, Region IV is likely described by spin-wave excitations. As mentioned above the transition between the two regions is gradual and is possibly a crossover rather than a phase transition with critical dynamics.

Refer to caption
Figure 7: Magnetic phase diagram constructed from the quantities measured by 23Na NMR. A contour plot of 1/T1T1/T_{1}T is also shown for comparison. The primed letters represent the high-field ordered phase with a possible reversal of the in-plane canted moment [45]. The lines between the phases and the regions are guides to the eyes. Right axis: field dependence of the spin stiffness constants ρs\rho_{s} and ρs\rho_{s}^{\,\prime} determined from the renormalized-classical analyses of 1/T11/T_{1} (see text). ρs\rho_{s}^{\,\prime} is scaled up by multiplying a factor of 1.17 to demonstrate that the field dependence of ρs\rho_{s} and ρs\rho_{s}^{\,\prime} is identical. The scale of the right axis is adjusted for ρs\rho_{s} at 1 T to coincide with the point for 1/T1T1/T_{1}T at 1 T plotted with respect to the left axis.

A transition to the high-field ordered phase (Regions III and IV) was detected around 6 T by the static quantity Bhf,pkB_{\mathrm{hf,pk}}, reflecting the magnetization jump associated with a reversal of canting moments [45]. The low-energy spin dynamics is essentially unchanged by this transition, although the field response of 1/T11/T_{1} is somewhat weakened. A spin-disordered phase does not appear up to 9 T.

IV Analysis of spin-lattice relaxation rate

IV.1 High-temperature limit

The exchange interactions between Co spins have been evaluated by several groups via powder INS techniques, but the results are not settled yet [41, 46, 42, 40]. Here we present an independent evaluation of the dominant interaction strength from the high-TT limiting value of 23Na 1/T11/T_{1} which is helpful in justifying a proper energy scale of this compound.

In the PM phase at temperatures much higher than the exchange interactions, the spin dynamics is modeled by Gaussian random modulation of individual spins under the influence of exchange-coupled neighbors. The characteristic (exchange) frequency ωe\omega_{e} is determined solely by the interaction strengths, leading to TT-independent 1/T11/T_{1} [79, 80]. In the presence of both dipolar and isotropic transferred hyperfine coupling, 1/T11/T_{1} in the high-TT limit is given as a sum of two contributions;

1T1\displaystyle\frac{1}{T_{1\infty}} =1T1,dip+1T1,tr.\displaystyle=\frac{1}{T_{1\infty,\mathrm{dip}}}+\frac{1}{T_{1\infty,\mathrm{tr}}}. (2)

The dipolar contribution is expressed as

1T1,dip\displaystyle\frac{1}{T_{1\infty,\mathrm{dip}}} =2π(γgμB)2lrl62S(S+1)3ωe,\displaystyle=\sqrt{2\pi}(\gamma g\mu_{B})^{2}\sum_{l}r_{l}^{-6}\frac{2S(S+1)}{3\omega_{e}}, (3)

where rlr_{l} is a distance between the nucleus and the ll-th electron spin, gg is the gg-factor. ωe\omega_{e} is given in terms of the exchange energy JijJ_{ij} between ii-th and jj-th electron spins and the number of jj-th spins zjz_{j} coupled to the ii-th spin as

ωe2\displaystyle\omega_{e}^{2} =23S(S+1)jzj(Jij)2.\displaystyle=\frac{2}{3}S(S+1)\sum_{j}z_{j}\biggl{(}\frac{J_{ij}}{\hbar}\biggr{)}^{2}. (4)

For the contribution of the transferred hyperfine coupling which we assume to come from the nearest-neighbor electron spins, we have

1T1,tr\displaystyle\frac{1}{T_{1\infty,\mathrm{tr}}} =π2(γgμBahf)22S(S+1)3znωe.\displaystyle=\sqrt{\frac{\pi}{2}}(\gamma g\mu_{B}a_{\mathrm{hf}})^{2}\frac{2S(S+1)}{3z_{\mathrm{n}}\omega_{e}}. (5)

Here ahfa_{\mathrm{hf}} is the hyperfine coupling constant, znz_{\mathrm{n}} is the number of nearest-neighbor spins coupled to the nucleus. We take the structural model of Xiao et al. [39] for simplicity and put zn=4z_{\mathrm{n}}=4.

Evaluating the sum lrl6\sum_{l}r_{l}^{-6} in Eq. (3) within a sphere of radius 100Å100\mathrm{~{}\AA} and using the value of ahfa_{\mathrm{hf}} determined from the Bhf,pkB_{\mathrm{hf,pk}}-MM scaling in Fig. 4(d), we obtain the ratio of the two contributions as T1,dip/T1,tr=ahf2/2znlrl60.02T_{1\infty,\mathrm{dip}}/T_{1\infty,\mathrm{tr}}=a_{\mathrm{hf}}^{2}/2z_{\mathrm{n}}\sum_{l}r_{l}^{-6}\approx 0.02. Hence we neglect the contribution 1/T1,tr1/T_{1\infty,\mathrm{tr}} and go on to evaluate ωe\omega_{e} using Eq. (3). Adopting the isotropic value g=4.33g=4.33 for the S=jeff=1/2S=j_{\mathrm{eff}}=1/2 manifold of Co2+ for simplicity [20], we finally obtain ωe=4.1×1012s1\omega_{e}=4.1\times 10^{12}\mathrm{~{}s^{-1}} from the observed value 1/T1=100s11/T_{1\infty}=100\mathrm{~{}s^{-1}}.

Since the contributions of JijJ_{ij}’s to ωe\omega_{e} are all additive, the maximum value of some JijJ_{ij} may be estimated by neglecting all the other contributions. By putting Jij=0J_{ij}=0 other than the nearest-neighbor Heisenberg coupling JJ and by putting z1=3z_{1}=3, we get |J|=26\lvert J\rvert=26 K (2.2 meV). Note that the sign of JijJ_{ij} cannot be determined by the present analysis. In estimating the Kitaev coupling KK, putting z1=1z_{1}=1 gives |K|=44\lvert K\rvert=44 K (3.8 meV). The obtained values of |J|\lvert J\rvert and |K|\lvert K\rvert fall in the same order as those evaluated from the INS experiments [41, 42, 40, 46, 47]. According to the microscopic model of the exchange interactions in Na2Co2TeO6 [35], this value of KK sets the energy scale t2/U1.1t^{2}/U\approx 1.1 meV where tt is the hopping amplitude and UU is Coulomb repulsion defined in Refs. 34, 35.

IV.2 Region with short-range spin correlations

At temperatures TN<T2TNT_{N}<T\lesssim 2T_{N} (Region II), the neutron diffuse scattering experiment has revealed that short-range spin correlations develop within the honeycomb planes [38]. The field suppression of 23Na 1/T11/T_{1} thus suggests a reduction of the in-plane spin correlation length in magnetic fields. In order to quantify the BB dependence of 1/T11/T_{1} from such a standpoint, we analyze the TT dependence of 1/T11/T_{1} based on a description of 2D Heisenberg antiferromagnets (HAFs) in terms of the quantum non-linear σ\sigma model. Similar analysis has recently been applied to 23Na 1/T11/T_{1} in a single crystal of Na2Co2TeO6 at a moderate field [49].

We assume that the system is in the renormalized-classical (RC) regime of 2D HAFs. For the system with collinear order [81, 82, 83], the correlation length ξ\xi grows exponentially with decreasing TT in the RC regime as ξexp(2πρs/T)\xi\propto\exp(2\pi\rho_{s}/T) where ρs\rho_{s} is the spin stiffness constant 666Including the correction term of O(T/2πρs)O(T/2\pi\rho_{s}) for ξ\xi introduces a minor change of parameters (\sim10% reduction of ρs\rho_{s} for example) but does not change the results qualitatively. The spin-lattice relaxation rate is given as 1/T1T3/2ξ1/T_{1}\propto T^{3/2}\xi, so that

1T1T3/2exp(2πρs/T).\displaystyle\frac{1}{T_{1}}\propto T^{3/2}\exp(2\pi\rho_{s}/T). (6)

Considering the possibility of triple-𝐪\mathbf{q} order [44], we also examine the non-linear σ\sigma model developed for the system with noncollinear order [85, 86, 87]. For this type of order, ξT1/2exp(4πρs/T)\xi\propto T^{-1/2}\exp(4\pi\rho_{s}^{\,\prime}/T) and 1/T1T7/2ξ1/T_{1}\propto T^{7/2}\xi, giving [88]

1T1T3exp(4πρs/T).\displaystyle\frac{1}{T_{1}}\propto T^{3}\exp(4\pi\rho_{s}^{\,\prime}/T). (7)

Here ρs\rho_{s}^{\,\prime} is the corresponding spin stiffness constant.

Refer to caption
Figure 8: Semi-logarithmic plots of 1/T1T3/21/T_{1}T^{3/2} versus 1/T1/T. The lines are fits to the renormalized-classical form 1/T1T3/2exp(2πρs/T)1/T_{1}\propto T^{3/2}\exp(2\pi\rho_{s}/T). Field dependence of ρs\rho_{s} is shown in the inset together with the fitting to the power law ρs(BcB)p\rho_{s}\propto(B_{c}-B)^{p}.

Figure 8 shows 1/T1T3/21/T_{1}T^{3/2} plotted against the inverse temperature 1/T1/T in a semi-logarithmic scale. 1/T11/T_{1} is found to obey the scaling relation Eq. (6) below 50 to 32 K (40 to 24 K at high fields) which ensures that the system is indeed in the RC regime. The scaling form Eq. (7) for noncollinear order is also satisfied in the same TT range (not shown). ρs\rho_{s} decreases monotonically with increasing field as shown in the inset of Fig. 8, showing a tendency to vanish at 10-12 T. The BB dependence of ρs\rho_{s}^{\,\prime} is identical to that of ρs\rho_{s} and is scaled with a multiplicative factor ρs/ρs=1.17\rho_{s}/\rho_{s}^{\,\prime}=1.17 [Fig. 7].

The spin stiffness constant characterizes the rigidity of an ordered state and is nonzero only in a phase with magnetic LRO. In the quantum non-linear σ\sigma model for 2D HAFs, it is renormalized by quantum fluctuations and vanishes on approaching a quantum critical point, beyond which a quantum disordered state appears [81]. The field-induced reduction of ρs\rho_{s} and ρs\rho_{s}^{\,\prime} thus suggests the system getting closer to the quantum disordered phase present above a certain critical field. This also implies that the high-field phase of Na2Co2TeO6 is not a QSL but a partially polarized phase showing bosonic excitations. As a rough estimate of the critical field BcB_{c}, we fitted the BB dependence of ρs\rho_{s} and ρs\rho_{s}^{\,\prime} to the power law ρs,ρs(BcB)p\rho_{s},\rho_{s}^{\,\prime}\propto(B_{c}-B)^{p}, getting Bc=10.4B_{c}=10.4 T and p=0.22p=0.22. The obtained value of BcB_{c} agrees well with the value Bc10B_{c}\approx 10 T determined from the magnetization measurement [48], although the agreement seems fortuitous considering the lack of our data closer to BcB_{c}.

The BB dependence of ρs\rho_{s} and ρs\rho_{s}^{\,\prime} is also shown in Fig. 7 for comparison with the phase diagram. Surprisingly enough, ρs\rho_{s} and ρs\rho_{s}^{\,\prime} trace TNT_{N} as a function of magnetic field with appropriate scale factors. (ρs0.5TN\rho_{s}\approx 0.5T_{N}.) This implies that the 2D spin correlations and the 3D magnetic LRO are characterized by a common energy scale that is renormalized by magnetic field. It is plausible that the magnetic field changes a balance of competing interactions to make the system more frustrated and suppress magnetic LRO, which may appear as a decrease of the characteristic energy and the spin stiffness constant, leading to the suppression of ξ\xi and 1/T11/T_{1}. In fact, frustration reduces the spin stiffness constant in 2D quantum HAFs [89, 90]. It is, however, puzzling that the macroscopic susceptibility scarcely depends on the magnetic field in the PM phase.

1/T11/T_{1} deviate from the RC scaling relations Eqs. (6) and (7) close to TNT_{N}, signaling an onset of 3D critical slowing down. One may expect a power-law dependence of 1/T11/T_{1} on the reduced temperature ϵ=T/TN1\epsilon=T/T_{N}-1 in the critical region, the exponent of which provides information on the universality class of the phase transition. We do not pursuit this subject because 23Na 1/T11/T_{1} is strongly modified by applying field in the critical region, which prevents us from extracting a reliable value of the critical exponent.

V Discussion

V.1 Low-energy spin dynamics in the antiferromagnetic phase

One of the most important characteristics of the spin dynamics in Na2Co2TeO6 is the existence of an intermediate temperature region in the AFM phase (Region III) in which Co spins exhibit unconventional dynamics. The magnetic excitation spectrum in Region III (TN/2T<TNT_{N}/2\lesssim T<T_{N}) comprises of a broad continuum centered at the ordering wave vector 𝐐=(12,0,0)\mathbf{Q}=(\frac{1}{2},0,0) rather than a distinct spin-wave mode [44]. At low energies probed via 1/T11/T_{1}, there remains an unusually large spectral weight of spin fluctuations enhanced strongly with magnetic field. In contrast, the spin dynamics at low TT in Region IV (TTN/2T\lesssim T_{N}/2) looks more conventional. The continuum is replaced with a gapped spin-wave mode, and 1/T11/T_{1} exhibits a rapid decrease on cooling which evidences the disappearance of low-energy spectral weight. The key ingredients in understanding the spin dynamics of Na2Co2TeO6 would thus be (i) the origin of a continuum and the large spectral weight at low energies in Region III, (ii) the origin of field enhancement of the low-energy spectral weight, and (iii) the trigger for the formation of a spin-wave mode in Region IV.

The most likely cause of the broad continuum accompanying a large spectral weight at low energies would be strong damping of spin waves. The appearance of long-lived spin waves at low TT is then understood as resulting from diminished damping. It is apparent that conventional mechanisms for spin-wave damping in collinear antiferromagnets do not apply because they yield too small damping to account for a broad feature of the excitation spectrum in Region III [91]. If allowed by symmetry, cubic anharmonicities in the magnon effective Hamiltonian enable the coupling between one- and two-magnon states which possibly leads to strong damping of the one-magnon mode degenerate with a two-magnon continuum [92, 93]. The cubic terms exist in noncollinear antiferromagnets as well as the Kitaev magnets with off-diagonal interactions and may lead to severe damping at relatively high energies, leaving a low-energy one-magnon mode almost untouched. This is not the case in Na2Co2TeO6 because there is no distinct excitation mode at low energies in Region III on the one hand, and the damping is not so severe at high energies in Region IV on the other [44]. Among other things, the fact that the damping changes rapidly its character around a certain temperature (TN/2\sim T_{N}/2) seems difficult to be accounted for by a known mechanism for damping which usually gives a smooth variation of magnon lifetime with TT.

Although a prime mechanism for the spin-wave damping is unidentified, temperature and field variation of the damping may be argued qualitatively based on the general expression of nuclear spin-lattice relaxation rate [79],

1T1T=2kB𝐪|A(𝐪)|2χ′′(𝐪,ω0)ω0.\displaystyle\frac{1}{T_{1}T}=2k_{\mathrm{B}}\sum_{\mathbf{q}}\lvert A(\mathbf{q})\rvert^{2}\frac{\chi^{\prime\prime}(\mathbf{q},\omega_{0})}{\omega_{0}}. (8)

Here χ′′(𝐪,ω)\chi^{\prime\prime}(\mathbf{q},\omega) is the imaginary part of the dynamical spin susceptibility, ω0\omega_{0} is the nuclear Larmor frequency, and A(𝐪)A(\mathbf{q}) is the hyperfine form factor determined by a geometry of the nuclear site. Equation (8) tells us that 1/T11/T_{1} is determined by a spectral weight of spin fluctuations at a very low frequency ω0108s1\omega_{0}\sim 10^{8}\mathrm{~{}s^{-1}} (ω00.1μeV\hbar\omega_{0}\sim 0.1\mathrm{~{}\mu eV}). Since there appears a staggered hyperfine field at the Na sites below TNT_{N} as revealed from the NMR line broadening, A(𝐐)A(\mathbf{Q}) at the ordering wave vector 𝐐\mathbf{Q} is nonzero and a dominant contribution to 1/T1T1/T_{1}T will come from 𝐪𝐐\mathbf{q}\sim\mathbf{Q} in the AFM phase.

The most intuitive view of the results not relying on the specific model is to interpret 1/T1T1/T_{1}T as a measure of low-energy spin excitations represented by χ′′(𝐪,ω0)\chi^{\prime\prime}(\mathbf{q},\omega_{0}). It is thus apparent from Figs. 6(b) and 7 that the low-energy spin excitations remain enhanced in Region III in the AFM phase, especially at high fields of 6-7 T as pointed out in the preceding section. It is also obvious that the active excitation channels survive to lower temperatures at higher fields. The excitation channels activated in Region III almost disappear on entering Region IV as evidenced by a steep decrease of 1/T1T1/T_{1}T below \sim10 K.

If we take a model of damped harmonic oscillator for spin-wave excitations [94], we may obtain semi-quantitative information on the spin-wave damping. According to the model, we have a contribution of the dynamical susceptibility to 1/T1T1/T_{1}T in the limit of ω00\omega_{0}\rightarrow 0 as χ′′(𝐪,ω0)/ω0γ𝐪/ω𝐪2\chi^{\prime\prime}(\mathbf{q},\omega_{0})/\omega_{0}\propto\gamma_{\mathbf{q}}/\omega_{\mathbf{q}}^{2} where ω𝐪\omega_{\mathbf{q}} is an undamped spin-wave frequency and γ𝐪\gamma_{\mathbf{q}} is a damping constant. ω𝐪\omega_{\mathbf{q}}’s are expected to depend only weakly on TT except in the vicinity of TNT_{N} and unless there is a drastic change of the magnetic structure. The TT dependence of 1/T1T1/T_{1}T at a constant BB is thus dominated by that of γ𝐪\gamma_{\mathbf{q}}. As to the BB dependence, both ω𝐪\omega_{\mathbf{q}} and γ𝐪\gamma_{\mathbf{q}} may vary with BB and affect the behavior of 1/T1T1/T_{1}T because of unknown effects of magnetic field on them.

As shown in Figs. 6(b) and 7, 1/T1T1/T_{1}T is peaked around 6-7 T at a constant TT in Region III. This suggests an increase of γ𝐪\gamma_{\mathbf{q}} and/or a decrease of ω𝐪\omega_{\mathbf{q}} with magnetic field. Notice that the latter matches the field-induced reduction of the characteristic energy inferred from the RC analysis of 1/T11/T_{1} in the PM phase. At a constant BB, on the other hand, the TT dependence of 1/T1T1/T_{1}T should be ascribed to that of γ𝐪\gamma_{\mathbf{q}} as mentioned above. Therefore, a broad hump of 1/T1T1/T_{1}T suggests enhancement of the spin-wave damping at high fields toward the boundary between Regions III and IV. The origin of such strong damping and its unusual TT and BB dependence is unclear. Frustration might play an important role in making the spin excitations incoherent as observed in the triangular-lattice antiferromagnet NaCrO2 below the spin-freezing temperature Tc41T_{c}\approx 41 K [95]. Note, however, that the excitation spectrum of NaCrO2 becomes dispersive below about 0.75Tc0.75T_{c} triggered by the onset of short-range 3D spin correlations, whereas in Na2Co2TeO6 both in-plane and out-of-plane spin correlations show no appreciable change across the boundary between Regions III and IV [38].

The presence of magnetic scattering centers is another possibility of strong spin-wave damping. This may provide a reasonable account for the broad continuum in Region III that changes to the dispersive mode in Region IV as described below. Structural disorder in Na layers and a related distribution of the interlayer magnetic coupling seem less important because they would give TT-independent damping. It is worth mentioning here that the qualitative change of the excitation spectrum with TT in Na2Co2TeO6 resembles the behavior observed in some geometrically frustrated magnets such as the kagomé staircase Ni3V2O8 [96] and the pyrochlore Gd2Ti2O7 [97]. These compounds have an intermediate temperature phase just below TNT_{N} in which the spin excitations are quasielastic or show only broad features and a low temperature phase with collective excitations. The intermediate phases are identified as a partially-disordered state where long-range ordered moments coexist with disordered (paramagnetic) moments, whereas all the sites are ordered in the low temperature phase. This suggests a vital role of partially-disordered moments in the spin-wave damping in these materials. Such a scenario may be realized in Na2Co2TeO6 if the AFM phase has triple-𝐪\mathbf{q} rather than single-𝐪\mathbf{q} zigzag order [44].

In the triple-𝐪\mathbf{q} ordered state shown in Fig. 1(b), three quarters of the Co atoms show noncollinear spin arrangement with a vortex-like texture in the honeycomb planes, while the remaining Co atoms become “spinless”, which means that the ordered moment is absent, or have only an out-of-plane Néel component. The mean fields originating from the in-plane component of the majority spins cancel at the minority site, which may allow the minority spin to fluctuate in a large amplitude. The absence of the signal wipeout implies that if the minority spins are not ordered in Region III, they fluctuate in a time scale much faster than the NMR time scale like a paramagnetic moment. Spin waves propagating on the majority sites would be strongly damped to give a broad continuum if the resulting quasielastic mode of the minority spins overlaps energetically with the spin-wave mode.

Since the minority spins are coupled via the third-neighbor interactions and possibly via the effective interactions mediated by the majority spins due to quantum fluctuations around the order [98, 99], they would participate in magnetic LRO at low enough temperatures exhibiting slowing of spin fluctuations. A broad hump of the magnetic specific heat observed around 10 K [45, 46] might be associated with such a change of the minority spin state. The spin-wave damping would be diminished as a spectral weight of the quasielastic mode shifts to lower energies, restoring collective excitations at low TT.

The low-energy spectral weight of the majority spins arises from the spin-wave damping and is reduced to give a decreasing contribution to 1/T1T1/T_{1}T at low TT. On the other hand, slowing of the minority spin fluctuations would contribute a peak or hump of 1/T1T1/T_{1}T like the case of a magnetic phase transition. The TT and BB dependence of 1/T1T1/T_{1}T should be determined by a balance between the two contributions. Actually, a feeble anomaly of 1/T1T1/T_{1}T around 15 K at low fields implies the crossover nature of slowing down rather than a sharp transition with critical dynamics as noted in the previous section. A gradual increase of the NMR line splitting below \sim15 K [49] might be related to this crossover and the resulting appearance of a static moment on the minority site. The broad hump of 1/T1T1/T_{1}T around 10 K at high fields should then be ascribed primary to the majority spins suffering strong damping even at that temperature for unknown reasons, although it might be possible that the minority spin dynamics becomes more critical at higher fields to contribute the hump. The TT and BB dependence of 1/T1T1/T_{1}T around the boundary between Regions III and IV is complex and is not fully understood in terms of the minority spin ordering. Further investigations are needed to clarify the field-dependent spin dynamics in the AFM phase of Na2Co2TeO6.

V.2 Comparison with Kitaev candidates

It has recently been argued whether Na2Co2TeO6 serves as a canonical example of the Kitaev magnet. As described in the preceding sections, the spin dynamics in Na2Co2TeO6 displays distinct features from those in other Kitaev candidates such as α\alpha-RuCl3 and Na2IrO3. The magnetic excitation spectrum in the AFM phase of Na2Co2TeO6 is characterized by the sole existence of a broad continuum or a distinct spin-wave mode, both of which has a dominant intensity around the M points of the 2D Brillouin zone (𝐐=(12,0,0)\mathbf{Q}=(\frac{1}{2},0,0) and the equivalents) [44]. On the other hand, a continuum in α\alpha-RuCl3 and Na2IrO3 is centered at the Γ\Gamma point (zone center) and coexists with spin-wave modes below TNT_{N} [23, 25]. The major excitations around the M point in Na2Co2TeO6 are possibly ascribed to large third-neighbor coupling J3J_{3} suggested from the powder INS [40, 46, 42, 47]. This is known to stabilize zigzag order but to counteract the formation of a Kitaev QSL [34, 35]. In the context of Kitaev physics, the honeycomb cobaltate Na3Co2SbO6 seems more promising because it exhibits intense excitations around the Γ\Gamma point probably due to smaller J3J_{3} [41, 42].

On the field evolution of the low-energy spin dynamics, it is interesting to compare our results of 1/T11/T_{1} with those for other Kitaev candidates exhibiting a similar field-induced transition from an AFM phase to a spin-disordered phase. To the best of the authors’ knowledge, α\alpha-RuCl3 is only one such example with extensive field-dependent NMR studies [55, 56, 57, 59]. The AFM phase of another well-studied candidate Na2IrO3 is robust against a field [43]. In fact, 1/T11/T_{1} at the Na sites in Na2IrO3 is insensitive to field and shows conventional behaviors above and below TNT_{N} [58].

α\alpha-RuCl3 seems to be the best reference as it shows zigzag order like Na2Co2TeO6. The in-plane critical field BcB_{c} to the high-field disordered phase is 7-8 T [28, 29, 30]. Most of the NMR experiments on α\alpha-RuCl3, however, focused on the behavior near and above BcB_{c} and the field evolution of 1/T11/T_{1} has not been investigated systematically in the low-field region. Although a direct comparison of the results is limited to a narrow field range, the field response of 1/T11/T_{1} has distinct differences between the two compounds. This may reflect the presence of the intermediate temperature region (Region III) characterized by a substantial low-energy spectral weight in Na2Co2TeO6 and the excitation continuum coexisting with spin-wave modes in α\alpha-RuCl3. Indeed, 1/T11/T_{1} at the 35Cl site in the AFM phase of α\alpha-RuCl3 is relatively insensitive to field except above the field Bc=7.1B_{c}^{\prime}=7.1 T (<Bc)(<B_{c}) where gapless magnon excitations have been suggested [59]. This shows a marked contrast to strong field enhancement of 23Na 1/T11/T_{1} in Region III of Na2Co2TeO6 starting far below the critical field Bc10B_{c}\approx 10 T. The field-insensitive response of 1/T11/T_{1} at B<BcB<B_{c}^{\prime} in α\alpha-RuCl3 would be due to gapped magnon excitations, but at T<4T<4 K well below TN=6.5T_{N}=6.5 K, 1/T11/T_{1} is contributed by a residual mode that grows with field on crossing BcB_{c}^{\prime} and becomes dominant above BcB_{c} [55, 59]. Such a mode was not detected in Na2Co2TeO6 down to 3.5 K and may be associated with the continuum around the Γ\Gamma point in α\alpha-RuCl3 identified as excitations inherent to Kitaev QSLs.

Despite the apparent differences in the low-energy spin dynamics probed by 1/T11/T_{1}, the two compounds have a lot of similarities in their response to magnetic field; closing of the AFM phase suggesting the existence of a quantum critical point, the possible appearance of a field-induced QSL phase, and so on. It is often encountered in strongly frustrated magnets that the magnetic LRO is controlled by sub-leading interactions instead of the leading one like the Kitaev coupling and by external perturbations such as magnetic field, pressure, and in some cases spin defects. Since the low-energy sector of a magnetic excitation spectrum is very much affected and reconstructed by these interactions and perturbations, there will be a wide variety of phases and behaviors in real materials which at first glance look very different. Concerning the present case, it might be possible that the unusual temperature and field evolution of low-energy spin dynamics in Na2Co2TeO6 is described in terms of a generalized Kitaev model by including necessary factors. It is worth noting that for the reported values of the Kitaev coupling KK [41, 42, 40, 47], Region III is around or lower than the crossover temperature TH0.375KT_{H}\sim 0.375K below which localized Z2Z_{2} fluxes and itinerant Majorana fermions are expected to emerge [100]. The close resemblance between 1/T1T1/T_{1}T and Cm/TC_{m}/T, however, implies confinement of the fractionalized particles to magnons. Our findings on the low-energy spin dynamics of Na2Co2TeO6 will thus provide new insights into Kitaev-derived spin models as well as more conventional models on the honeycomb lattice, promoting future studies in this research field. From an experimental side, field-dependent microscopic measurements complementing NMR, such as neutron, Raman, and terahertz spectroscopies, are highly required.

VI Summary and Conclusions

We have measured 23Na NMR in the honeycomb lattice antiferromagnet Na2Co2TeO6 to elucidate the phases and the underlying low-energy spin dynamics in a wide range of temperature and magnetic field. The magnetic phase diagram was constructed using the microscopic quantities measured by NMR. The persistence of AFM order up to a field of 9 T was confirmed from the magnetic shift and broadening of the NMR spectrum and the rapidly decreasing 1/T11/T_{1} at low TT.

The AFM phase is divided into two distinct temperature regions that exhibit contrasting low-energy dynamics and its field response. In the intermediate temperature region just below TNT_{N} (Region III), there exists an appreciable low-energy spectral weight of spin fluctuations that contributes to 1/T11/T_{1} and is enhanced strongly with magnetic field. The low temperature region below TN/2\sim T_{N}/2 (Region IV) is characterized by a loss of this low-energy weight as evidenced via a rapid decrease of 1/T11/T_{1} which is less field-dependent. The qualitative change of the low-energy spin dynamics across the boundary between the two regions is consistent with the fact that the magnetic excitation spectrum at higher energies displays an incoherent feature in Region III and a gapped dispersive mode in Region IV [44].

We interpreted the lack of a dispersive mode and the presence of a significant low-energy spectral weight in Region III as arising from strong spin-wave damping. The appearance of a dispersive mode in Region IV is then ascribed to weakening of the damping. As a possible scenario, we suggested a partially-disordered state in Region III with the triple-𝐪\mathbf{q} magnetic structure formed by superposing three equivalent zigzag patterns. In this scenario, the partially-disordered moment experiencing a vanishing mean field acts as a strong scatterer of spin waves propagating on the ordered sites in Region III and acquires an ordered moment to take part in the collective excitations in Region IV. The scattering hence weakens to restore spin-wave excitations at sufficiently low temperatures. The scenario, however, cannot fully account for the complex behavior of 1/T1T1/T_{1}T around the boundary between Regions III and IV and needs further investigation.

We also identified a temperature region with field-dependent 2D spin correlations in the PM phase near TNT_{N} (Region II). The TT dependence of 1/T11/T_{1} in Region II is well reproduced using the renormalized-classical scaling form for 2D quantum antiferromagnets. The field suppression of 1/T11/T_{1} due to a reduction of the in-plane correlation length is described by a monotonic decrease of the spin stiffness constant with magnetic field, suggesting the existence of a high-field disordered phase in the limit of vanishing spin stiffness. The fact that the spin stiffness constant scales with TNT_{N} as a function of magnetic field implies a common energy scale for the 2D spin correlations and 3D magnetic LRO. The magnetic phases and the spin dynamics may be controlled by field-tuning this energy scale, which is likely caused by cancellation of frustrating interactions including an effect of external magnetic fields.

Acknowledgements.
We thank Y. Itoh and H. Kusunose for valuable discussions. This work was partially supported by Grants-in-Aid for Scientific Research, MEXT, Japan (KAKENHI Grant No.15K05149).

*

Appendix A Inverse Laplace transform analysis of 1/T11/T_{1} in the antiferromagnetic phase

It is well known that the stretched exponential analysis of magnetization recovery dictates a specific form of the distribution function for 1/T11/T_{1} [72, 73] which may not reflect the true distribution of 1/T11/T_{1}. One should thus be careful in making concrete statements about results of 1/T11/T_{1} when there exist inequivalent nuclear sites and/or some domains exhibiting distinct spin-lattice relaxation. In this Appendix, we present analysis of magnetization recovery in the AFM phase where the stretching exponent β\beta as well as 1/T11/T_{1} is strongly TT dependent, based on the method of so-called inverse Laplace transform (ILT) which can deduce the probability distribution function P(1/T1)P(1/T_{1}), i.e., the histogram of 1/T11/T_{1}. This method has recently been applied successfully to analyze spatially-inhomogeneous spin-lattice relaxation in high-TcT_{c} cuprates [70, 71].

The ILT analysis assumes that each nucleus decays as a linear combination of normal modes with a definite relaxation rate, but the rate is heterogeneous over the sample and is described by a distribution function P(1/T1)P(1/T_{1}). For nuclei with I=3/2I=3/2, the magnetization recovery M(t)M(t) may be expressed in a discrete form for P(1/T1)P(1/T_{1}) as

M(t)=j=1N[1Ak=13αkeλkt/T1j]\displaystyle M(t)=\sum_{j=1}^{N}\Big{[}1-A\sum_{k=1}^{3}\alpha_{k}e^{-\lambda_{k}t/T_{1j}}\Big{]} P(1/T1j).\displaystyle P(1/T_{1j}). (9)

Here NN is the number of bins for P(1/T1)P(1/T_{1}), AA is a degree of inversion, {λk}={1,3,6}\{\lambda_{k}\}=\{1,3,6\} are mode eigenvalues, and {αk}\{\alpha_{k}\} are amplitudes of the corresponding modes satisfying kαk=1\sum_{k}\alpha_{k}=1. The summation jP(1/T1j)=M()\sum_{j}P(1/T_{1j})=M(\infty) is the equilibrium magnetization. The ILT analysis deduces {P(1/T1j)}\{P(1/T_{1j})\} numerically from recovery data {M(ti)}\{M(t_{i})\} (tit_{i} being the delay time) without assuming any functional form of P(1/T1)P(1/T_{1}). For technical details, see the Supplemental Material of Ref. 70 and references therein. We take 250 bins for P(1/T1)P(1/T_{1}) equally spaced on a logarithmic scale ranging from 102s11/T1j106s110^{-2}\mathrm{~{}s^{-1}}\leq 1/T_{1j}\leq 10^{6}\mathrm{~{}s^{-1}}. Tikhonov regularization method was employed to find the optimal solution. The resulting probability distribution is then normalized as jP(1/T1j)ΔP=1\sum_{j}P(1/T_{1j})\Delta_{P}=1 where ΔP\Delta_{P} is the logarithmic bin spacing.

When using polycrystals, one cannot determine {αk}\{\alpha_{k}\} uniquely because the initial (t=0t=0) populations of the nuclear level are not known exactly. We examined the following models for {αk}\{\alpha_{k}\} to perform ILT utilizing results of the fitting of {M(ti)}\{M(t_{i})\} to Eq. (1): Model A, {αk}\{\alpha_{k}\} taken as those determined at each temperature; Model B, {αk}\{\alpha_{k}\} fixed to the values at 30 K just above TNT_{N}; and Model C, {αk}\{\alpha_{k}\} taken as the average values in Region III (TN/2T<TNT_{N}/2\lesssim T<T_{N}) where {αk}\{\alpha_{k}\} are almost TT independent. We found that P(1/T1)P(1/T_{1}) is relatively insensitive to the choice of the above models for {αk}\{\alpha_{k}\}.

Refer to caption
Figure 9: Probability distribution function P(1/T1)P(1/T_{1}) deduced from the ILT analysis of the magnetization recovery at T=6T=6 K at B=3B=3 T. Mode amplitudes {αk}\{\alpha_{k}\} of each model are as follows; Model A (dotted line), {0.24,0,0.76}\{0.24,0,0.76\}; Model B (dashed-dotted line), {0.17,0.12,0.71}\{0.17,0.12,0.71\}; Model C (solid line), {0.17,0.22,0.61}\{0.17,0.22,0.61\}. The dashed line is P(1/T1)P(1/T_{1}) for the stretched exponential recovery with β=0.73\beta=0.73 and 1/T1=3.4s11/T_{1}=3.4\mathrm{~{}s^{-1}} calculated using Eq. (11) of Ref. 73.

Figure 9 displays the distribution functions obtained from the recovery at T=6T=6 K at B=3B=3 T shown in Fig. 6(d). The P(1/T1)P(1/T_{1})’s are almost identical, peaked at 1/T13s11/T_{1}\sim 3\mathrm{~{}s^{-1}} and having nearly a decade width, except that P(1/T1)P(1/T_{1}) for Model A exhibits sharper cutoff at the side of low relaxation rates than Models B and C. Wiggly subpeaks appearing at the side of high relaxation rates would be an oscillatory artifact [101]. In fact, a position of the subpeak depends on the details of the ILT analysis such as the number of bins and the choice of the regularization (smoothing) factor. The ILT fits to {M(ti)}\{M(t_{i})\} (not shown) were as good as the stretched exponential fit.

The P(1/T1)P(1/T_{1})’s deduced from the ILT analysis resemble the one for the stretched exponential function in that they are single-peaked and have a long tail at the side of high relaxation rates [72, 73]. For comparison, we calculated P(1/T1)P(1/T_{1}) numerically for the stretched exponential function using the expression given in Ref. 73. The exponent β=0.73\beta=0.73 and 1/T1=3.4s11/T_{1}=3.4\mathrm{~{}s^{-1}} obtained from the fitting to Eq. (1) were used. The overall line shape of the ILT P(1/T1)P(1/T_{1})’s is well reproduced by P(1/T1)P(1/T_{1}) for the stretched exponential function.

Refer to caption
Figure 10: Temperature evolution of the probability distribution function P(1/T1)P(1/T_{1}) below T=30T=30 K at B=3B=3 T. Filled bullets represent the location of the log means 1/T1lm1/T_{1}^{\mathrm{lm}} of P(1/T1)P(1/T_{1}). Open bullets are 1/T11/T_{1} determined by the stretched exponential fitting using Eq. (1).

Figure 10 shows TT evolution of P(1/T1)P(1/T_{1}) below T=30T=30 K at B=3B=3 T 777The mode amplitudes at 30 K above TNT_{N} were taken as {αk}={0.17,0.12,0.71}\{\alpha_{k}\}=\{0.17,0.12,0.71\} using the result of stretched exponential fitting.. We adopted Model C for {αk}\{\alpha_{k}\} based on the idea that {αk}\{\alpha_{k}\} would not depend strongly on TT in the AFM phase where the NMR line width does not vary strongly with TT. As the temperature is lowered, P(1/T1)P(1/T_{1}) becomes progressively broader, exhibiting an oscillatory tail at the side of high relaxation rates. No extra peak showing a distinct TT variation of 1/T11/T_{1} from the main peak appears. The log means 1/T1lm1/T_{1}^{\mathrm{lm}} of the distribution function defined by ln(1/T1lm)=jln(1/T1j)P(1/T1j)ΔP\ln(1/T_{1}^{\mathrm{lm}})=\sum_{j}\ln(1/T_{1j})P(1/T_{1j})\Delta_{P} [70] are marked as filled bullets on each P(1/T1)P(1/T_{1}) curve in Fig. 10. They are in good agreement with 1/T11/T_{1} determined by the stretched exponential fitting and shown as open bullets. The line shape is also consistent with the stretched exponential analysis at each temperature. These facts indicate that the distribution of 1/T11/T_{1} at the Na sites and its TT evolution are well captured by the phenomenological stretched exponential analysis, justifying the resulting 1/T11/T_{1} and β\beta as representing the average relaxation rate and the distribution of 1/T11/T_{1} in the AFM phase.

We performed the same analysis at B=7B=7 T. The results are qualitatively similar to those at B=3B=3 T. This means that the low-energy spin dynamics does not change much with field in a range covered in the present study as far as the distribution of 1/T11/T_{1} is concerned.

References

  • Anderson [1973] P. W. Anderson, Resonating valence bonds: A new kind of insulator?, Mater. Res. Bull. 8, 153 (1973).
  • Lacroix et al. [2011] C. Lacroix, P. Mendels, and F. Mila, eds., Introduction to Frustrated Magnetism (Springer, Berlin, 2011).
  • Balents [2010] L. Balents, Spin liquids in frustrated magnets, Nature 464, 199 (2010).
  • Savary and Balents [2016] L. Savary and L. Balents, Quantum spin liquids: a review, Rep. Prog. Phys. 80, 016502 (2016).
  • Broholm et al. [2020] C. Broholm, R. J. Cava, S. A. Kivelson, D. G. Nocera, M. R. Norman, and T. Senthil, Quantum spin liquids, Science 367, eaay0668 (2020).
  • Rastelli et al. [1979] E. Rastelli, A. Tassi, and L. Reatto, Non-simple magnetic order for simple Hamiltonians, Physica B+C 97, 1 (1979).
  • Fouet et al. [2001] J. B. Fouet, P. Sindzingre, and C. Lhuillier, An investigation of the quantum J1{J_{1}}-J2{J_{2}}-J3{J_{3}} model on the honeycomb lattice, Eur. Phys. J. B 20, 241 (2001).
  • Mulder et al. [2010] A. Mulder, R. Ganesh, L. Capriotti, and A. Paramekanti, Spiral order by disorder and lattice nematic order in a frustrated Heisenberg antiferromagnet on the honeycomb lattice, Phys. Rev. B 81, 214419 (2010).
  • Albuquerque et al. [2011] A. F. Albuquerque, D. Schwandt, B. Hetényi, S. Capponi, M. Mambrini, and A. M. Läuchli, Phase diagram of a frustrated quantum antiferromagnet on the honeycomb lattice: Magnetic order versus valence-bond crystal formation, Phys. Rev. B 84, 024406 (2011).
  • Reuther et al. [2011] J. Reuther, D. A. Abanin, and R. Thomale, Magnetic order and paramagnetic phases in the quantum J1{J}_{1}-J2{J}_{2}-J3{J}_{3} honeycomb model, Phys. Rev. B 84, 014417 (2011).
  • Li et al. [2012] P. H. Y. Li, R. F. Bishop, D. J. J. Farnell, and C. E. Campbell, Phase diagram of a frustrated Heisenberg antiferromagnet on the honeycomb lattice: The J1{J}_{1}-J2{J}_{2}-J3{J}_{3} model, Phys. Rev. B 86, 144404 (2012).
  • Kitaev [2006] A. Kitaev, Anyons in an exactly solved model and beyond, Ann. Phys. 321, 2 (2006).
  • Winter et al. [2017a] S. M. Winter, A. A. Tsirlin, M. Daghofer, J. van den Brink, Y. Singh, P. Gegenwart, and R. Valentí, Models and materials for generalized Kitaev magnetism, J. Phys.: Condens. Matter 29, 493002 (2017a).
  • Takagi et al. [2019] H. Takagi, T. Takayama, G. Jackeli, G. Khaliullin, and S. E. Nagler, Concept and realization of Kitaev quantum spin liquids, Nat. Rev. Phys. 1, 264 (2019).
  • Janssen and Vojta [2019] L. Janssen and M. Vojta, Heisenberg-Kitaev physics in magnetic fields, J. Phys.: Condens. Matter 31, 423002 (2019).
  • Motome and Nasu [2020] Y. Motome and J. Nasu, Hunting Majorana Fermions in Kitaev Magnets, J. Phys. Soc. Jpn. 89, 012002 (2020).
  • Trebst and Hickey [2022] S. Trebst and C. Hickey, Kitaev materials, Phys. Rep. 950, 1 (2022).
  • Jackeli and Khaliullin [2009] G. Jackeli and G. Khaliullin, Mott insulators in the strong spin-orbit coupling limit: From Heisenberg to a quantum compass and Kitaev models, Phys. Rev. Lett. 102, 017205 (2009).
  • Chaloupka et al. [2010] J. Chaloupka, G. Jackeli, and G. Khaliullin, Kitaev-Heisenberg model on a honeycomb lattice: Possible exotic phases in iridium oxides A2IrO3{A}_{2}\mathrm{IrO_{3}}Phys. Rev. Lett. 105, 027204 (2010).
  • Abragam and Bleaney [1970] A. Abragam and B. Bleaney, Electron Paramagnetic Resonance of Transition Ions (Oxford University Press, Oxford, 1970).
  • Chaloupka et al. [2013] J. Chaloupka, G. Jackeli, and G. Khaliullin, Zigzag Magnetic Order in the Iridium Oxide Na2IrO3\mathrm{Na_{2}IrO_{3}}Phys. Rev. Lett. 110, 097204 (2013).
  • Rau et al. [2014] J. G. Rau, E. K.-H. Lee, and H.-Y. Kee, Generic Spin Model for the Honeycomb Iridates beyond the Kitaev Limit, Phys. Rev. Lett. 112, 077204 (2014).
  • Banerjee et al. [2017] A. Banerjee, J. Yan, J. Knolle, C. A. Bridges, M. B. Stone, M. D. Lumsden, D. G. Mandrus, D. A. Tennant, R. Moessner, and S. E. Nagler, Neutron scattering in the proximate quantum spin liquid α\alpha-RuCl3\mathrm{RuCl_{3}}Science 356, 1055 (2017).
  • Banerjee et al. [2018] A. Banerjee, P. Lampen-Kelley, J. Knolle, C. Balz, A. A. Aczel, B. Winn, Y. Liu, D. Pajerowski, J. Yan, C. A. Bridges, A. T. Savici, B. C. Chakoumakos, M. D. Lumsden, D. A. Tennant, R. Moessner, D. G. Mandrus, and S. E. Nagler, Excitations in the field-induced quantum spin liquid state of α\alpha-RuCl3\mathrm{RuCl_{3}}npj Quantum Mater. 3, 8 (2018).
  • Kim et al. [2020] J. Kim, J. Chaloupka, Y. Singh, J. W. Kim, B. J. Kim, D. Casa, A. Said, X. Huang, and T. Gog, Dynamic Spin Correlations in the Honeycomb Lattice Na2IrO3\mathrm{Na_{2}IrO_{3}} Measured by Resonant Inelastic x-Ray Scattering, Phys. Rev. X 10, 021034 (2020).
  • Kubota et al. [2015] Y. Kubota, H. Tanaka, T. Ono, Y. Narumi, and K. Kindo, Successive magnetic phase transitions in α\alpha-RuCl3\mathrm{RuCl_{3}}: XY-like frustrated magnet on the honeycomb lattice, Phys. Rev. B 91, 094422 (2015).
  • Mehlawat et al. [2017] K. Mehlawat, A. Thamizhavel, and Y. Singh, Heat capacity evidence for proximity to the Kitaev quantum spin liquid in A2IrO3{A}_{2}\mathrm{IrO_{3}} (A=Na,Li{A}=\mathrm{Na,~{}Li}), Phys. Rev. B 95, 144406 (2017).
  • Johnson et al. [2015] R. D. Johnson, S. C. Williams, A. A. Haghighirad, J. Singleton, V. Zapf, P. Manuel, I. I. Mazin, Y. Li, H. O. Jeschke, R. Valentí, and R. Coldea, Monoclinic crystal structure of α\alpha-RuCl3\mathrm{RuCl_{3}} and the zigzag antiferromagnetic ground state, Phys. Rev. B 92, 235119 (2015).
  • Sears et al. [2017] J. A. Sears, Y. Zhao, Z. Xu, J. W. Lynn, and Y.-J. Kim, Phase diagram of α\alpha-RuCl3\mathrm{RuCl_{3}} in an in-plane magnetic field, Phys. Rev. B 95, 180411(R) (2017).
  • Wolter et al. [2017] A. U. B. Wolter, L. T. Corredor, L. Janssen, K. Nenkov, S. Schönecker, S.-H. Do, K.-Y. Choi, R. Albrecht, J. Hunger, T. Doert, M. Vojta, and B. Büchner, Field-induced quantum criticality in the Kitaev system α\alpha-RuCl3\mathrm{RuCl_{3}}Phys. Rev. B 96, 041405(R) (2017).
  • Hickey and Trebst [2019] C. Hickey and S. Trebst, Emergence of a field-driven U(1){U(1)} spin liquid in the Kitaev honeycomb model, Nat. Commun. 10, 530 (2019).
  • Liu and Khaliullin [2018] H. Liu and G. Khaliullin, Pseudospin exchange interactions in d7{d}^{7} cobalt compounds: Possible realization of the Kitaev model, Phys. Rev. B 97, 014407 (2018).
  • Sano et al. [2018] R. Sano, Y. Kato, and Y. Motome, Kitaev-Heisenberg Hamiltonian for high-spin d7{d}^{7} Mott insulators, Phys. Rev. B 97, 014408 (2018).
  • Liu et al. [2020] H. Liu, J. Chaloupka, and G. Khaliullin, Kitaev Spin Liquid in 3d3d Transition Metal Compounds, Phys. Rev. Lett. 125, 047201 (2020).
  • Liu [2021] H. Liu, Towards Kitaev spin liquid in 3d3d transition metal compounds, Int. J. Mod. Phys. B 35, 2130006 (2021).
  • Viciu et al. [2007] L. Viciu, Q. Huang, E. Morosan, H. Zandbergen, N. Greenbaum, T. McQueen, and R. Cava, Structure and basic magnetic properties of the honeycomb lattice compounds Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}} and Na3Co2SbO6\mathrm{Na_{3}Co_{2}SbO_{6}}J. Solid State Chem. 180, 1060 (2007).
  • Lefrançois et al. [2016] E. Lefrançois, M. Songvilay, J. Robert, G. Nataf, E. Jordan, L. Chaix, C. V. Colin, P. Lejay, A. Hadj-Azzem, R. Ballou, and V. Simonet, Magnetic properties of the honeycomb oxide Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Phys. Rev. B 94, 214416 (2016).
  • Bera et al. [2017] A. K. Bera, S. M. Yusuf, A. Kumar, and C. Ritter, Zigzag antiferromagnetic ground state with anisotropic correlation lengths in the quasi-two-dimensional honeycomb lattice compound Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Phys. Rev. B 95, 094424 (2017).
  • Xiao et al. [2019] G. Xiao, Z. Xia, W. Zhang, X. Yue, S. Huang, X. Zhang, F. Yang, Y. Song, M. Wei, H. Deng, and D. Jiang, Crystal growth and the magnetic properties of Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}} with quasi-two-dimensional honeycomb lattice, Cryst. Growth Des. 19, 2658 (2019).
  • Samarakoon et al. [2021] A. M. Samarakoon, Q. Chen, H. Zhou, and V. O. Garlea, Static and dynamic magnetic properties of honeycomb lattice antiferromagnets Na2M2TeO6\mathrm{Na_{2}}{M}_{2}\mathrm{TeO_{6}}, M=Co{M}=\mathrm{Co} and Ni\mathrm{Ni}Phys. Rev. B 104, 184415 (2021).
  • Songvilay et al. [2020] M. Songvilay, J. Robert, S. Petit, J. A. Rodriguez-Rivera, W. D. Ratcliff, F. Damay, V. Balédent, M. Jiménez-Ruiz, P. Lejay, E. Pachoud, A. Hadj-Azzem, V. Simonet, and C. Stock, Kitaev interactions in the Co honeycomb antiferromagnets Na3Co2SbO6\mathrm{Na_{3}Co_{2}SbO_{6}} and Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Phys. Rev. B 102, 224429 (2020).
  • Kim et al. [2022] C. Kim, J. Jeong, G. Lin, P. Park, T. Masuda, S. Asai, S. Itoh, H.-S. Kim, H. Zhou, J. Ma, and J.-G. Park, Antiferromagnetic Kitaev interaction in jeff=1/2j_{\mathrm{eff}}=1/2 cobalt honeycomb materials Na3Co2SbO6\mathrm{Na_{3}Co_{2}SbO_{6}} and Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}J. Phys.: Condens. Matter 34, 045802 (2022).
  • Ye et al. [2012] F. Ye, S. Chi, H. Cao, B. C. Chakoumakos, J. A. Fernandez-Baca, R. Custelcean, T. F. Qi, O. B. Korneta, and G. Cao, Direct evidence of a zigzag spin-chain structure in the honeycomb lattice: A neutron and x-ray diffraction investigation of single-crystal Na2IrO3\mathrm{Na_{2}IrO_{3}}Phys. Rev. B 85, 180403(R) (2012).
  • Chen et al. [2021] W. Chen, X. Li, Z. Hu, Z. Hu, L. Yue, R. Sutarto, F. He, K. Iida, K. Kamazawa, W. Yu, X. Lin, and Y. Li, Spin-orbit phase behavior of Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}} at low temperatures, Phys. Rev. B 103, L180404 (2021).
  • Yao and Li [2020] W. Yao and Y. Li, Ferrimagnetism and anisotropic phase tunability by magnetic fields in Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Phys. Rev. B 101, 085120 (2020).
  • Lin et al. [2021] G. Lin, J. Jeong, C. Kim, Y. Wang, Q. Huang, T. Masuda, S. Asai, S. Itoh, G. Günther, M. Russina, Z. Lu, J. Sheng, L. Wang, J. Wang, G. Wang, Q. Ren, C. Xi, W. Tong, L. Ling, Z. Liu, L. Wu, J. Mei, Z. Qu, H. Zhou, X. Wang, J.-G. Park, Y. Wan, and J. Ma, Field-induced quantum spin disordered state in spin-1/2 honeycomb magnet Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Nat. Commun. 12, 5559 (2021).
  • Sanders et al. [2022] A. L. Sanders, R. A. Mole, J. Liu, A. J. Brown, D. Yu, C. D. Ling, and S. Rachel, Dominant Kitaev interactions in the honeycomb materials Na3Co2SbO6\mathrm{Na_{3}Co_{2}SbO_{6}} and Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Phys. Rev. B 106, 014413 (2022).
  • Hong et al. [2021] X. Hong, M. Gillig, R. Hentrich, W. Yao, V. Kocsis, A. R. Witte, T. Schreiner, D. Baumann, N. Pérez, A. U. B. Wolter, Y. Li, B. Büchner, and C. Hess, Strongly scattered phonon heat transport of the candidate Kitaev material Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Phys. Rev. B 104, 144426 (2021).
  • Lee et al. [2021] C. H. Lee, S. Lee, Y. S. Choi, Z. H. Jang, R. Kalaivanan, R. Sankar, and K.-Y. Choi, Multistage development of anisotropic magnetic correlations in the Co-based honeycomb lattice Na2Co2TeO6\mathrm{Na_{2}Co_{2}TeO_{6}}Phys. Rev. B 103, 214447 (2021).
  • Andrew and Tunstall [1961] E. R. Andrew and D. P. Tunstall, Spin-Lattice Relaxation in Imperfect Cubic Crystals and in Non-cubic Crystals, Proc. Phys. Soc. 78, 1 (1961).
  • Narath [1967] A. Narath, Nuclear Spin-Lattice Relaxation in Hexagonal Transition Metals: Titanium, Phys. Rev. 162, 320 (1967).
  • Suter et al. [1998] A. Suter, M. Mali, J. Roos, and D. Brinkmann, Mixed magnetic and quadrupolar relaxation in the presence of a dominant static Zeeman Hamiltonian, J. Phys.: Condens. Matter 10, 5977 (1998).
  • McHenry et al. [1972] M. R. McHenry, B. G. Silbernagel, and J. H. Wernick, Nuclear spin-lattice relaxation in the La1cGdcAl2\mathrm{La}_{1-c}\mathrm{Gd}_{c}\mathrm{Al}_{2} intermetallic compounds, Phys. Rev. B 5, 2958 (1972).
  • Thayamballi and Hone [1980] P. Thayamballi and D. Hone, Nuclear relaxation in a randomly diluted Heisenberg paramagnet, Phys. Rev. B 21, 1766 (1980).
  • Baek et al. [2017] S.-H. Baek, S.-H. Do, K.-Y. Choi, Y. S. Kwon, A. U. B. Wolter, S. Nishimoto, J. van den Brink, and B. Büchner, Evidence for a field-induced quantum spin liquid in α\alpha-RuCl3\mathrm{RuCl_{3}}Phys. Rev. Lett. 119, 037201 (2017).
  • Zheng et al. [2017] J. Zheng, K. Ran, T. Li, J. Wang, P. Wang, B. Liu, Z.-X. Liu, B. Normand, J. Wen, and W. Yu, Gapless Spin Excitations in the Field-Induced Quantum Spin Liquid Phase of α\alpha-RuCl3\mathrm{RuCl_{3}}Phys. Rev. Lett. 119, 227208 (2017).
  • Janša et al. [2018] N. Janša, A. Zorko, M. Gomilšek, M. Pregelj, K. W. Krämer, D. Biner, A. Biffin, C. Rüegg, and M. Klanjšek, Observation of two types of fractional excitation in the Kitaev honeycomb magnet, Nat. Phy. 14, 786 (2018).
  • Takahashi et al. [2019] S. K. Takahashi, J. Wang, A. Arsenault, T. Imai, M. Abramchuk, F. Tafti, and P. M. Singer, Spin Excitations of a Proximate Kitaev Quantum Spin Liquid Realized in Cu2IrO3\mathrm{Cu_{2}IrO_{3}}Phys. Rev. X 9, 031047 (2019).
  • Nagai et al. [2020] Y. Nagai, T. Jinno, J. Yoshitake, J. Nasu, Y. Motome, M. Itoh, and Y. Shimizu, Two-step gap opening across the quantum critical point in the Kitaev honeycomb magnet α\alpha-RuCl3\mathrm{RuCl_{3}}Phys. Rev. B 101, 020414(R) (2020).
  • Yu et al. [2018] Y. J. Yu, Y. Xu, K. J. Ran, J. M. Ni, Y. Y. Huang, J. H. Wang, J. S. Wen, and S. Y. Li, Ultralow-Temperature Thermal Conductivity of the Kitaev Honeycomb Magnet α\alpha-RuCl3\mathrm{RuCl_{3}} across the Field-Induced Phase Transition, Phys. Rev. Lett. 120, 067202 (2018).
  • Abragam [1961] A. Abragam, The Principles of Nuclear Magnetism (Oxford University Press, Oxford, 1961).
  • Note [1] This plot is useful in determining the hyperfine coupling constant when magnetization shows non-linear increase at high fields so that the susceptibility χ\chi defined by M/BM/B depends on a magnetic field.
  • MacLaughlin and Alloul [1976] D. E. MacLaughlin and H. Alloul, Host Nuclear Resonance in a Spin-Glass: CuMnCu\mathrm{Mn}Phys. Rev. Lett. 36, 1158 (1976).
  • Mendels et al. [2000] P. Mendels, A. Keren, L. Limot, M. Mekata, G. Collin, and M. Horvatić, Ga NMR Study of the Local Susceptibility in Kagomé-Based SrCr8Ga4O19\mathrm{SrCr_{8}Ga_{4}O_{19}}: Pseudogap and Paramagnetic Defects, Phys. Rev. Lett. 85, 3496 (2000).
  • Olariu et al. [2006] A. Olariu, P. Mendels, F. Bert, B. G. Ueland, P. Schiffer, R. F. Berger, and R. J. Cava, Unconventional Dynamics in Triangular Heisenberg Antiferromagnet NaCrO2\mathrm{NaCrO_{2}}Phys. Rev. Lett. 97, 167203 (2006).
  • Note [2] We detected a dip in the uncorrected intensity (typically taken at τ=40μs\tau=40\mathrm{~{}\mu s}) around TNT_{N} as reported in Ref. \rev@citealpChen21 which is recovered by the extrapolation process.
  • Itoh [2015] Y. Itoh, 23Na Nuclear Spin-Lattice Relaxation Studies of Na2Ni2TeO6\mathrm{Na_{2}Ni_{2}TeO_{6}}J. Phys. Soc. Jpn. 84, 064714 (2015).
  • Beeman and Pincus [1968] D. Beeman and P. Pincus, Nuclear Spin-Lattice Relaxation in Magnetic Insulators, Phys. Rev. 166, 359 (1968).
  • Note [3] The stretched exponential function used in Ref. \rev@citealpChen21 is slightly different from conventional ones: the exponential terms are like exp[6(t/T1)β]\exp[-6(t/T_{1})^{\beta}] rather than exp[(6t/T1)β]\exp[-(6t/T_{1})^{\beta}]. If this form of stretched exponentials is applied to our recovery, the resulting 1/T11/T_{1} agrees quantitatively with their 1/T11/T_{1}.
  • Singer et al. [2020] P. M. Singer, A. Arsenault, T. Imai, and M. Fujita, La139{}^{139}\mathrm{La} NMR investigation of the interplay between lattice, charge, and spin dynamics in the charge-ordered high-Tc{T}_{c} cuprate La1.875Ba0.125CuO4\mathrm{La_{1.875}Ba_{0.125}CuO_{4}}Phys. Rev. B 101, 174508 (2020).
  • Arsenault et al. [2020] A. Arsenault, T. Imai, P. M. Singer, K. M. Suzuki, and M. Fujita, Magnetic inhomogeneity in charge-ordered La1.885Sr0.115CuO4\mathrm{La_{1.885}Sr_{0.115}CuO_{4}} studied by NMR, Phys. Rev. B 101, 184505 (2020).
  • Johnston et al. [2005] D. C. Johnston, S.-H. Baek, X. Zong, F. Borsa, J. Schmalian, and S. Kondo, Dynamics of magnetic defects in heavy fermion LiV2O4\mathrm{LiV_{2}O_{4}} from stretched exponential Li7{}^{7}\mathrm{Li} NMR relaxation, Phys. Rev. Lett. 95, 176408 (2005).
  • Johnston [2006] D. C. Johnston, Stretched exponential relaxation arising from a continuous sum of exponential decays, Phys. Rev. B 74, 184430 (2006).
  • Note [4] Vanishingly small quadrupolar splitting or negligible quadrupole coupling seems improbable for their missing of quadrupolar satellites because they can fit the magnetization recovery using the form for a quadrupolar-split center line.
  • Walstedt [1967] R. E. Walstedt, Spin-Lattice Relaxation of Nuclear Spin Echoes in Metals, Phys. Rev. Lett. 19, 146 (1967).
  • Auler et al. [1996] T. Auler, P. Butaud, and J. A. Gillet, A fictitious-spin calculation of the nuclear magnetic relaxation of nuclei of spin I{I} = 3/2 in the presence of strong quadrupole coupling, J. Phys.: Condens. Matter 8, 6425 (1996).
  • Kitagawa et al. [2018] K. Kitagawa, T. Takayama, Y. Matsumoto, A. Kato, R. Takano, Y. Kishimoto, S. Bette, R. Dinnebier, G. Jackeli, and H. Takagi, A spin-orbital-entangled quantum liquid on a honeycomb lattice, Nature 554, 341 (2018).
  • Note [5] Since the magnetic response of Na2Co2TeO6 is insensitive to the field applied perpendicular to the honeycomb planes, it seems reasonable to consider that the field dependence measured in powders is governed by in-plane field responses. This is further supported by the fact that in powders a grain with the field lying in the honeycomb plane is found more frequently than a grain with the field normal to the plane, because the probability of a field making an angle θ\theta with respect to the direction normal to the plane is proportional to sinθ\sin\theta.
  • Moriya [1956a] T. Moriya, Nuclear Magnetic Relaxation in Antiferromagnetics, Prog. Theor. Phys. 16, 23 (1956a).
  • Moriya [1956b] T. Moriya, Nuclear Magnetic Relaxation in Antiferromagnetics, II, Prog. Theor. Phys. 16, 641 (1956b).
  • Chakravarty et al. [1989] S. Chakravarty, B. I. Halperin, and D. R. Nelson, Two-dimensional quantum Heisenberg antiferromagnet at low temperatures, Phys. Rev. B 39, 2344 (1989).
  • Chakravarty and Orbach [1990] S. Chakravarty and R. Orbach, Electron and nuclear magnetic relaxation in La2CuO4\mathrm{La_{2}CuO_{4}} and related cuprates, Phys. Rev. Lett. 64, 224 (1990).
  • Imai et al. [1993] T. Imai, C. P. Slichter, K. Yoshimura, and K. Kosuge, Low frequency spin dynamics in undoped and Sr-doped La2CuO4\mathrm{La_{2}CuO_{4}}Phys. Rev. Lett. 70, 1002 (1993).
  • Note [6] Including the correction term of O(T/2πρs)O(T/2\pi\rho_{s}) for ξ\xi introduces a minor change of parameters (\sim10% reduction of ρs\rho_{s} for example) but does not change the results qualitatively.
  • Azaria et al. [1992] P. Azaria, B. Delamotte, and D. Mouhanna, Low-temperature properties of two-dimensional frustrated quantum antiferromagnets, Phys. Rev. Lett. 68, 1762 (1992).
  • Chubukov et al. [1994a] A. V. Chubukov, T. Senthil, and S. Sachdev, Universal magnetic properties of frustrated quantum antiferromagnets in two dimensions, Phys. Rev. Lett. 72, 2089 (1994a).
  • Chubukov et al. [1994b] A. V. Chubukov, S. Sachdev, and T. Senthil, Quantum phase transitions in frustrated quantum antiferromagnets, Nucl. Phys. B 426, 601 (1994b).
  • Itoh et al. [2009] Y. Itoh, C. Michioka, K. Yoshimura, K. Nakajima, and H. Sato, Critical Slowing Down of Triangular Lattice Spin-3/2 Heisenberg Antiferromagnet Li7RuO6\mathrm{Li_{7}RuO_{6}} via Li7{}^{7}\mathrm{Li} NMR Heisenberg Antiferromagnet Li7RuO6\mathrm{Li_{7}RuO_{6}} via Li7{}^{7}\mathrm{Li} NMR, J. Phys. Soc. Jpn. 78, 023705 (2009).
  • Einarsson and Schulz [1995] T. Einarsson and H. J. Schulz, Direct calculation of the spin stiffness in the J1{J_{1}}-J2{J_{2}} Heisenberg antiferromagnet, Phys. Rev. B 51, 6151 (1995).
  • Bishop et al. [2015] R. F. Bishop, P. H. Y. Li, O. Götze, J. Richter, and C. E. Campbell, Frustrated Heisenberg antiferromagnet on the honeycomb lattice: Spin gap and low-energy parameters, Phys. Rev. B 92, 224434 (2015).
  • Harris et al. [1971] A. B. Harris, D. Kumar, B. I. Halperin, and P. C. Hohenberg, Dynamics of an Antiferromagnet at Low Temperatures: Spin-Wave Damping and Hydrodynamics, Phys. Rev. B 3, 961 (1971).
  • Zhitomirsky and Chernyshev [2013] M. E. Zhitomirsky and A. L. Chernyshev, Colloquium: Spontaneous magnon decays, Rev. Mod. Phys. 85, 219 (2013).
  • Winter et al. [2017b] S. M. Winter, K. Riedl, P. A. Maksimov, A. L. Chernyshev, A. Honecker, and R. Valentí, Breakdown of magnons in a strongly spin-orbital coupled magnet, Nat. Commun. 8, 1152 (2017b).
  • Chaikin and Lubensky [1995] P. M. Chaikin and T. C. Lubensky, Principles of Condensed Matter Physics (Cambridge University Press, Cambridge, 1995).
  • Hsieh et al. [2008] D. Hsieh, D. Qian, R. Berger, R. Cava, J. Lynn, Q. Huang, and M. Hasan, Magnetic excitations in triangular lattice NaCrO2\mathrm{NaCrO_{2}}J. Phys. Chem. Solids 69, 3174 (2008).
  • Ehlers et al. [2015] G. Ehlers, A. A. Podlesnyak, M. D. Frontzek, A. V. Pushkarev, S. V. Shiryaev, and S. Barilo, Damped spin waves in the intermediate ordered phases in Ni3V2O8\mathrm{Ni_{3}V_{2}O_{8}}J. Phys.: Condens. Matter 27, 256003 (2015).
  • Paddison et al. [2021] J. A. M. Paddison, G. Ehlers, A. B. Cairns, J. S. Gardner, O. A. Petrenko, N. P. Butch, D. D. Khalyavin, P. Manuel, H. E. Fischer, H. Zhou, A. L. Goodwin, and J. R. Stewart, Suppressed-moment 2-𝐤\mathbf{k} order in the canonical frustrated antiferromagnet Gd2Ti2O7\mathrm{Gd_{2}Ti_{2}O_{7}}npj Quantum Mater. 6, 99 (2021).
  • Gonzalez et al. [2019] M. G. Gonzalez, F. T. Lisandrini, G. G. Blesio, A. E. Trumper, C. J. Gazza, and L. O. Manuel, Correlated partial disorder in a weakly frustrated quantum antiferromagnet, Phys. Rev. Lett. 122, 017201 (2019).
  • Seifert and Vojta [2019] U. F. P. Seifert and M. Vojta, Theory of partial quantum disorder in the stuffed honeycomb Heisenberg antiferromagnet, Phys. Rev. B 99, 155156 (2019).
  • Yoshitake et al. [2016] J. Yoshitake, J. Nasu, and Y. Motome, Fractional Spin Fluctuations as a Precursor of Quantum Spin Liquids: Majorana Dynamical Mean-Field Study for the Kitaev Model, Phys. Rev. Lett. 117, 157203 (2016).
  • Choi et al. [2021] H. Choi, I. Vinograd, C. Chaffey, and N. J. Curro, Inverse Laplace transformation analysis of stretched exponential relaxation, J. Magn. Reson. 331, 107050 (2021).
  • Note [7] The mode amplitudes at 30 K above TNT_{N} were taken as {αk}={0.17,0.12,0.71}\{\alpha_{k}\}=\{0.17,0.12,0.71\} using the result of stretched exponential fitting.