Significance
We report experimental evidence for a glassy state induced by quantum fluctuations, a spin jam, that is realized in SrCr9pGa12-9pO19 [SCGO(p)], a highly frustrated magnet, in which the magnetic Cr3+ (s = 3/2) ions form a quasi-two-dimensional triangular system of bipyramids. Our new experimental data and our theoretical spin jam model provide, for the first time, to our knowledge, a coherent understanding of the existing experimental data of this fascinating system. Furthermore, our findings strongly support the possible existence of purely topological glassy states.
Keywords: spin jam, frustration, glassy states
Abstract
Since the discovery of spin glasses in dilute magnetic systems, their study has been largely focused on understanding randomness and defects as the driving mechanism. The same paradigm has also been applied to explain glassy states found in dense frustrated systems. Recently, however, it has been theoretically suggested that different mechanisms, such as quantum fluctuations and topological features, may induce glassy states in defect-free spin systems, far from the conventional dilute limit. Here we report experimental evidence for existence of a glassy state, which we call a spin jam, in the vicinity of the clean limit of a frustrated magnet, which is insensitive to a low concentration of defects. We have studied the effect of impurities on SrCr9pGa12-9pO19 [SCGO(p)], a highly frustrated magnet, in which the magnetic Cr3+ (s = 3/2) ions form a quasi-2D triangular system of bipyramids. Our experimental data show that as the nonmagnetic Ga3+ impurity concentration is changed, there are two distinct phases of glassiness: an exotic glassy state, which we call a spin jam, for the high magnetic concentration region () and a cluster spin glass for lower magnetic concentration (). This observation indicates that a spin jam is a unique vantage point from which the class of glassy states of dense frustrated magnets can be understood.
Understanding glassy states found in dense frustrated magnets has been an intellectual challenge since peculiar low-temperature glassy behaviors were observed experimentally in the quasi-2D SrCr9pGa12-9pO19 (SCGO) (1–3) and in the 3D pyrochlore Y2Mo2O7 (4). Immediately following, theoretical investigations (5–9) were performed to see if an intrinsic spin freezing transition is possible in a defect-free situation, aided by quantum fluctuations, as in the order-by-fluctuations phenomenon (10, 11). Quantum fluctuations at T = 0 were shown to select a long-range ordered state in the 2D kagome isotropic antiferromagnet (AFM) (5, 6), later expanded to the isotropic pyroclore and SCGO (9). Anisotropic interactions were also considered as a possible origin of the glassy kagome AFM (7). For an XY pyrochlore AFM, thermal fluctuations were found to induce a conventional Neel order (8). Experimental works were also performed to investigate if the glassy states are extrinsic due to site defects or random couplings or intrinsic to the magnetic lattice (12, 13). The consensus is that the low-temperature spin freezing transitions in SCGO(p) near the clean limit (p ≈ 1) is not driven by site defects (13).
The nature of the frozen state in SCGO has been investigated by numerous experimental techniques, including bulk susceptibility (1–3), specific heat (2, 14), muon spin relaxation (μSR) (15), nuclear magnetic resonance (NMR) (13, 16), and elastic and inelastic neutron scattering (17). Observed are spin glassy behaviors, such as field-cooled and zero-field-cooled (FC/ZFC) hysteresis in bulk susceptibility (3), as well as non-spin-glassy behaviors, such as a quadratic behavior of specific heat at low T, (14), linear dependence of the imaginary part of the dynamic susceptibility at low energies, (17), and a broad but prominent momentum dependence of the elastic neutron scattering intensity (17). The interpretation of the frozen state below Tf is still controversial. One possibility suggested was a spin liquid with unconfined spinons or resonating valence bond state, based on NMR and μSR studies (15, 16). Many-body singlet excitations were also suggested to be responsible for the behavior (14).
Recently, some of us presented an alternative scenario involving a spin jam state by considering the effects of quantum fluctuations in the disorder-free quasi-2D ideal SCGO lattice with a simple nearest neighbor (NN) spin interaction Hamiltonian (18, 19). The spin jam framework provided a qualitatively coherent understanding of all of the low-temperature behaviors such as that a complex energy landscape is responsible for the frozen state without long-range order (18), and Halperin−Saslow (HS)-like modes for the and behaviors (5, 18). In this system, which we refer to as the ideal SCGO model (iSCGO), semiclassical magnetic moments (or spins) are arranged in a triangular network of bipyramids and interact uniformly with their NN (18, 19). The microscopic mechanism for the spin jam state is purely quantum mechanical. The system has a continuous and flat manifold of ground states at the mean field level, including locally collinear, coplanar, and noncoplanar spin arrangements. Quantum fluctuations lift the classical ground state degeneracy (order by fluctuations), resulting in a complex rugged energy landscape that has a plethora of local minima consisting of the locally collinear states separated from each other by potential barriers (18). Although the work of ref. 18 dealt with a similar phase space constriction by quantum fluctuations as the aforementioned other theoretical works did, we would like to stress here the difference between the two: Whereas the other works mainly focused on the selection of the long-range-ordered (LRO) energetic ground state, the work of ref. 18 showed that the short-range-ordered (SRO) states that exist at higher energies are long-lived, dominate entropically over the LRO states, and govern the low-T physics.
The introduction of nonmagnetic impurities into a topological spin jam state breaks some of the constraints in the system, and possibly allows local transitions between minima, with a time scale dependent on the density of impurities. At a sufficiently high vacancy concentration, the system exits the spin jam state and becomes either paramagnetic or an ordinary spin glass at lower temperatures. Here we try to identify and explore the spin jam regime in an experimentally accessible system. The three most important signatures we seek for the existence of a spin jam state, different from conventional spin glass states, are (i) a linear dependence of the imaginary part of the dynamic susceptibility at low energies, , (ii) intrinsic short range static spin correlations, and (iii) insensitivity of its physics to nonmagnetic doping near the clean limit. In the rest of the paper, we provide experimental demonstration of these properties.
Experimentally, there are, so far, two materials, SrCr9pGa12-9pO19 [SCGO(p)] (1–3, 13–17, 20) and qs-ferrites like Ba2Sn2ZnGa3Cr7O22 (BSZGCO) (21), in which the magnetic Cr3+ (3d3) ion surrounded by six oxygen octahedrally, form distorted quasi-2D triangular lattice of bipyramids (20, 21) as shown in Fig. 1A, and thus may realize a spin jam state. We would like to emphasize that these systems are very good insulators (resistivity cm at 300 K) and the Cr3+ ion has no orbital degree of freedom. Furthermore, the neighboring Cr ions share one edge of oxygen octahedral, and thus the direct overlap of the orbitals of the neighboring Cr3+ ions make the AFM NN Heisenberg exchange interactions dominant and further neighbor interactions negligible (22, 23), as found in Cr2O3 (24) and ZnCr2O4 (25).
Experimental Data
We have performed elastic and inelastic neutron scattering that directly probe spin−spin correlations and bulk susceptibility measurements on SCGO(p) with various values of p over spanning almost the entire region of p. To first characterize the samples and to construct the T−p phase diagram, we have performed dc magnetic susceptibility, , and elastic neutron measurements. The data obtained from two samples with p = 0.968(6) and 0.620(8), both of which are well above the percolation threshold for the triangular lattice of bipyramids, (9), are shown in Fig. 2 A and B, respectively (the data obtained from samples with other p values are shown in Fig. S3). For both samples, exhibits similar FC and ZFC hystereses below Tf = 3.68 K [p = 0.968(6)] and 1.76 K [p = 0.620(8)] that are much lower than their large Curie−Weiss temperatures = 504 K [p = 0.968(6)] and 272 K [p = 0.620(8)] (see Fig. S1). The high frustration index indicates the presence of strong frustration in both systems. Elastic neutron scattering intensity with an instrumental energy resolution of μeV starts developing at temperatures higher than their Tf determined by , as is expected for spin freezing for measurements with different energy resolution (26). Fig. 1B summarizes the results obtained from all of the samples studied.
We observe strikingly different behaviors in the low-energy spin dynamics of the p = 0.968(6) and 0.620(8) despite the similar temperature dependences of and of elastic neutron scattering intensity. As shown in Fig. 2C, in the frozen state of p = 0.968(6), the inelastic neutron scattering intensity is weak at very low energies below 0.25 meV and gets stronger as increases. In contrast, in the frozen state of p = 0.620(8), is stronger at very low energies below 0.25 meV and gets weaker as increases (see Fig. 2D). This stark difference hints that the frozen states of p = 0.968(6) and 0.620(8) are different in nature.
To investigate more carefully the difference between the low and high doping regions, we have also performed time-of-flight neutron scattering measurements on several samples with various doping concentration from p = 0.968(6) to 0.459(5) (see Fig. S4). The data exhibit a continuum spectrum centered in the vicinity of Å−1 that corresponds to (2/3,2/3,1.8), confirming that the kagome−triangular−kagome trilayer is responsible for the low-energy dynamic spin correlations (19). This resembles the energy continuum expected for spin liquids or cooperative paramagnets; however, in contrast to a spin liquid, in our system, static spin correlations develop below Tf as well. The Q dependence of the elastic magnetic neutron scattering intensity, , was obtained with an energy window of μeV, where the subtraction of the signal above the magnetic transition eliminates background from sources not related to the transition. As shown in Fig. 3, for p = 0.968(6), very close to the clean limit, exhibits a broad peak at Å−1. Its broadness indicates the short-range nature of the static spin correlations. Surprisingly, upon increasing nonmagnetic impurity concentration up to about 20% (p ∼0.8), the shape of remains the same. Only upon further doping does become broader, and the peak position starts shifting down.
To quantitatively investigate the correlation length and peak position as a function of p, we fit to a simple Lorentzian function for each p, where HWHM is the half width at half maximum of . As a measure of the static spin correlation length scale of the frozen state, we can use . Note that the powder averaging would introduce extrinsic broadening to , and thus is underestimated, in comparison with the correlation length determined from single-crystal data, Å for SCGO(p ∼0.67) (19). However, the p dependence of serves our search for an intrinsic state.
The resulting and are shown in Fig. 3B. Remarkably, both and exhibit a flat behavior near the clean limit up to 20% doping, which is a direct evidence for existence of a distinct phase over 1 < p < 0.8 where the intrinsic short-range static correlations are independent of nonmagnetic doping. This can be naturally described as a spin jam state at the clean limit. A spin jam state can be intrinsically short range, and has an intrinsic static correlation length, . Thus, when nonmagnetic doping is low and the typical distance between the nonmagnetic impurities is larger than , the spin correlations are not affected. Only upon significant doping would the spin correlations get disturbed to make shorter than when typical distance between impurities is short.
If the p > 0.8 phase is distinct from the lower p < 0.8 phase, then the low-energy spin excitations in those two phases should have different characteristics. To see this, we show in Fig. 4 the imaginary part of the dynamic susceptibility where and is the Boltzmann factor. For p = 0.968(6), upon cooling in the cooperative paramagnetic state from 20 K to 5 K (∼ Tf ), changes from being linear to almost flat (see Fig. 4A). The nearly -independent low-energy implies that a distribution of the characteristic spin relaxation rates, , is present, which is common in glassy transitions (27). A distribution of would yield where a term, , is ignored since meV, much larger than the -range of interest. Fitting the K data to the model yields a distribution of with meV (see the red solid line in Fig. 4A).
Upon further cooling into the frozen state, however, exhibits hardening: The weight gets depleted and becomes linear at low energies (see the 1.4 K data), which is consistent with a previous neutron scattering study of SCGO[p = 0.92(5)] (17). The 1.4 K data can still be fitted to with meV, which indicates that a distribution of is still present but with the larger minimum cutoff than that of 6 K. Also, for , is linear with .
For p = 0.459(5), on the other hand, we observe a fundamentally different behavior in the frozen state: As shown in Fig. 4F, below freezing, rather than hardening, the spectral weight at low frequencies exhibits a prominent increase peaked at ∼0.2 meV. Obviously, the data cannot be reproduced by alone. Instead it behaves more like a Lorentzian with one characteristic relaxation rate, .
The clear difference in the behavior of in the frozen state between the two regimes of is another evidence that the frozen state in the vicinity of the clean limit is indeed a distinct state. To see how the evolution of the state occurs as a function of , we have fitted measured in the frozen state of each to the sum of the two contributions, . The black solid lines in Fig. 4 A−F are the fitting results of the sum, and the dotted and dashed lines are the contributions of arctan and Lorentzian, respectively. The fraction of the contribution from the single Lorentzian to , , was determined in the vicinity to the pure limit to be zero; for and for . As decreases further, i.e., as nonmagnetic doping increases, however, gradually increases. As shown in Fig. 4 C and D, the term (dotted line) dominates for , whereas the Lorentzian term (dashed line) dominates for . The over a wide range of is plotted in Fig. 3C. This confirms, upon doping, a crossover from a frozen state near the clean limit to another frozen state at high nonmagnetic concentration limit.
What is then the nature of the frozen state in the vicinity of the clean limit? The hardening and linear behavior of the low-energy spin fluctuations in the frozen state of SCGO() can be explained as HS-type modes in a spin jam (5, 28, 29). HS modes are long-range collective modes that may be viewed as an analog of the Goldstone modes associated with continuous symmetry breaking in systems without long-range order (5). In contrast, for low values of p, as defect concentration is increased, the low-energy spectral weight is eventually dominated by contributions from a distribution of local spin clusters. This is consistent with the previous specific heat data that reported the behavior robust against dilution for (14), whereas, for , the exponent starts decreasing with decreasing p (30).
As further quantitative evidence of an HS mechanism for the specific heat, may be roughly estimated for a spin jam. HS theory with a dispersion gives a contribution to specific heat , where is the gyromagnetic ratio, is the Riemann Zeta function, and we have accounted for three modes for isotropic spins. Using experimental values of electromagnetic units (emu) per mole Cr (see Fig. 2A) and using the HS upper bound estimate adapted to iSCGO, meV, and , taking from ref. 17, we get a theoretical estimate of , which is consistent with the experimental value of (14).
Discussion
The salient features observed in SCGO() are consistent with the recent understanding of the iSCGO case with spin jam (18), where and are the intralayer and interlayer Heisenberg NN couplings, respectively. For real SCGO samples, , in general, due to lattice distortion. Furthermore, there are two different s in the kagome plane: and for the bonds within each bipyramid and for the bonds of the linking triangle connecting the bipyramids. However, at the mean field level, the case of yields the same ground states as the uniform case of because the ground state of the uniform case satisfies both the AFM constraints of the linking triangle and the bipyramid. Thus, we limit our discussion to the case of . For these situations, classical magnetic ground states have been described as a function of in ref. 19. These can be obtained from the case by coherent rotations of a subset of the spins. The classical manifold of ground states remains degenerate for , as can be seen from the flat zero-energy bands that are present in the linear spin wave analysis around long-range ordered states described in Fig. S5. As the ratio changes from 1 to 0, the system moves from the ideal trilayer to decoupled kagome layers (accompanied by a layer of noninteracting spins), at which point local zero-energy modes, such as weathervane modes (7), associated with kagome physics, appear. For the isolated semiclassical kagome, an extensive configurational entropy of local minima appears with kinetic barriers associated with the weathervane motion. At low enough temperatures, tunneling is suppressed and the system freezes. On general grounds, local modes such as the weathervane will become delocalized when the layers are coupled, suggesting that for , as for , the barriers between local minima remain nonlocal in nature and thus freezing is more effective than for the decoupled kagome.
The quantum fluctuations-induced spin jam scenario is consistent with the system freezing at temperatures much lower than . An energy scale for spin fluctuations in the clean limit can be determined from the potential barrier between local (ordered) minima and is given by (18). For SCGO(), K, which is close to the experimentally determined K K.
One may consider anisotropic interactions such as the Dzyaloshinskii−Moriya (DM) as the origin for the spin freezing. We believe that a simple DM interaction will not generate such a complex energy landscape and cannot explain coherently all of the experimental data as the quantum fluctuation-induced spin jam scenario does. However, theoretical confirmation about such an anisotropic scenario is needed, which is beyond the scope of this paper.
Upon weak nonmagnetic doping, the complex energy landscape is modified and the kinetic barriers become finite; however, the overall picture remains the same. Further doping will weaken the order by fluctuations mechanism and the selection of coplanar states, and a different glassy state will emerge. Indeed, our observation of a crossover as a function of doping is consistent with Henley (9), who, remarkably, speculated on the possibility of a defect-induced crossover as a function of doping, close to the clean limit, from a nongeneric phase (dominated by coplanar states) to a generic phase in large-spin kagome and SCGO systems.
It is worthwhile to discuss here the concept of jamming in a broader context. In granular systems, a transition occurs into a special type of glassy state called jammed state with the increase of the density of constituents due to a complex energy landscape (31). In analogy, one may view the approach of the disorder-free frustrated magnets as the addition of system constituents, i.e., removal of nonmagnetic impurities, and the state of the clean system as the “spin jammed” state. In such a system, a large set of mostly irregular spin configurations form local minima, the transition between which requires the simultaneous reorientation of a large number of spins. The recent theoretical work of ref. 18 has shown how a disorder-free quasi-2D frustrated magnet with a simple NN spin interaction Hamiltonian can exhibit a spin jam of topological nature, at low temperatures. It was also established that the locally collinear states differ by collective reorientations of spins, where the smallest units of the mean field zero-energy excitations are extended along lines, called spaghetti modes. The extended nature of the reorientations is responsible for particularly large dynamical barriers, of topological nature, to tunneling between minima (18). The effect of an order by fluctuations may also be a possible mechanism for forming a glassy state in kagome systems (5, 6, 32, 33), with less robust nature than in the iSCGO (9), as the presence of local tunneling through weathervane modes involving six spins may hasten approach to a global ground state.
In comparison, surprisingly, the 3D pyrochlore Y2Mo2O7 also exhibits a similar behavior at low temperatures but with D less than the value of 3 expected for a 3D system (34). The magnetic Mo4+ (4d2) ions form a 3D network of corner-sharing tetrahedra. If these magnetic moments are isotropic, and antiferromagnetically and uniformly interacting with their NN only, the system is supposed to yield the highest degree of frustration. One may speculate that the freezing in Y2Mo2O7 may also be explained in terms of a spin jam; however, we would like to point out crucial differences between the systems. In particular, the frustration index of Y2Mo2O7 is two orders of magnitude smaller than in SCGO; . This can be understood by the facts that, unlike SCGO, which is an excellent insulator, Y2Mo2O7 is semiconducting ( cm at 300 K), and the neighboring Mo ions share one corner of oxygen octahedra, which tends to result in nonnegligble longer-range magnetic interactions (22). More importantly, the magnetic Mo4+ () ions are orbitally degenerate (34, 35); it is well known that orbital degeneracy has a great tendency to modify the nature of a magnetic network, and, as a result, it reduces dimensionality of the magnetic interactions and frustration as well, as found in ZnV2O4 (36). Therefore, the spin glassy state of Y2Mo2O7 may be due to spatially random coupling constants induced by the orbital degrees of freedom rather than strong frustration. The effectively reduced dimensionality may be related to the observed with D < 3. A previous muon spin relaxation study (37) showed that the spin glassy state of Y2Mo2O7 seems to remain intact at 20% of nonmagnetic doping as in SCGO, which may indicate that the jamming physics is still relevant to Y2Mo2O7 as well. We believe that the concept of jamming and complex energy landscape can unify the seemingly different glassy states found in dense magnetic systems, such as SCGO, Y2Mo2O7 (30), spin ice (38), and even the dynamics of magnetic domain boundaries (39).
Conclusion
The search for glassiness that arises intrinsically without defects and randomness has been revived recently as such glassiness may bear intricate relations with topological order (40), lack of thermalization in many body localization (41), jamming in structural glasses (42), and glassiness in supercooled liquids (43). Our experiments indicate that quantum fluctuations, via an order by fluctuations mechanism, induce a glassy state, a spin jam, in the strongly frustrated SCGO(), which is robust, and extends to the clean limit. The findings strongly support the possible existence of purely topological glassy states.
SI Text
Methods and Materials.
Polycrystalline samples of SCGO(p) with 12 different values of p = 0.968(6), 0.960(9), 0.925(nominal), 0.917(9), 0.87(nominal), 0.844(8), 0.777(6), 0.736(6), 0.620(8), 0.5(nominal), 0.459(5), and 0.228(5) were made using standard solid-state synthesis techniques at the University of Virginia. Stoichiometric quantities of SrCO3, Ga2O3, and Cr2O3 were mixed, ground, and pelletized. Then, the pellets were heated in air in dense Al2O3 crucibles at 1400 °C for 96 h, with three intermediate grindings. The samples were then annealed at 1,000 °C for 7 d to promote cation ordering. The dc magnetic susceptibilities were measured using the Quantum design MPMS-XL7 with a Helium-3 insert in the temperature range from 0.5 K to 14 K at Eidgenössische Technische Hochschule (ETH) Zurich, and similar measurements were done from 2 K to 300 K at Kyoto University. Crystal structure and the actual Cr concentrations were determined by neutron diffraction. Neutron powder diffraction measurements were performed using BT1 diffractometer at the National Institute for Standards and Technology Center for Neutron Research (NCNR) located in Gaithersburg, MD, with the neutron wavelength of Å at room temperature for SCGO(p) samples with p = 0.968(6), 0.960(9), 0.917(9), 0.844(8), 0.777(6), 0.736(6), 0.620(8), 0.459(5), and 0.228(5).
Time-of-flight neutron scattering measurements were performed using the Disk Chopper Spectrometer (DCS) at NCNR with Å. SCGO(p) samples with p = 0.968(6), 0.960(9), 0.917(9), 0.844(8), 0.777(6), 0.736(6), and 0.620(8) were measured in a standard liquid Helium-4 cryostat, and p = 0.459(5) was measured with a Helium-3 insert. The low-energy inelastic magnetic neutron scattering spectrum was measured at various temperatures, T, around Tf for the samples. For each SCGO(p) sample, an energy-dependent background was determined by imposing the detailed balance constraint where is the Boltzmann constant, and it was subtracted from the data to obtain the magnetic contribution. Then, the imaginary part of the dynamic susceptibility was derived from the fluctuation dissipation theorem, .
Bulk Magnetic Susceptibility and Curie−Weiss Fitting.
Fig. S1 shows the reciprocal susceptibility as a function of temperature (1/χ vs. T) measured in a 0.1 Tesla field over a temperature range of 2–300 K. The Curie−Weiss temperatures (ΘCW) were extracted by fitting the 1/χ - T curves at T > 150 K with Curie−Weiss law [χ = C/(T − ΘCW)]. The ΘCW values range from −72 to −504 K, reflecting the strong antiferromagnetic interactions in the system. As shown in Fig. S1, Inset, |ΘCW| decreases gradually with decreasing Cr concentration from p = 0.960(9) to 0.228(5).
Neutron Powder Diffraction Refinement.
Fig. S2 shows the neutron powder diffraction data of SCGO [p = 0.968(6), 0.844(8), 0.620(8) and 0.495(5)]. The lines are the Rietveld refinement results of nuclear structure performed using General Structure Analysis System (GSAS) (38). The optimal parameters for their crystal structures are listed in Tables S1−S4 and are consistent with the previously published structure (11, 13).
Table S1.
Site | x | y | z | n, f.u. | , Å |
Sr 2d | 2/3 | 1/3 | 1/4 | 1 | 0.0086(8) |
Ga 2a | 0 | 0 | 0 | 0.036(18) | 0.0041(4) |
Cr 2a | 0 | 0 | 0 | 0.964(18) | 0.0041(4) |
Ga 2b | 0 | 0 | 1/4 | 1 | 0.0128(7) |
Ga 4fiv | 1/3 | 2/3 | 0.0275(1) | 2 | 0.0062(4) |
Ga 4fvi | 1/3 | 2/3 | 0.1911(2) | 0.056(30) | 0.0041(4) |
Cr 4fvi | 1/3 | 2/3 | 0.1911(2) | 1.944(30) | 0.0041(4) |
Ga 12k | 0.1685(3) | 0.3370(6) | −0.1081(1) | 0.192(60) | 0.0041(4) |
Cr 12k | 0.1685(3) | 0.3370(6) | −0.1081(1) | 5.808(60) | 0.0041(4) |
O 4e | 0 | 0 | 0.1520(2) | 2 | 0.0080(7) |
O 4f | 1/3 | 2/3 | −0.0546(2) | 2 | 0.0019(5) |
O 6h | 0.1813(3) | 0.3635(6) | 1/4 | 3 | 0.0086(4) |
O 12k | 0.1556(2) | 0.3112(4) | 0.0529(1) | 6 | 0.0037(3) |
O 12k | 0.5055(2) | 0.0111(4) | 0.1507(1) | 6 | 0.0059(3) |
Positions within space group P63/mmc and occupancies per formula unit (f.u.) of atoms in SCGO(p) at T = 298 K as determined by Rietveld analysis of the data shown in Fig. S2 using the GSAS. The lattice parameters are a = 5.7948(3) Å and c = 22.6625(2) Å. Isotropic Debye−Waller factors, , were used where is the mean squared displacement. The resulting overall reduced .
Table S4.
Site | x | y | z | n, f.u. | , Å |
Sr 2d | 2/3 | 1/3 | 1/4 | 1 | 0.0108(8) |
Ga 2a | 0 | 0 | 0 | 0.431(16) | 0.0036(3) |
Cr 2a | 0 | 0 | 0 | 0.569(16) | 0.0036(3) |
Ga 2b | 0 | 0 | 1/4 | 1 | 0.0159(7) |
Ga 4fiv | 1/3 | 2/3 | 0.0276(1) | 2 | 0.0050(4) |
Ga 4fvi | 1/3 | 2/3 | 0.1897(1) | 1.134(13) | 0.0037(3) |
Cr 4fvi | 1/3 | 2/3 | 0.1897(1) | 0.866(13) | 0.0037(3) |
Ga 12k | 0.1683(2) | 0.3366(4) | 0.1094(1) | 3.306(30) | 0.0037(3) |
Cr 12k | 0.1683(2) | 0.3366(4) | 0.1094(1) | 2.694(30) | 0.0037(3) |
O 4e | 0 | 0 | 0.1501(1) | 2 | 0.0097(6) |
O 4f | 1/3 | 2/3 | 0.0545(1) | 2 | 0.0024(5) |
O 6h | 0.1812(3) | 0.3623(5) | 1/4 | 3 | 0.0093(5) |
O 12k | 0.1554(2) | 0.3106(3) | 0.0526(7) | 6 | 0.0054(3) |
O 12k | 0.5047(2) | 0.0094(4) | 0.1500(1) | 6 | 0.0061(3) |
Positions within space group P63/mmc and occupancies per formula unit (f.u.) of atoms in SCGO(p) at T = 298 K as determined by Rietveld analysis of the data shown in Fig. S2 using the GSAS. The lattice parameters are a = 5.79566(3) Å and c = 22.7778(2) Å. Isotropic Debye−Waller factors, , were used where is the mean squared displacement. The resulting overall reduced .
Table S2.
Site | x | y | z | n, f.u. | , Å |
Sr 2d | 2/3 | 1/3 | 1/4 | 1 | 0.0098(7) |
Ga 2a | 0 | 0 | 0 | 0.131(17) | 0.0045(3) |
Cr 2a | 0 | 0 | 0 | 0.869(17) | 0.0045(3) |
Ga 2b | 0 | 0 | 1/4 | 1 | 0.0141(7) |
Ga 4fiv | 1/3 | 2/3 | 0.0272(1) | 2 | 0.0054(4) |
Ga 4fvi | 1/3 | 2/3 | 0.1906(2) | 0.321(28) | 0.0044(4) |
Cr 4fvi | 1/3 | 2/3 | 0.1906(2) | 1.679(28) | 0.0044(4) |
Ga 12k | 0.1686(3) | 0.3374(5) | −0.1089(1) | 0.948(54) | 0.0045(4) |
Cr 12k | 0.1686(3) | 0.3374(7) | −0.1089(1) | 5.052(54) | 0.0045(4) |
O 4e | 0 | 0 | 0.1514(1) | 2 | 0.0096(6) |
O 4f | 1/3 | 2/3 | −0.0551(1) | 2 | 0.0019(5) |
O 6h | 0.1809(2) | 0.3618(5) | 1/4 | 3 | 0.0094(5) |
O 12k | 0.1555(2) | 0.3109(3) | 0.0525(1) | 6 | 0.0043(3) |
O 12k | 0.5054(2) | 0.0108(4) | 0.1504(1) | 6 | 0.0049(3) |
Positions within space group P63/mmc and occupancies per formula unit (f.u.) of atoms in SCGO(p) at T = 298 K as determined by Rietveld analysis of the data shown in Fig. S2 using the GSAS. The lattice parameters are a = 5.79447(3) Å and c = 22.7120(2) Å. Isotropic Debye−Waller factors, , were used where is the mean squared displacement. The resulting overall reduced .
Table S3.
Site | x | y | z | n, f.u. | , Å |
Sr 2d | 2/3 | 1/3 | 1/4 | 1 | 0.0107(7) |
Ga 2a | 0 | 0 | 0 | 0.298(17) | 0.0047(3) |
Cr 2a | 0 | 0 | 0 | 0.702(17) | 0.0047(3) |
Ga 2b | 0 | 0 | 1/4 | 1 | 0.0159(7) |
Ga 4fiv | 1/3 | 2/3 | 0.0276(1) | 2 | 0.0044(4) |
Ga 4fvi | 1/3 | 2/3 | 0.1905(1) | 0.790(28) | 0.0047(3) |
Cr 4fvi | 1/3 | 2/3 | 0.1905(1) | 1.210(28) | 0.0047(3) |
Ga 12k | 0.1682(2) | 0.3366(4) | 0.1093(1) | 2.328(66) | 0.0047(3) |
Cr 12k | 0.1682(2) | 0.3366(4) | 0.1093(1) | 3.672(66) | 0.0047(3) |
O 4e | 0 | 0 | 0.1504(1) | 2 | 0.0102(6) |
O 4f | 1/3 | 2/3 | 0.0547(1) | 2 | 0.0019(5) |
O 6h | 0.1811(3) | 0.3622(5) | 1/4 | 3 | 0.0089(5) |
O 12k | 0.1556(2) | 0.3112(3) | 0.0526(1) | 6 | 0.0046(3) |
O 12k | 0.5047(2) | 0.0094(4) | 0.1502(1) | 6 | 0.0051(3) |
Positions within space group P63/mmc and occupancies per formula unit (f.u.) of atoms in SCGO(p) at T = 298 K as determined by Rietveld analysis of the data shown in Fig. S2 using the GSAS. The lattice parameters are a = 5.79676(3) Å and c = 22.7541(2) Å. Isotropic Debye−Waller factors, , were used where is the mean squared displacement. The resulting overall reduced .
Low-Temperature Magnetic Susceptibility and Elastic Neutron Intensity.
The low-temperature magnetic susceptibilities were measured in a field of 0.01 Tesla in the temperature range from 0.5 K to 14 K, as shown in Fig. S3 (blue squares). The susceptibility increases with decreasing temperature, exhibiting a low-temperature upturn. With further decreasing temperature, a divergence between the filed cooled (FC) and ZFC susceptibility, which is the typical behavior of spin glass transition, was observed for all six samples. The freezing temperature Tf was determined as the temperature of the maximum susceptibility for the ZFC curve, where the divergence between FC and ZFC susceptibility also occurs. Tf decreases with decreasing Cr concentration. Above Tf, no difference was observed between FC and ZFC susceptibilities for all of the samples.
The wave vector averaged magnetic elastic neutron scattering intensities normalized to the maximum intensity as a function of temperature are also shown in Fig. S3 (open circles). Neutron scattering data were obtained using the cold neutron DCS at the NCNR with λ = 6.0 Å. The temperature of the onset of magnetic neutron intensity is higher than the freezing temperature observed in bulk magnetic susceptibility data, and decreases with decreasing Cr concentrations, exhibiting a similar behavior to the one observed in the susceptibility data.
Contour Maps of Inelastic Neutron Scattering Data.
Fig. S4 shows the low-energy inelastic magnetic neutron scattering spectrum measured in the frozen states of several different SCGO(p) samples. In the frozen state of p = 0.968(6), the inelastic neutron scattering intensity is weak at very low energies below 0.25 meV and gets stronger as increases. In contrast, in the frozen state of p = 0.459(5), is strong at very low energies below 0.25 meV and gets weaker as increases. This stark difference hints that the frozen states of p = 0.968(6) and 0.459(5) are different in nature. The change of at very low energies below 0.25 meV is gradual with decreasing p and manifest when p = 0.777(6), indicating that a transition to a different frozen state occurs between p = 0.776(6) and 0.844(8).
Spin Wave Dispersion.
The spin wave calculations were done for the SCGO at the clean limit, p = 1. This was done by expanding around classically ordered ground state configuration to harmonic order, using the standard Holstein−Primakoff method (39).
The chosen states are coplanar spin configurations that are described in Fig. S5, Inset. The spin wave dispersion is plotted along the (H,H,1.8) momentum direction for various values of (5), going from for the ideal SCGO to for isolated kagome. The spin wave analysis shows a flat band of nine zero modes, indicating a large classical degeneracy.
Acknowledgments
J.Y. and S.H.L. thank Drs. Matthias Thede and Andrey Zheludev for their help during some of our bulk susceptibility measurements performed at Eidgenössische Technische Hochschule (ETH) Zurich. The work at University of Virginia was supported in part by the US National Science Foundation (NSF) Grant DMR-1404994 (to S.H.L.) and the NSF CAREER Grant DMR-0956053 (to I.K.). This work used facilities supported in part by the NSF under Agreement DMR-0944772, and was supported in part by NSF Grant PHYS-1066293 and the hospitality of the Aspen Center for Physics.
Footnotes
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1503126112/-/DCSupplemental.
References
- 1.Obradors X, et al. Magnetic frustration and lattice dimensionality in SrCr8Ga4O19. Solid State Commun. 1988;65(3):189–192. [Google Scholar]
- 2.Ramirez AP, Espinosa GP, Cooper AS. Strong frustration and dilution-enhanced order in a quasi-2D spin glass. Phys Rev Lett. 1990;64(17):2070–2073. doi: 10.1103/PhysRevLett.64.2070. [DOI] [PubMed] [Google Scholar]
- 3.Martínez B, et al. Magnetic dilution in the strongly frustrated kagome antiferromagnet SrGa12-xCrxO19. Phys Rev B Condens Matter. 1992;46(17):10786–10792. doi: 10.1103/physrevb.46.10786. [DOI] [PubMed] [Google Scholar]
- 4.Gingras MJP, Stager CV, Raju MP, Gaulin BD, Greedan JE. Static critical behavior of the spin-freezing transition in the geometrically frustrated pyrochlore antiferromagnet Y2Mo2O7. Phys Rev Lett. 1997;78(5):947–950. [Google Scholar]
- 5.Sachdev S. Kagomé- and triangular-lattice Heisenberg antiferromagnets: Ordering from quantum fluctuations and quantum-disordered ground states with unconfined bosonic spinons. Phys Rev B Condens Matter. 1992;45(21):12377–12396. doi: 10.1103/physrevb.45.12377. [DOI] [PubMed] [Google Scholar]
- 6.Chubukov A. Order from disorder in a kagomé antiferromagnet. Phys Rev Lett. 1992;69(5):832–835. doi: 10.1103/PhysRevLett.69.832. [DOI] [PubMed] [Google Scholar]
- 7.Chandra P, Coleman P, Ritchey I. The anisotropic kagome antiferromagnet: A topological spin glass? J Phys I France. 1993;3(2):591–610. [Google Scholar]
- 8.Bramwell ST, Gingras MJP, Reimers JN. Order by disorder in an anisotropic pyrochlore lattice antiferromagnet. J Appl Phys. 1994;75(10):5523–5525. [Google Scholar]
- 9.Henley CL. Effective Hamiltonians and dilution effects in kagome and related anti-ferromagnets. Can J Phys. 2001;79(11-12):1307–1321. [Google Scholar]
- 10.Villain J. Spin glass with non-random interactions. J Phys C Solid State Phys. 1977;10(10):1717–1734. [Google Scholar]
- 11.Henley CL. Ordering due to disorder in a frustrated vector antiferromagnet. Phys Rev Lett. 1989;62(17):2056–2059. doi: 10.1103/PhysRevLett.62.2056. [DOI] [PubMed] [Google Scholar]
- 12.Ramirez AP. Strongly geometrically frustrated magnets. Annu Rev Mater Sci. 1994;24:453–480. [Google Scholar]
- 13.Limot L, et al. Susceptibility and dilution effects of the kagome bilayer geometrically frustrated network: A Ga NMR study of SrCr9pGa12-9pO19. Phys Rev B. 2002;65:144447. [Google Scholar]
- 14.Ramirez AP, Hessen B, Winklemann M. Entropy balance and evidence for local spin singlets in a Kagome-like magnet. Phys Rev Lett. 2000;84(13):2957–2960. doi: 10.1103/PhysRevLett.84.2957. [DOI] [PubMed] [Google Scholar]
- 15.Keren A, et al. Magnetic dilution in the geometrically frustrated SrCr9pGa12-9pO19 and the role of local dynamics: A muon spin relaxation study. Phys Rev Lett. 2000;84(15):3450–3453. doi: 10.1103/PhysRevLett.84.3450. [DOI] [PubMed] [Google Scholar]
- 16.Bono D, Limot L, Mendels P, Collin G, Blanchard N. Correlations, spin dynamics, defects: The highly frustrated kagome bilayer. Low Temp Phys. 2005;31(8):704–721. [Google Scholar]
- 17.Lee SH, et al. Spin-glass and non–spin-glass features of a geometrically frustrated magnet. Europhys Lett. 1996;35(2):127–132. [Google Scholar]
- 18.Klich I, Lee SH, Iida K. Glassiness and exotic entropy scaling induced by quantum fluctuations in a disorder-free frustrated magnet. Nat Commun. 2014;5:3497. doi: 10.1038/ncomms4497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Iida K, Lee SH, Cheong SW. 2012. Coexisting order and disorder hidden in a quasi-two-dimensional frustrated magnet. Phys Rev Lett 108(21):217207, and erratum (2014) 113(7):079901. [DOI] [PubMed]
- 20.Lee S, et al. Isolated spin pairs and two dimensional magnetism in SrCr9pGa12-9pO19. Phys Rev Lett. 1996;76(23):4424–4427. doi: 10.1103/PhysRevLett.76.4424. [DOI] [PubMed] [Google Scholar]
- 21.Hagemann IS, Huang Q, Gao XPA, Ramirez AP, Cava RJ. Geometric magnetic frustration in Ba2Sn2Ga3ZnCr7O22: A two-dimensional spinel based Kagomé lattice. Phys Rev Lett. 2001;86(5):894–897. doi: 10.1103/PhysRevLett.86.894. [DOI] [PubMed] [Google Scholar]
- 22.Goodenough JB. Direct cation−cation interactions in several oxides. Phys Rev. 1960;117(6):1442–1451. [Google Scholar]
- 23.Motida K, Miahara S. On the 90° exchange interaction between cations (Cr3+, Mn2+, Fe3+ and Ni2+) in oxides. J Phys Soc Jpn. 1970;28(5):1188–1196. [Google Scholar]
- 24.Samuelson EJ, Hutchings MT, Shirane G. Inelastic neutron scattering investigation of spin waves and magnetic interactions in Cr2O3. Physica. 1970;48(15):13–42. [Google Scholar]
- 25.Lee S, Broholm C, Kim TH, Ratcliff W, 2nd, Cheong SW. Local spin resonance and spin-Peierls-like phase transition in a geometrically frustrated antiferromagnet. Phys Rev Lett. 2000;84(16):3718–3721. doi: 10.1103/PhysRevLett.84.3718. [DOI] [PubMed] [Google Scholar]
- 26.Murani AP. Spectral distribution of relaxation times in spin glasses. J Magn Magn Mater. 1981;22(3):271–281. [Google Scholar]
- 27.Mydosh JA. Spin Glasses: An Experimental Introduction. Taylor and Francis; London: 1993. [Google Scholar]
- 28.Halperin BI, Saslow WM. Hydrodynamic theory of spin waves in spin glasses and other systems with noncollinear spin orientations. Phys Rev B. 1977;16(5):2154–2162. [Google Scholar]
- 29.Podolsky D, Kim YB. Halperin-Saslow modes as the origin of the low-temperature anomaly in NiGa2S4. Phys Rev B. 2009;79(14):140402. [Google Scholar]
- 30.Ramirez AP, Espinosa GP, Cooper AS. Elementary excitations in a diluted antiferromagnetic Kagomé lattice. Phys Rev B Condens Matter. 1992;45(5):2505–2508. doi: 10.1103/physrevb.45.2505. [DOI] [PubMed] [Google Scholar]
- 31.Liu AJ, Nagel SR. Nonlinear dynamics: Jamming is not just cool any more. Nature. 1998;396(6706):21–22. [Google Scholar]
- 32.Chalker JT, Holdsworth PCW, Shender EF. Hidden order in a frustrated system: Properties of the Heisenberg Kagomé antiferromagnet. Phys Rev Lett. 1992;68(6):855–858. doi: 10.1103/PhysRevLett.68.855. [DOI] [PubMed] [Google Scholar]
- 33.Chernyshev AL, Zhitomirsky ME. Quantum selection of order in an XXZ antiferromagnet on a Kagome lattice. Phys Rev Lett. 2014;113(23):237202. doi: 10.1103/PhysRevLett.113.237202. [DOI] [PubMed] [Google Scholar]
- 34.Silverstein HJ, et al. Liquidlike correlations in single-crystalline Y2Mo2O7: An unconventional spin glass. Phys Rev B. 2014;89(5):054433. [Google Scholar]
- 35.Shinaoka H, Motome Y, Miyake T, Ishibashi S. Spin-orbital frustration in molybdenum pyrochlores A2Mo2O7 (A= rare earth) Phys Rev B. 2013;88(17):174422. [Google Scholar]
- 36.Lee SH, et al. Orbital and spin chains in ZnV2O4. Phys Rev Lett. 2004;93(15):156407. doi: 10.1103/PhysRevLett.93.156407. [DOI] [PubMed] [Google Scholar]
- 37.Dunsiger SR, et al. Low temperature spin dynamics of geometrically frustrated antiferromagnets Y2Mo2O7 and Y2Mo1.6Ti0.4O7 studied by muon spin relaxation. J Appl Phys. 1996;79(8):6636–6638. [Google Scholar]
- 38.Gingras MJP. In: Introduction to Frustrated Magnetism: Materials, Experiments, Theory. Lacroix C, Mendels P, Mila F, editors. Springer; New York: 2011. pp. 293–329. [Google Scholar]
- 39.Chen SW, et al. Jamming behavior of domains in a spiral antiferromagnetic system. Phys Rev Lett. 2013;110(21):217201. doi: 10.1103/PhysRevLett.110.217201. [DOI] [PubMed] [Google Scholar]
- 40.Chamon C. Quantum glassiness in strongly correlated clean systems: An example of topological overprotection. Phys Rev Lett. 2005;94(4):040402. doi: 10.1103/PhysRevLett.94.040402. [DOI] [PubMed] [Google Scholar]
- 41.De Roeck W, Huveneers F. Asymptotic quantum many-body localization from thermal disorder. Commun Math Phys. 2014;332(3):1017–1082. [Google Scholar]
- 42.Goodrich CP, Liu AJ, Nagel SR. Solids between the mechanical extremes of order and disorder. Nat Phys. 2014;10(8):578–581. [Google Scholar]
- 43.Tarjus G, Kivelson SA, Nussinov Z, Viot P. The frustration-based approach of supercooled liquids and the glass transition: A review and critical assessment. J Phys Condens Matter. 2005;17(50):R1143–R1182. [Google Scholar]