Skip to main content
Science Advances logoLink to Science Advances
. 2024 Oct 11;10(41):eadp4685. doi: 10.1126/sciadv.adp4685

A pyroxene-based quantum magnet with multiple magnetization plateaus

Lun Jin 1,*,, Shiyu Peng 2,, Aya Nakano Rutherford 3, Xianghan Xu 1, Danrui Ni 1, Chen Yang 1, Yen Ji Byeon 1, Weiwei Xie 4, Haidong Zhou 3, Xi Dai 2, Robert J Cava 1,*
PMCID: PMC11468919  PMID: 39392885

Abstract

Pyroxenes (AMX2O6) consisting of infinite one-dimensional edge-sharing MO6 chains and bridging XO4 tetrahedra are fertile ground for finding quantum materials. Thus, here, we have studied calcium cobalt germanate (CaCoGe2O6) and calcium cobalt silicate (CaCoSi2O6) crystals in depth. Heat capacity data show that the spins in both compounds are dominantly Ising-like, even after being manipulated by high magnetic fields. On cooling below the Néel temperatures, a sharp field–induced transition in magnetization is observed for CaCoGe2O6, while multiple magnetization plateaus beneath the full saturation moment are spotted for CaCoSi2O6. Our analysis shows that these contrasting behaviors potentially arise from the different electron configurations of germanium and silicon, in which the 3d orbitals are filled in the former but empty in the latter, enabling electron hopping. Thus, silicate tetrahedra can aid the interchain superexchange pathway between cobalt(II) ion centers, while germanate ones tend to block it during magnetization.


The observation of anisotropic field–induced magnetization plateaus in pyroxenes depends on the choice of bridging tetrahedra.

INTRODUCTION

Solid-state materials are one of the fundamental building blocks of modern society. Many modern technologies have changed in almost unimaginable ways in the past few decades, largely attributed to the discovery of solid-state materials providing sophisticated functionalities (18). Pyroxenes, with a general formula AMX2O6, are one of the most abundant minerals in Earth’s crust and are well studied in geoscience (912). The edge-sharing MO6 octahedra form infinite one-dimensional (1D) chains that propagate in a zigzag fashion, and the XO4 tetrahedra bridge adjacent 1D chains to form the structural motif of the lattice. The edge-sharing MO6 octahedra within each 1D chain naturally lead to shorter distances between adjacent metal centers within the chains. These metal centers communicate with metal centers in the neighboring chains through the bridging tetrahedra. This unique structural motif can have a marked impact on the physical properties. The fruitful choice of elements residing in the centers of the octahedra and tetrahedra provides degrees of freedom for manipulation of these properties. Therefore, this class of materials has attracted enduring interest from the science community because a wide variety of intriguing physical phenomena have been observed, including multiferroicity, magnetoelectric effects, etc. (1317), and the competing interchain and intrachain interactions within the highly versatile crystal structure of pyroxenes make them a fertile ground for finding quantum materials.

The electronic structure and magnetic properties of pyroxenes containing transition metals have been studied as low-dimensional magnets with 90° bonds via theoretical calculations (18). Despite the fact that an enormous number of studies have focused on pyroxenes bearing the magnetic transition metals Cr, Mn, and Fe (1317), the analogous Co-containing compounds are incompletely studied (1925). The Co-containing pyroxenes have been reported to serve as pigments (26, 27) or functional ceramics due to their thermal (28) and dielectric properties (29, 30) and, very recently, have been reported to serve as a playground for Kitaev physics and Ising model realization (31). We here show that on cooling through the Néel temperature TN of CaCoGe2O6 and CaCoSi2O6 in powder form, a sharp field–induced transition in magnetization is observed for CaCoGe2O6, while multiple magnetization plateaus beneath the full saturation moment are spotted for CaCoSi2O6 at reasonable applied fields. A series of CaCoGe2−xSixO6 (x = 0, 0.5, 1, 1.5, and 2) solid solutions has been synthesized as powders to show the continuous evolution of the observed phenomena in magnetization. To study these phenomena further, we grew macroscopic crystals of CaCoGe2O6 and CaCoSi2O6 by the floating-zone method. We found that the observed magnetization transitions are highly anisotropic, as expected from the crystal structure. Heat capacity data show that the spins in both compounds are dominantly Ising-like, even after being manipulated by high magnetic fields, which is unusual for Co2+ in extended lattices (32, 33). Beyond revealing the enormous orbital contribution to magnetism, our first-principles calculation results further justify the observed discrepancy in anisotropic magnetic saturation between two compounds, which stems from different crystal field hybridizations of Co 3d orbitals.

Magnetization plateaus pinned at fractional values of the full saturation moment during the magnetization of quantum magnets are intriguing and often induced by collective effects (20, 34). The findings in this study ultimately reveal that the choice of bridging tetrahedra (i.e., SiO4 versus GeO4) between adjacent 1D CoO6 chains in a pyroxene-based quantum magnet can locate magnetization plateaus in a measurable field range, potentially attributed to the empty 3d orbitals of Si that enable electron hopping. Hence, this work not only promotes pyroxenes as an excellent platform to study quasi-1D spin chain systems with nontrivial interchain coupling but also opens more research opportunities to further study those observed anisotropic plateaus during the magnetization process.

RESULTS

Structural characterization

The crystal structures of the entire series of magenta CaCoGe2−xSixO6 (x = 0, 0.5, 1, 1.5, and 2) compounds can be indexed by a monoclinic unit cell (space group C2/c), in good agreement with previous reports on the two end members of this series, CaCoGe2O6 (25) and CaCoSi2O6 (23). The lattice parameters extracted from the Le Bail fitting of the acquired powder x-ray diffraction (p-XRD) patterns at 300 K change continuously, confirming the purity of the as-made polycrystalline powder samples (fig. S1). In general, the unit cell gradually shrinks upon Si-for-Ge substitution as expected, since Ge4+ (0.53 Å) has a larger effective ionic radius than Si4+ (0.40 Å) (35). The end members CaCoGe2O6 and CaCoSi2O6 are of more interest in the present study, and, thus, macroscopic (i.e., a few millimeters on a side) single crystals suitable for anisotropic physical property measurements have been grown by the floating-zone method (fig. S2). Our single-crystal XRD data were collected from small pieces of each compound at ambient temperature. The refined structural parameters are perfectly aligned with those extracted from the abovementioned p-XRD data, as well as the literature (23, 25). Detailed crystallographic data, fitting statistics (table S1), atomic coordinates and equivalent isotropic atomic displacement parameters (table S2), and bond lengths (table S3) are presented in the Supplementary Materials.

The crystal structure is depicted in Fig. 1 (A to C). The 1D Co2+ (3d7) chains run parallel to c. The Ge and Si reside in the XO4 tetrahedra. These tetrahedra bridge adjacent infinite edge-sharing CoO6 1D chains to form the monoclinic pyroxene crystal structure. The interchain Co-Co distance is dominated by the size of these bridging tetrahedra, with the smallest value found in the all-Si case (d ≈ 6.635 Å) and the largest value found in the all-Ge case (d ≈ 6.772 Å). The intrachain Co-Co distance is less influenced by the size of bridging tetrahedron and only shows a slight contraction in the all-Si case (about 3.101 Å) compared to the all-Ge case (about 3.155 Å). To make sure that all the isotropic properties measured in the present work are reliable, we collected p-XRD data from the crushed and ground crystals of CaCoGe2O6 and CaCoSi2O6. The powder data were then refined against the unit cell we got from the single-crystal XRD results (Fig. 1, D and E).

Fig. 1. Crystal structure.

Fig. 1.

(A) Extended monoclinic unit cell of the CaCoGe2−xSixO6 compounds. (B) Infinite edge-sharing CoO6 1D chains run parallel to the c axis, separated by (Ge,Si)O4 tetrahedra. (C) The zigzag nature of the edge-sharing CoO6 1D chain. Le Bail fitting of the p-XRD data collected from crushed and ground crystals of (D) CaCoGe2O6 and (E) CaCoSi2O6. Insets: Pictures of single crystals on millimeter grids, with the flat crystal surface normal to the [010] axis. a.u., arbitrary units.

Isotropic magnetic properties

The isotropic zero field–cooled (ZFC) temperature-dependent dc magnetization data were collected from the crushed and ground crystals of CaCoGe2O6 and CaCoSi2O6, under an applied field of H = 0.1 T, and are plotted as magnetic susceptibility χ against temperature T (Fig. 2, A and B). The dc magnetic susceptibility data, over a suitable temperature range (selected as the straight line part of the 1/χ versus T curves, marked in red in Fig. 2, A and B) for each composition, were fitted to the Curie-Weiss law [χ = C/(T − θ) + χ0], to yield the Curie constants C, Weiss temperatures θ, and effective moments μeff. The Curie constants C for both compounds show a marginal difference [Ge, 4.040 (2) cm3·K/mol; Si, 4.002 (4) cm3·K/mol], and, hence, the value of effective moment μeff can be calculated [Ge, 5.69 μB/formula unit (f.u.); Si, 5.66 μB/f.u.]. Both values are larger than the spin-only expectation for S = 3/2 Co2+ (high-spin d7, 3.87 μB/f.u.), revealing that spin-orbit coupling (SOC) makes an enormous contribution to magnetism in both cases, which is expected (36). The Weiss temperatures θ [Ge, −21.2 (1) K; Si, −9.6 (2) K] were also determined. The negative sign of θ suggests that the dominant magnetic coupling between spins is antiferromagnetic (AFM), agreeing with the well-defined paramagnetic-to-AFM transition observed in both cases. Further, the magnitude of θ almost coincides with the Néel temperature TN (the local maximum in the magnetic susceptibility curve: Ge, 19.8 K; Si, 10.1 K) for both compounds. The Curie-Weiss fitting results of our isotropic magnetic susceptibility data in general agree well with the literature, with the only anomaly being the sign of θ for CaCoGe2O6, which can potentially be attributed to differences in sample preparation and fitted temperature range (23, 25). The analogous temperature-dependent data collected from the polycrystalline powder show that the migration of TN to a lower temperature is a continuous process throughout the whole range of the solid solution (Fig. 2C).

Fig. 2. Isotropic magnetization.

Fig. 2.

The temperature-dependent and field-dependent dc magnetization data collected from (A, B, D, and E) the crushed and ground crystals of CaCoGe2O6 and CaCoSi2O6, as well as (C and F) the polycrystalline powder form of the CaCoGe2−xSixO6 (x = 0, 0.5, 1, 1.5, and 2) series.

The field-dependent magnetization data were collected from the crushed and ground crystals of CaCoGe2O6 and CaCoSi2O6. At temperatures higher than TN, the magnetization (M) versus field (H) curves are linear and pass through the origin, while upon cooling to below TN, they exhibit features that indicate either field-induced spin-state transitions or spin reorientations (Fig. 2, D and E). For CaCoGe2O6, at T = 1.8 K, the M versus H curve shows a sharp kink under an external applied field of around 5.5 T. The isotherm shows no signs of saturation up to H = 9 T, and the largest magnetic moment achieved is approximately 2 μB/f.u. (Fig. 2D). In contrast, CaCoSi2O6 has a subtle magnetization plateau observed in the low-field regime. At T = 1.8 K, this subtle magnetization plateau arises when a relatively small field (0 ≤ H/T ≤ 1.2) is applied, and then a flattened saturation plateau at about 2.5 μB/f.u. starts to occur under a magnetic field of about 5.5 T (Fig. 2E). The analogous field-dependent magnetization data collected from the CaCoGe2−xSixO6 (x = 0, 0.5, 1, 1.5, and 2) series further illustrate the evolution of the field-induced magnetization plateaus/transitions in this system (Fig. 2F).

Anisotropic magnetic properties

The magnetization plateaus observed in the low-temperature field-dependent magnetization data, collected from the crushed and ground crystals of CaCoGe2O6 and CaCoSi2O6, were deemed to be worth further investigation. Low-temperature neutron diffraction data have been collected by others for both CaCoGe2O6 (25) and CaCoSi2O6 (23) powders. The long-range AFM unit cells adopt a similar arrangement, with spins aligned parallel to each other within the infinite edge-sharing CoO6 1D chain and antiparallel to those in adjacent chains. Spins are found to align in the ac plane (23, 25); hence, they should behave differently upon an external magnetic field applied in or out of plane. This could yield an intriguing anisotropy during the magnetization process.

By using Laue alignment, we identified that the flat crystal surface is normal to the [010] axis (fig. S3); thus, the naturally cleaved crystal surface is confirmed to be the ac plane that contains in-plane Co2+ spins. Hence, the anisotropic magnetization data were collected from these crystals by applying the external field parallel and perpendicular to the [010] axis (Fig. 3). After noting how the infinite edge-sharing CoO6 1D chains propagate along the c axis and how the adjacent chains are bridged together by XO4 tetrahedra along the a axis, together with the zero-field AFM magnetic unit cells determined in the literature (23, 25), the anisotropic magnetization data are less daunting to interpret.

Fig. 3. Anisotropic magnetization.

Fig. 3.

The (A and B) temperature-dependent and (C to F) field-dependent dc magnetization data collected from crystals of CaCoGe2O6 and CaCoSi2O6.

The anisotropic dc magnetic susceptibility data of CaCoGe2O6 and CaCoSi2O6 both show a substantial discrepancy between two field directions (Fig. 3, A and B), with details of the Curie-Weiss fitting provided in the Supplementary Materials (fig. S4). The anisotropic field–dependent magnetization data thus reveal the impact on magnetism brought by the difference in electron configuration of Si and Ge, which are in the superexchange pathway. For the CaCoGe2O6 crystal (Fig. 3, C and E), after deconvoluting the isotropic field–dependent magnetization data, it is clearly seen that an AFM-to-ferromagnetic transition has been induced by H = 7.3 T when the field is applied perpendicular to the [010] axis. The proper “square-S-shape” M versus H curve accompanied by a moment of nearly 3 μB/f.u. suggests that the system has reached a full (~100%) ferromagnetic saturation (Fig. 3E). Therefore, the spins in every other infinite CoO6 1D chain may have been flipped by the applied field and, thus, are all aligned parallel to each other along the field direction. It is necessary to emphasize that we do not rule out the possibility that the M versus H curve with H // [010] (Fig. 3C) will reach saturation at fields greater than 9 T. For the CaCoSi2O6 crystal (Fig. 3, D and F), both analogous isotherms collected reach a flattened saturation plateau of ~2.5 μB/f.u. at H ≈ 5.5 T (~83% of the expected full saturation moment) with potential further steps in the M versus H curve beyond the range of field measured to reach a full ferromagnetic saturation similar to the Ge case. Although the value of full saturation moment Ms for Co2+ varies from case to case in the literature, Ms ≈ 3 μB/Co2+ is reported for CoGeO3, a compound that is analogous to ours (20). Thus, we tentatively state that Ms ≈ 3 μB/Co2+ for our materials based on the current data, although we may observe further steps in the MH isotherms under extremely high fields and the study of which will be of future interest.

In addition, the previously mentioned subtle magnetization plateau observed in the low-field regime of the isotropic M versus H data collected from CaCoSi2O6 (Fig. 2E) is now deconvoluted. The presence of nonlinked 1D CoO6 chains along the [010] direction appreciably dilutes the contribution of the interaction that causes this subtle plateau, as evidenced by its visual absence in the M versus H curve collected with H // [010] (Fig. 3D). It is more important that this subtle magnetization plateau is not observed in the Ge case at all. Therefore, it is valid to postulate that this interaction may be due to the fact that Si resides in the bridging tetrahedra. Hence, this subtle magnetization plateau (~7% of the expected full saturation moment) may only be unveiled if the magnetic interaction is dominated by the superexchange pathway that accommodates an extensive amount of Si (e.g., the interchain coupling pathway in CaCoSi2O6). Given that the interchain Co-Co distance only varies ~2% between these two end members, which is too subtle to cause any qualitative changes in magnetism, the electron configuration of Si and Ge can be used to comprehend this phenomenon (Fig. 4A). The critical difference is that Si4+ has empty 3d orbitals that are available for electron hopping in the superexchange pathway, while Ge4+ has a filled 3d shell so it acts like a “shield”; thus, the communication between spins in the former is more effective than in the latter. That probably also explains why, in general, CaCoSi2O6 requires a much lower field to induce plateaus in its magnetization compared to CaCoGe2O6 (Fig. 4B). In summary, we observe that the choice of bridging tetrahedra between adjacent 1D CoO6 chains can markedly affect magnetization plateaus at finite fields in an anisotropic way. We argue that this is potentially because SiO4 can aid the interchain superexchange pathway between Co2+ centers, while GeO4 tends to block it during the magnetization process. This occurs as a consequence of the different electron configurations of Si and Ge.

Fig. 4. Electron configuration.

Fig. 4.

The schematic illustration for (A) the electron configuration of Si and Ge and (B) their impact on the manipulation of interchain AFM coupling by magnetic field.

Heat capacity

Heat capacity data were collected from sintered pellets of the crushed and ground crystals of CaCoGe2O6 and CaCoSi2O6, under various external applied fields (μ0H = 0, 3, 6, and 9 T), with details of background fitting (fig. S5) provided in Supplementary Text. The total heat capacity Ctotal data (Fig. 5, A and B) and the extracted magnetic contribution Cmag/T (Fig. 5, C and D) are plotted against temperature for both CaCoGe2O6 and CaCoSi2O6. The sharp peaks in the 0-T curves (Ge, ~20 K; Si, ~10 K) match well with the phase transitions in the temperature-dependent magnetization data. In the 3-T curves, the peaks are suppressed in magnitude, while their shapes remain almost intact, aligning perfectly with the AFM nature of this long-range magnetic ordering. However, the heat capacity data become more complicated under higher external magnetic fields (μ0H = 6 and 9 T), since these fields induce the plateaus observed in the magnetization data, hence severely interfering with the otherwise perfect AFM long-range ordering under low fields. For CaCoGe2O6, although the transition peaks in the 6- and 9-T curves have been suppressed extensively, the peak width only gets moderately broadened, accompanied by the sequential shift of peak position to a lower temperature in minor steps. In contrast, for CaCoSi2O6, the peaks in the 6- and 9-T curves are severely deformed, and the anomalies tend to move up in temperature, especially in the 9-T data. Thus, under high magnetic fields, the spins in CaCoGe2O6 tend to retain the original AFM arrangement or undergo a proper AFM-to-ferromagnetic transition (nearly 3 μB/f.u., full saturation for S = 3/2 Co2+), depending on the direction of field applied (Fig. 3, C and E). However, although the spins in CaCoSi2O6 communicate more effectively through the bridging SiO4 tetrahedra, maximum saturated magnetization plateaus reached are only ~2.5 μB/f.u. at 9 T in both field directions (Fig. 3, D and F), which indicates that a fully ordered spin arrangement cannot be achieved within our measurable field range. In summary, the heat capacity of CaCoGe2O6 shows qualitatively different features compared to CaCoSi2O6 under high fields and suggests that spins in the former tend to form longer-range ordering than in the latter, which coincides with their magnetization data. The magnetic entropy change ΔSmag can be used to diagnose whether the spins in the magnetic system are inclining toward the Ising model or the Heisenberg model. Our experimental data indicate that ΔSmag reaches a saturation value of ~5.41 J/mol/K for CaCoGe2O6 (Fig. 5E) and ~5.16 J/mol/K for CaCoGe2O6 (Fig. 5F), respectively. The Heisenberg spin prediction for an S = 3/2 (Co2+ high-spin d7) system should be Rln(2S + 1) = 11.53 J/mol/K, while the Ising spin prediction of Rln(2) = 5.76 J/mol/K is clearly closer to our experimental values (Ge, ~93.9%; Si, ~89.6%). Thus, we can tentatively conclude that the magnetic entropy released at the magnetic phase transition is associated with the bulk long-range ordering of spins in both compounds that are dominantly Ising-like, even after manipulated by high magnetic fields, which is unusual for Co2+ in extended lattices. Specific geometric arrangement of Co2+ (3d7) is often required, so that the interplay of crystal electric field and SOC could lead to an effective S = ½ ground state. Although Ising models have been adopted in some other compounds containing octahedral Co2+ 1D chains or 2D layers, the preserve of Ising-like behavior under high magnetic fields has been rarely spotted.

Fig. 5. Heat capacity.

Fig. 5.

(A and B) the total heat capacity Ctotal, (C and D) the magnetic contribution Cmag/T, and (E and F) the magnetic entropy Smag plotted against temperature for CaCoGe2O6 and CaCoSi2O6.

Theoretical calculations

In (37), we developed a single-ion model with proper crystal field extracted from the density functional theory calculations to describe the complicated spin-orbit moments caused by the interplay among the crystal field, Hund’s rule coupling, and the SOC. For such a model, the single-ion magnetic susceptibility tensor can be obtained by the exact diagonalization of the single-ion many-body Hamiltonian, from which three unique crystalline directions can be determined from the singular value decomposition as the “magnetic principal axes.” The external magnetic field along these directions will be parallel to its induced magnetization with the ratio between the two being the corresponding eigenvalues of the susceptibility tensor. In the present study, we applied the above method to CaCoGe2O6 and CaCoSi2O6. Our results indicate that for both CaCoGe2O6 and CaCoSi2O6, the principal axis y~ is along the crystallographic b axis, while x~ and z~ lie in the ac plane. We further calculated the Curie-Weiss effective magnetic moments along different principal axes, as shown in Table 1, with CaCoGe2O6 exhibiting moments of 5.07, 3.19, and 2.44 μB along x~, y~, and z~, respectively. In comparison, CaCoSi2O6 displays magnetic moments of 4.27, 4.32, and 1.95 μB along the corresponding axes. Our theoretical results obtained from the single-ion model indicate three important points. First, the large experimental isotropic effective magnetic moments can be well explained from our simulations. On the basis of the calculated magnetic moments along different principal axes, we can yield values of 6.47 μB (CaCoGe2O6) and 6.38 μB (CaCoSi2O6) for theoretical isotropic effective magnetic moments. They show an acceptable deviation (~13 to 14%) compared to those values obtained from experimental data (CaCoGe2O6, 5.69 μB; CaCoSi2O6, 5.66 μB). Second, both the size and anisotropy of the saturation magnetic moments in these two materials can be explained by the single-ion physics. As shown in Table 1, the Curie-Weiss saturation magnetic moment of CaCoGe2O6 manifests substantial anisotropy, with a large value of 2.93 μB along x~, a moderate value of 1.84 μB along y~, and a relatively small value of 1.41 μB along z~. These results are in good agreement with experimental anisotropic magnetization data shown in Fig. 3 (C and E). The 1.8-K MH isotherm in blue (Fig. 3C) does not show any signs of saturation, with the maximum magnetization reached ~1.3 μB at 9 T, which is much smaller compared to the analogous data in orange (~3 μB; Fig. 3E), consistent with the calculated values listed in Table 1. In contrast, for CaCoSi2O6, the theoretical values of saturation magnetic moment of x~ (2.47 μB) and y~ (2.49 μB) are only marginally different, which coincides with the fact that the 1.8-K MH isotherms in blue and orange saturate at almost the same plateau (~2.5 μB) in Fig. 3 (D and F). This discrepancy in the saturation magnetic moment between two materials stems from the different crystal field hybridization of dxz and dyz orbitals (17 meV for CaCoGe2O6 and 3 meV for CaCoSi2O6; referring to fig. S6), which results from varying degrees of distortion in the CoO6 octahedra. Third, in both materials, the magnetic moments contain considerably large orbital contribution, namely, 1.30 μB (26%) for x~ and 0.64 μB (20%) for y~ of CaCoGe2O6 and 0.95 μB (22%) for x~ and 0.97 μB (22%) for y~ of CaCoSi2O6, which are in good agreements with the experimental data. The large orbital component in the magnetic moment can be interpreted by the domination of Hunds’ coupling over the octahedral crystal field splitting, leading to large portion of t2g5eg2 rather than t2g6eg1 configurations in the many-body ground state. For the spin state with t2g5eg2 configuration, the formation of magnetic moment can be understood from LS coupling scheme with total orbital angular moment L = 1 and spin angular moment S = 3/2, which contains large portion of orbital magnetic moments.

Table 1. Calculated effective magnetic moments.

Materials Magnetic principal axis Total effective magnetic moments (μB) Saturation magnetic moments (μB) Orbital effective magnetic moments (μB) Spin effective magnetic moments (μB) Percentage of orbital component
CaCoGe2O6 x~ 5.07 2.93 1.30 3.78 26%
y~ 3.19 1.84 0.64 2.55 20%
z~ 2.44 1.41 0.35 2.10 14%
CaCoSi2O6 x~ 4.27 2.47 0.95 3.32 22%
y~ 4.32 2.49 0.97 3.35 22%
z~ 1.95 1.13 0.12 1.83 6%

DISCUSSION

In this work, the Co-bearing pyroxenes CaCoGe2O6 and CaCoSi2O6, as well as their solid solutions, have been investigated both theoretically and experimentally. Their crystal structures extracted from our XRD data are in good agreement with those reported in the literature (23, 25). The Curie-Weiss fits of the isotropic temperature-dependent magnetization data of both compounds indicate the local maxima of susceptibility curves as the onset of AFM ordering, consistent with the neutron powder diffraction data (23, 25). The calculated effective moments reveal substantial contributions due to SOC. Upon cooling down below TN, field-dependent magnetization data unveil a sharp field–induced transition for CaCoGe2O6 and multiple magnetization plateaus beneath the full saturation moment for CaCoSi2O6. These magnetization anomalies are highly anisotropic. Heat capacity data indicate that the spins in both compounds are dominantly Ising-like, even after being manipulated by high applied magnetic fields.

Beyond revealing the enormous orbital contribution to magnetism, our first-principles calculation results further justify the observed discrepancy in anisotropic magnetic saturation between two compounds, which stems from different crystal field hybridizations of Co 3d orbitals. Furthermore, despite the fact that our single-ion model manages to capture some substantial aspects when comparing these two compounds, additional phenomena such as the plateau observed only in the direction perpendicular to [010] under low magnetic fields for CaCoSi2O6, as well as the discrepancy in magnetic fields required to induce any saturation between these two compounds, cannot be adequately addressed if only the local interactions are taken into account. Hence, it ultimately proves that exchange interactions beyond our single-ion model, which are strongly affected by the bridging tetrahedra (SiO4 versus GeO4), play an important role during magnetization.

In summary, we conclude that the choice of bridging tetrahedra (i.e., SiO4 versus GeO4) between adjacent 1D CoO6 chains in a pyroxene-based quantum magnet can locate magnetization plateaus in a measurable field range, potentially attributed to the empty 3d orbitals of Si that enable electron hopping. Our research makes pyroxenes an excellent platform to study the dynamically intertwined lattice, orbital, charge, and spin degrees of freedom in the quantum regime, as a result of their highly versatile crystal structure and fruitful choice of elements. Further investigations of these materials including but not limited to various spectroscopic measurements (neutron, Raman, etc.) and magnetization/heat capacity under extreme high fields will be of future interest.

MATERIALS AND METHODS

Materials synthesis

The powders of magenta CaCoGe2−xSixO6 (x = 0, 0.5, 1, 1.5, and 2) were made through conventional solid-state synthesis. Stoichiometric amounts of CaCO3 (Alfa Aesar; 99.999%), CoO (Sigma-Aldrich; 99%), SiO2 (Sigma-Aldrich; 99.5%), and GeO2 (Alfa Aesar; 99.999%) were thoroughly ground together using an agate mortar and pestle and then transferred into a dense alumina crucible. The reaction mixture was first slowly (ramp rate = 1°C/min) heated to 1000°C in air, held overnight to decompose the carbonate, and then annealed in air at 1100° to 1200°C (ramp rate = 3°C/min) for two to three periods of 48 hours with intermittent grindings. The compounds with a high Si content required SiO2 in a fine particle size to facilitate the product formation.

Macroscopic single crystals of CaCoSi2O6 and CaCoGe2O6 were grown via the floating-zone method. The precursor was ground using an agate mortar and pestle for 10 min to achieve a fine and uniform polycrystalline state to facilitate rod making. The powder was then molded into a long thin balloon and pressed under 60 MPa in a hydraulic press. The rods were sintered in a high-temperature furnace at 1150° and 1250°C for 8 to 12 hours for the Si and Ge compounds, respectively. The optical floating zone furnace (Crystal Systems, HP-10000) has four halogen lamps. We used an airflow of 2 liters/min under 1-atm pressure. After a stable molten zone was formed, the feed rod and seed rod rotated in a reverse direction, each at 25 rpm. Since the material was quite easy to crystalize, a flat crystal surface formed at a low rotation speed disconnected the molten zone. Thus, a high rotating speed was necessary to maintain a uniform molten zone. The growth rate was kept at 5 mm/hour until the end of growth. Scanning electron microscopy images of as-grown crystals were collected using a Quanta 200 FEG environmental scanning electron microscopy.

Powder-XRD and single-crystal XRD measurements and refinements

The solid-state ceramic reaction progress was monitored using laboratory p-XRD data collected at room temperature on a Bruker D8 FOCUS diffractometer (Cu Kα) over the 2θ range between 5° and 70°. Once the reactions were deemed finished, laboratory p-XRD data with much better statistical significance, covering a 2θ range between 5° and 110°, were collected from each sample. The Le Bail fitting of the acquired p-XRD patterns was conducted via the TOPAS software.

Single-crystal XRD was carried out using a Bruker D8 Quest Eco diffractometer equipped with a Photon III CPAD detector and monochromated Mo Kα radiation (λ = 0.71073 Å). Frames were integrated and corrected for Lorentz and polarization effects using SAINT 8.40b and were corrected for absorption effects using SADABS V2016/2 (38). Space groups were assigned on the basis of systematic absences, the E-statistics, agreement factors for equivalent reflections, and the successful refinement of the structures. The structures were solved by direct methods, expanded through successive difference Fourier maps using SHELXT, and refined against all data using the SHELXL-2018 (39, 40) software package as implemented in Olex2 (41, 42). Weighted R factors, Rw, and all goodness-of-fit indicators are based on F2. Thermal displacement parameters for all atoms were refined anisotropically. The crystal structures were visualized by the VESTA program.

Physical property measurements

The magnetization data were collected using the VSM option of a Quantum Design Physical Property Measurement System. Temperature-dependent magnetization (M) data were collected under an applied external field (H) of 0.1 T. Magnetic susceptibility is defined as M0H. Field-dependent isothermal magnetization data between applied fields of +9 and −9 T were collected at various temperatures. Isotropic magnetization data were collected from crushed and ground FZ-grown crystals, while anisotropic magnetization data were collected with the applied external field perpendicular or parallel to the relatively flat surface of the FZ-grown crystals. Heat capacity was measured using the standard relaxation method in the Physical Property Measurement System over the temperature range of 1.8 to 50 K under an applied external field of 0, 3, 6, and 9 T, with the 0-T data extended to 150 K for the background phonon fitting.

Laue alignment for crystal orientations

Laue pattern images were produced using the HUBER Laue Imaging Plate x-ray diffractometer with a 1500-W Cu target. Samples were mounted parallel to beam axis at a distance of 10 mm from a collimator tube with an aperture of 0.8 mm. The fine-focus x-ray was operated at 40 mA and 20 kV in air at room temperature, and samples were measured for a typical exposure time of 1 to 3 min—longer times proving necessary for higher-quality patterns. These diffraction patterns were then transferred onto the OrientExpress software and oriented perpendicular with the [010] axis given lattice parameters derived from single-crystal XRD measurements.

Theoretical calculations

Local moments including spin and orbital components are crucial ingredients of the complicated magnetism in both CaCoGe2O6 and CaCoSi2O6. After extracting the crystal field and SOC terms from first principles, a single-ion model is derived to understand the formation mechanism of local moments of Co2+ ion in both materials, which reads

H^atom=H^kanamori+H^cf+H^soc

where the first term encodes both the Coulomb interaction (U) and Hund’s coupling (J) and the last two terms denote the local crystal field interaction and SOC, respectively. The Vienna ab initio simulation package (43, 44) is used to perform the first-principles calculations with the choice of the Perdew-Burke-Ernzerhof generalized gradient approximation for the exchange-correlation potential (43). An energy cutoff of 550 eV for the plane wave basis and a fine grid of 9 × 9 × 9 for k-point are adopted. Then, a tight-binding model is constructed with the help of Wannier90 software, which quantifies the crystal field interaction (∆Oct for octahedral splitting between t2g and eg orbitals, ∆1 for further splitting between dxz/dyz and dxy due to the symmetry breaking from C3 to C2 symmetry) and SOC (λsoc for the strength of SOC). Because the little group of the Co2+ ion is C2 with the two-fold axis along the crystallographic unique b axis, we select the local axes x and y to lie in the O-Co-O plane, with the constraint that x + y equals b. For CaCoGe2O6, ∆Oct ≈ 830 meV, ∆1 ≈ 33 meV, and λsoc ≈ 66 meV, while ∆Oct ≈ 910 meV, ∆1 ≈ 59 meV, and λsoc ≈ 67 meV for CaCoSi2O6. A typical value of 5 eV for the Hubbard U and 0.8 eV for the Hunds’ coupling J is set for the many-body interactions. Following the method we proposed in a previous paper (37), the lowest Kramers’ doubly degenerate eigenstates (pseudo-spin space) are obtained by solving the many-body atomic Hamiltonian within the Fock subspace of seven electrons, which is about 25 and 20 meV below the first excited states for CaCoGe2O6 and CaCoSi2O6, respectively. The theoretical discussion is based on the calculation within the aforementioned pseudo-spin subspace.

For convenience, we define the isotropic effective magnetic moment as

μeffiso=μeff,x~2+μeff,y~2+μeff,z~2

where μeff,x~, μeff,y~, and μeff,z~ are the calculated effective magnetic moment along each principal axis.

Acknowledgments

Funding: This work was supported by US Department of Energy Division of Basic Energy Sciences, through the Institute for Quantum Matter, grant no. DE-SC0019331 (to L.J. and R.J.C.); The Gordon and Betty Moore Foundation, EPiQS initiative, grant no. GBMF-9066 (to L.J., D.N., and R.J.C.); US Department of Energy, Office of Science, National Quantum Information Science Research Centers, Co-design Center for Quantum Advantage (C2QA), grant no. DE-SC0012704 (to X.X., C.Y., and R.J.C.); National Science Foundation through the Princeton University Materials Research Science and Engineering Center (PCCM), grant no. DMR-2011750; National Science Foundation, grant no. NSFDMR-2003117 (to A.N.R. and H.Z.); and US Department of Energy Division of Basic Energy Sciences, grant no. DE-SC0023648 (to W.X.).

Author contributions: Conceptualization: L.J. and R.J.C. Methodology: L.J. and S.P. Investigation: L.J., S.P., A.N.R., X.X., D.N., C.Y., and Y.J.B. Formal analysis: L.J. and S.P. Visualization: L.J. and S.P. Supervision: R.J.C., X.D., H.Z., and W.X. Writing—original draft: L.J. and S.P. Writing—review and editing: R.J.C., X.D., H.Z., and W.X. Funding acquisition: R.J.C., X.D., H.Z., and W.X.

Competing interests: The authors declare that they have no competing interest.

Data and materials availability: The CSD number 2368031-2368032 contains the supplementary crystallographic data for CaCoGe2O6 and CaCoSi2O6. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif. All other data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.

Supplementary Materials

This PDF file includes:

Supplementary Text

Figs. S1 to S7

Tables S1 to S3

sciadv.adp4685_sm.pdf (1.8MB, pdf)

REFERENCES AND NOTES

  • 1.Macdonald A. H., Schiffer P., Samarth N., Ferromagnetic semiconductors: Moving beyond (Ga,Mn)As. Nat. Mater. 4, 195–202 (2005). [Google Scholar]
  • 2.Schiffer P., Ramirez A. P., Bao W., Cheong S. W., Low temperature magnetoresistance and the magnetic phase diagram of La1−xCaxMnO3. Phys. Rev. Lett. 75, 3336–3339 (1995). [DOI] [PubMed] [Google Scholar]
  • 3.Karunadasa H., Huang Q., Ueland B. G., Schiffer P., Cava R. J., Ba2LnSbO6 and Sr2LnSbO6 (Ln = Dy, Ho, Gd) double perovskites: Lanthanides in the geometrically frustrating fcc lattice. Proc. Natl. Acad. Sci. U.S.A. 100, 8097–8102 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Bednorz J. G., Müller K. A., Possible high Tc superconductivity in the Ba−La−Cu−O system. Z. Phys. B 64, 189–193 (1986). [Google Scholar]
  • 5.Cohen R. E., Origin of ferroelectricity in perovskite oxides. Nature 358, 136–138 (1992). [Google Scholar]
  • 6.Lacorre P., Goutenoire F., Bohnke O., Retoux R., Lallgant Y., Designing fast oxide-ion conductors based on La2Mo2O9. Nature 404, 856–858 (2000). [DOI] [PubMed] [Google Scholar]
  • 7.Jin L., Ni D., Gui X., Straus D. B., Zhang Q., Cava R. J., Ferromagnetic double perovskite semiconductors with tunable properties. Adv. Sci. 9, 2104319 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Jin L., Varnava N., Ni D., Gui X., Xu X., Xu Y., Bernevig B. A., Cava R. J., Electron doping of a double-perovskite flat-band system. Proc. Natl. Acad. Sci. U.S.A. 120, 2218997120 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Warren J. M., Hauri E. H., Pyroxenes as tracers of mantle water variations. J. Geophys. Res. Solid Earth 119, 1851–1881 (2014). [Google Scholar]
  • 10.Boyd F. R., A pyroxene geotherm. Geochim. Cosmochim. Acta 37, 2533–2546 (1973). [Google Scholar]
  • 11.Lindsley D. H., Pyroxene thermometry. Am. Mineral. 68, 477–493 (1983). [Google Scholar]
  • 12.Morimoto N., Fabries J., Ferguson A. K., Ginzburg I. V., Ross M., Seifert F. A., Zussman J., Aoki K., Gottardi G., Nomenclature of pyroxenes. Am. Mineral. 73, 1123–1133 (1988). [Google Scholar]
  • 13.Jodlauk S., Becker P., Mydosh J. A., Khomskii D. I., Lorenz T., Streltsov S. V., Hezel D. C., Bohatý L., Pyroxenes: A new class of multiferroics. J. Phys. Condens. Matter 19, 432201– 432201 (2007). [Google Scholar]
  • 14.Kim I., Jeon B.-G., Patil D., Patil S., Nénert G., Kim K. H., Observation of multiferroic properties in pyroxene NaFeGe2O6. J. Phys. Condens. Matter 24, 306001 (2012). [DOI] [PubMed] [Google Scholar]
  • 15.Ackermann M., Andersen L., Lorenz T., Bohatý L., Becker P., Anisotropy study of multiferroicity in the pyroxene NaFeGe2O6. New J. Phys. 17, 013045 (2015). [Google Scholar]
  • 16.Ding L., Colin C. V., Darie C., Robert J., Gay F., Bordet P., One-dimensional short-range magnetic correlations in the magnetoelectric pyroxene CaMnGe2O6. Phys. Rev. B 93, 064423 (2016). [Google Scholar]
  • 17.Redhammer G. J., Roth G., Senyshyn A., Tippelt G., Pietzonka C., Crystal and magnetic spin structure of Germanium-Hedenbergite, CaFeGe2O6, and a comparison with other magnetic/magnetoelectric/multiferroic pyroxenes. Z. Kristallogr. Cryst. Mater. 228, 140–150 (2013). [Google Scholar]
  • 18.Streltsov S. V., Khomskii D. I., Electronic structure and magnetic properties of pyroxenes (Li,Na)TM (Si,Ge)2O6: Low-dimensional magnets with 90° bonds. Phys. Rev. B 77, 064405 (2008). [Google Scholar]
  • 19.Zhao L., Hu Z., Guo H., Geibel C., Lin H. J., Te Chen C., Khomskii D., Tjeng L. H., Komarek A. C., Single crystal growth and physical properties of pyroxene CoGeO3. Crystals 11, 378 (2021). [Google Scholar]
  • 20.Guo H., Zhao L., Baenitz M., Fabrèges X., Gukasov A., Melendez Sans A., Khomskii D. I., Tjeng L. H., Komarek A. C., Emergent 1/3 magnetization plateaus in pyroxene CoGeO3. Phys. Rev. B 3, L032037 (2021). [Google Scholar]
  • 21.Gori C., Tribaudino M., Mezzadri F., Skogby H., Hålenius U., Co2+-doped diopside: Crystal structure and optical properties. Phys. Chem. Miner. 45, 443–461 (2018). [Google Scholar]
  • 22.Ghose S., Wan C., Okamura F. P., Crystal structures of CaNiSi2O6 and CaCoSi2O6 and some crystal-chemical relations in C2/c clinopyroxenes. Am. Mineral. 72, 375–381 (1987). [Google Scholar]
  • 23.Durand G., Vilminot S., Rabu P., Derory A., Lambour J. P., Ressouche E., Synthesis, structure, and magnetic properties of CaMSi2O6(M= Co, Ni) compounds and their solid solutions. J. Solid State Chem. 124, 374–380 (1996). [Google Scholar]
  • 24.Tribaudino M., Aliatis I., Bersani D., Gatta G. D., Lambruschi E., Mantovani L., Redhammer G., Lottici P. P., High-pressure Raman spectroscopy of Ca(Mg,Co)Si2O6 and Ca(Mg,Co)Ge2O6 clinopyroxenes. J. Raman Spectrosc. 48, 1443–1448 (2017). [Google Scholar]
  • 25.Redhammer G. J., Roth G., Treutmann W., Paulus W., André G., Pietzonka C., Amthauer G., Magnetic ordering and spin structure in Ca-bearing clinopyroxenes CaM2+(Si, Ge)2O6, M = Fe, Ni, Co, Mn. J. Solid State Chem. 181, 3163–3176 (2008). [Google Scholar]
  • 26.Titorenkova R., Kostov-Kytin V., Dimitrov T., Synthesis, phase composition and characterization of Co-diopside ceramic pigments. Ceram. Int. 48, 36781–36788 (2022). [Google Scholar]
  • 27.Mantovani L., Tribaudino M., Dondi M., Zanelli C., Synthesis and color performance of CaCoSi2O6 pyroxene, a new ceramic colorant. Dyes Pigm. 120, 118–125 (2015). [Google Scholar]
  • 28.J. L. Schlenker, A Phenomenological Treatment of Thermal Expansion in Crystals of the Lower Symmetry Classes and the Crystal Structures of CaCoSi2O6 and CaNiSi2O6 (Virginia Polytechnic Institute and State University, 1976). [Google Scholar]
  • 29.Xiao M., Wei Y., Zhang P., The effect of sintering temperature on the crystal structure and microwave dielectric properties of CaCoSi2O6 ceramic. Mater. Chem. Phys. 225, 99–104 (2019). [Google Scholar]
  • 30.Joseph T., Lazar K. A., Divya D., Preparation and structural, dielectric characterization of CaBSi2O6 (B = Mg, Co, Ni, Mn, Zn) ceramics. IOP Conf. Ser. Mater. Sci. Eng. 1166, 012025 (2021). [Google Scholar]
  • 31.P. A. Maksimov, A. V. Ushakov, A. F. Gubkin, G. J. Redhammer, S. M. Winter, A. I. Kolesnikov, A. M. dos Santos, Z. Gai, M. A. McGuire, A. Podlesnyak, S. V. Streltsov, Cobalt-based pyroxenes: A new playground for Kitaev physics and Ising model realization. arXiv:2401.13550 [cond-mat.str-el] (8 September 2024).
  • 32.Mehmood S., Ali Z., Alharbi Y. T., Alderhami S. A., Almanqur L., Structural and magneto-elastic properties of the quadruple perovskites CaCu3B2Os2O12 (B = Mn-Ni): The Heisenberg model and DFT study. J. Electron. Mater. 52, 5872–5883 (2023). [Google Scholar]
  • 33.Lefrançois E., Songvilay M., Robert J., Nataf G., Jordan E., Chaix L., Colin C. V., Lejay P., Hadj-Azzem A., Ballou R., Simonet V., Magnetic properties of the honeycomb oxide Na2Co2TeO6. Phys. Rev. B 94, 214416 (2016). [Google Scholar]
  • 34.M. Takigawa, F. Mila, “Magnetization plateaus” in Introduction to Frustrated Magnetism: Materials, Experiments, Theory, C. Lacroix, P. Mendels, F. Mila, Eds. (Springer, 2011), pp. 241–267. [Google Scholar]
  • 35.Shannon R. D., Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 32, 751–767 (1976). [Google Scholar]
  • 36.Goodenough J. B., Spin-orbit-coupling effects in transition-metal compounds. Phys. Rev. 171, 466–479 (1968). [Google Scholar]
  • 37.Xu X., Hao Y., Peng S., Zhang Q., Ni D., Yang C., Dai X., Cao H., Cava R. J., Large off-diagonal magnetoelectricity in a triangular Co2+-based collinear antiferromagnet. Nat. Commun. 14, 8034 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Bruker, Bruker SAINT and SADABS (Bruker AXS Inc., 2016). [Google Scholar]
  • 39.Sheldrick G. M., A short history of SHELX. Acta Crystallogr. A 64, 112–122 (2008). [DOI] [PubMed] [Google Scholar]
  • 40.G. M. Sheldrick, SHELXS-97 and SHELXL-97, Program for Crystal Structure Solution and Refinement (University of Gottingen, 1997). [Google Scholar]
  • 41.Dolomanov H., Bourhis O. V., Gildea L. J., Howard R. J., Puschmann J. A. K., OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 42, 339–341 (2009). [Google Scholar]
  • 42.P. Müller, R. Herbst-Irmer, A. L. Spek, T. R. Schneider, M. R. Sawaya, Crystal Structure Refinement: A Crystallographer’s Guide to SHELXL (Oxford Univ. Press, 2006). [Google Scholar]
  • 43.Kresse G., Furthmüller J., Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169–11186 (1996). [DOI] [PubMed] [Google Scholar]
  • 44.Kresse G., Joubert D., From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999). [Google Scholar]

Associated Data

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

Supplementary Materials

Supplementary Text

Figs. S1 to S7

Tables S1 to S3

sciadv.adp4685_sm.pdf (1.8MB, pdf)

Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES