Previous Article in Journal
Components for an Inexpensive CW-ODMR NV-Based Magnetometer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of the Non-Magnetic Ion Doping on the Magnetic Behavior of MgCr2O4

1
Key Laboratory of Advanced Processing and Recycling of Nonferrous Metals, School of Materials Science and Engineering, Lanzhou University of Technology, Lanzhou 730050, China
2
Institute of High Energy Physics, Chinese Academy of Sciences (CAS), Beijing 100049, China
3
Spallation Neutron Source Science Center, Dongguan 523803, China
4
Guangdong Provincial Key Laboratory of Magnetoelectric Physics and Devices, School of Physics, Sun Yat-sen University, Guangzhou 510275, China
5
Centre for Physical Mechanics and Biophysics, and Centre for Neutron Science and Technology, School of Physics, Sun Yat-sen University, Guangzhou 510275, China
6
Department of Physics, Southern University of Science and Technology, Shenzhen 518055, China
7
Neutron Science Center, Songshan Lake Materials Laboratory, Dongguan 523808, China
8
Shenzhen Key Laboratory of Advanced Quantum Functional Materials and Devices, Southern University of Science and Technology, Shenzhen 518055, China
9
Beijing National Laboratory for Condensed Matter Physics, Institute of Physics, Chinese Academy of Sciences, Beijing 100190, China
10
Wenzhou Pump and Valve Engineering Research Institute, Lanzhou University of Technology, Wenzhou 325105, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Magnetism 2025, 5(3), 19; https://doi.org/10.3390/magnetism5030019
Submission received: 3 July 2025 / Revised: 13 August 2025 / Accepted: 20 August 2025 / Published: 25 August 2025
(This article belongs to the Special Issue Research on the Magnetism of Heavy-Fermion Systems)

Abstract

Geometrically frustrated magnets exhibit exotic excitations due to competing interactions between spins. The spinel compound MgCr2O4, a three-dimensional Heisenberg antiferromagnet, hosts both spin-wave and spin-resonance modes, but the origin of its resonant excitations remains debated. Suppressing magnetic order via non-magnetic doping can help isolate these modes in neutron scattering studies. We synthesized Ga3+ and Cd2+-doped MgCr2O4 via solid-state reaction and analyzed their structure and magnetism. Ga3+ doping (0–20%) causes anomalous lattice shrinkage due to site disorder from Ga3+ occupying both Mg2+ and Cr3+ sites. Magnetically, Ga3+ doping drives the system from the antiferromagnetic order to a spin-glass state, fully suppressing magnetic ordering at 20% doping. In contrast, Cd2+ replaces only Mg2+, expanding the lattice and meantime inducing strong spin-glass behavior. At 10% Cd2+, long-range antiferromagnetic order is entirely suppressed. Thus, 10% Cd-doped MgCr2O4 offers an ideal platform to study the resonant magnetic excitations without any spin-wave interference.

1. Introduction

Magnetic frustration emanating from competing exchange interactions at the microscopic scale engenders macroscopic degeneracy of spin configurations and exotic magnetic excitations [1,2]. Predominantly, such phenomena manifest in materials known as geometrically frustrated magnets, wherein magnetic ions occupy the vertices of specific lattices. Examples include YbMgGaO4 [3], which features a triangular lattice of Yb3+ ions and exhibits a quantum spin liquid state with itinerant excitations and quantum spin fluctuations. ZnCu3(OH)6Cl2 [4,5], which hosts a spin-1/2 kagome lattice of Cu2+ ions and whose ground state may lie near a quantum critical point or resemble a critical spin liquid, exhibits highly diffusive scattering intensity in neutron scattering experiments, as reported by Tian-Heng Han et al. [6], distinct from that of non-frustrated quantum magnets and indicative of strong quantum fluctuations—providing direct evidence of its frustrated antiferromagnetic nature. Neutron diffuse scattering measurements on the kagome lattice compound Y0.5Ca0.5BaCo4O7 reveal the presence of only a short-range chiral order, without the emergence of long-range antiferromagnetic ordering, making it a compelling realization of a frustrated antiferromagnet [7]. In addition, neutron scattering experiments on this system reveal the onset of spin-freezing at low temperatures, accompanied by a broad dynamic response with significant temperature dependence [8]. Additionally, Na4Ir3O8 [9] is present, a three-dimensional frustrated magnet with a hyperkagome lattice of Ir4+, whose ground state has been proposed as a rare 3D realization of a spin liquid. The pyrochlore lattice, another archetype of 3D geometry-frustrated lattice, has garnered enormous attention for decades as it could present a plethora of unique quantum behaviors, including spin ice, magnetic monopoles, quantum spin liquids, and spin resonances [10,11,12,13]. The lattice structure of magnetic ions is the motif of a 3D network of corner-sharing tetrahedra and is identified in two main categories of natural materials: pyrochlore oxides A2B2O7 [12,14,15,16] with dual pyrochlore lattices on both A and B atomic sites and spinel chalcogenides AB2X4 (X = O, S, Se) [17,18,19,20,21,22,23], where only the B-site cations form a pyrochlore lattice.
The magnesium trinary compound MgCr2O4, as a member of chromium-based spinel oxides ACr2O4 (A = Zn, Mg, Cd, Hg) [19,24,25,26,27,28,29], is considered as a classical Heisenberg pyrochlore antiferromagnet. The sole magnetic ion Cr3+ (3d3, S = 3 / 2 ) in MgCr2O4 [25], with half-filled t 2 g orbitals and quenched orbital momentum, resides on a pyrochlore lattice and is antiferromagnetically coupled with its six nearest Cr3+ neighbors. The Curie–Weiss temperature θ CW of MgCr2O4 varies from 346 to 430 K [25,26,30], indicating strong antiferromagnetic interactions among Cr3+, and an apparent hump around 50 K in the magnetic susceptibility provides evidence of short-range spin ordering. Upon cooling well below | θ CW | , MgCr2O4 undergoes intricate phase transitions with antiferromagnetic transition temperatures T N ranging from 12 to 16 K [25,31,32] and a concomitant decrease in lattice symmetry from cubic (space group: F d 3 ¯ m ) to tetragonal phase (space group: I 4 1 / a m d ). The large value of f = | θ CW | / T N > 20 clearly indicates that the whole spin system in MgCr2O4 is highly frustrated. Neutron powder diffraction (NPD) experiments have shown multiple magnetic propagation vectors within the antiferromagnetic phase of MgCr2O4 below T N [24,26,31]. Notably, these wave vectors exhibit various combinations across different NPD experiments, yet the magnetic moment of Cr3+ remains unsaturated and consistently smaller than the fully ordered moment of Cr3+ ( 3.87 μ B ) in most instances. Below the Néel temperature, magnetic Bragg reflections are indexed by multiple propagation vectors, including kL,1 = (1/2, 1/2, 0) and kL,2 = (1, 0, 1/2) [26,31], associated with a partially resolved “L phase” [26,31,33]. In addition, some samples exhibit a distinct “H phase” above T N but below T H 16 K, characterized by a propagation vector kH = (0, 0, 1) [26,32]. Some reports of k = ( 1 / 2 , 1 / 2 , 1 / 2 ) [24] further emphasize the complexity of the system’s magnetic ordering and suggest the coexistence of multiple magnetic phases. Additionally, sharp and non-dispersive spin resonances at energy transfer E = 4.5 , 9.0, 13.5 … meV observed in inelastic neutron scattering (INS) spectra were interpreted as transitions between two quantum levels of spin clusters, leading to the classification of MgCr2O4 as a potential molecular spin liquid [24,31,34]. However, this hypothesis has recently been contested from a classical perspective, and the debate surrounding its magnetic ground state remains unresolved [33,34].
Chemical doping is usually considered a promising method to understand complicated magnetic orders and underlying spin–lattice couplings in MgCr2O4, particularly by modulating its low-temperature antiferromagnetic behaviors [29,35,36,37]. Differences in ionic radius ratios and volatilization rates among different elements often hinder the formation of a pure phase, making detailed synthesis and doping studies essential. Moreover, understanding how doping affects the magnetic ion sites is crucial for identifying appropriate doping levels that can effectively suppress magnetic ordering. According to the report in (S. E. Dutton, PRB, 2011) [25], the magnetic properties of MgCr2O4 are profoundly sensitive to metal atoms’ nonstoichiometry both at Mg2+ and Cr3+ sites, and even minor alterations in the microscopic chemical environment significantly impact its Néel temperature TN and spin correlations above TN. A similar effect is also observed in its homologous compound, ZnCr2O4 [35,36,37,38,39], where Cd2+ substitution for Mg2+ and Ga3+ substitution for Cr3+ effectively suppress its TN. To our knowledge, there has been no systematic investigation of chemical doping in magnesium spinel oxide MgCr2O4.
In this work, we report the effects of Ga3+ and Cd2+ doping on the crystal structure and magnetic properties of MgCr2O4. Ga3+ doping effectively modulates the magnetic properties of the MgCr2O4 system by weakening its antiferromagnetic (AFM) order, inducing a transition to a spin-glass state, and eventually driving the system into a paramagnetic state at higher doping levels. Similarly, the introduction of Cd2+ also suppresses AFM interactions. However, due to the larger ionic radius of Cd2+ and the resulting stronger lattice distortion, this effect is more sensitive, enabling spin-freezing to occur at relatively low doping concentrations and temperatures.

2. Materials and Methods

The spinel chromium oxides Mg(Cr1−xGax)2O4 (x = 0, 0.05, 0.10, 0.15, and 0.20) and Mg1−xCdxCr2O4 (x = 0, 0.05, 0.10, 0.15, and 0.20) were synthesized via a conventional high-temperature solid-state reaction method [37,38]. Based on the X-ray powder diffraction (XRD) refinement results, we found that a 1% reduction in the Cr2O3 (99.99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China) stoichiometric amount was necessary to obtain a pure phase. Mg(Cr1−xGax)2O4 was prepared by mixing stoichiometric amounts of Cr2O3, Ga2O3 (99.99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China), and MgO (99.99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China) in air, with a 1% reduction in Cr2O3. The mixture was pre-sintered at 800 °C for 12 h in air. After cooling, it was reground and subsequently reacted at 1200 °C for 24 h to obtain Mg(Cr1−xGax)2O4(x = 0, 0.05, 0.10, 0.15, and 0.20). Similarly, MgO, Cr2O3, and CdO (99.99%, Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China) were mixed in stoichiometric ratios and subjected to the same synthesis procedure to obtain Mg1−xCdxCr2O4(x = 0, 0.05, 0.10, 0.15, and 0.20).
The crystallographic structures of the samples were characterized at room temperature using XRD (Rigaku Smartlab 3kW, Rigaku Corporation, Tokyo, Japan) with Cu Kα radiation (λ = 1.54 Å). The DC magnetic susceptibilities of all samples were measured using the Physical Property Measurement System (PPMS) from Quantum Design (San Diego, CA, USA), under the condition of zero-field cooling (ZFC) and field cooling (FC) processes. The time-of-flight neutron powder diffraction (NPD) measurements were conducted on Mg(Cr1−xGax)2O4 ( x = 0.20 ) samples using the General Purpose Powder Diffractometer (GPPD) at China Spallation Neutron Source (CSNS), and the diffraction data were refined by using the Rietveld method with FullProf (Version: January-2021).

3. Results and Discussions

3.1. MgCr2O4 Doped with Ga

To elucidate the structural modifications induced by Ga doping in the spinel compound MgCr2O4, a combination of X-ray and neutron diffraction analysis was employed. The phases and the cation site occupancy were carefully determined. Figure 1b shows the X-ray diffraction patterns of Mg(Cr1−xGax)2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20) measured at room temperature over the 2 θ range of 10°–80°. All diffraction peaks can be indexed to the cubic spinel structure with space group F d 3 ¯ m (No. 227), and no impurity phases were observed in any of the synthesized samples. A representative Rietveld refinement profile for the undoped MgCr2O4 sample is shown in Figure 1c. Initial XRD refinements with the model of Ga substitution only on Cr (B-site) yielded poor agreement, implying possible Ga on the A-site. However, due to the poor sensitivity of the light elements, such as Mg, XRD is not able to determine the ratio of Mg/Ga on the A-site. Thus, the neutron powder diffraction and refinement were employed because of the strong contrast of the neutron coherent scattering length of Mg (5.375 fm), Ga (7.288 fm), and Cr (3.635 fm) element [40]. The typical tof-NPD refinement for the sample Mg (Cr1−xGax)2O4, x = 0.20, is presented in Figure 1d, indicating that the actual composition is (Mg0.754Ga0.246)(Cr0.923Ga0.077)2O4. This demonstrates that Ga3+ ions substitute on both the A-site (Mg2+) and B-site (Cr3+), resulting in a complex compound with the nominal formula (Mg1−xGax)(Cr1−yGay)2O4 while preserving the spinel structure. The site disorder between Mg2+ and Ga3+ ions has been also observed in other compounds, such as the the nonmagnetic layers in YbMgGaO4 [3].
The obtained crystallographic parameters mainly from the XRD Rietveld refinement of the Gd-doped sample MgCr2O4 are presented in Figure 2. As shown in Figure 2a,b, with the increases in the named Ga3+ doping amount, both Cr3+ and Mg2+ occupancies of the A-site and B-site decrease. While the Cr occupancy nearly remains unchanged, the Mg occupancy increases drastically as the total doping amount of the Ga is constant. The above result suggests Ga ions prefer to stay at the A-site in the low doping level. Although Ga3+ doping does not change the overall symmetry of the spinel structure, the lattice constant a slightly decreases with Ga doping increasing due to the smaller ionic radius of Ga3+. A monotonic decrease in the Cr–Cr and Mg–Mg bond lengths is also observed, consistent with the decrease in lattice constant [Figure 2d,e]. Furthermore, Figure 2f shows that the decrease in the weighted average ionic radius at the A-site outweighs the slight increase at the B-site, suggesting that lattice shrinkage is primarily driven by substitution at the A-site. Such lattice contraction arises from the distinct size mismatch between dopant and host cations: Ga3+ in octahedral coordination (B-site) has a slightly larger ionic radius (62 pm) than Cr3+ (61.5 pm), while Ga3+ in tetrahedral coordination (A-site) is substantially smaller (47 pm) than Mg2+ (57 pm). In previously reported studies on Ga-doped ZnCr2O4 [25,37], no evidence of Ga incorporation at the A-site was observed, which may be attributed to the large ionic radius mismatch between Zn2+ (60 pm) and Ga3+ (47 pm) [41], resulting in an unfavorable substitution environment.
Figure 3 presents the temperature-dependent DC magnetic susceptibility, χ ( T ) , of Mg(Cr1−xGax)2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20) measured under zero-field-cooled (ZFC) and field-cooled (FC) conditions with an applied magnetic field of 0.2 T in the temperature range 2.0–300 K. The Néel temperature ( T N ) is determined by the drop temperature in χ ( T ) . For the undoped sample ( x = 0 ), χ ( T ) drops sharply at 12.6 K, indicating the antiferromagnetic ordering as reported [24,42]. With increasing Ga doping, T N gradually shifts toward lower temperatures and falls below 2.0 K for x = 0.20 (see Figure 4a), which is out of the measurable temperature range on a conventional PPMS. Further magnetic susceptibility measurement of this sample with He3 or Dilution refrigerator equipment is needed to clarify the possible antiferromagnetic order at extreme low temperature. Unlike the T N of MgCr2O4 increase with the lattice constants’ decrease under pressure [43], our results of Ga-doping reveal contrary behavior. One possibility is that the incorporation of Ga3+ ions partially replaces the magnetic Cr3+ ions, resulting in magnetic site dilution. This weakens the magnetic interactions and ultimately leads to a reduction in the T N value.
Furthermore, the ZFC and FC χ ( T ) curves of Ga-doped samples bifurcate below T N , with the downturn in the ZFC branch and the upturn in the FC branch, implying the emergence of spin-glass behaviors [44,45]. The inset in Figure 3 displays the inverse magnetic susceptibility curve and its Curie–Weiss fit for MgCr2O4. The data in the high-temperature range (200 K ≤ T ≤ 300 K) were fitted using the Curie–Weiss law, χ = C/(T θ CW ), where χ denotes the magnetic susceptibility, C represents the Curie constant, and θ CW stands for the Curie–Weiss temperature, as shown in Figure 4. The fitted Weiss temperature | θ CW | for MgCr2O4 is 516 K, which deviates significantly from previously reported values [25,30,46]. This discrepancy may be attributed to the modified synthesis process in which a 1% reduction in the Cr2O3 precursor mass leading to a pure phase MgCr2O4 product. Enhanced phase purity likely strengthens intrinsic antiferromagnetic interactions, thereby increasing the magnitude of θ CW .
With increasing Ga3+ doping, θ CW decreases monotonically [Figure 4b], indicating suppression of antiferromagnetic interactions. The effective magnetic moment per Cr3+ ion ( μ eff ) remains essentially unchanged upon doping [Figure 4c]. The magnetic frustration parameter, defined as f = | θ CW | / T N , further increases with Ga doping [Figure 4d], indicating enhanced magnetic frustration. This effect may arise from the disruption of lattice periodicity due to partial substitution of magnetic B-site ions, leading to the formation of spatially heterogeneous [Cr3Ga1] tetrahedral clusters. The competition between Cr–Cr superexchange interactions in undoped regions and Cr–Ga–Cr exchange pathways at dopant boundaries introduces additional magnetic frustration within the system.
The XRD refinement confirms that Ga3+ ions occupy both A and B-sites, where A-sites are typically non-magnetic, and Cr3+ ions at magnetic B-sites are diluted. This substitution leads to structural distortions in the Mg(Cr1−xGax)2O4 lattice. Figure 3 illustrates the evolution of magnetic phases with increasing Ga doping. Intermediate doping levels ( x = 0.05 –0.10) display spin-glass-like behavior, evidenced by the bifurcation between ZFC and FC curves. At higher doping levels ( x 0.15 ), the system gradually transitions into a paramagnetic state, evidenced by the disappearance of the ZFC–FC splitting.
In ZnCr2O4, Ga3+ ions predominantly substitute for Cr3+ ions at the B-site, with minimal incorporation into the A-site (Zn2+), resulting in a relatively well-defined doping process [37]. This “magnetic site dilution” disrupts the Cr3+-based magnetic tetrahedral network, progressively weakening the antiferromagnetic order and inducing a spin-glass state. In contrast, in MgCr2O4, due to the closer ionic radii of Mg2+ and Ga3+, Ga3+ ions not only occupy the B-site Cr3+ positions but also partially substitute for A-site Mg2+, leading to A/B-site mixing and cation disorder. This more complex site occupancy results in an anomalous lattice contraction and introduces stronger perturbations to the magnetic structure, thereby suppressing the antiferromagnetic order at lower doping levels and giving rise to more pronounced spin-glass behavior.
According to Ch. Kant et al. [42], both nearest-neighbor ( J 1 ) and next-nearest-neighbor ( J 2 ) magnetic exchange interactions are present among B-site Cr3+ ions, with J 2 generally weaker than J 1 . Their theoretical model indicates that the J 2 is antiferromagnet in MgCr2O4 and ZnCr2O4 while transitioning to ferromagnetic in highly structurally distorted analogs like CdCr2O4. These findings are corroborated by previous reports on Zn(Cr1−xGax)2O4 [37]. In Mg(Cr1−xGax)2O4, the structural distortion may be strong enough to trigger the change of the magnetic interactions. A ferromagnetic J 2 might be induced with the sample of x = 0.15 , leading to the observed ferromagnetic behavior in this system.

3.2. MgCr2O4 Doped with Cd

X-ray powder diffraction experiments were conducted at room temperature on polycrystalline Mg(Cr1−xCdx)2O4( x = 0 , 0.05, 0.10, 0.15, and 0.20). As shown in Figure 5, the diffraction patterns confirm phase purity for all samples, with no detectable secondary phases. Rietveld refinement of the X-ray data reveals crystallization in the cubic spinel structure with space group Fd 3 ¯ m (No. 227). As shown in Figure 5a, the normalized X-ray diffraction patterns suggest a systematic decrease in the intensity of the (111) reflection with increasing Cd content, indicating the successful Cd2+ substitution into the MgCr2O4 lattice.
Figure 5b presents a representative X-ray Rietveld refinement profile of Mg1−xCdxCr2O4 for x = 0.05 . The full refinement results for all Cd-doped samples are summarized in Figure 6. Figure 6a shows good agreement between the actual Cd concentrations and the nominal doping levels, confirming the accuracy of the composition control. Figure 6d displays the average ionic radius at the A-site, derived from the actual Cd doping levels. The radius increases linearly with doping, indicating effective substitution of Mg2+ by the larger Cd2+ ions. There is also an implication that no Cd ions take the site of Cr. Figure 6b,c present the refined lattice parameter a, along with the interatomic distances d 1 (Cr–Cr) and d 2 (Mg–Mg), respectively. All structural parameters increase monotonically with Cd content, indicating that Cd doping leads to lattice expansion and structural distortion in the MgCr2O4 system.
Figure 7 displays the ZFC and FC susceptibility curves of Mg1−xCdxCr2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20), measured under an external magnetic field of 0.2 T. As temperature decreases, larger bifurcation between ZFC and FC susceptibility curves were observed. A distinct bifurcation emerges at the spin-glass transition temperature ( T f ), which becomes more pronounced with higher Cd content. As shown in Figure 8a, T N decreases from 12.6 K to 6 K with increasing Cd doping. Figure 8b shows that the Weiss temperature ( θ CW ), obtained from Curie–Weiss fitting, decreases with increasing x in the range 0 x 0.125 but slightly increases for 0.125 x 0.20 . This suggests that the suppression of antiferromagnetic interactions saturates near x = 0.125 . Figure 8c shows that the frustration factor f = | θ CW | / T N reaches a minimum at x = 0.125 , indicating maximal suppression of magnetic frustration. The effective magnetic moment ( μ eff ) remains nearly constant across all doping levels. Previous investigations on Cd-doped ZnCr2O4 have demonstrated that the introduction of Cd2+ ions not only reduces the T N but also effectively suppresses long-range antiferromagnetic ordering, leading to a spin-glass-like state [38]. Similarly, in the present study on MgCr2O4, Cd2+ doping exhibits a comparable trend, with T N decreasing and AFM order gradually replaced by spin-glass behavior. This consistency across both Zn and Mg based spinels underscores the crucial role of A-site non-magnetic ion substitution in disrupting magnetic interactions and tuning the magnetic ground state in geometrically frustrated spinel systems.
Unlike Ga doping, XRD refinement confirms that Cd2+ substitutes exclusively at the A-site and does not affect the magnetic Cr3+ ions at the B-site. The χ ( T ) data reveal that even at x = 0.05 , Cd doping induces a magnetic transition from antiferromagnetic to spin-glass behavior, which persists even at high doping concentrations, similar to the previously reported behavior of Zn1−xCdxCr2O4 [38]. This early transition is likely driven by the large ionic radius mismatch (21 pm) between Cd2+ and Mg2+, compared to the smaller mismatches between Ga3+ and Cr3+ (0.5 pm) or Mg2+ (10 pm). The pronounced ionic size mismatch induces lattice distortions, which mediate the antiferromagnetic-to-spin-glass crossover at dilute doping concentrations. In studies on ZnCr2O4 [38], Ga3+ doping has been found to dilute the Cr3+ sublattice, introducing site disorder that suppresses the long-range antiferromagnetic ordering. In contrast, Cd2+ doping induces bond disorder by randomly modifying Cr–Cr exchange interactions, thereby promoting a transition into a spin-glass state. Given the high degree of similarity between MgCr2O4 and ZnCr2O4, it is reasonable to extend this interpretation to the MgCr2O4 system. Therefore, Cd2+ doping in MgCr2O4 effectively suppresses magnetic ordering while preserving a corner sharing tetrahedral lattice composed entirely of magnetic Cr3+ ions, making it an ideal candidate for investigating the origin of resonance magnetic excitations in MgCr2O4.

4. Conclusions

In summary, we successfully synthesized a series of Ga and Cd-doped MgCr2O4 spinel oxides via solid-state reaction. Doping with Ga3+ or Cd2+ ions preserved the cubic spinel structure of the parent MgCr2O4 system. Structural refinement of the XRD data reveals an anomalous decrease in the lattice constant of MgCr2O4 with increasing Ga3+ doping concentration. This behavior is attributed to cation site disorder and structural distortions induced by doping. Furthermore, NPD refinement performed on the x = 0.20 sample confirms that Ga3+ ions not only substitute for Cr3+ at the B-site but also partially occupy the A-site positions originally held by Mg2+, resulting in a complex composition with the chemical formula (Mg1−xGax)(Cr1−yGay)2O4. This provides direct experimental evidence for the dual-site occupation of Ga3+ in the MgCr2O4 lattice. Magnetic susceptibility measurements show that Ga3+ doping progressively transforms the system from an antiferromagnetic state into a spin-glass-like state in the x = 0 0.10 range and completely suppresses the antiferromagnetic transition at x = 0.20 , likely due to magnetic site dilution by Ga3+. In contrast, Cd2+ doping leads to lattice expansion, and Rietveld refinements confirm that Cd2+ exclusively substitutes for Mg2+ at the A-sites. Magnetic measurements demonstrate that even low levels of Cd2+ doping induce spin-glass-like behavior, with complete suppression of the antiferromagnetic transition at 10% doping level. This is attributed to lattice distortions caused by the large ionic radius of Cd2+, which disrupt magnetic exchange paths and destabilize long-range order. Accordingly, considering both the preserved Cr3+ magnetic tetrahedral lattice and the enhanced suppression of magnetic order, 10% Cd-doped MgCr3+ serves as an ideal material platform for future inelastic neutron scattering studies aimed at isolating and understanding the origin of resonant magnetic excitations. This work provides both sample materials and preliminary research foundations for understanding the origin of the resonant magnetic excitation modes in MgCr3+.

Author Contributions

Conceptualization, E.F. and Z.H.; investigation, F.Z., D.C., W.Z., and Z.F.; methodology, F.Z., Z.H., D.C., X.W., and Z.F.; resources, H.G., L.W., W.L., X.L., and L.H.; writing—original draft preparation, F.Z., Z.H., X.W., and P.Z.; writing—review and editing, E.F. and Y.Z.; funding acquisition, E.F. and Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

The project was supported by National Natural Science Foundation of China, No. 12375298, the Large Scientific Facility Open Subject of Songshan Lake, Dongguan, Guangdong, KFKT2022B05, and the China Postdoctoral Science Foundation under Grant No. 2024M751754.

Acknowledgments

This research used resources at the China Spallation Neutron Source (CSNS). We thank Guang Wang for supporting the laboratory work. We also acknowledge OpenAI for providing access to the GPT-4 language model, which was utilized for language refinement of the manuscript.

Conflicts of Interest

The authors declare no competing financial interests. Correspondence and requests for materials should be addressed to Erxi Feng.

Appendix A

Table A1. Wyckoff positions and atomic positions of each atom of Mg(Cr1−xGax)2O4 (x = 0.20) obtained by neutron powder diffraction.
Table A1. Wyckoff positions and atomic positions of each atom of Mg(Cr1−xGax)2O4 (x = 0.20) obtained by neutron powder diffraction.
AtomWyckoff Positionxyz
O32e0.260040.260040.26004
Cr16d0.500000.500000.50000
Ga(1)16d0.500000.500000.50000
Mg8a0.125000.125000.12500
Ga(2)8a0.125000.125000.12500
Table A2. Refined lattice constant a, site occupancies of Cr and Ga at the B-sites, site occupancies of Mg and Ga at the A-sites, and Cr Cr and Mg Mg bond lengths for the Mg(Cr1−xGax)2O4 series.
Table A2. Refined lattice constant a, site occupancies of Cr and Ga at the B-sites, site occupancies of Mg and Ga at the A-sites, and Cr Cr and Mg Mg bond lengths for the Mg(Cr1−xGax)2O4 series.
xa (Å) d Mg Mg ( Å ) d Cr Cr ( Å ) B-Site Occupancy (%)A-Site Occupancy (%)
CrGaMgGa
08.33333 (3)3.60844 (0)2.94628 (0)100.0 (1)0100.0 (1)0
0.058.33293 (17)3.60744 (8)2.94546 (7)95.7 (1)4.3 (1)98.6 (1)1.4 (1)
0.108.33111 (21)3.60678 (9)2.94492 (8)92.9 (1)7.1 (1)94.2 (1)5.8 (1)
0.158.32714 (19)3.60546 (9)2.94385 (7)92.8 (1)7.2 (1)84.4 (1)15.6 (1)
0.208.32183 (13)3.60275 (6)2.94163 (5)92.3 (2)7.7 (2)75.4 (2)24.6 (2)
Table A3. Magnetic phase transition temperature T N , Weiss temperature θ CW , effective magnetic moment μ e f f , and frustration factor f for Mg(Cr1−xGax)2O4.
Table A3. Magnetic phase transition temperature T N , Weiss temperature θ CW , effective magnetic moment μ e f f , and frustration factor f for Mg(Cr1−xGax)2O4.
Mg(Cr1−xGax)2O4 T N (K) θ CW (K) μ eff ( μ B )f
012.6−516.498 (11)4.04 (18)40.77 (9)
0.058.0−444.311 (25)3.92 (27)55.23 (31)
0.106.1−420.684 (25)3.93 (28)69.46 (41)
0.153.1−398.092 (23)3.93 (27)129.67 (75)
0.20<2.0−370.944 (19)3.90 (25)>123.49 (64)
Table A4. Refined lattice constant a, site occupancies of Mg and Cd at the A-sites, and Cr Cr and Mg Mg bond lengths for the Mg1−xCdxCr2O4 series.
Table A4. Refined lattice constant a, site occupancies of Mg and Cd at the A-sites, and Cr Cr and Mg Mg bond lengths for the Mg1−xCdxCr2O4 series.
xa (Å) d Cr Cr ( Å ) d Mg Mg ( Å ) A-Site Occupancy (%)
MgCd
08.33333 (3)2.94628 (0)3.60844 (0)100.0 (1)0
0.058.34472 (4)2.9503 (0)3.61337 (0)95.3 (2)4.7 (2)
0.108.35502 (9)2.95395 (4)3.61783 (4)91.6 (2)8.4 (2)
0.158.35776 (13)2.95491 (5)3.61902 (6)87.6 (5)12.4 (5)
0.208.38121 (14)2.96321 (6)3.62917 (7)80.9 (5)19.1 (3)
Table A5. Magnetic phase transition temperature T N , Weiss temperature θ CW , effective magnetic moment μ e f f , and frustration factor f for Mg1−xCdxCr2O4.
Table A5. Magnetic phase transition temperature T N , Weiss temperature θ CW , effective magnetic moment μ e f f , and frustration factor f for Mg1−xCdxCr2O4.
Mg1− xCdxCr2O4 T N (K) θ CW (K) μ eff ( μ B )f
012.667−516.5 (11)4.04 (18)40.77 (9)
0.0511.004−394.3 (19)4.07 (24)35.84 (17)
0.109.616−191.8 (17)4.10 (22)19.95 (17)
0.156.579−75.4 (15)4.09 (21)11.46 (23)
0.206.015−107.2 (13)3.08 (20)17.83 (21)

References

  1. Balents, L. Spin liquids in frustrated magnets. Nature 2010, 464, 199–208. [Google Scholar] [CrossRef] [PubMed]
  2. Moessner, R.; Ramirez, A.P. Geometrical frustration. Phys. Today 2006, 59, 24–29. [Google Scholar] [CrossRef]
  3. Rao, X.; Hussain, G.; Huang, Q.; Chu, W.J.; Li, N.; Zhao, X.; Dun, Z.; Choi, E.S.; Asaba, T.; Chen, L.; et al. Survival of itinerant excitations and quantum spin state transitions in YbMgGaO4 with chemical disorder. Nat. Commun. 2021, 12, 4949. [Google Scholar] [CrossRef]
  4. Helton, J.S.; Matan, K.; Shores, M.P.; Nytko, E.A.; Bartlett, B.M.; Qiu, Y.; Nocera, D.G.; Lee, Y.S. Dynamic scaling in the susceptibility of the spin-1/2 kagome lattice antiferromagnet Herbertsmithite. Phys. Rev. Lett. 2010, 104, 147201. [Google Scholar] [CrossRef]
  5. Khuntia, P.; Velazquez, M.; Barthélemy, Q.; Bert, F.; Kermarrec, E.; Legros, A.; Bernu, B.; Messio, L.; Zorko, A.; Mendels, P. Gapless ground state in the archetypal quantum kagome antiferromagnet ZnCu3(OH)6Cl2. Nat. Phys. 2020, 16, 469–474. [Google Scholar] [CrossRef]
  6. Han, T.H.; Helton, J.S.; Chu, S.; Nocera, D.G.; Rodriguez-Rivera, J.A.; Broholm, C.; Lee, Y.S. Fractionalized excitations in the spin-liquid state of a kagome-lattice antiferromagnet. Nature 2012, 492, 406–410. [Google Scholar] [CrossRef]
  7. Schweika, W.; Valldor, M.; Lemmens, P. Approaching the ground state of the kagomé antiferromagnet. Phys. Rev. Lett. 2007, 98, 067201. [Google Scholar] [CrossRef]
  8. Stewart, J.R.; Ehlers, G.; Mutka, H.; Fouquet, P.; Payen, C.; Lortz, R. Spin dynamics, short-range order, and spin freezing in Y0.5Ca0.5BaCo4O7. Phys. Rev. B—Condensed Matter Mater. Phys. 2011, 83, 024405. [Google Scholar] [CrossRef]
  9. Okamoto, Y.; Nohara, M.; Aruga-Katori, H.; Takagi, H. Spin-liquid state in the S = 1/2 hyperkagome antiferromagnet Na4Ir3O8. Phys. Rev. Lett. 2007, 99, 137207. [Google Scholar] [CrossRef] [PubMed]
  10. Gardner, J.S.; Gingras, M.J.P.; Greedan, J.E. Magnetic pyrochlore oxides. Rev. Mod. Phys. 2010, 82, 53–107. [Google Scholar] [CrossRef]
  11. Siddharthan, R.; Shastry, B.S.; Ramirez, A.P.; Hayashi, A.; Cava, R.J.; Rosenkranz, S. Ising pyrochlore magnets: Low-temperature properties, “ice rules,” and beyond. Phys. Rev. Lett. 1999, 83, 1854. [Google Scholar] [CrossRef]
  12. Gingras, M.J.P.; McClarty, P.A. Quantum spin ice: A search for gapless quantum spin liquids in pyrochlore magnets. Rep. Prog. Phys. 2014, 77, 056501. [Google Scholar] [CrossRef]
  13. Bramwell, S.T.; Gingras, M.J.P. Spin ice state in frustrated magnetic pyrochlore materials. Science 2001, 294, 1495–1501. [Google Scholar] [CrossRef] [PubMed]
  14. Matsuhira, K.; Hinatsu, Y.; Tenya, K.; Amitsuka, H.; Sakakibara, T. Low-temperature magnetic properties of pyrochlore stannates. J. Phys. Soc. Jpn. 2002, 71, 1576–1582. [Google Scholar] [CrossRef]
  15. Shapiro, M.C.; Riggs, S.C.; Stone, M.B.; de la Cruz, C.R.; Chi, S.; Podlesnyak, A.A.; Fisher, I.R. Structure and magnetic properties of the pyrochlore iridate Y2Ir2O7. Phys. Rev. B 2012, 85, 214434. [Google Scholar] [CrossRef]
  16. Xu, J.; Anand, V.K.; Bera, A.K.; Frontzek, M.; Abernathy, D.L.; Casati, N.; Siemensmeyer, K.; Lake, B. Magnetic structure and crystal-field states of the pyrochlore antiferromagnet Nd2Zr2O7. Phys. Rev. B 2015, 92, 224430. [Google Scholar] [CrossRef]
  17. Das, D.; Ghosh, S. Density functional theory based comparative study of electronic structures and magnetic properties of spinel ACr2O4 (A = Mn, Fe, Co, Ni) compounds. J. Phys. D Appl. Phys. 2015, 48, 425001. [Google Scholar] [CrossRef]
  18. Ali, A.; Singh, Y. A magnetocaloric study on the series of 3d-metal chromites ACr2O4 (A = Mn, Fe, Co, Ni, Cu and Zn). J. Magn. Magn. Mater. 2020, 499, 166253. [Google Scholar] [CrossRef]
  19. Kassem, M.A.; El-Fadl, A.A.; Nashaat, A.M.; Nakamura, H. Structure, optical and varying magnetic properties of insulating MCr2O4 (M = Co, Zn, Mg and Cd) nanospinels. J. Alloys Compd. 2019, 790, 853–862. [Google Scholar] [CrossRef]
  20. Moureen, C.K.; Stephanie, L.M.; Lucy, E.D.; Ram, S.; Matthew, R.S.; Daniel, P.S.; Katharine, P.; Joan, S. Structural ground states of (A,A)Cr2O4 (A = Mg, Zn; A = Co, Cu) spinel solid solutions: Spin-Jahn-Teller and Jahn-Teller effects. Phys. Rev. B 2014, 89, 174410. [Google Scholar]
  21. Ohgushi, K.; Okimoto, Y.; Ogasawara, T.; Miyasaka, S.; Tokura, Y. Magnetic, optical, and magnetooptical properties of spinel-type ACr2X4 (A = Mn, Fe, Co, Cu, Zn, Cd; X = O, S, Se). J. Phys. Soc. Jpn. 2008, 77, 034713. [Google Scholar] [CrossRef]
  22. Lau, G.C.; Ueland, B.G.; Freitas, R.S.; Dahlberg, M.L.; Schiffer, P.; Cava, R.J. Magnetic characterization of the sawtooth-lattice olivines ZnL2S4 (L = Er, Tm, Yb). Phys. Rev. B 2006, 73, 012413. [Google Scholar] [CrossRef]
  23. Zhandun, V.S. The magnetic, electronic, optical, and structural properties of the AB2O4 (A = Mn, Fe, Co; B = Al, Ga, In) spinels: Ab initio study. J. Magn. Magn. Mater. 2021, 533, 168015. [Google Scholar] [CrossRef]
  24. Ortega-San-Martin, L.; Williams, A.J.; Gordon, C.D.; Klemme, S.; Attfield, J.P. Low temperature neutron diffraction study of MgCr2O4 spinel. J. Phys. Condens. Matter 2008, 20, 104238. [Google Scholar] [CrossRef]
  25. Dutton, S.E.; Huang, Q.; Tchernyshyov, O.; Broholm, C.L.; Cava, R.J. Sensitivity of the magnetic properties of the ZnCr2O4 and MgCr2O4 spinels to nonstoichiometry. Phys. Rev. B 2011, 83, 064407. [Google Scholar] [CrossRef]
  26. Bai, X.; Paddison, J.A.M.; Kapit, E.; Koohpayeh, S.M.; Wen, J.-J.; Dutton, S.E.; Savici, A.T.; Kolesnikov, A.I.; Granroth, G.E.; Broholm, C.L.; et al. Magnetic excitations of the classical spin liquid MgCr2O4. Phys. Rev. Lett. 2019, 122, 097201. [Google Scholar] [CrossRef] [PubMed]
  27. Zhou, H.D.; Zhao, Z.Y.; Sun, X.F.; Suarez, M.N.; Rivas-Murias, B.; Tsurkan, V.; Deisenhofer, J.; Zapf, V.S.; Rivadulla, F. Low-temperature spin excitations in frustrated ZnCr2O4 probed by high-field thermal conductivity. Phys. Rev. B 2013, 87, 174436. [Google Scholar] [CrossRef]
  28. Matsuda, M.; Ueda, H.; Kikkawa, A.; Tanaka, Y.; Katsumata, K.; Narumi, Y.; Inami, T.; Ueda, Y.; Lee, S.-H. Spin–lattice instability to a fractional magnetization state in the spinel HgCr2O4. Nat. Phys. 2007, 3, 397–400. [Google Scholar] [CrossRef]
  29. Ueda, H.; Mitamura, H.; Goto, T.; Ueda, Y. Field-induced metamagnetic transitions in S = 3/2 Heisenberg antiferromagnets MCr2O4. Prog. Theor. Phys. Suppl. 2005, 159, 256–259. [Google Scholar] [CrossRef]
  30. Rudolf, T.; Kant, C.; Mayr, F.; Hemberger, J.; Tsurkan, V.; Loidl, A. Spin-phonon coupling in antiferromagnetic chromium spinels. New J. Phys. 2007, 9, 76. [Google Scholar] [CrossRef]
  31. Shaked, H.; Hastings, J.M.; Corliss, L.M. Magnetic structure of magnesium chromite. Phys. Rev. B 1970, 1, 3116. [Google Scholar] [CrossRef]
  32. Suzuki, H.; Tsunoda, Y. Spinel-type frustrated system MgCr2O4 studied by neutron scattering and magnetization measurements. J. Phys. Chem. Solids 2007, 68, 2060–2063. [Google Scholar] [CrossRef]
  33. Gao, S.; Guratinder, K.; Stuhr, U.; White, J.S.; Mansson, M.; Roessli, B.; Fennell, T.; Tsurkan, V.; Loidl, A.; Hatnean, M.C.; et al. Manifolds of magnetic ordered states; and excitations in the almost Heisenberg pyrochlore antiferromagnet MgCr2O4. Phys. Rev. B 2018, 97, 134430. [Google Scholar] [CrossRef]
  34. Tomiyasu, K.; Suzuki, H.; Toki, M.; Itoh, S.; Matsuura, M.; Aso, N.; Yamada, K. Molecular spin resonance in the geometrically frustrated magnet MgCr2O4 by inelastic neutron scattering. Phys. Rev. Lett. 2008, 101, 177401. [Google Scholar] [CrossRef]
  35. Yong, W.; Botis, S.; Shieh, S.R.; Shi, W.; Withers, A.C. Pressure-induced phase transition study of magnesiochromite (MgCr2O4) by Raman spectroscopy and X-ray diffraction. Phys. Earth Planet. Inter. 2012, 196, 75–82. [Google Scholar] [CrossRef]
  36. Lee, S.-H.; Ratcliff, W., II; Huang, Q.; Kim, T.H.; Cheong, S.-W. Néel to spin-glass-like phase transition versus dilution in geometrically frustrated ZnCr2−2xGa2xO4. Phys. Rev. B 2008, 77, 014405. [Google Scholar] [CrossRef]
  37. Yu, J.; Wang, L.; Ying, Y.; Zheng, J.; Li, W.; Qiao, L.; Cai, W.; Li, J.; Zhang, L.; Che, S. Magnetic Exchange Interactions in Geometrically Frustrated Antiferromagnet of ZnCr2−xGaxO4. J. Supercond. Nov. Magn. 2019, 32, 1095–1098. [Google Scholar] [CrossRef]
  38. Martinho, H.; Moreno, N.O.; Sanjurjo, J.A.; Rettori, C.; Âa-Adeva, A.J.G.; Huber, D.L.; Oseroff, S.B.; Ratcliff, W., II; Cheong, S.-W.; Pagliuso, P.G.; et al. Magnetic properties of the frustrated antiferromagnetic spinel ZnCr2O4 and the spin-glass Zn1−xCdxCr2O4 (x = 0.05, 0.10). Phys. Rev. B 2001, 64, 024408. [Google Scholar] [CrossRef]
  39. Ohta, H.; Okubo, S.; Kikuchi, H.; Ono, S. Millimetre-wave ESR (electron-spin resonance) measurements of frustrated system ZnCr2xGa2−2xO4. Can. J. Phys. 2001, 79, 1387–1391. [Google Scholar] [CrossRef]
  40. Sears, V.F. Neutron scattering lengths and cross sections. Neutron News 1992, 3, 26–37. [Google Scholar] [CrossRef]
  41. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Found. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  42. Kant, C.; Deisenhofer, J.; Tsurkan, V.; Loidl, A. Magnetic susceptibility of the frustrated spinels ZnCr2O4, MgCr2O4 and CdCr2O4. J. Phys. Conf. Ser. 2010, 200, 032032. [Google Scholar] [CrossRef]
  43. Nassar, L.S.; Lane, H.; Haberl, B.; Graves-Brook, M.; Winn, B.; Koohpayeh, S.M.; Mourigal, M. Pressure control of magnetic order and excitations in the pyrochlore antiferromagnet MgCr2O4. Phys. Rev. B 2024, 109, 064415. [Google Scholar] [CrossRef]
  44. Niu, X.; Liu, C.; Zong, B. Investigation on the microstructure and magnetic properties of Mg(Ga2−XFex)O4 Spinel Ferrites. J. Mater. Sci. Mater. Electron. 2023, 34, 118. [Google Scholar] [CrossRef]
  45. Das, A.; Ranaut, D.; Pal, P.; Pal, R.; Mandal, P.; Pal, A.N.; Moulick, S.; Das, M.; Topwal, D.; Mukherjee, K.; et al. Tuning of magnetic frustration and emergence of a magnetostructural transition in Mn1−XCdxCr2O4. Phys. Rev. B 2023, 108, 064426. [Google Scholar] [CrossRef]
  46. Koohpayeh, S.M.; Wen, J.-J.; Mourigal, M.; Dutton, S.E.; Cava, R.J.; Broholm, C.L.; McQueen, T.M. Optical floating zone crystal growth and magnetic properties of MgCr2O4. J. Cryst. Growth 2013, 384, 39–43. [Google Scholar] [CrossRef]
Figure 1. (a) The crystal structure of the MgCr2O4 compound; (b) the XRD patterns of the samples Mg(Cr1−xGax)2O4 (x = 0, 0.05, 0.10, 0.15, and 0.20); the combined Rietveld refinement of (c) XRD and (d) NPD of the sample Mg(Cr1−xGax)2O4 (x = 0.20), The Wyckoff positions of the atoms are in Table A1 of the Appendix A.
Figure 1. (a) The crystal structure of the MgCr2O4 compound; (b) the XRD patterns of the samples Mg(Cr1−xGax)2O4 (x = 0, 0.05, 0.10, 0.15, and 0.20); the combined Rietveld refinement of (c) XRD and (d) NPD of the sample Mg(Cr1−xGax)2O4 (x = 0.20), The Wyckoff positions of the atoms are in Table A1 of the Appendix A.
Magnetism 05 00019 g001
Figure 2. The occupancy of Ga at the Cr3+ and Mg2+ sites is presented in (a,b); the variations of the lattice constants for all Mg(Cr1−xGax)2O4 samples (c); (d,e) the changes in the Cr-Cr and Mg-Mg bond lengths, respectively; (f) the changes in the average ionic radii (a.i.r.) after Ga incorporation at the Cr3+ and Mg2+ sites. Detailed data can be found in Table A2 of the Appendix A.
Figure 2. The occupancy of Ga at the Cr3+ and Mg2+ sites is presented in (a,b); the variations of the lattice constants for all Mg(Cr1−xGax)2O4 samples (c); (d,e) the changes in the Cr-Cr and Mg-Mg bond lengths, respectively; (f) the changes in the average ionic radii (a.i.r.) after Ga incorporation at the Cr3+ and Mg2+ sites. Detailed data can be found in Table A2 of the Appendix A.
Magnetism 05 00019 g002
Figure 3. Temperature dependence of the DC magnetic susceptibilities χ ( T ) for Mg(Cr1−xGax)2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20) samples under ZFC and FC conditions. The inset shows the inverse magnetic susceptibility and the fitting of the Curie–Weiss law for the undoped sample MgCr2O4 compound.
Figure 3. Temperature dependence of the DC magnetic susceptibilities χ ( T ) for Mg(Cr1−xGax)2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20) samples under ZFC and FC conditions. The inset shows the inverse magnetic susceptibility and the fitting of the Curie–Weiss law for the undoped sample MgCr2O4 compound.
Magnetism 05 00019 g003
Figure 4. Extracted results from the DC magnetic susceptibilities for Mg(Cr1−xGax)2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20). Néel temperature T N (a), Curie–Weiss temperature | θ CW | (b), effective magnetic moment ( μ eff ) (c), and the frustration factor ( f = | θ CW | / T N ) (d) vs. Ga3+ doping level. Detailed data can be found in Table A3 of the Appendix A.
Figure 4. Extracted results from the DC magnetic susceptibilities for Mg(Cr1−xGax)2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20). Néel temperature T N (a), Curie–Weiss temperature | θ CW | (b), effective magnetic moment ( μ eff ) (c), and the frustration factor ( f = | θ CW | / T N ) (d) vs. Ga3+ doping level. Detailed data can be found in Table A3 of the Appendix A.
Magnetism 05 00019 g004
Figure 5. (a) The XRD patterns of the Mg1−xCdxCr2O4 (x = 0, 0.05, 0.10, 0.15, and 0.20) series of samples and (b) the Rietveld refinement plots for Mg1−xCdxCr2O4 (x = 0.05).
Figure 5. (a) The XRD patterns of the Mg1−xCdxCr2O4 (x = 0, 0.05, 0.10, 0.15, and 0.20) series of samples and (b) the Rietveld refinement plots for Mg1−xCdxCr2O4 (x = 0.05).
Magnetism 05 00019 g005
Figure 6. (a) The changes in the occupancy of Mg2+ and Cd2+ ions at the Mg-site, (b) the variation of the lattice constant, (c) the variations of Cr-Cr and Mg-Mg bond lengths, and (d) the variation in the average ionic radii (a.i.r.) at the Mg site. Detailed data can be found in Table A4 of the Appendix A.
Figure 6. (a) The changes in the occupancy of Mg2+ and Cd2+ ions at the Mg-site, (b) the variation of the lattice constant, (c) the variations of Cr-Cr and Mg-Mg bond lengths, and (d) the variation in the average ionic radii (a.i.r.) at the Mg site. Detailed data can be found in Table A4 of the Appendix A.
Magnetism 05 00019 g006
Figure 7. Temperature-dependent ZFC and FC DC magnetic susceptibility χ ( T ) curves for Mg1−xCdxCr2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20) measured under an applied magnetic field of 0.2 T. The inset displays the inverse susceptibility χ 1 ( T ) for Mg0.95Cd0.05Cr2O4 with Curie–Weiss fitting. As Cd2+ doping increases, the TN systematically shifts to lower temperatures, indicating a progressive suppression of AFM interactions. The increased bifurcation between ZFC and FC curves at higher doping levels suggests the emergence of spin-glass-like behavior.
Figure 7. Temperature-dependent ZFC and FC DC magnetic susceptibility χ ( T ) curves for Mg1−xCdxCr2O4 ( x = 0 , 0.05, 0.10, 0.15, and 0.20) measured under an applied magnetic field of 0.2 T. The inset displays the inverse susceptibility χ 1 ( T ) for Mg0.95Cd0.05Cr2O4 with Curie–Weiss fitting. As Cd2+ doping increases, the TN systematically shifts to lower temperatures, indicating a progressive suppression of AFM interactions. The increased bifurcation between ZFC and FC curves at higher doping levels suggests the emergence of spin-glass-like behavior.
Magnetism 05 00019 g007
Figure 8. (a) Magnetic transition temperature T N , (b) Weiss temperature θ CW , (c) effective magnetic moment μ eff , and (d) magnetic frustration coefficient f as a function of Cd2+ doping concentration in Mg1−xCdxCr2O4. Detailed data can be found in Table A5 of the Appendix A.
Figure 8. (a) Magnetic transition temperature T N , (b) Weiss temperature θ CW , (c) effective magnetic moment μ eff , and (d) magnetic frustration coefficient f as a function of Cd2+ doping concentration in Mg1−xCdxCr2O4. Detailed data can be found in Table A5 of the Appendix A.
Magnetism 05 00019 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, F.; He, Z.; Cheng, D.; Ge, H.; Zhang, W.; Wang, X.; Zhou, P.; Luo, W.; Fu, Z.; Liu, X.; et al. Effect of the Non-Magnetic Ion Doping on the Magnetic Behavior of MgCr2O4. Magnetism 2025, 5, 19. https://doi.org/10.3390/magnetism5030019

AMA Style

Zhou F, He Z, Cheng D, Ge H, Zhang W, Wang X, Zhou P, Luo W, Fu Z, Liu X, et al. Effect of the Non-Magnetic Ion Doping on the Magnetic Behavior of MgCr2O4. Magnetism. 2025; 5(3):19. https://doi.org/10.3390/magnetism5030019

Chicago/Turabian Style

Zhou, Fuxi, Zheng He, Donger Cheng, Han Ge, Wenjing Zhang, Xiao Wang, Pengfei Zhou, Wanju Luo, Zhengdong Fu, Xinzhi Liu, and et al. 2025. "Effect of the Non-Magnetic Ion Doping on the Magnetic Behavior of MgCr2O4" Magnetism 5, no. 3: 19. https://doi.org/10.3390/magnetism5030019

APA Style

Zhou, F., He, Z., Cheng, D., Ge, H., Zhang, W., Wang, X., Zhou, P., Luo, W., Fu, Z., Liu, X., Wu, L., He, L., Zhao, Y., & Feng, E. (2025). Effect of the Non-Magnetic Ion Doping on the Magnetic Behavior of MgCr2O4. Magnetism, 5(3), 19. https://doi.org/10.3390/magnetism5030019

Article Metrics

Back to TopTop