Next Article in Journal
Universal Dependence of Nuclear Spin Relaxation on the Concentration of Paramagnetic Centers in Nano- and Microdiamonds
Previous Article in Journal
Damage Evolution Characteristics of Back-Filling Concrete in Gob-Side Entry Retaining Subjected to Cyclical Loading
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two Magnetic Orderings and a Spin–Flop Transition in Mixed Valence Compound Mn3O(SeO3)3

1
School of Materials, Sun Yat-sen University, Guangzhou 510275, China
2
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China
3
Department of Chemistry, University of Science and Technology of China, Hefei 230026, China
*
Author to whom correspondence should be addressed.
Materials 2022, 15(16), 5773; https://doi.org/10.3390/ma15165773
Submission received: 7 July 2022 / Revised: 9 August 2022 / Accepted: 18 August 2022 / Published: 21 August 2022
(This article belongs to the Topic Materials for Energy Harvesting and Storage)

Abstract

:
A mixed-valence manganese selenite, Mn3O(SeO3)3, was successfully synthesized using a conventional hydrothermal method. The three-dimensional framework of this compound is composed of an MnO6 octahedra and an SeO3 trigonal pyramid. The magnetic topological arrangement of manganese ions shows a three-dimensional framework formed by the intersection of octa-kagomé spin sublattices and staircase-kagomé spin sublattices. Susceptibility, magnetization and heat capacity measurements confirm that Mn3O(SeO3)3 exhibits two successive long-range antiferromagnetic orderings with TN1~4.5 K and TN2~45 K and a field-induced spin–flop transition at a critical field of 4.5 T at low temperature.

1. Introduction

Mixed valence transition metal (TM) oxides with three-dimensional electronic configurations are of great significance in the fields of materials chemistry, electrochemical energy and condensed matter physics due to their diverse crystal structures and electronic configurations [1,2]. From the ancient application of Fe3O4 in the compass to today’s copper-based high temperature superconducting materials, mixed valence TM oxides exhibit exciting and unusual chemical and physical behaviors, including high-temperature superconductors [3], colossal magnetoresistance [4], ion deintercalation [5], metal-insulator transition [6], electrocatalysis/photocatalysis [7,8], etc. More specifically, copper oxides with bidimensional characters, together with the mixed valency of Cu+/Cu2+ or Cu2+/Cu3+, are responsible for superconducting properties [9,10]. The ferromagnetic (FM) material La0.67Sr0.33MnO3 exhibits metallic conductivity due to the Zener double exchange mechanism between Mn3+ and Mn4+ ions, but BaFe12O19 (also an FM material) is insulative due to the limitation of the ratio of Fe2+ and Fe3+ ions [11]. Compound K2Cr8O16 (hollandite), with a rare Cr3+/Cr4+ mixed valence state, exhibits a metal-insulator transition in a FM state [12]. The transition metal valence state of cathode material LiMO2 (M = Mn, Co, Ni) will switch back and forth between M2+ and M4+ during charging and discharging processes in Li-ion batteries [13,14]. X. Yu et al. reported the experimental observation of skyrmionic bubbles with various topological lattices in colossal magnetoresistive manganite La1−xSrxMnO3 [15]. In order to discover new materials with unusual physical/chemical properties, it is necessary to explore new mixed-valence transition metal compounds. The compound Mn3O(SeO3)3 (MnIIMnIII2O(SeO3)3) was first reported by Wildner [16]. Structure analysis confirmed that this compound shows a channel structure with a three-dimensional magnetic topological framework formed by the intersection of octa-kagomé spin sublattices and staircase-kagomé spin sublattices; however, there are few studies regarding its magnetic properties. In this paper, we report the discovery of a mixed valence manganese selenate Mn3O(SeO3)3. Magnetic measurements indicate that this compound possesses two successive antiferromagnetic (AFM) transitions at low-temperature. Moreover, a spin–flop transition is observed at 2 K with an applied magnetic field of ~4.5 T.

2. Experimental Section

2.1. Synthesis of Mn3O(SeO3)3

Single crystals of Mn3O(SeO3)3 were obtained using a conventional hydrothermal method. A mixture of 2 mmol Mn(NO3)2·xH2O (3 N, 0.3943 g), 2 mmol LiI (2 N, 0.2704 g), 1 mmol SeO2 (4 N, 0.1110 g) and 5 mL deionized water was sealed in an autoclave equipped with a Teflon liner (28 mL). The autoclave was gradually heated to 230 °C at a rate of 1 °C/min, held for 4 days and then naturally cooled to room temperature. The product contained the desired black noodle-like crystals with a 90% yield. The crystals’ sizes and morphologies were characterized using a stereomicroscope and field emission scanning electron microscopy (FE-SEM, SU8100, Hitachi, Tokyo, Japan). Figure S1 shows images of the crystals under the stereomicroscope and FE-SEM. It can be observed that the crystal size is approximately 0.3 × 0.08 × 0.05 mm. The product’s impurities were manually removed under a microscope. Powdered samples were prepared for physical measurement by crushing small single crystals; purity was confirmed by powder X-ray diffraction (XRD) analysis (Figure 1). Moreover, the reagent LiI acted as a mineralizer, as the quality of crystals was unsatisfactory without it.

2.2. Methods

XRD patterns were collected using a Bruker D8 diffractometer with Cu-Kα radiation (λ~1.5418 Å) at room temperature. Rieltveld refinement was performed using GSAS-EXPGUI software [17]. Refined crystal structures were analyzed using VESTA software [18]. Furthermore, element analysis was observed using FE-SEM with an X-ray energy-dispersive spectrometer (EDS). EDS analysis confirmed the molar ratio of Mn/Se as 3.1/2.0, which is in good agreement with the X-ray structure analysis. Thermogravimetric analysis (TGA) of Mn3O(SeO3)3 was collected on NETZSCH STA 449C instruments with an Al2O3 crucible from 50 to 900 °C at a rate of 10 °C/min under N2 atmosphere.
Magnetic measurements of a powdered sample of Mn3O(SeO3)3 were performed using a PPMS (Quantum Design, San Diego, CA, USA). The powdered sample (20.6 mg) was placed in a plastic capsule, which was suspended in a copper tube slot. Magnetic susceptibility was measured at 0.1 T from 2 to 300 K. Magnetization was measured at different temperatures at applied field from 0 to 9 T. Heat capacity was measured with the same PPMS system at zero field and determined using a relaxation method on a 5.6 mg sample.

3. Results and Discussion

The structure of compound Mn3O(SeO3)3 was first reported by Wildner [16]. Mn3O(SeO3)3 crystallizes in the monoclinic system with the space group C2/m. As shown in Figure S2, both Mn and Se atoms have three crystallographic sites. The oxidation state is +2 for Mn1 and +3 for Mn2/Mn3. All manganese atoms are coordinated by six oxygen atoms forming MnO6 distorted octahedra; Mn–O bond lengths range from 2.100(1) to 2.361(8) Å for Mn12+O6 octahedra and from 1.854(2) to 2.310(6) Å for Mn23+O6 and Mn33+O6, respectively. In other words, the degree of distortion for Mn12+O6 octahedra is smaller than that of Mn23+O6 and Mn33+O6. This is due to the Mn3+ (t2g3eg1) octahedron with a remarkable Jahn–Teller effect, which may induce a larger structure distortion than Mn2+ (t2g3eg2). All selenium atoms are in trigonal pyramid geometry with a stereoactive lone pair of 4 s2 in Se4+ ions; the Se-O bond lengths are approximately 1.70 Å. It should be noted that Se1/Se2/Se3 atoms are surrounded by 4/5/6 manganese atoms with a Se-O-Mn route, respectively. These 4/5/6 manganese atoms contain two Mn2+ atoms and 2/3/4 Mn3+ atoms, respectively.
As shown in Figure 2, Mn3O(SeO3)3 shows a tunnel structure along the b-axis, in which the framework is constituted by MnO6 octahedra and SeO3 trigonal pyramids. Mn1O6 octahedra share their edges (O5–O6) to form uniform [-Mn1-] chains along the b-axis. Mn3+O6 octahedra are interconnected via edge-sharing oxygen atoms, forming a two-dimensional [-Mn3+-] layered structure parallel to (001). The detailed linkage mode between manganese ions is shown in Figure 3. Two Mn2O6 octahedra connect to each other by edge-sharing oxygen atoms (O4–O4) to form a [Mn2O10] dimer along the a-axis. The Mn2-O4-Mn2 angle is 101.34(9)°. Mn3O6 octahedra are interconnected by corner-sharing oxygen atoms (O7) to form uniform [-Mn3-] chains along the b-axis. One Mn2O6 octahedron and two Mn3O6 octahedra are connected in an isosceles triangle configuration. The neighbored [-Mn3+-] layers are separated by [-Mn1-] chains and SeO3 trigonal pyramids. Furthermore, we noted that Mn1O6 octahedra are interconnected with Mn2O6 octahedra via conner-sharing O5 atoms, but Mn1O6 and Mn3O6 octahedra are connected by SeO3 groups in the manner of Mn1-O-Se-O-Mn3. After removing the nonmagnetic SeO32− groups, the topological arrangement of magnetic Mn ions is a three-dimensional framework (Figure 2b). Mn3+ ions form a two-dimensional octa-kagomé lattice parallel to (001) (Figure 2c). The adjacent octa-kagomé layers are connected by [-Mn1-] chains. It is significant that there is a staircase-kagomé lattice composed of Mn2+ and Mn3+ parallel to (100) in the magnetic topological framework (Figure 2d). The shortest Mn–Mn distance in both [-Mn1-] and [-Mn3-] chains is 3.332(9) Å. However, the detailed connection mode of MnO6 octahedra in [-Mn1-] chains are edge-sharing, whereas in [-Mn3-] chains it is corner-sharing. The Mn-O-Mn angles in [-Mn1-] and [-Mn3-] chains are 89.77(1)°/102.90(8)° and 127.61(5)°, respectively. The Mn2--Mn2 and Mn2–Mn3 distances in the octa-kagomé lattice are 3.150(0) Å and 3.069(6) Å, respectively, whereas the Mn1–Mn2 distance is 3.902(1) Å.
Figure 4a shows the temperature dependence of magnetic susceptibility χ(T) of Mn3O(SeO3)3 measured at 0.1 T. Magnetic susceptibility increases with decreasing temperature; two peaks can be observed at TN1~4.5 K and TN2~45 K., showing AFM transitions. At high temperature (80–300 K) inverse susceptibility χ−1(T) follows the Curie–Weiss law with a Weiss temperature of θ = −8.89 K and a Curie constant of C = 11.03 emu·mol−1·K. The effective magnetic moment is calculated to be μeff = 5.42(3) μB, obtained by μ2eff = 8C/n, where n = 3. This value of μeff is slightly smaller than the spin-only value of 5.91(6) μB for Mn2+ (3d5, high spin) and larger than the spin-only value of 4.89(9) μB for Mn3+ (3d4, high spin). As Mn ions are mixed-valent, the theoretical magnetic moment of the titled compound is μtheo = 5.25(9) μB obtained by the equation μ2eff = [μ2eff (Mn2+) + 2μ2eff (Mn3+)]/3. The value of μeff is quite close to that of μtheo, confirming that Mn ions in the structure are mixed valence. The negative value of θ suggests the presence of dominative AFM interactions between neighboring Mn ions. Figure S3 shows the χT-T curve, in which the value of χT decreases with decreasing temperature, which is characteristic of typical AFM interactions. As shown in Figure 4b, the heat capacity data of Mn3O(SeO3)3 show a λ-type peak at T~45 K and a corner-type transition at 4.5 K, providing concrete evidence for the two long-range magnetic orderings observed in the magnetic susceptibility curves.
To further investigate the magnetic properties of the system, magnetization (M) as a function of applied field (H) was observed at 30 K and 2 K. As shown in Figure 5, at 30 K, magnetization increased linearly with increasing field, and did not saturate at 9 T. Furthermore, no hysteresis or remanent magnetization was observed. These features of the MH curve suggest that the magnetic anomaly at T~45 K is the onset of an AFM ordering. At 2 K, the magnetization (M) shows a linear increase in magnetization at low field, indicative of a characteristic AFM ground state. A clear change in slope in the magnetization is observed at approximately 4.5 T, indicating field-induced magnetic transition. Furthermore, no hysteresis can be observed on the MH curve.
It is well-known that the magnetic properties of solid magnets are strongly related to their structural features. The three-dimensional manganese topological framework of Mn3O(SeO3)3 is formed by the intersection of octa-kagomé lattices and staircase-kagomé lattices. Firstly, we know that the kagomé-like lattices containing equilateral or isosceles triangle sublattices may exhibit strong frustrated magnetic properties [19,20,21]. We note the value of frustration factor, f =|θ|/TN~0.20, with Weiss temperature θ = −8.89 K and Neel temperature TN2~45 K, ruling out spin frustration in the system. It is well known that primary magnetic interactions originate from the superexchange of Mn-O-Mn. A detailed description of superexchange interactions is shown in Figure 2c,d and itemized in Table 1; there are five main magnetic exchange interactions, numbered J1J5, within the three-dimensional spin-lattice according to the Goodenough–Kanamori–Anderson rules (GKA rules) [22]. There are two isosceles triangular topological configurations composed by J1J4 on the staircase-kagomé spin sublattice. According to the GKA rules, J2 and J3 are AFM interactions. J1 and J4 both have two Mn-O-Mn superexchange interactions, as the corresponding MnO6 octahedra share their edges. In general, the spin exchange parameter, J, can be written as J = JFM + JAFM, where JFM indicates ferromagnetic exchange and JAFM indicates antiferromagnetic exchange (JFM > 0 and JAFM < 0) [23]. So, J1 = J1FM(O5) + J1AFM(O6), and the AFM interaction via O6 in J1 is negligible; this means that J1J1FM(O5) > 0. Using the same analytical method, the spin exchange parameter, J4, is ambiguous, as the magnitude of J4AFM(O7) is difficult to calculate. J5 should be a weak AFM interaction. This analysis indicates that AFM interactions are dominant in the system, which is consistent with the negative value of θ. It is significant that the neighboring octa-kagomé sublattices are separated by [-Mn1-] chains and that the neighboring staircase-kagomé sublattices are connected by [Mn2O10] dimers. It is safely said that these two long-range magnetic orders are driven by interlayer magnetic coupling. It is noteworthy that [-Mn1-] chains are composed of Mn2+ ions. If Mn2+ ions are replaced with nonmagnetic ions with a radius similar to Mn2+, such as Mg2+ or Zn2+ [24], an octa-kagomé lattice composed of Mn3+ ions might form. The expected compounds may exhibit spin-liquid or other quantum physical properties [25,26]. Exploratory synthesis of related compounds is underway.
As shown in Figure 6, Mn3O(SeO3)3 undergoes a slow weight gain of approximately 0.20% from 100 to 200 °C and a slow weight loss of approximately 0.40% from 300 to 400 °C. As the weight of the sample is only approximately 7 mg, slight weight gain/loss may be caused by instrument error. As the temperature rises, two successive steps of weight loss of approximately 57% occur from 450 to 630 °C, corresponding to the calculated 59% loss of 3SeO2. The 2% difference may also be caused by instrument error. Based on this analysis, the final residues of Mn3O(SeO3)3 should be Mn3O4; however, this was difficult to characterize, as the residues melted in the Al2O3 crucible after being heated to 900 °C. We re-selected the sample and sintered it in a smooth quartz crucible at 800 °C in a nitrogen atmosphere for 10 min. We then scraped off the sintered product and performed powder X-ray diffraction analysis. As shown in Figure S4, the residue was confirmed as Mn3O4 (PDF #80-0382). This result is consistent with the decomposition characteristics of most manganese-based compounds.

4. Conclusions

A mixed-valence compound, Mn3O(SeO3)3, was successfully synthesized using a conventional hydrothermal method. The reagent LiI acted as a mineralizer. Mn3O(SeO3)3 was shown to have a channel structure with a three-dimensional magnetic topological framework formed by the intersection of octa-kagomé spin sublattices and staircase-kagomé spin sublattices. Magnetic and specific heat data confirmed that Mn3O(SeO3)3 exhibits two successive long-range AFM orderings with TN1~4.5 K and TN2~45 K, and a field-induced spin–flop at a 4.5 T critical field at low temperature. Moreover, magnetic measurements confirmed that the ratio of Mn2+/Mn3+ ions in this compound is 1:2, which is consistent with the structural analysis. The exploratory synthesis of Mg2+ or Zn2+ replaced compounds is in progress.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15165773/s1, Figure S1. Single crystals of Mn3O(SeO3)3 obtained by a conventional hydrothermal method. Figure S2. The oxygen-coordination environments for (a) Mn1, (b) Mn2, (c) Mn3, (d) Se1, (e) Se2 and (f) Se3 atoms in Mn3O(SeO3)3. Figure S3. The variation in χT with the temperature of Mn3O(SeO3)3. With the decreasing temperature, the value of χT decreases. Figure S4. Powder X-ray diffraction pattern for the final residues of Mn3O(SeO3)3 after sintered at 800 °C for 10 min in nitrogen atmosphere.

Author Contributions

W.Z. and G.Q. conceived and designed the experiments. W.Z. prepared the materials and performed the XRD and magnetism testing. J.T. performed the thermogravimetric measurement. P.J. carried out the selection of pure phase crystals. W.Z., M.C. and G.Q. analyzed the data. W.Z. and G.Q. wrote the paper. M.C., J.T., P.J. and X.L. edited the paper. All authors discussed the results and commented on the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Hundreds of Talents program of Sun Yat-sen University and the Shenzhen Science and Technology Program (Grant No. RCBS20200714114820077).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Data is contained within the article or supplementary material.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Varma, C.M. Mixed-valence compounds. Rev. Mod. Phys. 1976, 48, 219–238. [Google Scholar] [CrossRef]
  2. Brown, D.B. Mixed-Valence Compounds: Theory and Applications in Chemistry, Physics, Geology, and Biology; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  3. Sleight, A.W. Chemistry of high-temperature superconductors. Science 1988, 242, 1519–1527. [Google Scholar] [CrossRef] [PubMed]
  4. Uehara, M.; Mori, S.; Chen, C.H.; Cheong, S.W. Percolative phase separation underlies colossal magnetoresistance in mixed-valent manganites. Nature 1999, 399, 560–563. [Google Scholar] [CrossRef]
  5. Tranquada, J.M. John Goodenough and the many lives of transition-metal oxides. J. Electrochem. Soc. 2022, 169, 010535. [Google Scholar] [CrossRef]
  6. Saha, R.A.; Bandyopadhyay, A.; Schiesaro, I.; Bera, A.; Mondal, M.; Meneghini, C.; Ray, S. Colossal electroresistance response accompanied by metal-insulator transition in a mixed-valent vanadate. Phys. Rev. B 2021, 104, 045149. [Google Scholar] [CrossRef]
  7. Wu, T.; Zhao, H.; Zhu, X.; Xing, Z.; Liu, Q.; Liu, T.; Gao, S.; Lu, S.; Chen, G.; Asiri, A.M.; et al. Identifying the origin of Ti3+ activity toward enhanced electrocatalytic N2 reduction over TiO2 nanoparticles modulated by mixed-valent copper. Adv. Mater. 2020, 32, 2000299. [Google Scholar] [CrossRef]
  8. Chen, H.; Liu, Y.; Cai, T.; Dong, W.; Tang, L.; Xia, X.; Wang, L.; Li, T. Boosting photocatalytic performance in mixed-valence MIL-53(Fe) by changing FeII/FeIII ratio. ACS Appl. Mater. Inter. 2019, 11, 28791–28800. [Google Scholar] [CrossRef]
  9. Michel, C.; Hervieu, M.; Borel, M.M.; Grandin, A.; Deslandes, F.; Provost, J.; Raveau, B. Superconductivity in the Bi-Sr-Cu-O system. Z. Phys. B 1987, 68, 421–423. [Google Scholar] [CrossRef]
  10. Raveau, B. Copper mixed valence concept: “Cu(I)–Cu(II)” in thermoelectric copper sulfides—an alternative to “Cu(II)–Cu(III)” in superconducting cuprates. J. Supercond. Nov. Magn. 2019, 33, 259–263. [Google Scholar] [CrossRef]
  11. Zinzuvadiya, S.; Pandya, R.J.; Singh, J.; Joshi, U.S. Low field magnetotransport behavior of barium hexaferrite/ferromagnetic manganite bilayer. J. Appl. Phys. 2021, 130, 024102. [Google Scholar] [CrossRef]
  12. Hasegawa, K.; Isobe, M.; Yamauchi, T.; Ueda, H.; Yamaura, J.; Gotou, H.; Yagi, T.; Sato, H.; Ueda, Y. Discovery of ferromagnetic-half-metal-to-insulator transition in K2Cr8O16. Phys. Rev. Lett. 2009, 103, 146403. [Google Scholar] [CrossRef] [PubMed]
  13. Lin, R.; Bak, S.M.; Shin, Y.; Zhang, R.; Wang, C.; Kisslinger, K.; Ge, M.; Huang, X.; Shadike, Z.; Pattammattel, A.; et al. Hierarchical nickel valence gradient stabilizes high-nickel content layered cathode materials. Nat. commun. 2021, 12, 2350. [Google Scholar] [CrossRef] [PubMed]
  14. Ran, Q.; Zhao, H.; Hu, Y.; Hao, S.; Liu, J.; Li, H.; Liu, X. Enhancing surface stability of LiNi0.8Co0.1Mn0.1O2 cathode with hybrid core-shell nanostructure induced by high-valent titanium ions for Li-ion batteries at high cut-off voltage. J. Alloy. Compd. 2020, 834, 155099. [Google Scholar] [CrossRef]
  15. Yu, X.; Tokunaga, Y.; Taguchi, Y.; Tokura, Y. Variation of topology in magnetic bubbles in a colossal magnetoresistive manganite. Adv. Mater. 2017, 29, 1603958. [Google Scholar] [CrossRef] [PubMed]
  16. Wildner, M. Crystal structure of Mn(II)Mn(III)2O(SeO3)3. J. Solid State Chem. 1994, 113, 252–256. [Google Scholar] [CrossRef]
  17. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef] [Green Version]
  18. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  19. Balents, L. Spin liquids in frustrated magnets. Nature 2010, 464, 199–208. [Google Scholar] [CrossRef]
  20. Aidoudi, F.H.; Aldous, D.W.; Goff, R.J.; Slawin, A.M.; Attfield, J.P.; Morris, R.E.; Lightfoot, P. An ionothermally prepared S = 1/2 vanadium oxyfluoride kagome lattice. Nat. Chem. 2011, 3, 801–806. [Google Scholar] [CrossRef] [Green Version]
  21. Chu, S.; McQueen, T.M.; Chisne, R.; Freedman, D.E.; Muller, P.; Lee, Y.S.; Nocera, D.G. A Cu2+ (S = 1/2) kagomé antiferromagnet: MgxCu4−x(OH)6Cl2. J. Am. Chem. Soc. 2010, 132, 5570–5571. [Google Scholar] [CrossRef] [Green Version]
  22. Goodenough, J.B. Magnetism and the Chemical Bond; John Wiley and Sons: New York, NY, USA, 1966. [Google Scholar]
  23. Whangbo, M.H.; Koo, H.J.; Dai, D.; Jung, D. Effect of metal-ligand bond lengths on superexchange interactions in Jahn-Teller d4 ion systems: Spin dimer analysis of the magnetic structure of Marokite CaMn2O4. Inorg. Chem. 2002, 41, 5575–5581. [Google Scholar] [CrossRef] [PubMed]
  24. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  25. Peng, C.; Ran, S.-J.; Liu, T.; Chen, X.; Su, G. Fermionic algebraic quantum spin liquid in an octa-kagomé frustrated antiferromagnet. Phys. Rev. B 2017, 95, 075140. [Google Scholar] [CrossRef] [Green Version]
  26. Tang, Y.; Peng, C.; Guo, W.; Wang, J.F.; Su, G.; He, Z. Octa-kagomé lattice compounds showing quantum critical behaviors: Spin gap ground state versus antiferromagnetic ordering. J. Am. Chem. Soc. 2017, 139, 14057–14060. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Rietveld refinement of powder X-ray (Cu Kα) diffraction patterns for Mn3O(SeO3)3. The refined lattice constants are a = 15.484(9) Å, b = 6.665(8) Å, c = 9.703(1) Å and β = 118.79(4)° with space group C2/m, which is consistent with the reported parameters of ref. [16].
Figure 1. Rietveld refinement of powder X-ray (Cu Kα) diffraction patterns for Mn3O(SeO3)3. The refined lattice constants are a = 15.484(9) Å, b = 6.665(8) Å, c = 9.703(1) Å and β = 118.79(4)° with space group C2/m, which is consistent with the reported parameters of ref. [16].
Materials 15 05773 g001
Figure 2. (a) The three-dimensional structure frameworks and (b) topological spin structures of Mn3O(SeO3)3; (c,d) show the octa-kagomé and staircase-kagomé spin sublattices, respectively. Here the blue, green, light blue and pink polyhedra represent Mn1O6, Mn2O6, Mn3O6 and SeO3, respectively. Balls of the above colors represent Mn1, Mn2 and Mn3 ions, respectively.
Figure 2. (a) The three-dimensional structure frameworks and (b) topological spin structures of Mn3O(SeO3)3; (c,d) show the octa-kagomé and staircase-kagomé spin sublattices, respectively. Here the blue, green, light blue and pink polyhedra represent Mn1O6, Mn2O6, Mn3O6 and SeO3, respectively. Balls of the above colors represent Mn1, Mn2 and Mn3 ions, respectively.
Materials 15 05773 g002
Figure 3. Detailed linkages of five main superexchange pathways in Mn3O(SeO3)3. (a) indicates J1 to J4 and (b) indicates J5, respectively.
Figure 3. Detailed linkages of five main superexchange pathways in Mn3O(SeO3)3. (a) indicates J1 to J4 and (b) indicates J5, respectively.
Materials 15 05773 g003
Figure 4. (a) The magnetic susceptibility χ(T) of Mn3O(SeO3)3 and its reciprocal. (b) The heat capacity of Mn3O(SeO3)3. The green solid line indicates Curie–Weiss fitting.
Figure 4. (a) The magnetic susceptibility χ(T) of Mn3O(SeO3)3 and its reciprocal. (b) The heat capacity of Mn3O(SeO3)3. The green solid line indicates Curie–Weiss fitting.
Materials 15 05773 g004
Figure 5. Isothermal magnetization (M) as a function of applied field (H) at 2 K and 30 K.
Figure 5. Isothermal magnetization (M) as a function of applied field (H) at 2 K and 30 K.
Materials 15 05773 g005
Figure 6. The thermogravimetric curve of Mn3O(SeO3)3, which shows that the residual product of decomposition above 630 °C is Mn3O4.
Figure 6. The thermogravimetric curve of Mn3O(SeO3)3, which shows that the residual product of decomposition above 630 °C is Mn3O4.
Materials 15 05773 g006
Table 1. Geometric parameters of dominant magnetic superexchanges in Mn3O(SeO3)3 *.
Table 1. Geometric parameters of dominant magnetic superexchanges in Mn3O(SeO3)3 *.
J1d(Mn-O) (Å)d(Mn-Mn) (Å)Mn-O-Mn Angle (°)Magnetism
J1Mn1-2.130(7)-O6-2.130(7)-Mn1
Mn1-2.361(4)-O5-2.361(4)-Mn1
3.332(9)Mn1-O6-Mn1…102.90(8)AFM-w
Mn1-O5-Mn1…89.77(1)FM-S
FM
J2Mn3-1.857(1)-O7-1.857(1)-Mn33.332(9)Mn3-O7-Mn3…127.61(5)AFM-SAFM
J3Mn1-2.361(4)-O5-2.278(1)-Mn23.902(1)Mn1-O5-Mn2…114.49(4)AFM-SAFM
J4Mn2-2.049(5)-O3-2.310(4)-Mn3
Mn2-1.854(2)-O7-1.857(1)-Mn3
3.069(6)Mn2-O3-Mn3…89.29(7)FM-S
Mn2-O7-Mn3…111.59(8)AFM-S
?
J5Mn2-1.896(5)-O4-2.169(5)-Mn2
Mn2-2.169(5)-O4-1.896(5)-Mn2
3.150(0)Mn2-O4-Mn2…101.34(9)AFM-w × 2AFM
* The S or W behind AFM or FM (in the Mn-O-Mn Angle (°) column) refers to the magnitude of the J value; S refers to strong and W refers to weak.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, W.; Cui, M.; Tian, J.; Jiang, P.; Qian, G.; Lu, X. Two Magnetic Orderings and a Spin–Flop Transition in Mixed Valence Compound Mn3O(SeO3)3. Materials 2022, 15, 5773. https://doi.org/10.3390/ma15165773

AMA Style

Zhang W, Cui M, Tian J, Jiang P, Qian G, Lu X. Two Magnetic Orderings and a Spin–Flop Transition in Mixed Valence Compound Mn3O(SeO3)3. Materials. 2022; 15(16):5773. https://doi.org/10.3390/ma15165773

Chicago/Turabian Style

Zhang, Wanwan, Meiyan Cui, Jindou Tian, Pengfeng Jiang, Guoyu Qian, and Xia Lu. 2022. "Two Magnetic Orderings and a Spin–Flop Transition in Mixed Valence Compound Mn3O(SeO3)3" Materials 15, no. 16: 5773. https://doi.org/10.3390/ma15165773

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop