Next Article in Journal
Zero-Pole Optimization of a Novel High-Quality-Factor Planar Helical Resonator
Previous Article in Journal
Review of Orbital Magnetism in Graphene-Based Moiré Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Two-Step Magnetic Ordering in Intercalated Niobium Dichalcogenide MnXNbS2

1
School of Physics, University of Exeter, Exeter EX4 4QL, UK
2
Osipyan Institute of Solid State Physics RAS, 142432 Chernogolovka, Russia
3
Faculty of Physics, HSE University, 101000 Moscow, Russia
*
Author to whom correspondence should be addressed.
Magnetism 2023, 3(3), 259-266; https://doi.org/10.3390/magnetism3030020
Submission received: 2 April 2023 / Revised: 6 July 2023 / Accepted: 22 August 2023 / Published: 4 September 2023

Abstract

:
Transition metal dichalcogenides are studied due to the possibility of creating nanoscale semiconductor devices, as well as fundamental issues of magnetic ordering. We researched the crystal structure and magnetic properties of niobium dichalcogenide Mn0.30NbS2. The results of the X-ray study showed the possible existence of an intermediate 2 3 a0·2 3 a0 structure between the “basic” superstructures. Also, two local maximums were found in the temperature dependence of the dynamic magnetic susceptibility. These features can indirectly confirm the presence of a transition superstructure and reflect the two-step nature of the magnetic ordering.

1. Introduction

Transition metal dichalcogenides TX2 (where T = Mo, Nb, Ta, Ti, W; X = S, Se, Te) have been relevant objects of study of solid-state physics for several decades. Today, the primary interest in these compounds is due to the discovery of monolayers of TX2. Technologies for assembling van der Waals heterostructures have led to the creation semiconductor nanodevices based on individual monolayers of TX2: nanoscale field effect transistors [1,2,3,4], low-dimensional rectifying contacts [5], phototransistor devices [6], spin-valley filters [7], battery applications [8,9], and photoluminescence effects [10,11].
At the same time, intercalated transition metal dichalcogenides provide a unique opportunity to study the fundamental regularities of magnetic ordering. Transition metal ions intercalated in an initially diamagnetic matrix of TX2 can form a magnetically ordered structure. The type and temperature of magnetic ordering depend on the concentration of intercalated ions and their valence and spin states, as well as the initial diamagnetic matrix. When tracking the dependence of the type and parameters of magnetic ordering on these parameters, one can find out some regularities that reflect the fundamental properties of magnetic ordering and its relationship with an electron structure. For example, with an increase in the number n of d-electrons, in a series of M1/3NbS2 compounds (where M = transition 3d metal) the transition from a paramagnet (M = Ti, n = 1) to a ferromagnet (M = V, n = 2) to a helimagnet (M = Cr, Mn, n = 3–4) to an antiferromagnet (M = Fe, Co, Ni, n = 5–7) is observed.
In an M1/3NbS2 series, Cr1/3NbS2 [12,13,14,15,16] and Mn1/3NbS2 [17,18,19] are two of the most interesting compounds for the physics of magnetism: a uniaxial helicoidal magnetic structure with the wave vector q directed along the hexagonal axis c is formed in them as a result of competition between symmetric and anti-symmetric exchange interactions. Despite a large number of works devoted to these compounds, the most important crystallographic and magnetic characteristics (the temperature of magnetic ordering, the critical fields of magnetic phase transitions, etc.) obtained by various authors differ significantly. Possible reasons for these differences may be the uncontrolled deviations of chemical composition from the stated one or the non-uniform distribution of intercalated ions in an initial matrix. It was shown in [20] that in Cr1/3NbS2 helimagnet, a decrease in the Cr concentration led to an increase in the structural disordering of intercalated ions. This caused a decrease in the temperature of magnetic ordering, as well as the formation of a ferromagnetic not helicoidal magnetic structure. It was shown in [21] that in a single FexTaS2 crystal, the extreme value of the magnetoresistance corresponds to x = 0.3. The present work aimed to study the peculiarities of crystal and magnetic structures in MnxNbS2 when the concentration of intercalated ions x was less than 0.33. It can be seen that this issue is of interest in the scientific community, so this year an article [22] was published on a similar topic.

2. Preparation of Samples and Experimental Methods of Study

Single-crystal Mn0.30NbS2 samples were prepared in two stages [23].
In the first stage, the powders of initial chemical elements (Nb, S, and Mn) taken in the required ratio were thoroughly mixed and sealed in an evacuated quartz ampoule. Then, the ampoule was slowly (5 °C/h) heated to 800 °C and subjected to this temperature for 14 days, after which the polycrystalline Mn0.30NbS2 sample was obtained.
In the second stage, Mn0.30NbS2 single crystals were obtained using the method of chemical transport. Iodine with a volume concentration of 5 mg/cm3 was used as a transport agent. The 25 cm long sealed quartz ampoule with weighed portions of iodine and the polycrystalline Mn0.30NbS2 sample obtained in the first synthesis stage was placed in a two-zone tube furnace with independent temperature control in each zone. During the growth, a constant temperature gradient T = 900–800 °C was maintained in the furnace. The weighed portion of the polycrystalline sample was in the zone with a higher temperature. After one week, the furnace was smoothly cooled down to room temperature at a rate of 10 °C/h. In the zone with a reduced temperature, non-transparent single crystals with the shape of thin hexagonal plates or pyramids were found on the walls of the quartz ampoule.
Four selected crystals with regular hexagonal faceting were cleaved along the basal plane into several thin plates. One plate of each crystal was used to determine the chemical composition using energy-dispersive X-ray spectroscopy. Three to five tests were conducted for each plate, and the chemical composition of the obtained crystals was determined, which was Mn0.30 ± 0.05Nb1.08 ± 0.03S1.96 ± 0.05.
The polycrystalline Mn0.30NbS2 samples and powdered single crystals were tested at room temperature by X-ray diffraction using an ARL X’TRA powder diffractometer, after the first stage and second stage, respectively. The obtained diffraction spectra were compared to the data of the online crystallographic database AtomWork [24].
The crystal structure of the single crystals grown was investigated using a P4 BRUKER single-crystal X-ray diffractometer (Cu Kα, λ = 0.154 nm).
To determine the magnetic properties of the crystals, the dynamic magnetic susceptibility χ was measured at a frequency of 100 kHz in the temperature range of T = 5–300 K. A sample was placed inside a pair of coaxial coils with a 6 mm in diameter. One of the coils served to excite an alternating magnetic field, while the other one was used for the measurement [25]. An identical pair of coils used to compensate for a signal in the absence of a sample was put nearby. By using a standard circuit of synchronous detection, we measured the imbalance signal. This signal was proportional to the magnetic moment M = χVHrf. Here Hrf ≈ 0.1 Oe is the amplitude of an alternating magnetic field produced by a coil, and V is the volume of a sample. In addition, we could put a sample in a constant magnetic field H, which had an amplitude of up to 150 Oe and was parallel to a high-frequency field.

3. Structural Studies

The crystal structure of intercalated MnxNbS2 (x = 0–0.33) dichalcogenides is formed by 2D NbS2 layers that consist of trigonal prisms. S atoms are located at the apexes of the prisms, and Nb atoms are located in the center of every second prism. Intercalated Mn ions take octahedral positions between two adjacent 2D layers. At certain concentrations, x, intercalated ions take ordered positions in the basal plane and form a crystal superstructure [26,27]. For example, at x = 1/4 a crystal superstructure of the 2a0·2a0 type with a space group P63/m (where a0 = 3.3 Å is the spacing of the lattice cell of initial non-intercalated NbS2) is formed. At the concentration x = 1/3, intercalated ions form a crystal superstructure of the 3 a0· 3 a0 type with a space group P6322 (see Table 1). Note that the crystallographic parameters of the compounds under study available in the literature often differ by several tenths of Å. Therefore, the accuracy of the estimations and comparisons provided in this work cannot exceed 0.2–0.3 Å.
A change in the concentration of intercalated ions affects both the crystal structure and the type of magnetic ordering. Thus, for Mn0.25NbS2 compounds magnetic ordering occurs at T = 100 K with the formation of an “easy-plane” ferromagnetic structure [27,28]. At the same time, in Mn0.33NbS2 a helicoidal incommensurate magnetic structure [17] with a period of 700 nm and a wave vector directed along the crystallographic axis c is formed at T = 40 K.
The X-ray powder diffraction spectrum of our Mn0.30NbS2 sample was compared to the spectra of Mn0.33NbS2 and Mn0.25NbS2 [24,26,27,29,30,31,32,33] (Figure 1). For all three compared compounds, the positions of the most intense peaks corresponding to scattering on the atoms of an initial NbS2 matrix are almost identical. At the same time, the positions of some low-intensity peaks in the spectrum of Mn0.30NbS2 cannot be related to any of the reflections of Mn0.33NbS2 or Mn0.25NbS2 structures. These reflections are marked with arrows in Figure 1. The compounds under discussion differ in the concentration x of intercalated Mn ions, which has little effect on the position of atoms in an initial NbS2 matrix. Therefore, it is reasonable to assume that the presence of additional peaks in the X-ray diffraction spectrum is due to the difference in the ordering of intercalated Mn ions in the basal plane.
To prove this assumption, single-crystal X-ray diffraction studies were carried out on individual plates cleaved from the selected single crystals. The following parameters of the crystal structure were obtained: a = b = 11.53 Å, c = 12.50 Å, α = β = 90°, γ = 120°, space group P-3m1. The spacing c of the crystal lattice along the hexagonal axis has an intermediate value between the values of this parameter for Mn0.33NbS2 and Mn0.25NbS2. This agrees well with the previously known dependence of interlayer distance on the concentration of intercalated ions [27]. At the same time, the spacing of a crystal structure in the basal plane a = b = 11.53 Å is much greater than that for Mn0.33NbS2 ( 3 a0 = 5.72 Å) and Mn0.25NbS2 (2a0 = 6.6 Å).
The discovered increase in the parameters a, b is most likely due to the formation of a new type of ordering of intercalated Mn ions, which differs drastically from the previously known 2a0·2a0 and 3 a0· 3 a0 superstructures. Our value of a, b is close with good accuracy to the value 2 3 a0 = 11.43 Å, corresponding to the crystal 2 3 a0·2 3 a0 superstructure. This type of superstructure was found earlier under the absorption of Sn atoms on the Si(111) surface [34]. The typical characteristics and conditions for the formation of this type of crystal structure have to be studied.

4. Magnetic Properties

The temperature dependencies of the dynamic magnetic susceptibility of Mn0.30NbS2 single crystal were measured for two different orientations of the crystal in an external magnetic field:
  • the magnetic field was in the basal plane ab (Hrfc);
  • the magnetic field was oriented along the hexagonal axis c (Hrfc).
In the Hrfc orientation, a decrease in the temperature below TC1 = 100 K leads to an abrupt increase in the susceptibility χ (Figure 2a). A further decrease in the temperature to 50 K leads to a slight decrease in the susceptibility. The second abrupt increase in the magnetic susceptibility χ occurs in the dependence χ(T) near TC2 = 40 K. It is obvious that these increases in the magnetic susceptibility are due to the phase transitions to magnetically ordered states.
The application of a constant magnetic field in the basal plane (Hc) has little effect on the position of the transition temperature TC1; however, it leads to an abrupt decrease in the dynamic magnetic susceptibility in the temperature range between TC1 and TC2. At the same time, the dependence χ(T) becomes smooth near TC2, and the phase transition expands to a wider temperature range. The external magnetic field up to H = 150 Oe has little effect on the magnetic susceptibility χ below the second critical temperature TC2.
In the Hrfc orientation, a phase transition near TC1 = 100 K is accompanied by only a slight increase in the susceptibility χ (Figure 2b). With a decrease in the temperature from 50 K to TC2 = 40 K, an abrupt increase in the magnetic susceptibility is observed. Below 40 K, χ(T) demonstrates nonmonotonic behavior with a minimum near 20 K. The constant magnetic field applied to the sample does not cause qualitative changes in the dependence χ(T).
Let’s consider the first phase transition. At TC1 = 100 K, an abrupt increase in the magnetic susceptibility in the basal plane and a small change in χ(T) along the hexagonal axis c occur. This indicates the transition to an anisotropic magnetically ordered ferro- or ferrimagnetic state. Then, a smooth decrease in χ with a temperature decrease below TC1 is due to the Hopkinson effect, which is observed in ferro- and ferrimagnetic materials. Below the magnetic ordering temperature TC1, the application of a small (H = 10 Oe) constant magnetic field in the basal plane leads to a decrease in the susceptibility by several times. These peculiarities indicate the easiness of magnetic ordering in the basal plane.
Thus, the experimental data indicate that an “easy-plane” ferromagnetic structure is formed at TC1. The critical temperature and the type of magnetic ordering are similar to those observed in the Mn0.25NbS2 compound.
The second magnetic phase transition at TC2 = 40 K is accompanied by an abrupt increase in the magnetic susceptibility both along the hexagonal axis and in the basal plane. Moreover, χ almost does not depend on the applied magnetic field and its direction. This fundamental difference in the dependencies χ(H) in the basal plane above and below the critical temperature TC2 provides evidence of a change in the type of magnetic ordering at this temperature.
The magnetic susceptibility χ in the basal plane is greater than susceptibility along the hexagonal axis both above and below TC2. Consequently, the transition at TC2 is not related to a change in the direction of easy magnetization, i.e., the “easy plane”–“easy axis” transition. It is known that a transition to a uniaxial helicoidal state occurs in Mn0.33NbS2 at about 40 K. As shown in [35], weak dependencies of susceptibility on the temperature and magnetic field are typical for uniaxial helimagnets away from phase transitions. Although our experimental data do not allow asserting with complete confidence that a magnetic structure formed below TC2 is similar to a uniaxial helicoidal phase in Mn0.33NbS2, the temperature of the phase transition coinciding within a few degrees and the dependence χ(T, H) typical for helimagnets are direct evidence for the formation of a helicoidal phase.
Thus, two magnetic phase transitions are observed in the obtained Mn0.30NbS2 compound. The first transition is observed at TC1 = 100 K and corresponds to the transition to an easy-plane ferromagnet phase at TC1 = 100 K [27,28]. This transition is similar to that in Mn0.25NbS2 with the crystal 2a0·2a0 superstructure. The second transition is observed at TC2 = 40 K and is similar to the magnetic ordering of a helicoidal phase in Mn0.33NbS2 with the crystal 3 a0· 3 a0 superstructure [17,18,19].
Two different interpretations of the presence of two critical temperatures in the compound under study can be proposed. On the one hand, in Mn0.30NbS2 the aggregation of intercalated Mn ions and the formation of the clusters of Mn0.33NbS2 and Mn0.25NbS2 phases are possible. In this case, the coexistence of clusters of two different phases in one crystal will result in the presence of two different critical temperatures of magnetic ordering. This interpretation is evidenced by a significant variation in the concentration of Mn ions, x = 0.30 ± 0.05, in the crystals under study.
On the other hand, both temperatures of magnetic ordering may correspond to the same crystallographic phase. This phase is an intermediate superstructure of the 2 3 a0·2 3 a0 type. The compound under consideration has a layered structure, which results in interlayer exchange interactions being significantly weaker than exchange interactions inside one layer. Then, the phase transition at TC1 = 100 K corresponds to the ferromagnetic ordering of intercalated ions located in the same layer, while the phase transition at TC2 = 40 K corresponds to interlayer ordering and the formation of a “simple spiral” helicoidal structure. A similar two-step magnetic transition was found earlier in DyNi2B2C [36]. The existence of the 2 3 a0·2 3 a0 superstructure is indicated by the presence of additional reflections in the X-ray diffraction spectra, as well as the parameters of the crystal lattice, which differ significantly from those of the 2a0·2a0 and 3 a0· 3 a0 superstructures. Moreover, in the Hc orientation the experimental dependence χ(T) (Figure 2a) contradicts the proposed “cluster” interpretation. In the case of the “cluster” origin of the critical temperatures TC1 and TC2, the susceptibility χ will be a sum of two terms χferro + χhelix. The first term, χferro, corresponds to the ferromagnetic phase of Mn0.25NbS2 clusters and decreases smoothly with a decrease in the temperature due to the Hopkinson effect. The second term, χhelix, corresponds to the helicoidal phase of Mn0.33NbS2 clusters and almost does not depend on the temperature at T < 40 K. Then, the resulting susceptibility χ will decrease with a decrease in the temperature both above and below TC2, and the contribution χhelix will lead to the emergence of a peak near TC2 = 40 K. In our experiments, χ(T) is almost independent of temperature below 40 K, which indicates the absence of the ferromagnetic contribution χferro below TC2. Consequently, the transition of a ferromagnetic phase to a helicoidal phase is observed at TC2.
Thus, the comparison of two critical temperatures TC1 and TC2 with one crystal 2 3 a0·2 3 a0 superstructure seems to be a more probable explanation of the experimental results.

5. Conclusions

Intercalated Mn0.30NbS2 crystal has been obtained. The crystal structure and magnetic properties of this crystal have been studied. The spacing of the crystal structure in the basal plane a = b = 11.53 Å exceeds the spacing of the known Mn0.33NbS2 and Mn0.25NbS2 compounds by several times. A new type of crystal superstructure 2 3 a0·2 3 a0 is most likely to be formed in the investigated compound; it is an intermediate type between “basic” 2a0·2a0 and 3 a0· 3 a0 superstructures. Two critical temperatures are found in the temperature dependence of the dynamic magnetic susceptibility χ. The first critical temperature, TC1 = 100 K, corresponds to the transition from a paramagnetic state to an “easy-plane” ferromagnetic phase. The second critical temperature, TC2 = 40 K, corresponds to the transition to a uniaxial helimagnet phase. Both critical temperatures may correspond to the 2 3 a0·2 3 a0 superstructure and reflect the two-step nature of magnetic ordering. The magnetic ordering of intercalated ions in the basal plane occurs at TC1, and the interlayer ordering with the formation of a helicoidal structure occurs at TC2. We believe that our exciting results are a good motivation for more in-depth studies of intercalated transition metal dichalcogenides near x = 0.3, in order to obtain both new material properties and fundamental knowledge. We also would like to emphasize the prospects of the MxTX2 series as excellent candidates for assembling van der Waals heterostructures with unusual magnetic properties.

Author Contributions

Conceptualization, methodology, software, validation, formal analysis, investigation, resources, data curation, writing—original draft preparation, visualization, supervision, funding acquisition, F.M., A.S. and D.S.; writing—review and editing, M.P.; project administration, A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Osipyan Institute of Solid State Physics RAS.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  2. Choi, W.; Kim, S. Thin-Film Transistors Based on Transition Metal Dichalcogenides. In Nanomaterials, Polymers, and Devices: Materials Functionalization and Device Fabrication; Kong, E.S.W., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015. [Google Scholar]
  3. Lan, Y.-W.; Torres, C.; Tsai, S.-H.; Zhu, X.; Shi, Y.; Li, Y.; Ming, Y.; Li, L.-J.; Yeh, W.-K.; Wang, K. Atomic-monolayer MoS2 band-to-band tunneling field-effect transistor. Small 2016, 12, 5676–5683. [Google Scholar] [CrossRef]
  4. Tang, L.; Xu, R.; Tan, J.; Yuting, L.; Zou, J.; Zongteng, Z.; Zhang, R.; Zhao, Y.; Lin, J.; Zou, X.; et al. Modulating Electronic Structure of Monolayer Transition Metal Dichalcogenides by Substitutional Nb-Doping. Adv. Func. Mater. 2020, 31, 2006941. [Google Scholar] [CrossRef]
  5. Szczęśniak, D.; Hoehn, R.D.; Kais, S. Canonical Schottky barrier heights of transition metal dichalcogenide monolayers in contact with a metal. Phys. Rev. B 2018, 97, 195315. [Google Scholar] [CrossRef]
  6. Pak, J.; Lee, I.; Cho, K.; Kim, J.-K.; Jeong, H.; Hwang, W.-T.; Ahn, G.; Kang, K.; Yu, W.; Javey, A.; et al. Intrinsic optoelectronic characteristics of MoS2 phototransistors via a fully transparent van der Waals heterostructure. ACS Nano 2019, 13, 9638–9646. [Google Scholar] [CrossRef]
  7. Szczęśniak, D.; Kais, S. Gap states and valley-spin filtering in transition metal dichalcogenide monolayers. Phys. Rev. B 2020, 101, 115423. [Google Scholar] [CrossRef]
  8. Zhang, J.; Du, C.; Dai, Z.; Chen, W.; Zheng, Y.; Li, B.; Zong, Y.; Wang, X.; Zhu, J.; Yan, Q. NbS2 nanosheets with M/Se (M = Fe, Co, Ni) codopants for Li+ and Na+ storage. ACS Nano 2017, 11, 10599–10607. [Google Scholar] [CrossRef]
  9. Zhou, J.; Shen, Y.; Lv, F.; Zhang, W.; Lin, F.; Zhang, W.; Wang, K.; Luo, H.; Wang, Q.; Yang, H.; et al. Ultrathin Metallic NbS2 Nanosheets with Unusual Intercalation Mechanism for Ultra-Stable Potassium-Ion Storage. Adv. Funct. Mater. 2022, 32, 2204495. [Google Scholar] [CrossRef]
  10. Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.-Y.; Galli, G.; Wang, F. Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10, 1271–1275. [Google Scholar] [CrossRef]
  11. Chernenko, A.; Brichkin, A.; Golyshkov, G.; Shevchun, A. Effect of the Quality of Interfaces on the Photoluminescence of Encapsulated MoSe2 Monolayers. Bull. Russ. Acad. Sci. Phys. 2023, 87, 161–164. [Google Scholar] [CrossRef]
  12. Miyadai, T.; Kikuchi, K.; Kondo, H.; Sakka, S.; Arai, M.; Ishikawa, Y.J. Magnetic properties of Cr1/3NbS2. Phys. Soc. Jpn. 1983, 52, 1394–1401. [Google Scholar] [CrossRef]
  13. Koyama, T.; Takayanagi, K.; Mori, S.; Kousaka, Y.; Akimitsu, J.; Nishihara, S.; Inoue, K.; Ovchinnikov, A.S.; Kishine, J. Chiral magnetic soliton lattice on a chiral helimagnet. Phys. Rev. Lett. 2012, 108, 107202. [Google Scholar]
  14. Ghimire, N.J.; McGuire, M.A.; Parker, D.S.; Sipos, B.; Tang, S.; Yan, J.-Q.; Sales, B.C.; Mandrus, D. Magnetic phase transition in single crystals of the chiral helimagnet Cr1/3NbS2. Phys. Rev. B 2013, 87, 104403. [Google Scholar] [CrossRef]
  15. Braam, D.; Tezok, S.; de Mello, E.V.L.; Li, L.; Mandrus, D.; Kee, H.-Y.; Sonier, J.E. Magnetic properties of the helimagnet Cr1/3NbS2 observed by μSR. Phys. Rev. B 2015, 91, 144407. [Google Scholar] [CrossRef]
  16. Sirica, N.; Vilmercati, P.; Bondino, F.; Pis, I.; Nappini, S.; Mo, S.-K.; Fedorov, A.V.; Das, P.K.; Vobornik, I.; Fujii, J.; et al. The nature of ferromagnetism in the chiral helimagnet Cr1/3NbS2. Commun. Phys. 2020, 3, 65. [Google Scholar] [CrossRef]
  17. Kousaka, Y.; Nakao, Y.; Kishine, J.; Akita, M.; Inoue, K.; Akimitsu, J. Chiral helimagnetism in T1/3NbS2 (T= Cr and Mn). Nucl. Instrum. Methods Phys. Res. A 2009, 600, 250–253. [Google Scholar] [CrossRef]
  18. Karna, S.K.; Womack, F.N.; Chapai, R.; Young, D.P.; Marshall, M.; Xie, W.; Graf, D.; Wu, Y.; Cao, H.; DeBeer-Schmitt, L.; et al. Consequences of magnetic ordering in chiral Mn1/3NbS2. Phys. Rev. B 2019, 100, 184413. [Google Scholar] [CrossRef]
  19. Dai, Y.; Liu, W.; Wang, Y.; Pi, L.; Zhang, L.; Zhang, Y.; Fan, J. Critical phenomenon and phase diagram of Mn-intercalated layered MnNb3S6. J. Phys. Condens. Matter 2019, 31, 195803. [Google Scholar] [CrossRef]
  20. Dyadkin, V.; Mushenok, F.; Bosak, A.; Menzel, D.; Grigoriev, S.; Pattison, P.; Chernyshov, D. Structural disorder versus chiral magnetism in Cr1/3NbS2. Phys. Rev. B 2015, 91, 184205. [Google Scholar] [CrossRef]
  21. Chen, C.-W.; Chikara, S.; Zapf, V.S.; Morosan, E. Correlations of crystallographic defects and anisotropy with magnetotransport properties in FexTaS2 single crystals (0.23 ≤ x ≤ 0.35). Phys. Rev. B 2016, 94, 054406. [Google Scholar] [CrossRef]
  22. Mao, Q.; Xu, Y.; Li, R.; Wang, Y.; Li, H.; Xue, L.; Hao, H.; Khan, R.; Chen, B.; Yang, J. Localized magnetism, magnetic entropy change and critical properties of Mn1/4+δNbS2 single crystals. Phys. B Condens. 2023, 649, 414469. [Google Scholar] [CrossRef]
  23. Anohin, V.Z.; Goncharov, E.G.; Kostrjukova, E.P.; Pshestanchik, V.R.; Marshakova, T.A. Praktikum po Himii i Tehnologii Poluprovodnikov; M.:Vysshaya shkola: Minsk, Belarus, 1978; p. 191. [Google Scholar]
  24. Inorganic Material Database (AtomWork), National Institute for Materials Science (Japan). Available online: http://crystdb.nims.go.jp (accessed on 1 January 2023).
  25. Neminsky, A.M.; Nikolaev, P.N.; Shovkun, D.V.; Laukhina, E.E.; Yagubskii, E.B. Temperature dependence of the magnetic field penetration depth in Rb3C60 measured by ac susceptibility. Phys. Rev. Lett. 1994, 72, 3092. [Google Scholar] [CrossRef]
  26. Anzenhofer, K.; Van Den Berg, J.M.; Cossee, P.; Helle, J.N. The crystal structure and magnetic susceptibilities of MnNb3S6, FeNb3S6, CoNb3S6 and NiNb3S6. J. Phys. Chem. Solids 1970, 31, 1057–1067. [Google Scholar] [CrossRef]
  27. Blanc-Soreau, A.L.; Rouxel, J.; Gardette, M.; Gorochov, O. Proprietes electriques et magnetiques de Mn0.25NbS2 et Mn0.33NbS2. Mater. Res. Bull. 1976, 11, 1061–1071. [Google Scholar] [CrossRef]
  28. Rahman, A.; Rehman, M.; Yousaf, M.; Kiani, M.; Zhao, H.; Wang, J.; Lu, Y.; Ruan, K.; Dai, R.; Wang, Z.; et al. RKKY-type in-plane ferromagnetism in layered Mn1/4NbS2 single crystals. Phys. Rev. B 2022, 105, 214410. [Google Scholar] [CrossRef]
  29. Hulliger, F.; Pobitschka, E. On the magnetic behavior of new 2H NbS2-type derivatives. J. Solid State Chem. 1970, 1, 117–119. [Google Scholar] [CrossRef]
  30. Voorhoeve, J.M.; Berg, V.D.; Robbins, M. Intercalation of the niobium-diselenide layer structure by first-row transition metals. J. Solid State Chem. 1970, 1, 134–137. [Google Scholar] [CrossRef]
  31. Anzenhofer, K.; de Boer, J.J. The crystal structure of MnNb4S8. Recl. Trav. Chim. Pays-Bas 1971, 90, 56–64. [Google Scholar] [CrossRef]
  32. Onuki, Y.; Ina, K.; Hirai, T.; Komatsubara, T. Magnetic properties of intercalation compound: Mn1/4MX2. J. Phys. Soc. Jpn. 1986, 55, 347–356. [Google Scholar] [CrossRef]
  33. Miwa, K.; Ikuta, H.; Hinode, H.; Uchida, T.; Wakihara, M. Synthesis and physical properties of MnxNbS2. J. Solid State Chem. 1996, 125, 178–181. [Google Scholar] [CrossRef]
  34. Ichikawa, T.; Cho, K. Structural Study of Si (111)(2√3 × 2√3) R30°–Sn Surfaces. Jpn. J. Appl. Phys. 2003, 42, 5239. [Google Scholar] [CrossRef]
  35. Borich, M.A.; Smagin, V.V.; Tankeev, A.P. Thermodynamics of a helicoidal magnetic structure in a spin-wave approximation. Phys. Met. Metallogr. 2014, 115, 737–743. [Google Scholar] [CrossRef]
  36. Kuznietz, M.; Takeya, H. Two-step magnetic ordering in as-cast and annealed DyNi2B2C samples. J. Phys. Condens. Matter. 1999, 11, 6561. [Google Scholar] [CrossRef]
Figure 1. The experimental X-ray diffraction spectrum of Mn0.30NbS2 powder and comparison spectra for (a) Mn0.33NbS2, crystal superstructure, (b) Mn0.25NbS2, crystal superstructure, and (c) Mn0.30NbS2 calculated according to the data of single-crystal X-ray diffraction.
Figure 1. The experimental X-ray diffraction spectrum of Mn0.30NbS2 powder and comparison spectra for (a) Mn0.33NbS2, crystal superstructure, (b) Mn0.25NbS2, crystal superstructure, and (c) Mn0.30NbS2 calculated according to the data of single-crystal X-ray diffraction.
Magnetism 03 00020 g001
Figure 2. Temperature dependencies of the magnetic susceptibility χ of Mn0.30NbS2 single crystals obtained in different magnetic fields in Hc (a) and Hc (b) orientations.
Figure 2. Temperature dependencies of the magnetic susceptibility χ of Mn0.30NbS2 single crystals obtained in different magnetic fields in Hc (a) and Hc (b) orientations.
Magnetism 03 00020 g002
Table 1. Comparison of the crystallographic data and magnetic properties of MnxNbS2.
Table 1. Comparison of the crystallographic data and magnetic properties of MnxNbS2.
xa, b (Å)c (Å)Space GroupType of the SuperstructureTC, K
0.256.64–6.9212.46–12.90P63/m2a0·2a0100
0.3011.5312.50P-3m12 3 a0·2 3 a0100; 40
0.335.78–5.8112.55–12.64 P6322 3 a0· 3 a040
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

Mushenok, F.; Shevchun, A.; Shovkun, D.; Prokudina, M. Two-Step Magnetic Ordering in Intercalated Niobium Dichalcogenide MnXNbS2. Magnetism 2023, 3, 259-266. https://doi.org/10.3390/magnetism3030020

AMA Style

Mushenok F, Shevchun A, Shovkun D, Prokudina M. Two-Step Magnetic Ordering in Intercalated Niobium Dichalcogenide MnXNbS2. Magnetism. 2023; 3(3):259-266. https://doi.org/10.3390/magnetism3030020

Chicago/Turabian Style

Mushenok, Fedor, Artem Shevchun, Dmitriy Shovkun, and Maria Prokudina. 2023. "Two-Step Magnetic Ordering in Intercalated Niobium Dichalcogenide MnXNbS2" Magnetism 3, no. 3: 259-266. https://doi.org/10.3390/magnetism3030020

Article Metrics

Back to TopTop