Next Article in Journal
Using Direct Current Potential Drop Technique to Estimate Fatigue Crack Growth Rates in Solid Bar Specimens under Environmental Assisted Fatigue in Simulated Pressurized Water Reactor Conditions
Next Article in Special Issue
Simultaneous Improvement in Strength and Ductility of TC4 Matrix Composites Reinforced with Ti1400 Alloy and In Situ-Synthesized TiC
Previous Article in Journal
Effect of Particle Size of MgO on the Sinter Strength before and after Reduction and Its Mechanism Analysis
Previous Article in Special Issue
Structure and Oxidation Behavior of NiAl-Based Coatings Produced by Non-Vacuum Electron Beam Cladding on Low-Carbon Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Peculiarities of the Electro- and Magnetotransport in Semimetal MoTe2

by
Alexandra N. Perevalova
*,
Sergey V. Naumov
and
Vyacheslav V. Marchenkov
M.N. Mikheev Institute of Metal Physics, UB RAS, 620108 Ekaterinburg, Russia
*
Author to whom correspondence should be addressed.
Metals 2022, 12(12), 2089; https://doi.org/10.3390/met12122089
Submission received: 30 September 2022 / Revised: 15 November 2022 / Accepted: 28 November 2022 / Published: 6 December 2022
(This article belongs to the Special Issue Intermetallic-Based Materials and Composites)

Abstract

:
Weyl semimetal MoTe2 single crystal was grown by the chemical vapor transport method. Electrical resistivity, magnetoresistivity, and Hall effect in MoTe2 were studied in detail. It was shown that both the electrical resistivity in the absence of a magnetic field and the conductivity in the field depend on temperature according to a quadratic law in a wide temperature range. It has been suggested that the quadratic temperature dependence of the conductivity in a magnetic field at low temperatures might be associated with the “electron-phonon-surface” interference scattering mechanism. The analysis of data on the Hall effect in MoTe2 was carried out using single-band and two-band models. Apparently, the two-band model is preferable in such systems containing different groups of current carriers.

1. Introduction

In recent years, topological materials have attracted a lot of attention because they have a non-trivial topology of the electronic band structure and unusual electronic properties. In addition to interest from the point of view of fundamental science, such materials are actively studied due to the possibility of their practical application in micro-, nanoelectronic and spintronic devices. Topological materials include topological semimetals [1,2] as well as previously discovered topological insulators [3,4]. Topological insulators have an energy gap in the bulk and topologically protected gapless surface states. At the same time, topological semimetals have unusual states both on the surface and in bulk. Such materials include Weyl semimetals [5]. In the bulk of a Weyl semimetal, two nondegenerate bands with linear dispersion cross near the Fermi level, forming Weyl nodes, which always exist in pairs with opposite chirality. Quasiparticles near such nodes behave such as massless Weyl fermions in high-energy physics. Similar to topological insulators, Weyl semimetals have topologically protected surface states called Fermi arcs. They are open curves in momentum space that connect the projections of bulk Weyl nodes with opposite chirality on the surface.
Weyl fermions can only be found in crystals in which either inversion symmetry or time-reversal symmetry is broken. The existence of such quasiparticles was first experimentally confirmed in TaAs in 2015 [6]. Soon, the authors of [7] predicted another type of band crossing with a tilted Weyl cone. As a result, the Lorentz invariance is violated, and the corresponding quasiparticles are called type-II Weyl fermions. Instead of the point-like Fermi surface in a type-I Weyl semimetal, the type-II Weyl node is the touching point between electron and hole pockets. The authors of [7] also theoretically predicted that the WTe2 compound is a type-II Weyl semimetal. This was experimentally confirmed in [8,9]. By analogy with WTe2, it was theoretically predicted [10] and experimentally confirmed [11] that MoTe2 is also a type-II Weyl semimetal.
WTe2 and MoTe2 belong to a large group of materials, layered transition metal dichalcogenides, with the general chemical formula MX2, where M is a transition metal atom, and X is a chalcogen atom [12]. However, unlike WTe2, for which the orthorhombic structure Td is stable with temperature, MoTe2 can crystallize into one of three phases: hexagonal (2H), monoclinic (1T’), or orthorhombic (Td) [13]. Depending on the synthesis conditions, MoTe2 can be obtained in the hexagonal (2H) or monoclinic (1T’) phase, which is also called the α- and β-phase, respectively [14]. The 2H-phase (space group P63/mmc) is semiconducting. Whereas the 1T’-phase is semimetallic and belongs to the centrosymmetric space group P21/m. A temperature-induced phase transition from the high-temperature 1T’-phase (β-MoTe2) to the low-temperature Td-phase of MoTe2 near 250 K was reported in [15]. The Td-phase has a similar crystal structure in the layer plane as the 1T’-phase but has a vertical (90°) packing and belongs to the noncentrosymmetric space group Pmn21. It is in Td-phase that MoTe2 is a type-II Weyl semimetal [10,11].
Due to the unusual topology of the electronic band structure, Weyl semimetals have unique electronic transport properties. Such features of electronic transport include extremely large unsaturated magnetoresistance [16,17], negative longitudinal magnetoresistance [18], low effective mass and ultrahigh mobility of current carriers [19,20], the intrinsic anomalous Hall effect, etc. In addition, quadratic temperature dependence of the electrical resistivity was observed in MoTe2 [21,22,23] and WTe2 [24,25] single crystals in a very wide temperature range. In particular, it was found in [21] that the electrical resistivity of semimetallic MoTe2 depends on temperature according to a quadratic law in the temperature range from 1.7 to 50 K. It was suggested (see [21] and references therein) that, in addition to electron-electron scattering, other possible scattering mechanisms can lead to the T2-dependence of the electrical resistivity of MoTe2; in particular, the contribution proportional to T2 can be associated with electron-phonon scattering, as this was shown for a semiconductor TiS2 compound [26]. However, further research is required to understand the role of all scattering mechanisms in MoTe2. As we previously showed for the WTe2 single crystal in [23], not only the electrical resistivity in the absence of a magnetic field but also the conductivity in the field depends on temperature according to a quadratic law in a wide temperature range from 12 K to ~70 K and ~55 K, respectively. Therefore, it is of interest to determine how the resistivity (conductivity) of MoTe2 in a magnetic field depends on temperature. In addition, when analyzing data on the Hall effect in topological semimetals, a single-band [19] or two-band model [27,28] is usually used. In [25], we showed that the estimates of the concentrations and mobilities of current carriers in WTe2 obtained using these two models are in good agreement with each other at 12 K. It is of interest to carry out such a comparison in a wider temperature range, in particular, using MoTe2 as an example.
This paper is devoted to the study of the features of the electro- and magnetotransport of the Weyl semimetal MoTe2 single crystal in order to establish the form of the temperature dependence of the resistivity (conductivity) in a magnetic field, compare the results of using single-band and two-band models to analyze its magnetotransport characteristics.

2. Materials and Methods

MoTe2 single crystal was grown by chemical vapor transport method [29] using bromine as a transport agent. Powders of molybdenum and tellurium in a stoichiometric ratio, as well as bromine, whose vapor density was 5 mg/cm3, were sealed into a quartz ampoule evacuated to a residual pressure of ~10−4 atm. Then the ampoule was placed into a horizontal tube furnace with a linear temperature gradient, where the temperatures T1 (“hot” zone) and T2 (“cold” zone) were 850 °C and 770 °C, respectively. The process of growing a single crystal was carried out for 500 h, followed by slow cooling to room temperature. It is known that MoTe2 undergoes a structural transition from the semiconductor α-phase to the metallic β-phase at a temperature of 820 °C for tellurium-rich samples and 880 °C for molybdenum-rich ones [14]. Therefore, the ampoule containing the grown crystal and evacuated to 10−4 atm. was heated to 910 °C and held for 3 h. In order to obtain the high-temperature phase, the quartz ampoule with crystal was rapidly cooled to room temperature by water quenching.
The grown crystal was investigated using X-ray diffractometer DRON-2.0 (Joint Stock Company “Bourevestnik”, Saint Petersburg, Russia.) with CrKα radiation. Fragments of the X-ray diffraction pattern taken from the surface of the sample immediately after growth, as well as after quenching, are shown in Figure 1a. Since all peaks can be indexed as (00l), the surface of the single crystal under study coincides with the (001) type plane. It can be seen that the intensity ratio of the lines from the (00l) planes changed after quenching. Note also that the lattice parameter c changed significantly from 13.93 ± 0.01 Å for the as-grown crystal to 13.81 ± 0.01 Å for the quenched one. The value of the lattice parameter c was estimated from the position of the (006) line. Figure 1b clearly shows the shift of the (006) line towards larger angles for the quenched sample. The obtained values of c before and after quenching are close to the values of lattice constant c for α-MoTe2 (hexagonal lattice) and β-MoTe2 (monoclinic lattice), respectively [14,30,31]. These results indicate a structural phase transition that occurred in MoTe2. The chemical composition of the sample was studied by energy-dispersive X-ray microanalysis using an Quanta 200 Pegasus scanning electron microscope (FEI Company, Eindhoven, the Netherlands) with an EDAX attachment at the Collaborative Access Center “Testing Center of Nanotechnology and Advanced Materials” of M.N. Mikheev Institute of Metal Physics of the Ural Branch of the Russian Academy of Sciences (IMP UB RAS). The chemical composition of the single crystal corresponds to stoichiometric MoTe2.
In this paper, we studied the quenched MoTe2 single crystal, which corresponded to the β-phase at room temperature and had the shape of a thin plate with dimensions ~6 × 1 × 0.2 mm3. The resistivity and Hall effect were measured by the four-contact method in the temperature range from 4.2 K to 290 K and in magnetic fields up to 9 T on a PPMS-9 setup (Quantum Design, San Diego, CA, USA) at the Collaborative Access Center “Testing Center of Nanotechnology and Advanced Materials” of IMP UB RAS. During the measurements, the electric current flowed in the (001) type plane, and the magnetic field was directed along the c axis, i.e., perpendicular to the plane of the plate. The residual resistivity ratio (RRR) in our sample is ρ300 K4.2 K ≈ 15, which is comparable to the RRR value in [21], but at the same time, much lower than the RRR in [27], which indicates a large number of defects in the crystal under study. For the convenience of interpretation, some of the experimental results obtained are presented in the form of conductivity σxx = [ρ(B) − ρ(0)]−1, where ρ(B) is the resistivity in magnetic field B.

3. Results and Discussion

3.1. Electrical Resistivity

The temperature dependence of the electrical resistivity ρ(T) of MoTe2, measured in the temperature range from 4.2 K to 290 K, is shown in Figure 2. It can be seen that the dependence ρ(T) has a “metallic” type, where ρ increases from 0.29 × 10−4 Ohm·cm to 4.2 × 10−4 Ohm·cm with increasing temperature. The ρ(T) curve shows a weak feature at a temperature of ~260 K. A similar behavior of the electrical resistivity near 250 K was observed in [21,22,23,27] and is associated with the structural phase transition from the monoclinic 1T’-phase (β-MoTe2) to the orthorhombic Td-one, which was reported in [15]. The inset shows the dependence ρ = f(T2). It can be seen that in the temperature range from 4.2 K to 45 K, the dependence ρ(T) can be represented as ρ = ρ0 + AT2. The coefficient A is 2.8 × 10−8 Ohm·cm·K−2 in our case, which coincides in order of magnitude with the value of 1.54 × 10−8 Ohm·cm·K−2 given in [21], where a quadratic temperature dependence of the electrical resistivity of MoTe2 was also observed in the temperature range from 1.7 to 50 K. Such a dependence ρ(T) at temperatures below ~10–15 K is usually explained by electron-electron scattering with a collision frequency proportional to T2. At higher temperatures, the electron-phonon scattering mechanism usually dominates. In this case, at T << ΘDD is the Debye temperature), the dependence ρ(T)~T5 can be observed, and at temperatures comparable to ΘD, ρ(T) is linear. In [27,32], where the transport characteristics of semimetallic MoTe2 were also studied, the data on the temperature dependence of the electrical resistivity were fitted, and it was found that there is a contribution to the resistivity proportional to T5. At the same time, for our crystal, as in [21,22,23], no contribution of ~T5 to the resistivity is observed at T << ΘD, where ΘD = 135 K was taken from [23].
The quadratic temperature dependence of the electrical resistivity in a wide temperature range was also observed in pure metals. Thus, tungsten single crystals were studied in the temperature range from 2 to 40 K in [33], where another mechanism was proposed that leads to the T2-dependence ρ(T), called the “electron-phonon-surface” interference scattering mechanism. In addition, the quadratic temperature dependence of resistivity (conductivity) in a magnetic field was observed in tungsten single crystals under conditions of the static skin effect [34]. In [25], we found that in the Weyl semimetal WTe2 single crystal, both the electrical resistivity in the absence of a magnetic field and the conductivity in a field depend on temperature according to a quadratic law in a wide temperature range from 12 K to ~70 K and ~55 K, respectively. By analogy with WTe2, it can be expected that the quadratic dependence of the resistivity (conductivity) of MoTe2 should also be observed in the presence of a magnetic field.

3.2. Resistivity (Conductivity) in Magnetic Field

Figure 3 shows the temperature dependence of the conductivity σxx = [ρ(B) − ρ(0)]−1 of the MoTe2 single crystal in a magnetic field of 9 T in the temperature range from 4.2 to 60 K. The inset shows the conductivity as σxx = f(T2). It can be seen that the value of σxx varies with temperature according to the quadratic law σxx = const + CT2 in a wide temperature range from 4.2 K to 60 K. In this case, two temperature ranges can be distinguished, namely (4.2–20) K and (20–60) K, where the coefficient C is 32.6 Ohm−1·cm−1·K−2 and 79.2 Ohm−1·cm−1·K−2, respectively. A similar change in the coefficients at T2 was observed by the authors of [33], where the electrical resistivity of tungsten single crystals was studied, and it was shown that the quadratic-in-temperature contribution to the resistivity at low temperatures is due to the scattering of conduction electrons by the sample surface, where the “electron-phonon-surface” interference scattering mechanism takes place. In addition, it was shown in [34] that in tungsten single crystals under conditions of a static skin effect, the conductivity in the magnetic field depends on temperature according to a quadratic law, which is also associated with the “electron-phonon-surface” scattering mechanism. Therefore, it can be assumed that the quadratic temperature dependence of σxx observed at temperatures from 4.2 K to 20 K in our single crystal might be associated with the “electron-phonon-surface” interference scattering mechanism [30,31]. In order to verify this, further studies are required, in particular, measurements of the resistivity of “sized” crystals (see, for example, [35] and references therein). Whereas the T2-dependence of the conductivity in a magnetic field from ~20 K to 60 K seems to be associated with contributions from various scattering mechanisms, primarily from the specific electron-phonon scattering process leading to T2.

3.3. Hall Effect

The analysis of data on the Hall effect in the MoTe2 single crystal under study was carried out in the framework of a single-band model. In order to calculate the Hall coefficient RH, concentration n, and mobility μ of main charge carriers, the following equations were used.
R H = ρ H B ,
n = 1 R H · e ,
μ = R H ρ ,
where ρH is the Hall resistivity; B is the magnetic field induction; e is the electron charge; ρ is the electrical resistivity in the absence of a magnetic field. Figure 4 shows the temperature dependences of RH, n, and μ in the MoTe2 single crystal. Since RH is negative, electrons are the majority of charge carriers. Their concentration and mobility at a temperature of 4.2 K are 2.6 × 1020 cm−3 and 0.8 × 103 cm2/(V·s), respectively. The value of n slightly changes with temperature. At the same time, the mobility μ decreases strongly with temperature, which can be explained by an increase in the scattering efficiency.
At the same time, it is known that MoTe2 contains carriers of both electron and hole types [10,11]. In such systems, the field dependences of the resistivity ρ in a magnetic field and Hall resistivity ρH are usually analyzed using a two-band model. In this case, the equations for ρ and ρH can be written in the form presented in [36]:
ρ = 1 e n h μ h + n e μ e + n h μ e + n e μ h μ h μ e B 2 n h μ h + n e μ e 2 + n h n e 2 μ h 2 μ e 2 B 2 ,
ρ H = B e n h μ h 2 n e μ e 2 + n h n e μ h 2 μ e 2 B 2 n h μ h + n e μ e 2 + n h n e 2 μ h 2 μ e 2 B 2 ,
where ne (μe) and nh (μh) are the concentration (mobility) of electrons and holes, respectively.
Figure 5a shows the field dependences of the resistivity ρ(B) in a magnetic field and Hall resistivity ρH(B) of the MoTe2 single crystal at temperatures of 4.2 K, 15 K, 25 K, and 50 K. The solid red lines correspond to fitting curves obtained using Equations (4) and (5) within the framework of the two-band model. It can be seen that the experimental data are well described by the fitting curves. The obtained values of the concentrations and mobilities of electrons and holes depending on the temperature are shown in Figure 5b. The values of ne are of the order of 1020 cm−3 and vary slightly with temperature, which is consistent with the data obtained within the single-band model (Figure 4). At the same time, the hole concentration nh decreases drastically with increasing temperature above 25 K. The electron mobility μe decreases with increasing temperature over the entire temperature range studied, as in the case of the single-band model. Whereas the hole mobility μh decreases with increasing temperature up to 25 K, at higher temperatures, an increase in the value of μh is observed. At T = 4 K, the values of μe and μh are 1.10 × 103 cm2/(V·s) and 0.58 × 103 cm2/(V·s), respectively. The geometric-mean mobility μ = μ e μ h is in good agreement with the carrier mobility obtained within the single-band model. It can be seen that the values of the concentration ne and mobility μe of electrons mainly exceed the values of nh and μh of holes. This means that electrons are the main type of carriers, which also agrees with the single-band model. Note that, in our case, the carrier mobility, according to the estimates made, is an order of magnitude lower than, for example, in [27]. Apparently, this is due to a large number of defects in our crystal and a lower RRR. At the same time, the qualitative behavior of the obtained ne (μe) and nh (μh) are in good agreement with the results given in [22,27], where the two-band model was used. In addition, theoretical calculations of the electronic structure at various temperatures were carried out in [22], and it was shown that at temperatures below 35 K, electron-hole compensation is observed in Td-MoTe2, whereas at higher temperatures, the Fermi surface is reconstructed mainly due to a decrease in the volume of hole pockets. This agrees with our results presented in Figure 5b, where at temperatures from 4.2 to 25 K nenh, while at higher temperatures, the hole concentration nh strongly decreases.
In [25], the concentrations and mobilities in WTe2 were also estimated using single-band and two-band models. It was shown that the values of concentrations and mobilities obtained using both models are in good agreement with each other. In [25] the Hall coefficient was calculated using the two-band model. The obtained value of RH is in good agreement with the value obtained from the experimental data based on the single-band model. Therefore, the Hall coefficient was also calculated for MoTe2 using the two-band model. The values obtained at temperatures of 4.2 K, 15 K, 25 K, and 50 K are plotted in Figure 4. It can be seen that the Hall coefficients calculated for both models are in good agreement. Apparently, this may indicate that the two-band model is applicable to systems similar to MoTe2, containing different groups of charge carriers that differ in sign, mobility, etc. Moreover, the two-band model is preferable to the single-band one.

4. Conclusions

  • The quadratic temperature dependence of the electrical resistivity of MoTe2 was observed in a wide temperature range from 4.2 K to 45 K, which is consistent with the experimental results previously reported.
  • The quadratic temperature dependence of the conductivity in a magnetic field was found in a wide temperature range. Moreover, two intervals can be distinguished: “low-temperature” and “high-temperature”. It has been suggested that in the low-temperature range, the quadratic dependence of the conductivity in a magnetic field might be associated with the “electron-phonon-surface” interference scattering mechanism.
  • The analysis of data on the Hall effect in MoTe2 was carried out using single-band and two-band models. The values of concentration and mobility of current carriers were estimated. The Hall coefficients calculated from both models are in good agreement. Apparently, the two-band model is preferable in such systems containing different groups of current carriers.

Author Contributions

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

Funding

The results of studies of electrical resistivity (Section 3.1) were obtained within the state assignment of the Ministry of Science and Higher Education of the Russian Federation (theme “Spin,” No. 122021000036-3), supported in part by the scholarship of the President of the Russian Federation to young scientists and graduate students (A.N.P., SP-2705.2022.1). Studies of magnetoresistivity (Section 3.2) and the Hall effect (Section 3.3) were supported by the Russian Science Foundation (project no. 22-42-02021).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The authors thank S.M. Podgornykh, V.V. Chistyakov, and E.B. Marchenkova for help in the measurements and study of the structure of the sample.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Armitage, N.P.; Mele, E.J.; Vishwanath, A. Weyl and Dirac semimetals in three-dimensional solids. Rev. Mod. Phys. 2018, 90, 015001. [Google Scholar] [CrossRef] [Green Version]
  2. Lv, B.Q.; Qian, T.; Ding, H. Experimental perspective on three-dimensional topological semimetals. Rev. Mod. Phys. 2021, 93, 025002. [Google Scholar] [CrossRef]
  3. Hasan, M.Z.; Kane, C.L. Colloquium: Topological insulators. Rev. Mod. Phys. 2010, 82, 3045–3067. [Google Scholar] [CrossRef] [Green Version]
  4. Liu, Y.; Chong, C.; Chen, W.; Huang, J.A.; Cheng, C.; Tsuei, K.; Li, Z.; Qiu, H.; Marchenkov, V.V. Growth and characterization of MBE-grown (Bi1− xSbx)2Se3 topological insulator. Jpn. J. Appl. Phys. 2017, 56, 070311. [Google Scholar] [CrossRef]
  5. Yan, B.; Felser, C. Topological Materials: Weyl Semimetals. Annu. Rev. Condens. Matter Phys. 2017, 8, 337–354. [Google Scholar] [CrossRef] [Green Version]
  6. Xu, S.-Y.; Belopolski, I.; Alidoust, N.; Neupane, M.; Bian, G.; Zhang, C.; Sankar, R.; Chang, G.; Yuan, Z.; Lee, C.-C.; et al. Discovery of a Weyl Fermion semimetal and topological Fermi arcs. Science 2015, 349, 613–617. [Google Scholar] [CrossRef] [Green Version]
  7. Soluyanov, A.A.; Gresch, D.; Wang, Z.; Wu, Q.; Troyer, M.; Dai, X.; Bernevig, B.A. Type-II Weyl semimetals. Nature 2015, 527, 495–498. [Google Scholar] [CrossRef] [Green Version]
  8. Wu, Y.; Mou, D.; Jo, N.H.; Sun, K.; Huang, L.; Bud’ko, S.L.; Canfield, P.C.; Kaminski, A. Observation of Fermi arcs in the type-II Weyl semimetal candidate WTe2. Phys. Rev. B 2016, 95, 121113. [Google Scholar] [CrossRef] [Green Version]
  9. Wang, C.; Zhang, Y.; Huang, J.; Nie, S.; Liu, G.; Liang, A.; Zhang, Y.; Shen, B.; Liu, J.; Hu, C.; et al. Observation of Fermi arc and its connection with bulk states in the candidate type-II Weyl semimetal WTe2. Phys. Rev. B 2016, 94, 241119. [Google Scholar] [CrossRef] [Green Version]
  10. Sun, Y.; Wu, S.-C.; Ali, M.N.; Felser, C.; Yan, B. Prediction of Weyl semimetal in orthorhombic MoTe2. Phys. Rev. B 2015, 92, 161107. [Google Scholar] [CrossRef]
  11. Deng, K.; Wan, G.; Deng, P.; Zhang, K.; Ding, S.; Wang, E.; Yan, M.; Huang, H.; Zhang, H.; Xu, Z.; et al. Experimental observation of topological Fermi arcs in type-II Weyl semimetal MoTe2. Nat. Phys. 2016, 12, 1105–1110. [Google Scholar] [CrossRef] [Green Version]
  12. Chernozatonskii, L.A.; Artukh, A.A. Quasi-two-dimensional transition metal dichalcogenides: Structure, synthesis, properties and applications. Phys. Usp. 2018, 61, 2. [Google Scholar] [CrossRef]
  13. Kim, H.-J.; Kang, S.-H.; Hamada, I.; Son, Y.-W. Origins of the structural phase transitions in MoTe2 and WTe2. Phys. Rev. B 2017, 95, 180101. [Google Scholar] [CrossRef] [Green Version]
  14. Vellinga, M.B.; de Jonge, R.; Haas, C. Semiconductor to Metal Transition in MoTe2. J. Solid State Chem. 1970, 2, 299–302. [Google Scholar] [CrossRef]
  15. Clarke, R.; Marseglia, E.; Hughes, H.P. A low-temperature structural phase transition in β-MoTe2. Philos. Mag. B 1978, 38, 121–126. [Google Scholar] [CrossRef]
  16. Ali, M.N.; Xiong, J.; Flynn, S.; Tao, J.; Gibson, Q.D.; Schoop, L.M.; Liang, T.; Haldolaarachchige, N.; Hirschberger, M.; Ong, N.P.; et al. Large, non-saturating magnetoresistance in WTe2. Nature 2014, 514, 205–208. [Google Scholar] [CrossRef] [Green Version]
  17. Keum, D.H.; Cho, S.; Kim, J.H.; Choe, D.-H.; Sung, H.-J.; Kan, M.; Kang, H.; Hwang, J.-Y.; Kim, S.W.; Yang, H.; et al. Bandgap opening in few-layered monoclinic MoTe2. Nat. Phys. 2015, 11, 482–486. [Google Scholar] [CrossRef]
  18. Li, P.; Wen, Y.; He, X.; Zhang, Q.; Xia, C.; Yu, Z.-M.; Yang, S.A.; Zhu, Z.; Alshareef, N.H.; Zhang, X.-X. Evidence for topological type-II Weyl semimetal WTe2. Nat. Commun. 2017, 8, 2150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Shekhar, C.; Nayak, A.K.; Sun, Y.; Schmidt, M.; Nicklas, M.; Leermakers, I.; Zeitler, U.; Skourski, Y.; Wosnitza, J.; Liu, Z.; et al. Extremely large magnetoresistance and ultrahigh mobility in the topological Weyl semimetal candidate NbP. Nat. Phys. 2015, 11, 645–649. [Google Scholar] [CrossRef]
  20. Liang, T.; Gibson, Q.; Ali, M.N.; Liu, M.; Cava, R.J.; Ong, N.P. Ultrahigh mobility and giant magnetoresistance in the Dirac semimetal Cd3As2. Nat. Mater. 2015, 14, 280–284. [Google Scholar] [CrossRef]
  21. Zandt, T.; Dwelk, H.; Janowitz, C.; Manzke, R. Quadratic temperature dependence up to 50 K of the resistivity of metallic MoTe2. J. Alloys Compd. 2007, 442, 216–218. [Google Scholar] [CrossRef]
  22. Chen, F.C.; Lv, H.Y.; Luo, X.; Lu, W.J.; Pei, Q.L.; Lin, G.T.; Han, Y.Y.; Zhu, X.B.; Song, W.H.; Sun, Y.P. Extremely large magnetoresistance in the type-II Weyl semimetal MoTe2. Phys. Rev. B 2016, 94, 235154. [Google Scholar] [CrossRef] [Green Version]
  23. Chen, F.C.; Luo, X.; Xiao, R.C.; Lu, W.J.; Zhang, B.; Yang, H.X.; Li, J.Q.; Pei, Q.L.; Shao, D.F.; Zhang, R.R.; et al. Superconductivity enhancement in the S-doped Weyl semimetal candidate MoTe2. Appl. Phys. Lett. 2016, 108, 162601. [Google Scholar] [CrossRef] [Green Version]
  24. Lv, Y.-Y.; Cao, L.; Li, X.; Zhang, B.-B.; Wang, K.; Pang, B.; Ma, L.; Lin, D.; Yao, S.-H.; Zhou, J.; et al. Composition and temperature dependent phase transition in miscible Mo1−xWxTe2 single crystals. Sci. Rep. 2017, 7, 44587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Perevalova, A.N.; Naumov, S.V.; Podgornykh, S.M.; Chistyakov, V.V.; Marchenkova, E.B.; Fominykh, B.M.; Marchenkov, V.V. Kinetic properties of topological semimetal WTe2 single crystal. Phys. Met. Metallogr. 2022; 123, in press. [Google Scholar]
  26. Kaveh, M.; Cherry, M.F.; Weger, M. New mechanism for the electrical resistivity of layer compounds: TiS2. J. Phys. C Solid State Phys. 1981, 14, L789–L795. [Google Scholar] [CrossRef]
  27. Zhou, Q.; Rhodes, D.; Zhang, Q.R.; Tang, S.; Schonemann, R.; Balicas, L. Hall effect within the colossal magnetoresistive semimetallic state of MoTe2. Phys. Rev. B 2016, 94, 121101. [Google Scholar] [CrossRef] [Green Version]
  28. Lee, S.; Jang, J.; Kim, S.I.; Jung, S.-G.; Kim, J.; Cho, S.; Kim, S.W.; Rhee, J.Y.; Park, K.-S.; Park, T. Origin of extremely large magnetoresistance in the candidate type-II Weyl semimetal MoTe2−x. Sci. Rep. 2018, 8, 13937. [Google Scholar] [CrossRef] [Green Version]
  29. Levy, F. Single-crystal growth of layered crystals. II Nuovo Cim. B 1977, 38, 359–368. [Google Scholar] [CrossRef]
  30. Brown, B.E. The Crystal Structures of WTe2 and High-Temperature MoTe2. Acta Cryst. 1966, 20, 268–274. [Google Scholar] [CrossRef]
  31. Al-Hilli, A.A.; Evans, B.L. The preparation and properties of transition metal dichalcogenide single crystals. J. Cryst. Growth 1972, 15, 93–101. [Google Scholar] [CrossRef]
  32. Bhattarai, N.; Forbes, A.W.; Dulal, R.P.; Pegg, I.L.; Philip, J. Transport characteristics of type II Weyl semimetal MoTe2 thin films grown by chemical vapor deposition. J. Mater. Res. 2020, 35, 454–461. [Google Scholar] [CrossRef] [Green Version]
  33. Startsev, V.E.; D’yakina, V.P.; Cherepanov, V.I.; Volkenshtein, N.V.; Nasyrov, R.S.; Manakov, V.G. Quadratic temperature dependence of the resistivity of tungsten single crystals. Role of surface scattering of electrons. Sov. Phys. JETP 1980, 52, 675–679. [Google Scholar]
  34. Marchenkov, V.V. Quadratic temperature dependence of magnetoresistivity of pure tungsten single crystals under static skin effect. Low Temp. Phys. 2011, 37, 852–855. [Google Scholar] [CrossRef]
  35. Marchenkov, V.V.; Weber, H.W. Size Effect in the High-Field Magnetoconductivity of Pure Metal Single Crystals. J. Low Temp. Phys. 2003, 132, 135–144. [Google Scholar] [CrossRef]
  36. Luo, Y.; Li, H.; Dai, Y.M.; Miao, H.; Shi, Y.G.; Ding, H.; Taylor, A.J.; Yarotski, D.A.; Prasankumar, R.P.; Thompson, J.D. Hall effect in the extremely large magnetoresistance semimetal WTe2. Appl. Phys. Lett. 2015, 107, 182411. [Google Scholar] [CrossRef]
Figure 1. Fragments of X-ray diffraction patterns taken from the surface of the MoTe2 single crystal: (a) The as-grown single crystal corresponds to α-MoTe2; The single crystal quenched from 910 °C corresponds to β-MoTe2; (b) Change of lattice parameter c of MoTe2 after quenching.
Figure 1. Fragments of X-ray diffraction patterns taken from the surface of the MoTe2 single crystal: (a) The as-grown single crystal corresponds to α-MoTe2; The single crystal quenched from 910 °C corresponds to β-MoTe2; (b) Change of lattice parameter c of MoTe2 after quenching.
Metals 12 02089 g001
Figure 2. Temperature dependence of the electrical resistivity of MoTe2 in the temperature range from 4.2 K to 290 K. The inset shows the dependence ρ = f(T2) at temperatures from 4.2 K to 60 K.
Figure 2. Temperature dependence of the electrical resistivity of MoTe2 in the temperature range from 4.2 K to 290 K. The inset shows the dependence ρ = f(T2) at temperatures from 4.2 K to 60 K.
Metals 12 02089 g002
Figure 3. Temperature dependence of the conductivity of MoTe2 in a magnetic field of 9 T in the temperature range from 4.2 K to 60 K. The inset shows the dependence σxx = f(T2) at temperatures from 4.2 K to 60 K.
Figure 3. Temperature dependence of the conductivity of MoTe2 in a magnetic field of 9 T in the temperature range from 4.2 K to 60 K. The inset shows the dependence σxx = f(T2) at temperatures from 4.2 K to 60 K.
Metals 12 02089 g003
Figure 4. Temperature dependences of the Hall coefficient RH, concentration n, and mobility µ of current carriers in MoTe2, obtained using the single-band model in a field B = 9 T. The red triangles show the values of the Hall coefficient obtained on the basis of the two-band model.
Figure 4. Temperature dependences of the Hall coefficient RH, concentration n, and mobility µ of current carriers in MoTe2, obtained using the single-band model in a field B = 9 T. The red triangles show the values of the Hall coefficient obtained on the basis of the two-band model.
Metals 12 02089 g004
Figure 5. Analysis of data on the Hall effect in MoTe2 using a two-band model: (a) Field dependences of the Hall resistivity ρH(B) (lower curve) and resistivity ρ (B) (upper curve) in a magnetic field for MoTe2 at temperatures of 4.2 K, 15 K, 25 K, and 50 K. Open symbols are experimental data; solid red lines are fitting curves obtained using the two-band model; (b) Temperature dependences of the concentration and mobilities of electrons and holes obtained on the basis of the two-band model.
Figure 5. Analysis of data on the Hall effect in MoTe2 using a two-band model: (a) Field dependences of the Hall resistivity ρH(B) (lower curve) and resistivity ρ (B) (upper curve) in a magnetic field for MoTe2 at temperatures of 4.2 K, 15 K, 25 K, and 50 K. Open symbols are experimental data; solid red lines are fitting curves obtained using the two-band model; (b) Temperature dependences of the concentration and mobilities of electrons and holes obtained on the basis of the two-band model.
Metals 12 02089 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Perevalova, A.N.; Naumov, S.V.; Marchenkov, V.V. Peculiarities of the Electro- and Magnetotransport in Semimetal MoTe2. Metals 2022, 12, 2089. https://doi.org/10.3390/met12122089

AMA Style

Perevalova AN, Naumov SV, Marchenkov VV. Peculiarities of the Electro- and Magnetotransport in Semimetal MoTe2. Metals. 2022; 12(12):2089. https://doi.org/10.3390/met12122089

Chicago/Turabian Style

Perevalova, Alexandra N., Sergey V. Naumov, and Vyacheslav V. Marchenkov. 2022. "Peculiarities of the Electro- and Magnetotransport in Semimetal MoTe2" Metals 12, no. 12: 2089. https://doi.org/10.3390/met12122089

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