Next Article in Journal
Thermal Conductivity and Orientation Structure of Liquid Crystalline Epoxy Thermosets Prepared by Latent Curing Catalyst
Previous Article in Journal
Investigation of the Electronic Properties of Silicon Carbide Films with Varied Si/C Ratios Annealed at Different Temperatures
Previous Article in Special Issue
Global Instability of Rod Eutectic Growth in Directional Solidification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Single Crystal Growth and Nano-Structure Study in a Topological Dirac Metal, CoTe2-δ

1
School of Physics and Materials Science, Guangzhou University, Guangzhou 510006, China
2
Department of Materials Science & Engineering, Monash University, Clayton, VIC 3800, Australia
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(1), 46; https://doi.org/10.3390/cryst14010046
Submission received: 24 November 2023 / Revised: 22 December 2023 / Accepted: 27 December 2023 / Published: 28 December 2023

Abstract

:
A single crystal of a topological material, CoTe2-δ, has been grown via the chemical vapor transport method for a structural and electronic transport study. Single-crystal X-ray diffraction, powder X-ray diffraction, and high-resolution scanning electron microscope measurements confirm the high quality of the as-grown single crystals. In a high-resolution scanning electron microscopy study, a clear layered feature of the trigonal CoTe2-δ crystal was observed. Fractal features and mosaic-type nanostructures were observed on the as-grown surface and cleaved surface, respectively. The trigonal CoTe2-δ demonstrates a metallic ground state in transport measurements, with a typical carrier’s concentration in a 1021 cm−3 magnitude and a residual resistivity ratio of 1.6. Below 10 K, trigonal CoTe2-δ contains quite complicated magnetoresistance behavior as a result of the competing effect between Dirac states and possible spin fluctuations.

1. Introduction

Transition metal dichalcogenides (TMDs) have drawn tremendous attention in both fundamental and applicational studies. In the TMD family, there are over 40 two-dimensional (2D) compounds with the general formula of MX2, where M is a transition metal and X is a chalcogen such as S, Se, or Te. Bulk TMDs have been widely studied over several decades for its crystal, electronic, and magnetic structures. The van der Waal bonds hold 2D layers along the c axis, in which the transition metal is sandwiched between two chalcogens, forming a stable structure. Therefore, like graphene exfoliated from graphite [1,2], a monolayer or few-layer TMD can be exfoliated from the bulk crystal, which could be effective wide-band semiconductors [3,4,5,6,7,8,9], highly efficient photoelectric units [10,11,12,13,14,15,16], 2D superconductors [7,17,18,19,20], building blocks in van der Waals heterostructures [21,22,23,24,25,26], and moiré superlattices [27,28,29,30,31,32,33,34,35,36,37,38].
The electronic structure of TMD bulk crystals are studied a lot in metallic or semi-metallic TMDs, such as WTe2, PdTe2, PtTe2, PtSe2, NiTe2, etc., in which the symmetry protected topological band structures are reported [39,40,41]. Topology is employed as a new type of category method on electronic band structures [42,43]. The band topology research started on the integer quantum Hall effect in two-dimensional electron gas (2DEG). Once the 2DEG is restricted in one dimension and works as an electron wave conductor, the Landau levels will change at the edges. In the center of the electron wave conductor, the LL energy is independent of the position, while near the edges, it is raised in response to the edge’s potential energy. Giving a certain Fermi energy EF, the boundary of the occupied LLs must pass through the EF, which leaves the edge conduction channels (or edge states). This effect can only be understood using the concept of topology. In the QHE states, the bulk states are insulating, and the edge modes are free-way-like one-dimensional (1D) conduction channels, which geographically separate the carriers with opposite momentum. In conventional 1D transport, the resistance originates from the backscattering mechanism, which mixes together the forward- and backward-moving charges. In integer quantum Hall states, however, the forward-moving charges and backward-moving charges are located at the different edges of the system, and therefore survive from the backscattering mechanism. A 2D band structure consists of mapping from the crystal momentum to the Bloch Hamiltonian. Gapped band structures can be classified topologically by considering the equivalence classes of Bloch Hamiltonian that can be continuously deformed into one another without closing the energy gap. These classes are distinguished by a topological invariant n called the Chern invariant. In quantum mechanics, a geometric phase is a phase difference acquired over a closed path, which could be nontrivial, and it was discovered by Sir Michael Berry, namely the Berry phase [44]. Another interesting quantized Hall state is the quantized version of the spin Hall effect, namely the quantum spin Hall effect (QSHE). The QSHE can be understood as the combination of spin-up QHE and spin-down QHE with its own chirality. Therefore, with the charging canceling, the total Hall conductance is zero for the QSE; however, it still presents a non-zero spin Hall conductance. At first, C. L. Kane and E. J. Mele [45] proposed a new spin orbital coupling (SOC) term in the Hamiltonian system related to Haldane’s model. The SOC Hamiltonian can be taken separately for the Sz = 1 spins, which are equivalent to Haldane’s model for spinless electrons, and also violate time reversal symmetry. This model suggests that the SOC introduces a periodic magnetic field with no net flux in the system, giving rise to a finite energy gap. At temperatures well below the gap, the double-spin Haldane’s model finally leads to a quantized Hall conductance to each spin. In this situation, the edge modes are helical spin modes, which are the spin-up electron and spin-down electron with opposite momentum. Provided that the time reversal symmetry of the system (TMS) is preserved, this acts to protect the edge modes from nonmagnetic impurity and disorder. The scattered electrons have a 2π phase difference for the presence of TMS. In quantum mechanics, when the spin 1/2 particles rotate by 2π, the wave function has a changed sign. However, in the presence of magnetic impurity, the time reversal symmetry is broken, as well as the backscattering immune transport. After the investigation in different quantum Hall states, topology was studied in solid-state materials, including, briefly, (1) a two-dimensional topological insulator, where the normal bands and inversed bands overlap in some materials (e.g., the CdTe/HgTe quantum well, in which the strong spin-orbit coupling (SOC) opens a gap at the bulk bands and leaves the bulk-symmetry protected metallic edge states); (2) a three-dimensional topological insulator (3DTI), which is the 3D form of a topological insulator, which possesses gapped bulk states and symmetry-protected metallic surface states; (3) Dirac/Weyl Semimetals, which have the bulk-symmetry protected four-fold band crossing (degeneration) points (Dirac points are found in a Dirac semimetals) and two-fold degenerate Weyl points. The Weyl points can be obtained by broken inversion symmetry, namely a Weyl semimetal, or by broken time-reversal symmetry, namely a (4) magnetic Weyl semimetal. The WTe2 is very famous for its type-Ⅱ Weyl semimetal states (with tilted Weyl cone pairs as signature) in bulk crystals [39,40], as well as 2D topological insulating states in monolayer form [46,47]. The PdTe2 and PtTe2 hold type-Ⅱ Dirac fermions with a tilted Dirac cone near its Fermi energy due to the violation of the Lorentz symmetry [41,48]. Similarly, the TMD compound CoTe2 has also attracted increasing attention recently [49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64]; it was first known as an effective electrocatalyst for hydrogen evolution reaction with low overpotentials and small Tafel slopes. The orthorhombic CoTe2 is reported as a quantum critical Dirac metal with strong spin fluctuations [65]. The ferromagnetic ordering is suppressed, even at low temperatures, down to 1.8 K, and instead, a strong spin fluctuation state is proposed. The electronic band structure in orthorhombic CoTe2 was predicted to harbor multiple Dirac-type nodal lines and conventional Dirac points [65]. A recent report also mentioned that the Jahn–Teller effect induces a slight distortion in the CoTe2 crystal, which enriches the active site exposure in lithium–sulfur batteries [66]. The trigonal CoTe2 has slightly different topology in its band structures, which is featured with a pair of type-Ⅱ Dirac nodes that are located along the Γ-A direction above the Fermi energy [67,68]. In addition, multiple bulk band inversions are located in the Brillouin zone, which give rise to a ladder of spin-polarized surface states over a wide range of energies. The surface states corresponding to the bulk band inversions form a Dirac cone at the Γ point, which has in-plane spin splitting due to Rashba-type spin–orbit interactions [67]. However, the magnetotransport properties in trigonal CoTe2 single crystals, which will further confirm the novel electronic states, have not reported in the literature.
In this article, we reported the single crystal growth, microstructural characterization, and electronic transport properties in CoTe2-δ. The mosaic-like nanostructure embedded in layered crystals is exceptionally rare in TMD crystals. Negative magnetoresistance is observed at 2 K, which possibly indicates the frustrated spin dynamics.

2. Experiments

The CoTe2-δ single crystals are grown using a chemical vapor transport (CVT) method, which is very popular in the single crystal growth of quantum materials [69,70,71,72,73], as demonstrated in Figure 1. The chemical vapor transport summarizes a variety of reactions that show one common feature: a solid-state material is volatilized in the presence of a gaseous reactant, the transport agent, and deposits elsewhere, usually in the form of single crystals. Usually, the chemical reactions happen in a closed environment, which relies on a suitable transport agent and the experimental condition. Therefore, it is a prerequisite for a transport effect in which the transport agent reacts with the source material with the formation of one or more gaseous compounds. In this case, halogens or halogen compounds are usually used as transport agents. Due to the chemical reaction that happens at the interface of solid and gas states, a relatively lower temperature condition is employed more than other crystal methods. Therefore, the CVT method shows significant advantage in high-melting-point crystal growth. In this experiment, we chose TeCl4 as the transport agent. Briefly, high-purity Co (99.9%) and Te (99.99%) powders in a 1:2 molar ratio (~1 g in total) are sealed in a silica tube in a high vacuum, together with 5 mg/mL of TeCl4 as transport agent. The actual transport agent’s mass (TeCl4) was calculated based on the volume of the sealed silica tube.
It is worth mentioning that the CVT method is highly flexible in quantum material designs. For example, Z. Li et al. [74] developed a CVT-based method to prepare a large-sized Weyl semimetal tantalum arsenide crystal, which has a higher melting point than its decomposition temperature. In this case, large-size crystal growth methods like the Bridgeman method are not allowed. The sample powders and I2 transport agent are sealed in an ampule, which is slightly tilted up to 30° from the horizon to enhance the growth rate. Particularly, one piece of polished tantalum foil was placed in the crystallization side to retard the spontaneous nucleation. In their report, a 1 cm TaAs single crystal was collected on the tantalum foil instead of small 1–3 mm crystals. Moreover, the CVT method can also be modified into a high-throughput method for doped quantum material system, e.g., topological semimetal Ta1-xNbxSb2 [69]. In their experiments, a doping ratio of Ta:Nb at 1:1 was employed in the starting materials at a low-temperature site. After a fine-tuned growth condition is established, single crystals with various Ta:Nb ratios can be obtained. In the crystal counting vs. chemical composition plot, a Gaussian-like distribution was observed, which supports the experimental research in quantum materials.
The sealed silica tube is placed in a dual temperature zone horizontal furnace. The Co and Te powders sit in the hot side (930 °C), which will be transported to the low-temperature site (870 °C) over one week into CoTe2 single crystals. The obtained single crystals are of hexagonal shape and have a metallic shiny color (Figure 2a), with a ~3 mm lateral dimension and ~0.3 mm thickness. Approximately 10 pieces of high-quality single crystals can be obtained from each silica tube. The X-ray diffraction (XRD) technique (Bruker D8 Advance Eco with cobalt radiation Kα = 1.79 Å) is employed to determine the crystal quality and orientation. The microstructure of the as-grown and cleaved crystal is characterized using scanning electron microscopy (SEM, ThermoFisher ScientificFEI Helios 5, Waltham, MA, United States, The elemental concentration mapping is conducted using an Oxford Instrument energy-dispersive spectroscopy (EDS) system equipped in Helios 5. Standard Hall-bar contacts are patterned on a freshly cleaved single crystal surface using room temperature cured silver paste for electronic transport measurements. A Physical Property Measurement System (PPMS, Quantum Design) is employed to conduct the temperature and magnetic field-dependent transport measurements in the CoTe2-δ single crystal.

3. Results and Discussion

There are three different types of CoTe2 that were investigated, namely trigonal CoTe2, cubic CoTe2, and orthorhombic CoTe2. The single-crystal XRD pattern taken from a freshly cleaved sample surface verifies the trigonal phase, as shown in Figure 2. Thus, the obtained single crystals are hexagonal thin plates, as demonstrated in Figure 2b. Trigonal CoTe2-δ crystallizes in a LiTiS2-type structure (space group P 3 ¯ m 1 ), which has a typical van der Waals layered structure, and can be exfoliated along the c axis [55,57,67,68]. The cleavage faces in CoTe2 is therefore verified as c surfaces, as indicated by the sharp (00l) patterns in Figure 2c. The powder XRD patterns of CoTe2-δ is shown in Figure 2d. The peak intensity ratio between (101) and (002) shows a preferred orientation in the prepared powder sample, which is due to the 2D crystal structure. The standard diffraction results in the PDF card #89-4061 are plotted in Figure 2e with clear indices. Due to the different X-ray source materials used in the current experiments and PDF card, Q = 4 π   s i n θ λ was employed in Figure 2. Figure 2b shows the low-magnification SEM image via the secondary electron mode, in which one can see the overall hexagonal shape in the spontaneous crystallization procedure in the CVT set-up.
To verify the chemical composition of the as-grown single crystal, the EDS spectrum is taken on a fairly large cleaved surface area, as shown in Figure 3. The second electron image and back-scattering electron image of the same area are shown in Figure 3a,b, with an applied accelerating voltage of 18 kV. A clear layered structure can be observed in Figure 3a, in which the highlighted lines indicate the edge of the layers, and the shadowing effect can be observed next to the highlighted region. Secondary electron (SE) imaging is carried out using the surface morphology sensitive technique, and therefore shows the details of the cleaving steps. At the same area, the back-scattering electron (BSE) image shows much lower contrast, which is due to the much deeper detecting depth in back-scattering mode. Note that the BSE mode is very sensitive to crystal orientation, chemical composition, and impurity phase. The similar feature with lower contrast in the BSE images implies that the cleaved surface of CoTe2 is from a single-crystalline phase without impurities within the resolution. In the EDS measurements, the detection is deeper than the BSE mode, and therefore it is nearly featureless in Figure 3e,f, which show the Co Kα1 and Te Kα1 mapping, respectively. The quantified analysis of the EDS spectrum on the whole surface area is shown in Figure 3c, in which the Co:Te atomic ratio of 36.1:63.9 (close to CoTe1.77) is confirmed. In a previous study, the stabilized layered structure was reported in a Te vacancy phase (CoTe1.8) [49], which supports the finding in our CVT-grown layered CoTe2. The reason for the off-stoichiometric ratio is possibly due to the deviation from a standard CdI2 structure, in which the metal–metal bonding and interlayer covalent interactions are detected [49]. This is quite unusual for a traditional layered compound.
The surface morphologies of the CoTe2-δ single crystals are further studied with the low-voltage (5 kV) SE method to demonstrate the details of the as-grown and cleaved surfaces (Figure 4). A low magnification SE image is shown in Figure 4a, in which the dark region beyond the hexagonal crystal edges is carbon tap. The crystal is cleaved using a blue tape before imaging, which reveals a cleaved area in the bottom left region, as indicated by the arrow. To show the zoomed-in view of the surface, a double-focused method (Emersion Mode) equipped in Helios 5 is employed in Figure 4b–d,f–h. The zoomed-in views of the as-grown surface are shown in Figure 4b–d, in which Panel b demonstrates the overall flatness in a relatively large area. There are a few contamination dots on the as-grown surface with a typical dimension in several tens to hundreds of nanometers. In a further zoomed-in view of the surface area, one can observe two typical surface morphologies, as demonstrated in Figure 4c,d, which were taken from the close-to-crystal-edge area and far-from-crystal-edge area, respectively. Figure 4c demonstrates the nanoscale morphology close to the crystal edges, in which the piled-up triangles with relatively thick edge areas are observed. A higher contrast in Figure 4c is observed in this region than the contrast in the middle area of the crystals, as shown in Figure 4d. Zig-zag patterns with 60° or 120° angles are observed in the few hundred nanometers region. These fractal hexagonal features on the as-grown single-crystal surface are dominant by the crystal growth dynamics, similar to the observation in the metal-flux method. The cleaved surface shows a quite different morphology compared with the as-grown surfaces in two aspects: (1) a large, flat region without nano-size contamination is observed; (2) a clear layer feature with steps is observed in Figure 4e. With a further zoomed-in view, nano-mosaic patterns are observed everywhere on the cleaved surfaces, as demonstrated in Figure 4f–h. The mosaics, typically 20–50 nm, contain curved boundaries with different widths, showing a dry-lake-bottom-like pattern. The reason for the mosaic feature contrast in Figure 4f–h is unknown, which is also observed in flat regions (Figure 4g) and is different with the step-induced contrast in Figure 4e. As mentioned by the referee in the peer-reviewing process, a complicated chemical origination of this nanostructure is deduced. Therefore, the elemental concentration analysis method with nanometer resolution are desired to conduct further characterizations in these crystals. The surface patterns observed in Figure 4 are universal for all CVT-grown CoTe2-δ single crystals using the aforementioned method. The rich surface nanostructures in trigonal CoTe2-δ could contribute to the excellent electrocatalysis properties [54,56,58,59,63].
We then study the electronic transport properties of the CoTe2 single crystals with nanoscale features using Hall-bar geometry patterned on a freshly cleaved crystal surface. During the measurements, the current was applied in the c plane, and the magnetic field was applied along the c axis of the crystal. Temperature (RT)- and magnetic field (MR)-dependent longitudinal resistivity are measured using the electronic transport function in PPMS, as shown in Figure 5 and Figure 6. The resistivity is ~205 µΩ cm at room temperature, and decreases monotonically with cooling to 2 K, yielding a residual resistivity ratio of ~1.6. The longitudinal room temperature resistivity is ~180 µΩ cm in a four-terminal measurement [65] in orthorhombic CoTe2, and ~2.5–42 mΩ cm in two-terminal measurements in trigonal CoTe2 nanoflakes [55]. The two terminal results contain contact resistance between CoTe2 and gold contacts, which is significantly higher than the observation here. Moreover, the similar metallic RT behavior was also reported in orthorhombic CoTe2, with comparable resistivity in the 2–300 K region [65]. The other similar compound, NiTe2, however, possesses a much larger RRR value (~109) and lower room temperature resistivity (~26 µΩ cm) [75]. The tellurium vacancy-induced structural change could be a reason for the different transport behaviors between NiTe2 and CoTe2-δ. We then measured the magnetic response of resistivity at low temperatures, as shown in Figure 6a. At 2 K, negative MR behavior is observed, which reaches ~−0.06% at ~4 T. The negative MR vanishes with heating, e.g., at 5 K, a positive MR up to 0.12% is observed. With further heating to 10 K, no obvious MR behavior observed. More low-field features can be observed at 2 K and 5 K, e.g., a slight upturn with magnetic fields applied from 0 to 0.5 T at 2 K, and a slightly negative MR observed at 5 K within 1 T, indicating the complicated magnetic interaction nature below 10 K in trigonal CoTe2-δ.
Figure 6b shows the Hall effect at 2 K, 5 K, and 10 K, which are linear lines coinciding with each other, implying that the aforementioned low-temperature magnetic interactions have negligible contributions on the Hall effect. From the Hall curves, one may obtain the Hall coefficient, RH~3.0 × 10−9 m3/C, and (hole) carrier’s density, nh~2 × 1021 cm−3. The carrier’s Hall mobility at 2 K can be calculated via µHall = σ·RH = 23 cm2/Vs, where σ is the conductivity. The complicated MR behavior can be understood based on the band topology study by ARPES and spin-ARPES [67]. The ionic balance of the chemical formula suggests that the d electrons from Co and the p electrons from Te contribute to the conduction carriers on the Fermi surface. The bulk bands show multiple band inversion features throughout the Brillouin zone, which harbor Dirac cone surface states at the Γ point [67]. Linear-dispersed topological surface states will contribute to a positive linear MR at low temperatures, together with quantum interference features, e.g., weak antilocalizations. However, the overall carrier’s concentration from bulk states is much higher than that of the Dirac surface states, which possibly leads to very weak MR responses. On the other hand, the absence of a hysteresis effect in MR suggests the absence of strong ferromagnetic interactions, which is similar to the observation in orthorhombic CoTe2. In Figure 6, no anomalous Hall effect was observed, even at 2 K, which also indicates the absence of long-range magnetic ordering at this temperature. However, the short-range interactions may form at this temperature range, which indicate moderate correlations between Co moments. In this case, the crystal is probably reaching the ordering transition point, and thus is easily affected by the applied magnetic field. This magnetic field-induced ordering will reduce the magnetic disorder-induced scatterings and thus contribute to a negative MR. A significant negative MR is usually observed slightly above the magnetic ordered phase as an indication of short-range ordering, e.g., in Ti-doped Cr2Se3 [73], NdCrSb3 [76]. In CoTe2-δ single crystals, the negative MR behavior is probably due to the short-range magnetic interactions, in which the moments will be aligned along the magnetic field.
To conclude, a trigonal CoTe2 single crystal with Te vacancies was successfully grown using the CVT method. The as-grown layered single crystals have smooth surfaces with hexagonal edges due to the crystal symmetry. Two features are observed on the as-grown single-crystal surface from the spontaneous crystal growth process: (1) piled up triangles with relatively thicker edges appear near to the edge region, and (2) a flat region with ~100 nm zig-zag stripes can be found everywhere on the surface. On a cleaved surface, mosaic-like patterns are observed with the typical dimension of a few tens of nanometers. The CoTe2 single crystal possesses metallic ground states, with a hole carrier concentration at a 1021 cm−3 magnitude. Below 10 K, CoTe2 shows transport features of mixed contributions from surface Dirac fermions and short-range magnetic interactions. The bulk states are much stronger than the Dirac surface states in a sub mm thick bulk crystal, which limits the observation of strong MR and quantum interference effects. Therefore, thin flakes down to few nanometers are desired in a future study, in which the topological surface states will contribute to a higher percentage in the transport behavior. Further, mosaic patterns in TMDs are rarely reported in high-quality single crystals, which will also inspire future studies in the TMD family.

Author Contributions

Investigation and first draft writing, L.C.; Supervision, W.Z. and R.-K.Z. All authors contribute to discussion. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Natural Science Foundation of Guangdong Province (Grant No. 2022A1515010583), the National Natural Science Foundation of China (Grant No. 11974155), and the Bureau of Education of Guangzhou Municipality (Grant No. 202255464).

Data Availability Statement

All data are included in the article, which will be available upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Katsnelson, M.I.; Grigorieva, I.V.; Dubonos, S.; Firsov, A.A. Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438, 197–200. [Google Scholar] [CrossRef]
  2. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.-E.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef]
  3. Bernardi, M.; Palummo, M.; Grossman, J.C. Extraordinary Sunlight Absorption and One Nanometer Thick Photovoltaics Using Two-Dimensional Monolayer Materials. Nano Lett. 2013, 13, 3664–3670. [Google Scholar] [CrossRef]
  4. Kang, K.; Xie, S.E.; Huang, L.J.; Han, Y.M.; Huang, P.Y.; Mak, K.F.; Kim, C.J.; Muller, D.; Park, J. High-mobility three-atom-thick semiconducting films with wafer-scale homogeneity. Nature 2015, 520, 656–660. [Google Scholar] [CrossRef]
  5. Li, H.; Wu, J.M.T.; Yin, Z.Y.; Zhang, H. Preparation and Applications of Mechanically Exfoliated Single-Layer and Multi layer MoS2 and WSe2 Nanosheets. Acc. Chem. Res. 2014, 47, 1067–1075. [Google Scholar] [CrossRef]
  6. Liu, W.; Kang, J.H.; Sarkar, D.; Khatami, Y.; Jena, D.; Banerjee, K. Role of Metal Contacts in Designing High-Performance Monolayer n-Type WSe2 Field Effect Transistors. Nano Lett. 2013, 13, 1983–1990. [Google Scholar] [CrossRef]
  7. Lv, R.; Robinson, J.A.; Schaak, R.E.; Sun, D.; Sun, Y.F.; Mallouk, T.E.; Terrones, M. Transition Metal Dichalcogenides and Beyond: Synthesis, Properties, and Applications of Single- and Few-Layer Nanosheets. Acc. Chem. Res. 2015, 48, 56–64. [Google Scholar] [CrossRef]
  8. Ugeda, M.M.; Bradley, A.J.; Shi, S.F.; da Jornada, F.H.; Zhang, Y.; Qiu, D.Y.; Ruan, W.; Mo, S.K.; Hussain, Z.; Shen, Z.X.; et al. Giant bandgap renormalization and excitonic effects in a monolayer transition metal dichalcogenide semiconductor. Nat. Mater. 2014, 13, 1091–1095. [Google Scholar] [CrossRef]
  9. Wang, H.; Yu, L.L.; Lee, Y.H.; Shi, Y.M.; Hsu, A.; Chin, M.L.; Li, L.J.; Dubey, M.; Kong, J.; Palacios, T. Integrated Circuits Based on Bilayer MoS2 Transistors. Nano Lett. 2012, 12, 4674–4680. [Google Scholar] [CrossRef]
  10. Choi, W.; Ahn, J.; Kim, K.T.; Jin, H.J.; Hong, S.; Hwang, D.K.; Im, S. Ambipolar Channel p-TMD/n-Ga2O3 Junction Field Effect Transistors and High Speed Photo-sensing in TMD Channel. Adv. Mater. 2021, 33, 2103079. [Google Scholar] [CrossRef]
  11. Kallatt, S.; Umesh, G.; Bhat, N.; Majumdar, K. Photoresponse of atomically thin MoS2 layers and their planar heterojunctions. Nanoscale 2016, 8, 15213–15222. [Google Scholar] [CrossRef]
  12. Liu, R.; Wang, F.K.; Liu, L.X.; He, X.Y.; Chen, J.Z.; Li, Y.; Zhai, T.Y. Band Alignment Engineering in Two-Dimensional Transition Metal Dichalcogenide-Based Heterostructures for Photodetectors. Small Struct. 2021, 2, 2000136. [Google Scholar] [CrossRef]
  13. Liu, W.J.; Liu, M.L.; Han, H.N.; Fang, S.B.; Teng, H.; Lei, M.; Wei, Z.Y. Nonlinear optical properties of WSe2 and MoSe2 films and their applications in passive Q-switched erbium doped fiber lasers. Photonics Res. 2018, 6, C15–C21. [Google Scholar] [CrossRef]
  14. Lu, J.P.; Liu, H.W.; Tok, E.S.; Sow, C.H. Interactions between lasers and two-dimensional transition metal dichalcogenides. Chem. Soc. Rev. 2016, 45, 2494–2515. [Google Scholar] [CrossRef]
  15. Si, K.Y.; Ma, J.Y.; Lu, C.H.; Zhou, Y.X.; He, C.; Yang, D.; Wang, X.M.; Xu, X.L. A two-dimensional MoS2/WSe2 van der Waals heterostructure for enhanced photoelectric performance. Appl. Surf. Sci. 2020, 507, 145082. [Google Scholar] [CrossRef]
  16. Zhuang, M.H.; Xu, G.L.; Gan, L.Y.; Dou, Y.B.; Sun, C.J.; Ou, X.W.; Xie, Y.Y.; Liu, Z.J.; Cai, Y.T.; Ding, Y.; et al. Sub-5 nm edge-rich 1T′-ReSe2 as bifunctional materials for hydrogen evolution and sodium-ion storage. Nano Energy 2019, 58, 660–668. [Google Scholar] [CrossRef]
  17. de la Barrera, S.C.; Sinko, M.R.; Gopalan, D.P.; Sivadas, N.; Seyler, K.L.; Watanabe, K.; Taniguchi, T.; Tsen, A.W.; Xu, X.D.; Xiao, D.; et al. Tuning Ising superconductivity with layer and spin-orbit coupling in two-dimensional transition-metal dichalcogenides. Nat. Commun. 2018, 9, 1427. [Google Scholar] [CrossRef]
  18. Shi, W.; Ye, J.T.; Zhang, Y.J.; Suzuki, R.; Yoshida, M.; Miyazaki, J.; Inoue, N.; Saito, Y.; Iwasa, Y. Superconductivity Series in Transition Metal Dichalcogenides by Ionic Gating. Sci. Rep. 2015, 5, 12534. [Google Scholar] [CrossRef]
  19. Sipos, B.; Kusmartseva, A.F.; Akrap, A.; Berger, H.; Forró, L.; Tutis, E. From Mott state to superconductivity in 1T-TaS2. Nat. Mater. 2008, 7, 960–965. [Google Scholar] [CrossRef]
  20. Zhou, B.T.; Yuan, N.F.Q.; Jiang, H.L.; Law, K.T. Ising superconductivity and Majorana fermions in transition-metal dichalcogenides. Phys. Rev. B 2016, 93, 180501(R). [Google Scholar] [CrossRef]
  21. Baranowski, M.; Surrente, A.; Klopotowski, L.; Urban, J.M.; Zhang, N.; Maude, D.K.; Wiwatowski, K.; Mackowski, S.; Kung, Y.C.; Dumcenco, D.; et al. Probing the Interlayer Exciton Physics in a MoS2/MoSe2/MoS2 van der Waals Heterostructure. Nano Lett. 2017, 17, 6360–6365. [Google Scholar] [CrossRef]
  22. Calman, E.V.; Fogler, M.M.; Butov, L.V.; Hu, S.; Mishchenko, A.; Geim, A.K. Indirect excitons in van der Waals heterostructures at room temperature. Nat. Commun. 2018, 9, 1895. [Google Scholar] [CrossRef]
  23. Epstein, I.; Terres, B.; Chaves, A.J.; Pusapati, V.V.; Rhodes, D.A.; Frank, B.; Zimmermann, V.; Qin, Y.; Watanabe, K.; Taniguchi, T.; et al. Near-Unity Light Absorption in a Monolayer WS2 Van der Waals Heterostructure Cavity. Nano Lett. 2020, 20, 3545–3552. [Google Scholar] [CrossRef]
  24. Ghiasi, T.S.; Ingla-Aynés, J.; Kaverzin, A.A.; van Wees, B.J. Large Proximity-Induced Spin Lifetime Anisotropy in Transition-Metal Dichalcogenide/Graphene Heterostructures. Nano Lett. 2017, 17, 7528–7532. [Google Scholar] [CrossRef]
  25. Kim, J.; Jin, C.H.; Chen, B.; Cai, H.; Zhao, T.; Lee, P.; Kahn, S.; Watanabe, K.; Taniguchi, T.; Tongay, S.; et al. Observation of ultralong valley lifetime in WSe2/MoS2 heterostructures. Sci. Adv. 2017, 3, e1700518. [Google Scholar] [CrossRef]
  26. Regan, E.C.; Wang, D.Q.; Paik, E.Y.; Zeng, Y.X.; Zhang, L.; Zhu, J.H.; MacDonald, A.H.; Deng, H.; Wang, F. Emerging exciton physics in transition metal dichalcogenide heterobilayers. Nat. Rev. Mater. 2022, 7, 778–795. [Google Scholar] [CrossRef]
  27. Brotons-Gisbert, M.; Baek, H.; Molina-Sanchez, A.; Campbell, A.; Scerri, E.; White, D.; Watanabe, K.; Taniguchi, T.; Bonato, C.; Gerardot, B.D. Spin-layer locking of interlayer excitons trapped in moire potentials. Nat. Mater. 2020, 19, 630–636. [Google Scholar] [CrossRef]
  28. Choi, J.H.; Hsu, W.T.; Lu, L.S.; Sun, L.Y.; Cheng, H.Y.; Lee, M.H.; Quan, J.M.; Tran, K.; Wang, C.Y.; Staab, M.; et al. Moire potential impedes interlayer exciton diffusion in van der Waals heterostructures. Sci. Adv. 2020, 6, eaba8866. [Google Scholar] [CrossRef]
  29. Enaldiev, V.V.; Zólyomi, V.; Yelgel, C.; Magorrian, S.J.; Fal’ko, V.I. Stacking Domains and Dislocation Networks in Marginally Twisted Bilayers of Transition Metal Dichalcogenides. Phys. Rev. Lett. 2020, 124, 206101. [Google Scholar] [CrossRef]
  30. Hennighausen, Z.; Kar, S. Twistronics: A turning point in 2D quantum materials. Electron. Struct. 2021, 3, 014004. [Google Scholar] [CrossRef]
  31. Li, H.Y.; Li, S.W.; Naik, M.H.; Xie, J.X.; Li, X.Y.; Wang, J.Y.; Regan, E.; Wang, D.Q.; Zhao, W.Y.; Zhao, S.H.; et al. Imaging moire flat bands in three-dimensional reconstructed WSe2/WS2 superlattices. Nat. Mater. 2021, 20, 945–950. [Google Scholar] [CrossRef]
  32. Maity, I.; Naik, M.H.; Maiti, P.K.; Krishnamurthy, H.R.; Jain, M. Phonons in twisted transition-metal dichalcogenide bilayers: Ultrasoft phasons and a transition from a superlubric to a pinned phase. Phys. Rev. Res. 2020, 2, 013335. [Google Scholar] [CrossRef]
  33. Marcellina, E.; Liu, X.; Hu, Z.H.; Fieramosca, A.; Huang, Y.Q.; Du, W.; Liu, S.; Zhao, J.; Watanabe, K.; Taniguchi, T.; et al. Evidence for Moire Trions in Twisted MoSe2 Homobilayers. Nano Lett. 2021, 21, 4461–4468. [Google Scholar] [CrossRef]
  34. Rosenberger, M.R.; Chuang, H.J.; Phillips, M.; Oleshko, V.P.; McCreary, K.M.; Sivaram, S.V.; Hellberg, C.S.; Jonker, B.T. Twist Angle-Dependent Atomic Reconstruction and Moire Patterns in Transition Metal Dichalcogenide Heterostructures. Acs Nano 2020, 14, 4550–4558. [Google Scholar] [CrossRef]
  35. Ruiz-Tijerina, D.A.; Fal’ko, V.I. Interlayer hybridization and moire superlattice minibands for electrons and excitons in heterobilayers of transition-metal dichalcogenides. Phys. Rev. B 2019, 99, 125424. [Google Scholar] [CrossRef]
  36. Sung, J.; Zhou, Y.; Scuri, G.; Zólyomi, V.; Andersen, T.I.; Yoo, H.; Wild, D.S.; Joe, A.Y.; Gelly, R.J.; Heo, H.; et al. Broken mirror symmetry in excitonic response of reconstructed domains in twisted MoSe2/MoSe2bilayers. Nat. Nanotechnol. 2020, 15, 750–754. [Google Scholar] [CrossRef]
  37. Wilson, N.P.; Yao, W.; Shan, J.; Xu, X.D. Excitons and emergent quantum phenomena in stacked 2D semiconductors. Nature 2021, 599, 383–392. [Google Scholar] [CrossRef]
  38. Xie, Y.M.; Zhang, C.P.; Hu, J.X.; Mak, K.F.; Law, K.T. Valley-Polarized Quantum Anomalous Hall State in Moire MoTe2/WSe2 Heterobilayers. Phys. Rev. Lett. 2022, 128, 026402. [Google Scholar] [CrossRef]
  39. 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]
  40. Soluyanov, A.A.; Gresch, D.; Wang, Z.J.; Wu, Q.S.; Troyer, M.; Dai, X.; Bernevig, B.A. Type-II Weyl semimetals. Nature 2015, 527, 495–498. [Google Scholar] [CrossRef]
  41. Yan, M.Z.; Huang, H.Q.; Zhang, K.N.; Wang, E.Y.; Yao, W.; Deng, K.; Wan, G.L.; Zhang, H.Y.; Arita, M.; Yang, H.T.; et al. Lorentz-violating type-II Dirac fermions in transition metal dichalcogenide PtTe2. Nat. Commun. 2017, 8, 257. [Google Scholar] [CrossRef] [PubMed]
  42. Hasan, M.Z.; Kane, C.L. Colloquium: Topological insulators. Rev. Mod. Phys. 2010, 82, 3045–3067. [Google Scholar] [CrossRef]
  43. Qi, X.L.; Zhang, S.C. Topological insulators and superconductors. Rev. Mod. Phys. 2011, 83, 1057. [Google Scholar] [CrossRef]
  44. Zhao, W.; Wang, X. Berry phase in quantum oscillations of topological materials. Adv. Phys. X 2022, 7, 2064230. [Google Scholar] [CrossRef]
  45. Fu, L.; Kane, C.L. Topological insulators with inversion symmetry. Phys. Rev. B 2007, 76, 045302. [Google Scholar] [CrossRef]
  46. Fei, Z.Y.; Palomaki, T.; Wu, S.F.; Zhao, W.J.; Cai, X.H.; Sun, B.S.; Nguyen, P.; Finney, J.; Xu, X.D.; Cobden, D.H. Edge conduction in monolayer WTe2. Nat. Phys. 2017, 13, 677–682. [Google Scholar] [CrossRef]
  47. Tang, S.J.; Zhang, C.F.; Wong, D.; Pedramrazi, Z.; Tsai, H.Z.; Jia, C.J.; Moritz, B.; Claassen, M.; Ryu, H.; Kahn, S.; et al. Quantum spin Hall state in monolayer 1T′-WTe2. Nat. Phys. 2017, 13, 683–687. [Google Scholar] [CrossRef]
  48. Chia, X.Y.; Adriano, A.; Lazar, P.; Sofer, Z.; Luxa, J.; Pumera, M. Layered Platinum Dichalcogenides (PtS2, PtSe2, and PtTe2) Electrocatalysis: Monotonic Dependence on the Chalcogen Size. Adv. Funct. Mater. 2016, 26, 4306–4318. [Google Scholar] [CrossRef]
  49. Chia, X.Y.; Sofer, Z.; Luxa, J.; Pumera, M. Unconventionally Layered CoTe2 and NiTe2 as Electrocatalysts for Hydrogen Evolution. Chem. A Eur. J. 2017, 23, 11719–11726. [Google Scholar] [CrossRef]
  50. Ganesan, V.; Nam, K.H.; Park, C.M. Robust Polyhedral CoTe2-C Nanocomposites as High-Performance Li- and Na-Ion Battery Anodes. ACS Appl. Energy Mater. 2020, 3, 4877–4887. [Google Scholar] [CrossRef]
  51. Gao, Q.; Huang, C.Q.; Ju, Y.M.; Gao, M.R.; Liu, J.W.; An, D.; Cui, C.H.; Zheng, Y.R.; Li, W.X.; Yu, S.H. Phase-Selective Syntheses of Cobalt Telluride Nanofleeces for Efficient Oxygen Evolution Catalysts. Angew. Chem. -Int. Ed. 2017, 56, 7769–7773. [Google Scholar] [CrossRef] [PubMed]
  52. Ji, L.; Wang, Z.C.; Wang, H.; Shi, X.F.; Asiri, A.M.; Sun, X.P. Hierarchical CoTe2 Nanowire Array: An Effective Oxygen Evolution Catalyst in Alkaline Media. Acs Sustain. Chem. Eng. 2018, 6, 4481–4485. [Google Scholar] [CrossRef]
  53. Liu, M.; Lu, X.Q.; Guo, C.; Wang, Z.J.; Li, Y.P.; Lin, Y.; Zhou, Y.; Wang, S.T.; Zhang, J. Architecting a Mesoporous N-Doped Graphitic Carbon Framework Encapsulating CoTe2 as an Efficient Oxygen Evolution Electrocatalyst. ACS Appl. Mater. Interfaces 2017, 9, 36146–36153. [Google Scholar] [CrossRef] [PubMed]
  54. Lu, T.H.; Chen, C.J.; Basu, M.; Ma, C.G.; Liu, R.S. The CoTe2 nanostructure: An efficient and robust catalyst for hydrogen evolution. Chem. Commun. 2015, 51, 17012–17015. [Google Scholar] [CrossRef] [PubMed]
  55. Ma, H.F.; Dang, W.Q.; Yang, X.D.; Li, B.; Zhang, Z.W.; Chen, P.; Liu, Y.; Wan, Z.; Qian, Q.; Luo, J.; et al. Chemical Vapor Deposition Growth of Single Crystalline CoTe2 Nanosheets with Tunable Thickness and Electronic Properties. Chem. Mater. 2018, 30, 8891–8896. [Google Scholar] [CrossRef]
  56. McKendry, I.G.; Thenuwara, A.C.; Sun, J.W.; Peng, H.W.; Perdew, J.P.; Strongin, D.R.; Zdilla, M.J. Water Oxidation Catalyzed by Cobalt Oxide Supported on the Mattagamite Phase of CoTe2. ACS Catal. 2016, 6, 7393–7397. [Google Scholar] [CrossRef]
  57. Muhler, M.; Bensch, W.; Schur, M. Preparation, crystal structures, experimental and theoretical electronic band structures of cobalt tellurides in the composition range CoTe1.3-CoTe2. J. Phys. -Condens. Matter 1998, 10, 2947–2962. [Google Scholar] [CrossRef]
  58. Nath, M.; De Silva, U.; Singh, H.; Perkins, M.; Liyanage, W.P.R.; Umapathi, S.; Chakravarty, S.; Masud, J. Cobalt Telluride: A Highly Efficient Trifunctional Electrocatalyst for Water Splitting and Oxygen Reduction. ACS Appl. Energy Mater. 2021, 4, 8158–8174. [Google Scholar] [CrossRef]
  59. Wang, K.; Ye, Z.G.; Liu, C.Q.; Xi, D.; Zhou, C.J.; Shi, Z.Q.; Xia, H.Y.; Liu, G.W.; Qiao, G.J. Morphology-Controllable Synthesis of Cobalt Telluride Branched Nanostructures on Carbon Fiber Paper as Electrocatalysts for Hydrogen Evolution Reaction. ACS Appl. Mater. Interfaces 2016, 8, 2910–2916. [Google Scholar] [CrossRef]
  60. Wang, X.; Huang, X.K.; Gao, W.B.; Tang, Y.; Jiang, P.B.; Lan, K.; Yang, R.Z.; Wang, B.; Li, R. Metal-organic framework derived CoTe2 encapsulated in nitrogen-doped carbon nanotube frameworks: A high-efficiency bifunctional electrocatalyst for overall water splitting. J. Mater. Chem. A 2018, 6, 3684–3691. [Google Scholar] [CrossRef]
  61. Wang, X.G.; Zhou, Z.; Zhang, P.; Zhang, S.Q.; Ma, Y.; Yang, W.W.; Wang, H.; Li, B.X.; Meng, L.J.; Jiang, H.N.; et al. Thickness-Controlled Synthesis of CoX2 (X = S, Se, and Te) Single Crystalline 2D Layers with Linear Magnetoresistance and High Conductivity. Chem. Mater. 2020, 32, 2321–2329. [Google Scholar] [CrossRef]
  62. Wang, Z.Q.; Ren, X.; Wang, L.; Cui, G.W.; Wang, H.J.; Sun, X.P. A hierarchical CoTe2-MnTe2 hybrid nanowire array enables high activity for oxygen evolution reactions. Chem. Commun. 2018, 54, 10993–10996. [Google Scholar] [CrossRef] [PubMed]
  63. Yin, Y.Q.; Xu, J.; Guo, W.; Wang, Z.; Du, X.D.; Chen, C.; Zheng, Z.J.; Liu, D.; Qu, D.Y.; Xie, Z.Z.; et al. A single-step fabrication of CoTe2 nanofilm electrode toward efficient overall water splitting. Electrochim. Acta 2019, 307, 451–458. [Google Scholar] [CrossRef]
  64. Zhang, B.; Zhang, Y.; Li, J.L.; Liu, J.; Huo, X.G.; Kang, F.Y. In situ growth of metal-organic framework-derived CoTe2 nanoparticles@nitrogen-doped porous carbon polyhedral composites as novel cathodes for rechargeable aluminum-ion batteries. J. Mater. Chem. A 2020, 8, 5535–5545. [Google Scholar] [CrossRef]
  65. Siegfried, P.E.; Bhandari, H.; Qi, J.N.; Ghimire, R.; Joshi, J.; Messegee, Z.T.; Beeson, W.B.; Liu, K.; Ghimire, M.P.; Dang, Y.L.; et al. CoTe2: A Quantum Critical Dirac Metal with Strong Spin Fluctuations. Adv. Mater. 2023, 35, 2300640. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, B.; Wang, L.; Ding, D.; Zhai, Y.; Wang, F.; Jing, Z.; Yang, X.; Kong, Y.; Qian, Y.; Xu, L. Zinc-Assisted Cobalt Ditelluride Polyhedra Inducing Lattice Strain to Endow Efficient Adsorption-Catalysis for High-Energy Lithium–Sulfur Batteries. Adv. Mater. 2022, 34, 2204403. [Google Scholar] [CrossRef] [PubMed]
  67. Chakraborty, A.; Fujii, J.; Kuo, C.N.; Lue, C.S.; Politano, A.; Vobornik, I.; Agarwal, A. Observation of highly anisotropic bulk dispersion and spin-polarized topological surface states in CoTe2. Phys. Rev. B 2023, 107, 085406. [Google Scholar] [CrossRef]
  68. Hu, Z.; Zhang, L.B.; Chakraborty, A.; D’Olimpio, G.; Fujii, J.; Ge, A.P.; Zhou, Y.C.; Liu, C.L.; Agarwal, A.; Vobornik, I.; et al. Terahertz Nonlinear Hall Rectifiers Based on Spin-Polarized Topological Electronic States in 1T-CoTe2. Adv. Mater. 2023, 35, 2209557. [Google Scholar] [CrossRef]
  69. Guo, L.; Xu, M.; Chen, L.; Huang, X.; Shi, X.Y.; Ying, J.S.; Zhang, T.; Zhao, W.Y.; Dong, S.; Zheng, R.K. Electronic Transport Properties of Nb1-xTaxSb2 Single-Crystal Semimetals Grown by a Chemical Vapor Transport Based High-Throughput Method. Cryst. Growth Des. 2021, 21, 653–662. [Google Scholar] [CrossRef]
  70. Guo, L.; Zhao, W.Y.; Ding, N.; Shi, X.Y.; Xu, M.; Chen, L.; Gao, G.Y.; Dong, S.; Zheng, R.K. Magnetotransport properties of square-net compounds of NbSiSb and NbGeSb single crystals. J. Phys. Condens. Matter 2020, 32, 435701. [Google Scholar] [CrossRef]
  71. Wu, J.; Zhang, C.L.; Yan, J.M.; Chen, L.; Guo, L.; Chen, T.W.; Gao, G.Y.; Fei, L.F.; Zhao, W.Y.; Chai, Y.; et al. Magnetotransport and magnetic properties of the layered noncollinear antiferromagnetic Cr2Se3 single crystals. J. Phys. Condens. Matter 2020, 32, 475801. [Google Scholar] [CrossRef] [PubMed]
  72. Xu, M.; Guo, L.; Chen, L.; Zhang, Y.; Li, S.S.; Zhao, W.Y.; Wang, X.L.; Dong, S.; Zheng, R.K. Emerging weak antilocalization effect in Ta0.7Nb0.3Sb2 semimetal single crystals. Front. Phys. 2023, 18, 13304. [Google Scholar] [CrossRef]
  73. Zhang, S.J.; Yan, J.M.; Tang, F.; Wu, J.; Dong, W.Q.; Zhang, D.W.; Luo, F.S.; Chen, L.; Fang, Y.; Zhang, T.; et al. Colossal Magnetoresistance in Ti Lightly Doped Cr2Se3 Single Crystals with a Layered Structure. ACS Appl. Mater. Interfaces 2021, 13, 58949–58955. [Google Scholar] [CrossRef] [PubMed]
  74. Li, Z.; Chen, H.; Jin, S.; Gan, D.; Wang, W.; Guo, L.; Chen, X. Weyl semimetal TaAs: Crystal growth, morphology, and thermodynamics. Cryst. Growth Des. 2016, 16, 1172–1175. [Google Scholar] [CrossRef]
  75. Liu, Q.; Fei, F.; Chen, B.; Bo, X.; Wei, B.; Zhang, S.; Zhang, M.; Xie, F.; Naveed, M.; Wan, X. Nontopological origin of the planar Hall effect in the type-II Dirac semimetal NiTe2. Phys. Rev. B 2019, 99, 155119. [Google Scholar] [CrossRef]
  76. Chen, L.; Zhao, W.; Wang, Z.; Tang, F.; Fang, Y.; Zeng, Z.; Xia, Z.; Cheng, Z.; Cortie, D.L.; Rule, K.C. Spin Reorientation Transition and Negative Magnetoresistance in Ferromagnetic NdCrSb3 Single Crystals. Materials 2023, 16, 1736. [Google Scholar] [CrossRef]
Figure 1. A sketch of the CVT crystal growth process.
Figure 1. A sketch of the CVT crystal growth process.
Crystals 14 00046 g001
Figure 2. The structural characterizations of CoTe2-δ samples. (a) Optical image of as-grown crystals. (b) The low-magnification secondary electron image. (c) The single-crystal and (d) powder XRD patterns of CoTe2-δ crystals. (e) The PDF card #89-4061.
Figure 2. The structural characterizations of CoTe2-δ samples. (a) Optical image of as-grown crystals. (b) The low-magnification secondary electron image. (c) The single-crystal and (d) powder XRD patterns of CoTe2-δ crystals. (e) The PDF card #89-4061.
Crystals 14 00046 g002
Figure 3. SEM and EDS measurements on a freshly cleaved CoTe2 surface. (a) Secondary electron image. (b) Back-scattering image. (c) The EDS spectrum over the whole imaging area. (d) The element maps and second electron image are plotted in superposition. The element maps are shown in Panels (e) (Co) and (f) (Te) at the same position.
Figure 3. SEM and EDS measurements on a freshly cleaved CoTe2 surface. (a) Secondary electron image. (b) Back-scattering image. (c) The EDS spectrum over the whole imaging area. (d) The element maps and second electron image are plotted in superposition. The element maps are shown in Panels (e) (Co) and (f) (Te) at the same position.
Crystals 14 00046 g003
Figure 4. The low-voltage SE images of CoTe2 single crystals. (a) Low-magnification SE image in a large area. The high magnification SE images of as-grown surface and cleaved surface are shown in Panels (bd) and (eh).
Figure 4. The low-voltage SE images of CoTe2 single crystals. (a) Low-magnification SE image in a large area. The high magnification SE images of as-grown surface and cleaved surface are shown in Panels (bd) and (eh).
Crystals 14 00046 g004
Figure 5. Temperature-dependent longitudinal resistivity behavior in 2–300 K region in a CoTe2 single crystal.
Figure 5. Temperature-dependent longitudinal resistivity behavior in 2–300 K region in a CoTe2 single crystal.
Crystals 14 00046 g005
Figure 6. Magnetotransport behavior in CoTe2 single crystal. (a) MR and (b) Hall effect are measured at 2 K, 5 K, and 10 K.
Figure 6. Magnetotransport behavior in CoTe2 single crystal. (a) MR and (b) Hall effect are measured at 2 K, 5 K, and 10 K.
Crystals 14 00046 g006
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

Chen, L.; Zhao, W.; Zheng, R.-K. Single Crystal Growth and Nano-Structure Study in a Topological Dirac Metal, CoTe2-δ. Crystals 2024, 14, 46. https://doi.org/10.3390/cryst14010046

AMA Style

Chen L, Zhao W, Zheng R-K. Single Crystal Growth and Nano-Structure Study in a Topological Dirac Metal, CoTe2-δ. Crystals. 2024; 14(1):46. https://doi.org/10.3390/cryst14010046

Chicago/Turabian Style

Chen, Lei, Weiyao Zhao, and Ren-Kui Zheng. 2024. "Single Crystal Growth and Nano-Structure Study in a Topological Dirac Metal, CoTe2-δ" Crystals 14, no. 1: 46. https://doi.org/10.3390/cryst14010046

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