Ihor Z.
Hlova‡
*a,
Prashant
Singh‡
a,
Serhiy Z.
Malynych
b,
Roman V.
Gamernyk
c,
Oleksandr
Dolotko
a,
Vitalij K.
Pecharsky
ad,
Duane D.
Johnson
ad,
Raymundo
Arroyave
e,
Arjun K.
Pathak
f and
Viktor P.
Balema
*a
aAmes Laboratory, U.S. Department of Energy, Iowa State University, Ames, IA 50011-2416, USA. E-mail: vbalema@ameslab.gov; ihlova@ameslab.gov
bHetman Petro Sahaidachnyi National Army Academy, Lviv, 79026, Ukraine
cIvan Franko National University of Lviv, Lviv, 79005, Ukraine
dDepartment of Materials Science and Engineering, Iowa State University, Ames, IA 50011-1096, USA
eDepartment of Materials Science & Engineering, Texas A&M University, College Station, TX 77843, USA
fDepartment of Physics, SUNY Buffalo State, Buffalo, NY 14222, USA
First published on 7th June 2021
A new family of heterostructured transition-metal dichalcogenides (TMDCs) with incommensurate (“misfit”) spatial arrangements of well-defined layers was prepared from structurally dissimilar single-phase 2H-MoS2 and 1T-HfS2 materials. The experimentally observed heterostructuring is energetically favorable over the formation of homogeneous multi-principle element dichalcogenides observed in related dichalcogenide systems of Mo, W, and Ta. The resulting three-dimensional (3D) heterostructures show semiconducting behavior with an indirect band gap around 1 eV, agreeing with values predicted from density functional theory. Results of this joint experimental and theoretical study open new avenues for generating unexplored metal-dichalcogenide heteroassemblies with incommensurate structures and tunable physical properties.
Bulk TMDCs are built from layers of covalently bonded metal and chalcogen atoms held together via weak van der Waals (vdW) forces (see ESI, Fig. S1†). As a result, they can be easily exfoliated into 2D-nanosheets, even down to single layers.8 Both bulk and 2D-TMDCs demonstrate a broad range of electronic transport properties that span from indirect and direct gap semiconductivity, semimetallic and metallic behavior, to low-temperature superconductivity, depending on the chemical composition and the spatial configurations of the material as well as external stimuli applied.1,9
While binary TMDCs are well known,10 the preparation of layered multi-principal element metal chalcogenides, where different metals (M) and chalcogens (X) share a common crystal lattice, remained challenging until our recent report.11 One of the intriguing outcomes of this earlier study was the observed heterostructuring of different group 5 and 6 binary TMDCs upon mechanical milling at room temperature. Thus-produced 3D-heterostructures are metastable and, when subjected to high-temperature annealing, they transform into uniform single-phase materials. However, if the starting metal chalcogenides possess different crystal structures and stoichiometries, e.g. TaS2 and SmS, or NbSe2 and LaSe, under similar processing conditions they form well-defined and thermodynamically stable heterostructures with incommensurate (misfit) spatial arrangements, where slabs of the mono-chalcogenide (SmS or LaSe) with cubic crystal structures alternate with hexagonal 2D layers of the TMDC (TaS2 or NbSe2).12
The latter discovery raised questions about possibility to design layered misfit materials from chemically related yet structurally dissimilar building blocks, such as hexagonal 2H-MoS213 and trigonal 1T-HfS2,14 by their simultaneous mechanochemical exfoliation and re-assembly into TMDCs heterostructures. If feasible, this would open a new avenue to an unexplored family of incommensurate 3D-heterostructures with tunable physical properties. Below we report on the successful implementation of this idea.
For DFT calculations of (Mo0.5W0.5)S2 and (Mo0.5W0.5)(S0.5Se0.5)2, we used (I) 24-atom supercell (2 × 2 × 1 unit cells of 2H-MoS2) to mimic disorder, whereas 2H-MoS2 was used to model 1 × 1 × 4 heterostructure supercell. For disorder cases: we use 5 × 5 × 1 and 11 × 11 × 5 k-mesh for structural relaxation and energy. For heterostructures: we use 3 × 3 × 1 and 11 × 11 × 1 k-mesh for structural relaxation and energy. For (Mo40W40Ta20)S2, (II) a 96-atom supercell (4 × 4 × 1 unit cells of 2H-MoS2) was used to mimic disorder, whereas 1 × 1 × 10 supercell of 2H-MoS2 was used to model an ordered 4(TaS2):8(MoS2):8(WS2) heterostructure. For disorder case: we use 5 × 5 × 3 and 8 × 8 × 6 k-mesh for structural relaxation and energy. For heterostructures: we use 3 × 3 × 1 and 9 × 9 × 1 k-mesh for structural relaxation and energy. A manual stacking approach was used to create vertical vdW heterostructures of 2H-MoS2 and 1T-HfS2 in the present study.
MoS2:HfS2 molar ratio | MoS2 (2H) lattice parametersb (Å) | HfS2 (1T) lattice parametersc (Å) | R p, % |
---|---|---|---|
a The Rp values correspond to the profile residuals. b Space group symmetry P63/mmc (#194). c Space group symmetry Pm1 (#164). d Processed twice using the same synthesis protocol as the previous material. | |||
1:0 | a = 3.160(1) Å | — | 6.07 |
c = 12.298(2) Å | |||
0.75:0.25 | a = 3.157(1) Å | a = 3.626(1) Å | 7.83 |
c = 12.286(1) Å | c = 5.852(1) Å | ||
0.5:0.5 | a = 3.162(1) Å | a = 3.634(1) Å | 7.68 |
c = 12.124(1) Å | c = 5.880(1) Å | ||
0.5:0.5d | a = 3.159(1) Å | a = 3.629(1) Å | 6.01 |
c = 12.074(1) Å | c = 5.881(1) Å | ||
0.25:0.75 | a = 3.163(1) Å | a = 3.634(1) Å | 6.52 |
c = 12.255(1) Å | c = 5.951(1) Å | ||
0:1 | — | a = 3.628(1) Å | 9.88 |
c = 5.854(1) Å |
Lattice parameters of both TMDC phases in the annealed sample reveal a slight reduction (∼1.4%) of the parameter c in the MoS2 phase and a smaller but noticeable expansion (∼0.5%) of the parameter c in the HfS2 constituent. Considering that rMo = 1.400 Å < rHf = 1.580 Å (metallic radii for coordination number 12), these anomalies cannot be attributed to minor substitutions on the metal sites. The changes that occur in opposite direction are not likely to be related to various concentrations of defects in two different structural motifs and, therefore, we ascribe them to large errors that arise from a number of factors. These are: strong diffuse scattering due to a few-nanometer thick slabs stacked along the c direction (see Fig. 1b–e); heavy overlap of the majority of the strongest Bragg peaks; and highly anisotropic broadening due to epitaxial strains that are unavoidable and cannot be removed by annealing. All of them combined, Le-Bail-refined unit-cell dimensions, and in particular those parallel to the stacking direction, become susceptible to random errors, even though the formal least squares standard deviations listed in Table 1 are low.
A minor impurity detected in the annealed sample almost certainly belongs to an off-stoichiometric hafnium oxysulfide that has formed in the sample even though all operations were carried out under high-purity argon or helium. The annealed material was further studied using High-Angle Annular Dark Field Scanning Transmission Electron Microscopy (HAADF-STEM) and Energy Dispersive Spectroscopy (STEM-EDS), which reveal a well-defined sandwich-like arrangement of the separate phases in the material as shown in Fig. 1b. The alternating phases have different Z-contrast, whereby the MoS2 slabs appear dark and HfS2 produces much brighter segments. The thickness of specific slabs varies, indicating stochastic nature of the mechanical exfoliation and self-assembly processes. Furthermore, the HAADF-STEM images reveal the presence of the Moiré pattern on the surface of the material (Fig. 1c), which is characteristic for TMDCs with lattice mismatch layers positioned on the top of each other,25i.e. incommensurate structural arrangements. STEM-EDS (Fig. 1d, e) confirms the 3D-heterostructured arrangement of the layers in the sample. The schematic diagram illustrating the formation of 3D-heterostructured TMDCs is shown in Fig. 1f.
Reprocessing of the annealed material by its milling for additional 30 hours, followed by annealing at 1000 °C for 72 hours, does not eliminate the phase separation in the sample. Its XRD, HAADF-STEM and STEM-EDS analyses clearly indicate that the sandwich-like arrangement of the MoS2 and HfS2 slabs is retained in the reprocessed material (Fig. 2, Table 1), although the slabs become markedly thinner.
Two other MoS2-HfS2 compositions that are rich in one or another component were prepared and investigated as well.
The XRD patterns and the structural parameters of the obtained samples are shown in Fig. 3 and Table 1. Also, in these cases, a distinctive formation of solid solutions could not be detected. The increased fraction of the HfOS impurity seen in the XRD pattern of Hf0.75Mo0.25S2 correlates with increased concentration of Hf in the material and indicates higher sensitivity of the Hf-rich material to oxygen. As discussed above, minor non-systematic changes in the c lattice parameters are related to contributions from diffuse scattering and varying anisotropic peak broadening.
The formation energies (Eform) calculated for equimolar. (5:5) HfS2/MoS2 heterostructures and the solid-solution (SS) compound are shown in Fig. 4c. Their comparison indicates that the 5(1T-HfS2)/5(2H-MoS2) arrangement is the most stable among evaluated structures, as its Eform is as low as −1.0902 eV per atom [which is 59.92, 73.45, and 84.32 meV below the Eform calculated, respectively, for 5(1T-HfS2)/5(1T-MoS2), 5(2H-HfS2)/5(2H-MoS2), and SS-(Mo0.5Hf0.5)S2].
In addition, two other families of x5HfS2/y5MoS2 materials were investigated. Their unit cells were built from five-layer HfS2 and MoS2 slabs taken in 1:3 (x = 1, y = 3) and 3:1 (x = 3, y = 1) stoichiometric proportions or SS, (MoyHfx)S2, layers. The calculated Eform values are plotted in Fig. 4c. Here, we also find that the x5(1T-HfS2)/y5(2H-MoS2) heterostructural arrangements are the most energy favorable among other evaluated cases. We also investigated several other possibilities, for example, solubility of Hf in 2H-MoS2 or Mo in 1T-HfS2 or anti-site defects (Mo and Hf at the interface were interchanged to see the effects on energetics), vacancies (Mo or Hf or S). However, none of these possibilities are thermodynamically stable compared to pure 2H-MoS2-1T-HfS2 interface.
The calculated partial density of states (DOS) and the charge-density for 5(1T-HfS2)/5(2H-MoS2) heterostructures, and the Mo-d, Hf-d, and S-sp bands are shown in Fig. 4d–f. The bands near the Fermi energy (EF in Fig. 4d-e) mainly consist of the S-p states that are hybridized with the Mo-d and Hf-d states, whereas the S-s orbitals emerge way below EF, and are separated from the other valence states by 8.0 eV, i.e., are chemically inactive. The strong intralayer hybridization between the d-orbitals of Mo and Hf, and the p-orbitals of S is also evident from the overlapping charge densities shown in Fig. 4f, which stabilizes the 5(1T-HfS2)/5(2H-MoS2) heterostructure. The charge density in both 1T-HfS2 and 2H-MoS2 layers is localized on the S atoms with the directional intralayer bonding toward Mo and Hf and the band gap of 1.01 eV.
The calculated band gaps for (Mo0.5W0.5)S2 and (Mo0.5W0.5)SSe are 1.02 eV and 1.17 eV, respectively. They are indirect in nature and follow the Γ–R high-symmetry direction. Surprisingly, the effect of Se-substitution is quite moderate in this case.
Finally, we also evaluated the formation energies and electronic structures of the single-phase material with the nominal composition of (Mo0.4W0.4Ta0.2)S2. Partial replacement of Mo and W by Ta in (Mo0.5W0.5)S2 can be seen as an injection of 0.2 holes per molecule into the system. Once again, the calculated Eform of the “hole-doped” (Mo0.4W0.4Ta0.2)S2 is substantially lower than that of the heterostructure material (−0.875 vs. −0.835 eV per atom), which explains the experimentally observed formation of the solid-solution phase in this case.11
The band structure and partial DOS for (Mo0.4W0.4Ta0.2)S2 are shown in Fig. 6. The presence of Ta is responsible for only a partial filling of bonding Mo/W d-states, while they are completely filled in (Mo0.5W0.5)S2 (Fig. 5a and b). This moves the bonding t2g and eg states (Fig. 6b and c) to EF, where the majority of states belong to Mo/W/Ta t2g-bands. Here, t2g is a combination of dxy, dyz, and dxz orbitals and eg represents dx2y2 and dz2 orbitals. The S p-bands near EF are situated in the same energy range as the Mo/W d-bands in (Mo0.5W0.5)S2 (Fig. 5b), and Ta causes reduced filling of S p-bands that also move closer to EF. Thus, the crossover of partially-filled Mo/W-d and S-p bands at EF caused by Ta predicts metallic behaviour of (Mo0.4W0.4Ta0.2)S2.
In summary, our DFT results reveal that combining structurally different 1T-HfS2 and 2H-MoS2 phases stabilizes the heterostructured arrangement over solid-solution-like single-phase (1T or 2H) states. At the same time, blending isostructural TMDCs, such as 2H-MoX2 and 2H-WX2 (X = S, Se), produces single-phase materials; even so, the metastable heterostructured intermediates are observed after the low-temperature stages of the previous experiments.11 Doping group 6 TMDCs with a group 5 metal (Ta) is not expected to affect the phase stability of the resulting compounds, but it changes transport behaviour from semiconducting to metallic.
The additional experimental details on measurements performed can be found in the ESI.† Several observations are worth noting. First, in the majority of the cases, the experimental values are in a good agreement with those predicted by DFT or published in the literature. The band gap values experimentally determined for both the HfS2/MoS2 and the multi-principal elements TMDCs shown in Table 2 are below those observed in the pure binary precursors, and obviously can be fine-tuned by altering the material's chemical and phase compositions.
The electronic transport behaviour of multi-principal element TMDCs can be further manipulated by doping with group 5 transition metals, such as Ta, that converts them from semiconductors into metallic-type conductors.30 To illustrate this, we measured temperature dependence of the electrical resistivity, ρ(T), of W0.4Mo0.4Ta0.2S2 using a Physical Property Measurements System (PPMS, Quantum Design, Inc.) employing a standard four-probe technique in magnetic field up to 120 kOe. The ρ vs. (T) measured during cooling and heating between 320 K and 1.8 K in the absence of magnetic field (H = 0) is shown in Fig. 7. Both the heating and cooling curves are practically identical. Consistent with the theoretical prediction, W0.4Mo0.4Ta0.2S2 demonstrates a weakly temperature-dependent metallic conductivity between 300 and 14 K. A minor increase in the resistivity observed at T ≤ 14 K can be attributed to the presence of electron transport barriers between crystallites in the sample that reduce its overall conductivity at cryogenic temperatures. The electrical resistivity measured as a function of the magnetic field up to 120 kOe (Fig. 7, inset) indicates very weak but positive magnetoresistance of ∼1.2% at T = 1.8 K without any sign of saturation.
Footnotes |
† Electronic supplementary information (ESI) available: Additional experimental details on precursor synthesis, Rietveld refinement, photoconductivity and magnetoresistance measurements, and density functional theory calculations. See DOI: 10.1039/d1na00064k |
‡ These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2021 |