Brought to you by:
Paper

Two-dimensional tantalum disulfide: controlling structure and properties via synthesis

, , , , , , , , and

Published 2 January 2018 © 2018 IOP Publishing Ltd
, , Citation Rui Zhao et al 2018 2D Mater. 5 025001 DOI 10.1088/2053-1583/aaa104

2053-1583/5/2/025001

Abstract

Tantalum disulfide (TaS2) is a transition metal dichalcogenide (TMD) that exhibits phase transition induced electronic property modulation at low temperature. However, the appropriate phase must be grown to enable the semiconductor/metal transition that is of interest for next generation electronic applications. In this work, we demonstrate direct and controllable synthesis of ultra-thin 1T-TaS2 and 2H-TaS2 on a variety of substrates (sapphire, SiO2/Si, and graphene) via powder vapor deposition. The synthesis process leads to single crystal domains ranging from 20 to 200 nm thick and 1–10 µm on a side. The TaS2 phase (1T or 2H) is controlled by synthesis temperature, which subsequently is shown to control the electronic properties. Furthermore, this work constitutes the first demonstration of a metal–insulator phase transition in directly synthesized 1T-TaS2 films and domains by electronic means.

Export citation and abstract BibTeX RIS

Introduction

The partial filling of 'd' orbitals in '1T' layered compounds enables fascinating structural and electronic behaviors [1]. For instance, the Peierls transition associated with charge density wave (CDW) instability is a unique feature of these 1T materials. CDW prompts new orbital ordering, which leads to unusual behavior like superconductivity or metal–insulator transition (MIT). Unlike the more heavily studied transition metal dichalcogenides (TMDs), 1T-TaS2 is metallic and exhibits an intriguing relationship of electron-lattice interactions [2]. The undistorted 1T-TaS2 is trigonal ($p\overline{3}m1$ ) and exhibits a metallic characteristic (figure 1(a) (band structure, (BS)) and figure 1(b) (density of states (DOS))) even when thinned to its monolayer limit [3]. However, below a critical temperature, the Fermi-surface instability spontaneously breaks the lattice symmetry of the crystal and the material forms the so-called 'Star-of-David' commensurate CDW (CCDW) phase with an in-plane periodic distortion of $\sqrt{13}\times \sqrt{13}$ R  =  13.5°. During this transition, bulk 1T-TaS2 exhibits interesting electronic properties; it behaves as a band insulator with an in-plane gap (Γ–MK–Γ direction; see figure 1(c)) of nearly 200 meV but it shows metallic characteristics in the out-of-plane direction (Γ–A direction) (figures 1(c) (BS) and (d) (DOS)). This strongly implies that the interlayer coupling, along with the CCDW formation, plays a major role in opening the in-plane bandgap in CDW multilayer TaS2 irrespective of the Coulombic repulsion in the crystal [1]. Similar to 1T-TaS2, undistorted hexagonal 2H-TaS2 (P63/mmc) also exhibits a metallic behavior (figures 1(e) and (f)). In the case of 2H, as the temperature is reduced, it undergoes a 3  ×  3 CDW superlattice transformation [2]. However, the energy lowering of the CDW distortion is quite small (approximately 2 meV per formula unit -f.u.) along with small atomic displacements (0.05 Å), indicating a similar electronic behavior to the undistorted one [4]. The calculation results for distorted 2H-TaS2 can be found in the supplemental material (SI) (stacks.iop.org/TDM/5/025001/mmedia).

Figure 1.

Figure 1. DFT calculation on both 1T-TaS2 and 2H-TaS2 bulk crystals. (a) BS and (b) density of states (DOS) of undistorted 1T-TaS2. (c) BS and (d) DOS of 1T-TaS2 within CCDW phase, when a $\sqrt{13}\times \sqrt{13}$ superstructure is formed. (e) BS and (f) DOS of undistorted 2H-TaS2. The energy lowering of $3\times 3$ superstructure for 2H-TaS2 phase is quite small below its phase transition temperature based on calculation. Its BS and DOS are put in supplemental materials.

Standard image High-resolution image

Recent progress in TaS2 relies on chemical vapor transport (CVT), followed by mechanical exfoliation onto different substrates [59]. With this route, stoichiometric Ta and S is sealed in a quartz tube and heated at elevated temperatures (>850 °C) for 10–14 d to create the bulk crystals [10, 11]. Besides the long growth time, the contamination during exfoliation and the transfer processes can degrade the material properties. Low yield and unpredictable flake size are also common problems. Thus, integrating TaS2 into large-scale complex structures is highly desirable, yet challenging to realize. In this paper, we demonstrate the powder vapor (PV) deposition (figure 2(a)) method for direct and controllable synthesis of both 1T-TaS2 and 2H-TaS2 films.

Figure 2.

Figure 2. Experiment setup and growth results. (a) PV deposition set-up. (b) Growth summary based on different tantalum precursors. (c) A typical AFM image demonstrating the 1T-TaS2 nanotextures through ALD  +  sulfurization route. (d) A typical AFM image showing 1T-TaS2 flakes by PVD route. (e) 1T-TaS2/2H-TaS2 phase ratios from each PVD growth as a function of temperature.

Standard image High-resolution image

Results and discussion

Powder vapor deposition (PV) (figure 2(a)) enables the controlled growth of 2H and 1T-TaS2. This is evident in figure 2(b), where multiple Ta precursors are evaluated with sulfur (S) vapors to realize TaS2. Tantalum films produced by thermal evaporation are 100% transformed into polycrystalline 2H-TaS2 with nano-sized domains. On the other hand, transformations from tantalum oxide (Ta2O5), tantalum carbide (TaC) or tantalum nitride (TaN) powders lead to very limited sulfide formation due to the high thermal and chemical stability of these compounds. In order to increase chemical reactivity, a TaOx film (with thickness smaller than 1 nm) is first deposited via atomic layer deposition (ALD). Sulfurizing ALD TaOx, using an Ar/H2 carrier gas, leads to successful transformation of the layer to 1T-TaS2 with triangular or hexagonal domains (figure 2(c)), but with lateral sizes smaller than 200 nm. The small lateral size is limited due to the starting ALD oxide that is amorphous or nanocrystalline, as well as the very low surface diffusion rates of Ta. In order to increase domain size, vaporization of the precursors is utilized rather than simple deposition and sulfurization of Ta-based films. The term 'PV' is utilized for this technique due to the inherent lack of control compared to true chemical vapor deposition (CVD). Directly sulfurizing Ta powder does not yield any growth due to the high chemical reaction barrier between Ta and S. Therefore, we introduce the use of tellurium (Te) to catalyze the vaporization and mobility of Ta in the PV process. By mixing Te with Ta as a solid precursor, the growth of TaS2 is greatly promoted as reported for several other TMDs [12, 13]. This is because the Te forms a binary eutectic with transition metals and thus lower the eutectic melting point significantly [12]. Additionally, Te also forms binary compounds with S that can more strongly reduce the transition metal. As a result, the reactivity between the chemical agents is largely increased [14]. Therefore, significantly larger TaS2 domains were achieved on the substrates with the help of Te catalyst compared to domains without the Te addition.

Using Te-assisted PV, consistent growths are achieved on multiple substrates (sapphire, SiO2/Si, epitaxy graphene, etc). A typical flake shape is shown in figure 2(d). The average flake size is 5–15 µm. Due to different crystal formation energy between 1T- and 2H-TaS2 [15], temperature is the key to control the reactant phases during PV. Figure 2(e) summarizes the averaged phase ratios from each growth as a function of temperature. Since 1T-TaS2 has a higher formation energy compared with 2H-TaS2 [16], higher temperatures favor 1T-TaS2 growth [17]. The transition from 2H to 1T-TaS2 growth occurs gradually from 850 °C to 975 °C, with a minimum of 975 °C to ensure a fully formed 1T phase. To stabilize the 1T phase following high temperature synthesis, the samples are cooled at a high rate (~60 °C/min) to  <300 °C; otherwise the phase transforms to the 2H phase during the cooling cycle.

Growth of TaS2 (both 1T and 2H) via Te-assisted PV enables the formation of crystalline domains with different chemical signatures. This is evidenced by high resolution scanning transmission electron microscopy (HRTEM) (figures 3(a) and (b)), where an intensity line is drawn along the Ta–Ta direction to highlight the Ta-S signal contrast. As shown, there is a stronger S signal from the 2H-TaS2 image than from the 1T-TaS2 image. This is due to difference in their crystal structure. In 1T-TaS2, S is octahedrally coordinated with Ta while in 2H-TaS2, S is trigonal pristimatically coordinated to Ta. As a result, S from both layers adds to the total intensity in the 2H-TaS2 image, while in the 1T-TaS2 the S signal is only collected from one layer resulting in a weaker contrast. X-ray photoemission spectroscopy (XPS) (figures 3(c) and (d)) of the Ta 4f peaks from 1T-TaS2 and 2H-TaS2 shows the peak is broader in 1T-TaS2 (figure 3(c)) than 2H-TaS2 (figure 3(d)). The peak broadening can be attributed to differeneces in their CDW properties. Although both are CDW materials, they have different CDW structures and transition temperatures. For 1T-TaS2, at room temperature, its structure is already partially distorted and forms a 'David Star' (this is known as the nearly commensurate CDW (NCCDW) phase) [9]. As a result, Ta atoms do not preserve a 'pure' chemical envionment [18, 19]. On the contrary, 2H-TaS2 remains in its pristine crystal structure until 75 K [20]. Below 75 K, it directly transitions into the commenrate CDW (CCDW) structure and forms a 3  ×  3 superstructure. As a result, at room temperature, all the Ta atoms share the same chemical bonding environment and result in sharp Ta 4f peaks. High resolution XPS survey scans (figure 3(e)) also confirm that no Te exists within the XPS detection limit.

Figure 3.

Figure 3. HRTEM and XPS data for 1T-TaS2 and 2H-TaS2. (a) HRTEM for 1T-TaS2. Inserted is the composition intensity along Ta–Ta direction. (b) HRTEM for 2H-TaS2. Inserted is the composition intensity along Ta–Ta direction. (c) XPS high resolution scan for Ta 4f peaks from 1T-TaS2 flake. (d) XPS high resolution data for Ta 4f peaks from 2H-TaS2 flake. (e) XPS high resolution survey for Te 4d peak detection. (f) Cross-section TEM demonstrating that the top layers (10%) of 1T-TaS2 has been oxidized when it is left at atmosphere environment for 20 h without protection. The scale bar is 2 nm.

Standard image High-resolution image

TaS2 is highly sensitive to the ambient environment [21]. Both XPS and cross-section TEM reveal heavy oxidation of as-grown TaS2 is possible within hours following growth. As observed in XPS, in addition to the Ta 4f peaks from the TaS2, a large fraction of the Ta 4f peak signal from Ta2O5 is detected. After 10–20 h exposure to ambient without capping or any other protection, cross-section TEM was conducted on these flakes (figure 3(f)). As shown in figure 3(f), the top 2–3 nm of the TaS2 is oxidized. Furthermore, after 3 d the as-grown TaS2 is nearly fully oxidized, leaving column-shaped oxidation traces vertically through the TaS2 (SI). This directly indicates that the oxidation process is not self-limiting, and that care needs to be taken when considering the use of TaS2.

PV deposited 1T-TaS2 layers exhibit similar temperature-dependent properties compared to mechanically exfoliated 1T-TaS2. Low-temperature Raman indicates that the 1T-TaS2 transforms from the NCCDW phase to the CCDW phase below 180 K [22], where all the Ta atoms group into a $\sqrt{13}\times \sqrt{13}$ superstructure (figure 1(c)), which are commensurate with the underlying lattice. This new structure changes the Raman features dramatically [3]. At room temperature, 1T-TaS2 has a characteristic sharp peak at 70 cm−1, while 2H-TaS2 has three broad peaks located at 200 cm−1, 310 cm−1 and 395 cm−1 (figures 4(a) and (b)). While decreasing temperature, there are no significant changes for the 1T-TaS2 Raman peaks down to 180 K; however, once the temperature goes below 180 K, the main peak splits into two sub-peaks abruptly, while a group of small but sharp peaks appear near 100 cm−1. This peak evolution is readily tracked by Raman spectroscopy and is used to represent the CDW phase transitions in 1T-TaS2 [9]. On the contrary, within the available measurement temperature range of 80–300 K, there are no new peaks detected from 2H-TaS2 Raman. The noticeable change is the slow disappearance of A1 mode while lowering temperature, which shows the decreasing intensity from the two phonon scattering process [23]. On the other hand, the increasing intensity ratio from A2 mode to E mode also suggests enhanced inter-layer interactions [4].

Figure 4.

Figure 4. Raman measurements for both 1T-TaS2 and 2H-TaS2. Room temperature Raman measurements on (a) 1T-TaS2 and (b) 2H-TaS2 grown on sapphire, SiO2/Si and eG substrates. Raman measurements as a function of temperature from (c) 1T-TaS2 and (d) 2H-TaS2 demonstrate that the NCCDW-CCDW phase transition in 1T leads to the formation of new Raman peaks, while the 2H phase Raman spectra remains relatively unchanged.

Standard image High-resolution image

Electronic transport of TaS2 confirms that the transition of 1T-TaS2 from NCCDW to CCDW leads to a large increase in resistivity, while the 2H phase remains highly conductive. This is confirmed via two-terminal TaS2 devices fabricated on sapphire substrates (figure 5(a)). Figure 5(b) plots resistivity change as a function of temperature. The resistivity is derived from $~\rho =\frac{RWt}{L}$ , where R is the extracted resistance, t is the TaS2 thickness, and W/L the width to length ratio of the device. The 1T-TaS2 displays the CCDW to NCCDW transition at ~180 K, with a hysteresis of 20 °C during the cooling cycle returning from the NCCDW to CCDW state. This hysteresis is 4×  lower than that reported from similar measurements for exfoliated 1T-TaS2 [24], however, there is also a 2.5×  reduction in the magnitude of the transition compared to the CVT flakes. The reduction in magnitude is indicative of a higher concentration of defects within PV grown flakes (See supplemental images) compared with CVT materials [25, 26], which can negatively impact the electronic transition. Importantly, electrical switching of the PV grown 1T-TaS2 is also possible while the temperature remains below 130 K (figure 5(c)). We have also observed an abrupt, reversible insulator/metal transition (IMT) associated with the CCDW to NCCDW transition with a ΔIMT of ~2×  . At 130 K, the metal-to-insulator (MIT) transition in the reverse sweep direction becomes more gradual due to the slower switching time in comparison to the IMT in the forward direction [24]. The 2H-TaS2 device, however, does not display the abrupt IMT in either the ρ versus T or the IV over these temperature ranges [20].

Figure 5.

Figure 5. Device fabrication and measurements. (a) A schematic of device fabricated on TaS2 grown on sapphire substrate. (b) Resistivity of 1T-TaS2 and 2H-TaS2 as a function of temperature from 250 K to 77 K. 1T-TaS2 shows phase transition at 180 K with resistivity increase in the cooling cycle. (c) IV curves of 1T-TaS2 and 2H-TaS2. Electric switching of 1T-TaS2 has been detected for 1T-TaS2 flakes.

Standard image High-resolution image

We have introduced a single-step tellurium-assisted synthesis process for controllably creating high crystallinity 1T-TaS2 and 2H-TaS2 domains via powder vaporization. Synthesis via this technique is universal, and can be carried out on a variety of substrates (sapphire, SiO2/Si, eG). The 1T-TaS2 undergoes a NCCDW/CCDW phase transition at 180 K, leading to dramatic changes in the phononic and electronic properties, as measured by Raman and electrical characterization. This study provides a unique way to directly synthesize CDW TMDs where the corresponding transition metal oxides have very high melting temperatures and hard to be directly sulfurized. Our work may open up possibilities to directly integrate TaS2 into further complex structures.

Methods

Density functional theory

The electronic properties of different crystals are calculated by using generalized gradient approximations (GGA) as implemented in QuantumWise, Atomistix Toolkit (ATK) [4]. Within the DFT, the valence band wave functions of different atoms are treated in terms of a linear combination of atomic orbitals (LCAO) and the electronic properties of core electrons is described by norm-conserving Troullier–Martins pseudopotentials. In the LCAO pseudopotential calculations, we consider the Perdew–Burke–Ernzerhof (PBE) approximation for the exchange-correlation functional along with double ζ—double polarized (DZDP) basis on the atoms which are usually comparable to well-converged plane wave basis sets. To incorporate the long range van der Waals correction in interlayer interaction within the GGA approximation, we have included Grimme's DFT-D2 functional with S6  =  0.75 under PBE functional. Moreover, to calculate the individual BS properties we use a k-point sampling of 21  ×  21  ×  21 in the Brillouin zone. The tolerance parameter was set to a value of 10−5 along with mesh cut-off energy of 150 Ry on a real space grid of charge density and potentials. The optimized lattice parameters of individual crystals are listed in table 1.

Table 1. The optimized lattice parameters of individual crystals of undistorted 1T-TaS2. CCDW 1T-TaS2, and undistorted 2H-TaS2.

  1T TaS2 CCDW 1T TaS2 2H TaS2
a 3.3023 Å 12.1327 Å  3.3492 Å
c 5.6754 Å  5.6855 Å 12.1529 Å

Acknowledgments

This work was supported by NSF Grant No. EFRI-1433307. The authors acknowledge Dr Yi Wang for helpful calculation discussions.

Please wait… references are loading.