Shear-Induced Isostructural Phase Transition and Metallization of

Feb 16, 2016 - TMD materials crystallize into layered structures made of a varying degree of covalency and ionicity in M–X bonds; these layered stru...
0 downloads 9 Views 3MB Size
Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Article

Shear-Induced Isostructural Phase Transition and Metallization of Layered Tungsten Disulfide Under Non-Hydrostatic Compression Sakun Duwal, and Choong-Shik Yoo J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.5b10759 • Publication Date (Web): 16 Feb 2016 Downloaded from http://pubs.acs.org on February 17, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry C is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Shear-induced Isostructural Phase Transition and Metallization of Layered Tungsten Disulfide under Non-hydrostatic Compression Sakun Duwal and Choong-Shik Yoo* Department of Chemistry and Institute of Shock Physics, Washington State University, Pullman, Washington 99164, USA.

ABSTRACT Pressure-induced structural and electronic transformations of tungsten disulfide (WS2) have been studied to 60 GPa, in both hydrostatic and non-hydrostatic conditions, using four-probe electrical resistance measurements, micro-Raman spectroscopy and synchrotron x-ray diffraction. The results show the evidence for an isostructural phase transition from hexagonal 2Hc phase to hexagonal 2Ha phase, which accompanies the metallization at ~37 GPa. This isostructural transition occurs displacively over a large pressure range between 15 and 45 GPa and is driven by the presence of strong shear stress developed in the layer structure of WS2 under non-hydrostatic compression. Interestingly, this transition is absent in hydrostatic conditions using He pressure medium, underscoring its strong dependence on the state of stress. We attribute the absence to the incorporation of He atoms between the layers, mitigating the development of shear stress. We also conjecture a possibility of magnetic ordering in WS2 that may occur at low temperature near the metallization.

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

INTRODUCTION Transition metal dichalcogenides (TMDs or MX2 where M = Mo, W and X = S, Se) are semiconductors with indirect bandgaps ranging from 1.0-1.4 eV for bulk forms. These indirect gaps are located slightly under direct gaps ranging from 1.6–2.5 eV.1 Because of the presence of tunable bandgaps in TMDs, they are widely used in mechanical and electronic applications, 2 and are often subjected to high-pressure investigations. TMD materials crystallize into layered structures made of a varying degree of covalency and ionicity in M-X bonds; these layered structures control the electronic band gaps of bulk and thin films. Yet, these layers are held together only by weak van der Waals forces, giving rise to interesting mechanical properties such as extremely low friction shears and high compressibilities. Combining the two, strong covalent bonds within layers and weak van der Waals interactions between layers, results in highly anisotropic crystal structures and intriguing optical and electronic properties, significant to various applications for catalysis, solar cells, batteries, and optoelectronic devices.3,4 MoS2, for example, is predicted to be an indirect bandgap semiconductor in a bulk crystal, but a direct bandgap semiconductor in its monolayer form.3 MoS2 has also shown great promise as an electro-catalyst for hydrogen evolution reactions.5 For similar reasons, various TMDs in bulk and multilayers have been subjected to extensive studies at both ambient and high pressures.3,6,7,8,9 The application of external pressure is an excellent way to tune mechanical, chemical and electronic properties of TMDs and investigate the nature of chemical bonding, crystal structure, and phase transformation. Interestingly, greatly diverse properties are often observed at high pressures among various TMD materials, despite their periodic and structural similarities. This

2

ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

divergence, on the other hand, underscores the significance of strongly anisotropic crystal structure and chemical bonding, which can change greatly depending on the state of stress, structural distortion and topological packing. Multilayer MoS2, for example, undergoes a semiconductor-to-metal transition at 19 GPa,9 whereas bulk MoS2 undergoes an isostructural transition followed by a semiconductor-to-metal transition above 28 GPa.10 Moreover, MoS2 even becomes a superconductor at 5 K and 90 GPa.11 The structure of MoSe2 continuously evolves from anisotropic 2D layers to an isotropic 3D network, as pressure increases above 30 GPa.12 MoSe2 is also found to transform into a metallic phase between 20 and 35 GPa.12 Despite the extensive studies on MoS2 and MoSe26-12, there have been only a few studies on WS2 at high pressures. Because both Mo and W belong to the same Periodic group, it is often assumed that the corresponding dichalcogenides possess similar properties. Contrary to this assumption, Bandaru et al.13 reported that bulk WS2 exhibits no structural phase transition up to 52 GPa in hydrostatic conditions using helium as a pressure medium. Nayak et al.14 reported that multilayered WS2 undergoes an isostructural semiconductor-to-metal transition at 22 GPa at 280 K. While these studies have furthered the understanding of high-pressure behavior on WS2, the results seem to show diverse behaviors of WS2 depending on the state of stress. To address this issue, we have investigated WS2 in both hydrostatic and non-hydrostatic conditions. The present results indeed indicate a strong stress dependence on the transition in WS2. WS2 undergoes an isostructural phase transition only in non-hydrostatic conditions, accompanied by a semiconductor-to-metal transition, as previously observed in other TMDs, such as MoS210 and WSe2.15

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

EXPERIMENTAL METHODS WS2 powder (99.8% purity from Alfa Aesar) was used in the present study. The powder samples were loaded into a small hole (140 µm in diameter) drilled on a Re gasket (200 µm thick pre-indented to ~40 µm) mounted between two diamond anvils (Type IA, 300 µm flats) of membrane-driven diamond anvil cell (DAC) without a pressure medium for non-hydrostatic experiments. For hydrostatic experiments, high-pressure helium gas (~2000 atmospheres) was used as a pressure-transmitting medium (PTM), loaded using a custom-built high-pressure gas loader. A few small chips of ruby crystals were also loaded together with the sample for pressure determination based on the peak position of R1 luminescence.16 Raman spectra were collected using a home-built, confocal micro-Raman system that consists of a Nd:Yag laser (532 nm, Verdi 6W, Coherent); a 20X objective lens (infinitive corrected, Edmund); a confocal device (a pair of lenses and a 2D knife-edge slit); a 0.5m spectrograph; a liquid N2-cooled, charge-coupled device (CCD) detector; and various optics such as a holographic diffraction beam splitter and a Raman notch filter (both from Kaiser Optics). Raman spectra were collected in a back-scattering geometry with a spectral resolution of ~0.3 cm-1. Electric resistance of the sample was measured using a standard four-probe technique for the DAC application. Alumina powder was used to insulate the gasket. A sample hole was drilled at the center of insulated alumina layer densely packed in the gasket hole. Four 5 µm thick platinum electrodes were placed on the top of insulating alumina, and Cu wires were used as an extension to measure the resistance. A fifth Cu wire was soldered onto the gasket to check

4

ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

for a short circuit while taking the measurements. The sample resistance was measured using a one µA-current power source (Lakeshore 120) and a voltmeter (Keithley 2000 DVM). For diffraction experiments, we used micro-focused (20 µm x 20 µm) synchrotron x-rays at 16BMD/HPCAT (λ = 0.4066 Å) at the Advanced Photon Source. The x-ray diffraction intensities were recorded over a large range of 2θ between 3 and 40 degrees, using a large 2D image plate detector (MAR CCD). We used the Fit2D software to convert the 2D diffraction image to the 1D angle-resolved x-ray diffraction (ARXD) pattern and, then, employed various diffraction analysis software including XRDA, GSAS and CRYSTAL MAKER to analyze the ARXD data.

RESULTS Figure 1 (a) shows the crystal structure of double-layered hexagonal (2H) WS2, which 4 belongs to the space group D6h (P63/mmc). In this structure, tungsten ions are coordinated with

six sulfur atoms forming edge-shared trigonal prismatic layers. These layers are then stacked up along the c-axis (or the [001] direction) bounded by weak van der Waals forces. A unit cell of WS2 is then composed of two metal atoms occupying sites with the point-group symmetry D3h. This structure with two WS2 per unit cell allows a total of 18 modes of lattice vibrations: Гvib = A1g + 2A2u + 2B2g +B1u+ E1g + 2E1u + 2E2g + E2u. These modes are then composed of seven Raman active (A1g, E1g and 2E2g); three IR active modes (A2u and E1u); three translational acoustic modes (A2u and E1u); and five inactive modes (2B2g, B1u and E2u). Vibrations of the 1 2 Raman active modes are represented in Fig. 1b. The 2-fold degenerate E2g (E2g + E2g ) modes

represent in-plane vibrations of tungsten and sulfur atoms (shear modes) 5

ACS Paragon Plus Environment

18

, whereas the A1g

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

mode corresponds to out-of-plane sulfur vibrations (a surface mode) along the c-axis. The IR active A2u and E1u modes are then understood in terms of the asymmetric counter parts of surface A1g and shear E1g modes, where tungsten and sulfur atoms move in the opposite directions. 1 Figure 2 shows Raman spectra of WS2 consisting of the A1g and E2g modes at several

pressures in (a) non-hydrostatic and (b) hydrostatic conditions. At ambient pressure, the A1g 1 mode is observed at ~421 cm-1. The E2g mode is observed at ~353 cm-1 as shown in Fig 2a and

2b. Even though the E1g mode is Raman active, this mode only appears as a very weak feature in the present back-scattering setup.1 This is because the E1g phonon is forbidden in a backscattering geometry on a surface perpendicular to the c-axis and, at high pressures, the layered structure develops a highly preferable orientation along the same [001] direction. The peak at 266 cm-1 corresponds to a higher order resonant Raman peak, previously assigned based on the neutron inelastic scattering data.19 Other weak Raman modes have also been assigned in Table 1 of Ref. 1; the peak at ~300 cm-1 to the resonant phonon mode of 2LA – 2 E2g2, the ~520 cm-1 2 peak to the E2g1 + LA mode, and the ~585 cm-1 peak to the A1g + LA mode.1 The other E2g mode

is known to appear at a very low wavenumber (~ 27 cm-1), which is outside the spectral window of our instrumentation. 1 The A1g mode is observed until 62 GPa, whereas the E2g mode disappeared at around 37

GPa, as shown in Fig. 2 (a) when no pressure-transmitting medium (PTM) was used. The higher 1 intensity of the A1g mode, compared to the E2g at 0.2 GPa, can be understood in terms of strong

coupling of A1g phonons with excited dz2 states, as the dz2 orbital points perpendicular to the basal plane. Hence, the displacement of atoms along the c-direction has a greater effect on polarizability, resulting in higher intensity for the A1g mode. On the other hand, the atomic

6

ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1 displacement in the E2g vibration is parallel to the plane, which gives a smaller change in 1 polarizability. So, the E2g mode is expected to weaken as pressure increases. However, at

pressures below 20 GPa, the relatively stronger intensity of these two modes keeps fluctuating. 1 is buried in the twice of longitudinal acoustic (LA) peak at 352 cmThis is likely because the E2g 1 18,20

.

At 0.4 GPa the intensity of the two peaks switches, and a shoulder on the A1g starts to

appear. Following the same interpretation by Zhao et al.20, we interpret this new feature on the A1g shoulder as a Raman-inactive B1u mode. The B1u mode is a part of Davydov doublet with the A1g mode, arising as a result of resonant effect.21 The appearance of the B1u mode can then be an indication of enhanced disorder in the layered structure introduced by shear frictions. Importantly, the intensity of Raman spectra rapidly decreases as pressure increases above 15 GPa. At 44 GPa nearly all Raman features disappear except a small remnant of the A1g mode. Therefore, this result appears to indicate a structural deformation occurring over a broad pressure range, starting from ~15 GPa and finishing above ~44 GPa, where WS2 becomes a metallic solid. Raman spectra of WS2 collected in hydrostatic conditions using He as a PTM (Fig. 2b) is consistent with those previously reported by Bandaru et al.13 As pressure increases, the A1g mode 1 shifts much faster at the rate of 2.27 (±0.06) cm-1/GPa than the E2g mode at 1.33 (±0.05) cm1

/GPa (see Fig. 3). This makes sense because the A1g mode corresponds to interlayer sulfur-

1 sulfur vibrations, which are bound by weak van der Waals interactions. The E2g mode

corresponds to tungsten-sulfur vibrations that are linked via strong covalent bonds. Furthermore, the present peak positions of WS2 agree well with those of Bandaru et al.,13 but are slightly higher than those of Nayak et al.14 for multi-layered WS2, as compared in Fig. 3. These spectral

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

changes are reversible upon pressure unloading in both hydrostatic and non-hydrostatic conditions. The present Raman data clearly shows that high-pressure behaviors of WS2 depend strongly on the state of stress in the sample. It also reconciles the different observations between Bandaru et al.13 reporting the absence of phase transition in hydrostatic conditions and Nayak et al.14 reporting a continuously occurring isostructural transition in non-hydrostatic conditions. It is believed that the presence of small helium atoms limits the slip motion of the layers and thereby inhibits the occurrence of shear-induced structural transformation. In fact, helium is known to resist the shear stress.22 In this respect, the presence of the A1g shoulder peak in non-hydrostatic conditions seems to support the occurrence of shear-driven structural change. At ambient pressure, WS2 is a semiconductor with an indirect bandgap of 1.35eV.23 We found that upon compression it transforms to a metallic state as shown in Fig. 4. The resistance of WS2 at ambient condition is ~104 Ω (or the estimated resistivity of ~2 Ωm) – the value after the initial compaction of powder samples and also after the pressure unloading. Upon compression, this resistance value decreases rapidly, at least by three orders, to a steady value of less than ~10 Ω at ~37 GPa, indicating a semiconductor-to-metal transition. For the current sample configuration in a DAC, the residual resistance of 10 Ω is typical for that of metal or semimetal. In fact, the temperature-dependent resistance change (shown in Fig. 4 inset) supports the metallic behavior above 37 GPa. Interestingly, the resistance curve at 37 GPa also shows an anomalous small bump at about 100 K, which could be an indication of structural changes and/or magnetic ordering at this temperature. The sample was further compressed to 110 GPa; however, no evidence for superconductivity was observed.

8

ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

In order to understand the structural change in non-hydrostatic conditions, we have obtained the x-ray diffraction data of WS2 to 60 GPa. The results are plotted in Fig. 5: (a) the measured ARXD data to 57 GPa and (b) the d-spacings of observed (hkl) lattice planes as a function of pressure. The calculated peak positions in Fig. 5a are based on a hexagonal structure in P63/mmc using the lattice parameters of a = 3.151 Å and c = 12.055 Å. The systematics of observed diffraction patterns are well matched to the calculated ones for the same structure, indicating no significant structure change under pressure. However, it is important to note that there are several subtle yet apparent changes in the diffraction patterns. For example, the (004) peak shifts rapidly and merges into the doublet of (100) and (101) peaks at 39 GPa. The doublet itself merges together, forming a single broad feature at high pressures. The (105) peak (marked by arrows) disappears at 45 GPa, while a new peak (marked by asterisks) emerges as a left shoulder of the (103) above 27 GPa. Importantly, these changes, the emergence of new feature and the disappearance of the (105), were not observed in the hydrostatic condition reported in ref. 13. The faster pressure shifts of both (002) and (004) indicate that the c-axis is much more compressible than the a-axis (see Figs. 5b and 6a). This is clearly due to relatively weak van der Waals interactions along the c-axis. Note that there is a small difference in the a-axis observed in between hydrostatic (open symbols) and nonhydrostatic (close symbols) conditions. This makes that the c/a ratio in nonhydrostatic conditions gradually decreases over the entire pressure range, whereas that in hydrostatic condition initially decreases at low pressures below 10-15 GPa and then stays unchanged at higher pressures. Therefore, we attribute the observed diffraction changes to a subtle structural transition occurring over a broad pressure range between ~20 GPa and 45 GPa. The sluggish nature of the transition and the large collapse of the c-axis are in support of a shear-driven, displacive phase transition.

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 27

The pressure-dependent structure change is illustrated in Fig. 6 in terms of (a) lattice parameters and (b) the pressure-volume compression curve. The cell parameters from the present study are in agreement with those of ref. 13. Note that the c-axis is substantially more compressible than the a- or b-axis. This anisotropic compression (also shown in the lower panel of Fig. 6a), in turn, results in a rapid decrease in interlayer S…S distances with respect to intralayer W-S and S-S distances and thereby a rapid increase of interlayer S…S repulsion with pressure. Therefore, we attribute this anisotropic compression to be responsible for an isostructural phase transition from hexagonal 2Hc phase to another hexagonal 2Ha phase, as observed previously in MoS2.9,10,26 In the inset of Fig. 6b, the 2Hc and 2Ha structures are presented in the ab-plane. Note that in the 2Ha structure, sulfur atoms in two nearby trigonal prismatic sheets are staggered along the c-axis, minimizing the interlayer S…S repulsion. This is in contrast to sulfur atoms facing the trigonal face centers in the 2Hc structure. Evidence for a subtle structure change in WS2 can also be found in the pressure-volume compression curve in Fig. 6b. The calculated volumes (red line) are based on the BirchMurnaghan equation of state (BM-EOS) fit to the measured ones. The BM-EOS fit gives B0 = 69 ± 2.0 GPa (Bo’ = 6.5 ± 0.1), which is almost comparable to the value obtained by Bandaru et al of B0 = 63 ± 1 GPa (B0’ = 6.5 ± 0.1) 13 and Selvi et al B0 = 61 ± 1 GPa (B0’ = 9.0 ± 0.3).2 It is important to note that the measured volumes above ~42 GPa deviate from the calculated EOS fit, suggesting a subtle structure change.

10

ACS Paragon Plus Environment

Page 11 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

DISCUSSION The present spectral and diffraction results suggest that WS2 undergoes an isostructural phase transition in non-hydrostatic conditions. The evidence for the structural transition includes the emergence of new Raman and diffraction peaks, as well as the deviation of specific volumes from the BM-EOS fit at high pressures. The absence of the same transition in hydrostatic conditions, on the other hand, seems to indicate that the transition is kinetically controlled by the presence of strong shear stress in this layered structure of WS2. We conjecture that the presence of helium atoms between the layers blocks the movement of the slip planes (i.e., the (110) layers). The sluggish nature of transition supports this mechanism, which occurs over a broad pressure range from ~15 GPa to ~45 GPa. The EOS plot shows a discontinuity, which is evidence for a first-order phase-transition. Furthermore, the cell parameters versus the pressure plot show (Fig. 6a) some discontinuity in the decreasing trend along the a-axis, but not along the c-axis. This illustrates that the pressureinduced structure change is likely related to the a-axis than the c-axis. This is consistent with the observed difference in the a-axis between nonhydrostatic and hydrostatic conditions (Fig. 6a). Each layer of S-W-S can slide over each other under high pressure.24 This kind of sliding can result in an isostructural phase transition from 2Hc (hexagonal, P63/mmc) to 2Ha (P63/mmc). In the 2Hc structure, the metal atoms occupy the positions at 2c (1/3 2/3 ¼), and the sulfur atoms at 4f (1/3 2/3 z).25 In the 2Ha, the metal atoms are at 2b (0 0 ¼) and the sulfur atoms at 4f (1/3 2/3 z).

25

This transition involves change in the interlayer stacking from (CaCAcA) in 2Hc to

(AbACbC) in 2Ha.25 The z value used in this study are taken from ref. 26. A similar transition has also been observed in MoS2.9, 10, 26 Structurally, the 2Ha structure is less anisotropic than the initial 2Hc. It has a smaller interlayer spacing and larger in-plane interatomic spacing in

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 27

comparison to the initial 2Hc phase.27 Furthermore, ref. 27 mentions that the demise of the 2Hc – MoS2 phase is signaled by the disappearance of (105)27, which is consistent with the present xray data for WS2. This further confirms that WS2 goes from 2Hc to 2Ha phase upon compression. In the P63/mmc structure, tungsten ions (d2) are coordinated with six sulfur atoms in a trigonal prismatic coordination (D3h), which splits the 5d-subshells of tungsten and causes a large energy gap between fully occupied dz2 and empty degenerated dxy and dx2-y2.28 Upon compression, this gap is expected to decrease as roughly proportional to the electron kinetic energy, ~ρ2/3. Furthermore, upon anisotropic compression, the gap decreases even faster, as sulfur 3pz orbitals hybridize with tungsten dxy and dx2-y2 orbitals more effectively than the dz2. Therefore, it is conceivable that the observed structural distortion in the ab-plane to 2Ha phase brings these states nearby or even pushes the dxy and dx2-y2 below the dz2, giving rise to a paramagnetic state. The anomalous resistivity change observed at ~100 K at 37 GPa (Fig.3 inset) may then be understood in terms of a magnetic ordering transition. The excitation of electrons occurs from this filled dz2 orbital to the empty dxy or dx2-y2 orbitals. Theoretical band structure calculations29,

30

suggest that the conduction band is

determined primarily by the transition metal, and the valence band is determined both by the transition metal and the chalcogenide. Upon compression, the separation between this band decreases and the charge carriers of sulfur 3pz orbitals from the valence band increase. As a result, resistance decreases with increasing pressure. This transition is similar to the semiconductor-to-semimetal transition observed in WSe215 at 38.1 GPa and the semiconductorto-metal transition in MoS2 at around 20 GPa.10 In the semi-metallic transition for WSe2 (ref.15), the transition was attributed to W-Se covalent bonding in the layer rather than van der Waals bonding between the layers. This same mechanism can be held true for the transition in WS2 as

12

ACS Paragon Plus Environment

Page 13 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

well. The increased mixing of W 5d and S 3p can result in an enhanced orbital hybridization in the layer, leading to the observed metallic conductivity in WS2. The same transition to metallic MoS2 occurs at a considerably lower pressure compared to that of WS2. This can be a result of the lower ionization potential of Mo than that of W. The lower potential makes it easier for Mo to pass on charges, resulting in the lower transition pressure.9 Substantially reduced Raman intensities above 25 GPa in Fig. 2a are likely due to a metallic nature of WS2. In the case of multi-layered WS2, the semiconductor-to-metal transition was observed at 22 GPa with a four order decrease in resistivity, 14 which is similar to the resistance trend observed in the present study. This semiconductor-to-metal transition in multilayered WS2 seems to occur at a substantially lower pressure than what we observe in bulk WS2 (~37 GPa). This difference could be due to the fact that the pressure at which metallization occurs decreases with an increase in the number of WS2 layers.14 In comparison, multi-layer MoS2 metallizes at 19 GPa,9 whereas bulk MoS2 metallizes above 28 GPa.10 The lower metallization pressures in MoS2 than those in WS2 may be understood in terms of a larger crystal field effect in W than Mo and a better hybridization of 3p orbitals of S with 4d orbitals of Mo than with 5d orbitals of W.

CONCLUSION We have presented the pressure-induced metallization in WS2 at ~37 GPa, following an isostructural 2Hc to 2Ha phase transition. The isostructural transition occurs displacively over a large pressure range between 15 and 45 GPa and is driven by the presence of strong shears developed in the layer structure of WS2 in non-hydrostatic conditions. We conjecture that the

13

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 27

structural distortion to the 2Ha structure may give rise to a magnetic ordering transition at low temperatures near the metallization.

AUTHOR INFORMATION Corresponding Author: Correspondence and requests for materials should be addressed to C. S. Yoo at [email protected], (509) 335-2712 Notes: The authors declare no competing financial interest.

ACKNOWLEDGEMENTS The present study has been performed in support of NSF-DMR (Grant No. 1203834) and DTRA (HDTRA1-12-01-0020). The x-ray work was performed at the HPCAT in support of CDAC. HPCAT operates in the supports of DOE-NNSA (DE-NA0001974) and DOE-BES (DEFG02-99ER45775). We thank Dr. Minseob Kim for his assistance in synchrotron x-ray diffraction experiments at the APS.

14

ACS Paragon Plus Environment

Page 15 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

REFERENCES 1. Bhatt, S. V.; Deshpande, M. P.; Sathe, V.; Rao, R.; Chaki, S. H. Raman Spectroscopic Investigations on Transition-metal Dichalcogenides MX2 (M = Mo, W; X = S, Se) at High Pressures and Low Temperature. J. Raman Spectrosc. 2014, 45, 971–979. 2. Selvi, E.; Ma, Y.; Aksoy, R.; Ertas, A.; White, A. High Pressure X-ray Diffraction Study of Tungsten Disulfide. J. Phys .Chem. Solids. 2006, 67, 2183-2186. 3. Kuc, A.; Zibouche, N.; Heine, T. Influence of Quantum Confinement on the Electronic Structure of the Transition Metal Sulfide. Phys. Rev. B 2011, 83, 245213 . 4.

Fleischauer, P. D.; Lince, J. R.; Bertrand, P.A.; Bauer , R. Electronic Structure and Lubrication Properties of MoS2: A Qualitative Molecular Orbital Approach. Langmuir. 1989, 5, 1009-1015.

5.

Nolan, H.; McEvoy, N.; O’Brien, M.; Berner, N. C.; Yim, C.; Hallam, T.; McDonald, A. R.; Duesberg, G. S. Molybdenum Disulfide/Pyrolytic Carbon Hybrid Electrodes for Scalable Hydrogen Evolution. Nanoscale. 2014, 6, 8185-8191.

6.

Frindt, R. F.; Yoffe, A. D. Physical Properties of Layer Structures: Optical Properties and Photoconductivity of Thin Crystals of Molybdenum Disulphide. Proc. R. Soc. A 1963, 273, 69-83.

7.

Frindt, R. F. Single Crystals of MoS2 Several Molecular Layers Thick. J. Appl. Phys. 1966, 37, 1928-1929.

8.

Yang, R.; Zheng, X.; Wang, Z.; Miller, C. J.; Feng, P. X. L. Multilayer MoS2 Transistors Enabled by a Facile Dry-Transfer Technique and Thermal Annealing. J. Vac. Sci. Technol. B 2014, 32, 061203-9.

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

9.

Page 16 of 27

Nayak, A. P.; Bhattacharya, S.; Zhu, J.; Liu, J.; Wu, X.; Pandey, T.; Jin, C.; Singh, A. K.; Akinwande, D.; Lin, J. F. Pressure-Induced Semiconducting to Metallic Transition in Multilayered Molybdenum Disulphide. Nat. Commun. 2014, 5.10.1038/ncomms4731.

10. Chi, Z. H.; Zhao, X. M.; Zhang, H.; Goncharov, A. F.; Lobanov, S. S.; Kagayama, T.; Sakata, M.; and Chen, X. J. Pressure-Induced Metallization of Molybdenum Disulfide. Phys. Rev Lett. 2014, 113, 036802-5. 11. Chi, Z., Yen, F.; Peng, F.; Zhu, J.; Zhang, Y.; Chen, X.; Yang, Z.; Liu, X.; Ma, Y.; Zhao, Y.; Kagayama, T.; Iwasa, Y. Ultrahigh Pressure Superconductivity in Molybdenum Disulfide. Arxiv. 2015, 1503, 05331. 12. Zhao, Z.; Zhang, H.; Yuan, H.; Wang, S.; Lin, Y.; Zeng, Q.; Xu, Z.; Liu, Z.; Solanki, G. K.; Mao, W. L. Pressure Induced Metallization With Absence of Structural Transition in Layered Molybdenum Diselenide. Nature Commun. 2015, 6, 7312. 13. Bandaru, N.; Kumar, R. S.; Baker, J.; Tschauner, O.; Hartmann, T.; Zhao, Y.; Venkat, R. Structural Stability of WS2 Under High Pressure. Int. J. Mod. Phys. B 2014, 28, 1450168. 14. Nayak, A. P.; Yuan, Z.; Cao, B.; Liu, J.; Wu, J.; Moran, S. T.; Li, T.; Akinwande, D.; Jin, C.; Lin, J.F. Pressure-Modulated Conductivity, Carrier Density, and Mobility of Multilayered Tungsten Disulfide. ACS Nano. 2015, 9, 9117-9123. 15. Liu, B.; Han, Y.; Gao, C.; Ma, Y.; Peng, G.; Wu, B.; Liu, C.; Wang, Y.; Hu, T.; Chi, X.; et al. Pressure Induced Semiconductor-Semimetal Transition in WSe2. J. Phys. Chem. C 2010, 114, 14253-14254. 16. Mao, H.; Xu, J.; Bell, P. Calibration of the Ruby Pressure Gauge to 800 Kbar Under QuasiHydrostatic Conditions. J. Geophys. Res. 1986, 91, 4673.

16

ACS Paragon Plus Environment

Page 17 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

17. Tonndorf, P.; Schmidt, R.; Böttger, P.; Zhang, X.; Borner, J.; Liebig, A.; Albrecht, M.; Kloc, C.; Gorgan, O.; Zahn, D. R. T.; et al. Photoluminescence Emission and Raman Response of Monolayer MoS2, MoSe2, and WSe2. Opt. Express. 2013, 21, 4908–4916. 18. Zhao, Y.; Luo, X.; Li, H.; Zhang, J.; Araujo, P. T.; Gan, C. K.; Wu, J.; Zhang, H.; Quek, S. Y.; Dresselhaus, M. S.; et al. Interlayer Breathing and Shear Modes in Few-Trilayer MoS2 and WSe2. Nano Lett. 2013, 13, 1007-1015. 19. Sourisseau, C.; Cruege, F.; Fouassier, M. Second-order Raman Effects, Inelastic Neutron Scattering and Lattice Dynamics in 2H-WS2. Chem. Phys. 1991, 150, 281-293. 20. Zhao, W.; Zohreh, G.; Kumar, A. K.; Pang, J. R.; Toh, M.; Zhang, X.; Kloc, C.; Tan, P. H.; Eda. G. Lattice Dynamics in Mono- and Few-Layer Sheets of WS2 and WSe2. Nanoscale. 2013, 5, 9677-9683. 21. Livneh, T.; Sterer, E. Resonant Raman Scattering at Exciton States Tuned by Pressure and Temperature in 2H-MoS2. Phys. Rev. B 2010, 81, 195209. 22. Singh, A. K. Strength of Solid Helium Under High Pressure. Journal of Physics: Conference Series. 2012, 377, 012007. 23. Kam, K. K; Parkinson, B. A. Detailed Photocurrent Spectroscopy of the Semiconducting Group VI Transition Metal Dichalcogenides. J. Phys. Chem. 1982, 86, 463-467. 24. Aksoy, R.; Ma, Y. Z.; Selvi, E.; Chyu, M. C.; Ertas, A.; White, A. X-ray Diffraction of Molybdenum Disulfide to 38.8 GPa. J. Phys. Chem. Solids. 2006, 67, 1914-1917. 25. Katzke, H.; Tolédano, P.; Depmeier, W. Phase Transitions Between Polytypes and Intralayer Superstructures in Transition Metal Dichalcogenides. Phys. Rev. B 2004, 69, 134111.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

26. Bandaru, N.; Kumar, R. S.; Sneed, D.; Tschauner, O.; Baker, J.; Antonio, D.; Luo, S. N.; Hartmann, T.; Zhao, Y.; Venkant, R. Effect of Pressure and Temperature on Structural Stability of MoS2. J. Phys. Chem. C 2014, 118, 3230-3235. 27. Hromadová, L.; Martonák, R.; Tosatti, E. Structure Change, Layer Sliding, and Metallization in High-Pressure MoS2. Phys. Rev. B: Condens. Matter Mater. Phys. 2013, 87, 144105. 28. Tsai, H.L.; Heising, J.; Schindler, J. L.; Kannewurf, C.R.; Kanatzidis, M.G. ExfoliatedRestacked Phase of WS2. Chem. Mater. 1997, 9, 879-882. 29. Jin, W.; Yeh, P.; Zaki, N.; Zhang, D.; Sadowski, J. T.; Al-Mahboob, A.; M. van der Zande, A.; Chenet, D. A.; Dadap, J. I.; Herman, I. P.; et al. Direct Measurement of the ThicknessDependent Electronic Band Structure of MoS2 Using Angle-Resolved Photoemission Spectroscopy. Phys. Rev. Lett. 2013, 111, 106801. 30. Guo, H.; Yang, T.; Tao, P.; Wang, Y.; Zhang, Z. High Pressure Effect on Structure, Electronic Structure and Thermoelectric Properties of MoS2. J. of Appl. Phys. 2013, 113, 013709.

18

ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

FIGURE CAPTIONS Fig. 1 (a) The crystal structure of WS2 in P63/mmc (or 2Hc phase) at ambient conditions. (b) Schematics of vibrational Raman modes (reproduced from ref.17). The red dotted line represents weak van der Waals bonds between the adjacent layers. Fig. 2 (a) Raman Spectra of WS2 at room temperature in non-hydrostatic condition, showing appearance of a shoulder peak in A1g mode and weakening of the Raman peaks above 25 GPa. (b) Raman Spectra of WS2 at room temperature using helium as a pressure transmitting medium (PTM). Fig. 3 (a) Pressure-induced spectral shifts of Raman modes observed in WS2. Solid and open circles correspond to the present data taken with no PTM and He PTM, respectively. Triangles correspond to the previous data by Bandaru et al.13, and short-dashes correspond to the previous data by Nayak et al.14 for multi-layered WS2. Fig. 4 Pressure-induced electric resistance change of WS2 showing a semiconductor-to-metal transition. The upper-right inset shows the temperature dependence of resistance at 37 and 82 GPa, showing a typical metallic behavior. The bottom-left inset is a microphotograph of WS2 sample with four Pt electrodes. Fig. 5 (a) Powder x-ray diffraction pattern of WS2 showing a structural phase transition over a broad pressure range of 15-42 GPa. The structural transition is evident by disappearance of the (105) peak at ~45 GPa and an emergence of small shoulder marked by asterisks above 15 GPa. Short vertical bars at the bottom are the calculated positions of (hkl) planes labeled on the diffraction pattern at 5 GPa. (b) Pressure dependence of d-spacings

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

of WS2 showing disappearance of (004), (101), and (105) peaks. The symbols represent the experimental data and the lines represent the polynomial fit. Fig. 6 (a) Pressure dependence of lattice parameters (a, c/2 and c/a in the top panel) and linear compressibilities (c/co and a/ao at the bottom), showing a highly anisotropic layer structure of WS2. The solid symbols represent data from the present study and the open symbols represent data from Bandaru et al.13 The lines represent the quadratic fits to the data. (b) Pressure-volume compression curve of WS2. The red line is the fit with BirchMurnaghan equation of state. The blue line is polynomial fit drawn at the phase transition region. The insets show crystal structures of 2Hc and 2Ha phases, showing the top view down the c-axis of a 3x2x1 supercell.

20

ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Graphic Table of Content and the Caption: Pressure-induced resistance (blue) and estimated resistivity (red) changes of WS2, showing a metallization and a shear-driven isostructural 2Hc to 2Ha phase transition (in the inset). A microphotograph of WS2 sample in a four-probe configuration for electric resistance measurements is also shown in the inset.

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

(a )

Page 22 of 27

(b )

ACS Paragon Plus Environment

F ig u re 1

Page 23 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

The Journal of Physical Chemistry

(a )

(b )

ACS Paragon Plus Environment

F ig u re 2

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

Page 24 of 27

ACS Paragon Plus Environment

F ig u re 3

Page 25 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

The Journal of Physical Chemistry

ACS Paragon Plus Environment

F ig u re 4

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

(a )

Page 26 of 27

(b )

ACS Paragon Plus Environment

F ig u re 5

Page 27 of 27

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

The Journal of Physical Chemistry

(a )

(b )

ACS Paragon Plus Environment

F ig u re 6