In situ synchrotron XRD experiments of compressed ZrS2 were conducted up to 32.3 GPa at room temperature. Figure 2a displayed representative XRD patterns during the compression. The diffraction patterns at pressures below 5.5 GPa are readily indexed to the trigonal phase (\(P\stackrel{-}{3}m1\), phase I) as illustrated in the inset of Fig. 2b with Re gasket and minor amount of ruby. All of Bragg peaks for the trigonal ZrS2 shifted to higher angles with increasing pressure, indicating conventional compression behavior. Pronounced changes in XRD pattern were observed when pressure was increased to 5.5 GPa, at which several new diffraction peaks appeared at 2θ angles of around 7.6°, 7.8°, 8.0°, 12.9°, 13.4° and 13.9°, suggesting the onset of structural phase transition (devoted as phase II). Further compression beyond 17.4 GPa results in the disappearance of several diffraction peaks, leaving only four broad peaks (as shown in Fig. S1 of the Supplemental Material). At this point, the crystalline of the sample began to diminish, and a degree of structural disordering (denoted by phase III) came into play. The limited number and severe broadening of diffraction peaks in phase III preclude determination of its lattice parameters.
We employed a Monte Carlo indexing algorithm25 to identify the lattice associated with the newly appeared diffraction peaks of phase II and refined its space group using the Crysfire software package26. The results pointed to an orthorhombic structure with the space group of Pmm2 with lattice parameters a = 3.297 (6) Å, b = 3.627 (7) Å, c = 9.509 (11) Å and V = 56.85 (34) Å3 at 5.5 GPa. Although the Pmm2 was not predicted as a stable phase by previous structural searching simulation22, it can be realized as a conjugate subgroup of the host tetragonal structure (I4/mmm), which was previously regarded as the ground state at above 25 GPa. Such lattice distortion was well documented for phase transitions under spontaneous strain27. The indexed twelve XRD peaks for this high-pressure phase are listed in Table 1.
Table 1
The indexed XRD peaks for Phase II at 5.5 GPa.
Phase II (5.5 GPa) | h | k | l | 2θ (degree) | 2θ (degree) | Δ 2θ |
experiment | refined |
Orthorhombic (Pmm2) a = 3.297 (6) Å b = 3.627 (7) Å c = 9.509 (11) Å V = 56.85 (34) Å3 | 1 | 0 | 0 | 7.599 | 7.6048 | -0.0058 |
0 | 0 | 3 | 7.818 | 7.8478 | -0.0298 |
1 | 0 | 1 | 8.012 | 8.0428 | -0.0308 |
1 | 0 | 2 | 9.226 | 9.2337 | -0.0077 |
1 | 0 | 3 | 10.975 | 10.9363 | 0.0387 |
1 | 1 | 3 | 12.885 | 12.9028 | -0.0178 |
0 | 2 | 1 | 13.919 | 13.9264 | -0.0074 |
0 | 1 | 5 | 14.733 | 14.7837 | -0.0507 |
2 | 0 | 1 | 15.443 | 15.4685 | -0.0255 |
0 | 2 | 4 | 17.275 | 17.2546 | 0.0204 |
2 | 1 | 4 | 19.809 | 19.7664 | 0.0426 |
1 | 0 | 7 | 19.935 | 19.9125 | 0.0225 |
The variation of unit cell volume with pressure of ZrS2 is plotted in Fig. 3a. The unit cell volume is abruptly decreased by 8.8% after the transformation from P\(\stackrel{-}{3}\)m1 to Pmm2 phase, thus attributes to first-order transition. The pressure-volume data for different phases were fitted using a third-order Birch-Murnaghan equation of state in EosFit7 program28. The derived bulk modulus K0 and zero-pressure unit cell volume V0 for phase I were 32.7 (88) GPa and 68.6 (10) Å3, respectively. As for phase II, we acquired the bulk modulus K0 = 66.1 (17) GPa, and V0 = 61.3 (1) Å3. The higher K0 value for phase II indicated that the eight-coordinated structure is less compressible. The pressure dependence of lattice parameter ratios is illustrated in Fig. 3b and 3c. For the ZrS2 phase I, the c axis is more compressible than the a axis due to the weak vdW forces existed in the c axial orientation. In contrast, the high-pressure phase II exhibited much weak anisotropy in comparison with the P\(\stackrel{-}{3}\)m1 phase.
The substantial volume collapse and severe lattice distortion associated with phase II makes it impracticable for the precise determination of atomic coordinates. However, considering Phase II as a distorted tetragonal phase with two formula units in a conventional unit cell, and assuming the atomic coordinates from the undistorted tetragonal phase22, each Zr atom would be coordinated by 8 neighboring S atoms, forming ZrS8 cuboids due to the elongation of b axis. This distortion maintains a substantial interlayer distance (4.75 Å at 5.5 GPa), Suggesting that the phase II may retains a layered structure. In particular, the interlayer structural transition in ZrS2 can be compared with other AB2-type TMDs with similar lattice structure, such as MoS2, WS2, and WSe215,29,30. All of such layered binary compounds were reported to undergo pressure-induced phase transition through layer sliding, which is associated with the lateral shift of adjacent atom layers, and thus those are isostructural phase transition and still remain the layered nature29. However, the phase transition of ZrS2 reconstructed the interlayer lattice to form an orthorhombic structure and the coordination number of Zr changed from six to eight. The partially disordered phase III may eventually collapse into a 3D structure by layer sliding. The postponed layer sliding in ZrS2 is possibly resulted from the relatively weak in-plane bonding in ZrS2.The absence of layer sliding in ZrS2 is possibly resulted from the relatively weak in-plane bonding in ZrS2.
We also conducted Raman spectroscopy with focus on the interlayer structure. Theoretical group analysis of lattice vibrations of 1T-ZrS2 at the Γ-point predicted two Raman-active modes represented as Γ = A1g + Eg + 2A2u + 2Eu 31. Figure 4a depicts Raman spectra collected at the frequency range from 100 cm− 1 to 600 cm− 1 in DAC up to 40.7 GPa at room temperature. At ambient conditions, the in-plane (Eg) and out-of-plane (A1g) Raman active modes of ZrS2 were clearly captured at positions of 248.8 cm− 1 and 332.7 cm− 1, respectively. A broadening peak near the A1g band appeared at 312.7 cm− 1, which can be explained by the non-harmonic effect induced by acoustic phonon despite coincidence with infrared-active A2u mode, as was concerned in some literature31. Moreover, we also detected another two peaks at 140.5 cm− 1 and 191.0 cm− 1, which are associated with the density of two phonon states of ZrS232. All of these peaks obtained from this study are in good accordance with previous Raman data33.
As shown in Fig. 4a, the ambient Raman spectra pattern of 1T-ZrS2 remained almost the same up to 3.3 GPa, above which some distinct variations were observed in the scattering profile. Several new peaks at 112.6, 134.4, 263.2 and 285.6 cm− 1 started to emerge along with the vanishment of two original bands at 140.5 and 191.0 cm− 1. The occurrences of new vibrational modes hint the structural rearrangement in 1T-ZrS2, from a trigonal phase (P\(\stackrel{-}{3}\)m1) to an orthorhombic phase (Pmm2) as confirmed by our XRD results. In particular, over the pressure range of 0–17.4 GPa, we observed an interesting phenomenon that the intensity of the acoustic phonon at 312.7 cm− 1 gradually increased with pressure, whereas the intensity of the A1g mode at 332.7 cm− 1 nearby the acoustic phonon decreased and finally vanished above 17.4 GPa. This indicates that the non-harmonic effect of out-of-plane (A1g) mode would be greatly enhanced by pressure engineering. Further compression up to 17.4 GPa, multiple new bands along with the disappearance of original peaks were observed corresponding to the II-III transition, reaching agreement with synchrotron XRD data. Moreover, it was noteworthy that beyond 17.4 GPa the Raman spectra profile was extremely lumpy with attenuated Raman peaks and higher frequency uncertainties, which reflected the introduction of structural disordering in phase III. Following decompression experiment found the recovered peaks matches those of metastable phase II. This suggests that the trigonal-to-orthorhombic phase transition is irreversible but we are able to preserve the vdW phase II after removing pressure.
The pressure dependence of Raman modes was plotted in Fig. 4b, and a liner equation was used to fit these mode frequencies. On the basis of the discontinuous changes in dω/dp, two obvious inflection points were located at 5.4 GPa and 17.4 GPa as marked by dashed lines, corresponding to two phase transitions in 1T-ZrS2 under high pressure. The frequencies of phonon modes and their pressure coefficient dω/dp in each phase of ZrS2 is described in the Supplemental material. For the initial trigonal phase (0–5.4 GPa), all the phonon modes shifted to higher wave numbers with increasing pressure, but the dω/dp value of A1g mode was relatively larger than that of the Eg mode. The A1g mode reflects the weak vdW interaction along c axis, while the Eg mode is related to the atomic vibration in the basic X-M-X monolayer along ab plane. As usual, the stretching force constant between the interlayer distances was greatly improved by pressure34, and thus the out-of-plane A1g mode is more sensitive to compression than the in-plane Eg mode. In fact, the divergence in the sensitivity of A and E modes to volume compression has also been reported by MoSe2 and InSe35,36. For high-pressure phases, most phonon modes exhibited positive pressure dependence except the one at 112.6 cm− 1 in phase II and mode of 157.8 cm− 1 in phase III, which were featured by mild phonon softening with negative pressure coefficient of -0.228 cm− 1/GPa and − 0.166 cm− 1/GPa, respectively. This phonon softening is likely interrelated with structural phase transitions in ZrS2. Some other TMDs with similar structure, like ReS237, were also found to exhibit phonon softening under high pressure. In addition, we also calculated mode Grüneisen parameters of A1g and Eg mode using the equation γ = K0/ω0(dω/dP), where K0 and ω0 are the zero-pressure bulk modulus and phonon frequency of Raman modes, respectively. The mode Grüneisen parameter of A1g mode is 0.95, which is larger than that of Eg mode (γ = 0.64). This implies a stronger anharmonicity in crystal structure in the vicinity of A1g vibration38 where acoustic phonon is likely to occur.
Structural phase transition has a profound effect on their electronic properties and we then investigated the electronic properties of ZrS2. The sample was connected by two Pt probes to measure the in situ AC impedance under high pressure, which has been widely applied in the measurement of electrical conductivity of materials39–42. Figure 5a–5c showed the typical impedance spectrum of ZrS2. These collected impedance spectrums were fitted by the equivalent circuit method to obtain the resistance of sample, which was then used to calculate the electrical conductivity with the equation σ = L/RS. It should be mentioned that we only selected the semicircular arcs at low frequency ranges to obtain the grain interior conduction. Figure 5d plotted the pressure dependence of EC during compression and decompression. Close to 1 atm, the electrical conductivity of ZrS2 was about 2.68 (13)×10− 4 S/cm, which is the typical value of semiconductor. The EC value slightly decreased upon pressurization 3.8 GPa, above which the EC jumped due to promoted charge carrier mobility in the orthorhombic phase II. Another electrical conductivity dump was observed at 17.6 GPa, at which the EC increased by roughly four orders of magnitude, indicating the occurrence of the partially disordered phase III. After 26.9 GPa, The EC remained a high value of 0.9 S/cm to the highest pressure studied, corresponding to a metal. The weak pressure dependence provided another evidence to support this metallization. As usual, the increase rate of carrier concentration with pressure would be substantially reduced with the progression of pressure-induced metallization.
We further conducted temperature dependence of EC (Fig. 6) to diagnose the semiconductor-to-metal transition. For typical insulator and semiconductors, the EC exhibits a positive correlation with temperature. In our experiments, the slope of log10(σ)/T approach zero between 23.0 and 30.2 GPa, indicating the pressure range within which metallization occurs. Our observed temperature dependence confirmed the phase II is still a semiconductor and pressure-induced metallization occur in phase III. The four-order-of-magnitude increase in EC during compression makes ZrS2 a promising material for designing optoelectronic switches controlled by pressure.
We finally decompressed the sample to examine the reversibility of this metallic phase. Initially, the EC remained high level, and rapidly dropped after 6.4 GPa. The hysteretic effect of transition pressure point between compression and decompression is a common phenomenon for structural phase transformation. After decompression to 1 atm, the EC value is 3.89 (21)×10− 3 S/cm, which is higher than the original state at ambient condition. This unrecoverable EC value is consistent with our Raman results.
Fig .5. a)–c) Typical impedance spectra for ZrS2 at high pressures in the frequency region of 10‒1–107 Hz during pressurization. The horizontal axe indicates the real part of the complex impedance, while vertical axe represents the imaginary part. d) The logarithm of electrical conductivity as a function of pressure in the process of compression and decompression.
Fig .6 a) The temperature dependence of the electrical conductivity (EC) of ZrS2 at various pressure points. b) The metallization at 30.2 GPa and 45.8 GPa.