The effect of pressure on Raman scattering (RS) in bulk H f S 2 is investigated under hydrostatic and non-hydrostatic pressure conditions. The RS line shape does not change significantly in the hydrostatic regime up to P = 9.6 GPa, showing a systematic blueshift of the spectral features. In a non-hydrostatic environment, seven peaks appear in the spectrum at P = 7 GPa, which dominate the RS line shape up to P = 10.5 GPa. The change in the RS line shape manifests a pressure-induced phase transition in H f S 2. The simultaneous observation of both low-pressure (LP) and high-pressure (HP) related RS peaks suggests the coexistence of two different phases over a wide pressure range. It is found that the HP-related phase is metastable and persists during the decompression cycle down to P = 1.2 GPa, while the LP-related features eventually recover at even lower pressure. The angle-resolved polarized RS performed under P = 7.4 GPa revealed a strong in-plane anisotropy of both the LP-related A 1 g mode and the HP peaks. The anisotropy is related to the possible distortion of the structure induced by the non-hydrostatic component of the pressure. The results are explained in terms of a distorted P n m a phase as a possible pressure-induced phase of H f S 2.

Transition metal dichalcogenides (TMDs) have emerged as an exciting class of materials with a layered van der Waals (vdW) structure. Their unique properties, which are strongly dependent on the thickness of the structure and the relatively easy exfoliation, have recently become a hot topic in materials research. Moreover, their layered structure makes them very sensitive to the interlayer spacing, which can be modulated by temperature or strain. This explains the interest in strain engineering of TMDs as an effective way to modify their properties.1–4 In this study, we address the properties of hafnium disulfide ( H f S 2), a less studied member of the TMD family. Interest in this material is driven by its promising electrical properties. According to theoretical calculations, the room-temperature electron mobility in H f S 2 is much higher than in the most studied MoS 2.5 Field-effect transistors based on few-layer H f S 2, fabricated by Kanazawa et al.,6 exhibit high drain currents and mobility, a transistor on/off current ratio greater than 10 4, and good responsivity ( 1.6  μ A W 1) in a transfer-free photodetector,7 making them notable for the development of thermoelectric and optoelectronic devices.8 

The properties of H f S 2 are highly susceptible to changes in the temperature and pressure, including pressure-induced phase transitions.9–13 However, there is an ongoing debate regarding the precise pressure at which the pressure-induced phase transitions occur, whether they are reversible upon pressure release, and what is the crystallographic structure of the pressure-induced phase. Under ambient conditions, H f S 2 belongs to the space group P 3 ¯ m 1 (No. 164). The onset of the phase transition was reported at P 11 GPa based on Raman scattering (RS) measurements in Ref. 10. The RS spectroscopy of H f S 2 permitted the earlier reporting on two phase transitions,11 which were observed at P 9.8 GPa and P 15.2 GPa. The distinctive feature of the study in which silicone oil was used as the pressure transmitting medium (PTM) was the emergence of several new high-pressure (HP) peaks in the RS spectra above the phase transformation. A reversible transformation to the 3D orthorhombic I m m m (No. 71) structure accompanied by semiconductor-to-semiconductor transition was reported in Ref. 14 at approximately 12 GPa of pressure, which was mediated by neon gas as the PTM. Notably, the authors observed that the RS measurements reflected the transition under pressure as low as 9.2 GPa, which corresponds to other reports.15,16 Hong et al.15 reported two structural phase transitions at P 8.0 GPa and P 15.2 GPa as well as metallization at P = 20.5 GPa using helium as the PTM. They also showed that upon decompression, the characteristic RS peaks, which emerged above the transformation, persisted down to P 1.4 GPa, indicating a significant pressure hysteresis effect for H f S 2. Quite recently, Zhang et al.16 have also reported two structural transitions at approximately P = 11.0 GPa and P = 35.5 GPa using silicone oil as the PTM. The former transformation was related to the transition to the orthorhombic P n m a (No. 62) phase, while the latter was linked to a structural transition from the P n m a phase to the tetragonal I 4 / m m m (No. 139) phase. The pressure-induced transition from the P 3 ¯ m 1 to P n m a phase under pressure was also corroborated by first-principles calculations.17 

The motivation of our work is to provide more information on the pressure-induced structural changes of H f S 2 upon compression by probing a local symmetry of the related Raman-active vibrations. To this end, we report on polarization-sensitive room-temperature RS measurements as a function of pressure up to P = 10.5 GPa. Our focus is on the effect of pressure hydrostaticity on the phase transition, with particular examination of the results of experiments performed with two PTMs. No apparent phase transition is observed when using the 4:1 methanol–ethanol (ME) mixture, which ensures hydrostatic conditions of pressure within the investigated pressure range. In contrast, a substantial change in the RS spectrum of H f S 2 was observed at approximately P = 7 GPa with silicone oil as the PTM. Similarly to the observation by Hong et al.,15 a large hysteresis effect was observed, and the phase transition is shown to be reversible upon full decompression. The angle-resolved polarized RS (ARPRS) measurements performed at P = 7.4 GPa revealed a strong in-plane anisotropy of the observed RS modes. We discuss the possible origin of the anisotropy, emphasizing the crucial effect of non-hydrostaticity on the phase transitions in H f S 2 under external pressure. The results of our study are explained in terms of a distorted P n m a phase as a possible pressure-induced phase of H f S 2.

There are six vibrational modes present in the total representation18 under ambient conditions in H f S 2 with the 1T structure,
(1)
Only the even modes A 1 g and E g are Raman active. The polarization dependence of the phonon mode intensities is an important factor for the analysis. The probed co- (XX) and cross-linear (XY) configurations correspond to the parallel and perpendicular polarization of the scattered light with respect to the polarization of the incident laser, respectively. The angle coordinates describe the in-plane spatial orientation of the laser polarization with respect to the H f S 2 layers. The polarization-sensitive RS spectra of H f S 2 measured in XX and XY polarization configurations can be appreciated in Fig. 1. The measurements were conducted with λ = 633 nm excitation [details of the optical setup can be found in Sec. V and in Fig. S1 in the supplementary material]. Four peaks were observed at 137, 262, 324, and 341  cm 1. These were referred to as ω 1, E g, A 2 u(LO), and A 1 g, respectively, and will be referred to as low-pressure (LP) modes. This is in agreement with our previous report.11 In addition to the even modes, two further features are clearly visible10,19–21 in the figure. The H f S 2 phonon dispersion allows us to propose that they are attributed to 2TA(M) ( ω 1) and A 2 u(LO) modes.
FIG. 1.

Raman scattering spectra of H f S 2 measured in the XX (red) and XY (black) configurations at ambient pressure. The most intense vibrational modes, A 1 g and E g, are depicted in the inset.

FIG. 1.

Raman scattering spectra of H f S 2 measured in the XX (red) and XY (black) configurations at ambient pressure. The most intense vibrational modes, A 1 g and E g, are depicted in the inset.

Close modal

The results of the measurements show that the A 1 g intensity is independent of the relative polarization of the incoming and scattered light for the XX configuration and practically vanishes for the XY configuration (although some signal can still be observed, which might be due to light traveling slightly out of the perpendicular direction). On the contrary, the E g intensity in both XX and XY configurations is practically independent of the polarization of the incoming light. The ARPRS polar plot of A 1 g and E g mode intensities at ambient pressure is shown in Fig. 2. The full and open circles represent the results of measurements for the XX and XY configurations, respectively.

FIG. 2.

The ARPRS intensity plots for the A 1 g (a) and E g (b) modes in both XX (colorful) and XY (black) configuration at ambient pressure. The solid lines represent the fittings.

FIG. 2.

The ARPRS intensity plots for the A 1 g (a) and E g (b) modes in both XX (colorful) and XY (black) configuration at ambient pressure. The solid lines represent the fittings.

Close modal

The evolution of RS spectra with pressure transmitted by the ME mixture, which generally provides hydrostatic conditions up to approximately P = 10 GPa,22 is presented in Fig. 3. The line shape of the RS spectra does not change significantly with pressure up to P = 9.6 GPa with a systematic blueshift of the observed RS peak energies. In particular, the A 1 g peak energy is characterized by a linear dependence throughout the entire range of applied pressures (see Fig. S2 in the supplementary material).

FIG. 3.

The evolution of Raman scattering spectra in H f S 2 during compression with ME PTM.

FIG. 3.

The evolution of Raman scattering spectra in H f S 2 during compression with ME PTM.

Close modal

The evolution of the RS spectra during compression with silicone oil as the PTM up to P = 10.5 GPa is shown in Fig. 4(a). The RS spectrum measured under P = 0.2 GPa, in the experiment with silicone oil PTM, is similar to that obtained under ambient conditions [compare Fig. 4(a) with Fig. 1]. The effect of increasing pressure up to P = 6.5 GPa is primarily manifested by the blueshift of the LP modes and slight changes in their relative intensities. In contrast, a substantial change in the RS line shape is observed at P = 7.4 GPa, with two well-resolved peaks: I (89  cm 1) and II (100  cm 1) emerging in the low-frequency range of the spectrum. The modification of the RS spectra is followed by the emergence of additional peaks: III (142  cm 1), IV (151  cm 1), V (291  cm 1), VI (324  cm 1), and VII (352  cm 1). The I–VII peaks will, henceforth, be referred to as HP modes. It is evident from Fig. 4(a) that under pressure exceeding P = 7.4 GPa, all HP peaks exhibit a notable increase in the intensity in comparison to their LP counterparts. With increasing pressure, the A 1 g and E g modes continuously disappear from the spectra, becoming barely discernible at P = 10.5 GPa and P = 8.1 GPa, respectively.

FIG. 4.

The evolution of Raman scattering spectra in H f S 2 during compression (a) and decompression cycle (b) with silicone oil as the PTM.

FIG. 4.

The evolution of Raman scattering spectra in H f S 2 during compression (a) and decompression cycle (b) with silicone oil as the PTM.

Close modal

To assess the reversibility of the structural transformation of H f S 2 under pressure, the RS spectra were also recorded during the decompression cycle. The corresponding selected RS spectra are displayed in Fig. 4(b). The HP RS modes can be observed in the spectra under pressure as low as P = 1.2 GPa [see the black line in Fig. 4(b)] at which A 1 g and E g recover. Although the A 1 g mode emerges as a separate feature at 350  cm 1, the quenching of the V peak is accompanied by the reemergence of the E g feature, as manifested by a change in the broadening of the peak at 270  cm 1. The energies of the RS peaks measured under the compression (full circles) and decompression (open squares) cycles are shown in Fig. 5. An inspection of the pressure evolution of the A 1 g energy reveals a slope change at approximately P = 5 GPa. An apparently non-monotonic evolution of peaks V and VII is observed at 280 and 340  cm 1 in the pressure range of 7.3 GPa < P < 8.9 GPa (for the determination of their energies, see Fig. S4 in the supplementary material). This behavior cannot be observed during decompression, as the energies of both peaks follow a monotonic dependence down to P = 1.2 GPa. In contrast, the evolution of I, II, IV, and VI during decompression reflects those of the compression procedure down to P = 1.2 GPa.

FIG. 5.

The pressure evolution of the Raman shifts obtained for the observed phonon modes. The colored spheres and open squares represent the results obtained for the compression and decompression processes, respectively. The solid blue lines represent linear fits to data collected during the decompression cycle.

FIG. 5.

The pressure evolution of the Raman shifts obtained for the observed phonon modes. The colored spheres and open squares represent the results obtained for the compression and decompression processes, respectively. The solid blue lines represent linear fits to data collected during the decompression cycle.

Close modal
Except for the non-monotonic behavior of peaks VII and V, the energies of the Raman modes can be approximated with a linear dependence,
(2)
where E ( 0 ) represents the zero-field frequency and α is the pressure coefficient. The fitting parameters for the observed modes are presented in Table I. It should be noted that under a silicone oil environment, A 1 g shows an inflection, the parameters calculated to be P < 5.5 GPa and P > 5.5 GPa are shown in blue and red, respectively. Similarly, during the decompression cycle, the HP peaks’ parameters were calculated to be P > 2.0 GPa (orange). The bulk modulus B 0 = 30.6 GPa was obtained from DFT calculations10 and the Grüneisen parameters through the expression γ i = B 0 α i / E i ( 0 ). As a non-hydrostatic condition is set to P > 5.5 GPa, γ i was calculated only to LP modes.
TABLE I.

Pressure-dependent parameters of the HfS2 phonon modes calculated for both PTM ME and silicone oil. The parameters E(0) and α represent the experimental zero-field frequencies and their pressure coefficients, respectively. Under a silicone oil environment, A1g shows an inflection, the parameters calculated to P < 5.5 GPa and P > 5.5 GPa are shown in blue and red, respectively. During the decompression cycle, the HP peak parameters were calculated to P > 2.0 GPa (orange).

MESilicone oil
CompressionCompressionDecompression
E(0)αE(0)αE(0)α
(cm−1)(cm−1/GPa)γ(cm−1)(cm−1/GPa)γ(cm−1)(cm−1/GPa)
ω 1 … … … 136.3 1.36 0.31 … … 
Eg 260.6 2.35 0.28 261.5 1.92 0.23 … … 
A2u(LO) … … … 322.4 2.13 0.20 … … 
A1g 341.8 4.11 0.37 340.7 3.84 0.35 340.9 3.84 
    347.8 2.33 … 351.7 2.04 
… … … 92 −0.44 … 88.3 −0.07 
II … … … 93.0 0.82 … 92.9 0.90 
III … … … 141.1 0.21 … 137.2 0.69 
IV … … … 144.3 0.87 … 145.7 0.85 
… … … … … … 268.5 1.70 
VI … … … … … … 310.4 1.50 
VII … … … … … … 330.1 2.01 
MESilicone oil
CompressionCompressionDecompression
E(0)αE(0)αE(0)α
(cm−1)(cm−1/GPa)γ(cm−1)(cm−1/GPa)γ(cm−1)(cm−1/GPa)
ω 1 … … … 136.3 1.36 0.31 … … 
Eg 260.6 2.35 0.28 261.5 1.92 0.23 … … 
A2u(LO) … … … 322.4 2.13 0.20 … … 
A1g 341.8 4.11 0.37 340.7 3.84 0.35 340.9 3.84 
    347.8 2.33 … 351.7 2.04 
… … … 92 −0.44 … 88.3 −0.07 
II … … … 93.0 0.82 … 92.9 0.90 
III … … … 141.1 0.21 … 137.2 0.69 
IV … … … 144.3 0.87 … 145.7 0.85 
… … … … … … 268.5 1.70 
VI … … … … … … 310.4 1.50 
VII … … … … … … 330.1 2.01 

To gain further insight into the H f S 2 structure under pressure with silicone oil as the PTM, the polarization-related properties of the observed RS modes were investigated under P = 7.4 GPa during the compression cycle and at P = 1.2 GPa during the decompression cycle. Let us first address the polarization dependence of the A 1 g intensity, which is still present in the RS spectrum at P = 7.4 GPa [see Fig. 6(a)]. A fourfold symmetry of the A 1 g intensity is observed for both the XX and XY configurations. The main axis of the A 1 g intensity for the XY configuration is rotated by 45 ° with respect to the main axis for the XX configuration. It should be noted that only minor deviations from the circular dependence of the A 1 g intensity can be observed at P = 1.2 GPa in the XX configuration. For the XY configuration, the A 1 g mode has a significantly lower intensity with a negligible angle dependence.

FIG. 6.

Polar plots for the A 1 g mode intensity in both the XX (red full circles) and XY (black open circles) configurations under excitation λ = 633 nm, detected upon compression at P = 7.4 GPa (a) and decompression at P = 1.2 GPa (b).

FIG. 6.

Polar plots for the A 1 g mode intensity in both the XX (red full circles) and XY (black open circles) configurations under excitation λ = 633 nm, detected upon compression at P = 7.4 GPa (a) and decompression at P = 1.2 GPa (b).

Close modal
The phonon mode intensity in the RS spectrum, calculated within the Placzek approximation, can be expressed as23–25 
(3)
where I 0 reflects experimental parameters such as laser intensity and integration time. The e ^ i and e ^ s are the electric field vectors of the incident and scattered light, respectively, and R represents the Raman tensor for different symmetry modes.

Let us restrict our discussion to light traveling along the z axis, perpendicular to the crystal planes, and adopt the counterclockwise chirality in our setup. In this case, the XX configuration sets the electric field of the incident light ( e ^ i) and the scattered beam ( e ^ s) as e ^ i = e ^ s = (cos θ, sin θ). The XY configuration is achieved by rotating the detection polarizer by 90 °, resulting in e ^ s = ( sin θ, cos θ).

In the case of a symmetric A 1 g out-of-plane mode of H f S 2 at ambient conditions, the Raman tensor is diagonal with equal matrix elements and reads
(4)
where a represents the absolute value of the tensor element. The tensor leads to a symmetric polar dependence of the RS in the XX configuration and to the quenching of the mode in the XY configuration, which corresponds to the results obtained under ambient conditions for A 1 g (see Fig. 2). It can be observed that the A 1 g intensity does not depend on the polarization direction in the XX configuration and practically vanishes in the XY configuration (although some signal can still be observed, which may be due to light traveling slightly out of the perpendicular direction). In contrast, the E g intensity is practically unaffected by the relative polarization of the incoming and scattered light.
A greater challenge is posed by the polar dependence of the RS feature, which evolved from the “ambient-condition” A 1 g mode at 7.4 GPa [Fig. 6(a)]. As can be seen in Fig. 6, the polar dependence of the A 1 g mode for both XX and XY configurations exhibits a fourfold symmetry. This clearly demonstrates that the polar evolution of the RS intensity cannot be described by the Raman tensor presented in Eq. (4). To account for the results shown in Fig. 6, we consider a more general form of the Raman tensor,26 which corresponds to A / A 1 / A g Raman-active modes in a crystal of the orthorhombic symmetry.27 The tensor can be described as
(5)
In the following, the lowercase letter will be used to represent the absolute value of the Raman tensor element, and φ i will be employed to indicate the related complex phase due to absorption effects. It should be noted that the complex nature of the Raman tensor elements is proposed to reflect the optical anisotropy of a crystal.28 As previously discussed in Ref. 27, we note that it is only possible to determine the relative values of b / a and φ b a = φ b φ a from experimental results. The polar dependence of the RS intensity expected for the Raman tensor for A / A 1 / A g modes in a crystal of orthorhombic symmetry in the XX and XY configurations can be found in Eqs. (S9)–(S11) in the supplementary material. As illustrated in Fig. 6, the results of the fit with b/ a = 0.94 and φ b a = 103 ° correspond closely to the experimental results. In our opinion, this strongly suggests that the symmetry of the H f S 2 crystal changes upon the compression cycle. A comparison of the results of experiments conducted with two different PTM suggests that the change in the H f S 2 symmetry is related to the non-hydrostaticity provided by the silicone oil. The non-hydrostatic component of pressure is confirmed by the increasing splitting between the R 1 and R 2 ruby-related luminescence lines [see Fig. S2(b) in the supplementary material]. The component affects the symmetry of the crystal and the observed Raman-active out-of-plane vibrations. As a result, the crystal structure becomes orthorhombic (with three crystallographic axes perpendicular to each other that have unequal length). The principal Raman-active out-of-plane vibration of the A 1 g symmetry in the 1T- H f S 2 crystal under ambient conditions becomes the A / A 1 / A g vibration, depending on the actual distortion of the crystal. However, for the sake of clarity, we will refer to the mode as A 1 g also above the transformation.
The polarization-dependent evolution of the HP RS modes can be observed in Fig. 7. The polar dependence for the vibrations obtained from Eq. (5) exhibits a fourfold symmetry in the XY configuration, which is the case for II (i), III (g), and IV (e) (fitting parameters can be found in Table II). Other modes exhibit a more complex dependence. To account for them, one must consider the Raman tensor, which corresponds to an A g mode in a triclinic crystal,27 which is given by
(6)
As previously stated, only the relative values can be determined from the experimental results, e.g., b / a and φ b a = φ b- φ a. In Fig. 7, the Raman tensor shown in Eq. (6) describes modes I (f), V (b), and VII (a) (see Table II for a summary of the fitting parameters).
FIG. 7.

Polar plots for the HP Raman scattering mode intensities: VII (a), V (b), IV (c), III (d), II (e), and I (f) at P = 7.4 GPa. The solid lines represent the fittings, and the full and open circles represent the XX and XY configurations, respectively.

FIG. 7.

Polar plots for the HP Raman scattering mode intensities: VII (a), V (b), IV (c), III (d), II (e), and I (f) at P = 7.4 GPa. The solid lines represent the fittings, and the full and open circles represent the XX and XY configurations, respectively.

Close modal
TABLE II.

Results of the fits of the polar intensity dependence of RS peaks measured under P = 7.4 GPa.

Peakb/ad/aφbaφdaMain axis
A1g 0.94 …  103 ° …  85 ° 
VII 0.67 0.12  270 °  285 °  160 ° 
0.27 0.35  348 ° 103 ° 170 ° 
IV 0.91 … 93 ° … 175 ° 
III 0.97 … 90 ° … 165 ° 
II 0.66 … 143 ° … 65 ° 
0.77 0.05 128 ° 100 ° 35 ° 
Peakb/ad/aφbaφdaMain axis
A1g 0.94 …  103 ° …  85 ° 
VII 0.67 0.12  270 °  285 °  160 ° 
0.27 0.35  348 ° 103 ° 170 ° 
IV 0.91 … 93 ° … 175 ° 
III 0.97 … 90 ° … 165 ° 
II 0.66 … 143 ° … 65 ° 
0.77 0.05 128 ° 100 ° 35 ° 

The analysis of our observations begins with an examination of the emergence of the anisotropic components of the Raman tensor. In our opinion, they must result from the anisotropy induced by the PTM used in our experiment. The silicone oil is known to introduce a non-hydrostatic component to the pressure. The anisotropy is removed during the decompression cycle at approximately P = 1.2 GPa, as clearly demonstrated in Fig. 6(b). The use of silicone oil in pressure experiments strongly affects the pressure value at which the phase transition occurs. We also found that with the use of the ME, the phase transition is not apparent up to P = 9.6 GPa (Fig. 3). The non-hydrostatic component of pressure clearly leads to the distortion of the 1T- H f S 2 crystallographic structure, as observed by the anisotropic RS spectrum of the A 1 g mode obtained at P = 7.4 GPa (Fig. 6). The Raman tensor, which can be used to interpret the results, suggests that H f S 2 has an orthorhombic crystal structure of the distorted. Similar anisotropy has been reported in a series of different layered materials, e.g., WTe 2,29 black phosphorus,30 black arsenic,31, α MoO 3,32, ReS 2 ,33 or GeS.34 

The Raman tensors, which describe the polar dependence of the II, III, and IV HP peaks, correspond to the orthorhombic structure, while those describing the I, V, and VII HP peaks correspond to the triclinic structure. We cannot offer a firm explanation for this observation. However, the energy coincidence of the III and ω 1 peaks may suggest that peaks III and IV are split features, which evolve from the ω 1 vibration in the unaffected H f S 2 crystal. Similarly, peak II may be attributed to another mode that is not optically active in the original H f S 2 crystal and became visible in the distorted structure. The fit for peak I is not very conclusive, as the off-diagonal element ( d / a) is much smaller than the diagonal ( b / a) one. Additionally, the intensity for peak VI was insufficient for the comprehensive analysis. This would leave peaks V and VII to plausibly attributed to a distorted orthorhombic symmetry of the HP phase of H f S 2. We propose that this is a distorted P n m a phase.

The orthorhombic P n m a phase has recently been proposed as an HP phase of H f S 2.17 There are 18 Raman-active modes in the P n m a phase of H f S 2,
(7)
Only A g and B 1 g are visible in the back-scattering experiment. Therefore, nine Raman-active modes for the P n m a phase are expected in the respective RS spectrum. The phonon dispersion in P n m a H f S 2 was calculated using density functional theory (DFT)—see Fig. 8(a). To better visualize the Raman-active modes at the Γ point of the Brillouin zone (BZ), the relative intensities of the modes are shown in Fig. 8.
FIG. 8.

(a) The phonon dispersion for the orthorhombic H f S 2 P n m a phase, calculated using LDA functional and (b) the corresponding relative intensities of the RS peaks. The gray stripes indicate the energy range, in which peaks V and VII were observed in the experiment.

FIG. 8.

(a) The phonon dispersion for the orthorhombic H f S 2 P n m a phase, calculated using LDA functional and (b) the corresponding relative intensities of the RS peaks. The gray stripes indicate the energy range, in which peaks V and VII were observed in the experiment.

Close modal

There are two strong A g modes with energies at 269 and 324  cm 1, while the other A g and B 1 g modes are much weaker. As illustrated in Table I, their energies coincide with the extrapolated zero-pressure energies of peaks V and VII. We note that the polarization-resolved dependence of the intensities of peaks V and VII corresponds well to the triclinic crystal symmetry, which may result from the pressure-induced distortion of the orthorhombic structure. Quantitative correspondence with the experimental results cannot be expected, as our calculations are made for the ambient pressure at T = 0 K. Moreover, the local density approximation (LDA) functional is used, which is known to underestimate the lattice constants that affect the frequencies and intensities of the modes. Additionally, possible resonant conditions are also neglected. Nevertheless, the observed correspondence strongly suggests that the HP phase should be attributed to the distorted P n m a phase of H f S 2.

The scattering of the bulk H f S 2 was investigated using the λ = 633 nm excitation at room temperature under external pressure up to P = 10.5 GP applied under hydrostatic and non-hydrostatic conditions, which were provided by ME and silicone oil, respectively. No significant RS spectrum line shape was observed with a systematic blueshift of the RS features under hydrostatic conditions. A phase transition in H f S 2 was observed at P 7 GP when a non-hydrostatic environment was provided. The phase transition manifested itself with the appearance of seven new modes in the RS spectrum. The angle- and polarization-dependent RS spectroscopic study conducted at P = 7.4 GPa clearly demonstrated anisotropic behavior for all observed RS modes. It was proposed that the original ambient-pressure 1T- H f S 2 phase was distorted under pressure, leading to the orthorhombic crystal structure. Concurrently, a new phase emerged, which was associated with a distorted P n m a phase. The pressure-induced phase coexisted with the ambient-condition phase up to P = 10.5 GPa and persisted during the decompression down to P = 1.2 GPa, which confirms its metastable behavior.

The H f S 2 crystals under investigation were obtained from 2D semiconductors. The iodine-related photoluminescence signal12 was observed in the 2D semiconductor crystal. However, in our opinion, the presence of iodine has no significant effect within the pressure range investigated. The x-ray diffraction measurements were found to be in perfect agreement with those of the octahedral 1T phase of H f S 2,18 although a minor amount of an unidentified phase was also detected.35 

Room-temperature optical measurements were conducted by exciting the sample with a He–Ne laser at λ = 633 nm (1.96 eV), which was focused by a 50 × long-working distance objective with 0.55 (NA). The excitation power focused on the sample was maintained at approximately 0.5 mW throughout all measurements, which resulted in temperature of the sample determined from the relative intensity of the anti-Stokes to the Stokes A 1 g modes equal to T = 291 K (see Fig. S4 in the supplementary material). The scattered light was collected in the back-scattering geometry, sent through a 0.75 m monochromator, and then detected by means of a liquid nitrogen-cooled CCD camera.

The polarization-sensitive spectroscopy was carried out by introducing a λ / 2 plate directly on top of the objective lens in order to simultaneously rotate the incoming and scattered light with reference to the investigated in-plane crystal orientation. The experiments were conducted with the incoming and scattered light traveling along the z axis. Both configurations, i.e., co- (XX) and cross-linear (XY), were probed. These correspond to the detection polarizer being aligned parallel or perpendicular to the excitation one, respectively. Further details and a schematic illustration of the experimental setup can be found in Fig. S1 of the supplementary material.

The high-pressure experiments were conducted at room temperature in a non-magnetic diamond anvil cell (SymmDAC 60 LT DAC) manufactured from the BeCu alloy.36 The pressure chamber was sealed using a BeCu gasket with a height of approximately. 75  μm, a diameter of 250  μm, and filled with either a methanol/ethanol mixture or silicone oil. The pressure was monitored by the ruby ( Al 2 O 3: Cr + 3) fluorescent method. The R 1 R 2-separation is used as a metric for uniaxial stress.37 

Density functional theory (DFT) calculations were performed in Quantum Espresso38,39 with scalar-relativistic norm-conserving pseudopotentials.40 The local density approximation (LDA) to the exchange-correlation functional was used. The wavefunction and charge density cutoff energies were set to 100 and 400 Ry, respectively. A 6 × 10 × 4 Γ-centered Monkhorst–Pack k-point grid was used. The geometrical structure was fully optimized using 10 5 Ry/Bohr and 0.01 kbar criteria for the interatomic forces and stress tensor components, respectively. Phonon and non-resonant Raman calculations were performed within the framework of density functional perturbation theory. A 2 × 2 × 2 q-point grid was found to be sufficient for the convergence of phonon dispersion and Raman intensities.

See the supplementary material for a schematic illustration of the experimental setup, fundamental pressure features, and details related to the fitting procedure and parameter tables.

The work was supported by the “Excellence Initiative—Research University at the University of Warsaw” program. Z.M. and W.Z. acknowledge the support from the National Natural Science Foundation of China (NNSFC) (Grant No. 62150410438), the International Collaboration Project (No. B16001), and the Beihang Hefei Innovation Research Institute (Project No. BHKX-19-02). This research was carried out with the support of the Interdisciplinary Centre for Mathematical and Computational Modelling, University of Warsaw, under computational Allocation No. G95-1773.

The authors have no conflicts to disclose.

Igor Antoniazzi: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Project administration (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Tomasz Woźniak: Data curation (equal); Formal analysis (equal); Software (equal); Writing – review & editing (equal). Amit Pawbake: Data curation (equal); Formal analysis (equal); Methodology (equal); Writing – review & editing (supporting). Natalia Zawadzka: Data curation (equal); Investigation (equal); Methodology (equal); Writing – review & editing (supporting). Magdalena Grzeszczyk: Data curation (supporting); Formal analysis (supporting); Writing – review & editing (supporting). Zahir Muhammad: Conceptualization (equal); Methodology (equal); Writing – review & editing (supporting). Weisheng Zhao: Methodology (equal); Writing – review & editing (supporting). Jordi Ibáñez: Conceptualization (equal); Writing – review & editing (equal). Clement Faugeras: Data curation (equal); Formal analysis (equal); Methodology (equal); Writing – review & editing (supporting). Maciej R. Molas: Conceptualization (lead); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Resources (equal); Writing – original draft (equal); Writing – review & editing (equal). Adam Babiński: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
S.
Deng
,
A. V.
Sumant
, and
V.
Berry
,
Nano Today
22
,
14
(
2018
).
2.
L. G.
Pimenta Martins
,
R.
Comin
,
M. J. S.
Matos
,
M. S. C.
Mazzoni
,
B. R. A.
Neves
, and
M.
Yankowitz
,
Appl. Phys. Rev.
9
,
011313
(
2023
).
3.
S.
Pei
,
Z.
Wang
, and
J.
Xia
,
Mater. Des.
213
,
110363
(
2022
).
4.
L.
Zhang
,
Y.
Tang
,
A. R.
Khan
,
M. M.
Hasan
,
P.
Wang
,
H.
Yan
,
T.
Yildirim
,
J. F.
Torres
,
G. P.
Neupane
,
Y.
Zhang
,
Q.
Li
, and
Y.
Lu
,
Adv. Sci.
7
,
2002697
(
2020
).
5.
W.
Zhang
,
Z.
Huang
,
W.
Zhang
, and
Y.
Li
,
Nano Res.
7
,
1731
(
2014
).
6.
T.
Kanazawa
,
T.
Amemiya
,
A.
Ishikawa
,
V.
Upadhyaya
,
K.
Tsuruta
,
T.
Tanaka
, and
Y.
Miyamoto
,
Sci. Rep.
6
,
22277
(
2016
).
7.
J.
Singh
,
J. H.
Shao
,
G. T.
Chen
,
H. S.
Wu
, and
M. L.
Tsai
,
Nanoscale Adv.
5
,
171
(
2022
).
8.
D.
Singh
and
R.
Ahuja
,
ACS Appl. Energy Mater.
2
,
6891
(
2019
).
9.
J.
Peng
,
S.
Najmaei
,
M.
Dubey
, and
P. W.
Chung
,
Mater. Today Commun.
26
,
101722
(
2021
).
10.
J.
Ibáñez
,
T.
Woźniak
,
F.
Dybala
,
R.
Oliva
,
S.
Hernández
, and
R.
Kudrawiec
,
Sci. Rep.
8
,
1
(
2018
).
11.
M.
Grzeszczyk
,
J.
Gawraczyński
,
T.
Woźniak
,
J.
Ibáñez
,
Z.
Muhammad
,
W.
Zhao
,
M.
Molas
, and
A.
Babiński
,
Acta Phys. Pol. A
141
,
95
(
2022
).
12.
N.
Zawadzka
,
T.
Woźniak
,
M.
Strawski
,
I.
Antoniazzi
,
M.
Grzeszczyk
,
K.
Olkowska-Pucko
,
Z.
Muhammad
,
J.
Ibanez
,
W.
Zhao
,
J.
Jadczak
,
R.
Stȩpniewski
,
A.
Babiński
, and
M. R.
Molas
,
Appl. Phys. Lett.
122
,
042102
(
2023
).
13.
I.
Antoniazzi
,
N.
Zawadzka
,
M.
Grzeszczyk
,
T.
Woźniak
,
J.
Ibáñez
,
Z.
Muhammad
,
W.
Zhao
,
M. R.
Molas
, and
A.
Babiński
,
J. Phys.: Condens. Matter
35
,
3055401
(
2023
).
14.
W.
Zhong
,
W.
Deng
,
F.
Hong
, and
B.
Yue
,
Phys. Rev. B
107
,
1
(
2023
).
15.
M.
Hong
,
L.
Dai
,
H.
Hu
,
X.
Zhang
,
C.
Li
, and
Y.
He
,
J. Mater. Chem. C
10
,
10541
(
2022
).
16.
S.
Zhang
,
H.
Wang
,
H.
Liu
,
J.
Zhen
,
S.
Wan
,
W.
Deng
,
Y.
Han
,
B.
Chen
, and
C.
Gao
,
Phys. Rev. Mater.
7
,
104802
(
2023
).
17.
R.
Zhang
,
Y.
Fang
,
F.
Wu
,
X.
Cao
,
Y.
Zhou
,
Z.-Z.
Zhu
, and
S.
Wu
,
J. Phys. D: Appl. Phys.
55
,
295304
(
2022
).
18.
G.
Lucovsky
,
R. M.
White
,
J. A.
Benda
, and
J. F.
Revelli
,
Phys. Rev. B
7
,
3859
(
1973
).
19.
T.
Iwasaki
,
N.
Kuroda
, and
N.
Y
,
J. Phys. Soc. Jpn.
52
,
2233
(
1982
).
20.
A.
Cingolani
,
M.
Lugara
,
G.
Scamarcio
, and
F.
Lévy
,
Solid State Commun.
62
,
121
(
1987
).
21.
L.
Roubi
and
C.
Carlone
,
Phys. Rev. B
37
,
6808
(
1988
).
22.
S.
Klotz
,
J.-C.
Chervin
,
P.
Munsch
, and
G. L.
Marchand
,
J. Phys. D: Appl. Phys.
42
,
075413
(
2009
).
23.
D.
Porezag
and
M. R.
Pederson
,
Phys. Rev. B
54
,
7830
(
1996
).
24.
B. R.
Carvalho
,
L. M.
Malard
,
J. M.
Alves
,
C.
Fantini
, and
M. A.
Pimenta
,
Phys. Rev. Lett.
114
,
136403
(
2015
).
25.
J. A.
Steele
,
P.
Puech
, and
R. A.
Lewis
,
J. Appl. Phys.
120
,
055701
(
2016
).
26.
M.
Cardona
and
G.
Guntherodt
,
Top. Appl. Phys.
50
,
45–50
(
1982
).
27.
M. A.
Pimenta
,
G. C.
Resende
,
H. B.
Ribeiro
, and
B. R.
Carvalho
,
Phys. Chem. Chem. Phys.
23
,
27103
(
2021
).
28.
C.
Kranert
,
C.
Sturm
,
R.
Schmidt-Grund
, and
M.
Grundmann
,
Phys. Rev. Lett.
116
,
127401
(
2016
).
29.
Q.
Song
,
X.
Pan
,
H.
Wang
et al.,
Sci. Rep.
6
,
29254
(
2016
).
30.
H. B.
Ribeiro
,
M. A.
Pimenta
,
C. J. D.
Matos
,
R. L.
Moreira
,
A. S.
Rodin
,
J. D.
Zapata
,
E. A. D.
Souza
, and
A. H. C.
Neto
,
ACS Nano
9
,
4270
(
2015
).
31.
Y.
Wang
,
F.
Chen
,
X.
Guo
,
J.
Liu
,
J.
Jiang
,
X.
Zheng
,
Z.
Wang
,
M. M.
Al-Makeen
,
F.
Ouyang
,
Q.
Xia
, and
H.
Huang
,
J. Phys. Chem. Lett.
12
,
10753
(
2021
).
32.
Y.
Gong
,
Y.
Zhao
,
Z.
Zhou
,
D.
Li
,
H.
Mao
,
Q.
Bao
,
Y.
Zhang
, and
G. P.
Wang
,
Adv. Opt. Mater.
10
,
2200038
(
2022
).
33.
D. A.
Chenet
,
B.
Aslan
,
P. Y.
Huang
,
C.
Fan
,
A. M.
van der Zande
,
T. F.
Heinz
, and
J. C.
Hone
,
Nano Lett.
15
,
5667
(
2015
).
34.
N.
Zawadzka
,
Ł.
Kipczak
,
T.
Woźniak
,
K.
Olkowska-Pucko
,
M.
Grzeszczyk
,
A.
Babiński
, and
M. R.
Molas
,
Nanomaterials
11
,
3109
(
2021
).
35.
J.
Ibáñez
, private information (
2022
).
36.
D. E.
Graf
,
R. L.
Stillwell
,
K. M.
Purcell
, and
S. W.
Tozer
,
High Pressure Res.
31
,
533
(
2011
).
37.
T.
Kenichi
,
J. Phys. Soc. Jpn.
76
,
202
(
2007
).
38.
P.
Giannozzi
,
S.
Baroni
,
N.
Bonini
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
G. L.
Chiarotti
,
M.
Cococcioni
,
I.
Dabo
,
A. D.
Corso
,
S.
de Gironcoli
,
S.
Fabris
,
G.
Fratesi
,
R.
Gebauer
,
U.
Gerstmann
,
C.
Gougoussis
,
A.
Kokalj
,
M.
Lazzeri
,
L.
Martin-Samos
,
N.
Marzari
,
F.
Mauri
,
R.
Mazzarello
,
S.
Paolini
,
A.
Pasquarello
,
L.
Paulatto
,
C.
Sbraccia
,
S.
Scandolo
,
G.
Sclauzero
,
A. P.
Seitsonen
,
A.
Smogunov
,
P.
Umari
, and
R. M.
Wentzcovitch
,
J. Phys.: Condens. Matter
21
,
395502
(
2009
).
39.
P.
Giannozzi
,
O.
Andreussi
,
T.
Brumme
,
O.
Bunau
,
M. B.
Nardelli
,
M.
Calandra
,
R.
Car
,
C.
Cavazzoni
,
D.
Ceresoli
,
M.
Cococcioni
,
N.
Colonna
,
I.
Carnimeo
,
A. D.
Corso
,
S.
de Gironcoli
,
P.
Delugas
,
R. A.
DiStasio
,
A.
Ferretti
,
A.
Floris
,
G.
Fratesi
,
G.
Fugallo
,
R.
Gebauer
,
U.
Gerstmann
,
F.
Giustino
,
T.
Gorni
,
J.
Jia
,
M.
Kawamura
,
H.-Y.
Ko
,
A.
Kokalj
,
E.
Küçükbenli
,
M.
Lazzeri
,
M.
Marsili
,
N.
Marzari
,
F.
Mauri
,
N. L.
Nguyen
,
H.-V.
Nguyen
,
A. O.
de-la Roza
,
L.
Paulatto
,
S.
Poncé
,
D.
Rocca
,
R.
Sabatini
,
B.
Santra
,
M.
Schlipf
,
A. P.
Seitsonen
,
A.
Smogunov
,
I.
Timrov
,
T.
Thonhauser
,
P.
Umari
,
N.
Vast
,
X.
Wu
, and
S.
Baroni
,
J. Phys.: Condens. Matter
29
,
465901
(
2017
).
40.