Raman spectroscopic measurements of the arsenolite–hydrogen inclusion compound As4O6·2H2 were performed in diamond anvil cells at high pressure and variable temperature down to 80 K. The experimental results were complemented by ab initio molecular dynamics simulations and phonon calculations. Observation of three hydrogen vibrons in As4O6·2H2 is reported in the entire temperature and pressure range studied (up to 24 GPa). While the experiments performed with protium and deuterium at variable temperatures allowed for the assignment of two vibrons as Q1(1) and Q1(0) transitions of ortho and para spin isomers of hydrogen trapped in the inclusion compound, the origin of the third vibron could not be unequivocally established. Low-temperature spectra revealed that the lowest-frequency vibron is actually composed of two overlapping bands of Ag and T2g symmetries dominated by H2 stretching modes as predicted by our previous density functional theory calculations. We observed low-frequency modes of As4O6·2H2 vibrations dominated by H2 “librations,” which were missed in a previous study. A low-temperature fine structure was observed for the J = 0 → 2 and J = 1 → 3 manifolds of hydrogen trapped in As4O6·2H2, indicating the lifting of degeneracy due to an anisotropic environment. A non-spherical distribution was captured by molecular dynamics simulations, which revealed that the trajectory of H2 molecules is skewed along the crystallographic ⟨111⟩ direction. Last but not least, low-temperature synchrotron powder x-ray diffraction measurements on As4O6·2H2 revealed that the bulk structure of the compound is preserved down to 5 K at 1.6 GPa.

Hydrogen is an emerging fuel of the future. It is ecological, as only water vapor is formed after its combustion, and very light offering high density of energy if a proper method of storage is applied.1 Nonetheless, there are still many challenges related to cheap hydrogen production and efficient storage. The basic research into intermolecular interactions of H2 molecules with host materials to elucidate factors responsible for their strength and directionality and, in turn, to help scientists and engineers design suitable hydrogen storage materials is necessary. Hydrogen inclusion compounds are good model compounds for such research. A few years ago, we discovered an arsenolite–hydrogen inclusion compound As4O6·2H2 that forms above 1.5 GPa and characterized its structure and properties using x-ray diffraction, Raman spectroscopy, and density functional theory (DFT) computations.2 The compound crystallizes in the cubic Fd3¯m space group (no. 227, origin choice 2) with As4O6 and H2 molecules each centered on special Wyckoff positions. Arsenolite molecules sit at the 8b site with 4¯3m (Td) point group symmetry, while hydrogen molecules are located at the 16c site with 3¯m (D3d) point group symmetry (see Fig. 1). We observed that the H2Q1(1) vibron is softened in the inclusion compound compared to fluid H2 over the entire pressure range studied, and the vibron continues to soften with increasing pressure. Interestingly, the vibron was observed to split into three components up to 4 GPa and into two components up to 10 GPa even though structural and computational results suggest that H2 molecules occupy only one Wyckoff site.

FIG. 1.

Crystal structure of the As4O6·2H2 arsenolite–hydrogen inclusion compound.2 As4O6 molecules are presented using a wireframe model with arsenic and oxygen denoted by green and red, respectively. Hydrogen molecules are presented using a ball-and-stick model and are enclosed within gray spheres centered at the 16c Wyckoff site.

FIG. 1.

Crystal structure of the As4O6·2H2 arsenolite–hydrogen inclusion compound.2 As4O6 molecules are presented using a wireframe model with arsenic and oxygen denoted by green and red, respectively. Hydrogen molecules are presented using a ball-and-stick model and are enclosed within gray spheres centered at the 16c Wyckoff site.

Close modal

Multiple and/or split hydrogen vibrons have been observed for related inclusion compounds and may indicate a variety of physical phenomena. For instance, clathrate hydrates of hydrogen exhibit multiple H2 vibrons due to the fact that various numbers of H2 molecules occupy large and small cavities therein and, additionally, each of these bands are split to reflect the population hydrogen’s two spin isomers.3,4 Owing to the fact that 1H nuclei are fermions, the total wavefunction of a hydrogen molecule must be antisymmetric with respect to nuclei permutation. This, combined with the symmetry of the electronic part of the wavefunction, leads to the fact that hydrogen at room temperature (RT) is composed of a mixture of two spin isomers that cannot interconvert quickly into one another. ortho-1H2 is the isomer with the symmetric nuclear spin wavefunction (I = 1) and occurs only in rotational states described by odd rotational (J) quantum numbers, whereas para-1H2 exhibits an antisymmetric nuclear spin wavefunction (I = 0) with even J. The situation is different for deuterium, in which the 2H nuclei are bosons. The ortho-2H2 isomer exhibits a symmetric nuclear spin wavefunction (I = 0 or 2) and even J, whereas para-2H2 possesses an antisymmetric nuclear wavefunction (I = 1) and odd J. The equilibrium distribution of ortho to para H2 is a result of an interplay between the degeneracy of nuclear spin and rotational wavefunctions as well as Boltzmann factor. In summary, the orthopara ratio for 2H2 (deuterium) at RT is 2:1 and increases upon cooling, whereas the ratio at RT for 1H2 (protium) is 3:1 and decreases upon cooling.3,5 It is noteworthy that the orthopara conversion is forbidden by selection rules and requires the exchange of angular momentum, which is typically a very slow process. The presence of an exchange catalyst, containing paramagnetic oxides like commercially available Fe2O3·2%H2O or CrO·SiO2, is often used to increase conversion rates.6 

Other examples of systems with multiple or split H2 vibrons include compounds such as Ar(H2)2,7 Xe(H2)x,8,9 SiH4(H2)2,10,11 GeH4(H2)2,12,13 (H2S)2H2,14 (H2O)2H2,15 Kr(H2)4,16 and H2 dissolved in solid Ne.17 The complex Raman spectra observed in compounds such as SiH4(H2)2, which exhibits at least seven vibrons, were explained based on slight perturbations of freely rotating H2 molecules leading to overall lowering of the crystallographic symmetry,11 similar to the statistical distribution of multiple local environments for H2 dissolved in rare gas solids.17 In other compounds, such as high-pressure (HP) (H2S)2(H2), multiple hydrogen vibrons reflect different crystallographic sites for hydrogen molecules.14 For the case of Xe(H2)x, a hydrogen 3 × 3 × 3 supercell was postulated.8,9,14

Herein, we present a detailed high-pressure (HP) Raman study of the arsenolite inclusion compound spanning from room temperature (RT) to low temperature (LT) both with hydrogen and deuterium to elucidate the physical origins of the observed hydrogen vibrons in As4O6·2H2. This paper is organized as follows: Sec. II explains the experimental and computational approach that we have applied, Sec. III describes the results from our studies, and, finally, Sec. IV contains discussion and conclusions.

Symmetric diamond anvil cells (DACs) typically equipped with diamonds having 400-µm culets together with Re gaskets pre-indented to a thickness of 50 µm–60 µm and with laser-drilled 250-µm circular holes were used. The DACs were capable of in situ high-pressure and low-temperature Raman measurements. Thin pellets of arsenolite powder were placed in the gasket hole alongside with 2–3 ruby chips used for pressure determination from ruby fluorescence shift.18,19 The pressure scale of Dewaele et al. along with the temperature corrections from Datchi et al. was applied, and fitting was carried out using the T-Rax program.20,21 The DACs were subsequently gas-loaded with protium 1H2 or deuterium 2H2, and Raman measurements were carried out. The samples were excited using a 532-nm diode laser. The laser was focused onto the sample using a 20× (10× in the case of LT measurements) objective lens. The backscattered light was detected using a Princeton Instrument spectrograph SP2750. An 1800-grooves/mm grating was used to disperse the Raman light onto a liquid-nitrogen-cooled CCD, which enabled a spectral resolution of ∼1 cm−1. Initial experiments indicated that the As4O6·2H2 inclusion compound is sensitive to intense laser light by the observation of elemental arsenic formed upon irradiation via the reduction of arsenolite by hydrogen. Consequently, subsequent measurements were performed with a laser power near 1 mW to record Raman spectra with a sufficient signal-to-noise ratio and to avoid sample damage. Neon emission lines were used to calibrate the spectrometer. The spectra were fitted with Lorentzian peaks using the fityk 1.3.1 program.22 In some LT runs, Fe2O3 powder, left in the air for a couple of days to absorb moisture, was also loaded into DACs to help facilitate the orthopara conversion of hydrogen. Computations of Raman modes were carried out according to the procedure outlined in Ref. 2.

HP, LT powder x-ray diffraction measurements on the arsenolite hydrogen inclusion compound were carried out at beamline 16-BMD, HPCAT, at the Advanced Photon Source synchrotron. A helium-cooled cryostat was used with DACs equipped with 600-µm culets to help minimize the pressure variation with temperature. Diffraction data were collected with a MAR 345 image plate. The raw images were integrated with the Dioptas 0.5.0 software, and Pawley fitting was carried out using GSAS-II.23,24

Ab initio molecular dynamics simulations for the As4O6·2H2 inclusion compound were performed with the NPT (N is the number of particles, P is the pressure, and T is the temperature) ensemble implemented in Vienna ab initio Simulation Package (VASP) code with Langevin dynamics.25,26 The all-electron projector-augmented wave (PAW) potential was adopted with the PAW potentials taken from the VASP library, where 4s24p3, 2s22p4, and 1s1 are treated as valence electrons for As, O, and H atoms, respectively.27 The plane wave was expanded to an energy cutoff of 400 eV, and Brillouin zone sampling with a 2 × 2 × 2 Monkhorst–Pack k-mesh was employed.28 Exchange and correlation effects were treated in the Perdew–Burke–Ernzerhof parameterization of the generalized gradient approximation.29 Dispersion interactions were accounted for using Grimme’s D3 correction with standard damping function—D3(zero).30,31 Trajectories of atomic movements were visualized using the VMD software.32 

Variable-temperature and pressure Raman spectra of powdered As4O6·2H2 are shown in Figs. 2 and 3. We found that the formation rate of the arsenolite–hydrogen inclusion compound is strongly dependent on the pressure at which the compound is formed. We were able to observe its formation for protium only above 1.5 GPa. The reaction takes tens of minutes to complete at this pressure, while at 1.8 GPa, it is completed within a few minutes (see Fig. S1 for the evolution of spectra as a function of time at ∼1.5 GPa). The formation of the inclusion compound is evidenced by a discontinuous change in the frequencies of As4O6 molecular vibrations and by the appearance of H2 vibrons in the Raman spectrum. The evolution of Raman spectra with pressure at room temperature, depicted in Figs. 2(a) and 3(a), confirms our earlier observations from the Raman spectra recorded for the single crystalline sample of the inclusion compound. However, careful acquisition of spectra from a thicker powder sample and subsequent fitting of the spectra revealed the presence of three 1H2 vibrons over the entire pressure range studied [up to 24.3(2) GPa] for the inclusion compound (see Figs. S2 and S3 for the frequency dependence of modes with pressure). Our interpretation of the previous Raman spectra pointed to the presence of three vibrons below ∼5 GPa and of two vibrons above this pressure.2 For clarity of presentation and discussion, the vibrons will be hereafter referred to as ν1, ν2, and ν3, and their numbering corresponds to increasing wavenumber. It is noteworthy that (1) all three components soften with pressure, while the opposite trend is observed for fluid/solid H2; (2) the frequency difference in between ν1 and ν2 decreases with pressure, and these bands merge above ∼8 GPa (above this pressure, fits of the spectra with two peaks are significantly better than with one peak only); (3) ν3 becomes very broad with diminished relative intensity at high pressure, and its frequency dependence with pressure deviates from the other two bands. The last two observations explain why we missed the presence of all three components at higher pressures in our initial study on As4O6·2H2.

FIG. 2.

Raman spectra of As4O6·21H2 and As4O6·22H2 in the 50 cm−1–900 cm−1 spectral range. The evolution of As4O6·21H2 spectra as a function of pressure at room temperature (a) and of temperature at 3.6(3) GPa (b). The evolution of As4O6·22H2 spectra as a function of pressure at room temperature (c) and of temperature at 2.6(3) GPa (d). Labels next to spectra denote pressure or temperature. Thinner lines slightly below inclusion compound spectra correspond to the spectra of fluid/solid H2 recorded at the same conditions.

FIG. 2.

Raman spectra of As4O6·21H2 and As4O6·22H2 in the 50 cm−1–900 cm−1 spectral range. The evolution of As4O6·21H2 spectra as a function of pressure at room temperature (a) and of temperature at 3.6(3) GPa (b). The evolution of As4O6·22H2 spectra as a function of pressure at room temperature (c) and of temperature at 2.6(3) GPa (d). Labels next to spectra denote pressure or temperature. Thinner lines slightly below inclusion compound spectra correspond to the spectra of fluid/solid H2 recorded at the same conditions.

Close modal
FIG. 3.

Raman spectra of As4O6·21H2 and As4O6·22H2 in the H2 vibron spectral range. The evolution of the As4O6·21H2 spectra as a function of pressure at room temperature (a) and of temperature at 3.6(3) GPa (b). The evolution of the As4O6·22H2 spectra as a function of pressure at room temperature (c) and of temperature at 2.6(3) GPa (d). Labels next to spectra denote pressure or temperature.

FIG. 3.

Raman spectra of As4O6·21H2 and As4O6·22H2 in the H2 vibron spectral range. The evolution of the As4O6·21H2 spectra as a function of pressure at room temperature (a) and of temperature at 3.6(3) GPa (b). The evolution of the As4O6·22H2 spectra as a function of pressure at room temperature (c) and of temperature at 2.6(3) GPa (d). Labels next to spectra denote pressure or temperature.

Close modal

In order to further understand the vibrational spectra observed in the inclusion compound, we carried out LT Raman measurements with protium, as shown in Figs. 2(a), 2(b), 3(a), and 3(b). Starting in the lower-frequency region below 1000 cm−1, vibrational modes associated with the arsenolite host lattice are clearly identified by comparison with spectra from the empty structure and with DFT computations. Given that all Raman measurements were performed in excess fluid hydrogen, roton excitations [e.g., S0(0) and S0(1)] from the pure fluid (or solid) are also observed in each spectrum, but are easily distinguished by comparing the spectra of pure hydrogen at the same conditions. With decreasing temperature, all Raman peak widths decrease as vibrations become more localized. Below ∼200 K, the rotational fine structure becomes apparent within the J = 0 → 2 and J = 1 → 3 manifolds near 350 cm−1 and 600 cm−1, respectively, and these bands broaden with increasing pressure. By 78 K, the splitting of these bands is significant [denoted by asterisks in Fig. 2(b)] and indicates at least partially hindered rotation of hydrogen molecules within the inclusion compound. These quantum rotor excitations are not captured by static phonon computations. Notably, computations do predict two “librational” modes with symmetries T2g and Eg, in which hydrogen atomic displacements dominate the overall character. These modes might be related to the two broad bands observed at ∼130 cm−1 and 340 cm−1 at 1.3 GPa and LT based on their appearance upon formation of the inclusion compound and their extreme pressure dependence, which agrees with the calculations.

Turning to the high-frequency region, all vibron peaks become sharper as temperature decreases (see Fig. S4), and the ν1 vibron clearly exhibits a shoulder on the high-frequency side indicating that it is, in fact, comprised of two very closely spaced ν1a and ν1b peaks. Given that DFT computations also predict two closely spaced Raman vibrons, one of which is of Ag and the other of T2g symmetry, it appears that the symmetry of the crystal structure explains the origin of these bands (see Tables S1 and S2). The relative intensities of the three vibrons also reveal distinct trends with temperature. At a constant pressure of 3.6(3) GPa, the relative intensity ratio of ν1 to all other vibrons (defined as the area of ν1a and ν1b over the total area of all vibrons) varies significantly, while the relative intensity ratio of the ν2 vibron to all other vibrons remains constant at ∼0.25 upon cooling from RT to 80 K. The relative intensity of the ν3 vibron diminishes from 0.36 at RT to 0.04 at ∼80 K (see Figs. 4 and S5 for an extended version including relative fluid/solid H2 roton intensities). The relative intensities of the ν13 vibrons recorded at RT do not exhibit any clear trends as a function of pressure.

FIG. 4.

Relative intensities of inclusion compound vibrons for protium (a) and deuterium (b) plotted as a function of temperature at 3.6(3) GPa and 2.6(3) GPa, respectively.

FIG. 4.

Relative intensities of inclusion compound vibrons for protium (a) and deuterium (b) plotted as a function of temperature at 3.6(3) GPa and 2.6(3) GPa, respectively.

Close modal

To further understand the observed vibrational spectra, we carried out analogous experiments with deuterium. In this case, the interaction potential should remain essentially constant, while the large isotopic mass shift and inverted nuclear spin isomers allow for unambiguous spectral assignments. A comparison of the spectra of the deuterium and protium inclusion compounds in the low-frequency range, shown in Figs. 2(a) and 2(c), allowed us to confirm all the host modes that were predicted by us previously using CRYSTAL09 calculations and also suggest that the calculated Eg and T2g libration modes of H2 molecules are related to the observed modes that exhibit very strong pressure dependence (see Fig. S2). While the frequencies of modes dominated by As4O6 molecules are predicted very well, the computed absolute frequencies of modes with dominant contributions from H2 molecules must be shifted by −160 cm−1 and −90 cm−1 for protium and deuterium, respectively, to compare with experimental values, but their evolution with pressure is predicted properly. Interestingly, we were able to obtain the inclusion compound with deuterium at only 1.09(13) GPa (compared with ∼1.5 GPa for hydrogen), which was evidenced by a broad 2H2 vibron in the inclusion compound clearly separated from the fluid 2H2 vibron. Attempts to obtain the inclusion compound with protium at similar pressure were not successful, even after 17 h at 1.03(9) GPa.

The 2H2 vibron in the inclusion compound is also composed of three components (ν1–ν3), but at lower frequencies due to the higher reduced mass of deuterium [Fig. 3(c)]. While the absolute frequency difference between these bands is lower than that observed for protium, fits of the 2H2 spectra recorded at RT with three Lorentzian peaks are always significantly better than with fits using one or two peaks. Moreover, the presence of three vibrons becomes evident at LT when the peaks become sharper [see Figs. 3(d) and S4]. The observed frequencies of the 2H2 vibrons, both in the inclusion compound and in the fluid/solid, are in very good agreement with the frequencies of the respective 1H2 vibrons when scaled by a factor of 1.39, which is close to the theoretical reduced mass ratio of 1.41 (see Fig. S6). At ∼2.5 GPa, the temperature dependence of the relative intensities of the 2H2 vibrons in the inclusion compound similar to 1H although the trends for ν1 and ν2 are obscured above 200 K due to significant overlap of these peaks [see Figs. 3(d), 4, and S5]. The relative intensity of ν3 falls from 0.55 at 260 K to 0.07 at 85 K, while the relative intensity of ν2 grows from RT to 200 K when it reaches ∼0.60 and remains constant down to 85 K. Similar to protium, ν1 for deuterium grows in intensity as temperature is lowered, while ν3 decreases in intensity over the same temperature range. We were not able to detect a shoulder on ν1 in the case of deuterium, but we suspect that this is caused by the very close spacing of peaks, even at 85 K. It is noteworthy that the relative intensities of ν1 and ν2 are inverted for deuterium as compared to protium.

Our attempts to determine the LT stability of the inclusion compound at ambient pressure were complicated by irreversible sticking of the DACs that prevented full pressure release at LT. We were, however, able to completely release the pressure during one LT experiment with deuterium, which revealed that the inclusion compound remains (meta)stable at ambient pressure and 85 K for at least 21 h (see Fig. S7 for the recorded Raman spectra).

Noting the observed vibron intensity trends between protium and deuterium and their frequency separations relative to the pure fluid/solid, we decided to carry out an additional HP–LT Raman experiment over a longer time to examine possible orthopara conversion and make definitive peak assignments. While orthopara conversion rates vary drastically under pressure,33,34 conversion in other inclusion compounds is typically slow,3,4 and we therefore utilized a catalyst to possibly help expedite conversion. Our goal was to study the inclusion compound at the lowest pressure possible to minimize pressure-induced intermolecular coupling, which causes dramatic changes in the relative intensities of the J = 0 and J = 1 vibrons.35 Due to the problems associated with releasing the DAC pressure at LT described above, we were only able to decrease the pressure down to 1.29(5) GPa, which was sufficient to clearly observe the Q1(1) and Q1(0) vibrons of solid 1H2 [see Fig. 5(d)].

FIG. 5.

Relative intensities of inclusion compound 1H2 vibrons (a), of solid 1H2 vibrons (b) and of solid 1H2 rotons (c) plotted as a function of time at 85(2) K and 1.29(5) GPa. The Raman spectrum of the inclusion compound in the vibron spectral range at these conditions at the beginning of the experiment (d). Vertical and dotted lines correspond to the Ag and T2g modes, respectively, for ortho and para H2, whereas the dashed lines illustrate that for ν3, there is no orthopara pair.

FIG. 5.

Relative intensities of inclusion compound 1H2 vibrons (a), of solid 1H2 vibrons (b) and of solid 1H2 rotons (c) plotted as a function of time at 85(2) K and 1.29(5) GPa. The Raman spectrum of the inclusion compound in the vibron spectral range at these conditions at the beginning of the experiment (d). Vertical and dotted lines correspond to the Ag and T2g modes, respectively, for ortho and para H2, whereas the dashed lines illustrate that for ν3, there is no orthopara pair.

Close modal

After 66 h at 84(1) K and 1.29(5) GPa, the relative intensities of the solid rotons and vibrons appear to reach constant values, indicating the approach to thermodynamic equilibrium [see Figs. 5(b) and 5(c)]. The relative intensities of the H2 vibrons for the inclusion compound also reached a plateau albeit within a shorter time span of 42 h. While the intensities of the solid rotons change significantly over time and reflect an ortho concentration moving toward thermal equilibrium, the intensity ratio of the Q1(1):Q1(0) fluid vibrons only decreases ∼3% reflecting the stronger vibron scattering for the J = 1 state. Given that the ν1 and ν2 relative intensities from the compound change by this same amount and exhibit a similar spacing, we assign the ν1 and ν2 vibrons as Q1(1) and Q1(0) vibrons of H2 trapped in the inclusion compound, respectively. This assignment is also supported by the inverse intensity trends observed for deuterium reflecting the difference between ortho and para for the different spin isomers. We note that while the ν1 and ν2 relative intensities mirror the ortho and para behavior of the fluid of time, the relative intensity of ν3 remained constant near 0.15 over the entire period [see Fig. 5(a)].

In order to analyze the Raman spectra that we recorded at LT, it is important to know whether any temperature-induced phase transitions occur upon cooling. We therefore collected powder x-ray diffraction patterns of the As4O6·21H2 inclusion compound at 1.6 GPa at LT down to 5 K at the 16-BMD beamline of the Advanced Photon Source. The analysis of the obtained diffraction patterns revealed that there is no phase transition down to this temperature. It is noteworthy that the x-rays used (wavelength of 0.41 Å) scatter mostly from arsenic atoms with negligible diffraction from H2 molecules. The diffraction patterns rule out a lowering of the symmetry of the host structure formed by As4O6 molecules, but do not rule out possible ordering of H2 molecules in supercells or lowering of the symmetry of the whole structure caused by H2 molecules (see Fig. S8).

Detailed analysis of H2 rotons from the inclusion compound is challenging due to the overlap with fluid/solid H2 present in the cell and overlap in the spectral range where strong As4O6 molecular vibrations are present [see Figs. 2(b) and S7(a)]. Nonetheless, it is clear that a low-temperature fine structure is observed for the J = 0 → 2 and J = 1 → 3 manifolds, indicating the lifting of degeneracy due to an anisotropic environment (which was observed for other inclusion compounds, e.g., for hydrogen clathrate hydrates).3,4 The energy splitting for these bands is quite significant. Partially resolved components for the J = 0 → 2 transitions centered about 360 cm−1 appear to span 40 cm−1, while components appear to span more than 200 cm−1 for the J = 1 → 3 transitions. At least partially hindered rotation, similar to the 2D rotor behavior for hydroquinone clathrate,36 could potentially explain such splitting. While being incapable of capturing the quantum nature of the rotor, molecular dynamics simulations performed on the As4O6·2H2 compound do indeed indicate a non-spherical classical probability distribution of H2 molecules at the 16c Wyckoff site wherein the distribution is skewed along the crystallographic ⟨111⟩ direction (Fig. S9). Previously, we suggested that H2 molecules might be aligned along the crystallographic ⟨111⟩ direction based on the calculated energies of an H2 molecule in different orientations enclosed in an (As4O6)6 cluster simulating the environment of the molecule at the 16c Wyckoff site in As4O6·2H2.2 Interestingly, the roton bands appeared much weaker in the spectra of the deuterium inclusion compound, but their presence could be detected in the quenched sample at 85 K [Fig. S7(a)].

Two additional modes were observed at low frequency and appear to be associated with hydrogen (deuterium). For 1H2, these bands appear at ∼130 cm−1 and 340 cm−1 at 1.3 GPa and 85 K. As the pressure is increased, they exhibit very strong pressure dependence that agrees with the T2g and Eg librational modes predicted by DFT computations. It is conceivable that these modes are actually also associated with the J = 0 → 2 manifold, but a shift to 130 cm−1 would represent an unprecedented energy splitting for a molecular inclusion compound. Additional experiments at lower temperature, including inelastic neutron scattering, are needed to fully understand the coupled translational–rotational behavior of hydrogen within As4O6·2H2.

The hydrogen vibrons in As4O6·2H2 are significantly softened from the fluid/solid hydrogen vibrons. From previously characterized H2 van der Waals compounds, we infer a trend that whenever H2 forms an inclusion compound with an electron donating substance, the H2 vibron is red-shifted due to the partial donation of electron density to the H2σ* antibonding orbital. This is the case for the herein studied As4O6·2H2 as well as (H2S)2H2,14 (H2Se)2H2,37 hydrogen clathrate hydrates,3,4 (HI)(H2)2 and (HI)(H2)13,38 in which all host molecules contain atoms with stereoactive lone electron pairs that donate some of its electron density to the H2 molecule. The inclusion compounds with He, Ne, Ar, and Kr exhibit blue-shifted H2 vibron, indicating repulsive interactions with hydrogen molecules and no charge transfer from noble gas atoms to H2 molecules.7,16,17 The effect is reversed in the inclusion compound of hydrogen with xenon where the H2 vibron is red-shifted, indicating the Xe valence electron density is partially transferred to the H2 antibonding orbital.8,9 This is in agreement with the fact that Xe is known for being much more reactive than lighter noble gases and forming numerous chemical compounds. An interesting observation can be made in the series of H2 van der Waals compounds with CH4, SiH4, and GeH4.10,12,39 While the H2 vibron is blue-shifted in the inclusion compound with methane, it is red-shifted for silane and germane. This is caused by the inversion of the X–H bond polarity. The C–H bonds are polarized toward carbon leaving partial positive charge on H atoms making donation of electron density to the H2σ* orbital impossible. From Si on, the X–H bonds are polarized toward H atoms leading to interactions similar to dihydrogen bonds that result in electron density being partially transferred to the H2σ* orbital and weakening the H–H bond.40 In the hydrogen–nitrogen van der Waals compounds (N2)6(H2)7 and (N2)(H2)2, the H2 vibron is blue-shifted, indicating only repulsive interactions between N2 and H2 molecules.41,42 It is noteworthy that the H2 vibron is softened when protium or deuterium is dissolved in solid matrices of Ar, Kr, and Xe at ambient pressure and LT (see Ref. 43 and references therein). Calculations revealed that the vibron is softened whenever the H2 molecule is located in the region of space where the intermolecular interaction potential is attractive.43 Taking this into account, we may conclude that there are two general mechanisms responsible for H2 vibron softening. One stems from the local intermolecular potential and attraction between host molecules/atoms and H2 molecules, while the other comes from partial charge transfer between host and H2 molecules. The latter mechanism dominates repulsive interactions between the host and guest at HP as has been shown for As4O6·2H2.2 

As for the multiple hydrogen vibrons observed for As4O6·2H2, our carefully collected spectra allowed us to conclude that there are three H2 vibrons from the inclusion compound up to the highest studied pressure. We assign ν1 and ν2 as Q1(1) and Q1(0) based on (1) the time dependence of their relative intensities at LT indicating slow orthopara conversion for protium and (2) the change in their relative intensities when going from protium to deuterium reflecting the difference in degeneracy of ortho and para spin isomers between 1H2 and 2H2. As for the shoulder present on ν1 in LT protium measurements, we attribute the ν1a main peak and the ν1b shoulder to Ag and T2g vibrations, respectively, both of which were predicted by DFT computations and may both exhibit orthopara contributions in the observed spectra [see dotted lines in Fig. 5(d)].

The origin of ν3 is more difficult. Our first hypothesis was that H2 occupies an additional site in the As4O6·2H2 crystal structure. This explanation seems unlikely to be the case since there should be a second pair of vibrons corresponding to ortho and para isomers occupying the second site, which is not the case here [see dashed vertical lines in Fig. 5(d)]. The occupation of a second site is also inconsistent with our MD simulations. We simulated the As4O6·2H2 crystal structure with H2 molecules initially positioned on the 16c site at RT and multiple pressure points (2 GPa, 4 GPa, and 6 GPa), and we did not observe H2 molecules migrate to any other sites (see Fig. S9). When H2 molecules were positioned on the 8a site, they quickly migrate to the 16c site, substantiating its energetic favorability. A possible temperature-dependent occupancy distribution of the 16c site (leading to multiple effective local environments) also seems unlikely based on the absence of orthopara pairs, the behavior with pressure, and previous single-crystal XRD measurement that give no indication for non-stoichiometry.

It is tempting to assign ν3 as an additional rovibrational contribution [e.g., Q1(3)] based on the temperature dependence of the relative intensities of ν1 and ν3 both for protium and deuterium (Fig. 4). Here, the depopulation of a higher-energy rotational state is consistent with the observed temperature dependence, as is the apparent transfer of intensity to the next low-lying transition with the same nuclear spin state. However, such an assignment would require an unusual rearrangement of energy levels such that Q1(3) is observed at a higher frequency than Q1(0) and Q1(1). This assignment is also unlikely in that Q1(2) is never observed and Q1(3) should not be populated at 80 K [the intensity of solid hydrogen S0(3) is zero, indicating that the J = 3 rotational state is not occupied in the pure phase].

It is noteworthy that the ν3 vibron is significantly broader than the ν1 and ν2 vibrons, which may suggest a different physical origin. The observed temperature dependence of the ν3 vibron is consistent with a Boltzmann factor and could indicate possible coupling with a phonon mode. For example, a librational mode or other translational state could potentially couple with symmetric stretching to yield a separate vibron. The final understanding of this vibron will require additional measurements to reveal the detailed energy levels of H2 trapped within arsenolite. HP solid-state 1H nuclear magnetic resonance (NMR), lower-temperature Raman, and inelastic neutron scattering measurements are all being pursued to reveal a greater understanding of the dynamics.

See the supplementary material for additional Raman spectra, trajectories from MD simulations, As4O6·2H2 powder diffraction patterns, and tables of calculated Raman frequencies for As4O6·21H2 and As4O6·22H2.

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

This work was supported by the Polish National Agency for Academic Exchange. Computations were performed using the Wroclaw Centre for Networking and Supercomputing (http://wcss.pl), Grant No. 260, and the Memex cluster of Carnegie Institution for Science. T.A.S. acknowledges the support from the U.S. Army Research Office under Grant No. W911NF-17-1-0604. Portions of this work were performed at HPCAT (Sector 16), Advanced Photon Source (APS), Argonne National Laboratory. We thank Curtis Kenney-Benson and Dmitry Popov for assistance with cryostat work and x-ray diffraction data collection during beamtime. HPCAT operations are supported by DOE–NNSA’s Office of Experimental Sciences. The Advanced Photon Source is a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357.

1.
S.
Sharma
and
S. K.
Ghoshal
,
Renewable Sustainable Energy Rev.
43
,
1151
(
2015
).
2.
P. A.
Guńka
,
M.
Hapka
,
M.
Hanfland
,
G.
Chałasiński
, and
J.
Zachara
,
J. Phys. Chem. C
123
,
16868
(
2019
).
3.
T. A.
Strobel
,
E. D.
Sloan
, and
C. A.
Koh
,
J. Chem. Phys.
130
,
014506
(
2009
).
4.
A.
Giannasi
,
M.
Celli
,
L.
Ulivi
, and
M.
Zoppi
,
J. Chem. Phys.
129
,
084705
(
2008
).
5.
J.
Van Kranendonk
,
Solid Hydrogen
, 1st ed. (
Plenum Press
,
New York
,
1983
).
6.
A. V.
Zhuzhgov
,
O. P.
Krivoruchko
,
L. A.
Isupova
,
O. N.
Mart’yanov
, and
V. N.
Parmon
,
Catal. Ind.
10
,
9
(
2018
).
7.
P.
Loubeyre
,
R.
Letoullec
, and
J.-P.
Pinceaux
,
Phys. Rev. Lett.
72
,
1360
(
1994
).
8.
M.
Somayazulu
,
P.
Dera
,
A. F.
Goncharov
,
S. A.
Gramsch
,
P.
Liermann
,
W.
Yang
,
Z.
Liu
,
H.-K.
Mao
, and
R. J.
Hemley
,
Nat. Chem.
2
,
50
(
2010
).
9.
M.
Somayazulu
,
P.
Dera
,
J.
Smith
, and
R. J.
Hemley
,
J. Chem. Phys.
142
,
104503
(
2015
).
10.
T. A.
Strobel
,
M.
Somayazulu
, and
R. J.
Hemley
,
Phys. Rev. Lett.
103
,
065701
(
2009
).
11.
Y.
Li
,
G.
Gao
,
Q.
Li
,
Y.
Ma
, and
G.
Zou
,
Phys. Rev. B
82
,
064104
(
2010
).
12.
T. A.
Strobel
,
X.-J.
Chen
,
M.
Somayazulu
, and
R. J.
Hemley
,
J. Chem. Phys.
133
,
164512
(
2010
).
13.
G.
Zhong
,
C.
Zhang
,
X.
Chen
,
Y.
Li
,
R.
Zhang
, and
H.
Lin
,
J. Phys. Chem. C
116
,
5225
(
2012
).
14.
T. A.
Strobel
,
P.
Ganesh
,
M.
Somayazulu
,
P. R. C.
Kent
, and
R. J.
Hemley
,
Phys. Rev. Lett.
107
,
255503
(
2011
).
15.
T. A.
Strobel
,
M.
Somayazulu
,
S. V.
Sinogeikin
,
P.
Dera
, and
R. J.
Hemley
,
J. Am. Chem. Soc.
138
,
13786
(
2016
).
16.
A. K.
Kleppe
,
M.
Amboage
, and
A. P.
Jephcoat
,
Sci. Rep.
4
,
4989
(
2014
).
17.
P.
Loubeyre
,
R.
LeToullec
, and
J. P.
Pinceaux
,
Phys. Rev. B
45
,
12844
(
1992
).
18.
H. K.
Mao
,
J.
Xu
, and
P. M.
Bell
,
J. Geophys. Res.: Solid Earth
91
,
4673
, (
1986
).
19.
K.
Syassen
,
High Pressure Res.
28
,
75
(
2008
).
20.
A.
Dewaele
,
M.
Torrent
,
P.
Loubeyre
, and
M.
Mezouar
,
Phys. Rev. B
78
,
104102
(
2008
).
21.
F.
Datchi
,
A.
Dewaele
,
P.
Loubeyre
,
R.
Letoullec
,
Y.
Le Godec
, and
B.
Canny
,
High Pressure Res.
27
,
447
(
2007
).
22.
M.
Wojdyr
,
J. Appl. Crystallogr.
43
,
1126
(
2010
).
23.
C.
Prescher
and
V. B.
Prakapenka
,
High Pressure Res.
35
,
223
(
2015
).
24.
B. H.
Toby
and
R. B.
Von Dreele
,
J. Appl. Crystallogr.
46
,
544
(
2013
).
25.
E.
Hernández
,
J. Chem. Phys.
115
,
10282
(
2001
).
26.
G.
Kresse
and
J.
Furthmüller
,
Phys. Rev. B
54
,
11169
(
1996
).
27.
G.
Kresse
and
D.
Joubert
,
Phys. Rev. B
59
,
1758
(
1999
).
28.
H. J.
Monkhorst
and
J. D.
Pack
,
Phys. Rev. B
13
,
5188
(
1976
).
29.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1996
).
30.
S.
Grimme
,
J.
Antony
,
S.
Ehrlich
, and
H.
Krieg
,
J. Chem. Phys.
132
,
154104
(
2010
).
31.
S.
Grimme
,
S.
Ehrlich
, and
L.
Goerigk
,
J. Comput. Chem.
32
,
1456
(
2011
).
32.
W.
Humphrey
,
A.
Dalke
, and
K.
Schulten
,
J. Mol. Graphics
14
,
33
(
1996
).
33.
A.
Driessen
,
E.
van der Poll
, and
I. F.
Silvera
,
Phys. Rev. B
30
,
2517
(
1984
).
34.
J. H.
Eggert
,
E.
Karmon
,
R. J.
Hemley
,
H.-K.
Mao
, and
A. F.
Goncharov
,
Proc. Natl. Acad. Sci. U. S. A.
96
,
12269
(
1999
).
35.
J. H.
Eggert
,
R. J.
Hemley
,
H. K.
Mao
, and
J. L.
Feldman
,
AIP Conf. Proc.
309
,
845
(
1994
).
36.
T. A.
Strobel
,
A. J.
Ramirez-Cuesta
,
L. L.
Daemen
,
V. S.
Bhadram
,
T. A.
Jenkins
,
C. M.
Brown
, and
Y.
Cheng
,
Phys. Rev. Lett.
120
,
120402
(
2018
).
37.
E. J.
Pace
,
J.
Binns
,
M.
Peña Alvarez
,
P.
Dalladay-Simpson
,
E.
Gregoryanz
, and
R. T.
Howie
,
J. Chem. Phys.
147
,
184303
(
2017
).
38.
J.
Binns
,
P.
Dalladay-Simpson
,
M.
Wang
,
G. J.
Ackland
,
E.
Gregoryanz
, and
R. T.
Howie
,
Phys. Rev. B
97
,
024111
(
2018
).
39.
M. S.
Somayazulu
,
L. W.
Finger
,
R. J.
Hemley
, and
H. K.
Mao
,
Science
271
,
1400
(
1996
).
40.
R.
Custelcean
and
J. E.
Jackson
,
Chem. Rev.
101
,
1963
(
2001
).
41.
D. K.
Spaulding
,
G.
Weck
,
P.
Loubeyre
,
F.
Datchi
,
P.
Dumas
, and
M.
Hanfland
,
Nat. Commun.
5
,
5739
(
2014
).
42.
D.
Laniel
,
V.
Svitlyk
,
G.
Weck
, and
P.
Loubeyre
,
Phys. Chem. Chem. Phys.
20
,
4050
(
2018
).
43.
J.
Vitko
and
C. F.
Coll
,
J. Chem. Phys.
69
,
2590
(
1978
).

Supplementary Material