Abstract
The structural and vibrational properties of the molecular units of sulfur hexafluoride crystal as a function of pressure have been studied by the Extreme Pressure Polarizable Continuum Model (XP-PCM) method. Within the XP-PCM model, single molecule calculations allow a consistent interpretation of the experimental measurements when considering the effect of pressure on both the molecular structure and the vibrational normal modes. This peculiar aspect of XP-PCM provides a detailed description of the electronic origin of normal modes variations with pressure, via the curvature of the potential energy surface and via the anharmonicity of the normal modes. When applied to the vibrational properties of the sulfur hexafluoride crystal, the XP-PCM method reveals a hitherto unknown interpretation of the effects of the pressure on the vibrational normal modes of the molecular units of this crystal.
Introduction
The study of structural and dynamic properties of both liquid and solid materials at high pressure, mainly carried out in the diamond anvil cell (DAC),1−3 is an active research field1,2,4−7 for its implication in both applied science and technology and knowledge of atomic and molecular properties. Indeed, in a recent paper by Cammi, Rahm, Hoffmann and Ashcroft8 it has been envisaged that basic electronic properties along the periodic table (like atomic electronic configuration and electronegativity) exhibit drastic changes at high pressures thus modulating general chemical behaviors.
In particular, in studying molecular crystals at high pressures, the general properties can be conveniently investigated with vibrational spectroscopy and some experimental findings can be achieved by comparison with computational models. A general approach to model molecular crystals involves periodic systems.9−12 Accurate determination of structural and spectroscopic properties by ab initio solid state calculations requires elevated computational resources.10,11 But, even when vibrational properties are obtained either by ab initio molecular dynamics simulations or by crystal calculations in the harmonic approximation10−12 subsequent extraction of high pressure effects on single molecule is not an easy task.
The recently introduced eXtreme Pressure Polarizable Continuum Model (XP-PCM)4,13−18 allows for ab initio calculations at high pressure on a single molecular unit of the system of interest at a reduced computational cost. In the XP-PCM calculations the main pressure effects on the vibrational properties can be easily taken into account and these are interpreted in terms of curvature and anharmonicity of the potential energy surface, providing useful insights on pressure effects on the intramolecular normal modes,16−18 a kind of analysis which is not currently available in periodic calculations. In the XP-PCM approach the interactions with the environment are modeled using the Polarizable Continuum Model (PCM),19,20 properly taking into account the pressure effect by increasing the Pauli repulsion due to the external medium on the molecule.4 This tuning of the pressure is obtained by reducing the volume of the cavity hosting the molecular unit.4,13,16,17,21
In the present paper the XP-PCM method has been applied to study the vibrational frequencies of sulfur hexafluoride (SF6) as a function of the pressure and how these effects can be properly analyzed. Sulfur hexafluoride is an interesting candidate, since the molecular structure of isolated SF6, as depicted in Figure 1, has a perfect octahedral arrangement (almost spherical). The high symmetry of the SF6 molecule (Oh) and the limited number of atoms allow to analyze the effect of pressure on vibrational properties at high level of theory and to compare the results with the available experimental measurements.22,23 In fact, several IR and Raman experiments22−26 are available on the molecular crystal for different phases both at ambient and high pressures.23,26
Figure 1.
Molecular structure of sulfur hexafluoride SF6.
The article is organized as follows. First, the XP-PCM theory is briefly reviewed in the “XP-PCM Theory” section, then the phase diagram is discussed in the “Phase Diagram of Sulfur Hexafluoride” section. The computational procedure is described in the “Computational Protocol and Details” section and finally the results of the calculations are presented in the “Results and Discussion” section, where a detailed analysis of the vibrational properties in terms of curvature and relaxation contributions are discussed. Final remarks are reported in the “Conclusions” section.
Methods
XP-PCM Theory
Let us consider a molecular systems embedded in the external medium of the XP-PCM model. The electronic energy determining its potential energy surface is given by eq 1
![]() |
1 |
where Ṽnn is the nuclei–nuclei interaction contribution in the presence of the external medium, Q̅(Ψ)·V̂ is the PCM electrostatic interaction term20 and V̂r is the Pauli repulsion term, which describes the exchange–repulsion contribution due to the overlap of the electron distribution of the studied molecular system with the mean electron distribution of the external medium (modeled by an uniform distribution outside the cavity following the PCM approach19,20). The repulsion term is expressed by eq 2
![]() |
2 |
where ρ̂(r) = ∑inδ(r – ri) is the electron density operator (over the n electrons of the molecular system).
is a factor
representing a step barrier
potential located at the boundary of the molecular cavity, as reported
in eq 3:
![]() |
3 |
where is the height
of the potential barrier
and ΘC(r) is a generalized
Heaviside step function with a value equal to zero inside the cavity
and equal to one outside of it (
denotes the domain of
the physical space
inside the cavity). The height of the potential barrier
depends on
the volume of the cavity, Vc,4 as shown in eq 4:
![]() |
4 |
where is the height corresponding to
a reference
upper value of the cavity volume, Vc0, and N is a semiempirical parameter that gauges the strength
of the Pauli repulsive barrier originated from the external medium,
and which can be estimated by comparison of the computed pressure–volume
results of XP-PCM with the available experimental pressure–volume
data as expressed by Murnaghan equation27 for different phases. The height of the potential barrier
given in eq 4 implicitly accounts for
the pressure dependence of
the properties of the external medium13 (i.e., the dielectric permittivity and average electron density).
The pressure p is computed as the negative of the derivative of the electronic energy, eq 5, with respect to the cavity volume Vc at constant number of particles n:
![]() |
5 |
The cavity is defined in terms of overlapping spheres centered on the atomic nuclei with radius related, by a uniform scaling factor (f), to the corresponding atomic van der Waals radii (RvdWi).28 The cavity volume may be freely reduced by decreasing the scaling factor with respect to the reference value f0 = 1.2.4,13,16−18 It has been recently shown29 that the XP-PCM model is able to faithfully reduce the van der Waals radii of elements in agreement with available experimental data, suggesting that the same procedure can be successfully extended in studying molecular species.
The Pauli repulsion potential depends indirectly on f. In fact, the repulsion potential is related both to the
numeral
density of the external medium and by a semiempirical parameter
, with 3 ≤ N ≤
6, which can be experimentally determined, where a0 and Eh are
conversion factors from Bohr and Hartree, respectively.
The equilibrium geometries of the molecule under pressure correspond to the minimum of the potential energy surface determined by the electronic functional Ger of eq 1.
The vibrational harmonic frequencies are obtained from the Hessian matrix of the second derivatives of Ger with respect to the atomic displacements obtained numerically by the gradients.
As mentioned in the Introduction, the XP-PCM helps to achieve relevant insights on the pressure effect on vibrational frequencies by a partition in curvature and relaxation (or anharmonicity) contributions;4,16,17 these contributions are referred to also as direct and indirect effects, respectively.
The Effect of the Pressure on the Equilibrium Geometry and Vibrational Frequencies and Its Direct and Indirect Contributions
Describing the molecule as a set of harmonic oscillators, the electronic energy Ger in eq 5 can be expanded as a function of the vibrational normal mode coordinates of the isolated molecule Q and of the pressure p, as shown in the Taylor expansion in eq 6:
![]() |
6 |
where Ger(Q, 0) is the electronic energy at zero pressure (i.e., that of the isolated molecule), as a function of the normal coordinates Q; Γi and Γii are partial derivatives of Ger and their physical meaning is summarized in the following.
Equation 7 shows that Γi is the mixed second order derivative:
![]() |
7 |
and it represents the direct effect of pressure on the forces on the nuclei along the normal coordinate Qi. Due to the isotropic pressure, Γi is forced to be different from zero only for totally symmetric (TS) normal mode coordinates. In particular, the variations of the equilibrium geometry along the TS normal coordinates {QiTS} are given by eq 8
![]() |
8 |
where ki is the harmonic force constants of the ith normal mode. Equation 8 shows also how the pressure affects the geometry not only through the coupling coefficients {Γi} but also with the normal mode force constants, {ki}, which describe the stiffness of the normal modes.
The Γii is expressed in eq 9 by the mixed third-order coefficient:
![]() |
9 |
and it represents the direct effect of the pressure on the force constant ki of the ith normal mode. Γii could be different from zero for any vibrational normal mode.
The total effect of pressure on the vibrational harmonic force constants is given by eq 10:
![]() |
10 |
where giij is the cubic anharmonic constant coupling the normal mode i with a totally symmetric normal mode j.
Furthermore, the effect of the pressure on the vibrational frequencies can be partitioned into a curvature (direct) and a relaxation (indirect) contribution,4,13,16,17 as shown in eq 11:
![]() |
11 |
with
![]() |
12 |
![]() |
13 |
where ν(p) denotes the vibrational frequencies evaluated at the equilibrium geometries and Q(0) the equilibrium geometry of the isolated molecule.
From eqs 11–13, the pressure effect on the harmonic force constants can be partitioned in a curvature (or direct) effect, represented by Γii, which relates the influence of pressure to changes in the electron density of the system, and a relaxation (or indirect) effect, involving the cubic anharmonicities giij, which is related to the pressure induced modification of the equilibrium geometry of the molecular system.
Equations 11–13 are of fundamental importance to analyze the effect of pressure on vibrational properties of molecules in the crystal. In essence, the curvature contribution, due to the variation of the second derivative of the potential energy surface, accounts for the confinement effect due to the repulsive intermolecular interactions, evaluated at the zero-pressure equilibrium geometry, while the relaxation contribution is related to the variation of the equilibrium geometry of the molecule through the anharmonic force constants. In more general terms, this analysis is motivated by the need to give answer to some questions like “Why do the molecules respond to the pressure in the way observed? Why does this response depend on the various vibrational normal modes? Is there a property of the normal modes that governs their response to the pressure?”. We remark that a similar partitioning, useful to rationalize observed frequency shifts with pressure, has previously been carried out with a semiclassical approach by Moroni et al.30 The XP-PCM method has the definite advantage of accurately estimating the two contributions for a general molecular system.
Phase Diagram of Sulfur Hexafluoride
The phase diagram of SF6 has been studied as a function of both temperature and pressure. At ambient pressure it has been found by neutron diffraction experiments31 that at 94.3 K a phase transition occurs from a high temperature orientationally disordered structure (space group Im3m, Z = 2) to a monoclinic structure (space group C2/m – C2h3 with Z = 6). The molecules are found in two different sites, with site symmetry 2/m (or C2h) or m (or Cs). The reduction of molecular symmetry and the intermolecular interactions in the crystal produce a multiplet structure in the vibrational spectra. The correlation diagram for the C2h site shows that the degeneracy of the normal modes is removed. Therefore, for molecules on this site it is expected that the A1g, Eg, and T2g should appear as singlet, doublet, and triplet, respectively, in the Raman spectrum. The same turns out to be the case for the two molecules on the Cs site since an additional Davydov component is only infrared active. For this latter lattice site each component of the infrared modes should become Raman active. In the Raman spectrum of Salvi and Schettino22 the A1g mode shows two components which should be taken as arising from the C2h and Cs sites. For the Eg and T2g modes 4 and 3 components are observed, respectively, and could, by analogy, be taken as arising from the molecules on different sites. No evidence is found of Raman activation of the infrared modes or of infrared activation of the Raman modes. This implies that in the low temperature phase the molecular deformation of the SF6 octahedra is actually very small. The model used in the present paper appears, therefore, suited for the system under study. As a whole the internal Raman active components should classify as 14 Ag ⊕ 7 Bg in the factor group.
The SF6 phase diagram as a function of pressure has been explored by X-ray diffraction up to 32 GPa.23 A phase transition to the same crystal structure of the low temperature phase described by Cockcroft and Fitch31 has been observed at 2 GPa (phase II). Further transitions have been observed to occur at 12 GPa (phase III) and 24 GPa (phase IV) as evidenced by the appearance of new lines in the diffraction pattern and changes in the Raman spectrum. It can be remarked that the multiplet structure observed in phase III is identical with that reported by Salvi and Schettino at low temperature and ambient pressure,22 apart from the improved resolution of the low temperature spectrum. It can be argued that also at high pressure deviations from the octahedral structure of the SF6 units are very small. Therefore, a geometry belonging to the Oh point group symmetry32−35 has been adopted in the initial configuration of DFT calculations. Subsequently, the pressure effects on both the structural and vibrational properties of SF6 have been simulated by increasing the Pauli exchange repulsion experienced by the target molecule properly reducing the cavity volume using the XP-PCM method.4,13,16−18
Computational Protocol and Details
All the calculations have been performed in the DFT framework adopting the 6-311G(d)36 basis set with a modified version of the Gaussian 09 rev.C.01 suite of programs:37 The exchange–correlation energy functional has been selected by verifying the accuracy in reproducing the experimental S–F bond length with some exchange and correlation functionals.
The bond length of the optimized SF6 molecule (in gas phase) obtained with the selected exchange and correlation functional is reported in Table 1; the computed bond lengths are in the range between 1.588 Å (PBE0 and WB97XD) and 1.638 Å (BLYP).
Table 1. Bond Length (Å) of Isolated SF6 Molecule for Selected Exchange and Correlation Functionals, Using the 6-311Gd) Basis Seta.
functional | S–F (Å) |
---|---|
exp38 | 1.561 |
SVWN39−42 | 1.591 |
BLYP43,44 | 1.638 |
B3LYP44,45 | 1.603 |
CAM-B3LYP46 | 1.589 |
O3LYP47 | 1.600 |
PBE48,49 | 1.622 |
PBE050 | 1.588 |
WB97XD51 | 1.588 |
The computed SF6 bond length is compared with experimental data in the gas phase.38
PBE0 and WB97XD provide results in better agreement with experiments.38 Therefore, we have decided to report the structural and spectroscopic analysis on the basis of calculations with the PBE0 functional, unless otherwise specified.
The XP-PCM protocol is the same employed with success in the studies on P4S317 and on As4S4.18 Argon has been used as solvent, while van der Waals radius for S and F have been set to 1.80 and 1.47 Å, respectively.28 The scaling factor f has been varied between 1.2 and 0.975, allowing to study the system in the pressure range between 0 and 10.2 GPa for N = 3 and between 0.0 and 24.6 GPa for N = 6, respectively. To ensure a smooth behavior of electronic and structural properties with pressure, the molecular cavities have been described using tesserae with an average size of 0.05 Å2 (the value has been imposed with the keyword TSARE and it is about an order of magnitude lower than the default value which is 0.2 Å2). However, it has been verified that a value of 0.075 Å2 represents a good compromise between accuracy and computational time for the present calculations.
Analogously to XP-PCM studies on P4S3,17 and As4S4,18 the protocol consists in a molecular geometry optimization performed with a self-consistent procedure until the structure convergence is reached for each scaling factor f (i.e., pressure value). Once the equilibrium geometry has been located, the procedure is repeated for a new cavity with a smaller scaling factor. The convergence criteria imposed in the calculations are Maximum force lower than 2 × 10–6Eh/a0, RMS force lower than 1 × 10–6Eh/a0, Maximum atomic displacement lower than 6 × 10–6a0 and RMS atomic displacement lower than 4 × 10–6a0. An ultrafine grid has been adopted for the evaluation of the numerical integrals. All the calculations on SF6 have been performed with values for the N parameter equal to both 3 and 6, as discussed in the following. This choice of using the limiting values for N is based on previous works4,13,15−17,52 on XP-PCM that proved how it was possible to interpret the experimental observable with pressure. No fitting procedures have been carried out to further improve the agreement between experimental and computed data.
The harmonic vibrational frequency calculation has been carried out for all the optimized geometries. The pressure effects on vibrational normal modes have been analyzed to determine both the curvature and relaxation contributions. The curvature contribution, eq 12, has been evaluated through the calculation of the vibrational frequencies, obtained by numerical differentiation, as a function of pressure at fixed equilibrium geometry in vacuo. The relaxation contribution can be evaluated numerically by eq 11 or analytically by determining the derivative of frequency variation with pressure as in eq 13.
Results and Discussion
IR and Raman Spectra at Ambient Pressure
The isolated SF6 molecule belongs to the Oh point group and consequently the vibrational normal modes can be grouped in the irreducible representations as
![]() |
14 |
The normal modes with symmetry A1g, Eg, and T2g are Raman active, whereas the two T1u modes are IR active. The normal mode of T2u symmetry is both IR and Raman inactive, but it has been identified and assigned by Salvi et al.22 analyzing combination bands in the IR spectrum of the crystal.
Calculated and experimental frequencies are compared in Table 2. It can be noted that the scaled calculated frequencies agree quite satisfactorily with the crystal frequencies.22,23 This may imply that the XP-PCM approach somehow accounts for the static field acting on the molecule albeit not (obviously) for the intermolecular coupling giving rise to the Davydov splitting.
Table 2. Normal Modes (cm–1) of SF6a.
symm | NM | scaled | expb | Δν | % | free | assignment |
---|---|---|---|---|---|---|---|
T2u | 328.2 | 346.9 | (345.7)c | 1.2 | 0.3 | 326.9 | ν6, antisymm degenerate bending |
T2g | 489.6 | 517.5 | 523.5 | 6 | 1.2 | 488.7 | ν5, symm degenerate bending |
T1u | 578.3 | 611.3 | (611) | 0.3 | 0.05 | 578.9 | ν4, antisymm degenerate bending |
Eg | 626.7 | 662.4 | 644 | 18.4 | 2.9 | 627.1 | ν2, symm degenerate stretching |
A1g | 729.6 | 771.2 | 775 | 3.8 | 0.5 | 728.0 | ν1, symm stretching |
T1u | 925.7 | 978.4 | (936) | 42.4 | 4.5 | 934.2 | ν3, antisymm degenerate stretching |
XP-PCM results with f = 1.20 at the PBE0/6-311G(d) level of theory. TSARE = 0.05. Vibrational frequencies (NM) have been scaled by a factor of 1.057, obtained through a fitting procedure for the calculated frequencies of the isolated molecule (without XP-PCM).
P.R. Salvi and V. Schettino.22
This frequency has been obtained from the analysis of combination bands. Δν represents the absolute error in cm–1, while with % are reported the absolute error in percentage. The assignment has been taken from Herzberg.53 Symmetries for the “free” molecule calculated frequencies at the same XP-PCM level of theory are reported.
The vibrational assignment is that reported by Herzberg53 and confirmed in several studies of the vibrational properties of SF6.22−26,54,55 In particular, the three lower vibrational frequencies are bending normal modes, whereas the others are stretching modes.
A graphical representation of the normal modes is reported in Figure 2.
Figure 2.
Graphical representation of the normal modes of SF6 with the assignment proposed by Salvi et al.22 Only one component of the each degenerate modes is shown.
Structural Properties
The first information on the response to pressure of SF6 molecule has been achieved by computing the Murnaghan27 and Birch–Murnaghan56 equation of state and comparing the results with experimental data.23
The relation between the scaling factor f and the pressure can be obtained assuming that the cavity volume can be initially described through the Murnaghan equation:27
![]() |
15 |
in which V0 is
the cavity volume for f = 1.20 and V(f) is the cavity volume for a certain scaling factor,
whereas and
.
The K0 and Kp parameters are obtained by using a fitting procedure of the electronic energy functional:
![]() |
16 |
A similar procedure has been performed using the Birch–Murnaghan equation of state:56
![]() |
17 |
The values of both the K0 and Kp Murnaghan parameters obtained by XP-PCM calculations are compared in Table 3 with experimental data by Rademacher et al.23 for both phase I and phase II. The XP-PCM calculations have been carried out with the N parameter equal to both 3 and 6. The Birch–Murnaghan equation of state obtained for phase II by Rademacher et al.23 with periodic DFT calculations (PBE+TS) provides results in excellent agreement with experiments.
Table 3. Comparison of K0 and Kp Murnaghan Parameters Obtained by XP-PCM Calculations on SF6 and Experiments23.
N = 3 | N = 6 | exp phase I | exp phase II | |
---|---|---|---|---|
K0 | 6.46 | 7.79 | 6.3(2) | 8.5(8) |
Kp | 5.01 | 7.19 | 4 | 7.4(9) |
The comparison of the fitting procedure with both the Murnaghan27 and Birch–Murnaghan56 equations of state is reported in Figure 3. It is possible to appreciate from the Figure 3 that the curves of the energy as function of the cavity volume are accurately reproduced with both the Murnaghan27 and Birch–Murnaghan56 equations of state, with appreciable differences only outside the range of the computed data (cavity volume <65 Å3).
Figure 3.
Fit of energy values with both Murnaghan27 and Birch–Murnaghan56 equations of state. The procedure has been carried out for XP-PCM calculations with both N = 3 and N = 6.
The fitting procedure has been carried out employing both the N = 3 and N = 6 values in the XP-PCM calculations, as in Table 3.
Since the SF6 molecular crystal undergoes phase transitions with pressure,23 an appropriate tuning of the N parameter allows to follow correctly the structural behavior in different pressure ranges, as it can be appreciated in Figure 4 where both computed and experimental Murnaghan27 equations of state are reported. The computed Murnaghan equation of state with N = 3 closely resemble the experimental one for phase I, whereas the XP-PCM calculations with N = 6 reproduce the behavior of phase II.
Figure 4.
Comparison between experimental and computed Murnaghan27 equation of states with volume. Cavity volume and molecular volume have been adopted for calculations and experiments, respectively.
Pressure Effects on Bond Lengths
The different effect due to the N parameter with increasing pressure provides further information on the compressibility of SF6. It is interesting to note (see Figure 5) that N = 3 and N = 6 parameters present a similar behavior for low pressure values, whereas a marked difference occurs for higher values, as confirmed by linear regression analysis. In fact, the correlation coefficients for the linear regression equations for PBE are R2 = 0.9907 and R2 = 0.9870 for N = 3 and N = 6, respectively, while the same coefficients for PBE0 are R2 = 0.9923 and R2 = 0.9989. By comparing the results obtained with both the PBE0 and PBE exchange and correlation functionals, it is possible to note that hybrid functionals provide a better agreement than GGA functionals with the experimental bond length38 and a similar trend with pressure. The bond length contraction has the same order of magnitude, but it presents a slightly underestimated value.23 Although marked volume variations are observed with the choice of N parameter as shown in Figure 3, the variation of the bond lengths are less affected by the N value. This behavior reflects on the calculated frequencies which are usually more accurate with PBE0 than PBE. This result further suggests to use hybrid functionals in simulation of structural and vibrational properties by performing calculations with the XP-PCM method.
Figure 5.
S–F bond length (in Å) with the pressure (in GPa). The results are reported for both PBE and PBE0 exchange and correlation functionals in conjunction with 6-311G(d) basis set. The equations of linear regression are PBE N3, y = −0.00068x + 1.62246; PBE N6, y = −0.00063x + 1.62216; PBE0 N3, y = −0.00070x + 1.58786; and PBE0 N6, y = −0.00055x + 1.58770.
As previously observed
in studying other systems with the XP-PCM
method,4,13,16,17 the effect of pressure on SF6 molecular
structure can be analyzed considering eq 8. In fact, the variation of molecular structure with
pressure are related both to pressure coupling coefficient Γi and to the stiffness of the normal modes.
This analysis can be carried out through two different approaches.
In the first case, the pressure coupling Γi (defined in eq 7) is estimated by a linear fitting of the nuclear gradient of Ger as function of pressure
evaluated at the equilibrium geometry of the isolated molecule. The
force constant ki is
determined from the calculation of the harmonic frequencies of the
isolated molecule. In the second case, the linear coefficient () has been obtained by
the projection of
the atomic displacement on the eigenvector of the totally symmetric
normal mode, as function of pressure. The two procedures have been
performed for N = 3 and N = 6, and
the final results, collected in Table 4, allow to verify eq 8.
Table 4. Pressure Coupling Constant Γi (Eha0–1 GPa–1), Defined in Eq 7, Harmonic Force Constant in the Gas Phase ki (Eha0–2), Ratio Γi/ki (Å GPa–1), and Pressure Coefficient dQi/dp (Å GPa–1) of SF6.
Γi | ki | ![]() |
![]() |
|
---|---|---|---|---|
N = 3 | 0.00123 | 0.7201 | –0.00173 | –0.00171 |
N = 6 | 0.00099 | 0.7201 | –0.00137 | –0.00134 |
Pressure Effects on the Vibrational Modes
The pressure effects on the vibrational frequencies have been obtained from calculations for the SF6 pressure optimized geometrical structures (scaling factor, f) for both the N = 3 and N = 6 parameters. The vibrational frequencies of the computed normal modes of SF6 as a function of pressure are collected in Table 5 and 6.
Table 5. Calculated Frequencies (in cm–1) at the Different Pressure Values (GPa) with Parameter N = 3.
symmetry | 0.0 | 0.7 | 1.9 | 2.8 | 3.9 | 5.4 | 7.4 | 10.2 |
---|---|---|---|---|---|---|---|---|
T2u | 346.9 | 348.4 | 350.9 | 352.9 | 355.7 | 359.2 | 364.2 | 371.2 |
T2g | 517.5 | 518.6 | 520.8 | 522.3 | 524.6 | 527.3 | 531.5 | 537.7 |
T1u | 611.3 | 611.7 | 612.8 | 613.6 | 615.0 | 616.5 | 619.1 | 622.8 |
Eg | 662.4 | 663.3 | 665.5 | 666.5 | 668.8 | 671.4 | 675.7 | 682.1 |
A1g | 771.2 | 772.6 | 775.5 | 777.1 | 780.0 | 783.0 | 788.0 | 795.7 |
T1u | 978.5 | 977.3 | 977.2 | 976.7 | 977.5 | 978.0 | 980.2 | 985.1 |
Table 6. Calculated Frequencies at the Different Pressure Values (GPa) with Parameter N = 6.
symmetry | 0.0 | 1.0 | 3.0 | 4.6 | 7.0 | 10.6 | 16.0 | 24.6 |
---|---|---|---|---|---|---|---|---|
T2u | 346.9 | 348.7 | 352.4 | 355.6 | 360.6 | 367.2 | 377.0 | 391.6 |
T2g | 517.5 | 518.8 | 521.9 | 524.5 | 528.7 | 534.0 | 542.3 | 555.3 |
T1u | 611.3 | 611.9 | 613.8 | 615.3 | 618.0 | 621.5 | 627.1 | 635.6 |
Eg | 662.4 | 663.6 | 667.0 | 669.3 | 673.7 | 679.3 | 688.5 | 702.8 |
A1g | 771.2 | 772.9 | 777.2 | 779.7 | 785.0 | 791.2 | 801.1 | 817.3 |
T1u | 978.5 | 977.6 | 978.6 | 979.1 | 982.2 | 985.4 | 992.3 | 1005.8 |
Comparison with Experiments
The experimental trend of vibrational frequencies of the T2g, Eg, and A1g modes as a function of pressure is compared with calculations in Figure 6, Figure 7, and Figure 8, respectively. The experimental data refer the measuremet results for phases II, III, and IV, while it has not been reported the vibrational frequencies relative to phase I.
Figure 6.
Experimental and computed vibrational frequencies of normal mode ν5(T2g). The XP-PCM calculations have been carried out with both N = 3 and N = 6. The experimental data have been taken by Rademacher et al.23
Figure 7.
Experimental and computed vibrational frequencies of normal mode ν2(Eg). The XP-PCM calculations have been carried out with both N = 3 and N = 6. The experimental data have been taken by Rademacher et al.23
Figure 8.
Experimental and computed vibrational frequencies of normal mode ν1(A1g). The XP-PCM calculations have been carried out with both N = 3 and N = 6. The experimental data have been taken by Rademacher et al.23
The pressure effect on vibrational frequencies
has been analyzed
by the numerical calculation of the frequency slope, , which differs for the
various normal modes
as found experimentally.23 The model depends
by the choice of N parameter as discussed above.
However, either N parameters provide a semiquantitative
description of the experimenatl data, also if the computed values
underestimate the experimental results, as can be appreciated in Table 7. A similar behavior
has been obtained in studying C60 and C70 fullerenes
under pressure.16
Table 7. Comparison
between Computed (for N = 3 and N = 6) and Experimental
in
a.
freq | symm | N = 3 | N = 6 | phase II | phase III | phase IV |
---|---|---|---|---|---|---|
517.5 | T2g | 1.98 | 1.55 | 2.73 ± 0.47 | 2.17 ± 0.61 | 1.18 ± 0.22 |
662.4 | Eg | 1.92 | 1.65 | 3.83 ± 0.31 | 3.13 ± 0.04 | 1.58 ± 0.08 |
771.2 | A1g | 2.38 | 1.87 | 4.64 ± 0.33 | 3.55 ± 0.22 | 1.86 ± 0.08 |
The experimental data refer to Raman measurements by Rademacher et al.23 for the different SF6 phases at high pressure.
Analysis of Pressure Effects on the Vibrational Frequencies
In this section we present
a detailed analysis of the vibrational
frequencies as a function of pressure in terms of curvature (direct)
and relaxation (indirect) contributions, as given from eq 11 to eq 13. The curvature contribution has been obtained
by calculating the vibrational frequencies of SF6 keeping
the geometry fixed to that of the isolated molecule but reducing the
cavity volume to increase the pressure. The term has been subsequently obtained through
a fitting procedure of the vibrational frequencies, which have been
reported in Table 8.
Table 8. Variations of Vibrational Frequencies (in cm–1) with Pressure (GPa) for N = 3 and N = 6 Parametersa.
N = 3 |
||||||
---|---|---|---|---|---|---|
freq | symm | g̃iij | ![]() |
![]() |
![]() |
![]() |
346.9 | T2u | –10.5 | 2.39 | –0.19 | 0.18 | 2.58 |
517.5 | T2g | –34.5 | 1.98 | 0.38 | 0.59 | 1.59 |
611.3 | T1u | –37.0 | 1.14 | 0.5 | 0.64 | 0.63 |
662.4 | Eg | –76.5 | 1.92 | 1.22 | 1.31 | 0.70 |
771.2 | A1g | –98.5 | 2.38 | 1.62 | 1.70 | 0.76 |
978.5 | T1u | –108.5 | 0.65 | 2.15 | 1.87 | –1.50 |
N = 6 |
||||||
---|---|---|---|---|---|---|
freq | symm | g̃iij | ![]() |
![]() |
![]() |
![]() |
346.9 | T2u | –10.5 | 1.83 | –0.12 | 0.15 | 1.95 |
517.5 | T2g | –34.5 | 1.55 | 0.31 | 0.47 | 1.24 |
611.3 | T1u | –37.0 | 1.00 | 0.37 | 0.50 | 0.63 |
662.4 | Eg | –76.5 | 1.65 | 0.96 | 1.04 | 0.70 |
771.2 | A1g | –98.5 | 1.87 | 1.29 | 1.35 | 0.58 |
978.5 | T1u | –108.5 | 1.14 | 1.61 | 1.48 | –0.47 |
Freq and symm refer to normal mode frequencies (in cm–1) and symmetries, respectively. g̃iij represents the cubic force constant (in cm–1). cur and rel subscripts indicate the curvature and relaxation terms, while anal and num superscripts refer to the calculation approaches.
The relaxation contribution has been obtained by both the numeric and analytic approaches, described in “Computational Protocol and Details” section. The results with N = 3 and N = 6 are summarized in Table 8, while a graphic representation is shown in Figure 9.
Figure 9.
Vibrational analysis in terms of curvature (orange) and
relaxation
(green) contribution to (blue). The upper panel
refers to the results
obtained with N = 3, while the lower panel refers
to the results obtained with N = 6.
The results of the vibrational frequency analysis can be
summarized
considering both the relaxation and curvature contributions. Independent
of the value of N parameter, the relaxation term, , computed with both numerical and analytic
approaches, grows with frequency of the normal modes at the expenses
of the curvature term. The results obtained with either approaches,
are similar and show that the
is positive with the exception of the lower
frequency normal mode with symmetry T2u. The same value is found for
; it can only be noted that for N = 6 the value
for T2u mode is negative,
but very small. In particular, the relaxation
term is related to the anharmonicity through the cubic force constants
as discussed in eq 13. In particular, the SF6 molecules has only one totally
symmetric normal mode (ν1 of Figure 2), consequently the relaxation contribution
has to be ascribed to the cubic anharmonic constants giij in eq 13; the anharmonic constants giij are reported in Table 8. The anharmonic character of the normal
modes is more pronounced for the higher than for the lower frequency
modes and it is more pronounced for stretching modes. This increasing
of the anharmonicity is the origin of the growth of the relaxation
term with frequencies.
The curvature term, , shows positive values with the exception
of the highest frequency mode (ν3) with symmetry T1u. While the positive values
of curvature term present a behavior similar to that found in other
systems,16,17 the negative value of ν3 is rather unusual. The behavior of ν3 with pressure
has been rationalized considering the Pauli repulsion due to pressure
on the electronic structure of SF6, especially in the region
along the S–F bond. The electron density variation has been
analyzed in terms of Mayer bond order,57,58 atomic charges,59−63 and maps of the electron density difference. The different methods
concord in the observed results. In fact, the Mayer bond order decreases
with pressure from 0.907 to 0.893 going from 0 to 24.6 GPa with a
difference of 0.014; the charges on fluorine atoms increases while
that of sulfur decreases (independently by the method, as summarized
in Table 9). These
results confirm an increase in ionic character of the bond. A similar
result has been obtained by the difference of electron density maps
depicted in Figure 10. This behavior substantially differs from that found in studying
other systems,13,16,17 which show an increase of electron density on the bond regions.
In the case of SF6, the observed electron structure rearrangement
with pressure induces a more pronounced variation in the normal modes
with asymmetric displacement with respect to the S–F bond.
Table 9. S and F Atomic Charges (e) Computed through Mulliken,59−61 Löwdin,62 and Hirshfeld63 Partition Schemesa.
S |
F |
|||||
---|---|---|---|---|---|---|
0.0 GPa | 24.6 GPa | Δ | 0.0 GPa | 24.6 GPa | Δ | |
Mulliken | 1.576 | 1.617 | –0.041 | –0.263 | –0.270 | 0.007 |
Löwdin | 1.285 | 1.303 | –0.018 | –0.214 | –0.217 | 0.003 |
Hirshfild | 0.598 | 0.604 | –0.006 | –0.100 | –0.101 | 0.001 |
The differences on atomic charges (Δ) of both S and F atoms have been computed considering the system at the pressure values of 0.0 and 24.6 GPa.
Figure 10.
a) Electron density difference between SF6 at 0.0 and 24.6 GPa (cutoff ±3 × 10–5). b) The red and blue colors on the map represent regions with increase and decrease of electron density, respectively. The isosurface cutoff is ±1.5 × 10–5.
Conclusions
The XP-PCM method has been applied to compute
the structural and
vibrational properties of SF6 with pressure. The results
of the XP-PCM method have been compared with both structural and spectroscopic
experimental data, showing a resonable agreement between experiments
and calculations. Pressure effects on the vibrational frequencies
have been analyzed by considering both the contribution of curvature, , and relaxation,
, obtaining useful information for discriminating
the importance of confinement and anharmonicity effects. In particular,
the curvature term for the ν3 normal modes has been
rationalized considering the variation of electronic structure of
SF6 with pressure. In fact, it has been observed that this
behavior can be related to the increase of ionic character of the
S–F bond under pressure. These results show how the XP-PCM
method could provide further useful information for the interpretation
of the experimental findings. The different pressure effects in determining
a more pronounced covalent or ionic character of bonds in a molecule
represent an interesting point to be analyzed in more detail to rationalize
the trend of vibrational frequencies with pressure not only using
the XP-PCM method but other single molecule computational approaches.6,64−69 Finally, it would be useful to extend this kind of analysis in periodic
DFT calculations on molecular crystalline systems, which explicitly
describe the intermolecular interactions to achieve further information
on spectroscopic properties.
Acknowledgments
The authors (M.B., M.P., G.C., and V.S.) thank MIUR-Italy (“Progetto Dipartimenti di Eccellenza 2018-2022” allocated to Department of Chemistry “Ugo Schiff”). R.C. is thankful for support the framework of the COMP-HUB Initiative of the Department of Chemistry, Life Sciences and Environmental Sustainability of the University of Parma, as part of the ‘Departments of Excellence’ program of the Italian Ministry for Education, University and Research (MIUR, 2018-2022) .
The authors declare no competing financial interest.
References
- Schettino V.; Bini R. Constraining Molecules at the Closest Approach: Chemistry at High Pressure. Chem. Soc. Rev. 2007, 36, 869–880. 10.1039/b515964b. [DOI] [PubMed] [Google Scholar]
- Bini R.; Schettino V.. Materials Under Extreme Conditions; Imperial College Press, 2014. [Google Scholar]
- Li B.; Ji C.; Yang W.; Wang J.; Yang K.; Xu R.; Liu W.; Cai Z.; Chen J.; Mao H.-k. Diamond anvil cell behavior up to 4 Mbar. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 1713–1717. 10.1073/pnas.1721425115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cammi R. In Frontiers of Quantum Chemistry; Wójcik M. J., Nakatsuji H., Kirtman B., Ozaki Y., Eds.; Springer Singapore: Singapore, 2018; pp 273–287. [Google Scholar]
- Grochala W.; Hoffmann R.; Feng J.; Ashcroft N. The Chemical Imagination at Work in Very Tight Places. Angew. Chem., Int. Ed. 2007, 46, 3620–3642. 10.1002/anie.200602485. [DOI] [PubMed] [Google Scholar]
- Stauch T. Quantum chemical modeling of molecules under pressure. Int. J. Quantum Chem. 2021, 121, e26208. 10.1002/qua.26208. [DOI] [PubMed] [Google Scholar]
- Chen B.; Hoffmann R.; Cammi R. The Effect of Pressure on Organic Reactions in Fluids–a New Theoretical Perspective. Angew. Chem., Int. Ed. 2017, 56, 11126–11142. 10.1002/anie.201705427. [DOI] [PubMed] [Google Scholar]
- Cammi R.; Rahm M.; Hoffmann R.; Ashcroft N. W. Varying Electronic Configurations in Compressed Atoms: From the Role of the Spatial Extension of Atomic Orbitals to the Change of Electronic Configuration as an Isobaric Transformation. J. Chem. Theory Comput. 2020, 16, 5047–5056. 10.1021/acs.jctc.0c00443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Califano S.; Schettino V.; Neto N.. Lattice Dynamics of Molecular Crystals; Springer: Berlin, Heidelberg, 1981. [Google Scholar]
- Martin R. M.Electronic Structure: Basic Theory and Practical Methods; Cambridge University Press, 2004. [Google Scholar]
- Marx D.; Hutter J.. Ab Inito Molecular Dynamics: Basic Theory and Advanced Methods; Cambridge Universty Press, 2009. [Google Scholar]
- Wang Y.; Shang S.-L.; fang H.; Liu Z.-K.; Chen L.-Q. First-principles calculations of lattice dynamics and thermal properties of polar solids. npj Comput. Mater. 2016, 2, 16006. 10.1038/npjcompumats.2016.6. [DOI] [Google Scholar]
- Cammi R.; Cappelli C.; Mennucci B.; Tomasi J. Calculation and analysis of the harmonic vibrational frequencies in molecules at extreme pressure: Methodology and diborane as a test case. J. Chem. Phys. 2012, 137, 154112. 10.1063/1.4757285. [DOI] [PubMed] [Google Scholar]
- Fukuda R.; Ehara M.; Cammi R. Modeling Molecular Systems at Extreme Pressure by an Extension of the Polarizable Continuum Model (PCM) Based on the Symmetry-Adapted Cluster-Configuration Interaction (SACCI) Method: Confined Electronic Excited States of Furan as a Test Case. J. Chem. Theory Comput. 2015, 11, 2063–2076. 10.1021/ct5011517. [DOI] [PubMed] [Google Scholar]
- Cammi R. A new extension of the polarizable continuum model: Toward a quantum chemical description of chemical reactions at extreme high pressure. J. Comput. Chem. 2015, 36, 2246–2259. 10.1002/jcc.24206. [DOI] [PubMed] [Google Scholar]
- Pagliai M.; Cardini G.; Cammi R. Vibrational frequencies of fullerenes C60 and C70 under pressure studied with a quantum chemical model including spatial confinement effects. J. Phys. Chem. A 2014, 118, 5098–5111. 10.1021/jp504173k. [DOI] [PubMed] [Google Scholar]
- Pagliai M.; Cammi R.; Cardini G.; Schettino V. XP-PCM Calculations of High Pressure Structural and Vibrational Properties of P4S3. J. Phys. Chem. A 2016, 120, 5136–5144. 10.1021/acs.jpca.6b00590. [DOI] [PubMed] [Google Scholar]
- Caratelli C.; Cammi R.; Chelli R.; Pagliai M.; Cardini G.; Schettino V. Insights on the Realgar Crystal Under Pressure from XP-PCM and Periodic Model Calculations. J. Phys. Chem. A 2017, 121, 8825–8834. 10.1021/acs.jpca.7b08868. [DOI] [PubMed] [Google Scholar]
- Tomasi J.; Persico M. Molecular Interactions in Solution: An Overview of Methods Based on Continuous Distributions of the Solvent. Chem. Rev. 1994, 94, 2027–2094. 10.1021/cr00031a013. [DOI] [Google Scholar]
- Tomasi J.; Mennucci B.; Cammi R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. 10.1021/cr9904009. [DOI] [PubMed] [Google Scholar]
- Labet V.; Hoffmann R.; Ashcroft N. W. A Fresh Look at Dense Hydrogen under Pressure. II. Chemical and Physical Models Aiding our Understanding of Evolving H-H Separations. J. Chem. Phys. 2012, 136, 074502. 10.1063/1.3679736. [DOI] [PubMed] [Google Scholar]
- Salvi P.; Schettino V. Infrared and Raman spectra and phase transition of the SF6 crystal. Anharmonic interactions and two-phonon infrared absorption. Chem. Phys. 1979, 40, 413–424. 10.1016/0301-0104(79)85154-X. [DOI] [Google Scholar]
- Rademacher N.; Friedrich A.; Morgenroth W.; Bayarjargal L.; Milman V.; Winkler B. High-pressure phases of SF6 up to 32 GPa from X-ray diffraction and Raman spectroscopy. J. Phys. Chem. Solids 2015, 80, 11–21. 10.1016/j.jpcs.2014.12.016. [DOI] [Google Scholar]
- Dows D. A.; Wieder G. M. Infrared intensities in crystalline SF6. Spectrochim. Acta 1962, 18, 1567–1574. 10.1016/S0371-1951(62)80284-7. [DOI] [Google Scholar]
- Rubin B.; McCubbin T.K.; Polo S.R. Vibrational Raman Spectrum of SF6. J. Mol. Spectrosc. 1978, 69, 254–259. 10.1016/0022-2852(78)90063-2. [DOI] [Google Scholar]
- Sasaki S.; Tomida Y.; Shimizu H. High-Pressure Raman Study of Sulfur Hexafluoride up to 10 GPa. J. Phys. Soc. Jpn. 1992, 61, 514–518. 10.1143/JPSJ.61.514. [DOI] [Google Scholar]
- Murnaghan F. D. The Compressibility of Media under Extreme Pressures. Proc. Natl. Acad. Sci. U. S. A. 1944, 30, 244–247. 10.1073/pnas.30.9.244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alvarez S. A cartography of the van der Waals territories. Dalton Trans. 2013, 42, 8617–8636. 10.1039/c3dt50599e. [DOI] [PubMed] [Google Scholar]
- Rahm M.; Ångqvist M.; Rahm J. M.; Erhart P.; Cammi R. Non-Bonded Radii of the Atoms Under Compression. ChemPhysChem 2020, 21, 2441–2453. 10.1002/cphc.202000624. [DOI] [PubMed] [Google Scholar]
- Moroni L.; Ceppatelli M.; Gellini C.; Salvi P. R.; Bini R. Excitation of Crystalline All Trans Retinal under Pressure. Phys. Chem. Chem. Phys. 2002, 4, 5761–5767. 10.1039/B207312A. [DOI] [Google Scholar]
- Cockcroft J. K.; Fitch A. N. The solid phases of sulphur hexafluoride by powder neutron diffraction. Z. Kristallogr. 1988, 184, 123–145. 10.1524/zkri.1988.184.1-2.123. [DOI] [Google Scholar]
- Dolling G.; Powell B.; Sears V. Neutron diffraction study of the plastic phases of polycrystalline SF6 and CBr4. Mol. Phys. 1979, 37, 1859–1883. 10.1080/00268977900101381. [DOI] [Google Scholar]
- Dove M. T.; Pawley G. S. A molecular dynamics simulation study of the plastic crystalline phase of sulphur hexafluoride. J. Phys. C: Solid State Phys. 1983, 16, 5969–5983. 10.1088/0022-3719/16/31/012. [DOI] [Google Scholar]
- Dove M. T.; Pawley G. S. A molecular dynamics simulation study of the orientationally disordered phase of sulphur hexafluoride. J. Phys. C: Solid State Phys. 1984, 17, 6581–6599. 10.1088/0022-3719/17/36/014. [DOI] [Google Scholar]
- Dove M.; Tucker M.; Keen D. Neutron total scattering method: simultaneous determination of long-range and short-range order in disordered materials. Eur. J. Mineral. 2002, 14, 331–348. 10.1127/0935-1221/2002/0014-0331. [DOI] [Google Scholar]
- Spitznagel G. W.; Clark T.; Schleyer P. v.; Hehre W. J. An Evaluation of the Performance of Diffuse Function-Augmented Basis Sets for Second Row Elements, Na-Cl. J. Comput. Chem. 1987, 8, 1109–1116. 10.1002/jcc.540080807. [DOI] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. A.; et al. Gaussian’09 Revision C.01; Gaussian Inc.: Wallingford, CT, 2010. [Google Scholar]
- Bartell L.; Doun S. Structures of hexacoordinate compounds of main-group elements: Part III. An electron diffraction study of SF6. J. Mol. Struct. 1978, 43, 245–249. 10.1016/0022-2860(78)80010-6. [DOI] [Google Scholar]
- Hohenberg P.; Kohn W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. 10.1103/PhysRev.136.B864. [DOI] [Google Scholar]
- Kohn W.; Sham L. J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. 10.1103/PhysRev.140.A1133. [DOI] [Google Scholar]
- Slater J. C.The Self-Consistent Field for Molecular and Solids, Quantum Theory of Molecular and Solids; McGraw-Hill: New York, 1974; Vol. 4. [Google Scholar]
- Vosko S. H.; Wilk L.; Nusair M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: a critical analysis. Can. J. Phys. 1980, 58, 1200–1211. 10.1139/p80-159. [DOI] [Google Scholar]
- Becke A. D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098–3100. 10.1103/PhysRevA.38.3098. [DOI] [PubMed] [Google Scholar]
- Lee C.; Yang W.; Parr R. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron-Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785–789. 10.1103/PhysRevB.37.785. [DOI] [PubMed] [Google Scholar]
- Becke A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. 10.1063/1.464913. [DOI] [Google Scholar]
- Yanai T.; Tew D. P.; Handy N. C. A new hybrid exchange-correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. 10.1016/j.cplett.2004.06.011. [DOI] [Google Scholar]
- Cohen A. J.; Handy N. C. Dynamic correlation. Mol. Phys. 2001, 99, 607–615. 10.1080/00268970010023435. [DOI] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple [Phys. Rev. Lett. 77, 3865 (1996)]. Phys. Rev. Lett. 1997, 78, 1396–1396. 10.1103/PhysRevLett.78.1396. [DOI] [PubMed] [Google Scholar]
- Adamo C.; Barone V. Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0Model. J. Chem. Phys. 1999, 110, 6158–6169. 10.1063/1.478522. [DOI] [Google Scholar]
- Chai J.-D.; Head-Gordon M. Long-range corrected hybrid density functionals with damped atom-atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. 10.1039/b810189b. [DOI] [PubMed] [Google Scholar]
- Cammi R.; Verdolino V.; Mennucci B.; Tomasi J. Towards the Elaboration of a QM Method to Describe Molecular Solutes under Effect of a Very High Pressure. Chem. Phys. 2008, 344, 135–141. 10.1016/j.chemphys.2007.12.010. [DOI] [Google Scholar]
- Herzberg G.Molecular Spectra and Molecular Structure II. Infrared and Raman Spectra of Polyatomic Molecules; Krieger: Malabar, FL, 1991. [Google Scholar]
- Edelson D.; McAfee K. Note on the infrared spectrum of sulfur hexafluoride. J. Chem. Phys. 1951, 19, 1311–1312. 10.1063/1.1748022. [DOI] [Google Scholar]
- Wagner N. L.; Wüest A.; Christov I. P.; Popmintchev T.; Zhou X.; Murnane M. M.; Kapteyn H. C. Monitoring molecular dynamics using coherent electrons from high harmonic generation. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 13279–13285. 10.1073/pnas.0605178103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Birch F. Finite Elastic Strain of Cubic Crystals. Phys. Rev. 1947, 71, 809–824. 10.1103/PhysRev.71.809. [DOI] [Google Scholar]
- Mayer I. Charge bond order and valence in the AB initio SCF theory. Chem. Phys. Lett. 1983, 97, 270–274. 10.1016/0009-2614(83)80005-0. [DOI] [Google Scholar]
- Mayer I. On bond orders and valences in the Ab initio quantum chemical theory. Int. J. Quantum Chem. 1986, 29, 73–84. 10.1002/qua.560290108. [DOI] [Google Scholar]
- Mulliken R. S. Electronic Population Analysis on LCAO-MO Molecular Wave Functions. I. J. Chem. Phys. 1955, 23, 1833–1840. 10.1063/1.1740588. [DOI] [Google Scholar]
- Mulliken R. S. Electronic Population Analysis on LCAOMO Molecular Wave Functions. II. Overlap Populations, Bond Orders, and Covalent Bond Energies. J. Chem. Phys. 1955, 23, 1841–1846. 10.1063/1.1740589. [DOI] [Google Scholar]
- Mulliken R. S. Criteria for the Construction of Good Self-Consistent-Field Molecular Orbital Wave Functions, and the Significance of LCAO-MO Population Analysis. J. Chem. Phys. 1962, 36, 3428–3439. 10.1063/1.1732476. [DOI] [Google Scholar]
- Löwdin Per-Olov On the Non-Orthogonality Problem Connected with the Use of Atomic Wave Functions in the Theory of Molecules and Crystals. J. Chem. Phys. 1950, 18, 365–375. 10.1063/1.1747632. [DOI] [Google Scholar]
- Hirshfeld F. Bonded-atom fragments for describing molecular charge densities. Theoret. Chim. Acta 1977, 44, 129–138. 10.1007/BF00549096. [DOI] [Google Scholar]
- Subramanian G.; Mathew N.; Leiding J. A generalized force-modified potential energy surface for mechanochemical simulations. J. Chem. Phys. 2015, 143, 134109. 10.1063/1.4932103. [DOI] [PubMed] [Google Scholar]
- Zaleśny R.; Góra R. W.; Luis J. M.; Bartkowiak W. On the particular importance of vibrational contributions to the static electrical properties of model linear molecules under spatial confinement. Phys. Chem. Chem. Phys. 2015, 17, 21782–21786. 10.1039/C5CP02865E. [DOI] [PubMed] [Google Scholar]
- Spooner J.; Smith B.; Weinberg N. Effect of high pressure on the topography of potential energy surfaces. Can. J. Chem. 2016, 94, 1057–1064. 10.1139/cjc-2016-0295. [DOI] [Google Scholar]
- Stauch T. A mechanochemical model for the simulation of molecules and molecular crystals under hydrostatic pressure. J. Chem. Phys. 2020, 153, 134503. 10.1063/5.0024671. [DOI] [PubMed] [Google Scholar]
- Chouj M.; Basiak B.; Bartkowiak W. Partitioning of the interaction-induced polarizability of molecules in helium environments. Int. J. Quantum Chem. 2021, 121, e26544. 10.1002/qua.26544. [DOI] [Google Scholar]
- Scheurer M.; Dreuw A.; Epifanovsky E.; Head-Gordon M.; Stauch T. Modeling Molecules under Pressure with Gaussian Potentials. J. Chem. Theory Comput. 2021, 17, 583–597. 10.1021/acs.jctc.0c01212. [DOI] [PubMed] [Google Scholar]