An infrared absorption spectroscopy study of the endohedral water molecule in a solid mixture of H2O@C60 and C60 was carried out at liquid helium temperature. From the evolution of the spectra during the orthopara conversion process, the spectral lines were identified as para-H2O and ortho-H2O transitions. Eight vibrational transitions with rotational side peaks were observed in the mid-infrared: ω1, ω2, ω3, 2ω1, 2ω2, ω1 + ω3, ω2 + ω3, and 2ω2 + ω3. The vibrational frequencies ω2 and 2ω2 are lower by 1.6% and the rest by 2.4%, as compared to those of free H2O. A model consisting of a rovibrational Hamiltonian with the dipole and quadrupole moments of H2O interacting with the crystal field was used to fit the infrared absorption spectra. The electric quadrupole interaction with the crystal field lifts the degeneracy of the rotational levels. The finite amplitudes of the pure v1 and v2 vibrational transitions are consistent with the interaction of the water molecule dipole moment with a lattice-induced electric field. The permanent dipole moment of encapsulated H2O is found to be 0.50 ± 0.05 D as determined from the far-infrared rotational line intensities. The translational mode of the quantized center-of-mass motion of H2O in the molecular cage of C60 was observed at 110 cm−1 (13.6 meV).

Endohedral fullerenes consist of atoms or molecules fully encapsulated in closed carbon cages. The remarkable synthetic route known as “molecular surgery” has led to the synthesis of atomic endofullerenes He@C601 and Ar@C602 and several molecular endofullerene species, including H2@C60,3 H2O@C60,4 HF@C60,5 CH4@C60,6 and their isotopologs.7–9 It is well established that these endohedral molecules do not form chemical bonds with the carbon cage and rotate freely.5,10,11 The rotation is further facilitated by the nearly spherical symmetry of the C60 cage. The thermal and chemical stability of A@C60 opens up the unique possibility of studying the dynamics and the interactions of a small molecule with carbon nano-surfaces. The trapping potential of dihydrogen H2 has been described with high accuracy using infrared (IR) spectroscopy,8,12 inelastic neutron scattering (INS),13 and theoretical calculations.14 The non-spherical shape of the small molecule and the quantized translational motion of its center of mass lead to coupled rotational and translational dynamics.10,15,16

The water-endofullerene H2O@C60 is of particular interest. The encapsulated water molecule possesses rich spatial quantum dynamics. It is an asymmetric-top rotor, supports three vibrational modes, displays nuclear spin isomerism (para-water and ortho-water), and has electric dipole and quadrupole moments.

Low-temperature dielectric measurements17 on solid H2O@C60 show that the electric dipole moment of the encapsulated water is reduced to 0.51 ± 0.05 D from the free water value of 1.85 D. The C60 carbon cage responds to the endohedral water molecule with a counteracting induced dipole, resulting in the lower total dipole moment.18 

The dynamics of isolated or encapsulated single water molecules have been studied before in other environments, such as noble gas matrices, solid hydrogen, and liquid helium droplets. Although the trapping sites in these matrices have high symmetry and allow water rotation, these systems exist only at low temperature19,20 or for a very short time.21,22 Water has also been studied in crystalline environments with nano-size cavities. However, in this case, the interactions with the trapping sites inhibit the free rotation of the water molecules.23–26 

Several spectroscopic techniques have been used to study H2O@C60, including nuclear magnetic resonance (NMR), INS, and IR spectroscopy4,11,27–29 and time-domain THz spectroscopy.30 The low-lying rotational states of the encapsulated molecule are found to be very similar to those of an isolated water molecule, with the notable exception of a 0.6 meV splitting in the J = 1 rotational state.11,31 This indicates that the local environment of the water molecule in H2O@C60 has a lower symmetry than the icosahedral point group of the encapsulating C60 cage. The splitting has been attributed to the interaction between the electric quadrupole moment of H2O and the electric field gradients generated by the electronic charge distribution of neighboring C60 molecules.32,33 The merohedral disorder present in solid C60 leads to two C60 sites with different quadrupolar interactions. This merohedral disorder also leads to splittings of the IR phonons in solid C60.34 There is also evidence from dielectric measurements that merohedral disorder leads to electric dipolar activity in solid C60.35 

Confined molecules exhibit quantization of their translational motion (“particle in a box”), in addition to their quantized rotational and vibrational modes. Quantized translational modes have been observed at 60–70 cm−1 for water in noble gas matrices.19,36 However, comparatively, little is known about the center-of-mass translational mode of H2O@C60. The fundamental frequency of the water translation mode in H2O@C60 has been predicted to occur at ∼160 cm−1.16 This relatively high frequency reflects the rather tight confinement of the water molecule in the C60 cage.

The energy level separation between the ground para rotational state and the lowest ortho rotational state in H2O@C60 has been determined to be 2.6 meV (21 cm−1) by INS.11 This energy level separation corresponds to a temperature of 28 K. The full thermal equilibration of H2O@C60 at temperatures below 30 K, therefore, requires the conversion of ortho-water into para-water. This conversion process takes between tens of minutes and several hours below 20 K in H2O@C60.11,17,30,37 The spin isomer conversion is much faster at ambient temperature, with a time constant of about 30 s reported for H2O@C60 dissolved in toluene.28 

In this paper, we report on a detailed low-temperature far-IR and mid-IR spectroscopic study of H2O@C60 and C60 solid mixtures. The IR technique allows us to measure the frequencies of rotational, vibrational, and translational modes and, from the line intensities, to determine the dipole moment of encapsulated water. In addition, the IR spectra reveal the interaction of endohedral water with the electrostatic fields present in solid C60.

The rest of this paper is organized as follows: Section II discusses the sample preparation, the recording procedures of the IR spectra, and the determination of the IR absorption cross sections for different filling factors, temperatures, and orthopara ratios. The quantum mechanical vibrating rotor model for the encapsulated water molecules is introduced in Sec. III A. We include in this model the interactions of the H2O electric dipole and quadrupole moments with the electrostatic fields present in solid C60. The theory of the IR line intensities is presented in Sec. III B. Section IV presents the measured IR spectra and the fitting of these data by the quantum mechanical model. The results are discussed in Sec. V, followed by a summary in Sec. VI.  Appendixes A–C contain a more detailed theory for the interaction of the water molecules with the electrostatic fields and the infrared radiation and more details on the fitting of the experimental data by the quantum mechanical model.

H2O@C60 was prepared by a multi-step synthetic route known as “molecular surgery.”4,9 The H2O-filled (number density N) and empty (number density N) C60 were mixed and co-sublimed to produce small solvent-free crystals with a filling factor f = N/(N + N). Five samples with filling factors f = 0.014, 0.052, 0.10, 0.18, and 0.80 were studied. The powdered samples were pressed into pellets under vacuum. The diameter of sample pellets was 3 mm, and the thickness d varied from 0.2 mm to 2 mm. The thinner samples were used in the mid-IR because of light scattering in the powder sample. The samples were thicker for lower filling factors and thinner for higher filling factors to avoid the saturation of absorption lines in the far-IR.

The far-IR measurements were done with a Martin–Puplett type interferometer and 3He cooled bolometer from 5 to 200 cm−1, as described in Ref. 11. The IR measurements between 600 and 12 000 cm−1 were performed with an interferometer Vertex 80v (Bruker Optics), as described in Ref. 12.

Two methods were used to record the H2O@C60 absorption spectra.

1. Method 1

The intensity through the sample, Is, was referenced to the intensity through a 3 mm diameter hole, I0. The sample was allowed to reach orthopara thermal equilibrium at a temperature of 30 or 45 K, and the temperature was rapidly reduced to 10 or to 5 K. The sample spectrum Is(t = 0) was recorded immediately after the temperature jump. Since the orthopara conversion process is slow, the ortho fraction was assumed to be preserved during the T jump, corresponding to the high temperature ortho fraction no ≈ 0.74. The absorption coefficient α was calculated from the ratio Tr = Is(t = 0)/I0 as α(t = 0) = −d−1 ln[(1 − R)−2Tr] where the factor (1 − R)2 with R = (η − 1)2(η + 1)−2 corrects for the losses of radiation, one reflection from the sample front and one from the back face. The refraction index of solid C60 was assumed to be given by η = 2.38 To identify para-water and ortho-water absorption peaks, the difference of two spectra was calculated, Δα = α(t = 0) − αt), where αt) is the absorption spectrum measured after the waiting time Δt. Only the para-H2O@C60 and ortho-H2O@C60 peaks show up in the differential absorption spectra, with the para-H2O@C60 and ortho-H2O@C60 peak amplitudes having different signs, ortho positive and para negative. This method was used for the far-IR and the mid-IR part of the spectrum.

2. Method 2

The sample was allowed to reach orthopara thermal equilibrium at a temperature of 30 or 45 K, leading to an ortho-rich state, as in the first method. The temperature was rapidly reduced to 10 or to 5 K, and a series of spectra were recorded at intervals of a few minutes starting immediately after the T jump, which continued until the orthopara equilibrium was reached. The differential absorption Δα = −d−1 ln[Is(0)/Is] was calculated, where Is(0) is the spectrum recorded immediately after the T jump and Is is the spectrum recorded when the low temperature equilibrium was reached. The equilibrium ortho fraction is ∼0.01 at 5 K. Method 2 was used for the far-IR part of the spectrum.

The absorption line area Aji(k) was determined by fitting the measured absorption αji(k)(ω) with Gaussian lineshape. i and j are the initial and final states of the transition, and k denotes para (k = p) or ortho (k = o) species. From these experimental line areas, Aji(k), a temperature and para (ortho) fraction independent line area Aji(k)¯ was calculated,

(1)

The population difference of initial and final states, pi(k)pf(k), is given by the sample temperature T, while the para (ortho) fraction nk depends on the history of the sample because of the orthopara conversion process. The normalized absorption line area Aji(k) was determined from the linear fit of Aji(k)¯(f) for each absorption line,

(2)

Thus, the normalized absorption line area Aji(k) is the absorption line area of a sample with a filling factor f = 1 and spin isomer fraction nk = 1 where all the population is in the initial state, pi(k)=1. We used Aji(k) to calculate a synthetic experimental spectrum for the spectral fit with the quantum mechanical model (Sec. IV B).

Furthermore, to compare the absorption cross sections of H2O@C60 in solid C60 and free H2O, a normalized absorption cross section σji(k) was obtained as

(3)

where η is the index of refraction of solid C60 and σji(k) is the absorption cross section of an endohedral water molecule [Eq. (20) in Sec. III B]. This H2O@C60 absorption cross section can be compared to the free water normalized cross section σji(k)=σji(k)nk1(pi(k)pj(k))1, where pi(k)pj(k) and nk(T) are given by the temperature of the water vapor in the experiment reporting σji(k).

We use the following Hamiltonian to model the endohedral water molecule in solid C60:

(4)

where HM=Hv+Hrot is the free-molecule rovibrational Hamiltonian and HES is the electrostatic interaction of H2O with the surrounding electric charges. The translational Hamiltonian HT consists of water center-of-mass kinetic and potential energies in the molecular cavity of the C60 molecule.

We neglect couplings between vibrational modes and between vibrational and rotational modes. In addition, the coupling between the translational motion and rotations is neglected. In HES, terms describing the coupling of the solid C60 crystal field to the electric dipole and quadrupole moments of H2O are included.

The fitting of IR absorption spectra (Sec. IV B) is done with the Hamiltonian where the translational part is excluded,

(5)

We employ three coordinate frames. The space-fixed coordinate frame is denoted A. M = {x, y, z} is the molecule-fixed coordinate frame (Fig. 1). The Euler angles ΩAM39,40 transform A to M. The crystal coordinate frame is C = {x′, y′, z′} with the z′ axis along the threefold symmetry axis of the S6 point group, the symmetry group of the C60 site in solid C60. The Euler angles ΩAC transform A to C and ΩCM transform C to M. The coordinate systems A and C are used because the radiation interacting with the molecule is defined in the space-fixed coordinate frame A, while the local electrostatic fields are defined by the crystal coordinate frame C, which in a powder sample has a uniform distribution of orientations relative to the space-fixed frame A.

FIG. 1.

(a) Molecule-fixed coordinate frame M = {x, y, z} and axes of principal moments of inertia, {a, b, c}. (b) Vibrations: v1, symmetric stretch; v2, symmetric bend; and v3, asymmetric stretch. (c) para-water and ortho-water rotational energy levels in the ground and excited vibrational states (see Sec. III A 5) and rovibrational IR transitions (arrows) between the levels. The rotational states are labeled by JKaKc and p (para) and o (ortho). The IR transitions are between the para states or between the ortho states. 1, 3, and 5 are ortho-H2O transitions, and 2 and 4 are para-H2O transitions. Transitions 1 and 2 are forbidden for a free water molecule. The transitions where the one-quantum excitation of the asymmetric stretch vibration v3 is involved are numbered 6 (para-H2O), and 7 and 8 (ortho-H2O transitions). The rotational and translational far-IR transitions in the ground vibrational states are shown in the inset of Fig. 2.

FIG. 1.

(a) Molecule-fixed coordinate frame M = {x, y, z} and axes of principal moments of inertia, {a, b, c}. (b) Vibrations: v1, symmetric stretch; v2, symmetric bend; and v3, asymmetric stretch. (c) para-water and ortho-water rotational energy levels in the ground and excited vibrational states (see Sec. III A 5) and rovibrational IR transitions (arrows) between the levels. The rotational states are labeled by JKaKc and p (para) and o (ortho). The IR transitions are between the para states or between the ortho states. 1, 3, and 5 are ortho-H2O transitions, and 2 and 4 are para-H2O transitions. Transitions 1 and 2 are forbidden for a free water molecule. The transitions where the one-quantum excitation of the asymmetric stretch vibration v3 is involved are numbered 6 (para-H2O), and 7 and 8 (ortho-H2O transitions). The rotational and translational far-IR transitions in the ground vibrational states are shown in the inset of Fig. 2.

Close modal

1. Vibrations

H2O has three normal vibrations: the symmetric stretch of O–H bonds, denoted by the quantum number v1, the bending motion of the H–O–H bond angle, v2, and the asymmetric stretch of O–H bonds, v3, 40 as sketched in Fig. 1(b). The vibrational state is denoted v, where the symbol v denotes the three vibrational quantum numbers, vv1v2v3, each of which takes values vi ∈ {0, 1, 2, …}. The vibrational energy for a harmonic vibrational potential is

(6)

where ωi is the vibrational frequency of the ith vibration mode, i ∈ {1, 2, 3} and [ωi] = cm−1. Throughout this paper, the reported ωi is the distance between the vibrational energy levels and is assumed to include anharmonic corrections.

2. Rotations

H2O has the rotational properties of an asymmetric top with principal moments of inertia Ia < Ib < Ic.40 The rotational states are indexed by three quantum numbers JKaKc, where J = 0, 1, 2, …, is the rotational angular momentum quantum number. Ka and Kc are the absolute values of the projection of J onto the a and c axes, in the limits of a prolate (Ia = Ib) and an oblate (Ib = Ic) top, respectively, Ka, KcJ.40 Each rotational state JKaKc,m is (2J + 1)-fold degenerate, where m ∈ {−J, −J + 1, …, J} is the projection of J onto the z′ axis of the crystal coordinate frame C.

The energies and the wavefunctions, JKaKc,m, of the free rotor Hamiltonian Hrot depend on the rotational constants of an asymmetric top, Av > Bv > Cv,40 

(7)

where Ĵi are the components of the angular momentum operator J^2 along the principal directions a, b, and c. The index v labels the rotational constants in the vibrational state v.

A molecule-fixed coordinate system M (axes x, y, and z, Fig. 1) with its origin at the nuclear center of mass is defined with the following orientations relative to the principal axes of the inertial tensor: xb, yc, and za, where y is perpendicular to the H–O–H plane and x points toward the oxygen atom. The rotational wavefunctions of an asymmetric top in that basis are39,40

(8)

where k is the projection of J on the a axis, axis z of M; for k > 0, k ∈ {1, 2, …, J} and each state is doubly degenerate. Here, J,k,m denotes a normalized rotational function,

(9)

where the Euler angles ΩCM transform the crystal-fixed coordinate frame C into the molecule-fixed coordinate frame M and DmkJ(Ω) is the Wigner rotation matrix element or the Wigner D-function.39,41

The correspondence between the asymmetric top wavefunctions J,k,m,± [Eq. (8)] and the asymmetric top wavefunctions JKaKc,m is given in Ref. 40. The latter notation of wavefunctions is useful for the symmetry analysis and can be used to relate the wavefunction to para and ortho states of the water molecule (see Sec. III A 5) and to define the electric dipole transition selection rules [Eqs. (26) and (27)].

3. Electrostatic interactions

We assume two contributions to the electrostatic interaction,

(10)

denoting the coupling of the quadrupole and dipole moments of H2O to the corresponding multipole fields created by the surrounding charges.

a. Quadrupolar interaction.

It was shown by Felker et al.32 that C60 molecules in neighboring lattice sites generate an electric field gradient at the center of a given C60 molecule. In H2O@C60, the electric field gradient couples to the electric quadrupole moment of the water molecule, lifting the threefold degeneracy of the J = 1 ortho-H2O rotational ground state.32 

The quadrupolar Hamiltonian may be expanded in rank-2 spherical tensors as follows: 39,42

(11)

where {Vm(2)} are the spherical components of the electric field gradient tensor and {QmC} are the quadrupole moments of the water molecule, both expressed in the crystal-fixed coordinate frame C.

The experimental value of the H2O quadrupole moment in the molecule-fixed coordinate frame M is given by {Qxx, Qyy, Qzz} = {−0.13, −2.50, 2.63} esu × cm2.43 Since |Qxx| ≪ |Qyy|, |Qzz|, it holds that Qzz ≈ −Qyy, and we may approximate the water quadrupole moment in spherical coordinates as follows:

(12)

The site symmetry of the C60 molecule in solid C60 is S6, with the threefold symmetry axis along the cubic [111] axis, which is chosen here to be the z′ axis of the crystal coordinate frame C. The spherical tensor component in frame C is

(13)

where m ∈ {−2, −1, 0, 1, 2} transforms like the fully symmetric Ag irreducible representation of the point group S6.44 After the transformation of the quadrupole moment (12) from the H2O-fixed molecular frame M to the crystal frame C [see Eq. (A5)], the quadrupolar Hamiltonian (11) is given by

(14)
b. Dipolar interaction.

Dielectric measurements of solid C60 have provided evidence for the existence of electric dipoles in solid C60.35 We propose that these electric dipoles are the source of finite electric fields at the C60 cage centers.

Consider a crystal electric field E with spherical coordinates {E,ϕE,θE} in the crystal-fixed frame C; see  Appendix A. For simplicity, we assume a homogeneous crystal field with uniform orientation in the crystal-fixed frame. The interaction of the electric dipole moment with the electric field is given by

(15)

where the dipole moment in the molecule-fixed frame M is given by

(16)

where μx is the permanent dipole moment of water in the Cartesian coordinates of frame M [Fig. 1(a)]. Since there are no other anisotropies than the axially symmetric electric field gradient tensor, the angle ϕE is arbitrary and we choose ϕE = 0.

The Hamiltonian (5) is diagonalized using the basis (8) up to J ≤ 4 for the ground vibrational state 000 and for the three excited vibrational states 100, 010, and 001. The ground state and the three excited vibrational states are assumed to have independent rotational constants Av, Bv, and Cv, where v = 000, 100, 010, or 001.

After separation of coordinates (see Appendix B 2), the quadrupole and dipole moments in Eqs. (14) and (15) are replaced by their expectation values, vQ0Mv, vQ±2Mv, and vμ±1Mv, in the ground and in the three excited vibrational states. We assume for simplicity that the dipole and quadrupole moments of H2O are independent of the vibrational state v.

4. Confined water translations: Spherical oscillator

The translational motion is the center-of-mass motion and is quantized for a confined molecule. The high icosahedral symmetry of the C60 cavity is close to spherical symmetry, and therefore, the translational motion of the trapped molecule can be described by the three-dimensional isotropic spherical oscillator model.45 For simplicity, we write the potential in the harmonic approximation as42 

(17)

where R is the displacement of the H2O center of mass from the C60 cage center. The C60 cage is assumed rigid, and its center of mass is fixed.

The frequency of the spherical harmonic oscillator is

(18)

where m is the mass of a molecule moving in the potential and [ωt0]=rads1. The energy of the spherical harmonic oscillator is quantized,

(19)

where N is the translational quantum number, N ∈ {0, 1, 2, …}. The orbital quantum number L takes values L = N, N − 2, …, 1(0) for odd (even) N. The energy of the harmonic spherical oscillator does not depend on L, and in the isotropic approximation, there is an additional degeneracy of each EN level in quantum number ML, taking 2L + 1 values, ML ∈ {−L, −L + 1, …, L}.

5. Nuclear spin isomers: para-water and ortho-water

The Pauli principle requires that the total quantum state is antisymmetric with respect to exchange of the two protons in water, which as spin-1/2 particles are fermions. This constraint leads to the existence of two nuclear spin isomers, with total nuclear spins I = 0 (para-H2O) and I = 1 (ortho-H2O), and different sets of rovibrational states. The antisymmetric nature of the quantum state has consequences on the IR spectra: only para to para and ortho to ortho transitions are allowed.

The ortho-H2O states have odd values of Ka + Kc, while the para-H2O states have even values of Ka + Kc in the ground vibrational state 000; see Fig. 1(c). The allowed rotational transitions are depicted in the inset of Fig. 2(a). The same rule applies to the excited vibrational states 100 and 010 [Fig. 1(c), upper left part]. However, the rules are inverted for the states 001, 101, and 011, which involve one-quantum excitation of the asymmetric stretch mode v3. In these cases, 40para-H2O has odd values of Ka + Kc, while ortho-H2O has even values of Ka + Kc [Fig. 1(c), upper right part].

FIG. 2.

Far-IR absorption spectra of H2O@C60 at 5 K. (a) Spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 44 h later (blue). The sample filling factor f = 0.05. Water rotational transitions corresponding to the absorption lines numbered 3 and 5 (ortho-water) and 4 (para-water) are shown in the inset. (b) Spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 5 h later (blue). The sample filling factor f = 0.18. The translational transitions N = 0 → N = 1 for para (Tp) and ortho (To1, To2) H2O@C60 are shown in the inset of panel (a). N is the quantum number of the spherical harmonic oscillator [Eq. (19)].

FIG. 2.

Far-IR absorption spectra of H2O@C60 at 5 K. (a) Spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 44 h later (blue). The sample filling factor f = 0.05. Water rotational transitions corresponding to the absorption lines numbered 3 and 5 (ortho-water) and 4 (para-water) are shown in the inset. (b) Spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 5 h later (blue). The sample filling factor f = 0.18. The translational transitions N = 0 → N = 1 for para (Tp) and ortho (To1, To2) H2O@C60 are shown in the inset of panel (a). N is the quantum number of the spherical harmonic oscillator [Eq. (19)].

Close modal

The energy difference between the lowest para rotational state 000 and the lowest ortho rotational state 101 is 2.6 meV (28 K).11 Above 30 K, the ratio of ortho and para molecules is no/np ≈ 3. Hence, if the sample is cooled rapidly to 4 K, the number of para molecules slowly grows in the subsequent time interval, while the number of ortho molecules slowly decreases to the thermal equilibrium value no ≈ 0. The full conversion takes several hours.11,17,30,37

The strengths of the transitions between the rotational states of a polar molecule are determined by the permanent electric dipole moment of the molecule and by the electric field of the infrared radiation, corrected by the polarizability of the medium. In principle, the polarizability χ of the solid depends on the fraction f of C60 cages that contain a water molecule, χ(f)=χC60+fχH2O@C60. However, we found that, within the studied range of filling factors, f = 0.1–0.8, the absorption cross section of H2O@C60 was independent of f. Hence, only the polarizability of solid C60 is relevant, χχC60, and the problem is similar to the optical absorption of an isolated impurity atom in a crystal.46 

Following Ref. 46, the electric field in the molecule embedded into the medium with an index of refraction η is Eeff=E(η2+2)/3, where E is the electric field of radiation in the vacuum. The refractive index of solid C60 is η = 2, 47 and hence, Eeff=2E.

In the following discussion, we use the index k ∈ {o, p} to indicate the ortho or para nuclear spin isomers. The absorption cross section40 for a given nuclear spin isomer k, including the effective field correction, is given by

(20)

where c0 is the speed of light in the vacuum and ε0 is the permittivity of vacuum; SI units are used, and the frequency ω is in number of waves per meter, [ω] = m−1. The integral in (20) is the area of the absorption line of the transition from the state i to j.

The square of the electric dipole matrix element is given by

(21)

where i and j are the eigenstates with corresponding energies Ei(k) and Ej(k). The symbol μσC denotes the dipole moment components of a water molecule in the crystal-fixed frame C. This form of Sji(k) is valid for random orientation of crystals in the powder sample and does not depend on the polarization of light; see  Appendix B.

The absorption cross section σji(k) was evaluated separately for para-water and ortho-water. The concentration of molecules is Nk=fNC60nk, where nk is the ortho (or para) fraction and np + no = 1. The filling factor is denoted by f, and the number density of C60 molecules in solid C60 is given by NC60=1.419×1021cm3.48 

pi(k) and pj(k) are the probabilities that the initial and final states are thermally populated,

(22)

where the statistical sum is

(23)

and E0(k) is the ground state energy of para (ortho) molecules. At the low temperatures considered in this work, only the vibrational ground state is significantly populated. Thus, the significantly thermally populated states are rotational states in the ground vibrational states, which can be written as linear combinations of basis states in Eq. (8). Since we expect that the water molecule is not in a spherically symmetric environment in H2O@C60, the degeneracy in quantum number m is lifted, in general. Therefore, pn(k) is the thermal population of a non-degenerate rotational state, and the eigenstates i and j in Eq. (21) include all possible m values for a given J.

When thermal equilibrium is reached between para-water and ortho-water, the fraction of the nuclear spin isomer k is

(24)

For the spin isomer k, the nuclear spin degeneracy g(k) = 2I + 1, with I = 0 for para and I = 1 for ortho.

The absorption line areas are calculated from Eq. (20), where the matrix elements in Eq. (21) are between the eigenstates Φrotv of the Hamiltonian given by Eq. (5). After separation of coordinates (see Appendix B 2), the matrix elements in the crystal-fixed coordinate frame C are

(25)

where the initial state is i=000Φrot000 and Φrotv are the linear combinations of states (8). The dipole moments vμσM000 are given by Eqs. (B13) and (B14) in Cartesian coordinates.

The following selection rules hold for the electric dipole moment along the z axis (the transition dipole moment of asymmetric stretch v3) and along the x axis (the permanent dipole moment and the transition dipole moments of v1 and v2):40 

(26)
(27)

where the convention of molecular axes is as shown in Fig. 1(a) and the definition of dipole moments is given by Eqs. (B13) and (B14). Corresponding transitions are sketched in Figs. 1(c) and 2.

The water IR absorption lines were identified unambiguously by taking advantage of the slow orthopara conversion at low temperature. After rapid cooling from 30 K to 5 K, the slow orthopara conversion causes the intensity of the para lines to slowly increase, while the intensity of the ortho lines decreases. The water absorption lines are readily identified and assigned to one of the two spin isomers, by taking the difference between spectra acquired shortly after cooling and spectra acquired after an equilibration time at the lower temperature.

A group of H2O lines, numbered 3, 4, and 5, is seen below 60 cm−1 [Fig. 2(a)]. These far-IR absorption lines have been reported earlier and correspond to the rotational transitions of H2O in the C60 cage.11 The rotational energy levels involved are shown in the inset of Fig. 2(a). Lines 3 and 5 are ortho-water rotational transitions starting from the ortho-water ground state 101. Line 4 is the para-water transition from the ground rotational state 000. No other rotational transitions were observed at 5 K, which is consistent with the selection rules for the electric dipole allowed rotational transitions from states 000 and 101.40 

Further H2O lines are observed around 110 cm−1 [Fig. 2(b)] and in six spectral regions above 600 cm−1, as shown in Figs. 36. Below, we address each wavenumber range separately and assign the spectral lines to the transitions shown in the energy schemes of Figs. 1(c) and 2(a). The absorption lines associated with transitions of free water, labeled 1–8, are listed in Table I. Line assignments are supported by the results of spectral fitting using the model of a vibrating rotor in a crystal field.

FIG. 3.

Absorption spectra of H2O@C60 vibrational, rovibrational, and vibration–translational transitions of the bending vibration v2 at 5 K. The spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 3 h later (blue). The sample filling factor f = 0.1. Lines numbered 1 and 2 are pure vibrational transitions and 3–5 are rovibrational transitions [Fig. 1(c)]. The left inset shows the para, line 1, and ortho, line 2, components of the pure vibrational transition difference spectrum with time delay 1 h of the f = 0.8 sample. The differential spectrum of v2 vibration–translational transitions Tp, To1, and To2, at 110 cm−1 higher frequency from 1 to 2, is shown in the right inset.

FIG. 3.

Absorption spectra of H2O@C60 vibrational, rovibrational, and vibration–translational transitions of the bending vibration v2 at 5 K. The spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 3 h later (blue). The sample filling factor f = 0.1. Lines numbered 1 and 2 are pure vibrational transitions and 3–5 are rovibrational transitions [Fig. 1(c)]. The left inset shows the para, line 1, and ortho, line 2, components of the pure vibrational transition difference spectrum with time delay 1 h of the f = 0.8 sample. The differential spectrum of v2 vibration–translational transitions Tp, To1, and To2, at 110 cm−1 higher frequency from 1 to 2, is shown in the right inset.

Close modal
FIG. 4.

Absorption spectra of H2O@C60 vibrational, rovibrational, and vibration–translational transitions of symmetric stretching, v1, and anti-symmetric stretching, v3, vibrations at 5 K. The spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 3 h later (blue dashed line multiplied by 4 and blue line). The sample filling factor f = 0.1. Lines numbered 1 and 2 are pure vibrational transitions and 3–5 are rovibrational transitions of mode v1, and lines 6–8 are rovibrational transitions of v3 [Fig. 1(c)]. (*) marks the v3 rovibrational transition at 3654 cm−1 from the thermally excited rotational state 000110 to the rovibrational state 001111. The inset shows the para, line 1, and ortho, line 2, components of the pure vibrational transition of mode v1 of the f = 0.8 sample as the difference spectrum with time delay 1 h.

FIG. 4.

Absorption spectra of H2O@C60 vibrational, rovibrational, and vibration–translational transitions of symmetric stretching, v1, and anti-symmetric stretching, v3, vibrations at 5 K. The spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 3 h later (blue dashed line multiplied by 4 and blue line). The sample filling factor f = 0.1. Lines numbered 1 and 2 are pure vibrational transitions and 3–5 are rovibrational transitions of mode v1, and lines 6–8 are rovibrational transitions of v3 [Fig. 1(c)]. (*) marks the v3 rovibrational transition at 3654 cm−1 from the thermally excited rotational state 000110 to the rovibrational state 001111. The inset shows the para, line 1, and ortho, line 2, components of the pure vibrational transition of mode v1 of the f = 0.8 sample as the difference spectrum with time delay 1 h.

Close modal
FIG. 5.

Absorption spectra of H2O@C60 at 10 K. (a) Overtone, 2ω2, and (b) combination, 2ω2 + ω3, rovibrational transitions. The spectrum α(0) was measured after the temperature jump from 45 to 10 K (black), and the difference Δα = α(0) − αt) was measured Δt = 2.75 h later (blue). Numbers 3–8 label the transitions shown in Fig. 1(c). The sample filling factor f = 0.8.

FIG. 5.

Absorption spectra of H2O@C60 at 10 K. (a) Overtone, 2ω2, and (b) combination, 2ω2 + ω3, rovibrational transitions. The spectrum α(0) was measured after the temperature jump from 45 to 10 K (black), and the difference Δα = α(0) − αt) was measured Δt = 2.75 h later (blue). Numbers 3–8 label the transitions shown in Fig. 1(c). The sample filling factor f = 0.8.

Close modal
FIG. 6.

Absorption spectra of H2O@C60 at 5 K. Rovibrational combination (a) ω2 + ω3 and (b) ω1 + ω3 transitions. The spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 0.83 h later (blue). Numbers 6–8 label the transitions shown in Fig. 1(c). (*) in (a) at 5202 cm−1 marks the ω2 + ω3 rovibrational transition from the thermally excited rovibrational state 000110 to 011111. The 2ω1 rovibrational transition at 7059 cm−1, marked (+) in (b), is from 000000 to 200111. The sample filling factor f = 0.8.

FIG. 6.

Absorption spectra of H2O@C60 at 5 K. Rovibrational combination (a) ω2 + ω3 and (b) ω1 + ω3 transitions. The spectrum α(0) was measured after the temperature jump from 30 to 5 K (black), and the difference Δα = α(0) − αt) was measured Δt = 0.83 h later (blue). Numbers 6–8 label the transitions shown in Fig. 1(c). (*) in (a) at 5202 cm−1 marks the ω2 + ω3 rovibrational transition from the thermally excited rovibrational state 000110 to 011111. The 2ω1 rovibrational transition at 7059 cm−1, marked (+) in (b), is from 000000 to 200111. The sample filling factor f = 0.8.

Close modal
TABLE I.

Rotational and rovibrational transition frequencies ωji (unit: cm−1) and normalized absorption cross sections ⟨σji⟩ (unit: cm per molecule) [Eq. (3)] from the ground para, 000, and ground ortho, 101, rotational states of H2O@C60 and of free H2O. The initial vibrational state is 000 for all transitions. The spectral lines are labeled by Nos. 1–8, each number associated with one pair of initial and final rotational states. The frequency ωjiC60 of lines 3–8 is the intensity-weighted average of line sub-component frequencies. ωjigas and σjigas are from Ref. 49. The paraortho index k in σji(k) has been dropped to simplify the notation.

v1v2v3jNo.JKaKciJKaKcjωjigasωjiC60σjigasσjiC60σjigas/σjiC60
000 101 110 18.6 16.8 (3.97 ± 0.08)10−17 (3.24 ± 0.26)10−18 12.1 
 000 111 37.1 33.6 (5.30 ± 0.11)10−17 (3.02 ± 0.19)10−18 17.3 
 101 212 55.7 51.1 (1.191 ± 0.024)10−16 (7.5 ± 0.4)10−18 16.0 
100 000 000 3657.1 3573.2  (3.4 ± 0.6)10−20  
 101 101      
 101 110 3674.7 3589.1 (5.03 ± 0.05)10−19 (2.89 ± 0.03)10−19 1.74 
 000 111 3693.3 3606.5 (3.01 ± 0.03)10−19 (1.93 ± 0.14)10−19 1.56 
 101 212 3711.1 3623.5 (3.97 ± 0.04)10−19 (1.57 ± 0.12)10−19 2.52 
010 000 000 1594.8 1569.3  (2.3 ± 0.6)10−19  
 101 101      
 101 110 1616.7 1588.5 (1.66 ± 0.03)10−17 (1.03 ± 0.11)10−18 16.1 
 000 111 1635.0 1605.3 (1.132 ± 0.023)10−17 (6.79 ± 0.23)10−19 16.7 
 101 212 1653.3 1623.4 (1.73 ± 0.04)10−17 (1.26 ± 0.09)10−18 13.7 
001 000 101 3779.5 3682.1 (8.08 ± 0.09)10−18 (1.73 ± 0.21)10−18 4.66 
 101 000 3732.1 3637.4 (8.64 ± 0.09)10−18 (1.31 ± 0.12)10−18 6.57 
 101 202 3801.4 3703.7 (1.526 ± 0.015)10−17 (2.18 ± 0.10)10−18 6.99 
011 000 101 5354.9 5228.0 (8.75 ± 0.09)10−19 (1.15 ± 0.06)10−19 7.63 
 101 000 5307.5 5183.6 (8.97 ± 0.09)10−19 (8.22 ± 0.04)10−20 10.9 
 101 202 5376.9 5249.2 (1.703 ± 0.017)10−18 (1.34 ± 0.06)10−19 12.7 
101 000 101 7273.0 7088.4 (5.5 ± 0.5)10−19 (1.261 ± 0.021)10−20 43.4 
 101 000 7226.0 7043.5 (5.7 ± 0.6)10−19 (1.044 ± 0.012)10−20 54.6 
 101 202 7294.1 7109.3 (1.07 ± 0.11)10−18 (1.85 ± 0.05)10−20 57.7 
020 000 111 3196.1 3142.0 (7.72 ± 0.07)10−20 (5.5 ± 0.4)10−21 14.1 
v1v2v3jNo.JKaKciJKaKcjωjigasωjiC60σjigasσjiC60σjigas/σjiC60
000 101 110 18.6 16.8 (3.97 ± 0.08)10−17 (3.24 ± 0.26)10−18 12.1 
 000 111 37.1 33.6 (5.30 ± 0.11)10−17 (3.02 ± 0.19)10−18 17.3 
 101 212 55.7 51.1 (1.191 ± 0.024)10−16 (7.5 ± 0.4)10−18 16.0 
100 000 000 3657.1 3573.2  (3.4 ± 0.6)10−20  
 101 101      
 101 110 3674.7 3589.1 (5.03 ± 0.05)10−19 (2.89 ± 0.03)10−19 1.74 
 000 111 3693.3 3606.5 (3.01 ± 0.03)10−19 (1.93 ± 0.14)10−19 1.56 
 101 212 3711.1 3623.5 (3.97 ± 0.04)10−19 (1.57 ± 0.12)10−19 2.52 
010 000 000 1594.8 1569.3  (2.3 ± 0.6)10−19  
 101 101      
 101 110 1616.7 1588.5 (1.66 ± 0.03)10−17 (1.03 ± 0.11)10−18 16.1 
 000 111 1635.0 1605.3 (1.132 ± 0.023)10−17 (6.79 ± 0.23)10−19 16.7 
 101 212 1653.3 1623.4 (1.73 ± 0.04)10−17 (1.26 ± 0.09)10−18 13.7 
001 000 101 3779.5 3682.1 (8.08 ± 0.09)10−18 (1.73 ± 0.21)10−18 4.66 
 101 000 3732.1 3637.4 (8.64 ± 0.09)10−18 (1.31 ± 0.12)10−18 6.57 
 101 202 3801.4 3703.7 (1.526 ± 0.015)10−17 (2.18 ± 0.10)10−18 6.99 
011 000 101 5354.9 5228.0 (8.75 ± 0.09)10−19 (1.15 ± 0.06)10−19 7.63 
 101 000 5307.5 5183.6 (8.97 ± 0.09)10−19 (8.22 ± 0.04)10−20 10.9 
 101 202 5376.9 5249.2 (1.703 ± 0.017)10−18 (1.34 ± 0.06)10−19 12.7 
101 000 101 7273.0 7088.4 (5.5 ± 0.5)10−19 (1.261 ± 0.021)10−20 43.4 
 101 000 7226.0 7043.5 (5.7 ± 0.6)10−19 (1.044 ± 0.012)10−20 54.6 
 101 202 7294.1 7109.3 (1.07 ± 0.11)10−18 (1.85 ± 0.05)10−20 57.7 
020 000 111 3196.1 3142.0 (7.72 ± 0.07)10−20 (5.5 ± 0.4)10−21 14.1 

1. Translational transitions

A group of absorption lines around 110 cm−1 is shown in Fig. 2(b). These lines do not correspond to any known water rotational transitions. We assign these peaks to the translational transitions (N = 0 → N = 1) of para-H2O and ortho-H2O, corresponding to the quantized center-of-mass vibrational motions of the water molecules in the encapsulating C60 cages. Here, N denotes the quantum number of a spherical harmonic oscillator.42 

The assignment of these peaks to water center-of-mass translational oscillations is supported by the presence of lines at 1680 cm−1, visible in the difference spectrum shown in the right-hand side inset of Fig. 3. These lines are 110 cm−1 higher than the vibrational transitions 1 and 2 of the v2 mode and correspond to the simultaneous excitation of the v2 vibration and the translational modes. A similar combination has been observed in H2@C60 where a group of lines between 4240 and 4270 cm−1 is the translational sideband to the H2 stretching vibration.10 

The translational side peak of the v1 vibrational mode is expected at about 3683 cm−1. However, this frequency coincides with a strong rovibrational absorption line 6 of the v3 mode (Fig. 4), which probably obscures the 3683 cm−1 translational side peak of the v1 vibration.

The translational ortho transitions display a splitting of 2.9 cm−1 in the ground vibrational state; see Fig. 2(b), and 2.7 cm−1 in the excited vibrational state 010; see Fig. 4. These splittings may be attributed to the coupling between the water translation and rotation, associated with the interaction of the non-spherical rotating water molecule with the interior of the C60 cage. The spectral structure of this type has been analyzed in detail for the case of H2@C60.8,10,12 The simplified theoretical model used here does not include translation–rotation coupling and cannot explain these splittings. A theoretical analysis of the translational peaks will be given in a later paper.

2. Vibrational and rovibrational transitions

The vibrational and rovibrational transitions are shown in Fig. 3. The three major features that distinguish the spectrum of H2O@C60 from the spectrum of free water are as follows.

a. Pure vibrational transitions.

Absorptions corresponding to pure vibrational transitions, i.e., without simultaneous rotational excitation, are present around ω1 = 3573 cm−1 and ω2 = 1570 cm−1. Both features are split into two components, labeled 1 and 2, identified from the difference spectra (Figs. 3 and 4) as para (1) and ortho (2) transitions.

Transition 1 is a transition from the ground vibrational state to the excited vibrational state without a change in the rotational state 000. Transition 2 is a vibrational excitation without a change in the rotational state 101. The corresponding transitions are forbidden because they violate the selection rule ΔKa = odd and ΔKc = odd [Eq. (27)]. As will be discussed in Sec. V E, the pure vibrational transitions 1 and 2 can be activated by the crystal electric field of solid H2O@C60.

The orthopara splitting, i.e., the separation of lines 1 and 2, is 0.5 cm−1 for the v1 vibrational mode and 1.8 cm−1 for the v2 vibrational mode.

b. Spectral splittings.

The rovibrational transitions 3–8 are split into two or more components (see Figs. 3 and 4). These splittings, as in the case of the rotational transitions, are absent for a water molecule in the gas phase. Moreover, transitions 3–5 have the same splitting pattern as the rotational transitions in the ground vibrational state, also labeled 3–5; see Fig. 2. Transitions 6 and 7 are special since they are between the 000 and 101 rotational states and, thus, reflect directly the splitting of the triply degenerate 101 state either in the ground vibrational state, transition 7, or in the excited vibrational state, transition 6 (Fig. 4).

We assign a weak ortho line at 3654 cm−1, marked by * in Fig. 4, to the v3 rovibrational transition from the thermally excited rovibrational state 000110 to 001111. This assignment is further confirmed by calculating the transition frequency39 with the parameters from Table III: ω3+E111E110=ω3+A001+C001(A000+B000)=3655cm1.

c. Red shifts.

The frequencies of vibrations are red-shifted relative to free H2O. The stretching mode frequencies are red-shifted by about 2.4%, while the bending mode frequencies are red-shifted by about 1.6%; see Table II.

TABLE II.

Frequencies of symmetric stretching (ω1), symmetric bending (ω2), and asymmetric stretching (ω3) modes and of their combinations 2ω2, ω2 + ω3, 2ω2 + ω3, 2ω1, and ω1 + ω3 measured from the ground vibrational state 000 for H2O@C60 (this work) and for free H2O.50ortho and para components of pure vibrational transitions v1 and v2 are indicated by superscripts o and p. The frequencies of transitions involving v3 in H2O@C60 are the average of frequencies of transitions 6 and 7; see Figs. 4, 5(b), and 6. The overtone 2ω2 is estimated from the frequency of line 4 [Fig. 5(a)], 2ω2+E111=3141cm1, where E111=A+C is the energy of the rotational state 111,39 in which A = A010 and C = C010 from Table III. The 2ω1 overtone frequency is estimated from the frequency of the para line 2ω1+E111=7059cm1 [Fig. 6(b)], where E111=A100+C100. The shift Δω=ωC60ωgas.

v1v2v3H2O@C60 cm−1Free H2O cm−1Δω cm−1Δω/ωgas
010 1569.2o 1594.8 −25.6 −0.016 
 1571.0p    
100 3573.2o 3657.1 −83.9 −0.023 
 3573.7p    
001 3659.6 3755.3 −95.7 −0.025 
020 3105.5 3151.6 −46.1 −0.015 
011 5205.2 5331.3 −126.1 −0.024 
021 6716.6 6871.5 −154.9 −0.023 
200 7027.0 7201.5 −173.5 −0.024 
101 7065.8 7249.8 −184.0 −0.025 
v1v2v3H2O@C60 cm−1Free H2O cm−1Δω cm−1Δω/ωgas
010 1569.2o 1594.8 −25.6 −0.016 
 1571.0p    
100 3573.2o 3657.1 −83.9 −0.023 
 3573.7p    
001 3659.6 3755.3 −95.7 −0.025 
020 3105.5 3151.6 −46.1 −0.015 
011 5205.2 5331.3 −126.1 −0.024 
021 6716.6 6871.5 −154.9 −0.023 
200 7027.0 7201.5 −173.5 −0.024 
101 7065.8 7249.8 −184.0 −0.025 

3. Overtone and combination rovibrational transitions

Overtone and combination vibrational transitions where two vibrational quanta are excited are presented for the 2ω2 transition in Fig. 5(a) and for the ω2 + ω3 and ω1 + ω3 transitions in Figs. 6(a) and 6(b). A three-quantum transition, 2ω2 + ω3, is shown in Fig. 5(b). Rotational levels involved are sketched in Fig. 1(c). Again, the splitting pattern of each higher order rovibrational transition is similar to the splitting of the rotational transition with Δvi = 0 (Fig. 2) and rovibrational transitions with Δvi = +1 (Figs. 3 and 4).

We assign a line marked by “+” at 7059 cm−1 to 2ω1 plus para-H2O rotational transition, 000 → 111 [Fig. 6(b)]. Another two rotational side peaks of 2ω1 are ortho transitions 3 and 5 expected at 2ω1+E110E101=7042cm1 and at 2ω1+E212E101=7076cm1, where E110E101=AC and E212E101=A+3C,39 with the approximation A = A200A100, C = C200C100. The numerical values of A100 and C100 are taken from Table III. The first line overlaps with line 7, rovibrational transition of ω1 + ω3. The second line is not observed, but this could be due to the low intensity of ortho line 5 relative to para line 4; see, for example, Fig. 5(a).

TABLE III.

Parameters obtained from the quantum mechanical model fit of IR spectra of H2O@C60 at T = 5 K (Fig. 7). The vibration frequencies ωi, rotational constants Av, Bv, Cv, and quadrupolar energy VQQzz are in units of cm−1; electric field E in 106 V m−1; θE in radians; and dipole moment μx(B13) and transition dipole moments μ0ix,z(B14) in D. It is assumed that μx and Qzz do not depend on the vibrational state. The parameters with zero error were not fitted.

ParameterValueError
no 0.7 
f 
ω1 3574.1 0.3 
ω2 1569.2 0.3 
ω3 3659.9 0.9 
A000 24.15 0.17 
B000 15.3 0.8 
C000 8.48 0.07 
A100 23.1 0.3 
B100 14.3 0.8 
C100 8.50 0.09 
A010 26.7 0.3 
B010 14.6 0.9 
C010 8.81 0.07 
A001 26 
B001 15 
C001 8.2 1.8 
μx 0.474 0.008 
μ01x 1.031 × 10−2 0.021 × 10−2 
μ02x 3.40 × 10−2 0.05 × 10−2 
μ03z 2.83 × 10−2 0.07 × 10−2 
E 110 
θE 1.4 1.0 
VQQzz −5.0 0.5 
ParameterValueError
no 0.7 
f 
ω1 3574.1 0.3 
ω2 1569.2 0.3 
ω3 3659.9 0.9 
A000 24.15 0.17 
B000 15.3 0.8 
C000 8.48 0.07 
A100 23.1 0.3 
B100 14.3 0.8 
C100 8.50 0.09 
A010 26.7 0.3 
B010 14.6 0.9 
C010 8.81 0.07 
A001 26 
B001 15 
C001 8.2 1.8 
μx 0.474 0.008 
μ01x 1.031 × 10−2 0.021 × 10−2 
μ02x 3.40 × 10−2 0.05 × 10−2 
μ03z 2.83 × 10−2 0.07 × 10−2 
E 110 
θE 1.4 1.0 
VQQzz −5.0 0.5 

All two-quantum and three-quantum vibrational transitions are red-shifted approximately by 2.4% except 2ω2 where the red shift is 1.5%; see Table II.

A synthetic spectrum consisting of Gaussian lines with full width at half maximum 1.5 cm−1 was calculated from the experimental normalized line areas Aji(k) [Eq. (2)] using f = 1, no = 0.7, np = 0.3, and T = 5 K. The parameters of the model Hamiltonian (from Sec. III A) and the transition dipole moments (from Sec. III B) were determined with a non-linear least squares method by minimizing the difference in synthetic and modeled spectra squared,  Appendix C. The reported parameter error is the average of errors calculated with +δaν and −δaν in Eq. (C9), where the parameter variation of the νth parameter at its best value aminν is |δaν|=0.005  aminν. The fit was applied to the rotational transitions in the ground vibrational state 000 and to the rovibrational transitions from the ground state to the vibrational states 010, 100, and 001. In total, 45 absorption lines were fitted.

The synthetic experimental spectra and the best fit spectra are shown in Fig. 7, with the best fit parameters given in Table III. The result of the fit overlaps well with the synthetic spectrum, except for transition 5 as seen in the first three panels of Fig. 7. While, for other transitions, one or two Gaussian components were sufficient, the experimental transition lineshape required three components to get a reliable fit of its line area. In addition, transitions 6 and 7 were represented by two components in the synthetic spectrum, although four peaks are seen in the experimental spectrum (Fig. 4). The additional structure of experimental peaks may originate from the merohedral disorder as discussed in Sec. V E.

FIG. 7.

Synthetic experimental spectra (black solid line) and spectra calculated with the best fit parameters from Table III (blue dashed line) with the filling factor f = 1 and ortho fraction no = 0.7 and at a temperature of 5 K. (a) Rotational transitions in the ground vibrational state. Vibrational (1 and 2) and rovibrational (3–5) transitions of (b) v2 and (c) v1. (d) Rovibrational transitions (6–8) of v3.

FIG. 7.

Synthetic experimental spectra (black solid line) and spectra calculated with the best fit parameters from Table III (blue dashed line) with the filling factor f = 1 and ortho fraction no = 0.7 and at a temperature of 5 K. (a) Rotational transitions in the ground vibrational state. Vibrational (1 and 2) and rovibrational (3–5) transitions of (b) v2 and (c) v1. (d) Rovibrational transitions (6–8) of v3.

Close modal

Tables IV and V list energies and the main components of rotational states in the ground vibrational state. The rotational energies are in qualitative agreement with recent computational estimates.51 The 2J + 1 degeneracy of rotational states is fully removed by the electrostatic field interacting with dipole and quadrupole moments of H2O.

TABLE IV.

para-H2O@C60 rotational energies and wavefunctions in the ground vibrational state calculated with the best fit parameters from Table III. Wavefunction components with the absolute value of the amplitude less than 0.1 are omitted; J,k,mJ,k¯,m¯.

JKaKcEnergy (cm−1)Wavefunction in the symmetric top basis J,k,m
000 −0.99 0,0,0 
111 33.07 0.49 1,1¯,1¯+1,1¯,11,1,1¯1,1,1 
 33.19 −0.69 1,1¯,01,1,0 
 33.75 0.48 1,1¯,1¯1,1¯,11,1,1¯+1,1,1 
202 67.04 −0.96 2,0,0+0.152,2¯,0+2,2,0 
 67.73 0.68 2,0,1¯+2,0,1 
 68.35 0.68 2,0,1¯2,0,1 
 71.27 0.68 2,0,2¯2,0,2 
 71.30 −0.68 2,0,2¯+2,0,2 
211 94.65 −0.4 2,1¯,0+2,1,0+0.412,1¯,2¯+2,1¯,2+2,1,2¯+2,1,2 
 94.72 −0.49 2,1¯,2¯2,1¯,2+2,1,2¯2,1,2 
 94.85 −0.53 2,1¯,0+2,1,00.242,1¯,2¯+2,1¯,2+2,1,2¯+2,1,2 
  0.212,1¯,1¯+2,1¯,12,1,1¯+2,1,1 
 94.91 0.45 2,1¯,1¯+2,1¯,12,1,1¯+2,1,10.212,1¯,0+2,1,0 
  0.152,1¯,2¯+2,1¯,2+2,1,2¯+2,1,2 
 95.01 0.49 2,1¯,1¯+2,1¯,1+2,1,1¯+2,1,1 
JKaKcEnergy (cm−1)Wavefunction in the symmetric top basis J,k,m
000 −0.99 0,0,0 
111 33.07 0.49 1,1¯,1¯+1,1¯,11,1,1¯1,1,1 
 33.19 −0.69 1,1¯,01,1,0 
 33.75 0.48 1,1¯,1¯1,1¯,11,1,1¯+1,1,1 
202 67.04 −0.96 2,0,0+0.152,2¯,0+2,2,0 
 67.73 0.68 2,0,1¯+2,0,1 
 68.35 0.68 2,0,1¯2,0,1 
 71.27 0.68 2,0,2¯2,0,2 
 71.30 −0.68 2,0,2¯+2,0,2 
211 94.65 −0.4 2,1¯,0+2,1,0+0.412,1¯,2¯+2,1¯,2+2,1,2¯+2,1,2 
 94.72 −0.49 2,1¯,2¯2,1¯,2+2,1,2¯2,1,2 
 94.85 −0.53 2,1¯,0+2,1,00.242,1¯,2¯+2,1¯,2+2,1,2¯+2,1,2 
  0.212,1¯,1¯+2,1¯,12,1,1¯+2,1,1 
 94.91 0.45 2,1¯,1¯+2,1¯,12,1,1¯+2,1,10.212,1¯,0+2,1,0 
  0.152,1¯,2¯+2,1¯,2+2,1,2¯+2,1,2 
 95.01 0.49 2,1¯,1¯+2,1¯,1+2,1,1¯+2,1,1 
TABLE V.

ortho-H2O@C60 rotational energies, counted from the para ground state 000, and wavefunctions in the ground vibrational state calculated with the best fit parameters from Table III. Wavefunction components with the absolute value of the amplitude less than 0.1 are omitted; J,k,mJ,k¯,m¯.

JKaKcEnergy (cm−1)Wavefunction in the symmetric top basis J,k,m
101 20.89 0.97 1,0,0 
 24.58 −0.69 1,0,1¯+1,0,1+0.161,1¯,0+1,1,0 
 25.51 −0.7 1,0,1¯1,0,1 
110 38.88 0.49 1,1¯,1¯1,1¯,1+1,1,1¯1,1,1 
 40.03 0.48 1,1¯,1¯+1,1¯,1+1,1,1¯+1,1,1+0.241,0,0 
 43.62 0.69 1,1¯,0+1,1,0 
212 72.29 0.71 2,1¯,02,1,0 
 73.12 −0.49 2,1¯,1¯+2,1¯,12,1,1¯2,1,1 
 73.36 −0.49 2,1¯,1¯2,1¯,12,1,1¯+2,1,1 
 75.87 −0.49 2,1¯,2¯2,1¯,22,1,2¯+2,1,2 
 75.88 −0.49 2,1¯,2¯+2,1¯,22,1,2¯2,1,2 
221 119.5 −0.49 2,2¯,2¯+2,2¯,22,2,2¯2,2,2 
 119.5 0.49 2,2¯,2¯2,2¯,22,2,2¯+2,2,2 
 122.2 0.49 2,2¯,1¯2,2¯,12,2,1¯+2,2,1 
 122.2 0.49 2,2¯,1¯+2,2¯,12,2,1¯2,2,1 
 123.1 0.71 2,2¯,02,2,0 
JKaKcEnergy (cm−1)Wavefunction in the symmetric top basis J,k,m
101 20.89 0.97 1,0,0 
 24.58 −0.69 1,0,1¯+1,0,1+0.161,1¯,0+1,1,0 
 25.51 −0.7 1,0,1¯1,0,1 
110 38.88 0.49 1,1¯,1¯1,1¯,1+1,1,1¯1,1,1 
 40.03 0.48 1,1¯,1¯+1,1¯,1+1,1,1¯+1,1,1+0.241,0,0 
 43.62 0.69 1,1¯,0+1,1,0 
212 72.29 0.71 2,1¯,02,1,0 
 73.12 −0.49 2,1¯,1¯+2,1¯,12,1,1¯2,1,1 
 73.36 −0.49 2,1¯,1¯2,1¯,12,1,1¯+2,1,1 
 75.87 −0.49 2,1¯,2¯2,1¯,22,1,2¯+2,1,2 
 75.88 −0.49 2,1¯,2¯+2,1¯,22,1,2¯2,1,2 
221 119.5 −0.49 2,2¯,2¯+2,2¯,22,2,2¯2,2,2 
 119.5 0.49 2,2¯,2¯2,2¯,22,2,2¯+2,2,2 
 122.2 0.49 2,2¯,1¯2,2¯,12,2,1¯+2,2,1 
 122.2 0.49 2,2¯,1¯+2,2¯,12,2,1¯2,2,1 
 123.1 0.71 2,2¯,02,2,0 

From our fit, the permanent dipole moment μx of the encapsulated water is given by the absorption cross section of the IR rotational transitions 3–5 in the ground vibrational state. With the value of μx in hand and by using the intensities of transitions 1 and 2, we were able to determine the internal static electric field in solid C60. The interaction of μx with the crystal electric field mixes rotational states within ground and excited vibrational states. For example, in case of ortho-water, the components of state 110 are mixed into the ground state 101, Table V. This mixing gives the oscillator strength to the pure vibrational transitions 1 and 2. As shown in Table III, the fitted value of the electric field at the C60 cage centers is (110 ± 5)106 V m−1.

A splitting of 4 cm−1 is observed for transition 7 and is due to the splitting of the ortho ground state 101 [Fig. 1(c)]. In principle, a splitting could be caused by the interaction of the water electric dipole with an electric field, or by the interaction of the water electric quadrupole moment with an electric field gradient. The electric field 110 × 106 V m−1 is too small to cause splitting of this magnitude. This electric field lifts the degeneracy of m = ±1 levels of 101, but the gap between m = 0 and m = ±1 levels is due to the quadrupolar interaction. As the splitting is determined by the product of VQ and Qzz, it is not possible to have an estimate of how much is the water quadrupole moment Qzz screened in C60.

All eight frequencies of the encapsulated H2O vibrations found in this work are red-shifted relative to those of free water; see Table II. The red shift of the vibrational frequency has been observed for other endofullerenes, H2@C60, 10,12 HD and D2@C60,8 and HF@C60.5 Six water modes have a relative shift between −2.3% and −2.5%. Two frequencies, namely, the bond-bending mode frequency ω2 and its overtone frequency 2ω2, are shifted by −1.6% and −1.5%, respectively.

The observed red shifts of the stretching mode frequencies ω1 and ω3 are only partially consistent with previous density functional theory (DFT) calculations. The DFT-based calculations published by Varadwaj et al.52 do predict vibrational red shifts, while some of the calculations reported by Farimani et al.53 predict blue shifts rather than red shifts.

The calculation predicts a blue shift of the bending mode ω2 although ten times less in absolute value than the predicted shift of stretching modes.52 The experimental shift of ω2 is less than that of stretching modes, but it is still red-shifted. The other method, fully coupled nine-dimensional calculation, predicts blue shifts for all three vibrational modes.51 

Table VI lists the measured translational energies from the ground to the first excited state, ωt, of small-molecule endofullerenes. It is known that the potential of di-hydrogen in C60 is anharmonic, 8,12 while the degree of anharmonicity of HF@C60 and H2O@C60 potentials is not known. For simplicity, we assume that the potential is harmonic for the current case of endohedral molecules, ωtωt0, and show its scaling relative to H2 in Table VI. In this approximation, the harmonic potential parameter is similar among the hydrogen isotopologs, but for HF and H2O, V2 is larger by a factor of 1.9 and 3.4, respectively. The steeper translational potential for HF and H2O, relative to dihydrogen, is consistent with the larger size of these molecules and, hence, their tighter confinement. The last line of Table VI is the frequency and the harmonic potential of H2O@C60 derived by Bacic and co-workers16,51 using Lennard-Jones potentials. The calculated potential is steeper than the experimentally determined potential.

TABLE VI.

Translational energies from the ground to the first excited state, ωt, of small-molecule endofullerenes and scaling of the harmonic spherical potential V2 for the translational motion of an endohedral molecule, of mass mi, relative to the H2@C60 potential V2H2. The anharmonicity is neglected, ωtωt0, where ωt0 is the frequency of an harmonic oscillator [Eq. (18)].

Moleculemiωt (cm−1)V2/V2H2References
H2 179.5 12  
HD 157.7 1.16 8  
D2 125.9 0.98 8  
HF 20 78.6 1.92 5  
H218 110 3.4 This work 
H218 162 7.3 Theory16,51 
Moleculemiωt (cm−1)V2/V2H2References
H2 179.5 12  
HD 157.7 1.16 8  
D2 125.9 0.98 8  
HF 20 78.6 1.92 5  
H218 110 3.4 This work 
H218 162 7.3 Theory16,51 

As seen in Fig. 2(b), the absorption line of the ortho-H2O translational mode is split by 2.9 cm−1. This splitting may be attributed to the coupling between the endohedral molecule translation and rotation, associated with the interaction of the non-spherical rotating molecule with the interior of the C60 cage, as seen in H2@C60.8,10,12 For the particular transition, shown in Fig. 2(b), it is the coupling between the translational state with N = 1, L = 1 and the rotational state 101 of ortho-water. Within spherical symmetry, a good quantum number is Λ = J + L. Thus, the translation–rotation coupled translational state L = 1 and rotational state J = 1 form three states with Λ-values 0, 1, and 2. The calculated energy difference of ortho-water Λ = 0 and Λ = 1 states is 8 cm−116 as compared to the experimental value 2.9 cm−1.

There are two possibilities why the rotational constants of water change when it is encapsulated. The first is that the bond length and angles of H2O change. The second is that H2O, because of confinement and being non-centrosymmetric, is forced to rotate about the “center of interaction,” which does not coincide with its nuclear center of mass.54 

The rotational constant relates to the moment of inertia Iaa as A=h(8π2c0Iaa)1, where c0 is the speed of light. Similarly, B=h(8π2c0Ibb)1 for the b-axis rotation and C=h(8π2c0Icc)1 for the c-axis rotation ([A] = m−1 in SI units and 0.01[A] = cm−1). The moments of inertia are Iαα=imi(βi2+γi2), where {αi, βi, γi} are the Cartesian coordinates of the ith nucleus with mass mi with the origin at the nuclear center of mass.

For non-centrosymmetric molecules, the translation–rotation coupling shifts the rotational energy levels. In quantum mechanical terms, the shift of rotational states is caused by the mixing of rotational and translational states by translation–rotation coupling, for example, HD@C60.8,55 Translation–rotation coupling was not included in our quantum mechanical model of H2O@C60. The rotational constants of the model were free parameters to capture the effect of translation–rotation coupling and the change in the H2O molecule geometry caused by the C60 cage. The rotational constants of free water are A0 = 27.88 cm−1, B0 = 14.52 cm−1, and C0 = 9.28 cm−1 in the ground vibrational state.56 The rotational constants of endohedral water have relative shifts −13%, 5.5%, −8.7% for A0, B0, and C0, in the ground vibrational state, Table III. In the following discussion, we use classical arguments to assess whether the shift of rotational levels is caused by the change in water molecule geometry or by the translation–rotation coupling.

From the symmetry of the H2O molecule, it is seen that the nuclear center of mass is on the b axis (Fig. 1). The shift of H2O center of rotation in the negative direction of b axis decreases A and C, while it does not affect B. If B changes, it must be due to the change in H–O–H bond angle and O–H bond length. The calculation predicts the lengthening of the H–O bond by 0.0026 Å and decrease in the H–O–H bond angle by 0.87° of caged water with respect to free water.52 With these parameters, the relative change in rotational constants A, B, and C is −2.5%, 0.65%, and −0.46%, an order of magnitude smaller than those derived from the IR spectra of H2O@C60. However, shifting the rotation center by −0.07 Å in the b direction (further away from oxygen) gives relative changes −14%, 0.65%, and −5.2%. This relative change in A000 and C000 is not very different from the values derived from the IR spectra, while the relative change in B000 is within the error limits, Table III. Thus, it is likely that the dominant contribution to the observed changes in rotational constants in the ground translational state comes not from the change in H2O molecule bond lengths and bond angle but from shifting its center of rotation away from the nuclear center of mass of the H2O molecule.

A comparison of the normalized absorption cross sections ⟨σji⟩ [Eq. (3)] for H2O@C60 and for free water is given in the last column of Table I. In general, for all observed transitions, the absorption cross section of endohedral water is smaller than that of free water. The v1 mode has the smallest relative change, while the largest relative change is for the combination mode ω1 + ω3.

The comparison of H2O@C60 and free H2O absorption cross sections (Table I) enables us to estimate independently from the spectral fit the permanent and transition dipole moments of encapsulated H2O,

(28)

The results are collected together with the dipole moments obtained from the fit of IR spectra in Table VII. The permanent dipole moment of encapsulated water is nearly four times smaller than that of free H2O. The results of the IR spectroscopy study are consistent with the dipole moments determined by the capacitance method, 0.51 ± 0.05 D.17 The reduction of the fullerene-encapsulated water dipole moment has been predicted by several theoretical calculations.18,52,60,61

TABLE VII.

Absolute values of the dipole moment, unit D, matrix elements of rotational [μx, Eq. (B13)] and rovibrational [μ01x,μ02x and μ03z, Eq. (B14)] transitions of gaseous H2O, as published, and of H2O@C60 calculated with Eq. (28) from the rotational and rovibrational absorption cross sections, Table I, and determined by the fit of IR spectra, Table III. The cross section-derived dipole moments are the cross section-error weighted averages of the three transitions, 3, 4, and 5 or 6, 7, and 8 in Table I.

GasReferencesCross sectionFit
μx 1.855 57  0.50 ± 0.05 0.474 ± 0.008 
μ01x 0.0153 58  (1.23 ± 0.13)10−2 (1.031 ± 0.021)10−2 
μ02x 0.1269 59  (3.38 ± 0.19)10−2 (3.40 ± 0.05)10−2 
μ03z 0.0684 58  (3.0 ± 0.3)10−2 (2.83 ± 0.07)10−2 
GasReferencesCross sectionFit
μx 1.855 57  0.50 ± 0.05 0.474 ± 0.008 
μ01x 0.0153 58  (1.23 ± 0.13)10−2 (1.031 ± 0.021)10−2 
μ02x 0.1269 59  (3.38 ± 0.19)10−2 (3.40 ± 0.05)10−2 
μ03z 0.0684 58  (3.0 ± 0.3)10−2 (2.83 ± 0.07)10−2 

The electric dipolar and quadrupolar interactions of the H2O@C60 molecule with the electrostatic field in solid C60 explain the oscillator strength of the pure v1 and v2 vibrational transitions and the splitting of the rotational states with J > 0.

The theoretical work of Felker et al.32 shows that the source of the quadrupole crystal field is the orientation of electron-rich double bonds of 12 nearest neighbors of C60 relative to the central H2O@C60. When the solid C60 is cooled below 90 K, a merohedral disorder is frozen where ∼85% of C60 have the electron-rich 6:6 bond (bond between the two hexagonal rings) facing pentagonal rings of a neighboring cage, the P-orientation.62 The rest are H-oriented where the 6:6 bond faces neighboring cage’s hexagonal rings. The calculated quadrupolar interaction for the P-oriented molecules is ten times bigger than that for the H-oriented molecules.32 The electric field gradient couples to the quadrupole moment of water and splits the 101 rotational state by 4.2 cm−1, where the m = ±1 doublet is above the m = 0 level.32 This theoretically predicted splitting of the 101 state for the P-oriented molecules is remarkably close to the observed experimental value, seen as a 4 cm−1 splitting of line 7, transition starting from the ortho ground state (Fig. 4). It is not possible to determine the crystal electric field gradient tensor and the encapsulated water quadrupole moment separately from our IR spectra.

Further splitting is possible if the symmetry is lower than S6, but the maximum number of components for J = 1 remains three. However, we see that line 7 consists of four components instead of three (Fig. 4). This suggests that there are two sites with different local electrostatic fields. Anomalous splitting of triply degenerate phonons into quartets has been seen in solid C60 by IR spectroscopy34 and was attributed to the merohedral disorder. Thus, our work and the IR study of phonons34 clearly show that there are two different quadrupolar interactions in solid C60. As was proposed by Felker et al.,32 the crystal field has a different magnitude for P-oriented sites and for H-oriented sites. The small population of H-oriented sites (about 15%) justifies our spectral fitting with a single quadrupolar interaction.

We assumed that there is an internal electric field in C60, and this field is a possible reason why the pure vibrational transitions 1 and 2 [see Fig. 1(c)] become visible in the IR spectrum. It is also plausible that 1 and 2 gain intensity through the translation–rotation coupling from the induced dipole moment of translational motion. However, there is evidence that local electric fields exist in solid C60 as a result of merohedral disorder C60.35 The estimate of un-balanced charge by Alers et al.35 was q = 6 × 10−3e assuming a dipole moment μ = qd0, where e is the electron charge and d0 = 0.7 nm is the diameter of a C60 molecule. Our estimate is that the electric field E=1.1×108 V m−1 at the center of the C60 cage is created by the dipole moment with charge q = 4.7 × 10−3e. These two estimates are very close.

C60 has six nearest-neighbor equatorial cages and three axial cages above and three axial cages below the equatorial plane, following the notation of Ref. 32. The z′ axis of the crystal field coordinate frame is normal to the equatorial plane. As our fit shows, the electric field is rotated away from the z′ axis by θE ≈ 80°, Table III, almost into the equatorial plane. It is possible that one of the six nearest-neighbors in the equatorial plane does not have P-orientation and this mis-oriented cage is the source of the electric field. θE has a large relative error consistent with the probability to have the mis-oriented cage in the equatorial or in the axial position.

The infrared absorption spectra of solid H2O@C60 samples were measured close to liquid He temperature, and rotational, vibrational, rovibrational, overtone, and combination rovibrational transitions of H2O were seen. The spectral lines were identified as para-water and ortho-water transitions by following the paraortho conversion process over the timescale of hours. The vibrational frequencies are shifted by −2.4% relative to free water, except bending mode frequency ω2 and its overtone 2ω2, where the shift is −1.6%. An absorption mode due to the quantized center-of-mass motion of H2O in the molecular cage of C60 was observed at 110 cm−1. The dipole moment of encapsulated water is 0.50 ± 0.05 D, which is approximately four times less than that of free water and agrees with previous estimates.17 

The rotational and rovibrational spectra were fitted with a quantum mechanical model of a vibrating rotor in the electrostatic field with dipolar and quadrupolar interactions. The quadrupolar interaction splits the J ≥ 1 rotational states of H2O. The source of quadrupolar interaction is the relative orientation of electron-rich chemical bonds relative to pentagonal and hexagonal motifs of C60 and its 12 nearest neighbors.32 Further IR study by using pressure to change the ratio of P-oriented and H-oriented motifs63 would provide more information on the quadrupolar crystal fields of these motifs. The finite oscillator strength of the fundamental vibrational transitions is attributed to a finite electric field at the centers of C60 cages due to merohedral disorder, as has been postulated in different contexts.35 Our results are consistent with an internal electric field of 108 V m−1. However, it is also plausible that the fundamental vibrational transitions gain intensity through the translation–rotation coupling from the dipole moment induced by the translational motion, something that can be addressed in further theoretical studies.

To conclude, H2O in the molecular cavity of C60 behaves as a vibrating asymmetric top, its dipole moment is reduced, and the translational motion is quantized. The splitting of rotational levels is caused by the quadrupolar interaction with the crystal field of solid C60. Evidence is found for the existence of a finite electric field at the centers of C60 cages in water-endofullerene due to merohedral disorder.

Two out of three components necessary for a rigorous, comprehensive description of the water translations, rotations, and vibrations inside the C60 molecular cage are now in place: first, the infrared spectroscopy data reported here and, second, the nine-dimensional quantum bound-state methodology51 plus the theory of symmetry breaking in solid C60.32,33 What is missing is a high-quality ab initio nine-dimensional potential energy surface for this system.

We thank Professor Zlatko Bačić for useful discussions. This research was supported by the Estonian Ministry of Education and Research through institutional research funding under Grant No. IUT23-3 and personal research funding under Grant No. PRG736 and the European Regional Development Fund under Project No. TK134. We acknowledge the EPSRC (UK) for support under Grant Nos. EP/P009980/1 and EP/T004320/1.

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

The dipole moment in spherical components is40 

(A1)

The dipole moment of water in the Cartesian molecule coordinate frame, as shown in Fig. 1, is μM = {−μx, 0, 0}, where we use a convention that the dipole moment is directed from the negative charge to the positive charge. Then, from Eq. (A1), the dipole moment in spherical components is

(A2)

We consider the coupling of the H2O dipole moment to the local electric field, E, with spherical coordinates {E,ϕE,θE} in the crystal frame C (frame where the electric field gradient tensor is defined) and in the coordinate frame E of electric field {EmE}=E{0,1,0}, i.e., along the zE axis. Corresponding Euler angles are ΩCE = {ϕE, θE, 0} and ΩEC = {π, θE, −πϕE}. The dipole moment of the molecule in frame E is

(A3)

Here, we used Wigner D-functions relating the components of a spherical rank j irreducible tensor Tjm in coordinate frames A and B,39 

(A4)

and

(A5)

where ΩAB = {ϕ, θ, χ} are Euler angles transforming coordinate frame A into frame B. The angles for the inverse transformation are ΩBA = {πχ, θ, −πϕ}.41 

The interaction of the molecular dipole moment μmM with the electric field is

The minus sign in front of sum in the formula above is consistent with Hed=Eμ if the spherical components of vectors E and μ are defined as in (A1).

Here, we derive the electric dipole transition matrix elements of a molecule, part (21) of the absorption cross section (20). The derivation of (21) starts from

(B1)

where

(B2)

where the electric field of radiation, e={emA}, and the dipole moment of a molecule, μmA, are in the space-fixed frame A; e2=m=11(emA)2.

The Euler angles of transformation of the crystal frame C to the space-fixed frame A, and vice versa, are ΩCA = {ϕR, θR, 0} and ΩAC = {π, θR, −πϕR}.

The dipole moment in frame A is

(B3)

where μmC is the dipole moment in the crystal frame C.

The absolute value of matrix element squared is

(B4)

1. Random orientation of crystals

We assume that all the crystals are identical and there are no other static fields outside the crystal that could break the directional isotropy. We average (B4) over the random orientation of crystal coordinate frames with respect to the space-fixed frame, ΩAC, and get

(B5)

where we used the property of rotation matrices,39 

(B6)

and

(B7)

If the sample is in the powder form, then it follows from Eq. (B5) that the absorption is independent of light polarization.

2. Transition matrix element and separation of coordinates

The absorption of radiation by a molecule [Eq. (20)] depends on the matrix elements of an electric dipole moment between the initial and final states,

(B8)

where μσA is the molecule dipole moment in the space-fixed coordinate frame. The molecule wavefunction consists of the nuclear spin wavefunction ImI, electron wavefunction Φe, electron spin wavefunction SmS, vibration wavefunction Φv, and rotation wavefunction Φr,40 

(B9)

Using these separable wavefunctions, the matrix element (B8) is40 

(B10)

where

(B11)

and we have taken into account that the electric dipole moment does not depend on nuclear and electron spin coordinates. Furthermore, if the energies of initial and final states of the transition are independent of spin projections mI and ms, the summation over initial and final states in the transition probability leads to degeneracy factors gI = 2I + 1 and gS = 2S + 1 in (20). The electron spin is zero in the ground electronic state of H2O, and thus, gS = 1. Degeneracy of para molecules (I = 0) is gI(p)=1, and that of ortho molecules (I = 1) is gI(o)=3.

If the electronic orbital does not change in the transition, then (B8) is the molecule electric dipole moment in the ground electronic state Φe,

(B12)

For the rest of the discussion, we use a shorthand notation μσM for the molecule dipole moment in the ground electronic state, μσM(eg).

Using Cartesian coordinates, the dipole moment in the ground vibrational state is

(B13)

In the quantum mechanical model of H2O@C60 that we used to fit the IR spectra, we set the dipole moment equal to μx in three excited vibrational states 100, 010, and 001.

The vibrational transition dipole moments are

(B14)

The relation between spherical and Cartesian dipole moment components is given by Eq. (A1).

We determined the Hamiltonian parameters and the dipole moments, parameter set a = {a1, …, aν, …, aM}, by finding the parameter set amin that gives the minimum value, χmin2, of function

(C1)

where S(νi) is the synthetic spectrum, with argument frequency νi, generated from the fit of the experimental spectra and f(νi; a) is the spectrum calculated from the model with M parameters a = {a1, …, aν, …, aM}; N is the number of points in the spectrum. The goal is to minimize χ2 over the parameter set a. The result is χmin2 and amin.

Let us define the matrix

(C2)

and the dispersion matrix V¯ [Eq. (12) of Ref. 64],

(C3)

The covariance matrix [Eq. (11) of Ref. 64]

(C4)

where σ2=χmin2/(NM) and V¯1 is the inverse matrix of V¯, V¯1V¯=1¯.

The estimated variance of parameter aν is

(C5)

The correlation matrix C¯ is

(C6)

The element of F¯ [Eq. (C2)] is

(C7)

where amin minimizes χ2 and δaν is a small variation of the parameter aν.

We change the sum for an integral over ν in (C1),

(C8)

where the experimental spectrum is available in several spectral ranges {νk1, νk2} indexed by k. Ak is the weight factor for each spectral range k. Ak is chosen so that the strongest lines for each range k are equal. The spectrum S(ν) is calculated with constant linewidth using line areas and frequencies from the fits of the experimental spectra.

By inserting (C7) into (C3) and using the continuum limit, we get the matrix V¯,

(C9)

We define the number of experimental points N as the number of lines fitted in the experimental spectra multiplied by two as each line has two parameters, area and frequency. We estimate the parameter error with Eq. (C5) using (C8) and (C9).

1.
Y.
Morinaka
,
F.
Tanabe
,
M.
Murata
,
Y.
Murata
, and
K.
Komatsu
, “
Rational synthesis, enrichment, and 13C NMR spectra of endohedral C60 and C70 encapsulating a helium atom
,”
Chem. Commun.
46
,
4532
4534
(
2010
).
2.
S.
Bloodworth
,
G.
Hoffman
,
M. C.
Walkey
,
G. R.
Bacanu
,
J. M.
Herniman
,
M. H.
Levitt
, and
R. J.
Whitby
, “
Synthesis of Ar@C60 using molecular surgery
,”
Chem. Commun.
56
,
10521
10524
(
2020
).
3.
K.
Komatsu
,
M.
Murata
, and
Y.
Murata
, “
Encapsulation of molecular hydrogen in fullerene C60 by organic synthesis
,”
Science
307
,
238
240
(
2005
).
4.
K.
Kurotobi
and
Y.
Murata
, “
A single molecule of water encapsulated in fullerene C60
,”
Science
333
,
613
616
(
2011
).
5.
A.
Krachmalnicoff
,
R.
Bounds
,
S.
Mamone
,
S.
Alom
,
M.
Concistrè
,
B.
Meier
,
K.
Kouřil
,
M. E.
Light
,
M. R.
Johnson
,
S.
Rols
,
A. J.
Horsewill
,
A.
Shugai
,
U.
Nagel
,
T.
Rõõm
,
M.
Carravetta
,
M. H.
Levitt
, and
R. J.
Whitby
, “
The dipolar endofullerene HF@C60
,”
Nat. Chem.
8
,
953
957
(
2016
).
6.
S.
Bloodworth
,
G.
Sitinova
,
S.
Alom
,
S.
Vidal
,
G. R.
Bacanu
,
S. J.
Elliott
,
M. E.
Light
,
J. M.
Herniman
,
G. J.
Langley
,
M. H.
Levitt
, and
R. J.
Whitby
, “
First synthesis and characterization of CH4@C60
,”
Angew. Chem., Int. Ed.
58
,
5038
(
2019
).
7.
A. J.
Horsewill
,
S.
Rols
,
M. R.
Johnson
,
Y.
Murata
,
M.
Murata
,
K.
Komatsu
,
M.
Carravetta
,
S.
Mamone
,
M. H.
Levitt
,
J. Y.-C.
Chen
,
J. A.
Johnson
,
X.
Lei
, and
N. J.
Turro
, “
Inelastic neutron scattering of a quantum translator-rotator encapsulated in a closed fullerene cage: Isotope effects and translation-rotation coupling in H2@C60 and HD@C60
,”
Phys. Rev. B
82
,
081410
(
2010
).
8.
M.
Ge
,
U.
Nagel
,
D.
Hüvonen
,
T.
Rõõm
,
S.
Mamone
,
M. H.
Levitt
,
M.
Carravetta
,
Y.
Murata
,
K.
Komatsu
,
X.
Lei
, and
N. J.
Turro
, “
Infrared spectroscopy of endohedral HD and D2 in C60
,”
J. Chem. Phys.
135
,
114511
(
2011
).
9.
A.
Krachmalnicoff
,
M. H.
Levitt
, and
R. J.
Whitby
, “
An optimised scalable synthesis of H2O@C60 and a new synthesis of H2@C60
,”
Chem. Commun.
50
,
13037
13040
(
2014
).
10.
S.
Mamone
,
M.
Ge
,
D.
Hüvonen
,
U.
Nagel
,
A.
Danquigny
,
F.
Cuda
,
M. C.
Grossel
,
Y.
Murata
,
K.
Komatsu
,
M. H.
Levitt
,
T.
Rõõm
, and
M.
Carravetta
, “
Rotor in a cage: Infrared spectroscopy of an endohedral hydrogen-fullerene complex
,”
J. Chem. Phys.
130
,
081103
(
2009
).
11.
C.
Beduz
,
M.
Carravetta
,
J. Y. C.
Chen
,
M.
Concistré
,
M.
Denning
,
M.
Frunzi
,
A. J.
Horsewill
,
O. G.
Johannessen
,
R.
Lawler
,
X.
Lei
,
M. H.
Levitt
,
Y.
Li
,
S.
Mamone
,
Y.
Murata
,
U.
Nagel
,
T.
Nishida
,
J.
Ollivier
,
S.
Rols
,
T.
Rõõm
,
R.
Sarkar
,
N. J.
Turro
, and
Y.
Yang
, “
Quantum rotation of ortho and para-water encapsulated in a fullerene cage
,”
Proc. Natl. Acad. Sci. U. S. A.
109
,
12894
12898
(
2012
).
12.
M.
Ge
,
U.
Nagel
,
D.
Hüvonen
,
T.
Rõõm
,
S.
Mamone
,
M. H.
Levitt
,
M.
Carravetta
,
Y.
Murata
,
K.
Komatsu
,
J. Y.-C.
Chen
, and
N. J.
Turro
, “
Interaction potential and infrared absorption of endohedral H2 in C60
,”
J. Chem. Phys.
134
,
054507
(
2011
).
13.
A. J.
Horsewill
,
K. S.
Panesar
,
S.
Rols
,
J.
Ollivier
,
M. R.
Johnson
,
M.
Carravetta
,
S.
Mamone
,
M. H.
Levitt
,
Y.
Murata
,
K.
Komatsu
,
J. Y.-C.
Chen
,
J. A.
Johnson
,
X.
Lei
, and
N. J.
Turro
, “
Inelastic neutron scattering investigations of the quantum molecular dynamics of a H2 molecule entrapped inside a fullerene cage
,”
Phys. Rev. B
85
,
205440
(
2012
).
14.
M.
Xu
,
F.
Sebastianelli
,
B. R.
Gibbons
,
Z.
Bačić
,
R.
Lawler
, and
N. J.
Turro
, “
Coupled translation-rotation eigenstates of H2 in C60 and C70 on the spectroscopically optimized interaction potential: Effects of cage anisotropy on the energy level structure and assignments
,”
J. Chem. Phys.
130
,
224306
(
2009
).
15.
M.
Xu
,
F.
Sebastianelli
,
Z.
Bačić
,
R.
Lawler
, and
N. J.
Turro
, “
Quantum dynamics of coupled translational and rotational motions of H2 inside C60
,”
J. Chem. Phys.
128
,
011101
(
2008
).
16.
P. M.
Felker
and
Z.
Bačić
, “
Communication: Quantum six-dimensional calculations of the coupled translation-rotation eigenstates of H2O@C60
,”
J. Chem. Phys.
144
,
201101
(
2016
).
17.
B.
Meier
,
S.
Mamone
,
M.
Concistrè
,
J.
Alonso-Valdesueiro
,
A.
Krachmalnicoff
,
R. J.
Whitby
, and
M. H.
Levitt
, “
Electrical detection of ortho–para conversion in fullerene-encapsulated water
,”
Nat. Commun.
6
,
8112
(
2015
).
18.
B.
Ensing
,
F.
Costanzo
, and
P. L.
Silvestrelli
, “
On the polarity of buckminsterfullerene with a water molecule inside
,”
J. Phys. Chem. A
116
,
12184
12188
(
2012
).
19.
J.
Ceponkus
,
P.
Uvdal
, and
B.
Nelander
, “
The coupling between translation and rotation for monomeric water in noble gas matrices
,”
J. Chem. Phys.
138
,
244305
(
2013
).
20.
M. E.
Fajardo
,
S.
Tam
, and
M. E.
DeRose
, “
Matrix isolation spectroscopy of H2O, D2O, and HDO in solid parahydrogen
,”
J. Mol. Struct.
695-696
,
111
127
(
2004
), Winnewisser Special Issue.
21.
C. M.
Lindsay
,
G. E.
Douberly
, and
R. E.
Miller
, “
Rotational and vibrational dynamics of H2O and HDO in helium nanodroplets
,”
J. Mol. Struct.
786
,
96
104
(
2006
).
22.
K. E.
Kuyanov
,
M. N.
Slipchenko
, and
A. F.
Vilesov
, “
Spectra of the ν1 and ν3 bands of water molecules in helium droplets
,”
Chem. Phys. Lett.
427
,
5
9
(
2006
).
23.
B. P.
Gorshunov
,
E. S.
Zhukova
,
V. I.
Torgashev
,
V. V.
Lebedev
,
G. S.
Shakurov
,
R. K.
Kremer
,
E. V.
Pestrjakov
,
V. G.
Thomas
,
D. A.
Fursenko
, and
M.
Dressel
, “
Quantum behavior of water molecules confined to nanocavities in gemstones
,”
J. Phys. Chem. Lett.
4
,
2015
2020
(
2013
).
24.
E. S.
Zhukova
,
V. I.
Torgashev
,
B. P.
Gorshunov
,
V. V.
Lebedev
,
G. S.
Shakurov
,
R. K.
Kremer
,
E. V.
Pestrjakov
,
V. G.
Thomas
,
D. A.
Fursenko
,
A. S.
Prokhorov
, and
M.
Dressel
, “
Vibrational states of a water molecule in a nano-cavity of beryl crystal lattice
,”
J. Chem. Phys.
140
,
224317
(
2014
).
25.
A. I.
Kolesnikov
,
G. F.
Reiter
,
N.
Choudhury
,
T. R.
Prisk
,
E.
Mamontov
,
A.
Podlesnyak
,
G.
Ehlers
,
A. G.
Seel
,
D. J.
Wesolowski
, and
L. M.
Anovitz
, “
Quantum tunneling of water in beryl: A new state of the water molecule
,”
Phys. Rev. Lett.
116
,
167802
(
2016
).
26.
M. A.
Belyanchikov
,
M.
Savinov
,
Z. V.
Bedran
,
P.
Bednyakov
,
P.
Proschek
,
J.
Prokleska
,
V. A.
Abalmasov
,
J.
Petzelt
,
E. S.
Zhukova
,
V. G.
Thomas
,
A.
Dudka
,
A.
Zhugayevych
,
A. S.
Prokhorov
,
V. B.
Anzin
,
R. K.
Kremer
,
J. K. H.
Fischer
,
P.
Lunkenheimer
,
A.
Loidl
,
E.
Uykur
,
M.
Dressel
, and
B.
Gorshunov
, “
Dielectric ordering of water molecules arranged in a dipolar lattice
,”
Nat. Commun.
11
,
3927
(
2020
).
27.
S.
Mamone
,
J. Y.-C.
Chen
,
R.
Bhattacharyya
,
M. H.
Levitt
,
R. G.
Lawler
,
A. J.
Horsewill
,
T.
Rõõm
,
Z.
Bačić
, and
N. J.
Turro
, “
Theory and spectroscopy of an incarcerated quantum rotor: The infrared spectroscopy, inelastic neutron scattering and nuclear magnetic resonance of H2@C60 at cryogenic temperature
,”
Coord. Chem. Rev.
255
,
938
948
(
2011
).
28.
B.
Meier
,
K.
Kouřil
,
C.
Bengs
,
H.
Kouřilová
,
T. C.
Barker
,
S. J.
Elliott
,
S.
Alom
,
R. J.
Whitby
, and
M. H.
Levitt
, “
Spin-isomer conversion of water at room temperature and quantum-rotor-induced nuclear polarization in the water-endofullerene H2O@C60
,”
Phys. Rev. Lett.
120
,
266001
(
2018
).
29.
Y.
Li
,
J. Y.-C.
Chen
,
X.
Lei
,
R. G.
Lawler
,
Y.
Murata
,
K.
Komatsu
, and
N. J.
Turro
, “
Comparison of nuclear spin relaxation of H2O@C60 and H2@C60 and their nitroxide derivatives
,”
J. Phys. Chem. Lett.
3
,
1165
1168
(
2012
).
30.
S. S.
Zhukov
,
V.
Balos
,
G.
Hoffman
,
S.
Alom
,
M.
Belyanchikov
,
M.
Nebioglu
,
S.
Roh
,
A.
Pronin
,
G. R.
Bacanu
,
P.
Abramov
,
M.
Wolf
,
M.
Dressel
,
M. H.
Levitt
,
R. J.
Whitby
,
B.
Gorshunov
, and
M.
Sajadi
, “
Rotational coherence of encapsulated ortho and para water in fullerene-C60 revealed by time-domain terahertz spectroscopy
,”
Sci. Rep.
10
,
18329
(
2020
).
31.
K. S. K.
Goh
,
M.
Jiménez-Ruiz
,
M. R.
Johnson
,
S.
Rols
,
J.
Ollivier
,
M. S.
Denning
,
S.
Mamone
,
M. H.
Levitt
,
X.
Lei
,
Y.
Li
,
N. J.
Turro
,
Y.
Murata
, and
A. J.
Horsewill
, “
Symmetry-breaking in the endofullerene H2O@C60 revealed in the quantum dynamics of ortho and para-water: A neutron scattering investigation
,”
Phys. Chem. Chem. Phys.
16
,
21330
21339
(
2014
).
32.
P. M.
Felker
,
V.
Vlček
,
I.
Hietanen
,
S.
FitzGerald
,
D.
Neuhauser
, and
Z.
Bačić
, “
Explaining the symmetry breaking observed in the endofullerenes H2@C60, HF@C60, and H2O@C60
,”
Phys. Chem. Chem. Phys.
19
,
31274
31283
(
2017
).
33.
Z.
Bačić
,
V.
Vlček
,
D.
Neuhauser
, and
P. M.
Felker
, “
Effects of symmetry breaking on the translation-rotation eigenstates of H2, HF, and H2O inside the fullerene C60
,”
Discuss. Faraday Soc.
212
,
547
567
(
2018
).
34.
C. C.
Homes
,
P. J.
Horoyski
,
M. L. W.
Thewalt
, and
B. P.
Clayman
, “
Anomalous splitting of the F1u(→ 3Fu) vibrations in single-crystal C60 below the orientational-ordering transition
,”
Phys. Rev. B
49
,
7052
7055
(
1994
).
35.
G. B.
Alers
,
B.
Golding
,
A. R.
Kortan
,
R. C.
Haddon
, and
F. A.
Theil
, “
Existence of an orientational electric dipolar response in C60 single crystals
,”
Science
257
,
511
(
1992
).
36.
L.
Abouaf-Marguin
,
A.-M.
Vasserot
,
C.
Pardanaud
, and
X.
Michaut
, “
Nuclear spin conversion of H2O@C60 trapped in solid xenon at 4.2 K: A new assignment of ν2 rovibrational lines
,”
Chem. Phys. Lett.
480
,
82
85
(
2009
).
37.
S.
Mamone
,
M.
Concistrè
,
E.
Carignani
,
B.
Meier
,
A.
Krachmalnicoff
,
O. G.
Johannessen
,
X.
Lei
,
Y.
Li
,
M.
Denning
,
M.
Carravetta
,
K.
Goh
,
A. J.
Horsewill
,
R. J.
Whitby
, and
M. H.
Levitt
, “
Nuclear spin conversion of water inside fullerene cages detected by low-temperature nuclear magnetic resonance
,”
J. Chem. Phys.
140
,
194306
(
2014
).
38.
C. C.
Homes
,
P. J.
Horoyski
,
M. L. W.
Thewalt
,
B. P.
Clayman
, and
T. R.
Anthony
, “
Effect of isotopic disorder on the Fu modes in crystalline C60
,”
Phys. Rev. B
52
,
16892
16900
(
1995
).
39.
R. N.
Zare
,
Angular Momentum
, Baker Lecture Series (
John Wiley & Sons, Inc.
,
1988
).
40.
P. P.
Bunker
and
P.
Jensen
,
Molecular Symmetry and Spectroscopy
, 2nd ed. (
NRC of Canada
,
1998
), p.
747
.
41.
D. A.
Varshalovich
,
A. N.
Moskalev
, and
V. K.
Khersonskii
,
Quantum Theory of Angular Momentum
(
World Scientific
,
1988
).
42.
C.
Cohen-Tannoudji
,
B.
Diu
, and
F.
Laloë
,
Quantum Mechanics
(
John Wiley & Sons
,
2005
).
43.
J.
Verhoeven
and
A.
Dymanus
, “
Magnetic properties and molecular quadrupole tensor of the water molecule by beam-maser Zeeman spectroscopy
,”
J. Chem. Phys.
52
,
3222
3233
(
1970
).
44.
S. L.
Altmann
and
P.
Herzig
,
Point-Group Theory Tables
, 2nd ed. (
University of Vienna
,
Vienna
,
2011
).
45.
W. H.
Shaffer
, “
Degenerate modes of vibration and perturbations in polyatomic molecules
,”
Rev. Mod. Phys.
16
,
245
259
(
1944
).
46.
D. L.
Dexter
, “
Absorption of light by atoms in solids
,”
Phys. Rev.
101
,
48
55
(
1956
).
47.

Solid C60 has strong optical phonons between 500 and 1400 cm−1. The index of refraction above 1400 cm−1 is 2.0 (Ref. 38). Using the phonon data from Table II (Ref. 38) and Eq. (1) (Ref. 38), we calculate the index of refraction at 50 cm−1, η=R[ε]=2.01. Thus, η = 2.0 is a good approximation for the index of refraction in the frequency range of H2O transitions reported here.

48.
S.
Aoyagi
,
N.
Hoshino
,
T.
Akutagawa
,
Y.
Sado
,
R.
Kitaura
,
H.
Shinohara
,
K.
Sugimoto
,
R.
Zhang
, and
Y.
Murata
, “
A cubic dipole lattice of water molecules trapped inside carbon cages
,”
Chem. Commun.
50
,
524
526
(
2014
).
49.
I. E.
Gordon
,
L. S.
Rothman
,
C.
Hill
,
R. V.
Kochanov
,
Y.
Tan
,
P. F.
Bernath
,
M.
Birk
,
V.
Boudon
,
A.
Campargue
,
K. V.
Chance
,
B. J.
Drouin
,
J.-M.
Flaud
,
R. R.
Gamache
,
J. T.
Hodges
,
D.
Jacquemart
,
V. I.
Perevalov
,
A.
Perrin
,
K. P.
Shine
,
M.-A. H.
Smith
,
J.
Tennyson
,
G. C.
Toon
,
H.
Tran
,
V. G.
Tyuterev
,
A.
Barbe
,
A. G.
Császár
,
V. M.
Devi
,
T.
Furtenbacher
,
J. J.
Harrison
,
J.-M.
Hartmann
,
A.
Jolly
,
T. J.
Johnson
,
T.
Karman
,
I.
Kleiner
,
A. A.
Kyuberis
,
J.
Loos
,
O. M.
Lyulin
,
S. T.
Massie
,
S. N.
Mikhailenko
,
N.
Moazzen-Ahmadi
,
H. S. P.
Müller
,
O. V.
Naumenko
,
A. V.
Nikitin
,
O. L.
Polyansky
,
M.
Rey
,
M.
Rotger
,
S. W.
Sharpe
,
K.
Sung
,
E.
Starikova
,
S. A.
Tashkun
,
J. V.
Auwera
,
G.
Wagner
,
J.
Wilzewski
,
P.
Wcisło
,
S.
Yu
, and
E. J.
Zak
, “
The HITRAN 2016 molecular spectroscopic database
,”
J. Quant. Spectrosc. Radiat. Transfer
203
,
3
69
(
2017
).
50.
J.
Tennyson
,
N. F.
Zobov
,
R.
Williamson
,
O. L.
Polyansky
, and
P. F.
Bernath
, “
Experimental energy levels of the water molecule
,”
J. Phys. Chem. Ref. Data
30
,
735
(
2001
).
51.
P. M.
Felker
and
Z.
Bačić
, “
Flexible water molecule in C60: Intramolecular vibrational frequencies and translation-rotation eigenstates from fully coupled nine-dimensional quantum calculations with small basis sets
,”
J. Chem. Phys.
152
,
014108
(
2020
).
52.
A.
Varadwaj
and
P. R.
Varadwaj
, “
Can a single molecule of water be completely isolated within the subnano-space inside the fullerene C60 cage? A quantum chemical prospective
,”
Chem. - Eur. J.
18
,
15345
15360
(
2012
).
53.
A. B.
Farimani
,
Y.
Wu
, and
N. R.
Aluru
, “
Rotational motion of a single water molecule in a buckyball
,”
Phys. Chem. Chem. Phys.
15
,
17993
18000
(
2013
).
54.
H.
Friedmann
and
S.
Kimel
, “
Theory of shifts of vibration–rotation lines of diatomic molecules in noble-gas matrices. Intermolecular forces in crystals
,”
J. Chem. Phys.
43
,
3925
3939
(
1965
).
55.
M.
Xu
,
F.
Sebastianelli
,
Z.
Bačić
,
R.
Lawler
, and
N. J.
Turro
, “
H2, HD, and D2 inside C60: Coupled translation-rotation eigenstates of the endohedral molecules from quantum five-dimensional calculations
,”
J. Chem. Phys.
129
,
064313
(
2008
).
56.
R. A.
Toth
, “
ν2 band of H216O: Line strengths and transition frequencies
,”
J. Opt. Soc. Am. B
8
,
2236
2255
(
1991
).
57.
S. A.
Clough
,
Y.
Beers
,
G. P.
Klein
, and
L. S.
Rothman
, “
Dipole moment of water from Stark measurements of H2O, HDO, and D2O
,”
J. Chem. Phys.
59
,
2254
2259
(
1973
).
58.
J. M.
Flaud
and
C.
Camy-Peyret
, “
Vibration-rotation intensities in H2O-type molecules application to the 2ν2, ν1, and ν3 bands of H216O
,”
J. Mol. Spectrosc.
55
,
278
310
(
1975
).
59.
C.
Camy-Peyret
and
J.-M.
Flaud
, “
Line positions and intensities in the ν2 band of H216O
,”
Mol. Phys.
32
,
523
537
(
1976
).
60.
C. N.
Ramachandran
and
N.
Sathyamurthy
, “
Water clusters in a confined nonpolar environment
,”
Chem. Phys. Lett.
410
,
348
351
(
2005
).
61.
K.
Yagi
and
D.
Watanabe
, “
Infrared spectra of water molecule encapsulated inside fullerene studied by instantaneous vibrational analysis
,”
Int. J. Quantum Chem.
109
,
2080
(
2009
).
62.
P. A.
Heiney
, “
Structure, dynamics and ordering transition of solid C60
,”
J. Phys. Chem. Solids
53
,
1333
1352
(
1992
).
63.
B.
Sundqvist
, “
Fullerenes under high pressures
,”
Adv. Phys.
48
,
1
134
(
1999
).
64.
D. L.
Albritton
,
A. L.
Schmeltekopf
, and
R. N.
Zare
, “
An introduction to the least-square fitting of spectroscopic data
,” in
Molecular Spectroscopy: Modern Research
(
Academic Press, Inc.
,
New York
,
1976
), pp.
1
67
.