Tanzeeha
Jafari
a,
George
Razvan Bacanu
b,
Anna
Shugai
a,
Urmas
Nagel
a,
Mark
Walkey
b,
Gabriela
Hoffman
b,
Malcolm H.
Levitt
b,
Richard J.
Whitby
b and
Toomas
Rõõm
*a
aNational Institute of Chemical Physics and Biophysics, Akadeemia tee 23, Tallinn 12618, Estonia. E-mail: toomas.room@kbfi.ee
bSchool of Chemistry, University of Southampton, SO17 1BJ Southampton, UK
First published on 21st April 2022
We studied the quantized translational motion of single He atoms encapsulated in molecular cages by terahertz absorption. The temperature dependence of the THz absorption spectra of 3He@C60 and 4He@C60 crystal powder samples was measured between 5 and 220 K. At 5 K there is an absorption line at 96.8 cm−1 (2.90 THz) in 3He@C60 and at 81.4 cm (2.44 THz) in 4He@C60, while additional absorption lines appear at higher temperature. An anharmonic spherical oscillator model with a displacement-induced dipole moment was used to model the absorption spectra. Potential energy terms with powers of two, four and six and induced dipole moment terms with powers one and three in the helium atom displacement from the fullerene cage center were sufficient to describe the experimental results. Excellent agreement is found between potential energy functions derived from measurements on the 3He and 4He isotopes. One absorption line corresponds to a three-quantum transition in 4He@C60, allowed by the anharmonicity of the potential function and by the non-linearity of the dipole moment in He atom displacement. The potential energy function of icosahedral symmetry does not explain the fine structure observed in the low temperature spectra.
Different methods have been used to synthesize endofullerenes. Endofullerenes may be synthesized by the arc discharge method in which carbon is evaporated at high temperature in the presence of a metal5 or an inert gas.6,7 Insertion of atoms into the fulllerene cage is possible by exposing the fullerene to high temperatures and high pressures of inert gas8 or by ion bombardment.9–11 However, these methods are not suited to capture small molecules and the production yield is small. In a major development, the group of Komatsu12 synthesized H2@C60 endofullerenes by a process known as molecular surgery.13 Following in the footsteps of Komatsu, the group of Murata successfully encapsulated a water molecule inside C60.14 Molecular surgery has been the most successful process to produce high-yield endofullerenes until today. A large number of endofullerenes are now available in macroscopic quantities, such as12 and its isotopologues,15 H2O@C60,14 HF@C6016 and CH4,17 and atoms like Ar,18 He19,20 and Ne.20
Endofullerenes are of great interest due to quantum effects which are more evident at cryogenic temperatures. The encapsulation isolates atoms and molecules from each other preventing them from forming liquid or solid condensates even at the lowest temperatures. This provides a unique opportunity to study their internal degrees of freedom at the lowest temperatures. The translational motion becomes quantized due to the confining potential. The confinement also prompts the coupling of translational and rotational motions of the encapsulated molecule. The combination of small molecular mass and tight confinement inside the nano cage gives rise to a discrete and well separated translation-rotational energy level structure. The most studied endohedral complex is H2@C60, which laid the basis for understanding the quantum dynamics of the isolated systems.4,21 Different spectroscopic techniques have been used to demonstrate the effect of the highly symmetric trapping potential on the quantum dynamics of the hydrogen molecule, including infrared spectroscopy (IR)22–24 and inelastic neutron scattering (INS).25–28
There is a growing interest in extending spectroscopic studies to noble gas endofullerenes. The research interest in noble gas endofullerenes is evident from the theoretical studies that have been done so far.29–36 The incarceration of large noble gas atoms results in the structural and electronic distortion of which has been examined by IR and Raman,34 NMR,37 X-ray38 and electronic spectroscopy.39
The first experimental evidence of a noble gas endofullerene was He@C60. It was spotted by mass spectrometry when the 4 atom was incorporated in C60 as the highly accelerated C60+ ions struck with helium gas9 and later found in fullerenes produced by arc discharge in the He gas.6 Despite that, and the 2010 synthesis of He@C60 by molecular surgery,19 the He endofullerene potential energy function study was restricted to theoretical explanations31,33,40 and the experimental data were inadequate for supporting the available information. Recently, the translational energies of He@C60 were determined by inelastic neutron scattering and THz spectroscopy studies and the experimentally derived potential was compared to estimates from quantum chemistry calculations and from sums of empirical two-body potentials.41
Here, as compared to ref. 41, we present a more detailed temperature-dependence study of THz absorption by endohedral 3 and 4 isotopes between 5 and 220 K. Also, we obtained a better fit of line intensities by including an r3 term in the expansion of the dipole moment in the helium atom displacement r. The temperature dependence and the new fit of line intensities confirms the assignment of the He atom translational energy levels and the accuracy of the derived potential energy surface reported in our previous paper.41
In general, a free He atom does not absorb electromagnetic radiation at THz frequencies due to the lack of an electric dipole moment. However, similar to the encapsulated hydrogen molecule,15,22,23 He atoms acquire a dipole moment from the interaction with the inner surface of C60. This interaction is modulated by the translational motion of the He atom causing the translational modes to become THz-active.
Ĥ = Ĥ0 + ![]() | (1) |
![]() | (2) |
![]() | (3) |
The harmonic Hamiltonian has eigenstates
|n,![]() ![]() ![]() | (4) |
![]() | (5) |
![]() | (6) |
The radial wavefunction is42
![]() | (7) |
![]() | (8) |
The eigenenergies of the harmonic Hamiltonian (3) are
![]() | (9) |
![]() | (10) |
The problem can be solved by diagonalizing the Hamiltonian (1) using the basis set of functions (4) where n ∈ {0, 1,…, nmax}. The angular momentum quantum number is given by
∈ {0, 2… n} for even n and
∈ {1, 3… n} for odd n.
It is practical to use the reduced matrix elements, 〈f‖Tk‖
i〉, of the spherical tensor operator Tkq of rank k, q ∈ {−k, −k + 1,…, k}, which are independent of m and q.46 The matrix element of Tkq is
![]() | (11) |
![]() | (12) |
For the isotropic spherical potential, eqn (2), the rank of potential spherical operator is k = 0. Therefore, the states with different and m are not mixed. Perturbation VNrN mixes states with different n and the eigenstate of Hamiltonian (1), |τ,
,m〉, is a linear superposition of states |n,
,m〉, eqn (4). Evaluation of matrix elements of the radial part of harmonic oscillator wavefunction, 〈Rnf
f|rN|Rni
i〉 with
f =
i, shows that non-zero elements are |nf − ni| ≤ N where ni and nf must have the same parity. Because each energy level of the anharmonic Hamiltonian (1) is (2
+ 1)-fold degenerate in m, it is sufficient to include only the |n,
,0〉 functions in the basis set. This reduces the number of states by a factor 2
+ 1 for each
.
![]() | (13) |
Using the Fermi golden rule we write the THz absorption line area for the radiation polarized linearly in the z direction, which couples to the dipole moment z component, dz = d10. Since the energy E of Hamiltonian (1) is degenerate in initial and final state quantum numbers m we perform the summation over mi and mf for a given transition frequency, ωfi = (Ef − Ei)/hc0. The absorption line area Sfi is
![]() | (14) |
![]() | (15) |
The sum over mi and mf in eqn (14) is
![]() | (16) |
The sublimed powdered sample was put inside the 3 mm diameter hole of cylindrical brass frame and pressed under the vacuum into a pellet. The mass and thickness of the 3He@C60 pellet were 28 mg and 2.16 mm and of the 4He@C60 pellet were 21 mg and 1.72 mm. The brass frame with the pellet was inserted into a sample chamber with two thin-film polypropylene windows and with a vacuum line for pumping and filling with helium heat exchange gas. The sample chamber was in a thermal contact with the cold finger of the cryostat. The cryostat was placed inside the interferometer Vertex 80v sample compartment. The cold finger with the sample chamber was moved up and down by letting the beam through the sample chamber or through a reference hole with 3 mm diameter. The transmission spectra were recorded up to 300 cm−1 using a Hg lamp, a 6 μm Mylar beam splitter, and a 4 K bolometer. The spectral resolution was 0.2 cm−1 which was found to be smaller than the width of the He absorption lines or their sub-components.
The absorption spectra were calculated from where α(ω) is the absorption coefficient,
is the transmission, and
is the amount of light lost in reflections from the pellet surfaces and the sample chamber windows. The measured transmission is
, where Is(ω) is the beam intensity at the bolometer with the sample chamber in the beam and Ir(ω) is the beam intensity at the bolometer with the reference hole in the beam. The reflection losses and the absorption in the sample chamber windows affect the background absorption but do not affect the absorption line areas of helium.
![]() | ||
Fig. 1 Temperature dependence of THz absorption spectra of 3He@C60 (a) and 4He@C60 (b). Absorption lines are numbered according to the transitions between the energy levels as shown in Fig. 3. Inset to (b) shows the 4He@C60 line no. 9 at 5 K. The line intensities are not corrected for the filling factor f0, see eqn (14). |
The spectra from the two isotopes share the same pattern of absorption spectra but with the 4He lines shifted to higher energy relative to the 4He lines. The line intensities of the two isotopes are similar. For 3He, a line at 97 cm−1 is the only line present at 5 K and its intensity decreases with increasing temperature. This indicates that it corresponds to a transition from the ground state. With the rise in temperature, energy levels above the ground state become thermally populated and we observe six more absorption lines: 106, 114, 120, 122, 128 and 138 cm−1.
In the case of 4He@C60, we observed two transitions from the ground state, at 81 cm−1 and at 284 cm−1. By increasing the temperature additional absorption lines become visible at 88, 95, 101, 106, 111, 114 and 126 cm−1. The line no. 9 starts from the ground state as is confirmed by its T dependence, Fig. 7 in the Appendix. This line cannot be detected in 3He@C60 THz spectra because of the strong absorption by the fullerene at its resonance frequency, 340.6 cm−1 in Table 4, but was observed in the inelastic neutron scattering spectrum, line no. e in ref. 41.
The lowest energy peak has a two-component structure which is visible at low temperatures. The separation of components is approximately 0.8 cm−1 for 3He@C60 and 0.7 cm−1 for 4He@C60, Tables 4 and 5 in the Appendix. We did not see any additional structure at the best resolution we used, 0.2 cm−1. The possible origin of line no. 1 and no. 9 splitting is discussed in Section 5.
The synthetic spectra were fitted with a spherical oscillator model, as described in the Appendix, and the result is shown in Fig. 2. The best spherical oscillator model fit parameters are listed in Table 2. As compared to our previous fit41 the extra term, A3r3, was added to the dipole moment expansion, eqn (13). This gives a better fit of the line intensities. The spherical oscillator fit revealed that line number 4 consists of three closely spaced transitions, and lines 5 and 6 of two transitions. The energy diagram together with the numbered transitions is shown in Fig. 3 and the energies of corresponding levels in Table 2. The list of line frequencies and intensities calculated with the spherical oscillator model best fit parameters for 3He (125 K) and 4He (100 K) is given in the Appendix, Tables 4 and 5.
![]() | ||
Fig. 2 Synthetic experimental THz spectrum (solid black line) and the best fit result (blue dashed line) of 3He@C60 at 125K (a) and 4He@C60 at 100 K (b). |
![]() | ||
Fig. 3 Energy level diagram and THz transitions of He@C60. The principal translational quantum number is n and the angular momentum quantum number is ![]() ![]() ![]() ![]() |
In Fig. 4 the areas of the first five THz absorption lines of 3He@C60 and 4He@C60 are compared with the temperature dependence of line areas derived from the spherical oscillator model fit. The line areas at each temperature are calculated using the parameters {V2,V4,V6,A1,A3} obtained from the anharmonic oscillator model fit of 125 K and 100 K spectra of 3He and 4He, respectively. For each line the line areas are normalized to the maximum of the theoretical intensity versus temperature curve. Same normalization factors are applied to each experimental line respectively.
![]() | ||
Fig. 4 Temperature dependence of normalized THz absorption lines areas Sn of 3He@C60, panels (a) and (b) and of 4He@C60, panels (c) and (d). Symbols are experimental line areas with errors from the fit with Gaussian lines and solid lines are theoretical areas calculated with the parameters from Table 1. Line areas have been normalized to the maximum value in the T dependence of theoretical area for each line. The lines are numbered according to the transitions shown in Fig. 3 and listed in Tables 4 and 5. |
![]() | ||
Fig. 5 (a) Potential energy curves of 3He and 4He, V (r) = V2r2 + V4r4 + V6r6, calculated with parameters from Table 1. The harmonic part of 3He potential curve, V2r2, is plotted as blue solid line. The potential curves of two isotopes, green solid line of 3He and red dashed line of 3He, are indistinguishable in this plot. (b) The difference between the potential curves of 3He and 4He, ΔV = V3 − V4, is less than ±0.5 cm−1. |
κ i | 3He@C60 | 4He@C60 | Unit |
---|---|---|---|
V 2 | (2.500 ± 0.015) × 10−3 | (2.46 ± 0.04) × 10−3 | meV pm−2 |
V 4 | (3.64 ± 0.03) × 10−7 | (3.77 ± 0.08) × 10−7 | meV pm−4 |
V 6 | (2.560 ± 0.017) × 10−11 | (2.46 ± 0.06) × 10−11 | meV pm−6 |
A 1 | (3.83 ± 0.10) × 10−4 | (3.73 ± 0.22) × 10−4 | D pm−1 |
A 3 | (1.7 ± 0.3) × 10−8 | (2.3 ± 0.6) × 10−8 | D pm−3 |
3He@C60 | 4He@C60 | ||||||
---|---|---|---|---|---|---|---|
E/cm−1 | ![]() |
n | |ξn|2 | E/cm−1 | ![]() |
n | |ξn|2 |
0 | 0 | 0 | 0.95 | 0 | 0 | 0 | 0.95 |
96.8 | 1 | 1 | 0.87 | 81.4 | 1 | 1 | 0.88 |
202.6 | 2 | 2 | 0.77 | 169.8 | 2 | 2 | 0.79 |
218.3 | 0 | 2 | 0.64 | 182.1 | 0 | 2 | 0.68 |
316.3 | 3 | 3 | 0.65 | 264.5 | 3 | 3 | 0.69 |
340.6 | 1 | 3 | 0.43 | 283.7 | 1 | 3 | 0.48 |
437.3 | 4 | 4 | 0.54 | 365.1 | 4 | 4 | 0.58 |
469.4 | 2 | 6 | 0.39 | 390.6 | 2 | 6 | 0.39 |
483.0 | 0 | 6 | 0.40 | 401.1 | 0 | 6 | 0.40 |
565.1 | 5 | 5 | 0.43 | 471.1 | 5 | 5 | 0.47 |
604.3 | 3 | 7 | 0.36 | 502.0 | 3 | 7 | 0.38 |
625.8 | 1 | 7 | 0.31 | 519.0 | 1 | 7 | 0.35 |
699.3 | 6 | 8 | 0.37 | 582.2 | 6 | 6 | 0.38 |
745.1 | 4 | 10 | 0.29 | 618.2 | 4 | 8 | 0.31 |
774.0 | 2 | 10 | 0.30 | 640.8 | 2 | 10 | 0.30 |
787.4 | 0 | 10 | 0.30 | 650.5 | 0 | 10 | 0.31 |
Experimentally determined translational energies of A@C60 endohedral complexes are summarized in Table 3. Although the anharmonic contributions to the potential have been determined experimentally,15,22,23 a more detailed comparison with is not meaningful as firstly, has translation–rotation coupling terms in the potential and secondly, it misses the V6 term in the potential fit. The shape of the potential curve is not known for other A species. We approximate the translational energy, ωT, as the difference of energies between the ground state and the first excited translational state,. Assuming harmonic approximation, VA2 ≈ MAω012/2, we scaled the potential of other species, VA2, relative to hydrogen, V2H2. The last column demonstrates that scaling of translational frequency of different A species has no predictive power. In general, the interaction of neutral A with can be separated into repulsive interaction and electrostatic interaction expanded in induction and dispersion terms.33 Since has no electric dipole nor quadrupole moment the induction terms are zero. The absence of induction terms may only partially explain the “softness” of potential because the dominant interactions are repulsion and dipole–dipole dispersion for endohedral atoms and molecules.33 In general, modern quantum chemistry calculations of He@C60 give a good description of V(r).41
To further validate the potential parameters of obtained from the fit of single high temperature spectra we compare the temperature dependence of line intensities of measured and calculated spectra, Fig. 4. The parameters determined from the 3He (4He) fit at 125 K (100 K) describe the temperature dependence from 5 to 200 K rather well except for the line no. 5. Other lines, no. 6, no. 7, no. 8 and no. 9, are weak and their intensities could not be determined reliably from the experimental spectra and therefore their T dependence was not analyzed. The intensity of the lowest frequency line no. 1, transition from the ground state, is overestimated by the fit at the lowest temperatures. The discrepancy of calculated and measured T dependence could be due the thermal motion of C60 not taken into account in our model.
The dipole moment of is induced by the displacement from the C60 cage center. The estimate for the displacement of the first transition can be made from the potential energy which is half of the total energy, V(r) = E1/2. Using the data from Table 2, E1 = 222.5 cm−1 (3He) and E1 = 187.9 cm−1 (4He), we obtain the displacement to be about 60 pm. The dipole moment induced by 60 pm displacement is d = A1r + A3r3 = 2.7 × 10−2 D, small compared to permanent dipole moments of molecules. For example, the permanent screened dipole moment of C60 endohedral water is about 0.5 D.49,50
Transition no. 9 starts from the ground state with = 0 and predominantly with n = 0 radial content. The final state is the
= 1 state, see Fig. 3 and Table 2, but with predominantly n = 3 radial part. It is an allowed electric dipole transition with the selection rule Δ
= +1 but three translation quanta are created, Δn = +3. All other lines, no. 1 to no. 8, are ordinary one quantum transitions, Δn = +1. The r3 term in the dipole moment expansion allows the Δn = + 3 excitation by radiation. However, because of the mixing of states by anharmonic potential, n = 0 state is mixed with n = 2 and n = 3 state is the mixed n = 1 state, the linear term in r of the dipole moment expansion with the selection rule Δn = +1 can induce the three-quantum transition. Thus, in there are two factors what activate the three-quantum transition, the anharmonicity of the potential and the r3 term in the dipole moment expansion.
Two absorption lines, no. 1 and no. 9, show partially resolved fine structure at 5 K, Fig. 1. These lines are the transitions from the ground state with = 0 to the state
= 1, see Fig. 3 and Table 2. In the icosahedral potential the first Ag symmetry representation after fully spherical potential, rank k = 0, is a combination of spherical harmonics of rank k = 6.51 Rank 6 term in the potential splits states with
≥ 3, which cannot explain the fine structure of helium line no. 1 and no. 9. To explain the spectral line splitting, a potential with symmetry lower than the icosahedral or two different exohedral environments must be invoked.
Fullerene molecules stop rotating below 90 K.52 This leads to two sites distinguished by the relative orientation of the central cage and its 12 nearest-neighbor C60.53 In the pentagonal orientation the electron-rich double bonds are facing pentagonal rings and in the hexagonal orientation the electron-rich double bonds are facing hexagonal rings of nearest-neighbor cages. It was proposed by Felker et al.54 that the orientational order creates electrostatic field that interacts with the quadrupole moment of endohedral molecule. Their model explains the splitting of J = 1 rotational state observed in H2@C60,22,23,28 HF@C6016 and H2O@C60.49,55,56 This mechanism is not applicable to the helium atom because it does not have rotational degrees of freedom nor quadrupole moment. The icosahedral symmetry of V(r) could still be disturbed by the nearest-neighbor C60 molecules. A splitting similar to the splitting of line no. 1 was observed in H2@C60 for the = 0 →
= 1 transition of para-H2@C60 in the J = 0 rotational state.22,23 Application of pressure changes the relative population of pentagon- and hexagon-oriented molecules.57 The analysis of the inelastic neutron scattering spectra of pressure-treated H2@C60 shows that the potential and the energy levels of are sensitive to the orientation of neighboring cages.28 Similar THz and infrared spectroscopy experiments are planned to verify the effect of orientational order on the potential energy function of endohedral molecules and atoms in C60.
In summary, with the THz absorption spectroscopy we have determined the energy level structure of quantized translational motion of single 3He and 4He atoms trapped in the C60 molecular cages. The fitted potential energy curves and the induced dipole moments of two isotopes overlap with high precision. However, there are deviations between the modeled and measured spectra. Firstly, the fine structure of spectral lines observed in the low temperature spectra cannot be explained by icosahedral symmetry of C60 molecule. Secondly, there are some discrepancies between the measured and modeled spectra in the line intensities and their temperature dependence. Both deviations could be due to the orientational order and the thermal motion of C60 molecules in the solid.
![]() | ||
Fig. 6 THz absorption spectrum of 3He at 125 K (a) and 4He at 100 K (b), black solid line, and the fit with Gaussian lineshapes, blue solid lines. Red dashed line is the sum of Gaussians. The line intensities are not corrected for the filling factor f0, see eqn (14). |
The synthetic spectra were fitted using the anharmonic spherical oscillator model. The reduced basis was limited to nmax = 18 providing 100 states in the reduced basis |n,〉. For a given model and basis, matrix elements of the Hamiltonian, eqn (1), and the dipole operator, eqn (13), were evaluated analytically in a symbolic form using the Mathematica software. At each step of minimizing chi squared,
, the Hamiltonian was diagonalized numerically. Here f(ωn,{κ}) is the theoretical spectrum with the same linewidth and lineshape as the synthetic experimental spectrum; {κ} = {V2,V4,V6,A1,A3} is the set of Hamiltonian and dipole operator fit parameters. Errors were calculated using the method described in ref. 49. Adding the V6 term to V4 in the potential reduced χ2 of 3He and 4He fit by three and two times, respectively. The frequencies and intensities of experimental spectra and spectra calculated using the best fit parameters are given in Tables 4 and 5.
Line no. | Experiment | Model fit | ||
---|---|---|---|---|
f/cm−1 | S/cm−2 | f/cm−1 | S/cm−2 | |
5 K | ||||
1 | 96.6 | 8.5 | 96.8 | 11.7 |
97.4 | 2.2 | — | — | |
9 | — | — | 340.6 | 0.4 |
125 K | ||||
---|---|---|---|---|
1 | 96.9 | 2.7 | 96.8 | 2.7 |
2 | 105.7 | 3.3 | 105.7 | 3.3 |
3 | 113.6 | 2.3 | 113.7 | 2.2 |
4 | 119.8 | 0.6 | 121.0 | 1.1 |
121.6 | 1.8 | 121.4 | 0.8 | |
— | — | 122.4 | 0.5 | |
5 | 128.1 | 0.3 | 127.8 | 0.5 |
— | — | 128.8 | 0.4 | |
7 | 137.7 | 0.5 | 138.0 | 0.6 |
Line no. | Experiment | Model fit | ||
---|---|---|---|---|
f/cm−1 | S/cm−2 | f/cm−1 | S/cm−2 | |
5 K | ||||
1 | 81.2 | 6.1 | 81.4 | 8.0 |
81.9 | 1.2 | — | — | |
9 | 283.3 | 0.3 | 283.7 | 0.3 |
284.3 | 0.4 | — | — |
100 K | ||||
---|---|---|---|---|
1 | 81.3 | 2.1 | 81.4 | 2.0 |
2 | 88.4 | 2.3 | 88.4 | 2.4 |
3 | 94.8 | 1.5 | 94.8 | 1.5 |
4 | 100.4 | 0.4 | 100.6 | 0.7 |
101.0 | 1.3 | 100.8 | 0.6 | |
— | — | 101.6 | 0.3 | |
5 | 105.9 | 0.5 | 106.0 | 0.3 |
— | — | 106.7 | 0.2 | |
6 | 110.6 | 0.1 | 111.0 | 0.09 |
— | — | 111.7 | 0.10 | |
7 | 113.9 | 0.4 | 113.9 | 0.4 |
8 | 125.7 | 0.2 | 125.9 | 0.2 |
This journal is © the Owner Societies 2022 |