Publication Cover
Molecular Physics
An International Journal at the Interface Between Chemistry and Physics
Volume 122, 2024 - Issue 1-2: Special Issue of Molecular Physics in Memory of Prof. Dieter Gerlich
979
Views
1
CrossRef citations to date
0
Altmetric
Festschrift in memory of Dieter Gerlich Special Issue

Gas-phase electronic spectroscopy of nuclear spin isomer separated H2O@C60+ and D2O@C60+

, , , , , & show all
Article: e2173507 | Received 29 Nov 2022, Accepted 18 Jan 2023, Published online: 14 Feb 2023

Abstract

Gas-phase electronic spectra of H2O@C60+ and D2O@C60+ are presented. These data were obtained by one-photon dissociation of weakly bound helium complexes synthesised in a 3 K ion trap. Measurements were recorded in the vicinity of the 2Ag,2BgX2Au electronic transitions of the C60+ cage. Two-colour hole burning experiments enabled nuclear spin isomer pure data to be obtained. The spectra are rich in structure with many absorptions attributed to internal excitation of the encapsulated molecule accompanying the C60+ electronic transition. The experimental data are complemented with density functional theory calculations using the B3LYP functional and 6-31++G** basis set.

GRAPHICAL ABSTRACT

1. Introduction

Molecular surgery has enabled the synthesis of macroscopic quantities of endohedral C60 fullerenes that contain small molecules such as H2  [Citation1], H2O  [Citation2, Citation3], HF [Citation4] and CH4  [Citation5]. These developments have allowed experimental investigation into their unique properties. The endohedral of relevance to the present work is H2O@C60. The encapsulated water is an asymmetric top rotor and displays nuclear spin isomerism, with total nuclear spin I=1 or I=0, referred to as ortho and para, respectively. The ortho isomer is threefold degenerate while para is non-degenerate, resulting in a 3:1 ortho:para ratio in the high-temperature limit. As protons are fermions, the Pauli principle dictates that the overall wave function (Ψtot=ΨeΨvΨrΨns) is antisymmetric with respect to exchange of the two H-nuclei. In the gas phase, ortho and para H2O do not readily interconvert allowing experiments to reveal differences in their chemical reactivity  [Citation6].

A consequence of nuclear spin statistics is the connection between the symmetry of rotational (Ψr) and nuclear spin (Ψns) wavefunctions. The rotations of an asymmetric top are denoted JKaKc, where Ka and Kc are projections in the prolate and oblate symmetric top limits. In the lowest vibrational level of the ground electronic state of gas-phase water and H2O@C60  [Citation7–9], the ortho nuclear spin isomer is associated with Ka+Kc= odd levels and para with Ka+Kc= even. These combinations can switch if the symmetry of one of the other contributing parts of the total wavefunction is changed, which happens, for example, when odd quanta of the asymmetric stretching mode of H2O are excited [Citation7] and upon ionisation to the ground state of H2O+  [Citation10].

Rotational transitions of the encapsulated H2O in H2O@C60 have been observed with IR  [Citation7], THz spectroscopy  [Citation11, Citation12] and inelastic neutron scattering  [Citation8, Citation9]. The experimental results resemble the rotational pattern of free H2O but the surrounding cage affected the rotational constants of H2O  [Citation7]. Furthermore, in the solid phase, the degeneracy of the rotational 101 ortho ground state of H2O is lifted due to crystal field effects  [Citation8, Citation9, Citation13, Citation14]. By observing the intensity change of the rotational transitions in the condensed phase it was possible to show that slow spin forbidden orthopara interconversion takes place and to assign the transitions to the different spin isomers  [Citation7, Citation8]. A change in dielectric constant was also observed upon ortho–para conversion [Citation15]. The decreased intensity of transitions originating from 101 was confirmed by a loss of NMR signal, because of the NMR-inactivity of the para spin isomer  [Citation15, Citation16].

The carbon cage resembles a spherical potential around H2O which is subject to strong quantum effects  [Citation17, Citation18]. For example, the confinement is responsible for the quantisation of H2O translational modes  [Citation7, Citation11, Citation12]. Additionally, computational work [Citation19] predicts frustrated vibrations and rotations, which are yet to be observed in experiments. However, there are features observed in THz spectroscopy [Citation11] attributed to hot bands based on temperature-dependent experiments [Citation12] but these remain unassigned to particular translational or rotational motions. The dipole moment of the H2O molecule is also shielded significantly by the C60 surrounding  [Citation15, Citation19, Citation20].

The influence of H2O on the cage has also been studied [Citation21, Citation22] and the UV-Vis spectrum of H2O@C60 showed negligible difference to the electronic transition(s) of the empty cage in solution  [Citation2]. 13C NMR studies of the endohedral showed no reduced symmetry of the cage in H2O@C60  [Citation23]. However, the singly charged cation H2O@C60+ has not been thoroughly investigated yet. As C60+ is an open shell system, it is expected to show more significant interactions between the encapsulated molecule and the surrounding carbon cage. In this contribution, gas-phase electronic spectra of H2O@C60+ are presented and contrasted to that of C60+  [Citation24]. The data were obtained using He tagging at cryogenic temperatures. The contributions of para and ortho H2O are separated by two-colour hole burning experiments. The assignment of features in the spectra is supported by DFT calculations.

Sections 2 and  3 describe the experimental and computational methods employed, respectively. Section 4 contains the electronic spectra of H2O@C60+ and D2O@C60+ and a discussion of the observed transitions along with their proposed assignments, while conclusions are given in Section 5.

2. Experimental

The H2O@C60 sample was synthesised following a synthetic route known as ‘molecular surgery’  [Citation2, Citation3]. A few mg of 80% filled H2O@C60 (D2O@C60) were heated in an oven to 350 C. Singly charged parent ions were produced from the neutral gas using electron impact at 40 eV. The experiments were carried out using the apparatus described in detail in Refs. [Citation25, Citation26], and only the parameters specific to the present work are given here.

In ion trapping experiments, species with m/z below 738 (740) were removed using the first quadrupole mass spectrometer. The nominal trap temperature was 4 K and the He buffer gas number density was 1015 cm3. These low trap temperatures and high He number densities, coupled with long interaction times (several hundred ms), allow H2O@C60+−Hen (D2O@C60+−Hen) complexes to form, which can be seen in Figure . For example, to generate the spectrum in Figure , H2O@C60+ cations were loaded into the trap for 200 ms by lowering the potential applied to the axial entrance electrode. During the first 500 ms they interact with helium buffer gas, which was subsequently pumped out for 420 ms before the trap contents were extracted and analysed by the second quadrupole mass spectrometer. The trapping process was repeated at a repetition rate of 1 Hz.

Figure 1. Mass spectrum recorded after storage of H2O@C60+ in helium buffer gas at 4 K.

Figure 1. Mass spectrum recorded after storage of H2O@C60+ in helium buffer gas at 4 K.

Photofragmentation spectra of H2O@C60+−He (D2O@C60+−He) were recorded by irradiation of the ion cloud with a continuous wave Ti:Sap laser for 30 ms at the end of the trapping cycle. The irradiation time was controlled by use of a mechanical shutter. The mass-channel of the H2O@C60+−He (D2O@C60+−He) complexes with m/z = 742 (744) is monitored on alternating cycles with (Ni) and without (N0) exposure to the radiation to obtain the depletion of complexes as a function of the laser frequency and account for fluctuations in the number of stored ions. For the two-colour experiments, a second near-infrared CW laser (diode) was used to continuously irradiate the ions during the entire trapping cycle at a fixed (resonant) frequency. This enabled the depletion of ions belonging to one nuclear spin isomer. In this case, the number of H2O@C60+−Hen (D2O@C60+−Hen) complexes remaining in the trap following exposure to both (Ni) lasers or only the diode laser (N0) were monitored to reveal nuclear spin isomer pure spectra.

3. Computational

The equilibrium structure of H2O@C60+ was determined with density functional theory in the unrestricted formalism, using the B3LYP functional and the 6-31++G** basis set. The B3LYP functional was selected for these calculations owing to its generally good performance for large conjugated systems and specifically because it was previously employed to study C60+  [Citation27], in conjunction with its generally smaller spin contamination compared to hybrid functionals with larger contributions of HF exchange  [Citation28]. In addition, weak dispersion-like intra- and inter-molecular noncovalent interactions were included via the D3(BJ) correction term  [Citation29–31]. An integral accuracy cutoff of 1012 in combination with an ultrafine integration grid were requested during the Hessian calculations. Geometry optimisation and vibrational frequency calculations were carried out using the Gaussian16 suite of programs  [Citation32]. Gaussian16 evaluates the molecular Hessian in Cartesian coordinates, and the vibrational frequencies of any isotopically substituted species may therefore easily be obtained on the basis of the full Hessian by changing the atomic masses. The vibrational frequencies for D2O@C60+ were then correlated with those of H2O@C60+ by computing the Duschinsky rotation matrix for the two sets of vibrational normal modes  [Citation33]. The optimised geometry displayed a modest spin contamination, with a final value of S2 resulting in 0.7539. The nature of the stationary point was assessed by vibrational frequency calculations that resulted to be all real. The final H2O@C60+ structure has no symmetry elements and is slightly distorted from C2v, with the two OH bond lengths being 0.965 and 0.966Å and the HOH angle measuring 104.9.

4. Results and discussion

4.1. Electronic spectrum of H2O@C60+: comparison with C60+

The gas-phase electronic spectrum of C60+ below 10 K is presented in Figure  in comparison to that of H2O@C60+. The former shows two intense bands at 10,378 and 10,438  cm1 and two weaker absorptions at 10,603 and 10,674  cm1. The two intense features were assigned to the origin bands of the 2BgX~2Au and 2AgX~2Au electronic transitions of C60+ based on calculations, which predicted a Jahn–Teller distortion of the C60+ cage  [Citation27]. The lowest energy vibrational mode (∼230 cm1) of C60+ in the 2Bg and 2Ag states causes the bands at 10,603 and 10,674  cm1. Comparison of the electronic spectra shown in Figure suggests that only one electronic transition is observed in the H2O@C60+ spectrum instead of the two observed for C60+. The strongest absorption feature of H2O@C60+ appears at 10,429 cm1, which lies between the two C60+ origin bands.

Figure 2. The electronic spectrum of C60+ (top) and H2O@C60+ (bottom), recorded under the same laboratory conditions. The data are photofragmentation spectra of C60+−He and H2O@C60+−He complexes.

Figure 2. The electronic spectrum of C60+ (top) and H2O@C60+ (bottom), recorded under the same laboratory conditions. The data are photofragmentation spectra of C60+−He and H2O@C60+−He complexes.

Additionally, the H2O@C60+ spectrum shows several less prominent features with varying intensities close to 10,429 cm1. The energy separation between the observed bands with respect to the strongest absorption is between 9 and 250 cm1. Most of these fall below the lowest frequency cage vibration and are also too low in energy to be associated with vibrational modes of (free) H2O. However, these energy separations are suggestive of rotational levels of gas-phase H2O or translational modes associated with confinement in the cage, which were observed in H2O@C60  [Citation7]. These features of the H2O@C60+ spectrum indicate that excitations of the encapsulated H2O molecule may accompany the electronic transition of the C60+ cage. The calculations reported here indicate that similar energy separations to those observed in the experimental spectrum are expected for vibrational and translational modes of H2O in the ground electronic state of H2O@C60+, as presented in Table .

Table 1. Vibrational and translational modes obtained from calculations.

Note that although these experiments are carried out on weakly bound H2O@C60+−He complexes, the exohedral helium is expected to be a minor perturbation to the electronic transitions of H2O@C60+, as was the case for C60+ (see supplementary information). The following discussion and assignment of the rich spectrum is made on this basis and the assumption that the exohedral, weakly bound, helium atom does not affect the symmetry of Ψe.

4.2. Ortho/para ratio

More information on the cause of the band pattern can be gained by observations of the absolute number of complexes depleted following irradiation on resonance because the broad width of the 10,429 cm1 band in the electronic spectrum shown in Figure indicates saturation. Even with a high laser fluence, a significant number of complexes remained in the trap after irradiating them at 10,438 cm1, as indicated in Figure . It was observed that the number of remaining ions varied depending on the particular transition excited.

Figure 3. Number of H2O@C60+−He complexes (Ni) as a function of laser fluence upon irradiation at 10,438 cm1. Without exposure to laser radiation, 1100±60 ions with m/z = 742 were in the trap. 800±30 complexes remain after irradiation, corresponding to 73±5% of the population.

Figure 3. Number of H2O@C60+−He complexes (Ni) as a function of laser fluence upon irradiation at 10,438 cm−1. Without exposure to laser radiation, 1100±60 ions with m/z = 742 were in the trap. 800±30 complexes remain after irradiation, corresponding to 73±5% of the population.

The conversion between nuclear spin states is forbidden and takes several hours in the condensed phase  [Citation8, Citation16]. In our gas phase experiment, the high spin temperature ortho/para ratio is expected to be conserved. This is consistent with the data shown in Figure , where 73±5% ions do not interact with the laser at 10,438 cm1. The 27±5% that absorb at this energy are therefore attributed to para H2O@C60+.

4.3. Separation of ortho and para nuclear spin isomers by 2 colour experiments

As it was possible to remove one group of nuclear spin isomers from the trap, a two-colour setup was used to record nuclear spin isomer pure electronic spectra. These are presented in Figure . The spectra were assigned to ortho and para based on the fraction of ions that were depleted. The ortho spectrum shows one intense transition at 10,429 cm1, two bands at 10,515 and 10,530  cm1, one band at 10,603 cm1, and one more at 10,666 cm1. Four bands of para H2O@C60+ were observed, all at energies below 10,500 cm1.

Figure 4. Electronic spectra of H2O@C60+ (top), p-H2O@C60+ (middle), and o-H2O@C60+ (bottom). The lower two traces were obtained in 2 colour experiments. The middle spectrum was recorded by fragmenting all complexes that absorb at 10,429 cm1. The bottom spectrum was recorded after irradiation at 10,438 cm1.

Figure 4. Electronic spectra of H2O@C60+ (top), p-H2O@C60+ (middle), and o-H2O@C60+ (bottom). The lower two traces were obtained in 2 colour experiments. The middle spectrum was recorded by fragmenting all complexes that absorb at 10,429 cm−1. The bottom spectrum was recorded after irradiation at 10,438 cm−1.

4.4. Assignment of the absorption features

To allow assignment of the bands in the spectrum, the energy differences between the band at 10,429 cm1 and the others are compared to the rotational energy levels of free water. The three transitions from the ortho and para rotational ground state are 110101, 111000 and 212101. The energy separation between these transitions resembles the pattern observed for H2O@C60+, as shown in Figure . Alternative explanations involving hot bands or Q-branch transitions were explored but do not agree with the observed energy separations.

Figure 5. PGopher simulations of the rotational pattern of an electronic transition of H2O. The details of the simulations are explained in the text. The middle trace uses the parameters based on a geometry optimisation (B3LYP/6-31++G**) while the upper one shows a fit involving a geometry change of H2O between ground and excited state.

Figure 5. PGopher simulations of the rotational pattern of an electronic transition of H2O. The details of the simulations are explained in the text. The middle trace uses the parameters based on a geometry optimisation (B3LYP/6-31++G**) while the upper one shows a fit involving a geometry change of H2O between ground and excited state.

To aid the assignment, simulations of H2O electronic transitions have been carried out using H2O rotational constants provided by calculations on H2O@C60+. These simulations, which ignore any effects of the C60+ cage, have been carried out using PGopher software [Citation34] in C2v symmetry. Based on the observed nuclear spin isomer ratios, the spin temperature was set to 300 K (high-temperature limit). B2 electronic symmetry in the ground and excited electronic state resulted in an intensity pattern in agreement with the experiments. A rotational temperature of 5 K and full width half maximum (FWHM) of 2.5 cm1 were used. The A000X000 energy was chosen in order to match the A111X000 transition to the 10,429 cm1 band of the experimental spectrum. Figure shows the resemblance of the three lowest energy bands in H2O@C60+ with the simulated spectrum using rotational levels of free H2O (middle trace). While these results are in qualitative agreement, they can be improved when a geometry change between the ground and excited state is considered.

A simulated electronic spectrum involving a geometry change of H2O between the ground and excited state is shown in the upper trace of Figure . The rotational constant CA of the excited state can be derived directly from the experimental spectrum  [Citation35]. Furthermore, the fitting capability of the PGopher software was used to extract the rotational constant BA for the excited state, and the A000X000 energy. This fit used the calculated constants of H2O in H2O@C60+ in the ground states as inputs, together with AX=AA. The constants used to obtain Figure are listed in Table .

Table 2. Rotational constants used for the simulation in Figure .

The proposed assignment indicated in Figure , with the 10,429 cm1 absorption originating from the 000 level, coupled with the observation that the associated nuclear spin isomer is responsible for 75% of the total population, implies that even Ka+Kc are ortho in H2O@C60+ (or H2O@C60+−He). This differs from H2O@C60, where even Ka+Kc rotational functions are associated with the para nuclear spin isomer, suggesting that the H2O@C60+ (or H2O@C60+−He) electronic wave function is antisymmetric. This interpretation is reached because the Pauli principle requires the overall wavefunction to be antisymmetric with respect to exchange of the two H-nuclei. As the transition does not originate from a vibrationally excited level, Ψe is the contributor to Ψtot=ΨeΨvΨrΨns that differs between H2O and H2O@C60+ (H2O@C60+−He) for the same rotational and nuclear spin states.

There are several possibilities for assignment of the bands to a higher energy than the three shown in Figure . For example, the constants derived from the simulation lead to a near coincidence in energy for the 10,484 cm1 (b) absorption with an A221X110 transition (10,483 cm1) or the 10,469 cm1 (a) absorption, when the A rotational constant changes by 20% between ground and excited state, which is observed for the other rotational constants. This hot band transition, however, is ruled out due to the absence of absorption to lower energy than 10,396 cm1 which would be expected to be observed for the A101X110 transition. Therefore it appears that H2O inside the C60+ cage is rotationally cold in our experiment.

Band b is also close to the predicted A303X101 (10,482 cm1) and A221X101 transition (10,500 cm1). The latter transition depends on AA and could also appear at 10,484 cm1 if the value for AA varies in a reasonable %-range after electronic excitation (i.e. comparable to the other rotational constants). Here the value for the excited state rotational constant would be 22 cm1 with a 10,400 cm1 A000X000 energy, which is similar to the constants for the excited vibrational level in neutral H2O@C60 (23.15 cm1)  [Citation7]. A change of AA would require a slight adjustment of the origin energy to maintain the agreement in energy with the previously assigned bands, but would otherwise not influence the shape of the spectrum. However, the A221X101 transition is not a dipole-allowed transition of asymmetric top rotors and the A303X101 has ΔJ=2.

The A220X101 and A211X101 transitions are also in the same energy region and are expected to have an energy of 10,489 and 10,460 cm1 if AA = 22 cm1. These are a and c-type transitions, respectively, while all the other assigned features are b-type. A component of the transition dipole along these axes would be required for such nuclear spin forbidden features.

The only transition below 10,500 cm1 in the spectrum of ortho-H2O@C60+ is assigned to A111X000. The higher energy bands c and d at 10,515 and 10,531  cm1 lie 119 and 135 cm1 from the predicted A000X000 energy. Previous DFT calculations [Citation19] at the PBEO/6-311++G** level of theory indicated that frustrated vibrations and translations of water inside neutral C60 could be observed. In DFT (B3LYP/6-311++G**) calculations of the charged H2O@C60+ system, three translational modes between 125 and 128 cm1 and three frustrated vibrational modes (34, 101 and 111 cm1) were found in the electronic ground state. Therefore the bands c and d are consistent with the excitation of either a frustrated vibrational mode or with translations that are quantised due to the spacial confinement. Translations have also been observed experimentally at 110 cm1 in H2O@C60  [Citation7], which is similar to the energy inferred here. Based on these computational results, it seems plausible that the bands c and d could be due to either quantised translations or frustrated vibrations.

The 10,603 and 10667 cm1 absorptions, e and f, have an energy separation of 206 and 270 cm1 from A000X000, which is close to the ∼230cm1 energy of the C60+ cage vibration. The ∼270 cm1 energy separation of absorption f also corresponds to twice the energy separation of band d with respect to the A000X000 energy. These two bands could therefore be caused by the excitation of two quanta of the same vibration or translation. A summary of the potential assignments is given in Table .

Table 3. Absorption energies and proposed assignment of bands in the H2O@C60+ and D2O@C60+ spectra.

4.5. Electronic spectrum of D2O@C60+: comparison with H2O@C60+

Further experiments on D2O@C60+ were performed to support the assignments of the bands in the H2O@C60+ electronic spectrum. Figure  shows the spectrum of D2O@C60+ recorded under similar conditions to those used for H2O@C60+. The intensity pattern of the bands is significantly different from the observations in H2O@C60+, and the energy separations between the bands are smaller as expected due to the isotope effect. After irradiation at 10,418 cm1, 37±3% of the complexes remain in the trap while at 10,425 cm1, 63±3% of the complexes remain. These values correspond well to the 2:1 o/p ratio of D2O observed in the high-temperature limit, as expected from the H2O@C60+ experiment.

Figure 6. A comparison of the electronic spectrum of H2O@C60+ (top) and D2O@C60+ (bottom) recorded under similar laboratory conditions. The labels are discussed in the text.

Figure 6. A comparison of the electronic spectrum of H2O@C60+ (top) and D2O@C60+ (bottom) recorded under similar laboratory conditions. The labels are discussed in the text.

Substitution of H with D in H2O@C60+ affects vibrations, rotations and translations of the encapsulated molecule differently. In D2O an energy of 12 of the same H2O vibrational mode would be expected. The rotational constants of D2O are lower than those of H2O by about a factor of 2, if both molecules have the same geometry. In translations, motion of the whole molecule in the constrained environment is involved, and the effect of the isotope substitution is less severe. Translations of D2O would have mH2OmD2O times the energy of those of H2O  [Citation36].

4.6. Confirmation of assignment(s) based on isotope effect

To confirm the assignment of the lowest energy rotational transitions in H2O@C60+, the same rotational transitions were simulated for D2O@C60+. The results are shown in Figure  and the constants are listed in Table . Values of the rotational constants in the ground state were 12 of those used for the simulation of H2O, and the A000X000 energy was changed accordingly. The excited state rotational constants were obtained from PGopher fits. The good agreement between the simulation and the experimental results appears to support the previous assignment of the rotational transitions in H2O@C60+. Two-colour spectra of D2O@C60+ are shown in Figure .

Figure 7. PGopher simulations of the rotational pattern of an electronic transition of D2O in comparison with the experimental results on D2O@C60+ (bottom). The rotational constants for the simulation are derived as described in the text.

Figure 7. PGopher simulations of the rotational pattern of an electronic transition of D2O in comparison with the experimental results on D2O@C60+ (bottom). The rotational constants for the simulation are derived as described in the text.

Figure 8. Electronic spectra of D2O@C60+ (top), para-D2O@C60+ (middle), and ortho-D2O@C60+(bottom). The middle trace was recorded by fragmenting all complexes that absorb at 10,418 cm1 and the bottom trace was recorded by irradiating at 10,426 cm1.

Figure 8. Electronic spectra of D2O@C60+ (top), para-D2O@C60+ (middle), and ortho-D2O@C60+(bottom). The middle trace was recorded by fragmenting all complexes that absorb at 10,418 cm−1 and the bottom trace was recorded by irradiating at 10,426 cm−1.

The simulation has some further implications for the interpretation of the spectra and the structure of the two molecules. The experimentally derived CA of the excited state of D2O@C60+ has a value of 2.7 cm1. This is smaller than the 3.2 cm1 which would be expected from 12CA. This indicates that the geometries of H2O@C60+ and D2O@C60+ are slightly different or that centrifugal distortion also contributes.

Another consequence of the simulation is that the A000X000 energy varies between the different isotopologues. Based on the assignments it is possible to estimate the energy for A000X000. This energy (10,409 cm1) is around 13 cm1 higher than the estimated A000X000 of H2O@C60+ at 10,396 cm1. It is assumed that differences in zero-point energies cause the blue shift of the A000X000 energy of D2O@C60+ with respect to H2O@C60+.

Based on the energy separations between the calculated A000X000 energy and the other bands in H2O@C60+, the corresponding rotation, translation and vibrational modes expected to be seen in measurements on D2O@C60+ could be estimated. The prediction can be compared to the experimental spectrum, to identify bands associated with translational, vibrational or rotational excitation.

In H2O@C60+, the bands marked as a and b lie 74 and 89 cm1 from A000X000. When dividing these energy separations by two and adding these values to the A000X000 transition of D2O@C60+, the corresponding rotational levels in D2O@C60+ are predicted at 10,444 and 10,452  cm1. The two observed bands at 10,447 and 10,453  cm1 in D2O@C60+ indicate that the bands at a and b are caused by rotational transitions of the encapsulated H2O molecule and, based on the isomer separation, that these transitions are excited from the X101 rotational level. As described for H2O@C60+, several forbidden rotational transitions are in a similar energy range, as presented in Table , but it can not be concluded which of these transitions is observed. However, it can be ruled out that these bands are due to translations or vibrations based on the predictions for D2O@C60+.

Bands c and d are observed in the spectrum of the ortho isomer of H2O@C60+. Rotational transitions are excluded from the possible explanations because there are no corresponding bands in D2O@C60+. However, predicted translational (10,521 and 10,535 cm1) and vibrational transitions (10,493 and 10,503 cm1), present two options for assignment. Bands IIIa and IIIb in D2O@C60+ are in a similar energy range but are not observed, as expected, in the para spectrum of D2O@C60+. On the other hand band IV at 10,506 cm1 is a close match to the predicted vibration at 10,504 cm1 of para-D2O@C60+ (band d in H2O@C60+).

Considering the isotope effect for band e, the equivalent vibration in D2O@C60+ would show at 10,555 cm1 and the rotation at 10,513 cm1. The predicted vibration is 15 cm1 lower than the observed band V and the rotation is also close to band IV in D2O@C60+, so it cannot be unambiguously assigned.

In the H2O@C60+ electronic spectrum the band f at 10,666 cm1 is close to the energy expected for the C60+-cage vibration. The measurement for D2O@C60+ only covered an energy range up to 10,650 cm1. If this band is caused by a vibrational mode of the C60+ cage the effect of the isotopes would be minimal and then it would show at similar energies for both isotopologues when including the blue shift of the A000X000 energy for D2O@C60+.

A summary of the above assignment discussion is presented in Table . The three lowest energy bands along with a and b can be classified as rotational transitions. The three bands in D2O@C60+ IIIa, IV and IIIb are only separated by a few wavenumbers and it is therefore difficult to assign them with certainty. Especially because band IIIa and IIIb of D2O@C60+ do not correspond to bands with the same rovibronic symmetry in H2O@C60+, they could also be caused by transitions not observed in H2O@C60+.

5. Conclusion

Electronic spectra of H2O@C60+ and D2O@C60+ were recorded in the gas phase at temperatures below 10K in the vicinity of the near-IR C60+ electronic transitions by photodissociation of weakly bound He complexes. Two-colour experiments allowed the removal of one nuclear spin isomer from the trap, allowing spin isomer pure spectra of these species to be obtained. The ortho/para ratio determined spectroscopically agrees with the high-temperature limit. The rich spectra obtained include contributions assigned to rotational excitation of the encapsulated H2O (D2O) molecules that appear to be relaxed to their ground states by buffer gas cooling in the trap. In contrast to C60+, only one of the two expected electronic transitions in this region was observed for the endohedral cations.

While not all of the observed spectral features have been assigned, the comparison of D2O@C60+ and H2O@C60+ spectra make it possible to exclude certain explanations and narrow down the possible assignments. For example, the isotope effect shows that there are more rotational transitions observed than the three allowed transitions from the rotational ground state of H2O (D2O). The results also imply that some of these bands could be caused by frustrated vibrational modes of the encapsulated molecule, which are yet to be observed in experiments on H2O@C60 or H2O@C60+.

Supplemental material

Supplemental Material

Download MS Word (2.5 MB)

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

The authors acknowledge financial assistance from the Royal Society [grant numbers RGF/EA/181035, URF/R1/180162] and Engineering and Physical Sciences Research Council [grant numbers EP/M001962/1, EP/P009980/1 and EP/T004320/1].

References

  • K. Komatsu, M. Murata and Y. Murata, Science 307, 238 (2005). doi:10.1126/science.1106185
  • K. Kurotobi and Y. Murata, Science 333, 613 (2011). doi:10.1126/science.1206376
  • A. Krachmalnicoff, M.H. Levitt and R.J. Whitby, Chem. Commun. 50, 13037 (2014). doi:10.1039/C4CC06198E
  • 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, Nat. Chem. 8, 953 (2016). doi:10.1038/nchem.2563
  • 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, Angew. Chem. Int. Ed. 58, 5038 (2019). doi:10.1002/anie.201900983
  • A. Kilaj, H. Gao, D. Rösch, U. Rivero, J. Küpper and S. Willitsch, Nat. Commun. 9, 2096 (2018). doi:10.1038/s41467-018-04483-3
  • A. Shugai, U. Nagel, Y. Murata, Y. Li, S. Mamone, A. Krachmalnicoff, S. Alom, R.J. Whitby, M.H. Levitt and T. Rõõm, J. Chem. Phys. 154, 124311 (2021). doi:10.1063/5.0047350
  • C. Beduz, M. Carravetta, J.Y.C. Chen, M. Concistrè, 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, Proc. Natl. Acad. Sci. U.S.A.109, 12894 (2012). doi:10.1073/pnas.1210790109
  • 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, Phys. Chem. Chem. Phys. 16, 21330 (2014). doi:10.1039/C4CP03272A
  • T.R. Huet, C.J. Pursell, W.C. Ho, B.M. Dinelli and T. Oka, J. Chem. Phys. 97, 5977 (1992). doi:10.1063/1.463735
  • 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, Sci. Rep. 10, 18329 (2020). doi:10.1038/s41598-020-74972-3
  • A. Melentev, S. Zhukov, V. Balos, G. Hoffman, S. Alom, M. Belyanchikov, E. Zhukova, M. Dressel, G.R. Bacanu, P. Abramov, M.H. Levitt, R. Whitby, B. Gorshunov and M. Sajadi, J. Phys. Conf. Ser. 1984, 012012 (2021). doi:10.1088/1742-6596/1984/1/012012
  • P.M. Felker, V. Vlček, I. Hietanen, S. Fitzgerald, D. Neuhauser and Z. Bačić, Phys. Chem. Chem. Phys. 19, 31274 (2017). doi:10.1039/C7CP06062A
  • O. Carrillo-Bohórquez, A. Valdés and R. Prosmiti, Chem. Phys. Chem. 23, e202200034 (2022). doi:10.1002/cphc.202200034
  • B. Meier, S. Mamone, M. Concistrè, J. Alonso-Valdesueiro, A. Krachmalnicoff, R.J. Whitby and M.H. Levitt, Nat. Commun. 6, 1 (2015). doi:10.1038/ncomms9112
  • 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, J. Chem. Phys. 140, 194306 (2014). doi:10.1063/1.4873343
  • M.H. Levitt, Philos. Trans. A Math. Phys. Eng. Sci. 371, 20130124 (2013). doi:10.1098/rsta.2012.0429
  • Z. Bačić, V. Vlček, D. Neuhauser and P.M. Felker, Faraday Discuss. 212, 547 (2018). doi:10.1039/C8FD00082D
  • A. Varadwaj and P.R. Varadwaj, Chem. Eur. J. 18, 15345 (2012). doi:10.1002/chem.201200969
  • B. Ensing, F. Costanzo and P.L. Silvestrelli, J. Phys. Chem. A 116, 12184 (2012). doi:10.1021/jp311161q
  • S.P. Jarvis, H. Sang, F. Junqueira, O. Gordon, J.E.A. Hodgkinson, A. Saywell, P. Rahe, S. Mamone, S. Taylor, A. Sweetman, J. Leaf, D.A. Duncan, T.L. Lee, P.K. Thakur, G. Hoffman, R.J. Whitby, M.H. Levitt, G. Held, L. Kantorovich, P. Moriarty and R.G. Jones, Commun. Chem. 4, 135 (2021). doi:10.1038/s42004-021-00569-0
  • S.J. Elliott, C. Bengs, K. Kouřil, B. Meier, S. Alom, R.J. Whitby and M.H. Levitt, ChemPhysChem 19, 251 (2018). doi:10.1002/cphc.201701330
  • G.R. Bacanu, G. Hoffman, M. Amponsah, M. Concistrè, R.J. Whitby and M.H. Levitt, Phys. Chem. Chem. Phys. 22, 11850 (2020). doi:10.1039/D0CP01282C
  • E.K. Campbell, M. Holz, D. Gerlich and J.P. Maier, Nature 523, 322 (2015). doi:10.1038/nature14566
  • E.K. Campbell, M. Holz, J.P. Maier, D. Gerlich, G.A.H. Walker and D. Bohlender, Astrophys. J. 822, 17 (2016). doi:10.3847/0004-637X/822/1/17
  • E.K. Campbell and P.W. Dunk, Rev. Sci. Instrum. 90, 103101 (2019). doi:10.1063/1.5116925
  • A.O. Lykhin, S. Ahmadvand and S.A. Varganov, J. Phys. Chem. Lett. 10, 115 (2018). doi:10.1021/acs.jpclett.8b03534
  • A.S. Menon and L. Radom, J. Phys. Chem. A 112, 13225 (2008). doi:10.1021/jp803064k
  • S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys. 132, 154104 (2010). doi:10.1063/1.3382344
  • L. Goerigk and S. Grimme, Modeling of Molecular Properties (John Wiley & Sons, Hoboken, 2011), pp. 1–16.
  • S. Grimme, S. Ehrlich and L. Goerigk, J. Comput. Chem. 32, 1456 (2011). doi:10.1002/jcc.21759
  • M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone, G.A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A.V. Marenich, J. Bloino, B.G. Janesko, R. Gomperts, B. Mennucci, H.P. Hratchian, J.V. Ortiz, A.F. Izmaylov, J.L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V.G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J.A. Montgomery, Jr. J.E. Peralta, F. Ogliaro, M.J. Bearpark, J.J. Heyd, E.N. Brothers, K.N. Kudin, V.N. Staroverov, T.A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A.P. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, J.M. Millam, M. Klene, C. Adamo, R. Cammi, J.W. Ochterski, R.L. Martin, K. Morokuma, O. Farkas, J.B. Foresman and D.J. Fox, Gaussian16 Revision A.03 (Gaussian Inc., Wallingford, CT, 2016).
  • F. Duschinsky, Acta Physicochim. URSS 7, 551 (1937).
  • C.M. Western and J. Quant, Spectrosc. Radiat. Transf. 186, 221 (2017). doi:10.1016/j.jqsrt.2016.04.010
  • P. Bunker and P. Jensen, Fundamentals of Molecular Symmetry (1st ed., CRC Press, Boca Raton, 2005), p. 101.
  • T. Rõõm, L. Peedu, M. Ge, D. Hüvonen, U. Nagel, S. Ye, M. Xu, Z. Bačić, S. Mamone, M.H. Levitt, M. Carravetta, J.Y. Chen, X. Lei, N.J. Turro, Y. Murata and K. Komatsu, Phil. Trans. A Math. Phys. Eng. Sci. 371, 20110631 (2013). doi:10.1098/rsta.2011.0631