Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

C20 and nitrogen-substituted fullerenes: anharmonic IR and UV-vis spectra for astrophysical environments

Venkata Lakshmi Karria, Ajay Chaudharib, Takashi Onakac and Mahadevappa Naganathappa*a
aDepartment of Physics, School of Science, GITAM (Deemed to be University), Hyderabad 502329, TS, India. E-mail: swamimahadev25@gmail.com
bDepartment of Physics, The Institute of Science, Dr Homi Bhabha State University, Madame Kama Road, FORT, Mumbai 400032, MH, India
cDepartment of Astronomy, Graduate School of Science, The University of Tokyo, 7-3-1 Hongo, Tokyo 113-0033, Japan

Received 21st July 2025 , Accepted 14th October 2025

First published on 28th October 2025


Abstract

Theoretical infrared (IR) and electronic absorption spectra of the C20 fullerene and its nitrogen-substituted heterofullerenes in gas and water solvent are studied and discussed in terms of astronomical observations. The replacement of a carbon atom by nitrogen results in two stable heterofullerenes, which is confirmed by their HOMO to LUMO energy gap. The ionization potential and electron affinity of these molecules are reported. Theoretical calculations performed at the B3LYP/6-311++G(d,p) level of density functional theory (DFT). The effect of water solvent is studied using the integral equation formalism polarized continuum model (IEFPCM) at the same level of theory. Effects of substitution on the electronic and absorption spectra of these molecules are studied. The results of the C20 fullerene and its heterofullerenes show spectra with peaks at 6.2, 6.67, 7.0, 7.7, 8.5, 11.3, and 12.8 μm, which have corresponding features in observed spectra of the planetary nebulae Tc1 and NGC 7027, and the reflection nebulae NGC 2023 and NGC 7023. The electronic absorption spectra of these molecules are also calculated by time-dependent DFT (TD-DFT) and discussed in relation to the ultraviolet bump feature at 217 nm in the interstellar extinction curve. We estimate the transition wavelength, oscillator strength, and symmetry using the AOMix program.


1 Introduction

Infrared (IR) spectra of various astronomical objects, from circumstellar environments and the interstellar medium of our galaxy to external galaxies, exhibit strong emission features at 3.3, 6.2, 7.7, 8.6, and 11.2 μm.1,2 These are attributed to large carbonaceous molecules composed of benzene rings, such as polycyclic aromatic hydrocarbons (PAHs).3–5 Another class of carbonaceous materials, fullerenes, is also responsible for several features in IR spectra.6–14 PAHs have planar structures of fused benzene rings, whereas fullerenes form closed carbon cages.

The discovery of C60 and C60+ in laboratory experiments15 and the development of synthesis methods for C60 and C70 (ref. 16) paved the way for their astrophysical identification. Cami et al.6 reported the first IR detection of C60 and weak features of C70 in the planetary nebula Tc1. Since then, features of fullerenes have been identified in various environments such as planetary nebulae,17,18 reflection nebulae,19,20 proto-planetary nebulae,21 and young stellar objects.22 The bands observed in NGC 7023 at 6.4, 7.1, 8.2, and 10.5 μm are attributed to C60+.23 IR features near 6.6, 9.8, and 20 μm have been reported in both galactic and extragalactic planetary nebulae18,24 and Sgr B2,25 resembling spectra of planar C24.26 C24 has also been suggested as a carrier of the 11.2 μm band in NGC 7027.27 The detection of both large (C60 and C70) and small (C24) fullerenes has broadened the understanding of carbonaceous chemistry in evolved stars.15

Several theoretical studies have focused on small or modified fullerenes. Adjizian et al.28 proposed that unassigned IR bands could arise from small fullerenes and modelled the IR spectra of C20 to C60 in various charge states using Density Functional Theory (DFT) calculations. Gómez-Muñoz et al.29 suggested that hydrogenated amorphous carbon grains formed in planetary nebulae could carry the 12 μm plateau. Fulleranes30 and nitrogen-doped fullerenes31 have also been studied in the context of IR and ultraviolet (UV) spectral features and chemical stability. Foing and Ehrenfreund first proposed that the Diffuse Interstellar Bands (DIBs) at 958 and 963 nm could be due to C60+.32 Laboratory confirmation33 and observational studies34–36 supported this proposal. Other DIBs have also been attributed to C60+ (ref. 37) and C70+.38 Iglesias-Groth et al.39 reported IR emission bands, which can be attributed to neutral, cationic, and anionic fullerenes in the IC 348 star-forming region. The cationic C60 was found to emit strongly at 11.21, 16.40, and 20–21 μm, in addition to the well-known 17.4 and 18.9 μm bands.40 Further studies investigated fullerene cage stability and astrochemical reactivity.41,42 The UV bump at 217.5 nm seen in the interstellar extinction curve—attributed to π → π* transitions in sp2 carbon systems—has been associated with PAHs and other carbonaceous materials.43–46 While PAHs are major contributors47,48 fullerene species may also play a role.26,49,50

Theoretical investigations, even before the first detection in space, explored the geometry and electronic structures of fullerenes across a wide size range from C20 to C720.51,52 C20 is the smallest fullerene, with a strained dodecahedral cage composed of 12 pentagons, lacking pentagon isolation and thus being less stable than C60.53,54 Alternative isomers, such as rings and chains, have also been studied.55–57 C20 was synthesized from dodecahedrane via debromination58,59 with ion beam irradiation.60 Its IR and UV spectra have been modelled.61–63

Substitution of carbon atoms with heteroatoms like nitrogen, boron, or oxygen yields heterofullerenes.64,65 The first nitrogen-substituted fullerene was identified using mass spectrometry.66,67 Li–fullerene interactions have also been studied,68 supporting stable ion-cage complexes. Theoretical investigations27,69 indicate that nitrogen substitution in small fullerenes such as C20 enhances their stability and modifies their electronic structure, making them more relevant for astrophysical environments than the pristine C20 cage. These results emphasize the importance of the study of small fullerenes with nitrogen substitution. Those small cyclic hydrocarbon species may be formed via reactions on the surface of dust grains with an ice mantle.70 Therefore, it is also of interest to study the effect of ice mantle on the spectroscopic properties of C20.

Fullerenes were first predicted theoretically and later confirmed experimentally. However, laboratory identification of fullerene species remains difficult due to challenges in studying them in an isolated condition. Thus, theoretical approaches remain essential for understanding and predicting their properties. Small fullerenes are generally less stable than larger ones, such as C60 and C70, and their detection is further complicated when they are part of a complex mixture of fullerenes, consisting of various sizes, charge states, and possible substitutions. These result in overlapping or weak spectroscopic features that hinder clear identification. The discovery of C60, C60+, and C70 fullerenes in interstellar and circumstellar environments suggests possibilities for the presence of other fullerenes and their derivatives. Identifying fullerenes in various astronomical environments relies on spectroscopic data analysis, which requires both laboratory and theoretical studies.

The present study investigates the vibrational and electronic spectroscopic properties of the C20 fullerene and its nitrogen-substituted derivatives (N10C10 and C12N8) in neutral, cationic, and anionic states using DFT. To simulate astrophysical environments, we also consider the spectra of C20 in a water solvent environment as an approximation of ice mantle conditions. While earlier studies employed harmonic DFT methods to model small fullerene IR spectra,24 this work presents-for the first time-anharmonic DFT calculations for both pristine and nitrogen-substituted C20 fullerenes across various charge states. In addition to IR spectra, we compute near-UV-visible absorption spectra to assess their potential contribution to the prominent 217.5 nm UV extinction bump. By combining charge state, nitrogen substitution symmetry, and solvent effects, this study provides a comprehensive understanding of the spectroscopic behavior and astrophysical relevance of small fullerenes. Recent radio observations have detected faint emission lines from nitrogen-containing small PAHs and cyclic hydrocarbons, suggesting that nitrogen is commonly found in interstellar carbon-based molecules.70–79 This supports the idea that nitrogen-substituted fullerenes may also exist in space. Including water as a solvent in our calculations helps us understand the effects of interstellar ices or polar environments on their infrared spectra.

2 Computational details

All the theoretical calculations for C20 and its nitrogen-substituted heterofullerenes (N10C10 and C12N8) are performed using the Gaussian 16 software package.80 An et al. and Gianturco et al.81,82 theoretically studied the molecular properties of C20 isomers in their neutral and anionic forms using various methods and basis sets. They confirmed that B3LYP/cc-pVTZ predicts the most accurate results for both the neutral and anionic low-lying isomers of C20. Saito et al.83 conducted their study at the B3LYP/6-311+G(d) level of theory, while Soleimani Amiri et al.69 used the B3LYP/aug-cc-pVTZ level of theory. In the present study, we optimized C20 molecules using the HF, B3LYP, MP2, and CCSD methods with 6-311++G(d,p), TZVP, aug-cc-pvtz, aug-cc-pvdz, and aug-cc-pvqz. Of these, the hybrid functional method B3LYP with the TZVP, aug-cc-pVTZ, and 6-311++G(d,p) basis sets is predicted to have the lowest energy for C20, respectively. Among these three basis sets, the variation in the obtained lowest energy is minimal, with the differences only in the decimal range, on the order of 0.01–0.02 eV. Taking into account a balance between accuracy and computational cost for both IR and UV spectra, we employ the B3LYP/6-311++G(d,p) level of theory in the present study. This level of theory is used to calculate the vibrational frequencies and electronic absorption spectra of the studied structures in their neutral, cationic, anionic, and water-solvated states. The B3LYP/6-311++G(d,p) approach has been extensively applied to PAH vibrational studies and has been shown to reliably reproduce experimental spectra.84,85 Although more recent hybrid functionals such as M06-2X and ωB97XD may provide improved accuracy for certain systems,86,87 benchmarking studies indicate that B3LYP remains a well-validated and widely accepted choice for PAHs and related carbon clusters, offering a good compromise between accuracy and efficiency. We use the integral equation formalism polarized continuum model (IEFPCM)88–90 with the dielectric constant of 78.4 (ref. 91) at the same level of theory to calculate IR and UV absorption spectra of water solvent states. The Cartesian coordinates of all optimized structures in their neutral, cationic, and anionic forms under both gas and solvent phases are provided in the SI (Table S2, SI).

In the present study, anharmonic vibrational spectra of the C20 fullerene and its nitrogen-substituted derivatives (N10C10 and C12N8) are computed using the Gaussian 16 software with the deperturbed second-order vibrational perturbation theory (DVPT2) method.92 This approach accounts for higher-order force constants, quadratic, cubic, and quartic terms, in the potential energy surface, enabling a more accurate representation of molecular vibrations. While a scaling factor of 0.9613 (ref. 93) is applied to harmonic frequencies to compare with experimental spectra, no such correction is needed for anharmonic calculations. We select N10C10 and C12N8 as representative nitrogen-substituted C20 since they have non-zero dipole moment and thus are expected to show appreciable changes in their spectra compared to those of C20 due to the large fraction of nitrogen. We perform anharmonic IR calculations for all species of the present study in their neutral, cationic, and anionic forms, both in the gas phase and in water solvent state. The calculated theoretical IR spectra are convolved with Lorentzian profiles of a full width at half maximum (FWHM) of 8 cm−1.94 We also estimate the electron affinity (E.A.) and ionization potential (I.P.) of these fullerenes, since these attributes are important parameters to control their chemical and physical properties.95 They are calculated by the following eqn (1) and (2).96

 
Ionization potential (I.P) = EcationEneutral (1)
 
Electron affinity (E.A) = EneutralEanion (2)
Here, Eneutral, Ecation, and Eanion represent the energies of optimized structures in their neutral, positively charged, and negatively charged forms, respectively. The electronic absorption spectra of these structures are reported using time-dependent density functional theory (TDDFT).97 The AOMix programme98 is used to identify the electronic transitions, oscillator strengths, and symmetry. The HOMO–LUMO energy gap is also obtained at the same level of theory.

3 Results and discussion

3.1 Molecular properties

The molecular properties such as the energy, ionization potential, electron affinity and symmetry of C20 and its heterofullerenes in their neutral and ionic forms in gas and water solvent states are reported in Table 1, and the corresponding geometries are shown in Fig. 1. Unlike the highly stable C60, the C20 fullerene lacks hexagons and violates the isolated pentagon rule, making it less stable. Our calculations show that N-substituted forms (N10C10 and C12N8) exhibit enhanced stability, being consistent with previous reports.31,69 Nitrogen substitution improves electronic stability by acting as both electron donors and acceptors, which modifies the molecular orbital distribution. Among the heterofullerenes studied here, the C12N8 anion and the N10C10 cation exhibit notable dipole moments of 1.56 D and 0.35 D, respectively. In contrast, the neutral forms of both molecules show near-zero dipole moments, despite their low-symmetry geometries (C1 and C2v point groups) in both gas and solvent phases. C20 fullerene exhibits the lowest symmetry Ci with zero dipole moment in neutral and ionic states. The fully optimized structure of the neutral C12N8 has Th symmetry, while its anionic state has the lowest symmetry C2v.
Table 1 Energy relative to N10C10, ionization potential (I.P), electron affinity (E.A), and symmetry of the C20 fullerene and its N-heterofullerenes and their ions
Molecules Neutral Solvent Cation Anion I.P (eV) E. A (eV)
Energy (eV) Sym Energy (eV) Sym Energy (eV) Sym Energy (eV) Sym
C20 4.526 Ci 4.527 Ci 4.525 Ci 4.527 Ci 6.95 2.32
N10C10 0.00 C1 0.00 C1 0.00 C1 0.00 C1 8.52 3.17
C12N8 9.02 Th 9.02 Th 9.01 Th 9.03 C2v 7.65 2.18



image file: d5ra05271h-f1.tif
Fig. 1 Structures of the C20 fullerene and its nitrogen-substituted heterofullerenes optimized at B3LYP/6-311++G(d,p) level of theory.

The average bond lengths of C20, C12N8, and N10C10 in their neutral, cationic, and anionic states (gas and solvent phases, harmonic and anharmonic levels) are summarized in Table S1. The pristine C20 cage shows uniform C–C bonds (∼1.45 Å), while nitrogen substitution introduces shorter C–N bonds (1.33–1.36 Å) and slightly perturbs C–C bonds (1.37–1.39 Å). In N10C10, N–N bonds appear (∼1.48 Å), further lowering the symmetry. Charge states influence bond lengths: in cations, C–C bonds contract by ∼0.01–0.02 Å, while in anions, bond alternation increases slightly. Solvent effects are minimal (<0.01 Å change). These results confirm that doping and charge localization significantly distort the C20 framework.

The results show that IR intensities are typically weak in the cationic states and significantly strong in the anionic forms, particularly for C–C stretching modes. Nitrogen substitution reduces molecular symmetry, leading to increased IR activity and changes the intensity patterns. The ionization potential (IP) of the nitrogen-substituted heterofullerenes N10C10 and C12N8 is higher than that of the C20 fullerene, amounting to 8.52 and 7.65 eV, respectively. It indicates that these molecules face strong resistance to losing electrons. In contrast, the neutral C20 fullerene exhibits a lower IP of 6.95 eV, indicating high nucleophilicity due to the concentration of positive charge at the centre of C20. The present calculation shows that C20 has an electron affinity of 2.32 eV, which agrees well with the experimental result of 2.25 eV.50 Miar et al.99 reported that nitrogen doping increases the electron affinity of the C20 fullerenes. In line with this, N10C10 exhibits a high electron affinity, indicating a strong tendency to accept electrons. Although N10C10 is overall neutral, its electron density is not evenly distributed. More electron density is concentrated near the center, which may affect how the molecule reacts. As a result, the carbon atoms show similar reactivity toward electrophiles.

The neutral C20, with its relatively low ionization potential (6.95 eV), can be described as nucleophilic, whereas nitrogen-substituted heterofullerenes (IP = 7.65–8.52 eV) are less nucleophilic and more electronically stable. The cationic species are electrophilic due to positive charge localization, while the anionic forms show strong nucleophilic character consistent with their higher electron density and electron affinities (Table 1).

3.2 Infrared spectra

Harmonic and anharmonic mid-infrared absorption spectra for the neutral forms in both gas and water solvent phases, as well as harmonic IR spectra for the ionic states, are computed at the B3LYP/6-311++G(d,p) level of theory. These static DFT calculations are performed at 0 K, and the resulting spectra are presented in Fig. 2 over the 2–20 μm region. Anharmonic vibrational frequencies along with their relative intensities and combination modes for C20, N10C10, and C12N8 in the neutral and ionic forms are given in Table 2. Table 2 lists selected vibrational modes to highlight spectroscopically relevant features. Modes with near-zero IR intensity or degenerate components are omitted; the full set of 54 modes is provided in the SI. The vibrational analysis provides valuable insight into the structural characteristics of C20 and its nitrogen-substituted derivatives. Each molecule exhibits 54 vibrational modes, with several intense features highlighted in the main text. The complete set of harmonic vibrational frequencies and intensities in both gas and solvent phases is provided in the SI (Tables S3–S8, SI). These results are consistent with earlier studies.28
image file: d5ra05271h-f2.tif
Fig. 2 Infrared vibrational spectra of the C20 fullerene and its nitrogen-substituted heterofullerenes in their neutral gas and water solvent state, and ionic states. The spectra are calculated at the B3LYP/6-311++G(d,p) level of theory.
Table 2 The Anharmonic infrared vibrational frequencies (in cm−1) and relative intensity I (in kM mol−1) along with fundamental and combination modes for C20, N10C10, and C12N8 in their neutral form in the gas and water solvent states and their ionic forms
Neutral in gas phase Neutral in water solvent Cation in the gas phase Anion in gas phase
Frequency (cm−1) Intensity (kM mol−1) Modea Frequency (cm−1) Intensity (kM mol−1) Modea Frequency (cm−1) Intensity (kM mol−1) Modea Frequency (cm−1) Intensity (kM mol−1) Modea
d denotes the doubly degenerate states.a Mode labels (e.g., V49) follow the normal mode numbering from Gaussian output. Only a representative subset of vibrational modes is shown. Modes with negligible IR intensity or degenerate components are omitted for clarity. The complete set of 54 modes for each species is provided in the SI (Tables S3–S8).
C20
1310.49 49 V49 13[thin space (1/6-em)]919.3 68 V51 1301.19 0.0001 V50 1360.98 0.0001 V54
    V1 + V49     V1 + V51     V19 + V50    
1297.9 27 V50 1270.56 1811 V44 1265.97 0.5 V48 1299.1 47 V52
    V2 + V50            
1289.65 29 V51 1237.39 173 V49 1265.11 0.6 V47 1254.22 30 V50
    V1 + V51     V1 + V49     V1 + V47     V2 + V50
1244.63 18 V44 1231.01 288 V38 1220.44 0.4 V44 1243.23 262 V49
    V1 + V44     V1 + V38     V22 + V44     V1 + V49
1240.26 29 V43 1208.54 434 V43 1203.04 9.1 V42 1191.12 62 V46
    V1 + V43     V1 + V43     V42 + V43     V2 + V46
1195.08 16 V42 1189 302 V50 1198.39 11 V41 1164.4 0.4 V39 + V48
    V2 + V42     V2 + V50         V1 + V39
        V2 + V36        
1192.26 2 V39 1166.12 51 V35 1159.02 0.7 V38 1159.26 26 V41
    V2 + V39     V1 + V35     V22 + V38    
1171.83 6 V38 1151.66 60 V42 1156.34 0.6 V39 1129.28 15 V38
    V1 + V38     V2 + V42     V22 + V39     V2 + V28
1159.21 3 V36 1117.83 6 V38 1104.53 0.001 V35
    V1 + V36     V1 + V38     V8 + V35    
1078.04 1.6 V31 1044.6 167 V31 1037.47 1.5 V31 1117.49 1.2 V35
    V1 + V31     V1 + V31         V5 + V35
905.91 15 V28 924.19 23 V28        
901.5 22 V27 858.29 218 V27 894.68 4 V28 893.87 33 V28
        V1 + V27     V28 + V52     V3 + V28
884.68 14 V26 898.83 87 V26 894 4 V27 877.68 27 V27
        V1 + V26     V2 + V27     V4 + V27
731.96 14 V21 740.84 150 V19 + V46 865.67 8 V26 866.3 44 V26
    V2 + V21     V1 + V19     V1 + V26     V5 + V26
726.54 12 V20 727.06 211 V20 679.45 0.3 V19 699.69 58 V21
        V1 + V20     V17 + V19    
636.41 0.4 V2 + V18 657.07 82 V18 667.04 0.9 V18 685.34 0.5 V20
        V2 + V18     V18 + V40     V1 + V20
607.06 4 V14 646.55 27 V17 609.79 1.1 V14 622.73 18 V17
    V1 + V14     V1 + V17     V14 + V52     V17 + V37
582.15 7 V13 682.81 22 V13 608.58 0.1 V13 599.52 12 V14
    V2 + V13     V13 + V21     V13 + V46     V3 + V14
591.66 2 V12 584.32 153 V8 585.78 0.2 V12 572.49 0.5 V11
    V1 + V12     V2 + V8     V2 + V11     V2 + V11
            V12 + V43     V11 + V19
571.12 0.1 V10 573.45 41 V10 + V22 585.48 0.2 V11 572.48 0.4 V11 + V53
    V10 + V22     V1 + V10     V2 + V43     V11 + V19
569.2 5 V11 559.3 103 V9 567.16 0.007 V9 557.87 4 V12
    V1 + V11     V1 + V9     V9 + V17     V1 + V12
568.78 0.5 V9 551.65 14 V6 567.82 0.002 V10 542.35 6 V9
    V9 + V22     V1 + V6     V10 + V45     V9 + V18
557.56 0.2 V8 481.85 554 V11 564.41 0.002 V8 526.95 4 V8
    V1 + V8     V11 + V21     V8 + V35     V8 + V18
543.71 8 V6 464.21 32 V12 549.19 4 V7 510.38 9 V7
    V6 + V17     V1 + V12        
[thin space (1/6-em)]
N10C10
d1517.13 0.002 V1 + V54 d1524.76 0.001 V1 + V51 1775.93 14 V54 1533.44 50 V52
    V1 + V53     V1 + V53 1774.31 2 V53     V5 + V52
d1492.59 0.0002 V20 + V51 1495.79 0.4 V21 + V52 1737.57 47 V51 1474.29 199 V51
    V20 + V50                 V1 + V51
1382.44 0.0013 V48 d1391.73 0.0002 V13 + V48 1642.53 165 V48 1357.05 50 V47
1346.6 2 V47     V22 + V47 1346.28 30 V41 1294.07 134 V46
1345.65 3 V46 1348.53 5.3 V46 1247.18 21 V36 1270.62 322 V53
1250.97 0.0002 V20 + V45 1257.6 0.009 V21 + V45 1205.57 331 V33 1261.02 128 V50
            V33 + V44     V5 + V50
984.91 0.0004 V2 + V43 1000.88 0.0001 V43 949.19 20 V27 1149.8 127 V48
            V27 + V44     V1 + V48
d925.87 0.1 V40 d929.06 1.1 V40 865.89 10.3 V24 948.12 8 V42
    V39     V39     V24 + V48     V41
859.31 5 V36 895.07 1.6 V36 860.46 3 V7 + V23 914 9 V37
733.75 26 V26 d726.59 72.2 V26 794.62 3.4 V17 + V20 832.22 16 V34
722.4 0.002 V29 d720.47 0.009 V29 723.49 6 V14 823.87 71 V35
d624.68 0.006 V19 d632.53 0.05 V19 618.8 1.1 V4 + V8 720.08 768 V26
592.18 9.3 V16 595.74 23 V16 537.19 2.6 V5 702.84 216 V2
485.94 0.3 V14 + V35 530.6 0.08 V5 + V14 518.65 5 V3 + V34 694.76 99 V25
            V3 + V27    
347.34 0.1 V5 352.82 0.023 V5 + V14 688.03 245 V28
                V1 + V28
305.82 0.002 V2 + V43 308.6 0.6 V3 + V44 512.14 82 V14
            443.4 219 V7
            360.91 223 V5
[thin space (1/6-em)]
C12N8
d1620.2 14 V50 d1619.15 23 V50 1353.49 22 V2 + V53 1618.55 13 V53
    V53     V51     1468.84 33 V50
    V52     V52         V6 + V54
1146.44 0.001 V4 + V47 1145.47 0.001 V4 + V47 1184.44 0 V47 1192.24 43 V4 + V47
1071.03 35 V42 1062.59 48 V41 1099.92 20.4 V40 1041.77 11 V7 + V42
d1069.4 42 V41 1061.96 56 V42 d1099.88 21 V41 1035.28 14 V8 + V37
    V40 1061.23 71 V40     V42 1009.14 5 V40
d938.51 3 V34 d938.55 5 V34 955.77 9.3 V35 913.27 0.3 V35
    V35     V35     V36     V2 + V35
    V36     V36 956.34 9 V34 909 1.1 V22 + V34
898.33 0 V23 + V31 899.13 0.1 V24 + V31 867.84 11 V1 + V31 819.29 15 V6 + V31
d759.07 6 V26 d761.2 11 V26 856.93 12 V3 + V29 763.93 9 V1 + V28
    V27     V27 853.58 14 V1 + V30     V28
d730.52 29 V23 + V31 d727.85 53 V23 759.75 2 V24 + V27 759.77 11 V3 + V26
    V24     V24 + V31 d752.41 1 V7 + V25 735.84 13 V24
    V25     V25 + V33     V23 725.42 14 V4 + V23
707.47 26 V2 + V15 d695.49 64 V27 706.49 8 V11 + V20
        694.67 66 V26 703.34 14 V21
        d684.82 1.7 V2 + V19 672.6 45 V17
d582.71 22 V11 582.82 36 V11 627.05 0.16 V11 + V15 537.22 12 V9 + V2
    V8 + V10     V7 + V10     V17 + V29     V9
    V7 + V9     V9     V7 + V25    
d433.24 91 V4 d427.12 160 V4 d444.4 188 V9 406.05 166 V6
    V5     V5     V8 394.97 172 V4 + V5
    V6     V6     V7 + V25 305.98 70 V4 + V23


Anharmonic effects lead to the appearance of combination bands and frequency shifts, as shown in Table 2. Unfortunately, experimental infrared spectra for the small fullerene C20 are not yet available. The following subsections present and discuss theoretically predicted anharmonic IR spectra of C20, N10C10, and C12N8 in their neutral, cationic, and anionic forms, both in the gas phase and in water solvent. We note that for some charged species, particularly the N10C10 anion, the HOMO–LUMO gap is relatively small (∼1.88 eV), which can raise concerns about low-lying excited states influencing the reliability of ground-state DFT vibrational spectra.41 However, all structures optimized in this study are confirmed to be true minima with no imaginary frequencies. The agreement of our calculated IR features with known vibrational bands in related carbonaceous molecules supports the validity of our results within the expected accuracy of the method. It is known that IR intensities often increase in fullerenes when they gain an electron (anion form), especially for C–C stretching modes. Our results follow this trend for C20 and its N-substituted forms. In contrast, PAH cations are known to show strong C–C stretches. In our case, nitrogen substitution changes the symmetry and causes the C–C stretch bands to become weaker or spread out, rather than becoming sharper.

3.2.1 C20. Neutral C20 harmonic vibrational spectra have two intense IR-active modes: CC stretching at 1298 (7.7 μm) and 1296 (7.716 μm) cm−1, and CCC bending modes at 706 (14.16 μm) and 705 cm−1 (14.12 μm). The anharmonic infrared spectrum of neutral C20 in the water solvent phase shows that the most intense band appears at 1270.56 (7.87 μm) in 2–20 μm, which is attributed to a CC stretching mode of the V41 fundamental mode. The significantly intense mode in the spectra of C20 in water solvent is at 1319.3 cm−1 (7.58 μm), corresponding to the V51 fundamental and V1 + V51 combination modes of CC stretching. The ring distortion mode at 481.85 cm−1 (20.75 μm) with a large intensity corresponds to the fundamental mode V11 and a linear combination of the states V11 + V21.
3.2.2 N10C10. The CN stretching mode has a strong intensity at 859.31 cm−1 (11.63 μm) and 733.75 cm−1 (13.62 μm), corresponding to the V36 and V26 fundamental modes, respectively. However, in the water solvent state, the fundamental mode V26 at 726.59 cm−1 (13.76 μm) has a quite intense peak. Other significant infrared features for this molecule in the cationic form are present in the CC stretching mode at 1642.53 cm−1 (6.09 μm) corresponding to the fundamental mode V48, and at 1205.57 cm−1 (8.279 μm) corresponding to the V33 + V44 combination and V33 fundamental modes. N10C10 in the anionic form shows significantly intense peaks at 1474.29 cm−1 (6.78 μm) and 1294.07 cm−1 (7.72 μm) corresponding to the V51 fundamental and V1 + V51 combinational modes, respectively. The ring distortion mode at 712 cm−1 becomes quite intense, which corresponds to the V26 fundamental mode.
3.2.3 C12N8. The most intense IR bands for neutral C12N8 are present in the 2–20 μm region at 1620.2 cm−1 (6.17 μm), which is attributed to the V50 fundamental mode of CC stretching. The strong IR feature is dominated by the CN stretching modes at 1071.03 cm−1 (9.33 μm) and 1069.4 cm−1 (9.35 μm), corresponding to the V41 and V42 fundamental modes. The ring distortion feature appears at 433.24 cm−1 (23.08 μm). In the water solvent state, the CC stretching typically occurs at 1619.15 cm−1 (6.176 μm) corresponding to the V50, V51, V52 fundamental degenerated modes. The substitution of eight nitrogen atoms in the C20 fullerene introduces intense peaks with degeneracy at ∼1062 cm−1 (9.41 μm), corresponding to the V40, V41, and V42 fundamental modes. The ring distortion of C12N8 in water solvent is quite intense at 427.12 cm−1 (23.41 μm), compared to the neutral gas phase spectra.

3.3 Electronic absorption spectra

Electronic absorption spectra in the UV-visible region are studied using TD-DFT in the neutral and ionic states. The wavelength of the electronic transition, absorbance, oscillator strength, and the HOMO to LUMO gap of these molecules are summarized in Table 3, and corresponding spectra are shown in Fig. 3. The strong absorption bands from 150 to 300 nm are attributed to π → π* transitions.
Table 3 Wavelength of electronic transitions, absorbance, oscillator strength, and transitions of C20 fullerene and its nitrogen-substituted heterofullerenes
Molecule Wavelength (nm) Absorbance Oscillator strength Transitions aH–L Energy gap (eV)
a H and L represent HOMO and LUMO, respectively.
Neutral in gas phase
C20 162.26 13[thin space (1/6-em)]325 0.0746 H−1 → L+11 1.9168
  232.96 37[thin space (1/6-em)]316 0.5118 H−3 → L+2  
  296.43 19[thin space (1/6-em)]378 0.1882 H → L+7  
N10C10 158.83 19[thin space (1/6-em)]306 0.0194 H−1 → L+15 4.5535
  255.16 5867 0.0409 H−1 → L+1  
C12N8 163.37 9115 0.0467 H−1 → L+17 2.8683
  195.46 14[thin space (1/6-em)]662 0.0177 H−7 → L  
  241.35 11[thin space (1/6-em)]257 0.0809 H−5 → L+1  
[thin space (1/6-em)]
Neutral in water solvent
C20 161.9 21[thin space (1/6-em)]006 0.0673 H−1 → L+11 1.9252
  239.27 44[thin space (1/6-em)]455 0.4408 H−3 → L+5  
  300.45 26[thin space (1/6-em)]651 0.2749 H → L+7  
N10C10 177.32 15[thin space (1/6-em)]156 0.1369 H → L+14 2.5565
  221.07 9446 0.1498 H−2 → L+4  
  310.8 11[thin space (1/6-em)]382 0.1189 H−3 → L+1  
C12N8 164.18 10[thin space (1/6-em)]855 0.0521 H−2 → L+17 2.9154
  192.05 17[thin space (1/6-em)]931 0.0913 H−8 → L  
  242.47 15[thin space (1/6-em)]284 0.1135 H−5 → L  
[thin space (1/6-em)]
Cation in gas phase
C20 236.14 45[thin space (1/6-em)]400 0.3153 H−4 → L+5 1.8647
N10C10 180.72 1138 0.0181 H−7 → L+3 2.4153
  222.59 11[thin space (1/6-em)]605 0.1042 H−2 → L+4  
  300.75 8484 0.0755 H−7 → L  
C12N8 157.58 1246 0.003 H−1 → L+13 2.3624
  215.27 11[thin space (1/6-em)]624 0.0599 H−6 → L+5  
  244.12 9214 0.0504 H → L+7  
  302.8 3909 0.03 H−2→L+4  
[thin space (1/6-em)]
Anion in gas phase
C20 242.93 30[thin space (1/6-em)]947 0.0503 H−5 → L+4 1.9282
  280.72 26[thin space (1/6-em)]785 0.1073 H−2 → L+4  
N10C10 204.88 12[thin space (1/6-em)]411 0.0727 H → L+14 1.8784
  236.03 10[thin space (1/6-em)]191 0.0548 H−2 → L+4  
  307.95 5758 0.0569 H−8 → L  
C12N8 220.34 14[thin space (1/6-em)]040 0.0162 H−6 → L+5 2.4
  263.38 9685 0.076 H → L+10  
  503.54 1517 0.0165 H → L+2  



image file: d5ra05271h-f3.tif
Fig. 3 UV-Visible spectra of C20 fullerene and, its nitrogen-substituted heterofullerenes in their neutral and ionic states in the gas phase, and neutral in the water solvent. The spectra are calculated at the B3LYP/6-311++G(d,p) level of theory.

The C20 fullerene in the gas phase and its nitrogen substitution heterofullerenes N10C10 and C12N8 possess high-energy transitions at 162.26, 158.83, and 163.37 nm with major contributions from H−1 → L+11, H−1 → L+15, and H−1 → L+17, respectively. These high-energy transitions often result in significant changes in electronic configuration, indicating strong electronic coupling and potentially highly reactive excited states, which can affect the photostability, ionization, and dissociation behavior of these molecules in astrophysical environments where intense UV radiation is present.100–102 Due to nitrogen's high electronegativity, the energy gap between states increases, which influences electronic transitions in nitrogen-substituted heterofullerenes. There is no effect of water solvent on high-energy transitions of C20 and its heterofullerenes. Almost all transitions are observed in water solvent similarly to the gas phase state. In the case of ionic states especially in anion, there is no single high-energy transition observed. This may be because core-level electrons are unaffected by the addition of extra electrons. The C20 cation and its nitrogen heterofullerenes (N10C10 and C12N8) show high-energy transitions with major contributions from H−7 → L+3 and H−1 → L+13, respectively.

The HOMO–LUMO energy gap is a crucial parameter in determining the electronic properties of a molecule. A large HOMO–LUMO gap indicates higher kinetic stability with lower reactivity, representing greater chemical hardness.103–105 Compared to the parent fullerene C20, the N10C10 heterofullerene in the neutral gas phase shows a relatively large gap of 4.55 eV, indicating enhanced stability and reduced reactivity. In contrast, in the water solvent state N10C10 exhibits low kinetic stability, with a much smaller gap of 2.56 eV. Low-energy transitions are observed for all structures in the water solvent due to extended conjugation, electron delocalization, nitrogen substitution, and the absence of core-level transitions. In the case of cations and anions, almost all molecules exhibit lower energy transitions. Highly reactive and lowest stable species is N10C10 in anion, which shows the lowest HOMO–LUMO gap at 1.8784 eV, while the C12N8 heterofullerene has unique electronic properties, including a significantly large HOMO-to-LUMO energy gap in the anionic state due to a strong polarization between the N–C bonds. The position and number of substituted heteroatoms in the C20 fullerene affect the heterofullerenes HOMO–LUMO energy gap. This energy gap is also affected by the ionic charge states of the heterofullerene.

4 Astrophysical implications

The present results show that nitrogen-substituted C20 fullerenes in their anionic form exhibit strong IR-active modes, particularly in the 6–15 μm range. This suggests that such species may contribute to the observed interstellar infrared absorption features in cold or dense regions, where anionic forms are more stable due to the shielding from UV radiation; however, their role in emission spectra is likely limited to environments with large UV excitation. The enhanced IR activity, resulting from the symmetry reduction via nitrogen substitution, increases their detectability compared to the pure C20. Furthermore, our calculated ionization potentials (IP) and electron affinities (EA) support the idea that anionic species are favored in shielded environments, while cationic forms are more likely to exist in UV-irradiated regions such as photodissociation regions. These findings suggest the variations in charge states and IR features of these molecules across different astrophysical environments. The detection of fullerenes C60 and C70 in the planetary nebula TC-1 by Cami et al.6 marked a major step forward in understanding fullerene chemistry in space. Although indene was recently detected in the TMC-1 molecular cloud,106 no gas-phase formation pathway for fullerene-like molecules is currently known under such low-temperature conditions. In support of this, Lorenzo et al.107 experimentally demonstrated the formation of the C20 fullerenes and small carbon clusters in laser-induced plasma, mimicking the conditions in the vicinity of the central star of planetary nebulae. These developments suggest that small fullerenes like C20 and its derivatives are likely to be present in space, and the present anharmonic calculations provide important spectral fingerprints to aid their future identification.

In dense molecular clouds, dust grains are covered with ice mantles primarily composed of H2O, CO, and CO2.108–111 Numerous observations using ISO-SWS and Spitzer confirm that H2O, CO, and CO2 are the most abundant and ubiquitous molecules frozen in mantles on interstellar grains.112,113 After absorbing cosmic rays or UV radiation, the surface reactions of frozen molecules could lead to the formation of carbonaceous material containing fullerenes.110 We observed that neutral anharmonic C20 fullerene has strong features at around 7.7, 8.0, 8.6, 9.2, and 11.3 μm in the gas and around 7.8, 8.0, 8.6, 9.6, and 10.8 μm in the water solvent state. Although the water solvent (ice mantle) has an insignificant effect on the peak positions of C20, this implies that distinguishing its presence in icy versus gas-phase environments solely based on vibrational band positions will be challenging, especially in absorption studies where precise shifts are crucial. The features of the cationic form appear at around 7.8, 8.2, 8.6, 11.2, and 17.1 μm with very weak intensity, while the anionic form exhibits features with significant intensity at 7.7, 8.03, 8.6, 11.2, and 18.9 μm. The vibrational spectra of the heterofullerene N10C10 display intense peaks at ∼6.1, ∼8.2, 10.5, and 12.6 μm in the cationic form. In the anion state, it shows strong features at 7.8, 8.6, and 10.5 μm. While the neutral form of C12N8 in the water solvent state shows strong features at around 6.2 and 10.5 μm, weak features are also observed at ∼8.7 and ∼11.2 μm. On the other hand, CN and CC stretching vibrational modes are quite intense, peaking at 6.8, 8.4 μm, and 14.86, 18.61 μm in the C12N8 anionic form. The differences in intensities among the vibrational spectra of nitrogen-containing molecules could be attributed to the position and number of heteroatoms in the parent fullerene C20. Experimental and theoretical studies by Mattioda et al., Hudgins et al., and Vats et al.114–116 suggest that nitrogen-containing PAHs (PANHs), in neutral, cationic, and anionic forms, may contribute to the aromatic infrared bands (AIBs) observed in the interstellar medium (ISM). Recent JWST observations of the Orion Bar have revealed high-resolution AIB spectra that help constrain potential carriers more precisely.117 Furthermore, carbon-rich dust containing nitrogen, with infrared features similar to those seen in novae, underscores the astrophysical importance of nitrogenated species.118

To explore the potential presence of C20 and its nitrogen-substituted derivatives in space, we selected two planetary nebulae (Tc 1 and NGC 7027) and two reflection nebulae (NGC 2023 and NGC 7023), based on their well-characterized mid-infrared emission features. These objects are ideal for comparison with the computed spectra of C20 fullerenes in neutral and ionic states. A summary of their key physical properties and observed IR bands is provided in Table 4. Their spectra are compared with the theoretical results in Fig. 4 to evaluate possible spectral matches. While some agreement is seen, further observational confirmation is required. A direct comparison of the computed spectra with the observed IR spectra of these nebulae is presented in Fig. 4. This comparison suggests that C20 and its N-substituted species could contribute to some of the observed features in these astronomical sources. However, due to the current computational limitations, the assignments remain tentative. Additional experimental and observational efforts are needed—particularly using high-resolution and high-signal-to-noise-ratio spectra-to confirm the presence of these species and refine their spectroscopic identification.

Table 4 Characteristics of the objects studied
Object Object Type Teff (K) Detected features (μm)
TC1 Planetary nebula 34[thin space (1/6-em)]700 (ref. 123) 6.23 7.0 (ref. 6) 7.7 8.51 8.6 11.3  
NGC 7027 Planetary nebula 200[thin space (1/6-em)]000 (ref. 124 and 125) 6.2 7.65 (ref. 126) 8.6 11.3 12.6
NGC 2023 Reflection nebula 22[thin space (1/6-em)]000 (ref. 5) 6.2 7.6 8.6 11.25  
NGC 7023 Reflection nebula 17[thin space (1/6-em)]000 (ref. 127) 6.2 7.7 8.6 11.25  



image file: d5ra05271h-f4.tif
Fig. 4 Theoretical vibrational spectra of the C20 fullerene and its nitrogen-substituted heterofullerenes in their neutral (gas phase and water solvent) and ionic states, compared with the observed spectra of four astronomical objects. The vertical dotted lines mark the positions of the interstellar aromatic infrared bands (AIBs), including the 7.0 μm band attributed to neutral C60. Observational spectra are plotted in surface brightness units (MJy sr−1), while theoretical spectra are shown as normalized intensities (a.u.) to facilitate direct comparison.

The UV bump at 217.5 nm in the interstellar extinction curve is attributed to the carbonaceous dust grains in space, although exact identification of the carrier remains an open question.63 Massa et al.119 show that the area of the 217.5 nm extinction bump and the strengths of the major AIB arise show a strong correlation for the same lines of sight, suggesting common carriers for both the UV extinction and the AIB emission. However, observational searches for PAH signatures in 400–700 nm have so far been unsuccessful in the interstellar extinction.66,120–122 Theoretical UV-visible spectra for the C20 fullerene and heterofullerenes show that the neutral C20 appears to have significant absorption at 232.96 nm, whereas in the water solvent, it is at 221.07 nm for N10C10. Apart from this, in the ionic state, the UV bump is observed at 222.59 nm for N10C10 and 215.27 nm for C12N8 in the cationic state, while in the anionic form, it is at 220.34 nm for C12N8. These results suggest that part of the 217.5 nm may have a contribution from C20 fullerene and heterofullerenes. Non-detection of the second strong feature around 300 nm suggests that the contribution should be limited. Further studies of the C20 fullerene, particularly in laboratory experiments, are needed to make a detailed study of the presence of the C20 fullerene in the ISM. The present study provides UV to IR spectra of the C20 fullerene and heterofullerenes for future studies of the possible presence of these small carbon clusters. These species may be able to survive in the presence of other fullerenes in the interstellar medium and circumstellar envelopes.

5 Summary

The discovery of C60, C60+, and C70 fullerenes in interstellar and circumstellar environments opened a new window to study the formation and evolution of carbonaceous species in the ISM. It suggests possibilities for the presence of other fullerenes and their derivatives. We conduct a theoretical study on the C20 fullerene and its nitrogen-substituted heterofullerenes for a comprehensive analysis of IR vibrational and electronic absorption spectral features. The analyses are performed for both the neutral and ionic states in the gas state, and also for the neutral water solvent state using the B3LYP/6-311++G(d,p) level of theory taking account of the anharmonic effects. We find significant effects in the vibrational and electronic absorption spectral features, when nitrogen-substitution occurs in the parent fullerene C20. The present results also show that C12N8 has a significant dipole moment in the anionic state with the lowest symmetry. The electron affinity of C20 is 2.32 eV, which is very close to the experimental measurement.

We compare the IR spectra observed in four astronomical objects with those of the C20 fullerene and heterofullerenes obtained in this study for the neutral and charge states. The wavelengths of strong vibrational modes for three molecules in their neutral and ionic states have peaks close to the peaks at 6.2, 6.6, 7.0, 7.7, 8.5, 8.6, 11.2, and 11.3 μm observed in astronomical objects. IR spectroscopic observations with the James Webb Space Telescope (JWST) will constitute a significant advancement in observational astronomy owing to their exceptional sensitivity and resolution. A vast array of astrophysical domains will be significantly impacted by the capacity to identify certain species, estimate their abundances or provide upper bounds on non-detections. This epoch holds the potential to augment our comprehension of the chemical composition of the universe, the mechanisms behind the development of stars and planets, the progression of galaxies, and the underlying essence of the cosmos.

We also report the electronic absorption spectra of these isomers using the TDDFT for all isomers in neutral and their ionic charge states. The heterofullerenes in the neutral form show 221.07 nm for N10C10 in the water solvent. In ionic states, these molecules have a broad absorption bump at 222.59 nm for N10C10 and 215.27 nm for C12N8 in cationic form, and in the anionic form at 220.34 nm for C12N8. The substitution of many heteroatoms in the C20 fullerene significantly affects the HOMO–LUMO energy gap in the gas phase, and this energy gap is also strongly affected by the water solvent and ionic charge states of the molecules. The HOMO-to-LUMO energy gap of a neutral N10C10 exhibits a high energy gap of 4.5535 eV in the gas phase, but it is observed in the water solvent as 1.9252 eV. Changes in the energy gap can influence the molecule's chemical stability and its ability to participate in electronic transitions. Larger gaps often correspond to more stable structures, while smaller gaps may enhance reactivity, which is important for catalysis or chemical sensing.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Data availability

The data that support the findings of this study are available within the article and its supplementary information (SI) files. Additional data, including computational input/output files, optimized geometries, and detailed numerical results, are available from the corresponding author upon reasonable request. Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra05271h.

Acknowledgements

The project was funded by ISRO, Bangalore, India, under the RESPOND Programme (Grant No. ISRO/RES/2/425/19-20). The authors are also thankful to IUCAA for references and library facilities. TO acknowledges the support by the Japan Society for the Promotion of Science (JSPS) KAKENHI Grant Number 24K07087.

References

  1. A. G. G. M. Tielens, Annu. Rev. Astron. Astrophys., 2008, 46, 289–337 CrossRef.
  2. A. Li, Nat. Astron., 2020, 4, 339–341 CrossRef.
  3. A. Léger and J. Puget, Astron. Astrophys., 1984, 137, L5–L8 Search PubMed.
  4. L. J. Allamandola, A. G. G. M. Tielens and J. R. Barker, Astrophys. J., Suppl. Ser., 1989, 71, 733–775 CrossRef.
  5. E. Peeters, A. G. G. M. Tielens, L. J. Allamandola and M. G. Wolfire, Astrophys. J., 2012, 747, 44–62 CrossRef.
  6. J. Cami, J. Bernard-Salas, E. Peeters and S. E. Malek, Science, 2010, 329, 1180–1182 CrossRef PubMed.
  7. K. Sellgren, M. W. Werner, J. G. Ingalls, J. Smith, T. Carleton and C. Joblin, Astrophys. J. Lett., 2010, 722, L54–L57 CrossRef.
  8. C. Joblin, A. G. G. M. Tielens and F. Pauzat, EAS Publ. Ser., 2011, 46, 75–91 CrossRef.
  9. O. Berné, G. Mulas and C. Joblin, Astron. Astrophys., 2013, 550, L4–L7 CrossRef.
  10. E. K. Campbell, M. Holz, D. Gerlich and J. P. Maier, Nature, 2015, 523, 322–323 CrossRef PubMed.
  11. E. K. Campbell, M. Holz, J. P. Maier, D. Gerlich, G. A. H. Walker and D. Bohlender, Astrophys. J., 2016, 822, 17 CrossRef.
  12. G. A. H. Walker, D. A. Bohlender, J. P. Maier and E. K. Campbell, Astrophys. J. Lett., 2015, 812, L8 CrossRef.
  13. R. Lallement, J. Capitanio, L. Babusiaux, F. Arenou, L. Vergely, M. Valette, M. Hottier and C. Gry, Astron. Astrophys., 2018, 614, A28 CrossRef.
  14. M. A. Cordiner, H. Linnartz, N. L. J. Cox, P. C. Myers, S. Bailleux, J. Cami, S. N. Milam and D. A. García-Hernández, Astrophys. J. Lett., 2019, 875, L28 CrossRef.
  15. H. W. Kroto, J. R. Heath, S. C. O'Brien, R. F. Curl and R. E. Smalley, Nature, 1985, 318, 162–163 CrossRef.
  16. W. Kratschmer, L. D. Lamb, K. Fostiropoulos and D. R. Huffman, Nature, 1990, 347, 354–358 CrossRef.
  17. D. A. García-Hernández, A. Manchado, P. García-Lario, L. Stanghellini, E. Villaver, R. A. Shaw and J. V. Perea-Calderón, Astrophys. J. Lett., 2010, 724, L39–L43 CrossRef.
  18. D. A. García-Hernández, S. Iglesias-Groth, J. A. Acosta-Pulido, A. Manchado, P. García-Lario, L. Stanghellini, E. Villaver, R. A. Shaw and F. Cataldo, Astrophys. J. Lett., 2011, 737, L30 CrossRef.
  19. K. Sellgren, M. W. Werner, J. G. Ingalls, J. Smith, T. Carleton and C. Joblin, Astrophys. J. Lett., 2010, 722, L54–L57 CrossRef CAS.
  20. E. Peeters, A. G. G. M. Tielens, L. J. Allamandola and M. G. Wolfire, Astrophys. J., 2012, 747, 44 CrossRef.
  21. Y. Zhang and S. Kwok, Astrophys. J., 2011, 730, 126 CrossRef.
  22. K. R. Roberts, K. T. Smith and P. J. Sarre, Mon. Not. R. Astron. Soc., 2012, 421, 3277 CrossRef CAS.
  23. O. Berné, G. Mulas and C. Joblin, Astron. Astrophys., 2013, 550, L4 CrossRef.
  24. D. A. García-Hernández, E. Villaver, P. García-Lario, J. A. Acosta-Pulido, A. Manchado, L. Stanghellini, R. A. Shaw and F. Cataldo, Astrophys. J., 2012, 760, 107 CrossRef.
  25. X.-H. Chen, F.-Y. Xiang, X.-J. Yang and A. Li, Res. Astron. Astrophys., 2019, 19, 141 CrossRef CAS.
  26. S. Kuzmin and W. Duley, arXiv, 2011, preprint, arXiv:1103.2989,  DOI:10.48550/arXiv.1103.2989.
  27. L. Bernstein, R. Shroll, D. Lynch and F. Clark, Astrophys. J., 2017, 836, 229 CrossRef.
  28. J.-J. Adjizian, A. Vlandas, J. Rio, J.-C. Charlier and C. P. Ewels, EAS Publ. Ser., 2016, 374, 20150323 Search PubMed.
  29. M. A. Gómez-Muñoz, D. A. García-Hernández, R. Barzaga, A. Manchado and H.-T. Roldán, Astron. Astrophys., 2024, 682, L18 CrossRef.
  30. Y. Zhang, S. Sadjadi and C. H. Hsia, Astrophys. Space Sci., 2020, 365, 67 CrossRef.
  31. M. Koohi, S. Soleimani Amiri and B. N. Haerizade, J. Phys. Org. Chem., 2017, 30, 3682–3689 CrossRef.
  32. B. Foing and P. Ehrenfreund, Nature, 1994, 369, 296 CrossRef.
  33. E. K. Campbell, M. Holz, D. Gerlich and J. P. Maier, Nature, 2015, 523, 322–323 CrossRef.
  34. G. A. H. Walker, D. A. Bohlender, J. P. Maier and E. K. Campbell, Astrophys. J. Lett., 2015, 812, L8 CrossRef.
  35. R. Lallement, J. Capitanio, L. Babusiaux, F. Arenou, L. Vergely, M. Valette, M. Hottier and C. Gry, Astron. Astrophys., 2018, 614, A28 CrossRef.
  36. M. A. Cordiner, H. Linnartz, N. L. J. Cox, P. C. Myers, S. Bailleux, J. Cami, S. N. Milam and D. A. García-Hernández, Astrophys. J. Lett., 2019, 875, L28 CrossRef.
  37. T. Misawa, P. Gandhi, A. Hida, T. Tamagawa and T. Yamaguchi, Astrophys. J., 2009, 700, 1988–2002 CrossRef.
  38. D. Majaess, T. A. Harriott, H. Seuret, C. Morera-Boado, L. Massa and C. F. Matta, Mon. Not. R. Astron. Soc., 2025, 538, 2392–2402 CrossRef.
  39. S. Iglesias-Groth, Mon. Not. R. Astron. Soc., 2019, 489, 1509–1516 Search PubMed.
  40. S. Sadjadi, Q. A. Parker, C. H. Hsia and Y. Zhang, Astrophys. J., 2022, 934, 75 CrossRef.
  41. A. Candian, M. Gomes Rachid, H. MacIsaac, V. N. Staroverov, E. Peeters and J. Cami, Mon. Not. R. Astron. Soc., 2019, 485, 1137–1149 CrossRef.
  42. S. Sadjadi and Q. A. Parker, Fullerenes, Nanotubes Carbon Nanostruct., 2021, 29, 620–628 CrossRef.
  43. R. Bauernschmitt, R. Ahlrichs, F. H. Hennrich and M. M. Kappes, J. Am. Chem. Soc., 1998, 120, 5052–5053 CrossRef.
  44. K. Balasubramanian, Polycyclic Aromat. Compd., 2022, 42, 1649–1674 CrossRef.
  45. H. Kroto, Nature, 1987, 329, 529 CrossRef.
  46. B. Zhang, C. Wang, K. Ho, C. Xu and C. T. Chan, J. Chem. Phys., 1992, 97, 5007–5010 CrossRef.
  47. M. Feyereisen, M. Gutowski, J. Simons and J. Almlöf, J. Chem. Phys., 1992, 96, 2926–2931 CrossRef.
  48. H. Kroto and K. McKay, Nature, 1988, 331, 328 CrossRef.
  49. C. Brabec, E. Anderson, B. Davidson, S. Kajihara, Q.-M. Zhang, J. Bernholc and D. Tomanek, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 46, 7326–7332 CrossRef PubMed.
  50. H. Prinzbach, A. Weiler, P. Landenberger, F. Wahl, J. Wörth, L. T. Scott, B. Issendorff and K. V. Klitzing, Nature, 2000, 407, 60–63 CrossRef PubMed.
  51. E. Sackers, T. Oßwald, K. Weber, M. Keller, D. Hunkler, J. Wörth, L. Knothe and H. Prinzbach, Chem.–Eur. J., 2006, 12, 6242–6250 CrossRef.
  52. Z. Wang, X. Ke, Z. Zhu, F. Zhu, M. Ruan, H. Chen, R. Huang and L. Zheng, Phys. Lett. A, 2001, 280, 351–356 CrossRef.
  53. M. Zamani, A. Motahari, H. A. Dabbagh and H. Farrokhpour, J. Nano Anal., 2014, 1, 31–38 Search PubMed.
  54. J. M. L. Martin, J. El-Yazal and J.-P. François, Chem. Phys. Lett., 1996, 248, 345–351 CrossRef.
  55. E. Małolepsza, H. A. Witek and S. Irle, J. Phys. Chem. A, 2007, 111, 6649–6656 CrossRef PubMed.
  56. J. Averdung, G. Torres-Garcia, H. Luftmann, I. Schlachter and J. Mattay, Fullerenes, Nanotubes Carbon Nanostruct., 1996, 4, 633–637 Search PubMed.
  57. J. C. Hummelen, C. Bellavia-Lund and F. Wudl, Fullerenes and Related Structures, 1999, vol. 93, pp. 1–17 Search PubMed.
  58. T. Pradeep, V. Vijayakrishnan, A. Santra and C. Rao, J. Phys. Chem., 1991, 95, 10564–10570 CrossRef.
  59. K.-C. Sun and C. Chen, J. Mol. Struct.: THEOCHEM, 1996, 360, 157–161 CrossRef.
  60. H. Kawabata and H. Tachikawa, C, 2017, 3, 15 Search PubMed.
  61. T. P. Stecher, Astron. J., 1965, 70, 693–698 CrossRef.
  62. C. Joblin, A. Léger and P. Martin, Astrophys. J. Lett., 1992, 393, L79–L82 CrossRef.
  63. B. T. Draine, Annu. Rev. Astron. Astrophys., 2003, 41, 241–289 CrossRef.
  64. G. Malloci, G. Mulas and C. Joblin, Astron. Astrophys., 2004, 426, 105–117 CrossRef.
  65. A. Li and J. M. Greenberg, Astron. Astrophys., 1997, 323, 566–584 Search PubMed.
  66. M. Steglich, Y. Carpentier and C. Jäger, et al., Astron. Astrophys., 2012, 540, A110 CrossRef.
  67. M. Braga, S. Larsson, A. Rosen and A. Volosov, Astron. Astrophys., 1991, 245, 232–238 Search PubMed.
  68. S. Iglesias-Groth, Astrophys. J. Lett., 2004, 608, L37–L40 CrossRef.
  69. S. Soleimani Amiri and B. Mirza, J. Phys. Org. Chem., 2016, 29, 514–520 CrossRef.
  70. G. Wenzel, D. A. García-Hernández, A. Manchado, S. Iglesias-Groth, J. A. Acosta-Pulido, F. Cataldo and R. A. Shaw, Astrophys. J. Lett., 2025, 984, L36 CrossRef.
  71. B. A. McGuire, A. M. Burkhardt, S. Kalenskii, C. N. Shingledecker, A. J. Remijan, E. Herbst and M. C. McCarthy, Science, 2018, 359, 202–205 CrossRef PubMed.
  72. B. A. McGuire, C. N. Shingledecker, A. M. Burkhardt, K. I. Öberg, A. J. Remijan and M. C. McCarthy, Science, 2021, 371, 1265–1270 CrossRef.
  73. M. C. McCarthy, A. J. Remijan, G. A. Blake, K. I. Öberg, A. M. Burkhardt and B. A. McGuire, Nat. Astron., 2021, 5, 176–180 CrossRef.
  74. K. L. K. Lee, H. Y. Chu, K. M. Mak and C. K. Yung, et al., Astrophys. J. Lett., 2021, 910, L2 CrossRef.
  75. M. K. Sita, D. A. García-Hernández, A. Manchado and F. Cataldo, Astrophys. J. Lett., 2022, 938, L12 CrossRef.
  76. J. Cernicharo, C. Cabezas, R. Fuentetaja, M. Agúndez, B. Tercero, J. Janeiro, M. Juanes, R. I. Kaiser, Y. Endo, A. L. Steber, D. Pérez, C. Pérez, A. Lesarri, N. Marcelino and P. de Vicente, Astron. Astrophys., 2024, 690, L13 CrossRef CAS.
  77. J. Cernicharo, C. Cabezas, M. Agúndez, B. Tercero and N. Marcelino, Astron. Astrophys., 2021, 655, L1 CrossRef CAS.
  78. G. Wenzel, R. A. Shaw, D. A. García-Hernández and A. Manchado, Science, 2024, 386, 810–815 CrossRef CAS PubMed.
  79. G. Wenzel, D. A. García-Hernández and A. Manchado, Nat. Astron., 2025, 9, 262–268 CrossRef.
  80. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, et al., Gaussian 16, Revision C.01, Gaussian, Inc., Wallingford, CT, 2016 Search PubMed.
  81. W. An, Y. Gao, S. Bulusu and X. C. Zeng, J. Chem. Phys., 2005, 122, 114 CrossRef.
  82. F. A. Gianturco, G. Y. Kashenock, R. R. Lucchese and N. Sanna, J. Chem. Phys., 2002, 116, 2811–2818 CrossRef CAS.
  83. M. Saito and Y. Miyamoto, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 65, 165434 CrossRef.
  84. C. W. Bauschlicher Jr, A. Ricca, A. L. Mattioda and L. J. Allamandola, Astrophys. J., 2009, 697, 311–327 CrossRef.
  85. A. Ricca, C. W. Bauschlicher Jr and L. J. Allamandola, Astrophys. J., 2012, 754, 75 CrossRef.
  86. L. Goerigk and S. Grimme, J. Chem. Theory Comput., 2011, 7, 291–309 CrossRef CAS.
  87. Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2008, 120, 215–241 Search PubMed.
  88. E. Cances, B. Mennucci and J. Tomasi, J. Chem. Phys., 1997, 107, 3032–3041 CrossRef CAS.
  89. B. Mennucci and J. Tomasi, J. Chem. Phys., 1997, 106, 5151–5158 CrossRef CAS.
  90. J. Tomasi, B. Mennucci and E. Cancès, J. Mol. Struct.: THEOCHEM, 1999, 464, 211–226 CrossRef CAS.
  91. CRC Handbook of Chemistry and Physics, ed. D. R. Lide, CRC Press, Boca Raton, FL, 86th edn, 2005 Search PubMed.
  92. V. Barone, J. Chem. Phys., 2005, 122, 014108 CrossRef.
  93. National Institute of Standards and Technology (NIST), Computational Chemistry Comparison and Benchmark Database, ed. R. D. Johnson III, NIST Standard Reference Database Number 101, Release 22, May 2022, Available at: http://cccbdb.nist.gov/ Search PubMed.
  94. M. Naganathappa and A. Chaudhari, Mon. Not. R. Astron. Soc., 2012, 425, 490–496 CrossRef CAS.
  95. J. E. Huheey, E. A. Keiter, R. L. Keiter, and O. K. Medhi, Inorganic Chemistry: Principles of Structure and Reactivity, Pearson Education India, 2006 Search PubMed.
  96. I. Dabo, A. Ferretti, N. Poilvert, Y. Li, N. Marzari and M. Cococcioni, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 115121 CrossRef.
  97. E. Runge and E. K. U. Gross, Phys. Rev. Lett., 1984, 52, 997–1000 CrossRef.
  98. S. Gorelsky, AOMix Program for Molecular Orbital Analysis, version 6.5, University of Ottawa, Ottawa, Canada, 2011 Search PubMed.
  99. M. Miar, A. Shiroudi, K. Pourshamsian, A. R. Oliaey and F. Hatamjafari, J. Chem. Res., 2021, 45, 147–154 CrossRef.
  100. I. N. Levine, Quantum Chemistry, Pearson Education, 7th edn, 2014 Search PubMed.
  101. F. Jensen, Introduction to Computational Chemistry, Wiley, 3rd edn, 2017 Search PubMed.
  102. A. Szabo and N. S. Ostlund, Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory, Dover Publications, 1996 Search PubMed.
  103. R. G. Pearson, Proc. Natl. Acad. Sci. U. S. A., 1986, 83, 8440–8441 CrossRef PubMed.
  104. R. G. Pearson, J. Org. Chem., 1989, 54, 1423–1430 CrossRef.
  105. Z. Zhou and R. G. Parr, J. Am. Chem. Soc., 1989, 111, 7371 CrossRef.
  106. J. Cernicharo, C. Cabezas, M. Agúndez, B. Tercero and N. Marcelino, Astron. Astrophys., 2021, 655, L1 CrossRef.
  107. T. Lorenzo, A. Torrisi and M. Cutroneo, Fullerenes, Nanotubes Carbon Nanostruct., 2025, 33, 111–120 CrossRef.
  108. S. A. Sandford and L. J. Allamandola, Astrophys. J., 1993, 417, 815–825 CrossRef PubMed.
  109. S. A. Sandford, Meteorit. Planet. Sci., 1996, 31, 449–452 CrossRef PubMed.
  110. M. P. Bernstein, J. P. Dworkin, S. A. Sandford and L. J. Allamandola, Meteorit. Planet. Sci., 2001, 36, 351–358 Search PubMed.
  111. F. Dulieu, L. Amiaud, E. Congiu, J. H. Fillion, E. Matar, A. Momeni and J. L. Lemaire, Astron. Astrophys., 2010, 512, A30 CrossRef.
  112. T. de Graauw, et al., Astron. Astrophys., 1996, 315, L345–L348 Search PubMed.
  113. K. M. Pontoppidan, A. C. Boogert, H. J. Fraser, E. F. van Dishoeck, G. A. Blake, F. Lahuis and C. Salyk, Astrophys. J., 2008, 678, 1005–1031 Search PubMed.
  114. A. L. Mattioda, D. M. Hudgins, C. W. Bauschlicher and L. J. Allamandola, J. Phys. Chem. A, 2003, 107, 1486–1498 Search PubMed.
  115. D. M. Hudgins, C. W. Bauschlicher Jr and L. J. Allamandola, Astrophys. J., 2005, 632, 316–332 CrossRef.
  116. A. Vats, A. Pathak, T. Onaka, M. Buragohain, I. Sakon and I. Endo, Publ. Astron. Soc. Jpn., 2022, 74, 161–181 Search PubMed.
  117. R. Chown, A. Sidhu and E. Peeters, et al., Astron. Astrophys., 2024, 685, A75 CAS.
  118. I. Endo, I. Sakon, T. Onaka, Y. Kimura, S. Kimura, S. Wada and S. Kwok, Astrophys. J., 2021, 917, 103 CAS.
  119. D. Massa, K. D. Gordon and E. L. Fitzpatrick, Astrophys. J., 2022, 925, 19 CAS.
  120. G. C. Clayton, K. D. Gordon and F. Salama, et al., Astrophys. J., 2003, 592, 947–952 CAS.
  121. F. Salama, G. A. Galazutdinov and J. Krełowski, et al., Astrophys. J., 2011, 728, 154 Search PubMed.
  122. R. Gredel, Y. Carpentier and G. Rouillé, et al., Astron. Astrophys., 2011, 530, A26 Search PubMed.
  123. S. R. Pottasch, R. Surendiranath and J. Bernard-Salas, Astron. Astrophys., 2011, 531, A23 Search PubMed.
  124. P. Atherton, T. Hicks, N. Reay, G. Robinson, S. Worswick and J. Phillips, Astrophys. J., 1979, 232, 786–798 CrossRef.
  125. W. B. Latter, A. Dayal, J. H. Bieging, C. Meakin, J. L. Hora, D. M. Kelly and A. G. G. M. Tielens, Astrophys. J., 2000, 539, 783–796 CrossRef.
  126. J. Bernard-Salas, S. R. Pottasch, D. A. Beintema and P. R. Wesselius, Astron. Astrophys., 2001, 367, 949–963 CrossRef.
  127. C. Moutou, L. Verstraete, K. Sellgren and A. Leger, in The Universe as Seen by ISO, ed. P. Cox and M. Kessler, ESA SP-427, ESA Publications Division, Noordwijk, The Netherlands, 1998, pp. 727–730 Search PubMed.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.