Building Carbon Bridges on and between Fullerenes in Helium

Apr 4, 2016 - We report the observation of sequential encounters of fullerenes with C atoms in an extremely cold environment. Experiments were perform...
1 downloads 6 Views 2MB Size
Letter pubs.acs.org/JPCL

Building Carbon Bridges on and between Fullerenes in Helium Nanodroplets Serge A. Krasnokutski,† Martin Kuhn,‡ Alexander Kaiser,‡ Andreas Mauracher,‡ Michael Renzler,‡ Diethard K. Bohme,§ and Paul Scheier*,‡ †

Laboratory Astrophysics Group of the Max Planck Institute for Astronomy at the Friedrich Schiller University Jena, Helmholtzweg 3, D-07743 Jena, Germany ‡ Institut für Ionenphysik und Angewandte Physik, Universität Innsbruck, Technikerstr. 25, A-6020 Innsbruck, Austria § Department of Chemistry, York University, 4700 Keele Street, Toronto M3J 1P3, Ontario, Canada S Supporting Information *

ABSTRACT: We report the observation of sequential encounters of fullerenes with C atoms in an extremely cold environment. Experiments were performed with helium droplets at 0.37 K doped with C60 molecules and C atoms derived from a novel, pure source of C atoms. Very high-resolution mass spectra revealed the formation of carbenes of the type C60(C:)n with n up to 6. Bridge-type bonding of the C adatoms to form the known dumbbell C60CC60 also was observed. Density functional theory calculations were performed that elucidated the carbene character of the C60(C:)n species and their structures. Mass spectra taken in the presence of water impurities and in separate experiments with added H2 also revealed the formation of the adducts C60Cn(H2O)n and C60Cn(H2)n probably by H−OH and H−H bond insertion, respectively, and nonreactivity for the dumbell. So C adatoms that form carbenes C60(C:)n can endow pristine C60 with a higher chemical reactivity.

S

with one or more C atoms.23 C additions to C60 were also predicted to catalyze Stone−Wales transformations.24 There has been a recent breakthrough in pure C atom production. Low-energy carbon atoms now can be produced with a purity >99% from a thin-walled, sealed tantalum tube containing carbon.19 We have adopted this source to investigate the reactivity of C60 toward C atoms in very low-temperature He droplets and modified our apparatus that allows C60 molecules to be added into He droplets (ca. 106 He atoms and five C60 molecules on average per droplet) where they can aggregate and be cooled to the droplet temperature of 0.37 K.25 Carbon atoms are then added and allowed to react with the C60 aggregates before the droplets are exposed to ionizing or attaching electrons that permit reactants and products to be analyzed with a mass spectrometer. While low-energy (22 eV) first generate either He+ or He*− that then ionizes neutrals inside the droplet.26−29 To vary the C input into the droplet, the electric power used to heat the tantalum tube was varied from ca. 60 to 162 W. Mass spectra recorded for adduct formation with up to six carbon atoms to C60 are shown in Figure 1. The two stable isotopes of carbon, 12C (98.93%) and 13C (1.07%), lead to a characteristic series of mass peaks when combined with

ince the discovery of fullerene in vaporized graphite and its assignment by Kroto et al. and confirmation by Krätschmer et al.,1,2 the growth mechanisms of fullerene formation under energetic conditions have been a matter of debate.3−6 Simultaneously, several routes to add functional parts to fullerenes have been highly successful, including cyclopropanation, Diels−Alder reactions, and carbene additions.7−9 Derivatized fullerenes have promising applications in material and life sciences.10−12 They are used as electron acceptor molecules in solution-processed organic photovoltaic cells,13 they are considered for biological and medical applications,14,15 for DNA manipulation,16 and for biosensors,17 and they have been successfully deposited on surfaces as novel self-assembled monolayers.18 The possibility of activating reactive fullerenes with C atoms to form carbenes that can promote subsequent fullerene derivatization has driven us to investigate these species more closely. We performed laboratory experiments designed to explore encounters of fullerenes with C atoms derived from a clean source of C atoms.19 Such encounters may also occur in natural environments such as interstellar space where C60 and C60+ have been detected.20,21 Quite recent experiments under energetic conditions with laser-ablated carbon vapor and C60 under a flow of helium suggest that the carbon vapor can promote C exchange, C60 isomerization, and C atom incorporation, leading to fullerene closed network growth under these conditions.6,22 Also, under the hot conditions in fullerene plasma experiments, Shvartsburg et al. proposed the existence of ball-and-chain type fullerene dimers connected © XXXX American Chemical Society

Received: February 26, 2016 Accepted: April 4, 2016

1440

DOI: 10.1021/acs.jpclett.6b00462 J. Phys. Chem. Lett. 2016, 7, 1440−1445

Letter

The Journal of Physical Chemistry Letters

that two carbon atoms adsorb next to each other at the same hexagonal or pentagonal face. The adatoms should be immobile once they are adsorbed at these low temperatures. Computations indicate that a single carbon adatom chemically binds to the C60 cage at both distinguishable bond positions, that is, between two hexagonal faces [6,6] with 3.09 eV or one hexagonal and a pentagonal face [6,5] with 2.83 eV.32 Later density functional theory (DFT) calculations identified the [6,5] bridge as the most stable favored by 0.3 eV.33 Our own DFT calculations with results summarized in Table 1S in the Supporting Information agree with this latter stability order; we found that the [6,5] bond is slightly preferred to the [6,6] bond with a bond energy De (dissociation energy of the carbon adatom) of 2.82 eV on the [6,5] bond. This compares with 3.52 eV for the cation and 3.39 eV for the anion. C60C also was seen to have carbene properties from a natural bond orbital analysis.34 The resulting doubly occupied lone-pair orbital of the 61st carbon atom in singlet C61 is shown in Figure 2a. Further calculations provided an ionization energy of 6.37 eV for C61, significantly smaller than 7.08 eV for C60, and an electron affinity of 2.43 eV, larger than 2.25 eV for C60, and a fairly large dipole moment of 1.2765 D (with μx = 1.1855 D, μy = 0.0000 D, and μz = 0.4735 D). In the case of two carbon adatoms, the reference structure is shown in Figure 2b with two [6,6] C adatoms that are situated normal to each other with respect to the fullerene center. The adatoms can also adsorb very close to each other on neighboring bonds of a single carbon hexagon (Figure 2c); this structure is only 0.013 eV less stable than the reference structure, an amount that is not significant for DFT. We also considered other structures. A high-energy gain of 3.97 eV compared with the structure shown in Figure 2b can be reached by adsorbing the two carbon atoms in a bent chain on a [6,6] position, as shown in Figure 2d. Adsorption of two carbon atoms at opposite bonds of a pentagon face leads without barrier to a very stable (−5.3 eV) C62 cage structure with two heptagons and a tetragon (almost square, Figure 2e). In this noncarbene cage all carbon atoms are bound to three neighbors, and lone pairs are absent in the natural bond orbital analysis. For C63 (C64) we find that adsorption of a C−C−C (C−C− C−C) chain on a [6,6] position is energetically favorable by −7.13 eV (−10.83 eV) compared with distributed single adatoms. The C−C−C structure for C63 is shown in Figure 2f and that of the distributed single adatom structure for C64 is shown in Figure 2g. The perfectly symmetric structure for C66 with six single carbon adatoms at [6,6] positions (Figure 2h) is energetically unfavorable by only 0.038 eV compared with 6 C adatoms at [6,5] positions, as shown in Figure 2i. We did not calculate chain structures for more than four adatoms. Adding an additional C atom to the perfectly symmetric C66 gives C67. The dissociation energies for distributed adatoms from C61 to C66 lie in a range of 2.49 to 2.82 eV for one to six carbons. For seven adatoms the dissociation energy decreases to 1.94 eV. In the latter case the C60 cage distorts more. The typical bond length of the [6,6] bond, where C is adsorbed, is 1.582 Å for C66. For two of the seven C adatoms in C67, this bond length is stretched to 1.618 Å. In summary, the computations predict the existence of a variety of different stable isomers for which fullerene cage growth6,22 is energetically favorable compared with C−C−··· chain attachment, which is, in turn, favorable to the wellseparated addition of up to at least seven single C atoms. The

Figure 1. Mass spectra recorded for C60Cn+ (bottom) and C60Cn− ions (top) for He droplets (nozzle temperature 9.5 and 9.7 K, respectively) containing C60 and 13C atoms. The ionizing energies are 70 and 22 eV, the C60 oven temperatures were 292 and 288 °C, and the power settings for the 13C source were 158 and 162 W, respectively. Note the presence of water impurities. (See the text.) The inset shows enlarged the range of the dication (C60)2C2+. The thin line shows the corresponding monocation plotted with the mass scale divided by two.

molecules or clusters (binomial statistics). The peak pattern of the mass spectrum of the negatively charged ions from m/z = 720 to 725 perfectly matches the calculated isotopic pattern of C60, whereas the subsequent groups of mass peaks assigned to C6013Cn− (n = 1 to 6) exhibit enhanced intensity for the fifth isotopologue 12C5513Cn+5, which can be attributed to water, replacing one 13C (12C6013Cn−1H2O). Remarkably, the observed mass distribution is quite similar for positive and negative ions despite the substantial differences in the mode of ionization. Positively charged ions are predominantly formed via charge transfer from an initially formed He+ ion, which approaches the C60Cn via resonant holehopping,30 whereas the dominant process for anion formation at 22 eV is due to He*−, which is resonantly formed at this electron energy and attracted by the polarizable C60Cn.31 The electronic energy of He+ and He*− is almost the same, and thus similarity in the spectra is expected if the C atoms attach to C60 in neutral reactions first and the products are subsequently ionized. The occurrence of dissociative ionization accompanied by loss of one or more C atoms from higher members of the series could not be excluded. Similarity in the positive and negative ion spectra is expected if the C attaches to C60 in neutral reactions first, and the products are subsequently ionized. The structures of the C60(C)n molecules that are formed are not known from these experiments; sequential attachment of C atoms can proceed either only on the surface of the C60 by bridge addition (Figure 2a−c,g−i) or in the formation of a C chain (Figure 2d,f) extending away from the surface, as in C60CCCC CC:, or in an intermediate fashion. Fullerene growth (Figure 2e) by C2 insertion can be largely excluded because it is unlikely 1441

DOI: 10.1021/acs.jpclett.6b00462 J. Phys. Chem. Lett. 2016, 7, 1440−1445

Letter

The Journal of Physical Chemistry Letters

Figure 2. (a) Lone pair orbital at the C-adatom in singlet C61 [6,5] from a natural bond orbital analysis.34 Structures computed with DFT calculations for C62 (b−e), C63 (f), C64 (g), and C66 (h,i).

well-separated addition is proposed to occur in our experiments in a cold environment due to rapid cooling accompanied by trapping in local minima on the potential energy surface,35 whereas cage growth is likely favored at higher temperatures as in the experiments of Dunk et al.6,22 The carbene character of bridge-type C60(C:)n is clearly manifested by the natural bond orbital analysis that is displayed in Figure 2a for C60(C:)1. Carbene character can be inferred for all experimentally observed C adducts C60(C)n from the observed water impurity adduct C60Cn(H2O)n± ion peaks in Figure 1. According to the presence of these peaks, C60(C)n molecules with n > 0 pick up H2O, as would be expected from a carbene reaction, presumably by H−OH bond insertion, while C60 itself adds H2O only by physisorption. Furthermore, the results of our separate experiments with added H2 dopant (see Figure 3 and Figures 1S and 3S in the Supporting Information) indicate H2 additions for each C60(C)n with n = 1,2,3,4,5, presumably by H−H bond insertion. These results with H2O and H2 strongly indicate sequential surface addition of at least the first five carbon atoms to form carbenes of the type C60(C:)n with n = 1−5 instead of energetically favorable C chains such as C60C CCC:, for which always only one H2O or H2 would add chemically (H2 at T = 0.37 K does not add chemically to C60 because of a large barrier predicted to be 1.69 eV).36 This result shows that C-atom additions can transform a chemically inert

Figure 3. Mass spectra recorded for 12C6013CnHn+ ions for droplets containing C60, 13C, and H2. The ionizing energy is 70 eV, the C60 oven temperature is 292 °C, the power setting for the 13C source was 162 W. The ambient pressure of H2 was 0.5 mPa.

(against reaction with H2 and H2O) fullerene C60 into chemically reactive carbenes C60(C:)n. Carbenes are known from terrestrial chemistry to be among the most versatile and synthetically useful reactive intermediates. Carbene additions to fullerenes have already attracted considerable attention.8 1442

DOI: 10.1021/acs.jpclett.6b00462 J. Phys. Chem. Lett. 2016, 7, 1440−1445

Letter

The Journal of Physical Chemistry Letters

mixture of C60 and C60CBr2 or the neutron irradiation of pure C60.40 Photochemical and photophysical properties of the three bridged isomers [6,6]−[6,6], [5,6]−[6,6], and [5,6]−[5,6] also have been reported.41,42 Under energetic conditions C and C2 addition to fullerenes leads to closed network growth as demonstrated by Kroto et al.,6,22 whereas our observed surface C additions at low temperature enhance the chemical activity of the surface of C60 by carbene formation. The latter may pave the way toward a generation of a new class of fullerene derivatives. Reactive C60 carbene species may also be formed in interstellar and circumstellar environments containing C and C60, where they could add H2 and H2O, as well as C60, to form derivatized fullerene molecules such as C60(CH(H))n, C60(CH(OH))n, C60(CCO)n, and even C60(CC60)n.

Carbenes should also play a similar role in interstellar chemistry.37 Our experiments demonstrate the reactivity of C60(C:)n toward H2 and H2O. We expect similar reactivities toward many other molecules as well as other carbenes such as :CO, :CS, :Cn:, :CnO, and :CNH, for example.37 Our other recorded mass spectra show evidence of the presence of C adduct ions for clusters of fullerenes of the type (C60)n(C)m± with n = 1 to 9 and m = 1−5 including the dumbbell. Excerpts are shown in Figure 4 for (C60)2(C)m± with



EXPERIMENTAL SECTION

He droplets were formed by expanding helium (Messer purity 99.9999%) from a stagnation pressure of 2.4 MPa through a 5 μm nozzle, cooled by a closed-cycle refrigerator (Sumitomo Heavy Industries LTD, model RDK-415D) to ∼9.5 K, into vacuum. The average size of the He droplets is ∼106 atoms.43 The resulting supersonic beam was skimmed by a 0.8 mm conical skimmer located 1 cm downstream from the nozzle and flew through two 10 cm long, differentially pumped pick-up regions. Small amounts of C60 (MER, purity 99.9%) were vaporized at an oven temperature between 292 and 308 °C and picked up by the He droplets with almost unity efficiency upon collisions. In the second differentially pumped pickup chamber atomic (purity >99%) vapor of 13C emitted from a heated sealed tantalum tube containing carbon is picked up by the He droplets.19 H2 (Messer, purity 99.999%) was introduced via an electronically controlled leak valve into the first pickup chamber. Water was present in the machine as a trace impurity from the residual gas. The doped helium droplets were ionized by collisions with electrons at energies ranging from 0 to 100 eV. The ions were accelerated to 40 eV into the extraction region of a commercial orthogonal time-of-flight mass spectrometer equipped with a reflectron (Tofwerk AG, model HTOF). For anions the mass resolution was m/Δm = 2500 and for cations 3500. The base pressure in the mass spectrometer was 10−5 Pa. The ions were detected by a microchannel plate and recorded via a time to digital converter. Additional experimental details have been described elsewhere.44 Special home-built software was utilized to deduce ion intensities from the mass spectra, taking into account all possible isotopologues, as shown in Figure 2S in the Supporting Information.45 Energetics and geometries of neutral C60 with carbon adatoms were calculated by means of DFT. All structures were fully optimized to account for deformations of the fullerene due to the presence of the carbon adatoms. We used the B3LYP46,47 hybrid density functional and a standard 631g(d)48 basis set as implemented in the Gaussian 09 program.49 Energies were corrected for zero-point vibrations, and all reported structures are true local minima. The accuracy of isomerization energies and heats of formation of B3LYP with the 6-31g(d) basis set was estimated with 0.13 eV for a large set of neutral, closed-shell organic compounds containing C, H, N, and O.50

Figure 4. Upper part: Mass spectra recorded for C60Cn+/(C60)2Cn+ ions (left) and C60Cn−/(C60)2Cn− ions (right) for He droplets (nozzle temperature 9.5 and 9.6 K, respectively) containing C60 and 13C atoms. The ionizing energies are 90 and 22 eV, the C60 oven temperatures were 308 and 300 °C, and the power settings for the 13C source were 141 and 150 W, respectively. Note the presence of water impurities (see text). Lower part: Mass spectra recorded for (C60)213CnHn+ ions for droplets containing C60, 13C, and H2. The ionizing energy is 70 eV, the C60 oven temperatures is 292 °C, and the power setting for the 13C source was 162 W. The ambient pressure of H2 was 0.5 mPa.

m = 0−3. The positive and negative ion spectra both show a remarkable predominance of ionized (C60)2C, suggesting the occurrence of neutral chemistry before ionization and a special stability for neutral (C60)2C, presumably the dumbbell C60 CC60. The concomitant observation of the dication (C60)2C2+ (see insert in Figure 1) as well as experiments with additional H2 dopant completely support this assignment. The dication appears to be stable against Coulomb explosion; presumably the two charges are located apart on the two fullerenes. H2 is seen in Figure 4 to add to (C60)2C2, in which the second C atom is attached to one of the fullerenes and so is available to insert into the H−H bond but not (C60)2C, in which the C atom is not accessible in the dumbbell configuration. The stability of the bridged dumbbell structure C60CC60 is well known from terrestrial chemistry.38 C121 has been previously detected in the laboratory by mass spectrometry39 of the products obtained in the synthesis of C61Br2 from C60 and C6H5HgCBr3 and has been isolated using thermolysis of a 1443

DOI: 10.1021/acs.jpclett.6b00462 J. Phys. Chem. Lett. 2016, 7, 1440−1445

Letter

The Journal of Physical Chemistry Letters



(16) Nakamura, E.; Isobe, H.; Tomita, N.; Sawamura, M.; Jinno, S.; Okayama, H. Functionalized fullerene as an artificial vector for transfection. Angew. Chem., Int. Ed. 2000, 39, 4254−4257. (17) Afreen, S.; Muthoosamy, K.; Manickam, S.; Hashim, U. Functionalized fullerene (C60) as a potential nanomediator in the fabrication of highly sensitive biosensors. Biosens. Bioelectron. 2015, 63, 354−364. (18) del Carmen Gimenez-Lopez, M.; Räisänen, M. T.; Chamberlain, T. W.; Weber, U.; Lebedeva, M.; Rance, G. A.; Briggs, G. A. D.; Pettifor, D.; Burlakov, V.; Buck, M.; et al. Functionalized Fullerenes in Self-Assembled Monolayers. Langmuir 2011, 27, 10977−10985. (19) Krasnokutski, S. A.; Huisken, F. A simple and clean source of low-energy atomic carbon. Appl. Phys. Lett. 2014, 105, 113506. (20) Sellgren, K.; Werner, M. W.; Ingalls, J. G.; Smith, J. D. T.; Carleton, T. M.; Joblin, C. C60 in Reflection Nebulae. Astrophys. J., Lett. 2010, 722, L54−L57. (21) Campbell, E. K.; Holz, M.; Gerlich, D.; Maier, J. P. Laboratory confirmation of C60+ as the carrier of two diffuse interstellar bands. Nature 2015, 523, 322−323. (22) Dunk, P. W.; Adjizian, J.-J.; Kaiser, N. K.; Quinn, J. P.; Blakney, G. T.; Ewels, C. P.; Marshall, A. G.; Kroto, H. W. Metallofullerene and fullerene formation from condensing carbon gas under conditions of stellar outflows and implication to stardust. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 18081−18086. (23) Shvartsburg, A. A.; Hudgins, R. R.; Gutierrez, R.; Jungnickel, G.; Frauenheim, T.; Jackson, K. A.; Jarrold, M. F. Ball-and-Chain Dimers from a Hot Fullerene Plasma. J. Phys. Chem. A 1999, 103, 5275−5284. (24) Lee, I.-H.; Jun, S.; Kim, H.; Kim, S. Y.; Lee, Y. Adatom-assisted structural transformations of fullerenes. Appl. Phys. Lett. 2006, 88, 011913. (25) Grebenev, S.; Toennies, J. P.; Vilesov, A. F. Superfluidity within a small helium-4 cluster: The microscopic Andronikashvili experiment. Science 1998, 279, 2083−2086. (26) Scheidemann, A.; Schilling, B.; Toennies, J. P. Anomalies in the Reactions of He+ with SF6 Embedded in Large Helium-4 Clusters. J. Phys. Chem. 1993, 97, 2128−2138. (27) Mauracher, A.; Daxner, M.; Huber, S. E.; Postler, J.; Renzler, M.; Denifl, S.; Scheier, P.; Ellis, A. M. The interaction of He− with fullerenes. J. Chem. Phys. 2015, 142, 104306. (28) Mauracher, A.; Daxner, M.; Huber, S. E.; Postler, J.; Renzler, M.; Denifl, S.; Scheier, P.; Ellis, A. M. Formation of Dianions in Helium Nanodroplets. Angew. Chem., Int. Ed. 2014, 53, 13794−13797. (29) Daxner, M.; Denifl, S.; Scheier, P.; Ellis, A. M. Electron-Driven Self-Assembly of Salt Nanocrystals in Liquid Helium. Angew. Chem., Int. Ed. 2014, 53, 13528−13531. (30) Ellis, A. M.; Yang, S. Model for the charge-transfer probability in helium nanodroplets following electron-impact ionization. Phys. Rev. A: At., Mol., Opt. Phys. 2007, 76, 032714. (31) Mauracher, A.; Daxner, M.; Postler, J.; Huber, S. E.; Denifl, S.; Scheier, P.; Toennies, J. P. Detection of Negative Charge Carriers in Superfluid Helium Droplets: The Metastable Anions He*− and He2*−. J. Phys. Chem. Lett. 2014, 5, 2444−2449. (32) Eggen, B. R.; Heggie, M. I.; Jungnickel, G.; Latham, C. D.; Jones, R.; Briddon, P. R. Autocatalysis During Fullerene Growth. Science 1996, 272, 87−90. (33) Tsetseris, L.; Pantelides, S. T. Intermolecular bridges and carrier traps in defective C60 crystals. Phys. Rev. B: Condens. Matter Mater. Phys. 2011, 84, 195502. (34) Glendening, E. D.; Reed, A. E.; Carpenter, J. E.; Weinhold, F. NBO, version 3.1; Theoretical Chemistry Institute, University of Wisconsin: Madison, WI, 1998. (35) Nauta, K.; Miller, R. E. Formation of Cyclic Water Hexamer in Liquid Helium: The Smallest Piece of Ice. Science 2000, 287, 293−295. (36) Kaiser, A.; Leidlmair, C.; Bartl, P.; Zöttl, S.; Denifl, S.; Mauracher, A.; Probst, M.; Scheier, P.; Echt, O. Adsorption of hydrogen on neutral and charged fullerene: Experiment and theory. J. Chem. Phys. 2013, 138, 074311. (37) Bohme, D. K. Ionic Origins of Carbenes in Space. Nature 1986, 319, 473−474.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpclett.6b00462. Table 1S: Results of density functional theory calculations. Figure 1S: Comparison of the spectra with and without H2. Figure 2S: Isotope fitting procedure and the composition of the peaks. Figure 3S: H2 addition in the presence and absence of atomic carbon. (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the Austrian Science Fund FWF (P26635, I978). S.A.K. gratefully acknowledges support for a short-term scientific mission by the COST Action CM1401 “Our Astro-Chemical History”. Support also was provided by the Austrian Ministry of Science BMWF as part of the UniInfrastrukturprogramm of the Focal Point Scientific Computing at the University of Innsbruck. We thank Harold Linnartz for stimulating discussions.



REFERENCES

(1) Kroto, H. W.; Heath, J. R.; O’Brien, S. C.; Curl, R. F.; Smalley, R. E. C60: Buckminsterfullerene. Nature 1985, 318, 162−163. (2) Krätschmer, W.; Lamb, L. D.; Fostiropoulos, K.; Huffman, D. R. Solid C60: a new form of carbon. Nature 1990, 347, 354−358. (3) Goroff, N. S. Mechanism of Fullerene Formation. Acc. Chem. Res. 1996, 29, 77−83. (4) James, R. H. Synthesis of C60 from Small Carbon Clusters. In Fullerenes; American Chemical Society, 1992; Vol. 481, pp 1−23. (5) von Helden, G.; Gotts, N. G.; Bowers, M. T. Experimental evidence for the formation of fullerenes by collisional heating of carbon rings in the gas phase. Nature 1993, 363, 60−63. (6) Dunk, P. W.; Kaiser, N. K.; Hendrickson, C. L.; Quinn, J. P.; Ewels, C. P.; Nakanishi, Y.; Sasaki, Y.; Shinohara, H.; Marshall, A. G.; Kroto, H. W. Closed network growth of fullerenes. Nat. Commun. 2012, 3, 855. (7) Taylor, R. The Chemistry of Fullerenes; World Scientific: Singapore, 2003. (8) Yamada, M.; Akasaka, T.; Nagase, S. Carbene Additions to Fullerenes. Chem. Rev. 2013, 113, 7209−7264. (9) Bingel, C. Cyclopropanierung von Fullerenen. Chem. Ber. 1993, 126, 1957−1959. (10) Yan, W.; Seifermann, S. M.; Pierrat, P.; Brase, S. Synthesis of highly functionalized C60 fullerene derivatives and their applications in material and life sciences. Org. Biomol. Chem. 2015, 13, 25−54. (11) Nakamura, E.; Isobe, H. Functionalized Fullerenes in Water. The First 10 Years of Their Chemistry, Biology, and Nanoscience. Acc. Chem. Res. 2003, 36, 807−815. (12) Diederich, F.; Gomez-Lopez, M. Supramolecular fullerene chemistry. Chem. Soc. Rev. 1999, 28, 263−277. (13) Clarke, T. M.; Durrant, J. R. Charge Photogeneration in Organic Solar Cells. Chem. Rev. 2010, 110, 6736−6767. (14) Bosi, S.; Da Ros, T.; Spalluto, G.; Prato, M. Fullerene derivatives: an attractive tool for biological applications. Eur. J. Med. Chem. 2003, 38, 913−923. (15) Huang, Y. Y.; Sharma, S. K.; Yin, R.; Agrawal, T.; Chiang, L. Y.; Hamblin, M. R. Functionalized Fullerenes in Photodynamic Therapy. J. Biomed. Nanotechnol. 2014, 10, 1918−1936. 1444

DOI: 10.1021/acs.jpclett.6b00462 J. Phys. Chem. Lett. 2016, 7, 1440−1445

Letter

The Journal of Physical Chemistry Letters (38) Dragoe, N.; Tanibayashi, S.; Nakahara, K.; Nakao, S.; Shimotani, H.; Xiao, L.; Kitazawa, K.; Dragoe, N.; Kitazawa, K.; Achiba, Y.; et al. Carbon allotropes of dumbbell structure: C121 and C122. Chem. Commun. 1999, 85−86. (39) Osterodt, J.; Vogtle, F. C 61 Br 2 : a new synthesis of dibromomethanofullerene and mass spectrometric evidence of the carbon allotropes C121 and C122. Chem. Commun. 1996, 547−548. (40) Dragoe, N.; Shimotani, H.; Wang, J.; Iwaya, M.; de BettencourtDias, A.; Balch, A. L.; Kitazawa, K. First Unsymmetrical Bisfullerene, C121: Evidence for the Presence of Both Homofullerene and Methanofullerene Cages in One Molecule. J. Am. Chem. Soc. 2001, 123, 1294−1301. (41) Ren, T. X.; Sun, B. Y.; Chen, Z. L.; Qu, L.; Yuan, H.; Gao, X. F.; Wang, S. K.; He, R.; Zhao, F.; Zhao, Y. L.; et al. Photochemical and photophysical properties of three carbon-bridged fullerene dimers: C121 (I, II, III). J. Phys. Chem. B 2007, 111, 6344−6348. (42) Fujitsuka, M.; Ito, O.; Dragoe, N.; Ito, S.; Shimotani, H.; Takagi, H.; Kitazawa, K. Photophysical and Photochemical Processes of an Unsymmetrical Fullerene Dimer, C121. J. Phys. Chem. B 2002, 106, 8562−8568. (43) Gomez, L. F.; Loginov, E.; Sliter, R.; Vilesov, A. F. Sizes of large He droplets. J. Chem. Phys. 2011, 135, 154201. (44) An der Lan, L.; Bartl, P.; Leidlmair, C.; Schöbel, H.; Jochum, R.; Denifl, S.; Märk, T. D.; Ellis, A. M.; Scheier, P. The submersion of sodium clusters in helium nanodroplets: Identification of the surface → interior transition. J. Chem. Phys. 2011, 135, 044309. (45) Ralser, S.; Postler, J.; Harnisch, M.; Ellis, A. M.; Scheier, P. Extracting cluster distributions from mass spectra: IsotopeFit. Int. J. Mass Spectrom. 2015, 379, 194−199. (46) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648. (47) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (48) Ditchfield, R.; Hehre, W. J.; Pople, J. A. Self-Consistent Molecular-Orbital Methods. IX. An Extended Gaussian-Type Basis for Molecular-Orbital Studies of Organic Molecules. J. Chem. Phys. 1971, 54, 724−728. (49) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A., et al. Gaussian 09, revision D.01; Gaussian, Inc.: Wallingford, CT, 2009. (50) Tirado-Rives, J.; Jorgensen, W. L. Performance of B3LYP Density Functional Methods for a Large Set of Organic Molecules. J. Chem. Theory Comput. 2008, 4, 297−306.

1445

DOI: 10.1021/acs.jpclett.6b00462 J. Phys. Chem. Lett. 2016, 7, 1440−1445