Redispersion and Self-Assembly of C60 Fullerene in Water and

May 26, 2017 - ... open access article published under an ACS AuthorChoice License, ... with the formation of two different types of self-assembled st...
0 downloads 4 Views 7MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article http://pubs.acs.org/journal/acsodf

Redispersion and Self-Assembly of C60 Fullerene in Water and Toluene Antonio Cid,*,†,‡ Ó scar A. Moldes,‡ Mário S. Diniz,† Benito Rodríguez-González,§ and Juan C. Mejuto‡ †

UCIBIO, REQUIMTE, Departamento de Química, Faculdade de Ciências e Tecnología, Universidade NOVA de Lisboa, 2829-516 Caparica, Portugal ‡ Departamento de Química-Física, Facultade de Ciencias de Ourense, Universidade de Vigo, Campus de As Lagoas S/N, 32004 Ourense, Spain § Departamento de Química Física, Universidade de Vigo, 36310 Vigo, Spain ABSTRACT: This work aims at assessing the influence of two different solvents, bidistilled water and toluene, on dispersions of carbon-based engineered nanomaterials, namely, fullerenes, and their self-assembly behavior. The obtained self-assembled carbon-based materials were characterized using UV−vis spectrophotometry and transmission electron microscopy techniques. The results obtained were unexpected when toluene was used for dispersing fullerene C60, with the formation of two different types of self-assembled structures: fullerene C60 nanowhiskers (FNWs) and a type of quasispherical nanostructure. The FNWs ranged between 1 and 6 μm in length, whereas the quasispherical fullerene C60 nanoaggregates ranged between 10 and 50 nm in diameter. Aggregates obtained in toluene showed a well-formed crystal structure. When using water, the obtained aggregates were amorphous and showed a no well-defined shape. Their sizes ranged between 20 and 40 nm for nanosized structures and between 0.4 and 4.8 μm for micron-sized self-aggregates.

1. INTRODUCTION Engineered carbon-based nanomaterials are being used in a large number of biomedical applications, such as drug-delivery systems,1 cancer therapy,2 and theranostics, as well as in electronic devices. It is well known that buckminsterfullerene or fullerene C60 is a truncated icosahedron (Ih) formed by sp2hybridized atoms of carbon. Its peculiar external shape is due to a combination of 20 hexagons and 12 pentagons. Allotropic forms of carbon show significant properties, such as optoelectrical, optoelectronic, and redox properties; effective n-doping of organic semiconductors; and high conductivity,3 which make of them one of the most promising elements in chemistry. Self-assembled nanostructures show electron-transfer improvements when compared with planar electron acceptors. This property allows the development of organic electron devices (e.g., field-effect transistors and solar cells)4,5 and optoelectronic devices. Different syntheses approaches of self-assembled fullerene derivatives have been reported, such as supramolecular approaches,6,7 controlled precipitation,8−12 template methods,13 and photoassisted growth.11 In this article, we report on a method for obtaining of fullerene C60 self-assembled nanostructures that avoids liquid−liquid interface precipitation (LLIP) and several other issues associated with previously reported preparation routes.14 Instead of using isopropyl alcohol as a solvent in the LLIP method9 or modifications for faster preparation, as reported by Jin et al.,15 we chose a simple © 2017 American Chemical Society

method based on self-assembly. This new method yielded a bimodal population of carbon-based materials and finally enabled us to assess the aggregation behavior in two solvents. Our aim in this work was to explore the self-assembly behavior as a function of solvent redispersion nature (solubility) and the structural implications in the final products. Supramolecular assemblies are defined as chemical structures organized by means of noncovalent bonds, for instance, hydrogen bonds,16,17 π−π interactions,18−20 van der Waals interactions,21,22 and coordination bonds.23,24 Our supramolecular self-aggregates involve quasispherical aggregates of fullerene C60 nanoparticles (FC60 NPs), with a diameter ranging from 10 to 50 nm; on the other hand, we also found fullerene C60 nanowhiskers (FNWs),a with a length ranging from 1 to 6 μm.

2. RESULTS AND DISCUSSION Typical transmission electron microscopy (TEM) low-magnification images of the samples obtained by redispersing C60 in bidistilled water are shown in Figure 1. It appears that in this dispersion medium C60 fullerenes form amorphous aggregates. Additionally, spontaneous production of a slight brownish Received: January 13, 2017 Accepted: May 10, 2017 Published: May 26, 2017 2368

DOI: 10.1021/acsomega.7b00049 ACS Omega 2017, 2, 2368−2373

ACS Omega

Article

Figure 1. TEM images of fullerene C60 redispersed in bidistilled water. Scale bar: (a) 1 μm, (b) 100 nm, and (c) 50 nm.

Figure 2. TEM images of fullerene C60 redispersed in toluene forming (a) FC60 NPs and (b) FNWs, and (c) selected area electron diffraction (SAED) pattern obtained from a single nanowhisker.

Figure 3. UV−vis normalized absorbance spectra of fullerene C60 in toluene at several concentrations (10−50 μL aliquots in 2 mL of toluene). Inset: Enhanced image of the peak found at 407 nm. Naked-eye observation of C60 redispersed in toluene.

yellow coloration was observed in the aqueous solution of fullerene C60. A large variety of particles sizes and shapes were identified. The particles show irregular edges and differences in contrast due to different thicknesses. Despite the small size of the fullerene seeds (1−2 nm in diameter), a mixture of large

(micron-sized) and small (20−40 nm) structures was obtained. The internal structure of the aggregates was investigated by electron diffraction, and the obtained data showed a diffraction pattern that corresponds to that of an amorphous structure. This result clearly points toward aggregates that have a disordered internal structure. 2369

DOI: 10.1021/acsomega.7b00049 ACS Omega 2017, 2, 2368−2373

ACS Omega

Article

Figure 4. (a) HRTEM image showing the core−shell structure of the FNWs. (b) HRTEM lattice image showing the structure of the core obtained from the area depicted in (a). (c) FT of (b) with the indexes of the spots calculated on the basis of the fcc crystalline structure. (d, e) Growth axes of the core and shell, respectively. (f) FT obtained from (e), indexed on the basis of the body-centered tetragonal (bct) crystalline structure.

observations, consistent with the findings of previous works.32 Aggregation of the fullerene C60 nanoparticles continued, which finally rendered a nanowhisker structure; upon storing the C60 nanoparticles for 2 weeks at 2 °C, we obtained FNWs as large as 60.13 μm, as measured by electron microscopy. HRTEM images showing a core−shell structure of FNWs are shown in Figure 4a. In this image, the growth axis of the core (Figure 4b) was identified to be in the ⟨011⟩ crystalline direction; the corresponding Fourier-transform (FT) displaying the [011] zone axis of the cubic fcc crystalline structure of the FNW core is shown in Figure 4c. However, analyses of the FT obtained from the shell area confirmed the [010] zone axis, which established the ⟨200⟩ crystalline direction as the growth axis of the shell parallel to the larger axis of the nanowhiskers and confirmed the presence of a bct tetragonal structure on the shell. Coincidentally, as was previously reported by Miyazawa et al.,33 the same growth axis, ⟨200⟩, was determined previously in the bct tetragonal structure of the FNWs. In the present case, HRTEM images also show clear differences between the inner structure (core cubic crystalline structure) and shell (tetragonal crystalline structure). HRTEM images and FT also confirmed a core@ shell structure formed by two different crystalline structures, with planes and growth axes different in both the core and shell. These differences were also previously observed in nanowhiskers by Miyazawa34 and have been reported in the LLIP method of FNWs synthesis. We also found a difference of 0.5% dhkl in the Miller indices when compared to previously reported data on pristine fullerene C60 crystals.35 The influence of solvent36 on the redispersion and coassembly of fullerene37 has been previously reported, and its significance has also been underlined,38 but here, we intend to stress on the influence of electronic interactions on the assembly process. EELS allowed us to determine the nature of C−C bonds present in the internal structure of C60 fullerenes found in both FC60 NPs and FNWs. As this bimodal population was formed as consequence two different self-assembly pathways, it was important to be sure that these two different

Quite a different rearrangement was observed when C60 fullerenes were redispersed in toluene. This different behavior was confirmed and characterized using UV−vis spectroscopy, TEM, high-resolution electron microscopy (HRTEM), and electron energy-loss spectroscopy (EELS). In this case, the aggregates of fullerene C60 showed a crystalline face-centered cubic (fcc)27 structure, forming two different types of aggregates, namely, FC60 NPs and FNWs. Figure 2 shows TEM images of the self-assembled nano- and micro-structures obtained on redispersing fullerene C60 in toluene. Clearly, the self-assembled nanostructures, including FC60 NPs (20−40 nm) and FNWs (0.4−4.8 μm), have a bimodal shape and size. FNWs have often been obtained using the LLIP method; from now on, we will report on the different behaviors as a function of solvent dispersion nature.28 UV−vis absorbance spectra of FNWs at several concentrations in toluene are shown in Figure 3. As can be seen, we found a small peak near λAbs = 407 nm, but the maximum absorbance intensity was observed at about 335−336 nm; this is different from previously reported data on pure fullerene C60 redispersed in several solvents, like hexane, wherein a maximum λAbs near 329 nm was reported.29 The two peaks found in the UV−vis absorption spectra were assigned to (a) the effects exerted by the environment on fullerene C60 molecules (λAbs = 335−336 nm) and (b) the vibronic structure and effects of the surrounding environment on fullerene C60 molecules (λAbs = 407 nm).30 The UV−vis spectra obtained with water as the solvent showed maximum peaks at 265 and 345 nm, as previously reported by Scharff et al.31 We associated this small shift in the spectra, when compared to those with toluene, to the hydration of fullerene C60 by water molecules. The heating caused by sonication in small areas of the initial fullerene C60− toluene solution can induce a change in the small amount of fullerene C60 seeds to a point at which the polymerization process starts between neighboring molecules. Following sonication at room temperature, enhanced by sunlight (UV− vis radiation), the formation of crystalline FNWs and FC60 NPs was confirmed by optical and transmission microscopy 2370

DOI: 10.1021/acsomega.7b00049 ACS Omega 2017, 2, 2368−2373

ACS Omega

Article

Figure 5. Electron energy-loss spectra of self-aggregated FC60 NPs (black solid circles) and FNWs (blue solid circles).

C60 fullerene (0.5−40 mg mL−1/6.94 × 10−4−5.55 × 10−2 M) were redispersed as saturated solutions of engineered nanomaterials in bidistilled water and toluene. We did not perform any further treatment; no filtration of the stock solutions was carried out; and we did not use any other polymer or surfactant to stabilize the suspension or obtain pristine carbon nanomaterials. Using a Branson 3510 ultrasonic bath, frequency 40 kHz, redispersion of the C60 fullerene was carried out over 5 min at 298 K, and then, the solutions were stored at room temperature until morphological observations were carried out. First, all fullerene C60 samples in water and toluene were exposed to ultrasonication at room temperature (298 K) under sunlight (it is known that the crystalline structures of fullerene C60 trend to grow under UV−vis radiation). These samples were stored at room temperature until morphological measurements by HRTEM were carried out (ca. 1 h). Then, the samples were stored at 2 °C and light-protected for 3 weeks until morphological characterization was carried out again. Samples for UV−vis absorbance spectroscopy were freshly prepared using the same ultrasound bath. Characterization was carried out in triplicate for every measurement to confer reproducibility of the experiments. TEM analyses were performed using a JEOL JEM1010 TEM operating at an accelerating voltage of 100 kV, whereas HRTEM was performed using a JEOL JEM2010F HRTEM operating at 200 kV. EEL spectra were recorded following standard and well-known methods. In brief, measurements were carried out at an operation voltage of 200 kV in the STEM mode using a Gatan GIF Quantum spectrometer, with an energy resolution of 1.75 eV (FWHM zero loss peak), 0.5 eV/channel energy dispersion, and an EELS collection semiangle of 16 mrad. To avoid the contribution of the grid carbon foil, the EELS spectra were collected from areas of the sample situated over a hole. The background EEL spectra were subtracted using standard routines of Digital Micrograph. Samples for TEM and HRTEM were prepared by dropping and evaporating the obtained colloid on top of a carbon-coated copper grid. The authors emphasize that the sonication time was selected on the basis of macroscopic observations. The observation of a change in the turbidity and color of the solution after sonication for 5 min suggested a change in the optical properties of the solution; this was the reason behind exploring

types of particles had the same electronic configuration, as evidenced by the indexing of SAED. Results are shown in Figure 5, and they suggest that the bonds established between C60 fullerenes as building units are responsible for the formation of both aggregates. Basically, the presence of shoulders with the same shape in EELS confirmed the existence of σ and π bonds in the internal structure of the FC60 NPs and in the FNWs.39 In view of this, it is also clear that the structure of individual fullerene molecules remained unchanged within a nanowhisker.

3. CONCLUSIONS In this work, we studied the influence of solvent in a simple one-step method to redisperse the fullerene C60 avoiding the use of a polymeric stabilizer, by means of self-assembly of these carbon-based nanomaterials. In toluene, we obtained two different types of crystalline structures: FNWs, having a length of 1−6 μm, and FC60 NPs, with a diameter of 10−50 nm. In water, we obtained two amorphous populations, one with a size of 20−40 nm and the second with a size of 0.4−4.8 μm, which are larger than the former. Both optical and structural characterizations confirmed the formation of these selfassembled structures. The formation of these shape allotropes of carbon can be due to steric and electrostatic rearrangements of the fullerene C60 involved in the redispersion as a function of solvent nature. Previous work of Hughes et al. underlined that C60 shows very low solubility in water, whereas the solubility of C60 in toluene is 2.8 mg/mL, which is an important factor to take into account in our experiments and in further studies.40 4. EXPERIMENTAL SECTION 4.1. Reagents. C60 fullerene was purchased from SigmaAldrich at the maximum commercially available purity of 99.5% (Germany),41 with no further treatment. Toluene was commercially obtained from Panreac (Spain) (99.9%). Bidistilled water showing 0.10−0.50 μS cm−1 conductance was used for synthesis. 4.2. Carbon-Based Nanomaterial Redispersion: Synthesis of FNWs. In a one-step simple ultrasound-assisted synthesis method, we used bidistilled water and toluene to test the influence of the nature of two solvents on the redispersion of carbon-based nanomaterials.42,43 Briefly, different amounts of 2371

DOI: 10.1021/acsomega.7b00049 ACS Omega 2017, 2, 2368−2373

ACS Omega

Article

(11) Tachibana, M.; Kobayashi, K.; Uchida, T.; Kojima, K.; Tanimura, M.; Miyazawa, K. Photo-assisted growth and polymerization of C60 “Nano”whiskers. Chem. Phys. Lett. 2003, 374, 279−285. (12) Sathish, M.; Miyazawa, K.; Sasaki, T. Nanoporous fullerene nanowhiskers. Chem. Mater. 2007, 19, 2398−2400. (13) Liu, H.; Li, Y.; Jiang, L.; Luo, H.; Xiao, S.; Fang, H.; et al. Imaging As-Grown [60] Fullerene Nanotubes by Template Technique. J. Am. Chem. Soc. 2002, 124, 13370−13371. (14) Guo, Y. G.; Li, C. J.; Wan, L. J.; Chen, D. M.; Wang, C. R.; Bai, C. L.; Wang, Y. G. Well-Defined Fullerene Nanowire Arrays. Adv. Funct. Mater. 2003, 13, 626−630. (15) Jin, Y.; Curry, R. J.; Sloan, J.; Hatton, R. A.; Chong, L. C.; Blanchard, N.; Stolojan, V.; Kroto, H. W.; Silva, S. R. P. Structural and optoelectronic properties of C60 rods obtained via a rapid synthesis route. J. Mater. Chem. 2006, 16, 3715−3720. (16) Seki, T.; Yagai, S.; Karatsu, T.; Kitamura, A. Formation of Supramolecular Polymers and Discrete Dimers of Perylene Bisimide Dyes Based on Melamine−Cyanurates Hydrogen-Bonding Interactions. J. Org. Chem. 2008, 73, 3328−3335. (17) Madueno, R.; Raisanen, M. T.; Silien, C.; Buck, M. Functionalizing hydrogen-bonded surface networks with selfassembled monolayers. Nature 2008, 454, 618−621. (18) Percec, V.; Glodde, M.; Bera, T. K.; Miura, Y.; Shiyanovskaya, I.; Singer, K. D.; et al. Self-organization of supramolecular helical dendrimers into complex electronic materials. Nature 2002, 419, 384−387. (19) Laschat, S.; Baro, A.; Steinke, N.; Giesselmann, F.; Hägele, C.; Scalia, G.; et al. Discotic liquid crystal: from tailor-made synthesis to plastic electronics. Angew. Chem., Int. Ed. 2007, 46, 4832−4887. (20) Zang, L.; Che, Y.; Moore, J. S. One-Dimensional Self-Assembly of Planar π-Conjugated Molecules: Adaptable Building Blocks for Organic Nanodevices. Acc. Chem. Res. 2008, 41, 1596−1608. (21) Hosseini, M. W. Molecular Tectonics: From Simple Tectons to Complex Molecular Networks. Acc. Chem. Res. 2005, 38, 313−323. (22) Hoeben, F. J. M.; Jonkheijm, P.; Meijer, E. W.; Schenning, A. P. H. J. About Supramolecular Assemblies of π-Conjugated Systems. Chem. Rev. 2005, 105, 1491−1546F. (23) Kelley, R. F.; Lee, S. J.; Wilson, T. M.; Nakamura, Y.; Tiede, D. M.; Osuka, A.; et al. Intramolecular Energy Transfer within ButadiyneLinked Chlorophyll and Porphyrin Dimer-Faced, Self-Assembled Prisms. J. Am. Chem. Soc. 2008, 130, 4277−4284. (24) Li, S.-S.; Northrop, B. H.; Yuan, Q.-H.; Wan, L.-J.; Stang, P. J. Surface Confined Metallosupramolecular Architectures: Formation and Scanning Tunneling Microscopy Characterization. Acc. Chem. Res. 2009, 42, 249−259. (25) Miyazawa, K.; Kuwasaki, Y.; Obayashi, A.; Kuwabara, M. C60 Nanowhiskers Formed by the Liquid−liquid Interfacial Precipitation Method. J. Mater. Res. 2002, 17, 83−88. (26) Fortner, J. D.; Lyon, D. Y.; Sayes, C. M.; Boyd, A. M.; Falkner, J. C.; Hotze, E. M.; Alemany, L. B.; Tao, Y. J.; Guo, W.; Ausman, K. D.; Colvin, V. L.; Hughes, J. B. C60 in Water: Nanocrystal Formation and Microbial Response. Environ. Sci. Technol. 2005, 39, 4307−4316. (27) Li, Z. G.; Fagan, P. J. Coexistence of fcc and hcp phases of C60. Microstructural characterization of C60 and metal complexes of C60. Chem. Phys. Lett. 1992, 194, 461−466. (28) Miyazawa, K. Synthesis of fullerene nanowhiskers using the liquid-liquid interfacial precipitation method and their mechanical, electrical and superconducting properties. Sci. Technol. Adv. Mater. 2015, 16, 1−10. (29) Lieber, C. M.; Chen, C.-C. Preparation of fullerene and fullerene-based materials. Solid State Phys. 1994, 48, 109−148. (30) Bensasson, R. V.; Bienvenue, E.; Dellinger, M.; Leach, S.; Seta, P. C60 in model biological systems. A visible-UV study of solventdependent parameters and solute aggregation. J. Phys. Chem. 1994, 98, 3492−3500. (31) Scharff, P.; Risch, K.; Carta-Abelman, L.; Dmytuk, I. M.; Bilyi, M. M.; Golub, O. A.; et al. Structure of C60 fullerene in water: spectroscopic data. Carbon 2004, 42, 1203.

the optical properties and morphology. Further experiments (either kinetically or thermodynamically controlled) are necessary to clarify and explain the formation process of the above-mentioned structures in detail.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Phone: +34 687 858 923. ORCID

Antonio Cid: 0000-0002-0507-3394 Benito Rodríguez-González: 0000-0002-8885-788X Juan C. Mejuto: 0000-0001-8396-1891 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Dr. A.C. acknowledges the post-doctoral grants SFRH/BD/ 78849/2011 granted to Requimte and UID/MULTI/04378/ 2013 granted to Unidade de Ciências Biomoleculares aplicadas (UCIBIO), both from the Portuguese Foundation for Science and Technology.

■ ■

ABBREVIATIONS FNWs, fullerene C60 nanowhiskers FC60 NPs, fullerene C60 quasispherical nanoparticles

ADDITIONAL NOTE Nanowhiskers have been defined as needlelike crystals of fullerene C60 with a submicron diameter,25 and quasispherical nC60 have been previously reported by Fortner et al.26 as a structure formed as a function of rate of solvent addition and pH. a



REFERENCES

(1) Cataldo, F.; Da Ros, T., Eds. Medicinal Chemistry and Pharmacological Potential of Fullerenes and Carbon Nanotubes; Springer: Dordrecht, Netherlands, 2008; Vol. 1. (2) Prylutska, S. V.; Burlaka, A. P.; Prylutskyy, Y. I.; Ritter, U.; Scharff, P. Pristine C(60) fullerenes inhibit the rate of tumor growth and metastasis. Exp. Oncol. 2011, 33, 162−164. (3) Li, C. Z.; Chueh, C.-C.; Yip, H.-L.; Ding, F.; Li, X.; Jen, A. K.-Y. Solution-processible highly conducting fullerenes. Adv. Mater. 2013, 25, 2457−2461. (4) Hasobe, T. Supramolecular nanoarchitectures for light energy conversion. Phys. Chem. Chem. Phys. 2010, 12, 44−57. (5) Segura, J. L.; Martín, N.; Guldi, D. M. Materials for organic solar cells: the C60/π-conjugated oligomer approach. Chem. Soc. Rev. 2005, 34, 31−47. (6) Sawamura, M.; Kawai, K.; Matsuo, Y.; Kanie, K.; Kato, T.; Nakamura, E. Stacking of conical molecules with a fullerene apex into polar columns in crystals and liquid crystals. Nature 2002, 419, 702− 705. (7) Babu, S. S.; Möhwald, H.; Nakanishi, T. Recent progress in morphology control of supramolecular fullerene assemblies and its applications. Chem. Soc. Rev. 2010, 39, 4021−4035. (8) Miyazawa, K.; Kuwasaki, Y.; Obayashi, A.; Kuwabara, M. C60 nanowhiskers formed by the liquid − liquid interfacial precipitation method. J. Mater. Res. 2002, 17, 83−88. (9) Miyazawa, K.; Hamamoto, K.; Nagata, S.; Suga, T. Structural investigation of the C60/C70 whiskers fabricated by forming liquid− liquid interfaces of toluene with dissolved C60/C70 and isopropyl alcohol. J. Mater. Res. 2003, 18, 1096−1103. (10) Minato, J. I.; Miyazawa, K. Solvated structure of C 60 nanowhiskers. Carbon 2005, 43, 2837−2841. 2372

DOI: 10.1021/acsomega.7b00049 ACS Omega 2017, 2, 2368−2373

ACS Omega

Article

(32) Li, Q.; Xie, B.; Hwang, Y. S.; Xu, Y. Kinetics of C60 Fullerene Dispersion in Water Enhanced by Natural Organic Matter and Sunlight. Environ. Sci. Technol. 2009, 43, 3574−3579. (33) Miyazawa, K.; Kuwasaki, Y.; Hamamoto, K.; Nagata, S.; Obayashi, A.; Kuwabara, M. Structural characterization of C60 nanowhiskers formed by the liquid/liquid interfacial precipitation method. Surf. Interface Anal. 2003, 35, 117−120. (34) Kato, R.; Miyazawa, K. Cross-sectional structural analysis of C60 nanowhiskers by transmission electron microscopy. Diamond Relat. Mater. 2011, 20, 299−303. (35) McCready, D.; Alnajjar, M. Powder Diffraction File, 44-558; International Centre for Diffraction Data: Newton Square, PA, 1994. (36) Sathish, M.; Miyazawa, K.; Hill, J. P.; Ariga, K. Solvent engineering for shape-shifter pure Fullerene (C60). J. Am. Chem. Soc. 2009, 131, 6372−6373. (37) Qu, Y.; Liang, S.; Zou, K.; Li, S.; Lumei, L.; Zhao, J.; et al. Effect of solvent type on the formation of tubular fullerene nanofibers. Mater. Lett. 2011, 65, 562−564. (38) Miyazawa, K.; Hotta, K. The effect of solvent ratio and water on the growth of C60 nanowhiskers. J. Cryst. Growth 2010, 312, 2764− 2770. (39) Hamon, A.-L.; Verbeeck, J.; Schryvers, D.; Benedikt, J.; Sanden, R. M. C. M. v. d. ELNES study of carbon K-edge spectra of plasma deposited carbon films. J. Mater. Chem. 2004, 14, 2030−2035. (40) Padma, L. N.; Ke, Y.; Pradeep, K. D.; Shridhara, A.; Jayant, K.; Sukant, K. T. Polyelectrolyte-Containing Fullerene I: Synthesis and Characterization of the Copolymers of 4-Vinylbenzoic Acid with C60. Chem. Mater. 1998, 10, 2058−2066. (41) Ruoff, R. S.; Tse, D. S.; Malhotra, R.; Lorents, D. C. Solubility of C60 in a Variety of Solvents. J. Phys. Chem. 1993, 97, 3379−3383. (42) Marcus, Y.; Smith, A. L.; Korobov, M. V.; Mirakyan, A. L.; Avramenko, N. V.; Stukalin, E. B. Solubility of C60 fullerene. J. Phys. Chem. B 2001, 105, 2499−2506. (43) Prylutskyy, Y. I.; Petrenko, V. I.; Ivankov, O. I.; Kyzyma, O. A.; Bulavin, L. A.; Litsis, O. O.; et al. On the Origin of C60 Fullerene Solubility in Aqueous Solution. Langmuir 2014, 30, 3967−3970.

2373

DOI: 10.1021/acsomega.7b00049 ACS Omega 2017, 2, 2368−2373