Open access peer-reviewed chapter

Rational Synthesis of Fullerenes

Written By

Konstantin Amsharov

Submitted: 05 October 2017 Reviewed: 19 December 2017 Published: 07 February 2018

DOI: 10.5772/intechopen.73251

From the Edited Volume

Fullerenes and Relative Materials - Properties and Applications

Edited by Natalia V. Kamanina

Chapter metrics overview

1,355 Chapter Downloads

View Full Metrics

Abstract

Fullerenes are a unique family of carbon-based cage molecules, which attract interest because of their remarkable properties and potential applications. Most effort so far has been focused on the study of C60 and C70, whereas other members of the huge fullerene family remain poorly explored. One of the main challenges in this field is the developing of the synthetic methods, which are suitable for the production of these unique materials in isomer-pure form in macroscopic amounts. Here, we review studies toward the rational synthesis of fullerenes from molecular precursors that have been published to date. The scope and limitation of the zipping strategy are discussed. The relevance and prospects for construction of the fullerene cages and related carbon-based nanostructures via cyclodehydrofluorination (C─F bond activation) are highlighted.

Keywords

  • fullerenes
  • bottom-up synthesis
  • zipping approach
  • C─F bond activation

1. Introduction

Fullerenes represent a unique class of sp2-carbon-based compounds with a spherical π-system, displaying intriguing physical and chemical properties [1, 2]. The conventional methods of fullerene production are based on various graphite evaporation techniques which result in the heterogeneous fullerene containing soot [1, 2]. Unfortunately, none of the existing techniques provide access to fullerenes in isomer-pure form, which is crucial for their possible applications. The random nature of the fullerene formation results in a highly heterogeneous mixture containing mainly the most stable C60-Ih and C70-D5h cages. Both fullerenes can be rather easily obtained in pure form by single-step HPLC separation of the fullerene soot extract (Figure 1). So far, only these two fullerenes have been the subject of in-depth investigations although the fullerene family represents a huge class of carbon-cage molecules. As soon as C60 became available in bulk amounts, a number of unique properties such as superconductivity [3], ferromagnetism [4, 5], and high electron affinity [6] have been disclosed. C60 and C70 fullerenes were found to be ideal material for the development of solar-energy conversion systems, in which the fullerene plays the role of effective electron acceptor [7]. Besides these two fullerene members, the fullerene soot typically contains about 1% of the so-called higher fullerenes (C2n with n > 35). In contrast to C60 and C70, the higher fullerenes have several structural isomers obeying the isolated pentagon rule (IPR), stating that all pentagons have to be completely surrounded by hexagons [8]. Some of isomers of C7684 can be obtained in small amounts by multi-step HPLC using recycling HPLC. However, the HPLC technique reaches its limit already on the separation of individual isomers of C84. The separation of isomers of higher fullerenes with more than 84 carbon atoms is extremely complicated and these fullerenes remain highly unexplored (Figure 1). Moreover, the number of possible IPR isomers surges exponentially with increasing number of carbon atoms in the cage, which make the separation of higher fullerenes practically impossible. For example, C100 has 450 IPR isomers whereas for C200 the number of possible IPR isomers is 15.655.672. Furthermore, fullerenes that satisfy the IPR rule represent only a tiny fraction of possible carbon cages that in principle can be formed. Thus, C60 has 1811 non-IPR isomers whereas C200 has already more than 214 millions of isomers [8]. The non-IPR cages are remarkably less stable than IPR fullerenes, since adjacency of two pentagons resulted in high local strain and therefore high reactivity of these species. Nevertheless, it has been demonstrated that non-IPR cages exist and can be isolated at least in the form of exohedral or endohedral derivatives [9]. Furthermore, the possibility to introduce heteroatoms (heterofullerenes) [10], and/or to incorporate rings of different sizes (non-classical fullerenes) into the carbon cage [11], provides virtually unlimited number of possible fullerene structures. Each individual fullerene molecule possesses its own set of unique properties such as HOMO-LUMO gap, electron affinity, optical and magnetic characteristics. Despite intense research in the fullerene field during last decades, only a small fraction of possible fullerenes has been isolated in pure form and characterized. The higher fullerenes, as well as non-IPR and non-classical cages remain poorly explored due to the extremely low content in the soot and difficulties connected with purification. Further development in this field requires alternative synthetic methods which are suitable for the macroscopic production of desired fullerenes in isomer-pure form.

Figure 1.

(Left) Several members of the fullerene family and the number of possible IPR isomers. (Right) Typical HPLC profile of the fullerene soot extract (Buckyprep column, toluene as eluent, UV detection). Two strong signals at 8 and 10 min correspond to C60-Ih and C70-D5h fullerenes, which can be easily isolated. The fraction eluting between 17 and 25 min corresponds to abundant higher fullerenes—C76-C84 (mixture of about 30 fullerene species). The huge number of small overlapping peaks (tR > 28 min) corresponds to the mixture of a large number of unknown higher fullerenes.

2. The concept of the rational synthesis of fullerenes

Rational synthesis of fullerenes is of practical interest as a promising method for the synthesis of isomer-pure fullerenes, including fullerenes which cannot be obtained by conventional techniques. One of the first attempts of the “rational synthesis“ of fullerenes was undertaken in 1993 [12]. It has been proposed that C60-Ih fullerene cage can be constructed by the fusion of six naphthalene moieties. Indeed, small amounts of C60 were detected in the pyrolysis products of the naphthalene performed in the argon atmosphere at 1000°C [12]. The proposed mechanism suggests a step-by-step assembly of naphthalene molecules via cyclodehydrogenation leading to various oligo-naphthalenes. Some of hexamers with defined orientation of naphthalene fragments possess the required for C60 connectivity, and thus can act as a precursor of C60. The intramolecular cyclization of such precursors should unambiguously lead to the C60 cage formation. One of the possible naphthalene hexamers (C60H28), which is likely to form C60 trough intramolecular condensation, is shown in Figure  2. The low yield of C60 in this case might be explained by the low probability of the formation of the naphthalene hexamer with required orientation of subunits during accidental condensation. Indeed, as it was shown later, the pyrolysis of properly connected oligo-naphthalene improves the C60 yield [13]. In spite of the fact that this hypothesis does not explain the simultaneous formation of C70, which was also observed in the pyrolysis products (direct assembly of C70 from naphthalene fragments is impossible), the given mechanism suggests that fullerenes can be built from well-defined aromatic fragments—the so-called fullerene precursors. These polycyclic aromatic hydrocarbons (PAHs), already containing required carbon connectivity, appear to be the key intermediates which exhaustive cyclization should lead to the strictly predefined carbon cage. In 2001, this methodology was successfully employed by Scott et al. for C60 fullerene synthesis under laser ablation conditions [14]. The rigid C60 fullerene precursor, containing all 60 carbon atoms at required positions, was obtained through a multi-step organic synthesis and subjected to the LDI-induced “pyrolysis” (laser desorption ionization). According to MS analysis the LDI-induced cyclization led the formation of C60 (Figure 2) [14]. Although the cyclization was found to be rather inefficient, the demonstrated principal transformation of the preprogrammed PAH precursor to the desired carbon cage paves the way for the rational construction of the fullerenes and other related carbon-based nanostructures.

Figure 2.

(Top) Schematic presentation of the C60 fullerene formation during naphthalene pyrolysis showing formation of the naphthalene hexamer as a key intermediate. (Bottom) Synthesis of C60 fullerene by zipping of the rigid atomically precise precursor under LDI conditions. The new C─C bond, which forms during cyclization, is coded by dashed lines.

The general strategy of the rational synthesis of fullerenes is based on the synthesis of specially “designed” PAH-precursor molecules containing the carbon framework required for the formation of the target fullerene cage. The respective precursor can be “zipped” to the desired fullerene by the tandem intramolecular Aryl─Aryl coupling. The characteristic feature of this approach is a zipper mechanism of cyclization in the final synthetic step. Since the regiospecificity of each condensation step can be strictly predefined by the precursor design, the cyclization will lead to the preprogrammed isomer of fullerene exclusively. This final cyclization step represents the key transformation in the rational synthesis of fullerenes by the zipping approach.

2.1. Wet-chemical synthesis

Since a big number of new C─C bonds have to be established during precursor zipping, the efficiency of each condensation has to be very high in order to achieve reasonable conversion. Unfortunately, methods developed for the synthesis of the planar PAHs via intramolecular Aryl–Aryl coupling are usually uneffective in the case of highly strained fullerenes. The strain energy of the fullerenes stems mainly due to pyramidalization of carbon atoms, as a result of deviation from energetically ideal values observed in graphene lattice. The perturbation of the π-system causes enormous enhancement of the energy, which has to be overcome during cyclization. Thus for example, various polyphenylenes can readily undergo an intramolecular cyclization under UV irradiation in the presence of an oxidant [15]. The photocyclization is an effective and frequently used technique for the synthesis of large planar polycyclic systems. Alternatively, oxidative intramolecular cyclodehydrogenation can be carried out under Scholl condition [16]. Recently, this approach has become of high practical interest for the synthesis of large PAHs such as nanographenes and nanoribbons [17, 18]. Although both cyclodehydrogenation methods show excellent efficiency in the case of planar PAHs, these techniques appear to be not effective for the synthesis of bowl-shaped PAHs and fullerenes. Several alternative intermolecular aryl–aryl coupling techniques have been developed to overcome these disadvantages. Here the palladium(0)-catalyzed direct arylation utilizing bromo and chloro derivatives has been found to be effective for the synthesis of small fullerene fragments [19]. The largest fullerene fragments obtained so far contain 50 carbon atoms and formally represent more than 83% of the C60 fullerene surface (Figure 3) [20].

Figure 3.

The synthesis of bowl-shaped PAH (C60 fullerene fragment) by Pd-catalyzed microwave-assisted intramolecular arylations.

Although the largest synthesized C50 bowl is more curved than C60 [20], the rational synthesis of fullerene following this route still remains a challenge. This is mainly connected with difficulties to perform the Pd-catalyzed cyclization in a domino fashion. Secondly, the reaction conditions appear to be aggressive with respect to fullerene species. The recent discovery of effective intramolecular aryl–aryl coupling via C-F bond activation under transition-metal free conditions might be a solution to this problem. Recently, Siegel et al. [21], Ichikawa et al. [22], and Amsharov et al. [23, 24] have demonstrated the efficiency of the C─F bond activation strategy for the synthesis of various bowl-shaped PAHs via formal HF elimination step (cyclodehydrofluorination). In the last case, the effective cyclization was achieved utilizing easily accessible aluminum oxide, which makes the approach highly attractive for preparative-scale synthesis. Moreover, it has been found that intramolecular Aryl-Aryl coupling of fluoroarenes via γ-aluminum oxide-mediated C─F bond activation can be realized under mild conditions with unprecedentedly high chemoselectivity [25, 26]. Using this technique an extended bowl-shaped PAHs representing more than 75% of the C60 fullerene connectivity was obtained with close to quantitative yield (Figure 4) [24], demonstrating high potential for generating extended nonplanar carbon-based nanostructures including fullerenes. Although the rather big fullerene fragments can be successfully produced by wet-chemical methods, the solution-synthesis of fullerenes remains a challenge.

Figure 4.

(Left) Synthesis of the extended fullerene fragments via alumina-assisted cyclodehydrofluorination. (Right) Superposition of C60 fullerene (dashed lines) and buckybowl (highlighted by gray) structures showing the same atom topology.

2.2. Surface-assisted synthesis

The direct on-surface synthesis of nanostructures is a highly promising strategy. The on-surface studies are usually conducted on single-crystal surfaces under ultra-high vacuum conditions that offer extraordinary degree of control in the experiments [27]. Moreover, the usage of atomically clean and flat crystal surfaces allows the observation of the reaction process at the single molecule level by means of scanning tunneling microscopy (STM) [28]. Recent progress in this field demonstrates a high potential for the synthesis of complex architectures via Aryl-Aryl coupling of organic precursors on single-crystal metal surfaces [29, 30]. In the case of planar PAHs, the intramolecular cyclodehydrogenation can be performed effectively on Cu(111), Ag(111), and Au(111) surfaces. Intramolecular cyclization on Au(111) surfaces typically occurs at around 400°C [31]. On the more reactive copper surfaces, cyclization takes place earlier at 200°C [32]. However, attempts to perform a similar cyclization to form non-planar PAHs were unsuccessful because of the intermolecular merging of the precursor molecules. Therefore, the synthesis of fullerenes requires harder conditions, including using more reactive Pt-group metal surfaces and applying higher temperatures of 450–600°C [33]. The first surface-assisted synthesis of the bowl-shaped PAH was reported by Rim et al. on the example of the hemispherical PAH C48, which was generated on the Ru(0001) surface via exhaustive cyclodehydrogenation of hexabenzocoronene precursor (Figure 5) [34]. The six-fold cyclodehydrogenation on the periphery of precursor was achieved by annealing of the precursor molecule at 600°C. Note that Ru surface serves not only as a catalyst for the cyclodehydrogenation it also prevents intermolecular coupling due to efficient bonding to the surface. Finally, the Ru surface acts as a support for precursors and products enabling STM investigation. The STM analysis has revealed a high selectivity and exceptionally high conversion ratio of the deposited precursors into non-planar PAH structure (Figure 5). However, due to high reactivity of the Ru surface and high temperature applied, the dehydrogenation process does not stop on the formation of the C48H12 bowl. The bowl-shaped PAH undergoes further dehydrogenation resulting in all-carbon C48 bowl covalently bonded to the Ru surface by 12 C─Ru bonds (Figure 5) [34].

Figure 5.

Surface-assisted synthesis of the hemispherical C48 buckybowl via exhaustive cyclodehydrogenation of hexabenzocoronene on Ru surface at 600°C.

Soon after, the efficiency of the surface-assisted approach has been demonstrated by Otero et al. on the synthesis of C60 fullerene on Pt surface [35]. In their study, C60H30 fullerene precursor molecules (Figure 6, precursor A) were deposited onto a Pt(111) surface and annealed at around 480°C. The annealing step led to the formation of C60 cages with an efficiency close to 100%. The STM analysis shows that the molecular coverage does not change during temperature-induced reaction, and all of the precursor molecules converted into fullerenes [35]. Later it was shown that the process is highly selective in nature and is not accomplished by skeletal rearrangements. This important question has been addressed by the investigation of surface-catalyzed cyclodehydrogenation of a specifically designed precursor molecule (Figure 6, precursor B) “preprogrammed” to the formation of the C60-bowl-shaped PAH with unique triangular shape [36]. The on-surface cyclodehydrogenation of two isomeric C60H30 precursors (Figure 7) gave convincing evidence that the reaction occurs in a highly selective manner [36]. In contrast to the condensation of the precursor A whose cyclodehydrogenation leads to the spherical C60, the zipping of the precursor B resulted in a triangular-shaped C60-bowls whose formation was distinguished by the STM analysis.

Figure 6.

Structures of the C60 fullerene precursor (A and the C60 buckybowl precursor (B) molecules and their cyclodehydrogenation products.

Figure 7.

Surface-assisted synthesis of fullerene C60-Ih (top) and C60 buckybowl (bottom) via cyclodehydrogenation of respective precursors on Pt(111) surface. STM images of the precursor before and after annealing at 480°C (reprint with permission from Ref [36]).

Those works demonstrated that the surface-assisted cyclodehydrogenation on Pt-group metal surfaces is highly selective and can be used for fabrication of various specifically designed carbon-based nanostructures including fullerenes.

Finally, the approach was successfully applied for the first rational synthesis of isomer-pure higher fullerene C84(20) [36]. Upon annealing of specially designed C84H42 precursor at 550°C on Pt(111) surface, the shape of all of precursor molecules changed and all species became spherical with a lateral size expected for C84 fullerene (Figure 8). The formation of the C84 fullerene cages was additionally confirmed by MS analysis [36]. Moreover, the approach was found to be suitable for the synthesis of highly interesting hetero fullerenes. The fabrication of an azafullerene by surface-assisted approach has been demonstrated by Otero et al. on the example of C57N3 synthesis. In this study, a structurally related precursor with three incorporated nitrogen atoms was quantitatively converted to the hetero fullerene C57N3 upon annealing on Pt(111) surface at 477°C [35].

Figure 8.

Surface-assisted synthesis of fullerene C84(20) via cyclodehydrogenation of C84H42 precursors on Pt(111) surface. STM images of the precursor before and after annealing at 550°C (reprint with permission from Ref. [36]).

Although the technique allows to perform effective zipping to the target carbon nanostructures [37, 38], the resulting nano-carbons are covalently bonded to the metal surface and cannot be desorbed without decomposition. Those difficulties could be probably overcome by using less active metal oxide surfaces. Since the metal oxides show low activity in cyclodehydrogenation, the cyclization on such surfaces would require much higher temperatures leading to loss of selectivity. However, introduction of “activating” groups into the precursor structures could significantly reduce the activation barriers. In particular, it has been found that fluorine functionality appears to be an excellent ring-closure promoter, which has demonstrated unexpectedly high efficiency. It has been found that coupling can be realized in fluoroarenes on thermally activated aluminum oxide at very mild conditions [23, 24, 25, 26]. Moreover, the alumina-supported cyclodehydrofluorination (cyclization via formal HF elimination) has demonstrated unprecedentedly high chemoselectivity and regiospecificity. The intramolecular Aryl–Aryl coupling on aluminum oxide was achieved with close to quantitative yield even in the case of large bowl-shaped PAHs. Since fluorine can promote the ring closure only if hydrogen is placed neighboring in space in the precursor structure, it was possible to perform the cyclization in domino fashion. This makes the cyclodehydrofluorination more attractive in comparison to cyclodehydrogenation where the selectivity must be predefined by the rigid structure of precursor. Further, it was found that metal oxides of II, III and IV groups are active in the cyclodehydrofluorination of fluoroarenes. In particular, zirconium, titanium and indium oxides were found to be effective for cyclization via C─F bond activation [39]. These results point the way for fabrication of higher fullerenes and other related carbon-based nanostructures in a fully controllable manner directly on metal oxide surfaces.

2.3. Gas phase synthesis

The rational construction of fullerenes in the gas phase seems to be very logic, since in all conventional techniques the cage fabrication occurs in the gas phase. First of all, fullerenes, including higher, non-IPR and heterofullerenes, are stable in the gas phase. Second, the undesired intermolecular interactions can be effectively suppressed in the gas phase, providing excellent conditions for intramolecular processes. Not surprising that gas-phase synthesis appears to be the most promising approach.

The flash vacuum pyrolysis (FVP) represents one of the most powerful techniques for the synthesis of highly strained architectures via intramolecular cyclization. It consists in the heating of precursor molecules in the gas phase to a very high temperature for a very short time. The high temperature applied in FVP causes hemolytic cleavage of C─H bonds leading to the formation of radicals possessing enough energy to cause intramolecular cyclization yielding highly strained products. Pioneering by Scott et al., variety of small bowl-shaped PAHs (fullerene fragments) have been successfully synthesized in macroscopic amounts via intramolecular cyclization of quasi-planar PAH precursors under FVP conditions [40, 41, 42]. It was found that introduction of chlorine or bromine functionalities into the initial precursor is essential to facilitate the cyclization via free radical mechanism (Figure 9). Finally, this methodology has been successfully applied for the first rational synthesis of C60 utilizing partially chlorinated C60 fullerene precursor (Figure 9) [43]. Despite the low yield of C60 (0.1–1%) the power of the strategy has been clearly demonstrated.

Figure 9.

(Left) Synthesis of fullerenes and fullerene fragments by FVP approach from chlorinated PAH precursors. (Right) The cyclization mechanism under FVP condition for activated (chlorinated) and non-activated PAHs.

The possibility of the synthesis of higher fullerenes by FVP has been shown on the examples of C78 [44] and C84 [45] by pyrolysis of specially programmed precursors. In both cases the size-selective formation of fullerenes was observed as indicated by the MS analysis (Figure 10). However, the rates of conversion to the target molecules have remained disappointingly low because of the low efficiency of intramolecular condensation of non-activated precursors. The usually employed for activation chlorine reach their limits in the case of large molecules because of decomposition of precursors during sublimation. The molar masses will generally be high, since a large number of new C─C bonds that need to be formed necessitate the introduction of a big number of promoter groups. Moreover, the radical nature of the condensation could drastically affect the selectivity of the process.

Figure 10.

Schematic representation of the intramolecular cyclodehydrogenation of fullerene precursors to fullerenes C78(4) and C84(20). Dashed lines indicate where the new bonds have to be established in order to create a closed fullerene cage.

Although the FVP approach has been proven to be very prolific for the synthesis of small non-planar PAHs, the high-yield synthesis of large bowls and isomerically pure fullerenes has remained challenging. Recently, an efficient intramolecular fluorine-promoted ring closure in benzo[c]phenanthrenes under FVP conditions via HF elimination has been reported [46]. It has been shown that HF elimination is a synchronous process leading directly to the C─C bond formation without any intermediates, thus producing no side products [46]. The small size and low molecular weight of fluorine, as well as high thermostability of the C─F bond make fluorine a perfect activating group for the rational synthesis of fullerene by zipping approach. The feasibility of the approach has been demonstrated on the successful synthesis of C60 fullerene from the respective fluorinated precursor which was effectively folded into the fullerene under laser ionization. The respective 15-fold cyclization was realized in a domino fashion via consecutive HF elimination without observation of side reactions or undesired fragmentations [47]. As it is shown in Figure 11, the precursor molecule undergoes highly selective HF loss (with simultaneous C─C bond formation) leading to the open-cage which spontaneously collapses to the target C60 via H2 loss. The regioselectivity of the domino-like cyclization and the formation of bowl-shaped intermediates and finally spherical C60-Ih cage were recently confirmed by ion-mobility experiments [48].

Figure 11.

Mass spectrum of the C60 fullerene precursor (C60H21F9) obtained under laser ablation (negative-ion mode) showing an effective domino-like HF elimination yielding C60 fullerene via 15 selective C─C bond formation steps. The mechanism of single cyclodehydrofluorination involving a four-centered transition state is shown in the insert on the example of benzo[c]phenanthrene model.

Due to exceptionally high efficiency, it was assumed that the same approach can be extended to the synthesis of higher fullerenes as well as to the synthesis of highly attractive non-IPR cages. Recently, it has been demonstrated that non-IPR fragments can be indeed fabricated via cyclodehydrofluorination in the gas phase. It was shown that despite high strain the respective precursor undergoes rather smooth four-fold cyclodehydrofluorination, leading to the non-IPR fullerene fragment C26H10 (Figure 12) [49]. These results demonstrate the general possibility of rational fabrication of the non-IPR architectures via cyclodehydrofluorination of the specially preprogrammed fluoroarene-based precursors. Two possible precursors for generation of non-IPR fullerenes C38 and C42 are shown as example. Taking into account the present state-of-the-art of MS-based synthesis/deposition technique, which allows multilayer deposition of such compounds, the future scale-up appears to be feasible. Among non-IPR cages, the laser-induced cyclodehydrofluorination approach seems to be also applicable for the rational synthesis of higher and non-classical carbon cages.

Figure 12.

Mass spectrum of model precursor C26H14F4 obtained under laser ablation showing tandem cyclodehydrofluorination yielding the non-IPR C26H10 fragment. The DFT optimized geometry of the non-IPR bowl (side and top views) is shown in the inset. (Right) Two examples of possible non-IPR fullerene precursor.

Advertisement

3. Conclusion

The zipping strategy had been proved to be prolific for the rational synthesis of the fullerene fragments and the fullerene cages. The progress in this field shows the general possibility to fabricate elusive fullerenes in a fully controllable manner. Production of these unique materials in isomer-pure form in macroscopic amounts does no longer appear elusive, and there is promising prospect that such compounds will become available in the near future. Since there are virtually no limiting factors, there are no principle obstacles standing against extending the zipping approach to highly interesting related carbon-based architectures.

Advertisement

Acknowledgments

The author thanks the Deutsche Forschungsgemeinschaft (DFG-SFB 953 “Synthetic Carbon Allotropes”).

References

  1. 1. Hirsch A, Brettreich M. Fullerenes: Chemistry and Reactions. Weinheim, FRG: Wiley-VCH Verlag GmbH & Co. KGaA; 2004. 423 pp. DOI: 10.1002/3527603492
  2. 2. Langa F, Nierengarten JF. Fullerenes: Principles and Applications. Cambridge: RSC Publishing; 2011. 634 pp. DOI: 10.1039/9781849732956
  3. 3. Hebard AF, Rosseinsky MJ, Haddon RC, Murphy DW, Glarum SH, Palstra TTM, Ramirez AP, Kortan AR. Superconductivity at 18 K in potassium-doped C60. Nature. 1991;350:600-601. DOI: 10.1038/350600a0
  4. 4. Allemand PM, Khemani KC, Koch A, Wudl F, Holczer K, Donovan S, Gruner G, Thompson JD. Organic molecular soft ferromagnetism in a fullerene C60. Science. 1991;253:301-302. DOI: 10.1126/science.253.5017.301
  5. 5. Stephens PW, Cox D, Lauther JW, Mihaly L, Wiley JB, Allemand PM, Hirsch A, Holczer K, Li Q, Thompson JD, Wudl F. Lattice structure of the fullerene ferromagnet TDAE–C60. Nature. 1992;355:331-332. DOI: 10.1038/355331a0
  6. 6. Reed CA, Bolskar RD. Discrete fulleride anions and fullerenium cations. Chemical Reviews. 2000;100:1075-1119. DOI: 10.1021/cr980017o
  7. 7. Lu L, Zheng T, Wu Q, Schneider AM, Zhao D, Yu L. Recent advances in bulk heterojunction polymer solar cells. Chemical Reviews. 2015;115:12666-12731. DOI: 10.1021/acs.chemrev.5b00098
  8. 8. Fowler PW, Manolopoulos DE. An Atlas of Fullerenes. Oxford: Clarendon Press; 1995. 392 pp
  9. 9. Tan YZ, Xie SY, Huang RB, Zheng LS. The stabilization of fused-pentagon fullerene molecules. Nature Chemistry. 2009;1:450-460. DOI: 10.1038/nchem.329
  10. 10. Vostrowsky O, Hirsch A. Heterofullerenes. Chemical Reviews. 2006;106:5191-5207. DOI: 10.1021/cr050561e
  11. 11. Gan LH, Zhao JQ, Hui Q. Nonclassical fullerenes with a heptagon violating the pentagon adjacency penalty rule. Journal of Computational Chemistry. 2010;31:1715-1721. DOI: 10.1002/jcc.21459
  12. 12. Taylor R, Langley GJ, Kroto HW, Walton DRM. Formation of C60 by pyrolysis of naphthalene. Nature. 1993;366:728-731. DOI: 10.1038/366728a0
  13. 13. Amsharov KY, Simeonov K, Jansen M. Formation of fullerenes by pyrolysis of 1,2′-binaphthyl and 1,3-oligonaphthylene. Carbon. 2007;45:337-343. DOI: 10.1016/j.carbon.2006.09.013
  14. 14. Boorum MM, Vasil’ev YV, Drewello T, Scott LT. Groundwork for a rational synthesis of C60: Cyclodehydrogenation of a C60H30 polyarene. Science. 2001;294:828-831. DOI: 10.1126/science.1064250
  15. 15. Mallory FB, Mallory CW. Photocyclization of stilbenes and related molecules. Organic Reactions. Hoboken, New Jersey: John Wiley & Sons, Inc.; 2005. 456 pp. DOI: 10.1002/0471264180.or030.01
  16. 16. Grzybowski M, Skonieczny K, Butenschon H, Gryko DT. Comparison of oxidative aromatic coupling and the Scholl reaction. Angewandte Chemie, International Edition. 2013;52:9900-9930. DOI: 10.1002/anie.201210238
  17. 17. Chen L, Hernandez Y, Feng XL, Mullen K. From nanographene and graphene nanoribbons to graphene sheets: Chemical synthesis. Angewandte Chemie, International Edition. 2012;51:7640-7654. DOI: 10.1002/anie.201201084
  18. 18. Kivala M, Wu D, Feng X, Li C, Mullen K. Cyclodehydrogenation in the synthesis of graphene-type molecules. In: Materials Science and Technology. Weinheim: Wiley-VCH; 2013. pp. 373-420. DOI: 10.1002/9783527603978.mst0423
  19. 19. Alberico D, Scott ME, Lautens M. Aryl−aryl bond formation by transition-metal-catalyzed direct arylation. Chemical Reviews. 2007;107:174-238. DOI: 10.1021/cr0509760
  20. 20. Steinberg BD, Jackson EA, Filatov AS, Wakamiya A, Petrukhina MA, Scott LT. Aromatic π-systems more curved than C60. The complete family of all indenocorannulenes synthesized by iterative microwave-assisted Intramolecular arylations. Journal of the American Chemical Society. 2009;131:10537-10545. DOI: 10.1021/ja9031852
  21. 21. Allemann O, Duttwyler S, Romanato P, Baldridge KK, Siegel JS. Proton-catalyzed, silane-fueled Friedel-Crafts coupling of fluoroarenes. Science. 2014;332:574-577. DOI: 10.1126/science.aaa4609
  22. 22. Fuchibe K, Mayumi Y, Zhao N, Watanabe S, Yokota M, Ichikawa J. Domino synthesis of fluorine-substituted polycyclic aromatic hydrocarbons: 1,1-Difluoroallenes as synthetic platforms. Angewandte Chemie, International Edition. 2013;52:7825-7828. DOI: 10.1002/anie.201302740
  23. 23. Amsharov KY, Merz P. Intramolecular aryl-aryl coupling of fluoroarenes through Al2O3-mediated HF elimination. The Journal of Organic Chemistry. 2012;77:5445-5448. DOI: 10.1021/jo300783y
  24. 24. Amsharov KY, Kabdulov MA, Jansen M. Facile bucky-bowl synthesis by regiospecific Cove-region closure by HF elimination. Angewandte Chemie, International Edition. 2012;51:4594-4597. DOI: 10.1002/anie.201200516
  25. 25. Papaianina O, Akhmetov VA, Goryunkov AA, Hampel F, Heinemann FW, Amsharov KY. Synthesis of rationally halogenated buckybowls via chemoselective aromatic C─F bond activation. Angewandte Chemie, International Edition. 2017;56:4834-4838. DOI: 10.1002/anie.201700814
  26. 26. Steiner AK, Amsharov KY. The rolling-up of oligophenylenes to nanographenes by a HF-zipping approach. Angewandte Chemie, International Edition. 2017;56:14732-14736. DOI: 10.1002/anie.201707272
  27. 27. Tait SL. Function follows form: Exploring two-dimensional supramolecular assembly at surfaces. ACS Nano. 2008;2:617-621. DOI: 10.1021/nn800207w
  28. 28. Hla SW, Rieder KH. STM control of chemical reaction: Single-molecule synthesis. Annual Review of Physical Chemistry. 2003;54:307-330. DOI: 10.1146/annurev.physchem.54.011002.103852
  29. 29. Gourdon A. On-surface covalent coupling in ultrahigh vacuum. Angewandte Chemie, International Edition. 2008;47:6950-6953. DOI: 10.1002/anie.200802229
  30. 30. Mendez J, Lopez MF, Martin-Gago JA. On-surface synthesis of cyclic organic molecules. Chemical Society Reviews. 2011;40:4578-4590. DOI: 10.1039/C0CS00161A
  31. 31. Cai J, Ruffieux P, Jaafar R, Bieri M, Braun T, Blankenburg S, Muoth M, Seitsonen AP, Saleh M, Feng X, Müllen K, Fasel R. Atomically precise bottom-up fabrication of graphene nanoribbons. Nature. 2010;466:470-473. DOI: 10.1038/nature09211
  32. 32. Treier M, Pignedoli CA, Laino T, Rieger R, Müllen K, Passerone D, Fasel R. Surface-assisted cyclodehydrogenation provides a synthetic route towards easily processable and chemically tailored nanographenes. Nature Chemistry. 2011;3:61-67. DOI: 10.1038/nchem.891
  33. 33. Amsharov K. Rational synthesis of fullerenes, buckybowls, and single-walled carbon nanotubes by a surface-assisted approach. Advances in polymer science. In: Müllen K, Feng X, editors. From Polyphenylenes to Nanographenes and Graphene Nanoribbons. Cham: Springer; 2017. pp. 127-145. DOI: 10.1007/12_2017_7
  34. 34. Rim KT, Siaj M, Xiao SX, Myers M, Carpentier VD, Liu L, Su CC, Steigerwald ML, Hybertsen MS, McBreen PH, Flynn GW, Nuckolls C. Forming aromatic hemispheres on transition-metal surfaces. Angewandte Chemie, International Edition. 2007;46:7891-7895. DOI: 10.1002/anie.200701117
  35. 35. Otero G, Biddau G, Sanchez-Sanchez C, Caillard R, Lopez MF, Rogero C, Palomares FJ, Cabello N, Basanta MA, Ortega J, Mendez J, Echavarren AM, Perez R, Gomez-Lor B, Martin-Gago JA. Fullerenes from aromatic precursors by surface-catalysed cyclodehydrogenation. Nature. 2008;454:865-868. DOI: 10.1038/nature07193
  36. 36. Amsharov KY, Abdurakhmanova N, Stepanow S, Rauschenbach S, Jansen M, Kern K. Towards the isomer-specific synthesis of higher fullerenes and buckybowls by the surface-catalyzed cyclodehydrogenation of aromatic precursors. Angewandte Chemie, International Edition. 2010;49:9392-9396. DOI: 10.1002/anie.201005000
  37. 37. Sanchez-Valencia JR, Dienel T, Groning O, Shorubalko I, Mueller A, Jansen M, Amsharov K, Ruffieux P, Fasel R. Controlled synthesis of single-chirality carbon nanotubes. Nature. 2014;512:61-64. DOI: 10.1038/nature13607
  38. 38. Abdurakhmanova N, Mueller A, Stepanow S, Rauschenbach S, Jansen M, Kern K, Amsharov KY. Bottom up fabrication of (9,0) zigzag and (6,6) armchair carbon nanotube end-caps on the Rh(111) surface. Carbon. 2015;84:444-447. DOI: 10.1016/j.carbon.2014.12.038
  39. 39. Amsharov KY. Cyclodehydrofluorination of fluoroarenes on metal oxides. Towards bottom-up synthesis of carbon nanostructures on insulating surfaces. Physica Status Solidi B: Basic Solid State Physics. 2016;253:2473-2477. DOI: 10.1002/pssb.201600211
  40. 40. Scott LT. Methods for the chemical synthesis of fullerene. Angewandte Chemie, International Edition. 2004;43:4994-5007. DOI: 10.1002/anie.200400661
  41. 41. Tsefrikas VM, Scott LT. Geodesic polyarenes by flash vacuum pyrolysis. Chemical Reviews. 2006;106:4868-4884. DOI: 10.1021/cr050553y
  42. 42. Petrukhina MA, Scott LT. Fragments of Fullerenes and Carbon Nanotubes. Hoboken, NJ, USA: John Wiley & Sons, Inc.; 2012. 440 pp. DOI: 10.1002/9781118011263
  43. 43. Scott LT, Boorum MM, McMahon BJ, Hagen S, Mack J, Blank J, Wegner H, de Meijere A. A rational chemical synthesis of C60. Science. 2002;295:1500-1502. DOI: 10.1126/science.1068427
  44. 44. Amsharov KY, Jansen M. A C78 fullerene precursor: Toward the direct synthesis of higher fullerenes. The Journal of Organic Chemistry. 2008;73(7):2931-2934. DOI: 10.1021/jo7027008
  45. 45. Amsharov KY, Jansen M. Synthesis of a higher fullerene precursor—An “unrolled” C84 fullerene. Chemical Communications. 2009;(19):2691-2693. DOI: 10.1039/B901496A
  46. 46. Amsharov KY, Kabdulov MA, Jansen M. Homo-elimination of HF—An efficient approach for intramolecular aryl-aryl coupling. Chemistry--A European Journal. 2010;16:5868-5871. DOI: 10.1002/chem.201000374
  47. 47. Kabdulov M, Jansen M, Amsharov KY. Bottom-up C60 fullerene construction from a fluorinated C60H21F9 precursor by laser-induced tandem cyclization. Chemistry--A European Journal. 2013;19:17262-17266. DOI: 10.1002/chem.201303838
  48. 48. Greisch JF, Amsharov KY, Weippert J, Weis P, Böttcher A, Kappes MM. From planar to cage in 15 easy steps: Resolving the C60H21F9 → C60-transformation by ion mobility mass spectrometry. Journal of the American Chemical Society. 2016;138:11254-11263. DOI: 10.1021/jacs.6b06205
  49. 49. Amsharov KY. Towards rational synthesis of non-IPR fullerenes via tandem cyclodehydrofluorination of fluoroarene-based precursors. Physica Status Solidi B: Basic Solid State Physics. 2017;254:1-5. DOI: 10.1002/pssb.201700170

Written By

Konstantin Amsharov

Submitted: 05 October 2017 Reviewed: 19 December 2017 Published: 07 February 2018