Next Article in Journal
Evaluation of the Composite Mechanism of Nano-Fe2O3/Asphalt Based on Molecular Simulation and Experiments
Previous Article in Journal
Concrete Modular Pavement Structures with Optimized Thickness Based on Characteristics of High Performance Concrete Mixtures with Fibers and Silica Fume
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Capture of Fullerenes in Cages and Rings by Forming Metal-π Bond Arene Interactions

1
Instituto de Investigaciones en Materiales, Universidad Nacional Autónoma de México, Circuito Exterior s/n, Ciudad Universitaria, Coyoacán, México City 04510, Mexico
2
Facultad de Ciencias, Universidad Nacional Autónoma de México, Circuito Exterior s/n, Ciudad Universitaria, Coyoacán, México City 04510, Mexico
*
Author to whom correspondence should be addressed.
Materials 2021, 14(12), 3424; https://doi.org/10.3390/ma14123424
Submission received: 29 April 2021 / Revised: 4 June 2021 / Accepted: 11 June 2021 / Published: 21 June 2021
(This article belongs to the Special Issue Theoretical Chemistry of Fullerenes and Related Materials)

Abstract

:
Nowadays, the task of the selectively capture of fullerene molecules from soot is the subject of several studies. The low solubility of fullerenes represents a drawback when the goal is to purify them and to carry out chemical procedures where they participate. There are different molecules that can act as a kind of cocoon, giving shelter to the fullerene cages in such a way that they can be included in a solution or can be extracted from a mix. In this work, a theoretical study of some known and new proposed organic molecules of this kind is presented. In all cases, the interaction occurs with the help of a metallic atom or ion which plays the role of a bridge, providing a place for a metallocene like interaction to occur. The thermodynamic arguments favoring the formation of this adduct species are addressed as well as the nature of the bond by means QTAIM parameters and frontier molecular orbitals analysis.

1. Introduction

Fullerenes have invaded several fields of science and technology. The first report of a carbon nanostructure was C60 done by Kroto and his coworkers in 1985 [1] and it represented in those days one of the major surprises in the history of chemistry due to its unique properties, such as high surface to volume ratio, as well as high thermal and chemical stability, they can easily either accept or donate electrons and they show a singular symmetric structure [2]. These features make fullerenes attractive for different applications not only in chemistry but also in engineering, biology, physics and even medicine and its industrial production has been pursued [3,4,5,6,7,8,9]. In this respect, Murayama and coworkers [2] proposed a process to obtain fullerenes in large quantities that can be considered as an industrial product. However, the result is a heterogeneous mass, 80% of soot and 20% of a toluene soluble mixture of C60 (60%), C70 (25%) and large fullerenes (15%). The poor solubility of fullerenes in different solvents is still a drawback [10,11] and the purification of these species has been performed by instrumental exhaustive methods in lab [12], however there is a strong tendency to find chemical methods to overcome this challenge [13] an example is the functionalization of the cage with soluble chains [14]. Another approach is not to solve fullerene, itself but to transport it inside other larger chemical species to deliver the cages into a determined environment. Those chemical species should be molecular hosts with certain properties: (a) they should be wide enough to support the presence of a fullerene in their cage; (b) they should have some electronic characteristics that allow the interaction with the outer wall of fullerene; and (c) they should be soluble and stable in the chosen solvent.
There is an excellent review in which several metallosupramolecular receptors have been proposed as fullerene hosts [15]. This report recovers information of supramolecular systems that can allocate and release fullerene species based on weak interatomic forces. The encapsulation and immobilization of fullerenes has received much attention because it can enhance their properties to develop specific properties. In this sense, the confinement of fullerenes has led to structures [16] with different features, such as photosensitization [17]. They can be used to tune conductivity of MOFs [18], to develop supramolecular architectures [19] and electric and magnetic nano switches [20]. The present work deals in part with this problem, the long-range interaction of fullerene with different systems carrying electron rich centers is theoretically analyzed. Another approach to this feature would be a strong interaction between fullerene and the molecular trap. In this case, the formation of almost genuine covalent bonds is proposed. However, in this case, the species will not be directly joined, but they will contain an intermediate transition metal atom to form covalent coordinated bonds.
The obvious disadvantage of this kind of compound is the difficult isolation of fullerene which makes a chemical separation necessary. On the other hand, a huge advantage is that fullerene would acquire all the electronic characteristics and will be able to provide them in the new compounds and indeed fullerene would be carried by the MOF into crystals, solvents, etc., giving place to a material with special electronic characteristics.
The molecular traps under study in this contribution have the common characteristic of being large, covalent, organic frameworks [21]. They show a polyhedral or polygonal shape containing aromatic rings which will work as the electronic traps for capturing either the fullerene cages or the intermediate transition metal atom. Some of those molecules have similar analogs which have been already prepared, as will be indicated below, but there is at least one case of a totally local designed species.

2. Methods

All calculations were carried out by applying a DFT method based on the combination of Becke’s gradient corrections [22] for exchange and Perdew-Wang’s for correlation [23]. This is the scheme for the B3PW91 method which forms part of the Gaussian16 [24] package. This functional was chosen because it has been demonstrated that it is an excellent option to work with systems in which π bonds of organometallic species are involved [25,26,27]. The calculations were performed using the 6–31G** basis set. In order to confirm that the optimized structures were at a minimum of the potential surfaces, frequency calculations were carried out at the same level of theory. Grimme’s empirical dispersion corrections (G-3) used for evaluating the hydrogen bond energy were carried out using the DFT-D3 method [28] and single point PBE [29] calculations.
Quantum theory of atom-in-molecules analysis [30,31] was executed using Multiwfn 3.7 software [32] to assess the bonding type between the transition metal atoms with the coronene unit on the one hand, and with the fullerene on the other hand.

3. Results and Discussion

The three molecular traps are shown in Figure 1.
Molecule a is that prepared by Beuerle and his coworkers [33] without changes, whereas molecule b is a domestic design of a square framework with coronene fragments in the edges. Molecule c is inspired by a macrocycle synthesized by Bain and coworkers [34]. However, the system proposed in this work holds coronene units instead of the porphyrinic ones of the original.
Therefore, there are two ways to trap fullerene cages. Firstly, there is the simple interaction by dispersion forces, mainly π-stacking, which compels the fullerene to make a long-distance interaction with the aromatic fragment of the supramolecular species. Secondly, fullerene is captured in an organometallic complex through a transition metal center and the same aromatic fragment. These two cases will be presented in the next sections.

3.1. Π-Stacking

One of the most popular intermolecular interactions that has been studied is the so-called π-stacking [35,36]. This phenomenon is defined as the long-distance electronic communication between the π orbitals of two aromatic species through the space. It is important in supramolecular chemistry [33] because two fragments which are attracted by this force can join to form a more elaborated structure. These interactions can be quantified by computational methods as the consideration of dispersion corrections [37]. The mentioned strategy was adopted in this work.
In all the cases, the dispersion energy results suggest that there is some interaction between fullerene and the molecular trap. The fullerene complexes show 33.9, 35.7, and 36.3 Kcal/mol dispersion energies for molecules a, b, and c, respectively. Besides, the nearest distance between a six-membered ring from the fullerene surface and its counterpart in the trap molecule is 5.1 Å on average. This value matches well with the experimental result of 8.09 from an aromatic ring to the geometric center of the sphere of fullerene [38]
Molecule c belongs to the Oh point group in a static image. Its molecular orbital shapes are shown in Figure 2. It is important to highlight that the LUMO is actually a four-folded degenerated set, whereas the HOMO is a larger five folded degenerated set. This strong electronic accumulation arises from the generation of two sets of accidental degeneration, where eg, eu, and t2g as irreducible representations come together to form both sets.
The first interesting result is focused on the degenerated set corresponding to the LUMO. Practically all the probability is found on the aromatic rings which lie on the edges, so these regions will accept electrons from other species in search for a place to form a bond.
The HOMO in other context is a set of more diffuse functions. It is found in all the edges with no preference for a special aromatic ring.
The inclusion of fullerene, (see Figure 3) does not cause strong modifications, that is to say the fullerene molecule reaches the inner of the cubic cage without electronic changes in it, nor in the fullerene itself. Therefore, the sets HOMO and LUMO belong to the fullerene alone and the nearest probability function to trap the fullerene is the LUMO + 6, although it is relatively near in terms of energy, with a difference of 4.3 eV from the HOMO.
In the case of the compound c, there is not a defined spatial group. However, all the probabilities in virtual as well as occupied orbitals are centered on the coronene fragments. Therefore, these regions are expected to either accept or donate electrons depending on the species acting as the counterparts of the reaction. This behavior is almost the same in compound b, which also bears coronene units as catchers. An important fact is that, in both cases, the molecules show accidental degeneracy on the HOMO. The cause of this situation is not symmetry, but the periodical arrangements which establish similar coronene units with the same energy and behavior in the chain of the molecule. This fact allows to suggest that each molecular trap can capture several fullerene molecules.
Interestingly the presence of fullerene does not contribute to avoid this property. In fact, the complex fullerene-compound c (or b) shows a molecular orbital diagram with full participation of the HOMO and LUMO combinations of both counterparts, but the important point is that fullerene can also be easily trapped in this jail. The arrangement of the trapped fullerene in the hollow of compound c is shown in Figure 4.

3.2. Formation of Metallocenes

In this case, there is a third participant in the interaction, a transition metal ion. Therefore, the molecular trap again captures fullerene but not directly. Instead, the capture is achieved through the metal atom which plays the role of a bridge. Indeed, the molecular trap follows a similar process to the one analyzed before, but now the captured species is an exohedral C60M.
The capability of fullerene to bond transition metal atoms has been object of study of several groups [39,40,41]. There are examples of six different hapticities [41], although η5 [42,43] and η2 [44] are the most common cases. The η6 case is difficult to find. However, there are some experimental [39] and theoretical [45,46,47,48,49] examples to acknowledge. In the present work, two possibilities are considered: chromium η6 and Nickel η2 complexes.
Figure 5 shows the shape of the complex arising from the interaction between C60Cr fragment and the cubic compound a. The corresponding molecular orbital shapes are presented in Figure 6. The most remarkable is that the fragment C60M exhibits the behavior of an isolated species because its frontier molecular orbitals are practically transferred to the resultant complex with little change in energy.
The behavior of the C60M-molecule c is something different because the ring contributes with the HOMO (which is accidentally degenerated and shares a function on each coronene ring in every MO, including the one supporting the fullerene), whereas the fullerene metal complex contributes with the LUMO to produce the electronic interaction. The strong difference in this case is that the trapping is now generated by the formation of a covalent bond (see Figure 7).
The Hirschfield charges analysis shows there are interesting negative values in the fullerene surface. The values are 0.251 in the case of the chromium compound and 0.260 for the nickel one. This fact can be considered together the Wiberg bond order results, in which case the results for the coordination bond between chromium atom and the aromatic ring on the MOF is 1.63, whereas the same value for the bond between the fullerene surface and the chromium atom is 1.58. The same results for the nickel complex are 1.75 and 1.70. Such evidence would suggest there is enough charge transfer to give place to strong covalent coordinated bonds.
Some interesting differences are observed between both C60M-molecule b compounds induced by the Ni and Cr hapticities, e.g., in the case of η2-C60Ni, the HOMO (a degenerated set) is localized on the edges of the square where transition metal atoms are placed with their coordination sphere (see Figure 8), while the interaction between fullerene and coronene fragments is found at HOMO-4 (the immediate function after the degenerated set of HOMO). In constrast, in η6-C60Cr, this interaction is found at HOMO and HOMO-2 (see Figure 9), which are quasi-degenerated too (−5.130 and −5.155 eV respectively). In both cases, LUMO is localized again on the coordination complexes of the edges (Figure 8 and Figure 9). Since the OM analysis confirms the interaction between fullerene and coronene unit via transition metal atoms, the question now is what kind of interaction is this? Looking at the QTAIM parameters (see Table 1), it can be seen that in all cases the electronic density (ρ) and Laplacian ( 2 ρ ) at the bond critical point on the fullerene-M and M-coronene bond paths are ρ < 0.10   e/bohr3 and 2 ρ > 0 . Therefore, all bonds correspond to close-shell interactions. Following the classification by Nakanishi et al. [50,51], two categories of close-shell bonding can be identified based on the calculated value for the |V(r)|/G(r) ratio (which reflects the magnitude of the covalent interaction), being V(r) the potential energy density and G(r) the kinetic energy density: if | V ( r ) | / G ( r ) < 1 , no covalence is expected and the so called pure close-shell interaction is obtained (ionic or van der Waals bonding), if | V ( r ) | / G ( r ) > 1 the interaction is called regular close-shell, and some degree of covalence is expected (e.g., molecular complex through charge transfer). In Table 1, in all cases | V ( r ) | / G ( r ) > 1 , therefore the fullerene-M and M-coronene bonds are regular close-shell interactions, with a small covalence contribution [51,52].
Thus, from the above analysis, we can conclude that these compounds can work as molecular traps of fullerenes. However, its main activity consists in the development of a kind of electrical flux around its perimeter.
An additional feature to note is that the energy gaps of these species (containing the trapped fullerene) are 1.143 and 1.113 eV, respectively. These values allow to classify these substances as good semiconductors or even soft conductors. Therefore, the cages would be useful to trap as well as to generate electronic interesting species.

4. Conclusions

The designed supramolecular cages can trap fullerene C60 either by π-π stacking interaction or by the formation of coordinated covalent bonds. The cages belong to the MOP family of compounds and have edges that contain strong aromatic centers which constitute the main trap regions because of its large electron clusters. The idea is to achieve the formation of links that can transform weak bonds to real covalent coordinate bonds. The nature of the covalent interaction was studied by means the QTAIM method and it was found that the bonds between the metal atom and fullerene or coronene fragments correspond to a regular close-shell interaction, with a small covalence contribution following the Nakanishi classification. Therefore, the studied species show that real weak covalent bonds and the capture of fullerene units are achieved.

Author Contributions

Conceptualization, investigation, methodology, validation and writing—review and editing, C.R., B.M., and R.S.; writing—original draft preparation and formal analysis, B.M. and R.S.; supervision and administration, R.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

Authors would like to acknowledge Oralia L Jiménez A., María Teresa Vázquez, Alejandro Pompa, Alberto López-Vivas and Caín González for their technical support. B.M. acknowledges the Dirección General de Cómputo y de Tecnologías de la Información (DGTIC-UNAM) for the computer facilities on “Miztli.” Project number CAD-UNAM-DGTIC-298.

Conflicts of Interest

The authors declare no conflict of interest.

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. [Google Scholar] [CrossRef]
  2. Murayama, H.; Tomonoh, S.; Alford, J.M.; Karpuk, M.E. Fullerene production in tons and more: From science to industry. Fuller. Nanotub. Carbon Nanostructures 2004, 12, 1–9. [Google Scholar] [CrossRef]
  3. Delgado, J.R.; Fillipone, S.; Giacolone, F.; Herranz, M.A.; Illescas, B.; Perez, E.M.; Martin, N. Bucckyballs, Polyarenes II. In Topics in Current Chemistry; Siegel, J.S., Wu, Y.T., Eds.; Springer: Berlin, Germany, 2014; Volume 350, pp. 1–64. [Google Scholar]
  4. Acquah, S.F.A.; Penkova, A.V.; Markelov, D.A.; Semisalova, A.S.; Leonhardt, B.E.; Magi, J.M. Review—The beautiful molecule: 30 years of C60 and its derivatives. ECS J. Sol. State Sci. Tech. 2017, 6, M3155–M3162. [Google Scholar] [CrossRef]
  5. Ganesamoorthy, R.; Sathiyan, G.; Sakthivel, P. Review: Fullerene based acceptors for efficient bulk heterojunction organic solar cells applications. Sol. Energy Mater. Sol. Cell 2017, 161, 102–148. [Google Scholar] [CrossRef]
  6. Coro, J.; Suarez, M.; Silva, L.S.R.; Eguiluz, K.I.B.; Salazar-Banda, G.R. Fullerene applications in fuel cells: A review. Int. J. Hydrog. Energy 2016, 41, 17944–17959. [Google Scholar] [CrossRef]
  7. Gupta, V.K.; Saleh, T.A. Sorption of pollutants by porous carbon, carbon nanotubes and fullerene—An overview. Environ. Sci. Pollut. Res. 2013, 20, 2828–2843. [Google Scholar] [CrossRef] [PubMed]
  8. Mousavi, S.Z.; Nafisi, S.; Maibach, H.I. Fullerene nanoparticle in dermatological and cosmetic applications. Nanomed. Nanotech. Biol. Med. 2017, 13, 1071–1087. [Google Scholar] [CrossRef] [PubMed]
  9. Castro, E.; Hernández-García, A.; Zavala, G.; Echegoyen, L. Fullerenes in biology and medicine. J. Mater. Chem. B 2017, 5, 6523–6535. [Google Scholar] [CrossRef]
  10. Ruoff, R.S.; Tse, D.S.; Malhotra, R.; Lorents, D.C. Solubility of fullerene (C60) in a variety of solvents. J. Phys. Chem. 1993, 97, 3379–3383. [Google Scholar] [CrossRef]
  11. Sivaraman, N.; Dhamodaran, R.; Kaliappan, I.; Srinivasan, T.G.; Vasudeva Rao, P.R.; Mathews, C.K. Solubility of C60 in organic solvents. J. Org. Chem. 1992, 57, 6077–6079. [Google Scholar] [CrossRef]
  12. Astefanei, A.; Nunez, O.; Galceran, M.T. Characterization and determination of fullerenes; a critical review. Anal. Chim. Acta 2015, 882, 1–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. García-Simón, C.; García-Borras, M.; Gómez, L.; Parella, T.; Osuna, S.; Juanhuix, J.; Imaz, I.; Maspoch, D.; Costas, M. Sponge-like molecular cage for purification of fullerenes. Nat. Comm. 2014, 5, 1–9. [Google Scholar] [CrossRef]
  14. 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. [Google Scholar] [CrossRef]
  15. García-Simón, C.; Costas, M.; Ribas, X. Metallosupraamolecular receptors for fullerene binding and release. Chem. Soc. Rev. 2015, 45, 40–62. [Google Scholar] [CrossRef]
  16. Ubasart, E.; Borodin, O.; Fuertes-Espinosa, C.; Xu, Y.; García-Simón, C.; Gómez, L.; Juanhuix, J.; Gándara, F.; Imaz, I.; Maspoch, D.; et al. A three-shell supramolecular complex enables the symmetry-mismatched chemo- and regioselective bis-functionalization of C60. Nat. Chem. 2021, 13, 420–427. [Google Scholar] [CrossRef]
  17. Purba, P.C.; Maity, M.; Bhatacharyya, S.; Mukherjee, P.S. A self-assembled palladium(II) barrel for binding of fullerenes and photosensitization ability of the fullerene-encapsulated barrel. Angew. Chem. 2021. [Google Scholar] [CrossRef]
  18. Ray, D.; Goswami, S.; Duan, J.; Hupp, J.T.; Cramer, C.J.; Gagliardi, L. Tuning the conductivity of hexa-zirconium(IV) metal–organic frameworks by encapsulating heterofullerenes. Chem. Mater. 2021, 33, 1182–1189. [Google Scholar] [CrossRef]
  19. Zhang, J.; Zhao, C.; Meng, H.; Nie, M.; Li, Q.; Xiang, J.; Zhang, Z.; Wang, C.; Wang, T. Size-selective encapsulation of metallofullerenes by cycloparaphenylene and dissociation using metal-organic framework. Carbon 2020, 161, 694–701. [Google Scholar] [CrossRef]
  20. Villalva, J.; Develioglu, A.; Montenegro-Pholhammer, N.; Sánchez de Armas, R.; Gamonal, A.; Rial, E.; García-Hernández, M.; Ruiz-González, L.; Sánchez-Costa, J.; Calzado, C.J.; et al. Spin-state-dependent electrical conductivity in single-walled carbon nanotubes encapsulating spin-crossover molecules. Nat. Commun. 2021, 12, 1578–1585. [Google Scholar] [CrossRef]
  21. Beuerle, F.; Gole, B. Covalent organic frameworks and cage compounds: Design and applications of polymeric and discrete organic scaffolds. Angew. Chem. 2018, 57, 4850–4878. [Google Scholar] [CrossRef]
  22. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  23. Perdew, J.P.; Wang, Y. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B 1992, 45, 13244–13249. [Google Scholar] [CrossRef]
  24. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scal-mani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian16, Revision A.03; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  25. Sumimoto, M.; Kawashima, Y.; Hori, K.; Fujimoto, H. Theoretical study on the stability of double-decker type metal phthalocyanines, M(Pc)2 and M(Pc)2+ (M = Ti, Sn and Sc): A critical assessment on the performance of density functionals. Phys. Chem. Chem. Phys. 2015, 17, 6478–6483. [Google Scholar] [CrossRef] [PubMed]
  26. Masuda, J.D.; Boeré, R.T. Metal carbonyl complexes of phosphaamidines. Coordinative integrity detected in C-amino(λ3,σ2)-phosphaalkene isomers coordinated through n(P) HOMO-1 donor orbitals. Dalton Trans. 2016, 45, 2102–2115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Coimbra, D.F.; Ortolan, A.O.; Orenha, R.P.; da Silva, V.B.; Parreira, R.L.T.; Caramori, G.F. Shedding light on the bonding situation of triangular and square heterometallic clusters: Computational insight. New J. Chem. 2020, 44, 5079–5087. [Google Scholar] [CrossRef]
  28. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H.J. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [Green Version]
  29. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Bader, R.F.W.; Nguyen-Dang, T.T.; Tal, Y. A topological theory of molecular structure. Rep. Prog. Phys. 1981, 44, 894–948. [Google Scholar] [CrossRef]
  31. Matta, C.F.; Boyd, J.R. The Quantum Theory of Atoms in Molecules. From Solid State to DNA and Drug Design; Chapter 3; Matta, C.F., Boyd, R.J., Eds.; Wiley-VCH: Weinheim, Germany, 2007. [Google Scholar]
  32. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  33. Klotzbach, S.; Scherpf, T.; Beuerle, F. Dynamic covalent assembly of tribenzotriquinacenes into molecular cubes. Chem. Commun. 2014, 50, 12454–12457. [Google Scholar] [CrossRef]
  34. Calik, M.; Auras, F.; Salonen, L.M.; Bader, K.; Grill, I.; Handloser, M.; Medina, D.D.; Dogru, M.; Löbermann, F.; Trauner, D.; et al. Extraction of photogenerated electrons and holes from a covalent organic framework integrated heterojunction. J. Am. Chem. Soc. 2014, 136, 17802–17807. [Google Scholar] [CrossRef]
  35. Chipot, C.; Jaffe, R.; Maigret, B.; Pearlman, D.A.; Kollman, P.A. Benzene dimer: A good model for π-π interactions in proteins? A comparison between the benzene and the toluene dimmers in the gas phase and in an aqueous solution. J. Am. Chem. Soc. 1996, 118, 11217–11224. [Google Scholar] [CrossRef]
  36. Chen, T.; Li, M.; Liu, J. π-π stacking interaction: A nondestructive and facile means in material engineering for bioaplications. Cryst. Growth Des. 2018, 18, 2765–2783. [Google Scholar] [CrossRef]
  37. Grimme, S.; Antony, J.; Schwabe, T.; Muck-Lichtenfeld, C. Density functional theory with dispersion corrections for supramolecular structures, aggregates and complexes of (bio) organic molecules. Org. Biomol. Chem. 2007, 5, 741–758. [Google Scholar] [CrossRef]
  38. Henne, S.; Bredenkötter, B.; Dehghan-Baghi, A.A.; Schmid, R.; Volkmer, D. Advanced buckyball joints: Synthesis, complex formation and computational simulations of centrohexaindane-extended tribenzotriquinacene receptors for C60 fullerene. Dalton Trans. 2012, 41, 5995–6002. [Google Scholar] [CrossRef] [PubMed]
  39. Balch, A.L.; Olmstead, M.M. Reactions of transition metal complexes with Fullerenes (C60, C70, etc.) and related materials. Chem. Rev. 1998, 98, 2123–2165. [Google Scholar] [CrossRef] [PubMed]
  40. Soto, D.; Salcedo, R. Coordination modes and different hapticities for fullerene organometallic complexes. Molecules 2012, 17, 7151–7168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Axet, M.R.; Dechy-Cabaret, O.; Durand, J.; Gouygou, M.; Serp, P. Coordination chemistry on carbon surfaces. Coord. Chem. Rev. 2016, 308, 236–345. [Google Scholar] [CrossRef]
  42. Nakamura, E.; Sawamura, M. Chemistry of η5-fullerene metal complexes. Pure Appl. Chem. 2001, 73, 355–359. [Google Scholar] [CrossRef] [Green Version]
  43. Nakamura, E. Bucky ferrocene: Hybrid of ferrocene and fullerene. Pure Appl. Chem. 2003, 75, 427–434. [Google Scholar] [CrossRef] [Green Version]
  44. Sokolov, V.I. Fullerenes as a new type of ligands for transition metals. Russ. J. Coord. Chem. 2007, 33, 723–737. [Google Scholar] [CrossRef]
  45. Salcedo, R. Fullerenocene. Polyhedron 2009, 28, 431–436. [Google Scholar] [CrossRef]
  46. Molina, B.; Pérez-Manríquez, L.; Salcedo, R. Exohedral complexes of large fullerenes, a theoretical approach. J. Mol. Mod. 2017, 23, 171. [Google Scholar] [CrossRef] [PubMed]
  47. Anafche, M.; Naderi, F.; Khodadadi, Z.; Ektefa, F.; Ghafouri, R.; Zahedi, M. Computational study for the circular redox reaction of N2O with CO catalyzed by fullerometallic cations C60Fe+ and C70Fe+. J. Mol. Graph. Mod. 2017, 72, 50–57. [Google Scholar] [CrossRef]
  48. Gal’pern, E.G.; Sabirov, A.R.; Stankevich, I.P. C60 fullerene as an η6 ligand in π complexes of transition metals. Phys. Sol. State 2007, 49, 2330–2334. [Google Scholar] [CrossRef]
  49. Jemmis, E.D.; Manoharan, M.; Sharma, P.K. Exohedral η5 and η6 Transition-Metal Organometallic Complexes of C60 and C70:  A theoretical study. Organometallics 2000, 19, 1879–1887. [Google Scholar] [CrossRef]
  50. Nakanishi, W.; Hayashi, S. Role of dG/dw and dV/dw in AIM analysis: An approach to the nature of weak to strong interactions. J. Phys. Chem. A 2013, 117, 1795–1803. [Google Scholar] [CrossRef]
  51. Pilmé, J.; Renault, E.; Bassal, F.; Amaouch, M.; Montavon, G.; Galland, N. QTAIM analysis in the context of quasirelativistic quantum calculations. J. Chem. Theory Comput. 2014, 10, 4830–4841. [Google Scholar] [CrossRef]
  52. Cremer, D.; Kraka, E. Chemical bonds without bonding electron density—Does the difference electron-density analysis suffice for a description of the chemical bond? Angew. Chem. 1984, 23, 627–628. [Google Scholar] [CrossRef]
Figure 1. The molecular traps under study. (a) A molecule prepared by Beuerle and his coworkers [33]. (b) A domestic designed structure with coronene fragments in the edges. (c) A molecule inspired by a macrocycle synthesized by Bain and coworkers [34].
Figure 1. The molecular traps under study. (a) A molecule prepared by Beuerle and his coworkers [33]. (b) A domestic designed structure with coronene fragments in the edges. (c) A molecule inspired by a macrocycle synthesized by Bain and coworkers [34].
Materials 14 03424 g001
Figure 2. Frontier molecular orbitals of molecule a.
Figure 2. Frontier molecular orbitals of molecule a.
Materials 14 03424 g002
Figure 3. π-stacking interaction between fullerene and molecule a.
Figure 3. π-stacking interaction between fullerene and molecule a.
Materials 14 03424 g003
Figure 4. Fullerene molecule trapped into the hollow of molecule c.
Figure 4. Fullerene molecule trapped into the hollow of molecule c.
Materials 14 03424 g004
Figure 5. Fullerene and molecule a complex with metallocene bond.
Figure 5. Fullerene and molecule a complex with metallocene bond.
Materials 14 03424 g005
Figure 6. Frontier molecular orbitals of the complex between C60M fragment and molecule a.
Figure 6. Frontier molecular orbitals of the complex between C60M fragment and molecule a.
Materials 14 03424 g006
Figure 7. Frontier molecular orbitals of the complex between molecule c and C60M fragment.
Figure 7. Frontier molecular orbitals of the complex between molecule c and C60M fragment.
Materials 14 03424 g007
Figure 8. Frontier molecular orbitals of the complex between molecule b and C60Ni fragment.
Figure 8. Frontier molecular orbitals of the complex between molecule b and C60Ni fragment.
Materials 14 03424 g008
Figure 9. Frontier molecular orbitals of the complex between molecule b and C60Cr fragment.
Figure 9. Frontier molecular orbitals of the complex between molecule b and C60Cr fragment.
Materials 14 03424 g009
Table 1. QTAIM parameters for C60M-Molecule b compounds. Bond Critical Point (BCP), electronic density (ρ), Laplacian ( 2 ρ ), potential energy density (V(r)) and the kinetic energy density (G(r)).
Table 1. QTAIM parameters for C60M-Molecule b compounds. Bond Critical Point (BCP), electronic density (ρ), Laplacian ( 2 ρ ), potential energy density (V(r)) and the kinetic energy density (G(r)).
Bond Critical Point Numberρ
u.a.
2 ρ
u.a.
G(r)
u.a.
|V(r)|/G(r)
u.a.
Ni-CoroneneC60Ni-Molecule b
BCP10.0600.1740.0641.322
BCP20.0590.17800.0651.309
BCP30.0610.1790.0661.322
BCP40.0560.1900.0631.248
BCP50.0610.1680.0641.350
BCP60.0570.1880.0641.265
Ni-Fullerene
BCP10.1180.1970.1141.569
BCP20.1190.1990.1151.567
Cr-coroneneC60Cr-Molecule b
CP10.0630.2460.0721.143
CP20.0590.2460.0701.119
CP30.0600.2430.0691.123
CP40.0620.2450.0711.138
CP50.0620.2410.0701.142
CP60.0600.2460.0701.121
Cr-Fullerene
CP10.0540.2230.0631.114
CP20.0530.2160.0611.118
CP30.0540.2230.0631.111
CP40.0520.2120.0601.118
CP50.0530.2200.0621.113
CP60.0530.2150.0611.116
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rios, C.; Molina, B.; Salcedo, R. Capture of Fullerenes in Cages and Rings by Forming Metal-π Bond Arene Interactions. Materials 2021, 14, 3424. https://doi.org/10.3390/ma14123424

AMA Style

Rios C, Molina B, Salcedo R. Capture of Fullerenes in Cages and Rings by Forming Metal-π Bond Arene Interactions. Materials. 2021; 14(12):3424. https://doi.org/10.3390/ma14123424

Chicago/Turabian Style

Rios, Citlalli, Bertha Molina, and Roberto Salcedo. 2021. "Capture of Fullerenes in Cages and Rings by Forming Metal-π Bond Arene Interactions" Materials 14, no. 12: 3424. https://doi.org/10.3390/ma14123424

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop