Next Article in Journal
On a Relation between the Perfect Roman Domination and Perfect Domination Numbers of a Tree
Next Article in Special Issue
Wiener–Hosoya Matrix of Connected Graphs
Previous Article in Journal
Entropy Measures for Plithogenic Sets and Applications in Multi-Attribute Decision Making
Previous Article in Special Issue
On the Aα-Spectral Radii of Cactus Graphs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Skeletal Rearrangements of the C240 Fullerene: Efficient Topological Descriptors for Monitoring Stone–Wales Transformations

1
Laboratory of Mathematical Chemistry, Institute of Petrochemistry and Catalysis, Russian Academy of Sciences, 450075 Ufa, Russia
2
Actinium Chemical Research Institute, 00182 Rome, Italy
*
Author to whom correspondence should be addressed.
Mathematics 2020, 8(6), 968; https://doi.org/10.3390/math8060968
Submission received: 28 April 2020 / Revised: 27 May 2020 / Accepted: 5 June 2020 / Published: 12 June 2020
(This article belongs to the Special Issue Graph Theory at Work in Carbon Chemistry)

Abstract

:
Stone–Wales rearrangements of the fullerene surface are an uncharted field in theoretical chemistry. Here, we study them on the example of the giant icosahedral fullerene C240 to demonstrate the complex chemical mechanisms emerging on its carbon skeleton. The Stone–Wales transformations of C240 can produce the defected isomers containing heptagons, extra pentagons and other unordinary rings. Their formations have been described in terms of (i) quantum-chemically calculated energetic, molecular, and geometric parameters; and (ii) topological indices. We have found the correlations between the quantities from the two sets that point out the role of long-range topological defects in governing the formation and the chemical reactivity of fullerene molecules.

Graphical Abstract

1. Introduction

Most chemical studies on fullerenes are focused on the reactions and compounds of C60 and C70, the two most abundant fullerenes. Nevertheless, the diversity of the sizes and shapes within the fullerenes’ family is much wider. Indeed, there are many fullerene structures, smaller [1] and larger [2], which are produced in smaller amounts. The chemistry of such small and large fullerenes has growing experimental and theoretical interests.
We pay attention to the giant C240 fullerene with icosahedral symmetry. Its structure is the next after C60 in the series of icosahedral Goldberg polyhedra [3,4,5] having 60a2 vertices (where a is the integer). In the fundamental aspect, such chemical structures are very close to their mathematical models, so direct bridges between the chemistry and mathematics could be built, which means the possible correlations between the structural and topological (mathematical) indices and molecular (chemical) properties of the molecules [4,5,6]. It is noteworthy that the C240 structure obtained with quantum chemical calculations corresponds to the idealized one, i.e., the highest symmetry (Ih) allowed by the topology of the molecule (there is a possibility of the Jahn–Teller symmetry reduction in the case of other icosahedral fullerenes, e.g., C20 and C80) [6]. In practical aspects, such giant cages as C240 attract as they have a large size and may be classified as nano-objects, but they are still molecules. Due to their size, giant fullerenes are considered as nano-vessels for gas storage [7] and the closed structure makes C240 a molecular Faraday cage screening the guest atoms/molecules from the external electric fields [8]. Note that C240 (Ih) has not been yet synthesized nor isolated. Under arc-discharge fullerene synthesis, it should be produced in extremely minute quantities (it is also potentially insoluble compound due to its size; this makes an obstacle of its extraction) [9,10]. However, its complexes like C60@C240, C240@C560, and C60@C240@C560 have been obtained and visualized with atom-force microscopy [11,12,13,14]. As follows from these studies, the C240 molecule is most likely icosahedral. The thermodynamic [5,15,16] and dielectric properties [8,17,18,19] of C240 and other giant fullerenes have been studied with relevant computational techniques.
Giant fullerenes relate to the processes of classical fullerene formation [20]. They are proposedly formed from the C2 gas and convert to C60 and C70 [21]. The presence of a large cavity (much larger than in C60) makes giant fullerenes chemically unstable [21] and sensitive to various transformations. The Stone–Wales isomerization is one of such topological processes that may occur on the fullerene surface [22,23]. Since 1986 [24], it is well-known that Stone–Wales (SW) rotations have a theoretical role in connecting the isomers of a given Cn fullerene.
Under the action of the SW5|6 operator (Figure 1a), the 1812 isomers of C60 are grouped in 13 sets. The larger one includes 1709 cages connected to C60 (Ih) through one or more SW5|6 transformations whereas 31 isomers remain separated from the rest of the isomers [24]. Later, the introduction of the so-called generalized Stone-Wales transformations (Figure 1b) overcame this limitation offering a method of generating the complete C60 isomeric space starting from just one C60 isomers [25]. The instrumental role of chemical graphs in studying fullerene rearrangements was immediately evident [25]. The description of fullerene molecule Cn as a planar graph G made by n nodes (carbon atoms), B = 3 n / 2 edges (carbon–carbon bonds) forming f5 = 12 pentagons and f6 = n/2 – 10 hexagons is a convenient and powerful tool that is also suitable for computerized algorithms. Chemical graph G represents the topological skeleton of the molecule and allows a clear picture of the Stone–Wales isomerization mechanism. In Figure 1a, the SW5|6 transformation (the so-called pyracylene rearrangement) that is capable of bridging Cn isomers with different symmetries is shown in the direct (bottom) and dual graph (top). In G the rotation of the central bond is obtained by cutting the two graph edges indicated by the red arrows and then by connecting, pairwise, the green and the orange nodes. This topological transformation is perfectly isomeric since the numbers of nodes, edges, and faces are conserved. The generation of the SW5|6 rotation became even simpler when occurring on the dual fullerene graph G ¯ composed by n ¯ = n 2 + 2 dual nodes, which correspond to the faces of the direct Cn fullerene and B = 3 n / 2 edges. Both fullerene graphs, direct G and dual G ¯ , have the same number of edges B . The top part of Figure 1a illustrates the pyracylene transformation in the dual graph G ¯ . In this case the SW5|6 rearrangement is obtained just by rotating the central edge (red arrow) evidencing a general feature of the SW transformations the dual space G ¯ :
  • In the dual space, a SW rotation consists in the reversible rotation of just one dual bond.
This straightforward mechanism allows the generation of extended topological defects, so-called generalized Stone-Wales transformations (gSW) that modify the fullerene surfaces in a peculiar fashion.
Figure 1b shows the typical gSW involving two 5|6 pairs (shown in Figure 1a) and η pairs of the internal faces (white circles) included between those 5|6. The white circles represent indifferently hexagons or pentagons. The surface transformation of the dual fullerene G ¯     (Figure 1b) corresponds therefore to the gSWη=3 of size η = 3 whereas the generalized rotation gSWη = 0 of order η = 0 represents instead to the pyracylene rearrangement SW5|6. Operators like gSW depend on integer parameter η (Figure 1b) and constitute only one set of the infinite class S of reversible non-local topological rearrangements. The extension (or the size) of the gSW defect is expressed by integer η.
  • The gSW size η corresponds to the number of pairs of internal faces (white circles) included between the two 5|6 pairs.
All SW operators represent full-isomeric rearrangements of the fullerene surface, which therefore preserve the numbers of direct nodes n, dual nodes n ¯ , bonds B and faces. They also preserve the type of internal faces included in a dislocation dipole. Figure 1 gives nice examples of these topological modifications.
In S another interesting group of operators is represented by the specific transformations that generate and then propagate 5|7 pairs in a hexagonal network [26,27]. In a regular hexagonal mesh, this original mechanism produces a peculiar linear extended defect, the so-called SW wave (SWw) with variable length η. Figure 2 describes in the dual graph the sequence of specific SW operators that create (SW6|6) and move (SW6|7) that wave in a graphenic lattice. The first rotation SW6|6 (Figure 2a) of the arrowed dual edge creates the 5|7 double pairs (Figure 2b). The second rotation SW6|7 separates them by the insertion of a couple of hexagons (Figure 2c). At this point, the propagation of the 5|7 pair has begun and successive SW6|7 transformations propagate it in the lattice producing the topological SW wave. This topological defect SWw is known in the literature also under the name of dislocation dipole, whose length η corresponds to the minimal number of hexagon pairs between the 5|7 pairs. Figure 2 shows the dislocation dipoles (SWw) with η = 0, 1, and 2.
Theoretically, the existence of extended topological defects such as SWw with η ≥ 3 has been documented by molecular dynamics and ab initio studies [28,29] reporting about the presence of 5|7 dislocations in pristine graphene and at the grain boundaries of polycrystalline graphenic lattices. In the latter case, the authors state that these linear defects cannot be annealed by local reorganization of the honeycomb mesh. Specifically, work [28] affirmed that “neither an isolated pentagon, or a heptagon, or their pair 5|7” cannot be corrected by lattice rearrangements – an erroneous conclusion not considering that the SWw mechanism allows the 5|7 dislocations to be annealed by moving the 5|7 pair backward by applying the SW transformations (Figure 2) in the reverse direction.
Electronic properties of honeycomb planar systems indicate that single heptagon–pentagon dislocations are stable defects whereas 5|7 adjacent pairs are dynamically unstable [29]. The energy Ef of the formation of the 5|7 double pairs (Figure 2b) determined by the extended Hückel [30] and molecular mechanical simulations [31] depends on multiple structural factors:
(i) the size of the honeycomb flat lattice around the defect, with Ef that assumes the minimum value Ef ≅ 6 eV for “hexagonal layers” made by 9,000 carbon atoms [31];
(ii) the presence a cylindrical curvature in the honeycomb mesh that facilitates furthermore the creation of the dislocation dipole by lowering the energy barrier to Ef ≅ 3 eV for carbon nanotubes. Formation energy of the 5|7 double pair drops even more, down to Ef ≅ 2 eV, when interstitial defects or ad-atoms are present in the networks [32,33].
Pentagon–heptagon pairs have been moreover detected by high-resolution TEM studies [34,35,36,37] whereby energetic particles, such as electrons and ions, generate 5|7 pairs in graphite layers or single-walled CNTs because of atom displacements.
We consider that both appearing elements favor the creation of dislocation dipoles, i.e., curvature and presence of hexagonal regions of variable size, are specific characteristics of fullerenes’ morphology. Therefore, the aim of this work is addressed, mainly from a topological corner, to the following basal theoretical questions:
  • In which ways is the fullerene surface modified by SWw topological defects?
  • Is the creation of SWw defects energetically favored?
In the following, the first attempt for answering the above points is made by taking into account the C240 (Ih) molecule as a case study. The reasons for this choice are the high symmetry of the molecule and the previous of well-documented ab initio [16] and topological studies [27] about its stability. The next paragraph is devoted to the description of the mechanisms for the formation of dislocation dipoles on the C240 (Ih) surface; then the energetic and topological considerations complete this work.

2. SW Waves on the C240 (Ih) Fullerene

The theoretical evidences of the stability of 5|7 defects in various hexagonal systems [28,29,30,31,32,33,34,35,36] suggest the possibility to create extended defects even on the surface of a fullerene large enough to have graphene-like zones tiled only with hexagons. This section provides a concise introduction to the topological wave-like mechanism for the creation and the diffusion (or annihilation) of the pentagon–heptagon linear defect on the fullerene surface. The present investigation, combining extended topological rearrangements of the carbon networks and specific modifications of fullerene topology, sheds new light on possible mechanism that could be a base of fullerene formation and chemical properties.
Figure 3 shows the icosahedral C240 cage with f6 = 110 hexagonal rings and the symmetry-independent rearrangements of type SW6|6, which may take place on its surface (images sourced from the web [38]). Due to the icosahedral structure of the C240 molecule, one has in fact just three symmetry-independent ways A, B, and C of selecting quartets of hexagons admitting the SW6|6 rotation. The three regions A, B, and C are shown in Figure 3a prior to applying the SW operator according to the mechanisms illustrated above (Figure 2). The SW6|6 operator rotates the dual bonds pictured in black in each diamond in Figure 3a.
Our density functional theory (DFT) calculations show that, among the three isomers of the C240 (Ih) fullerene generated by the SW6|6 operator applied to the quartets A, B, and C (Figure 3a), the energy effect favors the rotation of the quartet A. Indeed, the calculated ΔE values for modes A, B, and C equal 147.9, 323.5, and 347.4 kJ/mol (see Appendix A for the details of the calculations). Note that the structures obtained via the rotations of quartets A and B have the CS symmetry whereas the structure derived by the quartet C rearrangement has no symmetry. The configuration of the molecule after the SW6|7 rotation of quartet A is given in Figure 3b, which evidences the new 5|7 double pair. At this point, the SW wave with η = 0 has been created on the surface of the new isomer C 240 η = 0 with f5 = 14, f6 = 106, and f7 = 7. The two extra pentagons and the two heptagons are labeled with the number of respective edges in Figure 3b. According to the propagation mechanism (Figure 2c,d), the second SW6|7 operator rotates the dual bond between the heptagon and the nearby hexagon (the arrowed dual bonds in Figure 2b,c), moving the 5|7 pair along the fullerene network. At each step η, the 5|7 pair swaps its place with a 6|6 one (see also Figure 3b); after η iterations of the SW6|7 rearrangement, the C 240 η isomer is formed. Figure 4 shows the isomer with η = 2 . The described propagation process is applied to the isolated 5|7 dislocation monopole as well as to the 5|7 double pair arising from the SW6|6 rearrangement. It is worth mentioning that similar topological tools are used in other disciplines such as biology whereby wave-like diffusion processes and icosahedral patterns of viruses are normal mechanisms to model biochemical processes [39,40,41,42].
From the pure topological point of view, it is noteworthy that each fullerene network in the current study could be considered as the result of an instantaneous transformation generated by a single non-local generalized SW rotation. This kind of transformations, previously proposed in [25], constitutes a potentially infinite class of global rearrangements aiming to generate the entire isomeric space of a given Cn fullerene starting from a limited number (even only one) of inequivalent cages.
The final part of this section briefly introduces the topological invariants and how they measure the relative stability of fullerene isomers C 240 η modified by SW waves. Graph-theoretical results are presented and compared with energy values of the systems coming from quantum chemical simulations in the next sections. According to the topological modeling (TM) methods [43], chemical structures with n atoms and B bonds are schematically described as simple graphs G with n nodes and B edges. Geometrical, structural, and other physicochemical properties (maximal molecular symmetry being in the opinion of the authors the prominent, and somehow surprising, one) are embedded in the graph topology as the study of distance-based invariants evidenced some decades ago in pioneering studies [44,45]. In this context, distances dij are integer numbers computed by counting the number of edges connecting two atoms i and j with the shortest path in G. The distances are graph invariants (i.e., they do not depend on the labeling of the nodes) and generate a vast parade of distance-based graph descriptors. Graph diameter is defined as M = max{dij}.
TM techniques mainly make use of two invariants W (1) and ρ E (3) to measure graph compactness and roundness, respectively. Assuming dii = 0, the Wiener index W is derived as the semi-sum of all distances in G:
W G = 1 2 i , j = 1 n d i j = i = 1 n w i   ,
counting the contributions w i arising from each node in the graph:
w i = 1 2 j = 1 n d i j ,
Intuitively, Equation (2) admits the minimum ( w _ ) and maximum ( w ¯ ) values for central (peripheral) nodes in G. The quotient ρ E between the two w i extremals:
ρ E = w ¯ / w _ ,
called extreme topological roundness denotes the topological symmetry of the graph: the lower (3) is the more symmetric G is (with the general constraint ρ E 1 ). We assign the role of the topological potential of the system described by G to the Wiener index and extreme topological roundness. Indeed, both quantities obey the minimum principle. Therefore, the TM model assumes that similar carbon systems arrange their structures to minimize W and ρ E . Heuristically, this approach is confirmed by the fact that icosahedral fullerenes C60 and C240 minimize both indices in the respective isomeric sets. Recent works [46,47,48] on C66 and C84 molecules confirm that stable fullerenes combine high topological symmetry (low ρ E ) with high compactness (low W).
This fast and elegant computational approach is applicable to the evolution of the SWw-defected fullerene molecules C 240 η . The electronic configurations of the atoms constituting the C240 isomers with defective structures in Figure 3; Figure 4 keep the sp2-hybridization state. On the other hand, surface reconstructions induced by SWw cause complex modifications of “the electronic charge density and this varies the bond lengths within and around” [49], with effects also on the local curvature of the molecular cage.
We have performed the preliminary density functional theory (DFT) computations to monitor the topology effect on the geometrical parameters and molecular properties of the defected C240 cages. The DFT computations have been performed with standard procedures; the details are collected in Appendix A. We have computationally studied SWw on the C240 surface up to η = 6 (Figure 5). The SW6|6 operator implies replacing six original hexagons with the emergence of four “defected” polygons (two 5|7 pairs). From the point of structural chemistry, we approximate SWw as a movement of one “defected” pair from the other due to similar operations. Thus, each C 240 η with η = 0–3 contains 12 original pentagons, 2 heptagons and 2 pentagons, static and migrating, generated by the wave (the remaining faces being hexagons). In Figure 5, static and dynamic defects are colored differently for clarity. The C 240 η = 4 structure contains one migrating tetragon resulting from involving an original pentagon by SWw. In further structures, this tetragon converts to a new static pentagon. In other words, the wave shifts the position of the original pentagon in two operations. Note that the structures with η = 3 and η = 5 contain pentagon–pentagon fusions. These structural features are presented in Table 1.
Local curvature (k) is a key parameter we pay attention as it strongly correlates with the stability [1,2,50] and reactivity [51,52,53] of the fullerene species. The SWw transformations occur locally on the C240 surface. However, the calculated curvatures indicate the change in the shape throughout the whole surface (Equation (A2), Appendix A). To demonstrate it, we have chosen the curvature in the pentagon regions, both original and emerged upon SWw. As follows from Table 1, all pentagons have the same curvature in original C240 but SWw makes the curvatures more diversified. Accordingly, some of pentagons become more flattened as the others become more curved. Herewith, the structures with a tetragon ( C 240 η = 4 ) and fused pentagons ( C 240 η = 3 and C 240 η = 5 ) manifest the highest curvatures in the set.
Sphericity (Ψ) is another parameter indicating the changes in the shape (Equation (A1), Appendix A). The calculated sphericities decrease from the starting value 0.9829 as the C240 system loses its initial icosahedral symmetry. Note that original C240 (Ih) and C 240 η = 0 have very close values because the structure with η = 0 still has a low symmetry (CS). The further changes are more pronounced. Nevertheless, Ψ does not become lower than 0.97 though SWw perturbs the shape. For comparison, Ψ may be significantly reduced (Ψ < 0.9) in highly functionalized fullerene compounds (e.g., C60 halogenides [54]). Sphericity (with associated quantity of volume, V) does not change monotonically with SWw propagation, i.e., there are no (and should not be) correlations between Ψ (or V) and η. However, we demonstrate below the correlations between these geometric parameters with topological ones.
Pristine and almost spherical C240 (Ih) has n = 240 direct nodes and B = 360, three symmetry-distinct sets of atoms (orbits) having multiplicities 60, 60, and 120, and 240 atoms with the same w ¯ = w _ = 2312 . This last condition implies W = 277,400 and ρ E = 1 , indicating that the C240 (Ih) fullerene is a maximal roundness molecule even in presence of non-symmetry equivalent atoms (a special condition featured by tetrahedral C40 too [49]). We would like to remember here that the complete individuation of all families of non-transitive fullerenes with ρ E = 1 is still an open task.

3. Topological Simulations and Electronic Structure

3.1. Topological Modeling

The first step of the SW wave on the C240 (Ih) fullerene graph consists of a SW6|6 bond flip that produces the 5|7 double pair (Figure 3b). This surface transformation decreases the lattice compactness to W = 277,122 derivable from the direct computation of the graph chemical distances dij and Equation (1). In our approximated model, this –0.11% variation of the Wiener index represents the gain in topological compactness allowed by the creation of the SW defect in the molecule C 240 η = 0 .
The subsequent SW6|7 rotation (η = 1) splits the 5|7 pairs by inserting in between one pair 6|6 according to the topological mechanism described in (Figure 2a,b). This decreases by another –0.14% the value of the topological descriptor W = 276,730 for newly formed C 240 η = 1 . This “gaining” behavior holds for each propagation steps in the interval η = [0, 3] of SWw over the fullerene surface, increasing the topological compactness of the system.
The reduction of the topological index W follows a quasi-parabolic trend (Figure 6). The curve evidences that the creation of an extended topological defect up to η = 3 is favored; the topology of the fullerene network allows the topological diffusion of the 5|7 pair over a quite extended region. The SWw propagation is then impeded by a small +0.02% increment of the topological index W having a local peak for η = 4. The Wiener index curve in Figure 6 evidences, and this result is reported here for the first time for the fullerene molecules, that the 5|7 defects are able to migrate on the fullerene surface to form extended linear defects. This is a specific behavior that matches similar findings previously reported for graphene planar layers [26] and curved carbon nanotubes and nanotori [33]. The present TM simulations show that creation of SW waves is allowed also on quasi-spherical surfaces tiled by hexagons. Based on this outcome, we may therefore propose the following important conjecture: the formation and propagation of the SW waves are allowed on surfaces tiled with hexagons independently of the surface genus.
This effect has a pure topological root and strongly correlates with the connectivity properties of the pentagon–heptagon pairs embedded in the hexagonal mesh and the “edge” effect induced by the presence of pristine pentagons on the surface. We note in fact here that the small barrier effect encountered by the 5|7 pair at η = 4 (Figure 6) is an effect generated by SWw colliding with one of the original pentagons of the C240 (Ih) fullerene. A different situation is encountered in the variations of the topological roundness (Equation (3)) of the system upon SWw diffusion, see Figure 6.
The generation of the 5|7 double pairs reduces the topological symmetry and moves the ρ E curve upward after the first SW6|6 rotation, with a topological increase ρ E = 1.0241 for C 240 η = 0 compared with the unitary values of the original C240 (Ih). The successive SW6|7 flip starts the dipole propagation still increasing the topological invariant ρ E . The diffusion of the wave is opposed by the action of the invariant ρ E , which shows a barrier effect confirmed at the successive steps in Figure 5. At η = 2 and 3, the ρ E quantity show an almost flat curve in the same region where W shows a local minimum (Figure 6). This result is quite interesting since it confirms that system, when the dislocation dipole has size η = 2 and 3, stops the obstacle to the creation of defective molecules ( C 240 η = 2 represented in Figure 4).
Topological invariants are listed in Table 2 and represented in Figure 6. We have found the inverse correlations between the extremal roundness (Table 2), volumes and sphericities of the C240 cages (Figure 7a). The found correlations are the same as in the case of the C84 isomeric series (made up with the isomers obeying the isolated-pentagon rule) [55] and similar to the findings of comprehensive work [56] on all possible isomers of C60. The geometry and topology of the molecules obtain direct linkage in the relations of Ψ and ρ E . Indeed, Ψ → 1– for sphere-like species whereas ρ E → 1+ in the case of structures with “equilibrated connectivity”. These terms, in the case of the fullerenes, coincide with the notion that is manifested with the found correlations. As for the correlation between the volume and extremal roundness of the C240 cages, it follows from the interdependence of the Ψ and V values (Figure 7b).

3.2. DFT Simulations on Energetic and Molecular Parameters of the Defected C240 Cages

We have assessed the possibility of SWw transformations on the C240 cage from the thermodynamic point of view. For this purpose, we have operated with the DFT-computed total energies of the original and defected cages (see Appendix A for the details of the DFT study). As expected, the energies of all SWw-generated C240 cages are much higher as compared with the pristine one (Table 3). However, the energy effects of the stepwise transitions between C 240 η = 0 shows that some of the transitions may be favorable. Indeed, the transitions are endothermic up to C 240 η = 3 C 240 η = 4 and exothermic for the two next steps (Table 3). The exothermicity is explainable in terms of the structural chemistry of fullerenes. The structure C 240 η = 5 , a participant of the two exothermic steps, has two adjacent pentagons. This is the instability factor making the conversion C 240 η = 5 C 240 η = 6 favorable. As for η = 4, the precursor of C 240 η = 5 , it has the tetragon–heptagon junction, and exothermicity C 240 η = 4 C 240 η = 5 indicates that such an introduction of tetragon to the C240 cage makes the structure less favorable as compared with one containing the fused pentagons.
The HOMO–LUMO gap is another indicator of the fullerene stability [55] (HOMO and LUMO are the highest occupied and lowest unoccupied molecular orbitals, respectively). It is calculated as the difference in the HOMO and LUMO energies. In our isomeric C240 series, only the original structure manifests the largest gap value that indicates its stability. Other (defected) cages, as follows from the orbital energies, should be kinetically unstable. We have tried to collate the obtained HOMO–LUMO gaps with the results of the TM modeling and found a correlation between the gap and extreme topological roundness values (Figure 8). Indeed, both values, being different in their origin, should reflect the stability of the C240 cages. However, the correlation is weak (R2 = 0.861) and requires further investigation, which includes a larger number of the defected fullerene structures.

4. Discussion

The results of the TM simulations are summarized in Table 2 and Figure 6. The analysis of the evolution of the topological compactness W of the isomers C 240 η (Figure 6) allows observing how the generation η = 1 of the SW dipole causes an immediate decrease in the network invariant W = 277,122 (Equation (1)). In our topological model, this negative variation in the Wiener index represents the topological gain in the graph compactness in respect to the pristine C240 (Ih) fullerene. The creation of the SW defect is therefore able to make the fullerene cage more compact. The successive SW6|7 rotations further augment the topological compactness of the structure, the reduction in W showing an almost parabolic trend for the first propagation steps η = 1, 2, 3. This characteristic behavior of the W index, which favors the expansion of the SW wave, has a pure topological root and it relates to the connectivity properties of the hexagonal regions partially covering the fullerene surface. This result, in fact, matches with previously reported behaviors of SW waves when they propagate in pure honeycomb lattices [26]. The topological diffusion of the 5|7 pair with creating extended dislocation dipoles is stopped at η = 4 and W reaching a local maximum (W = 276,607). Here, the SW wave “collapses” with one of the pentagons of the pristine icosahedral fullerene with the creation of a 4-ring of carbon atoms.
Figure 6 shows instead that the topological order gets diminished by the creation and the propagation of the SW wave. The increase in the extreme topological roundness ρ E (Equation (3)) represents, in fact, an evident “topological response” of the cage to the diffusion of the pentagon-heptagon pairs over the fullerene surface. The evolution of ρ E demonstrates a characteristic plateau (η = 3, 4 with ρ E ≅ 1.05 and η = 4, 5, 6 with ρ E ≅ 1.06) where the topological symmetry of the C 240 η cages remains pretty uncanged, opening the door for the presence of extended dislocation dipoles embedded in the fullerene mesh.
In summary, the main outcomes of the TM simulations state that isomers C 240 η have greater compactness (lower W) than the parent C240 (Ih), favoring the creation and the propagation of the SWw defects over the fullerene surface. In these molecules, however, at each step of the SWw propagation, the ρ E index increases (lowering their extreme topological roundness) and obstructs the diffusion of the SWw itself; this barrier effect is reduced in the case of the large η values.
The DFT-estimated energy parameters, in general, indicate endothermicity of the transformations towards the defected fullerene isomers. Thus, such SW processes may take place under high-energy (nonequilibrium) conditions (plasma, interstellar environments etc. [20,21]). The computed HOMO–LUMO gaps also reveal the kinetic instability of the defected cages (in contrast to the parent one having the largest gap). We do not deeply discuss the HOMO–LUMO vs. ρ E plot (Figure 8). There values seem correlated, but the justification of the trend requires extending the correlation field.
The title defect introduced to the honeycomb carbon-containing system is called a dislocation dipole. In the case of the fullerenes, we associate it with nonzero dipole moments (Table 3) of the defected fullerene cages (indeed, it is impossible to make such attribution in the case of introducing a dislocation dipole to the infinite systems, like graphene [57], as their dipole moment is not defined). The influence of the defects on the polarizability of fullerenes has been studied in a few works concluding the increase with the topological defects (see in review [58]). In further studies, we will investigate this important property in the context of SW waves.
As usually considered, the title defect also generates negative-curvature regions on the fullerene surface. In our estimates deduced from the pyramidality angles (Equation (A2)), we have found the atoms with decreased curvatures of the surroundings but which are still positive. Thus, the presence of the pair 5|7 is not the only condition of the negatively curved surface. Additionally, upon SWw, more curved regions also emerge, especially with tetragon inclusion or pentagon–pentagon fusions. Curvature is an important parameter for assessing the reactivity of fullerenes and their derivatives [1,50,51,52,53]. We think that our estimates may be useful for assessing chemical properties of the giant fullerenes, hollow [4,5,6,50] or filled (as a part of carbon nano-onions [59,60]).
The calculated curvature values correspond to the local changes in the C240 structures upon SWw. Sphericity is another parameter reflecting the general change in the shape of the molecule. As expected, it goes down with introducing the defects. More important here, sphericity, a geometric structural parameter, is correlated with extreme roundness, a topological structural parameter. This indicated the deducibility of the structural and energetic parameters from the topology of the molecules.

5. Conclusions and Prospective

In the present work, topological, energetic and structural modifications induced on the C240 fullerene by the presence of a propagating expanding SWw have been investigated for the first time. Although cage compactness is enhanced for the dislocation-dipole isomers, their energies increase with the size η of the SW dislocation. This means a relatively low probability of the formation of these defective fullerenes. However, the existence of these defective isomers, especially those molecules with large η, which combine low W with non-increasing ρ E , may relate to special thermodynamic conditions like during fullerene growth or in interstellar media.
As found, the extreme topological roundness correlate with volumes and sphericities of the C240 cages. Such correlations are general for fullerene structures (as they have also been found for other isomeric series). Considering these regularities, we consider the efficiency of the presented topological approach for structural studies on fullerenes and deducing their molecular properties.
These results are perfectly in line with the topological determination of the relative stability of smaller fullerenes. The ability of W in selecting the most compact cage as the physically stable one is documented by the non-trivial case of the C60 fullerene [44] that, among its 1812 non-isomorphic isomers, minimizes the Wiener index in correspondence to the C60 (Ih) with W = 8340 and ρ E = 1 . In more complex C66 case [46], the stability of the whole set of 4478 isomers has been ranked with maximizing two graph properties, i.e., topological compactness (low W) and topological roundness (low ρ E ). As concluded, the most stable C66 isomers minimize both graph invariants and, hence, may be found in the local minima of the (W,   ρ E ) plane. We recently found that this purely topological approach works also for the C84 fullerene. Its most stable forms have minimal graph invariants W and   ρ E [47,48]. The latter result also evidences the computational convenience of the TM simulations that rank just in a matter of minutes the topological stability of 51,592 distinct C84 isomers. Thus, we highlight that the TM approximated method is generally applicable to the formation and stabilization mechanisms of complex sp2 carbon structures, like large fullerenes. As an additional advance, topological studies allow quick sieving the possible structures before detailed but time-consuming ab initio computations.

Author Contributions

Conceptualization, D.S.S. and O.O.; Investigation, D.S.S. and O.O.; Methodology, D.S.S. and O.O.; Visualization, D.S.S. and O.O.; Writing—original draft, D.S.S. and O.O.; Writing—review & editing, D.S.S. and O.O. The authors have contributed equally. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partly supported by Russian Foundation for Basic Research (project number 19-03-00716).

Acknowledgments

The friendly support and suggestions of Franco Cataldo (Actinium Chemical Research Institute, Italy) and Mihai V. Putz (West University of Timisoara, Romania) are warmly recognized. The work was performed under the theme “Novel theoretical approaches and software for modeling complex chemical processes and compounds with tunable physicochemical properties” (registration number AAAA-A19-119022290011-6, Institute of Petrochemistry and Catalysis of RAS).

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Details of Quantum Chemical and Auxiliary Calculations

All optimizations were performed by the density functional theory method PBE/3ζ implemented in the Priroda program [61]. The 3ζ basis set describes the electronic configuration of carbon atoms as contracted Gaussian-type functions (11s,6p,2d)/[6s,3p,2d] combined with uncontracted ones (10s,3p,3d,1f). Previously, this method was used in diverse computational studies on fullerenes, including structure simulations [62,63,64], DFT-assisted spectral characterization of fullerene compounds [62,65,66,67], assessing their molecular and energetic parameters [68,69,70,71], kinetic and thermodynamic studies of fullerene reactions [62,72,73,74,75,76,77,78]. The method is also applied to related compounds such as polycyclic aromatic hydrocarbons [68,79,80] and endofullerenes [69,70,71,81]. The accuracy and reliability of PBE/3ζ were discussed in the mentioned papers. Based on the previous advances, we consider the suitability of the method to the purpose of this work (e.g., the method works well in the case of giant fullerenes [16], perfectly reproduces experimental IR and NMR spectra of fullerene compounds [62,65,66,67] and volumes of fullerene cages [64]).
We used the Cartesian coordinates of the atoms for calculating the volumes, sphericities, and curvatures of the fullerene cages. The term volume of the fullerene cage means the volume of the polyhedron made up with nuclei of the atoms of the cage (i.e., nuclear volume). The algorithm for the volume calculations has been previously presented (in general, it implies the triangulation of the fullerene cage, i.e., its partition over the disjoint pyramid primitives having one common vertex in the center of mass of the cage) [64]. The sphericities (Ψ) of the C240 cages are calculated as previously described [47,54]:
Ψ = π 1 / 3 6 V 2 / 3 A
where V and A are the volume and the surface area of the fullerene cage. Sphericity is the ratio of the surface area of a sphere, having the same volume as the given particle, to the surface area of the particle.
The local curvature of each site (atom) is calculated using the optimized geometry of a fullerene as [63]
k = 2sin θP/a,
where a is the average distance between the site and its three neighbors, and θP is the pyramidality angle of the site. The local curvature of polygons are the average values of the constituting atoms.

References

  1. Lu, X.; Chen, Z. Curved pi-conjugation, aromaticity, and the related chemistry of small fullerenes (<C60) and single-walled carbon nanotubes. Chem. Rev. 2005, 105, 3643–3696. [Google Scholar] [CrossRef]
  2. Kovalenko, V.I.; Khamatgalimov, A.R. Regularities in the molecular structures of stable fullerenes. Russ. Chem. Rev. 2006, 75, 981–988. [Google Scholar] [CrossRef]
  3. Goldberg, M. A class of multi-symmetric polyhedra. Tohoku Math. J. First Ser. 1937, 43, 104–108. [Google Scholar]
  4. Schewerdtfeger, P.; Wirz, L.N.; Avery, J.E. The topology of fullerenes. WIREs Comput. Mol. Sci. 2015, 5, 96–145. [Google Scholar] [CrossRef]
  5. Yoshida, M.; Ōsawa, E. Molecular mechanics calculations of giant- and hyperfullerenes with eicosahedral symmetry. Fullerene Sci. Tech. 1992, 1, 55–74. [Google Scholar] [CrossRef]
  6. Sabirov, D.S.; Ōsawa, E. Information entropy of fullerenes. J. Chem. Inf. Model. 2015, 55, 1576–1584. [Google Scholar] [CrossRef]
  7. Suyetin, M.V.; Vakhrushev, A.V. Guided carbon nanocapsules for hydrogen storage. J. Phys. Chem. C 2011, 115, 5485–5491. [Google Scholar] [CrossRef]
  8. Zope, R.R. Electronic structure and static dipole polarizability of C60@C240. J. Phys. B: At. Mol. Opt. Phys. 2008, 41, 085101. [Google Scholar] [CrossRef]
  9. Hirsch, A.; Brettreich, M. Fullerenes: Chemistry and Reactions; John Wiley & Sons: New York, NY, USA, 2006. [Google Scholar]
  10. Taylor, R. Lecture Notes on Fullerene Chemistry: A Handbook for Chemists; Imperial College Press: London, UK, 1999. [Google Scholar]
  11. Mordkovich, V.Z.; Umnov, A.G.; Inoshita, T.; Endo, M. The observation of multiwall fullerenes in thermally treated laser pyrolysis carbon blacks. Carbon 1999, 37, 1855–1858. [Google Scholar] [CrossRef]
  12. Mordkovich, V.Z.; Shiratori, Y.; Hiraoka, H.; Umnov, A.G.; Takeuchi, Y. A path to larger yields of multishell fullerenes. Carbon 2001, 39, 1929–1941. [Google Scholar] [CrossRef]
  13. Mordkovich, V.Z.; Takeuchi, Y. Multishell fullerenes by laser vaporization of composite carbon–metal targets. Chem. Phys. Lett. 2002, 355, 133–138. [Google Scholar] [CrossRef]
  14. Mordkovich, V.Z.; Shiratori, Y.; Hiraoka, H.; Takeuchi, Y. Synthesis of multishell fullerenes by laser vaporization of composite carbon targets. Phys. Solid State 2002, 44, 603–606. [Google Scholar] [CrossRef]
  15. Noël, Y.; De La Pierre, M.; Zicovich-Wilson, C.M.; Orlando, R.; Dovesi, R. Structural, electronic and energetic properties of giant icosahedral fullerenes up to C6000: Insights from an ab initio hybrid DFT study. Phys. Chem. Chem. Phys. 2014, 16, 13390–13404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Pankratyev, E.Y.; Khatymov, R.V.; Sabirov, D.S.; Yuldashev, A.V. On the upper bound of the thermodynamic stability of fullerenes from small to giant. Physica E 2018, 101, 265–272. [Google Scholar] [CrossRef]
  17. Zope, R.R.; Baruah, T.; Pederson, M.R.; Dunlap, B.I. Static dielectric response of icosahedral fullerenes from C60 to C2160 characterized by an all-electron density functional theory. Phys. Rev. B 2008, 77, 115452. [Google Scholar] [CrossRef] [Green Version]
  18. Langlet, R.; Mayer, A.; Geuquet, N.; Amara, H.; Vandescuren, M.; Henrard, L.; Maksimenko, S.; Lambin, P. Study of the polarizability of fullerenes with a monopole–dipole interaction model. Diamond Relat. Mater. 2007, 16, 2145–2149. [Google Scholar] [CrossRef]
  19. Calaminici, P.; Carmona-Espindola, J.; Geudtner, G.; Köster, A.M. Static and dynamic polarizability of C540 fullerene. Int. J. Quantum Chem. 2012, 112, 3252–3255. [Google Scholar] [CrossRef]
  20. Irle, S.; Zheng, G.; Wang, Z.; Morokuma, K. The C60 formation puzzle “solved”: QM/MD simulations reveal the shrinking hot giant road of the dynamic fullerene self-assembly mechanism. J. Phys. Chem. B 2006, 110, 14531–14545. [Google Scholar] [CrossRef]
  21. Ōsawa, E. Formation mechanism of C60 under nonequilibrium and irreversible conditions—An annotation. Fuller. Nanotub. Carbon Nanostruct. 2012, 20, 299–309. [Google Scholar] [CrossRef]
  22. Stone, A.J.; Wales, D.J. Theoretical studies of icosahedral C60 and some related species. Chem. Phys. Lett. 1986, 128, 501–503. [Google Scholar] [CrossRef]
  23. Monthioux, M.; Charlier, J.C. Giving credit where credit is due: The Stone–(Thrower)–Wales designation revisited. Carbon 2014, 75, 1–4. [Google Scholar] [CrossRef]
  24. Kumeda, Y.; Wales, D.J. Ab initio study of rearrangements between C60 fullerenes. Chem. Phys. Lett. 2003, 374, 125–131. [Google Scholar] [CrossRef]
  25. Babić, D.; Bassoli, S.; Casartelli, M.; Cataldo, F.; Graovac, A.; Ori, O.; York, B. Generalized Stone–Wales transformations. Mol. Simul. 1995, 14, 395–401. [Google Scholar] [CrossRef]
  26. Ori, O.; Cataldo, F.; Putz, M.V. Topological anisotropy of Stone–Wales waves in graphenic fragments. Int. J. Mol. Sci. 2011, 12, 7934–7949. [Google Scholar] [CrossRef] [PubMed]
  27. Ori, O.; Putz, M.V.; Gutman, I.; Schwerdtfeger, P. Generalized Stone–Wales transformations for fullerene graphs derived from Berge’s switching theorem. In Ante Graovac—Life and Works; Gutman, I., Pokrić, B., Vukičević, D., Eds.; University of Kragujevac: Kragujevac, Serbiba, 2014; pp. 259–272. [Google Scholar]
  28. Liu, Y.; Yakobson, B.I. Cones, pringles, and grain boundary landscapes in graphene topology. Nano Lett. 2010, 10, 2178–2183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Carpio, A.; Bonilla, L.L.; de Juan, F.; Vozmediano, M.A.H. Dislocations in graphene. New J. Phys. 2008, 10, 053021. [Google Scholar] [CrossRef]
  30. Zhoua, L.G.; Shib, S.Q. Formation energy of Stone–Wales defects in carbon nanotubes. Appl. Phys. Lett. 2003, 83, 1222–1224. [Google Scholar] [CrossRef] [Green Version]
  31. Collins, P.G. Defects and disorder in carbon nanotubes. In Oxford Handbook of Nanoscience and Technology: Frontiers and Advances; Narlikar, A.V., Fu, Y.Y., Eds.; Oxford University Press: Oxford, UK, 2011; Volume 2. [Google Scholar] [CrossRef] [Green Version]
  32. Ewels, C.P.; Heggie, M.I.; Briddon, P.R. Adatoms and nanoengineering of carbon. Chem. Phys. Lett. 2002, 351, 178–182. [Google Scholar] [CrossRef] [Green Version]
  33. Samsonidze, G.G.; Samsonidze, G.G.; Yakobson, B.I. Energetics of Stone–Wales defects in deformations of monoatomic hexagonal layers. Comput. Mater. Sci. 2002, 23, 62–72. [Google Scholar] [CrossRef]
  34. Nordlund, K.; Keinonen, J.; Mattila, T. Formation of ion irradiation induced small-scale defects on graphite surfaces. Phys. Rev. Lett. 1996, 77, 699–702. [Google Scholar] [CrossRef] [Green Version]
  35. Krasheninnikov, A.V.; Nordlund, K.; Sirviö, M.; Salonen, E.; Keinonen, J. Formation of ion-irradiation-induced atomic-scale defects on walls of carbon nanotubes. Phys. Rev. B 2001, 63, 245405. [Google Scholar] [CrossRef] [Green Version]
  36. Hashimoto, A.; Suenaga, K.; Gloter, A.; Urita, K.; Iijima, S. Direct evidence for atomic defects in graphene layers. Nature 2004, 430, 870–873. [Google Scholar] [CrossRef]
  37. Kotakoski, J.; Krasheninnikov, A.V.; Kaiser, U.; Meyer, J.C. From point defects in graphene to two-dimensional amorphous carbon. Phys. Rev. Lett. 2011, 106, 105505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. RobertLovesPi: Polyhedra, tessellations, and more. Available online: https://robertlovespi.net/c240-fullerene-1/ (accessed on 27 April 2020).
  39. Pyshnov, M.B. Topological solution for cell proliferation in intestinal crypt. I. Elastic growth without cell loss. J. Theor. Biol. 1980, 87, 189–200. [Google Scholar] [CrossRef]
  40. Branden, C.I.; Tooze, J. Introduction to Protein Structure, 2nd ed.; Garland Science: New York, NY, USA, 1999; pp. 1–302. [Google Scholar]
  41. Twarock, R. Mathematical virology: A novel approach to the structure and assembly of viruses. Philos. Trans. R. Soc. A 2006, 364, 335–3373. [Google Scholar] [CrossRef]
  42. Dechant, P.-P.; Wardman, J.; Keef, T.; Twarock, R. Viruses and fullerenes – symmetry as a common thread? Acta Crystallogr. A 2014, 70, 162–167. [Google Scholar] [CrossRef] [Green Version]
  43. Iranmanesh, A.; Ashrafi, A.R.; Graovac, A.; Cataldo, F.; Ori, O. Wiener index role in topological modeling of hexagonal systems-from fullerenes to graphene. In Distance in Molecular Graphs – Applications; Gutman, I., Furtula, B., Eds.; University of Kragujevac: Kragujevac, Serbia, 2012; pp. 135–155. [Google Scholar]
  44. Ori, O.; D’Mello, M. A topological study of the structure of the C76 fullerene. Chem. Phys. Lett. 1992, 197, 49–54. [Google Scholar] [CrossRef]
  45. Babić, D.; Klein, D.J.; Sah, C.H. Symmetry of fullerenes. Chem. Phys. Lett. 1993, 211, 235–241. [Google Scholar] [CrossRef]
  46. Vukicevic, D.; Cataldo, F.; Ori, O.; Graovac, A. Topological efficiency of C66 fullerene. Chem. Phys. Lett. 2011, 501, 442–445. [Google Scholar] [CrossRef]
  47. Sabirov, D.S.; Ori, O.; László, I. Isomers of the C84 fullerene: A theoretical consideration within energetic, structural, and topological approaches. Fuller. Nanotub. Carbon Nanostruct. 2018, 26, 100–110. [Google Scholar] [CrossRef]
  48. Dobrynin, A.A.; Ori, O.; Putz, M.V.; Vesnin, A.Y. Generalized topological efficiency – case study with C84 fullerene. Fuller. Nanotub. Carbon Nanostruct. 2020, 28, 545–550. [Google Scholar] [CrossRef]
  49. Chen, Q.; Robertson, A.W.; He, K.; Gong, C.; Yoon, E.; Lee, G.D.; Warner, J.H. Atomic level distributed strain within graphene divacancies from bond rotations. ACS Nano 2015, 9, 8599–8608. [Google Scholar] [CrossRef] [PubMed]
  50. Haddon, R.C. Chemistry of the fullerenes: The manifestation of strain in a class of continuous aromatic molecules. Science 1993, 261, 1545–1550. [Google Scholar] [CrossRef] [PubMed]
  51. Astakhova, T.Y.; Vinogradov, G.A.; Gurin, O.D.; Menon, M. Effect of local strain on the reactivity of carbon nanotubes. Russ. Chem. Bull. 2002, 51, 764–769. [Google Scholar] [CrossRef]
  52. Sabirov, D.S.; Bulgakov, R.G.; Khursan, S.L. Indices of the fullerene reactivity. ARKIVOC 2011, 2011, 200–224. [Google Scholar] [CrossRef] [Green Version]
  53. Sabirov, D.S.; Bulgakov, R.G. Reactivity of fullerene derivatives C60O and C60F18 (C3v) in terms of local curvature and polarizability. Fuller. Nanotub. Carbon Nanostruct. 2010, 18, 455–457. [Google Scholar] [CrossRef]
  54. Sabirov, D.S.; Garipova, R.R. The increase in the fullerene cage volume upon its chemical functionalization. Fuller. Nanotub. Carbon Nanostruct. 2019, 27, 702–709. [Google Scholar] [CrossRef]
  55. Fowler, P.W.; Manolopoulos, D.E. An Atlas of Fullerenes; Dover Publications Inc.: Mineola, NY, USA, 1995; pp. 1–392. [Google Scholar]
  56. Sure, R.; Hansen, A.; Schwerdtfeger, P.; Grimme, S. Comprehensive theoretical study of all 1812 C60 isomers. Phys. Chem. Chem. Phys. 2017, 19, 14296–14305. [Google Scholar] [CrossRef]
  57. Klyavlina, A.I.; Rysaeva, L.K.; Murzaev, R.T. Dislocation dipole in graphene at finite temperatures. J. Phys. Conf. Ser. 2020, 1435, 012063. [Google Scholar] [CrossRef]
  58. Sabirov, D.S. Rules of fullerene polarizability. Fuller. Nanotub. Carbon Nanostruct. 2020, 28, 71–77. [Google Scholar] [CrossRef]
  59. Plonska-Brzezinska, M.E. Carbon nano-onions: A review of recent progress in synthesis and applications. ChemNanoMat. 2019, 5, 568–580. [Google Scholar] [CrossRef]
  60. Bartkowski, M.; Giordani, S. Supramolecular chemistry of carbon nano-onions. Nanoscale 2020, 12, 9352–9358. [Google Scholar] [CrossRef] [PubMed]
  61. Laikov, D.N.; Ustynyuk, Y.A. PRIRODA-04: A quantum-chemical program suite. New possibilities in the study of molecular systems with the application of parallel computing. Russ. Chem. Bull. 2005, 54, 820–826. [Google Scholar] [CrossRef]
  62. Shestakov, A.F. Reactivity of fullerene C60. Russ. J. Gen. Chem. 2008, 78, 811–821. [Google Scholar] [CrossRef]
  63. Sabirov, D.S.; Khursan, S.L.; Bulgakov, R.G. 1,3-Dipolar addition reactions to fullerenes: The role of the local curvature of carbon surface. Russ. Chem. Bull. 2008, 57, 2520–2525. [Google Scholar] [CrossRef]
  64. Sabirov, D.S.; Zakirova, A.D.; Tukhbatullina, A.A.; Gubaydullin, I.M.; Bulgakov, R.G. Influence of the charge on the volumes of nanoscale cages (carbon and boron-nitride fullerenes, Ge9z Zintl ions, and cubic Fe4S4 clusters). RSC Adv. 2013, 3, 1818–1824. [Google Scholar] [CrossRef]
  65. Tulyabaev, A.R.; Kiryanov, I.I.; Samigullin, I.S.; Khalilov, L.M. Are there reliable DFT approaches for 13C NMR chemical shift predictions of fullerene C60 derivatives? Int. J. Quant. Chem. 2017, 117, 7–14. [Google Scholar] [CrossRef] [Green Version]
  66. Pankratyev, E.Y.; Tulyabaev, A.R.; Khalilov, L.M. How reliable are GIAO calculations of 1H and 13C NMR chemical shifts? A statistical analysis and empirical corrections at DFT (PBE/3z) level? J. Comput. Chem. 2011, 32, 1993–1997. [Google Scholar] [CrossRef] [PubMed]
  67. Sabirov, D.S.; Kinzyabaeva, Z.S. Sonochemical synthesis of novel C60 fullerene 1,4-oxathiane derivative through the intermediate fullerene radical anion. Ultrason. Sonochem. 2020, 67, 105169. [Google Scholar] [CrossRef]
  68. Sabirov, D.S.; Garipova, R.R.; Cataldo, F. Polarizability of isomeric and related interstellar compounds in the aspect of their abundance. Mol. Astrophys. 2018, 12, 10–12. [Google Scholar] [CrossRef]
  69. Sabirov, D.S.; Terentyev, A.O.; Shepelevich, I.S.; Bulgakov, R.G. Inverted thermochemistry of “norbornadiene–quadricyclane” molecular system inside fullerene nanocages. Comput. Theor. Chem. 2014, 1045, 86–92. [Google Scholar] [CrossRef]
  70. Sabirov, D.S.; Tukhbatullina, A.A.; Bulgakov, R.G. Compression of methane endofullerene CH4@C60 as a potential route to endohedral covalent fullerene derivatives: A DFT study. Fuller. Nanotub. Carbon Nanostruct. 2015, 23, 835–842. [Google Scholar] [CrossRef]
  71. Zakirova, A.D.; Sabirov, D.S. Volume of the fullerene cages of endofullerenes and hydrogenated endofullerenes with encapsulated atoms of noble gases and nonadditivity of their polarizability. Russ. J. Phys. Chem. A 2020, 94, 963–971. [Google Scholar] [CrossRef]
  72. Shestakov, A.F. Role of fullerene–nitrogen complexes of alkali metals in C60-catalyzed nitrogen fixation. Russ. J. Phys. Chem. A 2020, 94, 919–924. [Google Scholar] [CrossRef]
  73. Pimenova, A.S.; Kozlov, A.A.; Goryunkov, A.A.; Markov, V.Y.; Khaverl, P.A.; Avdoshenko, S.M.; Vorobiev, V.A.; Ioffe, I.N.; Sakharov, S.G.; Troyanov, S.I.; et al. Preparation and structures of [6,6]-open difluoromethylene [60]fullerenes: C60(CF2) and C60(CF2)2. Dalton Trans. 2007, 5322–5328. [Google Scholar] [CrossRef]
  74. Ignat’eva, D.V.; Goryunkov, A.A.; Tamm, N.B.; Ioffe, I.N.; Sidorov, L.N.; Troyanov, S.I. Isolation and structural characterization of the most highly trifluoromethylated C70 fullerenes: C70(CF3)18 and C70(CF3)20. New J. Chem. 2013, 37, 299–302. [Google Scholar] [CrossRef]
  75. Lukonina, N.S.; Semivrazhskaya, O.O.; Apenova, M.G.; Belov, N.M.; Troyanov, S.I.; Goryunkov, A.A. CF2-functionalized trifluoromethylated fullerene C70(CF3)8(CF2): Structure, electronic properties, and spontaneous oxidation at the bridgehead carbon atoms. Asian J. Org. Chem. 2019, 8, 1924–1932. [Google Scholar] [CrossRef]
  76. Diniakhmetova, D.R.; Friesen, A.K.; Kolesov, S.V. Quantum chemical analysis of the mechanism of the participation of C60 fullerene in the radical polymerization of styrene and mma initiated by benzoyl peroxide or azobisisobutyronitrile. Russ. J. Phys. Chem. B 2017, 11, 492–498. [Google Scholar] [CrossRef]
  77. Diniakhmetova, D.R.; Friesen, A.K.; Kolesov, S.V. Quantum chemical modeling of the addition reactions of 1-n-phenylpropyl radicals to C60 fullerene. Int. J. Quant. Chem. 2016, 116, 489–496. [Google Scholar] [CrossRef]
  78. Diniakhmetova, D.R.; Friesen, A.K.; Yumagulova, R.K.; Kolesov, S.V. Simulation of potentially possible reactions at the initial stages of free-radical polymerization of styrene and methyl methacrylate in the presence of fullerene C60. Polymer Sci. B 2018, 60, 414–420. [Google Scholar] [CrossRef]
  79. Khatymov, R.V.; Muftakhov, M.V.; Tuktarov, R.F.; Raitman, O.A.; Shokurov, A.V.; Pankratyev, E.Y. Fragmentation and slow autoneutralization of isolated negative molecular ions of phthalocyanine and tetraphenylporphyrin. J. Chem. Phys. 2019, 150, 134301. [Google Scholar] [CrossRef] [PubMed]
  80. Khatymov, R.V.; Shchukin, P.V.; Muftakhov, M.V.; Yakushchenko, I.K.; Yarmolenko, O.V.; Pankratyev, E.Y. A unified statistical RRKM approach to the fragmentation and autoneutralization of metastable molecular negative ions of hexaazatrinaphthylenes. Phys. Chem. Chem. Phys. 2020, 22, 3073–3088. [Google Scholar] [CrossRef] [PubMed]
  81. Kuznetsov, V.V. Stereochemistry of simple molecules inside nanotubes and fullerenes: Unusual behavior of usual systems. Molecules 2020, 25, 2437. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Bottom Stone–Wales reversible transformation SW5|6 on the fullerene direct graph G; SW5|6 rotates the central bond between green-orange pair. Red arrows indicate the bonds involved in the topological rearrangement. Top, SW5|6 acting in the dual space G ¯ . (b) The generalized Stone–Wales reversible rotation gSW of order η = 3 in the dual space. In both cases, blue (yellow) elements represent hexagonal (pentagonal) rings; white circles may be hexagonal or pentagonal faces indifferently and they are preserved before/after the gSW rotation.
Figure 1. (a) Bottom Stone–Wales reversible transformation SW5|6 on the fullerene direct graph G; SW5|6 rotates the central bond between green-orange pair. Red arrows indicate the bonds involved in the topological rearrangement. Top, SW5|6 acting in the dual space G ¯ . (b) The generalized Stone–Wales reversible rotation gSW of order η = 3 in the dual space. In both cases, blue (yellow) elements represent hexagonal (pentagonal) rings; white circles may be hexagonal or pentagonal faces indifferently and they are preserved before/after the gSW rotation.
Mathematics 08 00968 g001
Figure 2. (a) SW6|6 rotation of the arrowed bond creates in the pristine graphene dual lattice (white circles represent dual hexagons) the 5|7 double pair on the left. (b) A view of the 5|7 double pair; SW6|7 applied to the arrowed bond moves one 5|7 pair one step down in the hexagonal mesh (c). After the second iterated SW6|7 rotation SWw of order η = 2 is formed (d). Green (yellow) elements represent heptagonal (pentagonal) rings. The 5|7 double pair in (b) represents SWw of order η = 0. All Stone–Wales rearrangements are reversible.
Figure 2. (a) SW6|6 rotation of the arrowed bond creates in the pristine graphene dual lattice (white circles represent dual hexagons) the 5|7 double pair on the left. (b) A view of the 5|7 double pair; SW6|7 applied to the arrowed bond moves one 5|7 pair one step down in the hexagonal mesh (c). After the second iterated SW6|7 rotation SWw of order η = 2 is formed (d). Green (yellow) elements represent heptagonal (pentagonal) rings. The 5|7 double pair in (b) represents SWw of order η = 0. All Stone–Wales rearrangements are reversible.
Mathematics 08 00968 g002
Figure 3. (a) Pristine C240 (Ih) fullerene has three symmetry-independent quartets of hexagons A, B, and C delimited by yellow diamonds whose corners correspond to the centers of the hexagons. (b) The C 240 η = 0 structure generated by applying the SW6|6 rotation to quartet A in (a). The double pair 5|7 and SWw with length η = 0 are shown. Yellow (green) polygons represent pentagonal (heptagonal) rings. The pair 6|6 is ready for the SW6|7 rotation and swap place with the nearby pair 5|7.
Figure 3. (a) Pristine C240 (Ih) fullerene has three symmetry-independent quartets of hexagons A, B, and C delimited by yellow diamonds whose corners correspond to the centers of the hexagons. (b) The C 240 η = 0 structure generated by applying the SW6|6 rotation to quartet A in (a). The double pair 5|7 and SWw with length η = 0 are shown. Yellow (green) polygons represent pentagonal (heptagonal) rings. The pair 6|6 is ready for the SW6|7 rotation and swap place with the nearby pair 5|7.
Mathematics 08 00968 g003
Figure 4. The molecule of the C 240 η = 2 fullerene. It exhibits the extended dislocation dipole generated by applying two SW6|7 rotation to the configuration shown in Figure 3b. The topological defect consists of two pairs of 5|7 and two pairs of 6|6 in between (the faces are numbered according to the number of carbon atoms in the ring). The molecule is ready for another SW6|7 rearrangement involving the lower pair 5|7 and two neighboring hexagons (marked with yellow diamond) to produce the C 240 η = 3 fullerene (not represented).
Figure 4. The molecule of the C 240 η = 2 fullerene. It exhibits the extended dislocation dipole generated by applying two SW6|7 rotation to the configuration shown in Figure 3b. The topological defect consists of two pairs of 5|7 and two pairs of 6|6 in between (the faces are numbered according to the number of carbon atoms in the ring). The molecule is ready for another SW6|7 rearrangement involving the lower pair 5|7 and two neighboring hexagons (marked with yellow diamond) to produce the C 240 η = 3 fullerene (not represented).
Mathematics 08 00968 g004
Figure 5. The DFT-optimized C 240 η fullerene structures. Blue and red colorings correspond to the static and migrating (Stone–Wales wave) defects, respectively (except for the case of η = 0 whereby the defect polygons are fused).
Figure 5. The DFT-optimized C 240 η fullerene structures. Blue and red colorings correspond to the static and migrating (Stone–Wales wave) defects, respectively (except for the case of η = 0 whereby the defect polygons are fused).
Mathematics 08 00968 g005
Figure 6. Changes in the Wiener index (W) and extreme topological roundness ( ρ E ) upon the dislocation dipole propagation over the C240 surface, from the original Ih fullerene to η = 6.
Figure 6. Changes in the Wiener index (W) and extreme topological roundness ( ρ E ) upon the dislocation dipole propagation over the C240 surface, from the original Ih fullerene to η = 6.
Mathematics 08 00968 g006
Figure 7. (a) Correlations between the extremal roundness, volume and sphericity of the C240 cages. (b) Linear correlation between the volume and sphericity of the C240.
Figure 7. (a) Correlations between the extremal roundness, volume and sphericity of the C240 cages. (b) Linear correlation between the volume and sphericity of the C240.
Mathematics 08 00968 g007
Figure 8. Correlation between the HOMO–LUMO gap and extreme topological roundness of the C240 cages.
Figure 8. Correlation between the HOMO–LUMO gap and extreme topological roundness of the C240 cages.
Mathematics 08 00968 g008
Table 1. Structural parameters of original and defected C240 fullerenes (DFT computations).
Table 1. Structural parameters of original and defected C240 fullerenes (DFT computations).
Fullerene CageNumber of Polygons 1Volume (Å 3)Sphericity, ΨCurvatures of Original Pentagons (Å–1)Curvatures of Defected Pentagons (Å–1)
StaticMigrating
C240 (Ih)12 110 0 01433.730.98290.2365nonenone
C 240 η = 0 14 106 2 01436.100.98280.1850–0.24750.16410.1257
C 240 η = 1 14 106 2 01430.110.97970.1763–0.25860.21120.1537
C 240 η = 2 14 106 2 01425.730.97740.1763–0.25020.21710.1915
C 240 η = 3 14 106 2 01420.880.97510.1763–0.26800.21510.2722
C 240 η = 4 12 107 2 11417.490.97620.1754–0.32770.21350.3932 2
(tetragon)
C 240 η = 5 14 106 2 01420.490.97500.1765–0.25890.2140
(0.2679)
0.2754
C 240 η = 6 14 106 2 01421.940.97590.1578–0.25880.2140
(0.2383)
0.1939
1 In the order: pentagons, hexagons, heptagons, tetragons. 2 For tetragon. 3 Curvatures of static pentagons, which resulted due to SWw, are in parentheses.
Table 2. Topological descriptors of the original and defected C240 fullerenes.
Table 2. Topological descriptors of the original and defected C240 fullerenes.
Fullerene CageWM 1 w _ w ¯ ρ E
C240 (Ih)277,44019115611561
C 240 η = 0 277,1221911401167.51.0241
C 240 η = 1 276,730191125.51171.51.0409
C 240 η = 2 276,607191118.51178.51.0536
C 240 η = 3 276,5571911241185.51.0547
C 240 η = 4 276,607191118.51192.51.0662
C 240 η = 5 276,5601911251199.51.0662
C 240 η = 6 276,5511911201196.51.0683
1 M is the graph diameter M = max{dij}.
Table 3. Energetic and molecular parameters of original and defected C240 fullerenes (DFT computations).
Table 3. Energetic and molecular parameters of original and defected C240 fullerenes (DFT computations).
Fullerene CageRelative Energy, Erel (kJ/mol) 1Energy Effect, ΔE (kJ/mol) 2Dipole Moment, μ (D)εHOMO (eV)εLUMO (eV)HOMO–LUMO Gap (eV)
C240 (Ih)0n/a0–5.53–4.301.23
C 240 η = 0 147.8147.80.86–5.26–4.460.80
C 240 η = 1 315.5167.70.51–5.18–4.410.76
C 240 η = 2 430.4114.91.57–5.30–4.570.72
C 240 η = 3 510.580.12.53–4.98–4.700.28
C 240 η = 4 642.6132.10.83–4.89–4.610.27
C 240 η = 5 489.2–153.54.62–4.94–4.830.10
C 240 η = 6 384.0–105.23.57–5.03–4.790.24
1 Calculated as E r e l = E C 240 η E C 240 o r i g i n a l . 2 Calculated as Δ E = E C 240 η E C 240 η 1 .

Share and Cite

MDPI and ACS Style

Sabirov, D.S.; Ori, O. Skeletal Rearrangements of the C240 Fullerene: Efficient Topological Descriptors for Monitoring Stone–Wales Transformations. Mathematics 2020, 8, 968. https://doi.org/10.3390/math8060968

AMA Style

Sabirov DS, Ori O. Skeletal Rearrangements of the C240 Fullerene: Efficient Topological Descriptors for Monitoring Stone–Wales Transformations. Mathematics. 2020; 8(6):968. https://doi.org/10.3390/math8060968

Chicago/Turabian Style

Sabirov, Denis Sh., and Ottorino Ori. 2020. "Skeletal Rearrangements of the C240 Fullerene: Efficient Topological Descriptors for Monitoring Stone–Wales Transformations" Mathematics 8, no. 6: 968. https://doi.org/10.3390/math8060968

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