Next Article in Journal
Sn(IV)porphyrin-Anchored TiO2 Nanoparticles via Axial-Ligand Coordination for Enhancement of Visible Light-Activated Photocatalytic Degradation
Next Article in Special Issue
QSPR and Nano-QSPR: Which One Is Common? The Case of Fullerenes Solubility
Previous Article in Journal
Circular Dichroism Spectroscopic Studies on Solution Chemistry of M(II)-Monensinates in Their Competition Reactions
Previous Article in Special Issue
Rational Design of Fullerene Derivatives for Improved Stability of p-i-n Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Matching Polynomial-Based Similarity Matrices and Descriptors for Isomers of Fullerenes

by
Krishnan Balasubramanian
School of Molecular Sciences, Arizona State University, Tempe, AZ 85287-1604, USA
Inorganics 2023, 11(8), 335; https://doi.org/10.3390/inorganics11080335
Submission received: 17 July 2023 / Revised: 9 August 2023 / Accepted: 11 August 2023 / Published: 13 August 2023
(This article belongs to the Special Issue Advances in Fullerene Science)

Abstract

:
I have computed the matching polynomials of a number of isomers of fullerenes of various sizes with the objective of developing molecular descriptors and similarity measures for isomers of fullerenes on the basis of their matching polynomials. Two novel matching polynomial-based topological descriptors are developed, and they are demonstrated to have the discriminating power to contrast a number of closely related isomers of fullerenes. The number of ways to place up to seven disjoint dimers on fullerene isomers are shown to be identical, as they are not structure-dependent. Moreover, similarity matrices that provide quantitative similarity measures among a given set of isomers of fullerenes are developed from their matching polynomials and are shown to provide robust quantitative measures of similarity.

1. Introduction

Fullerene cages, their isomers, stabilities, structures, aromaticities, electronic and magnetic properties, and spectra have been the subject of intense scrutiny over the years [1,2,3,4,5,6,7] ever since the pioneering work of Smalley and coworkers [1] that resulted in the discovery of another state of carbon with a dome-shaped icosahedral structure of C60, named buckminsterfullerene. Subsequent discovery of carbon nanotubes [8] further fueled a plethora of research papers related to fullerenes and carbon nanomaterials. Fullerenes are cage-like closed structures that contain 12 pentagons and a varied number of hexagons. Numerous isomers are possible for a fullerene with a given molecular formula; for example, there are 1812 isomers for C60 while there are 8149 isomers for the C70 fullerene.
Fullerenes and related polycyclic aromatic compounds of various kinds have attracted several theoretical and mathematical studies due to the subject matter of aromaticity, local aromaticity, global aromaticity, ring currents, electronic and magnetic properties, and so forth [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23]. Due to a large number of isomers for larger fullerenes, it is quite challenging to carry out ab initio computations on each one of them to gain insights into their structures, properties, similarities, and stabilities. Consequently, mathematical techniques primarily derived from combinatorics and graph theory such as the conjugated circuits, enumeration of Kekulé structures, sextet polynomials, matching polynomials, etc., of fullerenes and related polycyclic aromatic compounds have been studied over the decades [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23]. Furthermore, giant fullerenes pose even more computational challenges for ab initio quantum chemical studies. The existence of multiple low-lying isomers and minima in their potential energy surfaces has caused further complexity in such high-level quantum chemical studies which can be computationally quite intensive. Quantum chemical studies have been made on some of the fullerenes including their vibrational spectra [24,25,26,27]. Although fullerenes that exhibit isolated pentagon structures have been generally attributed to be more stable, recent studies have revealed the existence of stable non-isolated pentagon structures, for example, the C72(C2v)-11188 isomer [28]. Consequently, there is a clear and compelling need for the topological or graph theoretical characterization of fullerene cages, as such studies cumulatively can provide viable alternatives for gaining insights into their structures, properties, stabilities, and spectra. Although there is no direct correlation between the number of Kekulé structures and the stabilities of fullerene isomers, there appears to be a very good correlation between the overall topological resonance energies, conjugated circuits, and sextet polynomials with stabilities. Another application of graph theory is the enumeration of isomerization or rearrangement pathways that convert one isomer to the other as demonstrated in Stone–Wales rearrangement graphs [27] and the internal rotation isomerization graphs of isomers of alkanes [29] as well as water clusters and other fluxional molecules.
Topological characterization of fullerenes through the development of topological indices of various kinds has been the topic of several studies [30,31,32,33,34,35,36,37]. A number of structural invariants such as the Wiener indices, Mostar indices, and several other vertex-degree and distance-based indices have been developed to characterize the isomers of fullerenes so that they can be employed in QSPR/QSAR relations. Several graph theoretical polynomials and their spectra such as characteristic polynomials, graph spectra, matching polynomials, distance polynomials, enumeration of walks, spanning trees, Laplacians, graph automorphisms, and combinatorial enumerations of isomers of polysubstituted fullerenes, etc., have been considered over the years [38,39,40,41,42,43,44,45,46,47,48,49,50,51,52]. Among these matching polynomials of fullerenes, lattices and various other graphs and related graph polynomials have been the subject matter of several studies [40,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64]. Various graph polynomials, Laplacians, and the enumeration of spanning trees have been considered for the isomers of fullerenes [65,66,67] and holey nanographenes [68]. Furthermore, such graph theoretical techniques including the enumeration of matchings have been stimulated by applications to phase-transition phenomena and statistical mechanics [69,70,71,72,73].
The present study is stimulated by several applications of combinatorial and graph theoretical techniques for the characterization of fullerene cages. In the present study, I propose topological invariants based on matching polynomials inspired by the pioneering studies of Hosoya and coworkers [55,56,57,58,59,60,61]. In the current study, while analyzing the coefficients of matching polynomials of fullerene isomers that contain only pentagons and hexagons, it was discovered that the first several coefficients were identical for the isomers and, hence, a reduced Z-index was developed to compare the isomers. A new similarity matrix was developed to provide quantitative similarity measures among a given set of isomers of fullerenes.

2. Preliminaries and Computational Methods

The adjacency matrix of a graph is defined as:
A i j = 1   i f   v e r t i c e s   i   a n d   j   a r e   c o n n e c t e d 0   o t h e r w i s e
The characteristic polynomial of the graph, PG, is given by the secular determinant of the adjacency matrix A:
P G x = A x I = C n x n + C n 1 x n 1 + + C 1 x + C 0
where the coefficient Ck in the characteristic polynomial of a graph yields several combinatorial quantities pertinent to the structure as per Sach’s theorem:
C k = g G i ( 1 ) c ( g ) 2 r ( g )
where Gis are Sach’s subgraphs of G containing k vertices, c(g) is the disjoint components in g, and r(g) is the number of cycles in the subgraph. For example, for a fullerene, the coefficient of x10 term would be comprised of 5 disjoint dimers, 1 6-membered ring and 2 disjoint dimers not contained in the ring, 2 isolated pentagons, a 10-membered ring arising from 2 fused hexagons, and so forth. The matching polynomial, which is also referred to as the acyclic polynomial of a graph G, is defined as
M G x = k = 0 [ n 2 ] 1 k p G , k x n 2 k
where p(G, k) enumerates the number of ways to place k disjoint dimers on the graph, and n is the number of vertices while [n/2] is the greatest integer contained in n/2. For fullerenes, as n is even, it can be readily seen that the upper limit is n/2 for any fullerene Cn. The characteristic and matching polynomials of trees are completely equivalent. Likewise, the matching polynomials of monocyclic rings as well as rings with pending bonds are all readily obtained. The constant coefficient or Cn/2 of the matching polynomials enumerates the number of perfect matchings. Consequently, the constant coefficient in the matching polynomial enumerates the number of Kekulé structures for a fullerene. We note that the number of Kekulé structures alone does not provide a direct measure of the relative stability of a fullerene, although it could be used as a preliminary indicator for further perusal.
While it is well known that the computation of the matching polynomial of a highly clustered graph is both CPU and disk intensive, several techniques have been developed specific to computing the matching polynomials and perfect matchings of fullerenes over the years [74,75]. One of the important outcomes is that the labeling of the graph or alternatively the order in which the edges of the graph are to be deleted in recursive reduction is critical to the intensity of the required computations. Although the matching polynomials are invariant to the labeling of the vertices, the order in which the edges are to be chosen for recursive pruning influences the evolution and dynamics of the recursive process and, hence, the overall computational time. In another investigation, Salvador et al. [50] made use of computer linguistic tools comprising theses, lines, and grammar to compute the matching polynomials of fullerenes, although their coefficients are limited to double precision or less than 15 digits. In the present study, I employ a combination of optimal vertex labeling and recursive reduction in conjunction with quadruple precision arithmetic. Furthermore, the characteristic polynomials of all line graphs up to the needed orders, monocyclic graphs, and other recurring fragments are computed upfront and stored in a data file so that they need not be repeatedly computed to generate the polynomials. The characteristic polynomials are computed using the author’s previously developed codes enhanced further for efficiency and quadruple precision arithmetic. Hence, all the polynomial coefficients are accurate to 33–35 digits.

3. Results and Discussion

3.1. Matching Polynomials of Fullerenes

I have chosen a variety of isomers of fullerenes of varied sizes. In order to consider a contrasting case, I also included a relatively stable fullerene isomer of C58 with Cs symmetry that contains 1 heptagon and 13 pentagons. This isomer, denoted as C58(Cs)-hept, is although strictly not a fullerene, several workers [74,75] have considered this as an energetically viable low-lying isomer compared to the C58(C3v)-1 fullerene. Consequently, I have included this isomer also for the derivation of our matching polynomial-based similarity matrices. The fullerene isomers that are included in the present study are shown in Figure 1. I designate each fullerene by the number of carbon atoms, its symmetry, and a standard label as per fullerene library designations. Although there are several more isomers for each fullerene compared to the ones shown in Figure 1, I chose the isomers on the basis of their stabilities, differing symmetries, or shapes so that the similarity analysis would be meaningful and provide contrasting comparisons in order to assess the efficacy of the matching-based similarity analysis of these isomers of fullerenes. I have computed the matching polynomials of all of the fullerenes shown in Figure 1. As mentioned in the previous section, I employed a combination of recursive techniques and a binary database of previously computed and stored polynomials of the common fragments generated during the pruning process.
Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, Table 11 and Table 12 show the matching polynomials of the various isomers of fullerenes organized according to their formula. In each table, the various columns provide the matching polynomials of the isomers of a given constitution. The tables are constructed in the same order as the structures appear in Figure 1. All the results shown in Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, Table 11 and Table 12 were computed with quadruple precision accuracy and, hence, every digit in these tables is valid. Consider Table 1, which shows the matching polynomials of two isomers of C28, namely C28(D2) and C28(Td). They have different symmetries but their shapes are somewhat similar (see, Figure 1). The Td structure has less strain compared to the D2 structure. As a result of the close similarity between the Td and D2 isomers of C28, their matching polynomials are also quite similar, as can be seen in Table 1. The identical nature of the first eight coefficients of the matching polynomials of the isomers of fullerenes has nothing to do with the symmetry of the structure of C28. This arises from the fact that the first 8 coefficients of all fullerenes, that is, for the cage structures with 12 pentagons and any number of hexagons, do not depend on the structures but only on the number of carbon atoms. I shall discuss this in depth subsequently. However, it is noted that other coefficients for the C28(D2) versus C28(Td) structures also differ very little, consistent with the similarity of the shapes and other structural features, as seen from Figure 1.
Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8, Table 9 and Table 10 display the computed matching polynomials of a number of isomers of fullerenes, C30 through C58. Among these, fullerenes C36, C40, and C50 were considered for five isomers with contrasting symmetries and shapes in Table 3 and Table 5, respectively (see Figure 1 for the corresponding structures of the isomers). Table 11 and Table 12 display the matching polynomials of two isomers, C60 and C72, where for each case, two isomers of contrasting shapes or symmetries were considered. In the case of C72, the two isomers as well as C70 have been considered in quantum chemical studies [76,77]. A critical analysis of all matching polynomials displayed in the Tables reveals that for all fullerenes containing only pentagons and hexagons, the first eight coefficients are identical for the isomers in that these coefficients do not exhibit any structural dependence. That is, they vary as polynomials of n. As discussed earlier [52], the exact analytical expressions for the first few coefficients of fullerene cages can be derived through a combination of Sach’s theorem and the coefficients of the corresponding terms in the characteristic polynomials. The resulting expressions are shown below:
p(Cn:Full,0) = 1
p(Cn:Full,1) = −3n/2
p(Cn:Full,2) = 3n(3n−10)/8
p ( C n : Full , 3 ) = 1 16 ( 9 n 3 90 n 2 + 232 n )
p ( C n : Full , 4 ) = c 8 1 4 3 n 24 n 20 + 2 n 5 ( 2 )
p ( C n : Full , 5 ) = c 10 n 5 2 3 n 30 + 2 n 6 2 4 n 5 ( 2 )
where cn is the corresponding coefficient in the characteristic polynomial of the fullerene, n l ( k ) is the number of ways of choosing k adjacent l-membered rings in the fullerene whereas n l ( k ) is the number of ways to choose k disjoint l-membered rings from the fullerene.
The coefficients of the first 8 terms in the matching polynomials of all cages containing 12 pentagons and varied number of hexagons are the same for the isomers, as can be inferred to be identical from Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, Table 11 and Table 12. The only exception to this is the C58(Cs)-hept structure which is comprised of 13 pentagons and 1 heptagon and, thus, the ring structures are different compared to the C58(C3v)-1 fullerene, which contains 12 pentagons and no heptagons. Even then, the first five coefficients of the matching polynomials of the two isomers of C58 are identical, with the sixth coefficient differing only by unity. Although the results for C72(D6d) in Table 12 were derived from [50] and hence they lack the accuracy of C72(C2v)-11188 computed here, the similarity indices computed subsequently for C72 do not suffer from the accuracy issue, as the similarity measures are based on a natural logarithmic scale.
The constant coefficients of the matching polynomials yield the number of Kekulé structures of fullerene isomers, although there exists no direct correlation between the stability of the fullerene structure and the number of resonance structures. However, a number of related topological indices have been derived and used from the coefficients of the matching polynomials as well as their spectra. For example, the sum of the absolute values of the coefficients of the matching polynomials is the well-known Hosoya’s topological index [78] while the sum of the difference in the eigenvalues of the characteristic and matching polynomials yields the topological resonance energy; the latter has been employed as a measure of the relative stabilities of isomers of fullerenes. The isomers that exhibit extremal values of Hosoya’s topological Z-index [78] are also of interest. It can thus be inferred that if two isomers of fullerenes exhibit Z-indices close to each other, then they can be viewed as candidates for further investigations by a higher level of computations in order to assess further their relative stabilities. Although many such variants have been proposed, up to now, no similarity measures have been developed for comparing the isomers of fullerenes or other structures. As the first eight coefficients are identical, I have proposed the reduced Z-indices for fullerenes which consider only the differing coefficients of the matching polynomials in deriving the Z-indices. I have further introduced natural logarithms and scaling techniques for deriving the indices proposed in the next section for both the comparison and similarity analysis of fullerenes.

3.2. The Similarity Matrices of Fullerenes and Reduced Z-Indices

As can be seen from Table 1, Table 2, Table 3, Table 4, Table 5, Table 6, Table 7, Table 8, Table 9, Table 10, Table 11 and Table 12, the matching polynomials of isomers of fullerenes exhibit similarities and, hence, I develop quantitative similarity measures in terms of the similarity matrices that would have the capability to offer a contrast among isomers as well as across the platform of fullerenes. These matrices are defined using the coefficients of the matching polynomials with a scaling incorporated into them. Hence, I define the similarity matrix based on matching polynomials as follows:
S ij ( M ) = 1 n e l n ( k = 8 n 2 | p G i , λ k p G j , λ k | ,   i     j
where p G i , λ k is the kth coefficient of the matching polynomial of fullerene isomer Gi, while p G j , λ k is the kth coefficient of the matching polynomial of fullerene isomer Gj. The absolute differences of the corresponding coefficients are taken and, thus, the difference is always positive so as to maintain this as a true difference without regard for the sign variations of the alternate terms of the matching polynomials. We obtain a matrix element Sij for any two members (i, j) among a set of isomers considered for comparison. The diagonal elements of the similarity matrix are set to 0 as the similarity distance between two identical isomers is 0. Consequently, the larger the similarity matrix element, the greater is the dissimilarity between the isomers i and j, while a small value would then imply that the two isomers are very similar. I have computed the similarity matrices for all of the isomers of fullerenes considered in this study, and the computed similarity matrices are shown in Figure 2 for each fullerene considered here.
As the first eight coefficients of the matching polynomials of isomers of fullerenes are identical, I have introduced a scaled, natural logarithmic version of the reduced Z-index, ZR, as follows:
S - ln ( Z R ( M ) ) = 1 n e l n ( k = 8 n 2 p G , λ k )
A primary advantage of the reduced-scaled version is that it facilitates a comparison of isomers of fullerenes across the platform. Hence, I have shown in Figure 2 both S-ln(ZR) as well as ZR for comparing isomers, where ZR is simply the sum of the absolute coefficients starting with the eighth coefficient of the matching polynomials.
As seen from Figure 2, the computed similarity measures are in a logarithmic scale and the matrix elements vary between 0.137 and 0.313 where the lowest value corresponding to the most similar structures are for the first two isomers of C36 (Figure 1) which are C36(C2)-12 and C36(C2v)-9. As can be seen from both Figure 1 and Table 3, the two isomers are very similar in multiple ways. Their overall shapes and structural similarities are striking. At a quantitative level, an inspection of Table 3 reveals that the first 10 coefficients in the matching polynomial are identical while the 11th coefficient differs only by unity. Several other subsequent coefficients are also close to each other. This is in turn reflected by the similarity matrix element of 0.137484542 for the two isomers. Likewise, the isomers 5 and 3, which correspond to C36-D5h-15 and C36-D2d-14, exhibit remarkable similarity both in terms of their shape, structures, and matching polynomials. That is, the arrangements of pentagons and hexagons are such that they provide very similar combinatorial matchings. I note that other structures which exhibit such similarities are the two isomers of C28 in Figure 1; the two isomers have a similarity measure of 0.1677151 on the basis of their combinatorial matchings. Likewise, two isomers of C30 also exhibit comparable similarity measures (see Figure 2). The first isomers of C40 (Figure 1) have comparable similarity measures of 0.1627944 and this is corroborated by the corresponding matching polynomials shown in Table 5 where I find that the first 12 coefficients of the two isomers are identical, with the 13th coefficient differing by only 12.
Although most of the other isomers of fullerenes exhibit similarity indices close to 0.2, the two isomers of C58 are important cases to be noted for the dramatic similarity contrast. First, as noted before, their similarity index is the highest among all the isomers considered here with a striking value of 0.312761 given that this is a logarithmic scale. The contrasting similarity measure is fully consistent with the fact that the first isomer of C58 is a true fullerene containing 12 pentagons and hexagons while the second one designated as C58(Cs)-hept contains 1 heptagon and 13 pentagons. This contrasting juxtaposition shown in Figure 1 as well as Table 10 is truly echoed in their similarity index measure introduced here. This is a direct validation of the similarity matrix measure that I have developed in that the measure faithfully reflects the variations and dissimilarities as well as similarities among the structures. Moreover, with the values shown in Figure 2, I now have a reference platform to evaluate the similarities among isomers through such quantitative similarity measures.
To shed further light into the similarity matrix invariants, let us consider the five isomers of C50 shown in Figure 1 with their matching polynomials displayed in Table 8. Let us consider the computed similarity matrix which is highlighted below for the five isomers of C50 in the order:
 
C50(C2)-269, C50(C2v)-13, C50(D3)-270, C50(D5h)-271, and C50(D3h)-3.
C50(C2)-269C50(C2v)-13 C50(D3)-270 C50(D5h)-271 C50(D3h)-3
C50(C2)-2690.00000000000.27830027790.26107471090.26626304420.2850277531
C50(C2v)-130.27830027790.00000000000.28153549070.28283499700.2726842541
C50(D3)-2700.26107471090.28153549070.00000000000.25117350660.2870732104
C50(D5h)-2710.26626304420.28283499700.25117350660.00000000000.2879451010
C50(D3h)-30.28502775310.27268425410.28707321040.28794510100.0000000000
The above array suggests that the smallest matrix element (0.251174) is between the isomers 3 and 4, while the largest matrix element is between the isomers 4 and 5 (0.2879451010). I now refer to Figure 1, where indeed I find the isomers 3 and 4, C50(D3)-270 and C50(D5h)-271, which are quite similar in their shapes and overall structural features. On the other hand, the isomers C50(D5h)-271 and C50(D3h)-3 are extremely dissimilar in that the latter is an oblate spheroid while the former is more spherical. Likewise, as can be seen from the fifth row of the similarity matrix, the oblate spheroidal C50(D3h)-3 stands out in having larger matrix elements with the entire array of other isomers of C50 considered here. This is consistent with the fact that the C50(D3h)-3 isomer is conspicuous among the five isomers of C50 in being an oblate spheroid while the other four isomers are closer to spherical structures (See Figure 1).
I note from Figure 2 that although ZR increases rapidly as a function of the number of atoms in fullerenes, the scaled-logarithmic version can be used to make comparisons. As pointed out by Hosoya [78], the Z-index by itself does not correlate with the aromaticity or stability of polycyclic aromatics. However, the reduced index ZR can provide first-order information on the total number of resonance structures and possible full and partial matchings. If I consider the two isomers of C60, their ZR values are 1,417,033,687,086,496 and 1,417,370,147,605,744 for the Ih and D3 isomers, respectively. Although the numbers of the resonance structures of the Ih and D3 structures are 12,500 and 9622, respectively, their ZR indices exhibit an opposite trend with the Ih isomer exhibiting an overall lower ZR index. The lower overall ZR for the Ih isomer together with the greater number of resonance structures for the Ih structure suggests a considerably enhanced stability for the Ih isomer. This is consistent with the DFT quantum chemical studies on these isomers which reveal that the D3 isomer of C60 is higher in energy [76]. I find a similar correlation for other fullerenes such as C50 and C36 with the cautionary note that there is no direct correlation between the relative stability and the ZR indices as well as the total number of resonance structures.
Finally, there appears to be a correlation between the shapes of fullerene structures and the combinatorial matching-based similarity indices. For example, nearly spherical structures have very close similarity indices while a fullerene isomer with an oblate spheroid structure exhibits a numerically larger value of the similarity index when compared to more spherical structures. Likewise, two oblate spheroid isomers have closer similarity and, thus, a smaller similarity matrix element. The subject matter of quantifying shapes and QShAR has received attention over the years [79,80]. Consequently, the present similarity matrices derived from the matchings add yet another novel dimension to the shape similarity problem. The similarity indices derived here based on combinatorial matchings could find applications in water clusters [81] where the hydrogen bonds between any two water molecules could become matchings. Moreover, dimer covers could also model placing dimers such as transition metal dimers [82] that avoid being neighbors and, thus, could also serve as models for the chemisorption or substitution of dimeric molecules on fullerene cages and nanotubes.

4. Conclusions

In retrospect, I have developed powerful similarity measures using matrix invariants derived from the matching polynomials. These similarity matrices were applied to isomers of fullerenes, and it was demonstrated that the similarity matrix measures are quite robust in providing quantitative measures of similarity of two fullerene isomers. It also seems that the techniques provide some indirect measures of shape similarities of fullerene isomers. There are a few limitations that should be pointed out. The techniques developed might not provide much contrast for isospectral graphs and isospectral trees. In particular, for isospectral trees, the matching polynomials and characteristic polynomials become degenerate. Likewise, some isospectral structures that contain rings with pending fragments might not be contrasted by the matching polynomial-based methods. Babić [83] has shown the existence of isospectral benzenoid graphs containing 33 vertices and 9 hexagons. Likewise, the author and Basak [84] have illustrated isospectral benzenoid graphs with pendant bonds. Yet, the techniques based on matching polynomials appear to provide considerable promise for molecular structures containing several rings, three-dimensional fullerene cages, and carbon nanotubes. Graph theoretical techniques analogous to the ones developed here in conjunction with group theory and combinatorics can also be applied to NMR, ESR, and vibrational spectroscopies [85], thus paving the way for the applications of the emerging field of artificial intelligence.

Funding

This research received no funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data used in the manuscript are contained in the manuscript.

Conflicts of Interest

The author declares 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. Kroto, H.W.; Heath, J.R.; O’Brien, S.C.; Curl, R.F.; Smalley, R.E. Long carbon chain molecules in circumstellar shells. Astrophys. J. 1987, 314, 352–355. [Google Scholar] [CrossRef]
  3. Diederich, F.; Whetten, R.L.; Thilgen, C.; Ettl, R.; Chao, I.T.O.; Alvarez, M.M. Fullerene isomerism: Isolation of C2v,-C78 and D3-C78. Science 1991, 254, 1768–1770. [Google Scholar] [CrossRef] [PubMed]
  4. Simeonov, K.S.; Amsharov, K.Y.; Jansen, M. Chlorinated Derivatives of C78-Fullerene Isomers with Unusually Short Intermolecular Halogen–Halogen Contacts. Chem.-A Eur. J. 2008, 14, 9585–9590. [Google Scholar] [CrossRef]
  5. Simeonov, K.S.; Amsharov, K.Y.; Krokos, E.; Jansen, M. An epilogue on the C78-fullerene family: The discovery and characterization of an elusive isomer. Angew. Chem. Int. Ed. 2008, 47, 6283–6285. [Google Scholar] [CrossRef] [PubMed]
  6. Diederich, F.; Ettl, R.; Rubin, Y.; Whetten, R.L.; Beck, R.; Alvarez, M.; Koch, A. The higher fullerenes: Isolation and characterization of C76, C84, C90, C94, and C70O, an oxide of D5h-C70. Science 1991, 252, 548–551. [Google Scholar] [CrossRef] [PubMed]
  7. Terrones, M.; Terrones, G.; Terrones, H. Structure, Chirality, and Formation of Giant Icosahedral Fullerenes and Spherical Graphitic Onions. In Science of Crystal Structures; Springer: Cham, Switzerland, 2015; pp. 101–112. [Google Scholar]
  8. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. [Google Scholar] [CrossRef]
  9. Aihara, J. Graph theory of aromatic stabilization. Bull. Chem. Soc. Jpn. 2016, 89, 1425–1454. [Google Scholar] [CrossRef] [Green Version]
  10. Aihara, J. Graph-theoretical formulation of London diamagnetism. J. Am. Chem. Soc. 1979, 101, 5913–5917. [Google Scholar] [CrossRef]
  11. Aihara, J.; Horikawa, T. Graph-theoretical formula for ring currents induced in a polycyclic conjugated system. Bull. Chem. Soc. Jpn. 1983, 56, 1853–1854. [Google Scholar] [CrossRef] [Green Version]
  12. Aihara, J. Circuit resonance energy: A key quantity that links energetic and magnetic criteria of aromaticity. J. Am. Chem. Soc. 2006, 128, 2873–2879. [Google Scholar] [CrossRef] [PubMed]
  13. Aihara, J. Topological resonance energy, bond resonance energy, and circuit resonance energy. J. Phys. Org. Chem. 1979, 21, 79–85. [Google Scholar] [CrossRef]
  14. Dias, J.R. Valence-bond determination of diradical character of polycyclic aromatic hydrocarbons: From acenes to rectangular benzenoids. J. Phys. Chem. A. 2013, 117, 4716–4725. [Google Scholar] [CrossRef]
  15. Dias, J.R. The Formula Periodic Table for Benzenoid Hydrocarbons and the Unifying Theory of a Periodic Table Set. Polycycl. Aromat. Compd. 1994, 4, 87–106. [Google Scholar] [CrossRef]
  16. Dias, J.R. Search for singlet-triplet bistability or biradicaloid properties in polycyclic conjugated hydrocarbons: A valence-bond analysis. Mol. Phys. 2012, 111, 735–751. [Google Scholar] [CrossRef]
  17. Dias, J.R. What do I know about C24H14 benzenoid, fluoranthenoid, and indacenoid hydrocarbons? Polycycl. Aromat. Compd. 2014, 34, 177–190. [Google Scholar] [CrossRef]
  18. Dias, J.R. Nonplanarity index for fused benzenoid hydrocarbons. Polycycl. Aromat. Compd. 2014, 34, 161–176. [Google Scholar] [CrossRef]
  19. Aihara, J.; Makino, M.; Ishida, T.; Dias, J.R. Analytical study of superaromaticity in cycloarenes and related coronoid hydrocarbons. J. Phys. Chem. A 2013, 117, 4688–4697. [Google Scholar] [CrossRef]
  20. Aihara, J. On the number of aromatic sextets in a benzenoid hydrocarbon. Bull. Chem. Soc. Jpn. 1976, 49, 1429–1430. [Google Scholar] [CrossRef] [Green Version]
  21. Randić, M. Aromaticity of polycyclic conjugated hydrocarbons. Chem. Rev. 2003, 103, 3449–3606. [Google Scholar] [CrossRef]
  22. Aihara, J. Aromaticity and superaromaticity in cyclopolyacenes. J. Chem. Soc. Perkin Trans. 1994, 2, 971–974. [Google Scholar] [CrossRef]
  23. Randić, M.; Balaban, A.T. Local aromaticity and aromatic sextet theory beyond Clar. Int. J. Quant. Chem. 2018, 118, e25657. [Google Scholar] [CrossRef]
  24. Slanina, Z.; Jerzy, M.; Rudziński, M.T.; Ōsawa, E. Quantum-chemically supported vibrational analysis of giant molecules: The C60 and C70 clusters. J. Mol. Struct. THEOCHEM 1989, 202, 169–176. [Google Scholar] [CrossRef]
  25. Slanina, Z.; Lee, S.L.; Yuf, C.H. Computations in treating fullerenes and carbon aggregates. Rev. Comput. Chem. 2009, 8, 1. [Google Scholar]
  26. Balasubramanian, K. Combinatorics of Supergiant Fullerenes: Enumeration of Polysubstituted Isomers, Chirality, Nuclear Magnetic Resonance, Electron Spin Resonance Patterns, and Vibrational Modes from C70 to C150000. J. Phys. Chem. A 2020, 124, 10359–10383. [Google Scholar] [CrossRef]
  27. Ōsawa, E.; Ueno, H.; Yoshida, M.; Slanina, Z.; Zhao, X.; Nishiyama, M.; Saito, H. Combined topological and energy analysis of the annealing process in fullerene formation. Stone–Wales interconversion pathways among IPR isomers of higher fullerenes. J. Chem. Soc. Perkin Trans. 1998, 2, 943–950. [Google Scholar] [CrossRef]
  28. Moreno-Vicente, A.; Abella, L.; Azmani, K.; Rodriguez-Fortea, A.; Poblet, J.M. Formation of C2v-C72 (11188)Cl4: A Particularly Stable Non-IPR Fullerene. J. Phys. Chem. A 2018, 122, 2288–2296. [Google Scholar] [CrossRef]
  29. Balasubramanian, K. Enumeration of Internal-Rotation Reactions and Their Reaction Graphs. Theor. Chim. Acta 1979, 53, 129–146. [Google Scholar] [CrossRef]
  30. Balasubramanian, K. Topological peripheral shapes and distance-based characterization of fullerenes C20-C720: Existence of isoperipheral fullerenes. Polycycl. Aromat. Compd. 2022, 42, 1649–1667. [Google Scholar] [CrossRef]
  31. Balaban, A.T.; Liu, X.; Klein, D.J.; Babić, D.; Schmalz, T.G.; Seitz, W.A.; Randi’c, M. Graph invariants for fullerenes. J. Chem. Inf. Comput. Sci. 1995, 35, 396–404. [Google Scholar] [CrossRef]
  32. Bača, M.; Horváthová, J.; Mokrišová, M.; Suhányiová, A. On topological indices of fullerenes. Appl. Math. Comput. 2015, 251, 154–161. [Google Scholar] [CrossRef]
  33. Alizadeh, Y.; Iranmanesh, A.; Mirzaie, S. Computing Schultz polynomial, Schultz index of C60 fullerene by gap program. Dig. J. Nanomater. Biostructures 2009, 4, 7–10. [Google Scholar]
  34. Ghorbani, M. Connective eccentric index of fullerenes. J. Math. Nanosci. 2011, 1, 43–50. [Google Scholar]
  35. Behmaram, A.; Yousefi-Azari, H.; Ashrafi, A.R. Wiener polarity index of fullerenes and hexagonal systems. Appl. Math. Lett. 2012, 25, 1510–1513. [Google Scholar] [CrossRef] [Green Version]
  36. Yousefi-Azari, H.; Khalifeh, M.H.; Ashrafi, A.R. Calculating the edge-Wiener and Szeged indices of graphs. J. Comput. Appl. Math. 2011, 235, 4866–4870. [Google Scholar] [CrossRef] [Green Version]
  37. Vakili-Nezhaad, G.R.; Mansoori, G.A.; Ashrafi, A.R. Symmetry property of fullerenes. J. Comput. Theor. Nanosci. 2007, 4, 1202–1205. [Google Scholar] [CrossRef]
  38. Balasubramanian, K. Characteristic Polynomials of Fullerene Cages. Chem. Phys. Lett. 1992, 198, 577–586. [Google Scholar] [CrossRef]
  39. Balasubramanian, K. Distance spectra and distance polynomials of fullerenes. J. Phys. Chem. 1995, 99, 10785–10796. [Google Scholar] [CrossRef]
  40. Ramaraj, R.; Balasubramanian, K. Computer Generation of Matching Polynomials of Chemical Graphs and Lattices. J. Comput. Chem. 1985, 6, 122–141. [Google Scholar] [CrossRef]
  41. Balasubramanian, K. Computer Generation of the Characteristic-Polynomials of Chemical Graphs. J. Comput. Chem. 1984, 5, 387–394. [Google Scholar] [CrossRef]
  42. Balasubramanian, K. Computer Generation of Distance Polynomials of Graphs. J. Comput. Chem. 1990, 11, 829–836. [Google Scholar] [CrossRef]
  43. Balasubramanian, K. Characteristic-Polynomials of Organic Polymers and Periodic Structures. J. Comput. Chem. 1985, 6, 656–661. [Google Scholar] [CrossRef]
  44. Manoharan, M.; Balakrishnarajan, M.; Venuvanalingam, P.; Balasubramanian, K. Topological resonance energy predictions of the stability of fullerene clusters. Chem. Phys. Lett. 1994, 222, 95–100. [Google Scholar] [CrossRef]
  45. Babić, D.; Ori, O. Matching polynomial and topological resonance energy of C70. Chem. Phys. Lett. 1995, 234, 240–244. [Google Scholar] [CrossRef]
  46. Liu, J.; Shan, B.Z. The Computation of the Matching Polynomials of Fullerene Graphs. Croat. Chem. Acta. 1997, 70, 971–977. [Google Scholar]
  47. Aihara, J.I.; Babić, D.; Gutman, I. Matching spectra of fullerenes. MATCH Commun. Math. Comput. Chem. 1996, 33, 7–16. [Google Scholar]
  48. Herndon, W.C.; Radhakrishnan, T.P.; Zivković, T.P. Characteristic and matching polynomials of chemical graphs. Chem. Phys. Lett. 1988, 152, 233–238. [Google Scholar] [CrossRef]
  49. Liu, Y.; Ma, C.; Yao, H.; Wang, X. Computing the forcing and anti-forcing numbers of perfect matchings for graphs by integer linear programmings. MATCH Commun. Math. Comput. Chem. 2022, 87, 561–575. [Google Scholar] [CrossRef]
  50. Salvador, J.M.; Hernandez, A.; Beltran, A.; Duran, R.; Mactutis, A. Fast partial-differential synthesis of the matching polynomial of C72-100. J. Chem. Inf. Comput. Sci. 1988, 38, 1105–1110. [Google Scholar] [CrossRef]
  51. Shi, Y.; Dehmer, M.; Li, X.; Gutman, I. (Eds.) Graph Polynomials; CRC Press: Boca Raton, FL, USA, 2017; p. 239. [Google Scholar]
  52. Balasubramanian, K. Matching Polynomials of Fullerene Clusters. Chem. Phys. Lett. 1993, 201, 306–314. [Google Scholar] [CrossRef]
  53. Balasubramanian, K. Graph-Theory and the PPP Method. J. Math. Chem. 1991, 7, 353–362. [Google Scholar] [CrossRef]
  54. Balasubramanian, K.; Ramaraj, R. Computer Generation of King and Color Polynomials of Graphs and Lattices and Their Applications to Statistical Mechanics. J. Comput. Chem. 1985, 6, 447–454. [Google Scholar] [CrossRef]
  55. Hosoya, H. Matching and symmetry of graphs. Comput. Math. 1986, 12, 271–290. [Google Scholar]
  56. Hosoya, H.; Motoyama, A. An effective algorithm for obtaining polynomials for dimer statistics. Application of operator technique on the topological index to two-and three-dimensional rectangular and torus lattices. J. Math. Phys. 1985, 26, 157–167. [Google Scholar] [CrossRef]
  57. Hosoya, H. On some counting polynomials in chemistry. Discrete Appl. Math. 1988, 19, 239–257. [Google Scholar] [CrossRef] [Green Version]
  58. Cash, C.G. Relationship between the Hosoya polynomial and the hyper-Wiener index. Appl. Math. Lett. 2002, 15, 893–895. [Google Scholar] [CrossRef]
  59. Hosoya, H. Topological index. A newly proposed quantity characterizing the topological nature of structural isomers of saturated hydrocarbons. Bull. Chem. Soc. Jpn. 1971, 44, 2332–2339. [Google Scholar] [CrossRef] [Green Version]
  60. Hosoya, H.; Balasubramanian, K. Computational Algorithms for Matching Polynomials of Graphs from the Characteristic Polynomials of Edge-Weighted Graphs. J. Comput. Chem. 1989, 10, 698–710. [Google Scholar] [CrossRef]
  61. Hosoya, H.; Balasubramanian, K. Exact Dimer Statistics and Characteristic-Polynomials of Cacti Lattices. Theor. Chim. Acta 1989, 76, 315–329. [Google Scholar] [CrossRef]
  62. Balasubramanian, K. Tree Pruning Method and Lattice Statistics on Bethe Lattices. J. Math. Chem. 1988, 2, 69–82. [Google Scholar] [CrossRef]
  63. Balasubramanian, K. Spectra of Chemical Trees. Int. J. Quant. Chem. 1982, 21, 581–590. [Google Scholar] [CrossRef]
  64. Balasubramanian, K. Spectra of Matching Polynomials. Chem. Phys. Lett. 1993, 208, 219–224. [Google Scholar] [CrossRef]
  65. Balasubramanian, K. Topological characterization of five C78 fullerene isomers. Chem. Phys. Lett. 1993, 206, 210–216. [Google Scholar] [CrossRef]
  66. Balasubramanian, K. Laplacians of Fullerenes (C42–C90). J. Chem. Phys. Chem. 1995, 99, 6509–6518. [Google Scholar] [CrossRef]
  67. Balasubramanian, K. Computation of Spanning Tree Generators of Fullerenes. Mol. Phys. 1995, 85, 335–348. [Google Scholar] [CrossRef]
  68. Balasubramanian, K. Density Functional and Graph Theory Computations of Vibrational, Electronic and Topological Properties of Porous Nanographenes. J. Phys. Org. Chem. 2022, e4435. [Google Scholar] [CrossRef]
  69. Kasteleyn, P.W. Graph Theory and Theoretical Physics; Harary, F., Ed.; Academic Press: London, UK, 1967. [Google Scholar]
  70. Percus, J.K. Combinatorial Methods; Springer Science & Business Media: New York, NY, USA, 1969. [Google Scholar]
  71. Temperley, H.N.V. Phase Transition and Critical Phenomena; Domb, C., Green, M.S., Eds.; Academic Press: London, UK, 1972; Volume I, p. 227. [Google Scholar]
  72. Park, J.; Newman, M.E.J. Statistical mechanics of networks. Phys. Rev. B 2004, 70, 066117. [Google Scholar]
  73. Kaatz, F.H.; Bultheel, A. Informational thermodynamic model for nanostructures. J. Math. Chem. 2014, 52, 1563–1575. [Google Scholar] [CrossRef] [Green Version]
  74. Chen, D.L.; Tian, W.Q.; Feng, J.K.; Sun, C.C. Structures, stabilities, and electronic and optical properties of C58 fullerene isomers, ions, and metallofullerenes. ChemPhysChem 2007, 8, 1029–1036. [Google Scholar] [CrossRef]
  75. Bihlmeier, A.; Klopper, W. Hydrides and dimers of C58 fullerenes: Structures and stabilities. Phys. Chem. Chem. Phys. 2009, 11, 1050–1059. [Google Scholar]
  76. Zettergren, H.; Alcamí, M.; Martín, F. First-and second-electron affinities of C60 and C70 isomers. Phys. Rev. A 2007, 76, 043205. [Google Scholar] [CrossRef]
  77. Slanina, Z.; Ishimura, K.; Kobayashi, K.; Nagase, S. C72 isomers: The IPR-satisfying cage is disfavored by both energy and entropy. Chem. Phys. Lett. 2004, 384, 114–118. [Google Scholar] [CrossRef]
  78. Hosoya, H. Aromaticity Index Can Predict and Explain the Stability of Polycyclic Conjugated Hydrocarbons. Mon. Für Chem. 2005, 36, 1037–1054. [Google Scholar]
  79. Walker, P.D.; Maggiora, G.M.; Johnson, M.A.; Petke, J.D.; Mezey, P.G. Shape group analysis of molecular similarity: Shape similarity of six-membered aromatic ring systems. J. Chem. Inf. Comput. Sci. 1995, 35, 568–578. [Google Scholar]
  80. Mezey, P.G. Fuzzy electron density fragments in macromolecular quantum chemistry, combinatorial quantum chemistry, functional group analysis, and shape-activity relations. Acc. Chem. Res. 2014, 47, 2821–2827. [Google Scholar] [PubMed]
  81. Balasubramanian, K. Nonrigid group theory, tunneling splittings, and nuclear spin statistics of water pentamer:(H2O)5. J. Phys. Chem. A 2004, 108, 5527–5536. [Google Scholar]
  82. Balasubramanian, K.; Liao, D.W. Spectroscopic properties of low-lying electronic states of rhodium dimer. J. Phys. Chem. 1989, 93, 3989–3992. [Google Scholar]
  83. Babić, D. Isospectral benzenoid graphs with an odd number of vertices. J. Math. Chem. 1993, 12, 137–146. [Google Scholar] [CrossRef]
  84. Balasubramanian, K.; Basak, S.C. Characterization of isospectral graphs using graph invariants and derived orthogonal parameters. J. Chem. Inf. Comput. Sci. 1998, 38, 367–373. [Google Scholar] [CrossRef]
  85. Balasubramanian, K. Symmetry, Combinatorics, Artificial Intelligence, Music and Spectroscopy. Symmetry 2021, 13, 1850. [Google Scholar] [CrossRef]
Figure 1. Structures of fullerene isomers C28-C72 considered in this study for similarity matrices.
Figure 1. Structures of fullerene isomers C28-C72 considered in this study for similarity matrices.
Inorganics 11 00335 g001aInorganics 11 00335 g001bInorganics 11 00335 g001cInorganics 11 00335 g001dInorganics 11 00335 g001eInorganics 11 00335 g001f
Figure 2. Similarity matrices and reduced Z-indices and scaled reduced ln(Z-indices) of fullerene isomers considered in this study.
Figure 2. Similarity matrices and reduced Z-indices and scaled reduced ln(Z-indices) of fullerene isomers considered in this study.
Inorganics 11 00335 g002aInorganics 11 00335 g002b
Table 1. Matching polynomials of two isomers of C28 fullerene.
Table 1. Matching polynomials of two isomers of C28 fullerene.
kC28(D2)C28(Td)
011
1−42−42
2777777
3−8344−8344
457,70857,708
5−269,628−269,628
6868,440868,440
7−1,932,444−1,932,444
82,932,0102,932,008
9−2,944,736−2,944,708
101,859,7961,859,652
11−678,656−678,312
12123,782123,387
13−8492−8274
149075
Table 2. Matching polynomials of isomers of C30 fullerene.
Table 2. Matching polynomials of isomers of C30 fullerene.
kC30(D5h)-1C30(C2v)-2C30(C2v)-3
0111
1−45−45−45
2900900900
3−10,560−10,560−10,560
480,82080,82080,820
5−424,392−424,392−424,392
61,566,0651,566,0651,566,065
7−4,091,265−4,091,265−4,091,265
87,524,7707,524,7687,524,767
9−9,568,000−9,567,966−9,567,949
108,137,5518,137,3278,137,216
11−4,388,255−4,387,529−4,387,172
121,377,4201,376,1981,375,602
13−217,960−216,930−216,418
1413,26512,86712,661
15−151−117−107
Table 3. Matching polynomials of isomers of C36 fullerene.
Table 3. Matching polynomials of isomers of C36 fullerene.
kC36(C2)-12C36(C2v)-9C36(D2d)-14C36(D3h)-13C36(D6h)-15
011111
1−54−54−54−54−54
213231323132313231323
3−19,476−19,476−19,476−19,476−19,476
4192,321192,321192,321192,321192,321
5−1,346,910−1,346,910−1,346,910−1,346,910−1,346,910
66,898,0196,898,0196,898,0196,898,0196,898,019
7−26,255,052−26,255,052−26,255,052−26,255,052−26,255,052
874,743,46874,743,46874,743,46774,743,47074,743,467
9−158,920,900−158,920,900−158,920,874−158,920,952−158,920,874
10250,185,492250,185,493250,185,213250,186,053250,185,213
11−286,863,270−286,863,284−286,861,644−286,866,564−286,861,644
12233,454,871233,454,925233,449,143233,466,479233,449,135
13−129,759,156−129,759,130−129,746,290−129,784,686−129,746,178
1446,513,09746,512,60546,494,50946,548,27946,494,021
15−9,838,170−9,837,524−9,822,212−9,867,396−9,821,428
161,057,1031,057,2531,050,7961,070,0071,050,468
17−43,008−43,278−42,320−45,186−42,288
18289312288364272
Table 4. Matching polynomials of isomers of C38 fullerene.
Table 4. Matching polynomials of isomers of C38 fullerene.
kC38(C2)-13C38(C2)-17C38(C2)-6C38(C3v)-16
01111
1−57−57−57−57
21482148214821482
3−23,294−23,294−23,294−23,294
4247,323247,323247,323247,323
5−1,877,511−1,877,511−1,877,511−1,877,511
610,521,46110,521,46110,521,46110,521,461
7−44,311,485−44,311,485−44,311,485−44,311,485
8141,457,329141,457,328141,457,331141,457,329
9−342,789,923−342,789,894−342,789,983−342,789,924
10627,517,764627,517,410627,518,522627,517,788
11−858,202,534−858,200,138−858,207,810−858,202,755
12860,996,587860,986,593861,018,781860,997,544
13−616,691,139−616,664,107−616,749,327−616,692,840
14303,148,423303,100,171303,243,068303,147,594
15−96,506,463−96,451,021−96,598,748−96,498,792
1618,197,16618,159,92318,247,44718,186,867
17−1,750,486−1,738,799−1,763,805−1,745,700
1864,19063,20465,38563,675
19−386−382−385−378
Table 5. Matching polynomials of isomers of C40 fullerene.
Table 5. Matching polynomials of isomers of C40 fullerene.
kC40(C2)-35C40(C2)-36C40(C2v)-37C40(D2)-38C40(D5d)-39
011111
1−60−60−60−60−60
216501650165016501650
3−27,580−27,580−27,580−27,580−27,580
4313,335313,335313,335313,335313,335
5−2,563,260−2,563,260−2,563,260−2,563,260−2,563,260
615,606,39015,606,39015,606,39015,606,39015,606,390
7−72,094,680−72,094,680−72,094,680−72,094,680−72,094,680
8255,308,426255,308,426255,308,426255,308,425255,308,425
9−695,619,674−695,619,674−695,619,674−695,619,640−695,619,640
101,455,391,4941,455,391,4941,455,391,4941,455,391,0021,455,391,002
11−2,321,341,062−2,321,341,062−2,321,341,066−2,321,337,096−2,321,337,100
122,786,393,2302,786,393,2422,786,393,3292,786,373,6862,786,373,750
13−2,468,240,914−2,468,241,118−2,468,241,896−2,468,180,232−2,468,180,640
141,568,689,2301,568,690,5741,568,694,2491,568,571,6261,568,572,975
15−687,082,056−687,086,328−687,096,074−686,946,860−686,949,378
16195,552,995195,559,583195,573,824195,471,357195,473,975
17−33,038,938−33,043,186−33,053,906−33,020,784−33,022,120
182,846,5002,847,2042,851,2692,848,7662,849,295
19−93,008−92,940−93,740−94,080−94,470
20493473513518562
Table 6. Matching polynomials of isomers of C44 fullerene.
Table 6. Matching polynomials of isomers of C44 fullerene.
kC44(C2)-1C44(D2)-2C44(D3d)-3
0111
1−66−66−66
2201320132013
3−37,664−37,664−37,664
4483,978483,978483,978
5−4,531,152−4,531,152−4,531,152
632,000,46232,000,46232,000,462
7−174,145,908−174,145,908−174,145,908
8739,662,349739,662,351739,662,351
9−2,468,621,824−2,468,621,902−2,468,621,902
106,487,128,8116,487,130,1456,487,130,145
11−13,393,750,298−13,393,763,496−13,393,763,496
1221,594,289,60621,594,373,48921,594,373,484
13−26,906,332,278−26,906,691,854−26,906,691,732
1425,516,302,64925,517,364,44725,517,363,112
15−18,028,774,350−18,030,941,516−18,030,933,108
169,216,119,4619,219,142,4729,219,110,019
17−3,272,339,730−3,275,146,364−3,275,069,730
18761,369,684763,026,684762,920,396
19−106,313,844−106,892,800−106,813,188
207,699,3887,807,2797,780,449
21−215,950−225,134−222,470
2289210911170
Table 7. Matching Polynomials of Isomers of C48 Fullerene.
Table 7. Matching Polynomials of Isomers of C48 Fullerene.
kC48(C2)-1C48(D2)-2
011
1−72−72
224122412
3−49,944−49,944
4716,238716,238
5−7,554,444−7,554,444
660,745,32260,745,322
7−380,928,456−380,928,456
81,890,083,4851,890,083,487
9−7,486,060,102−7,486,060,192
1023,775,570,46023,775,572,268
11−60,611,207,684−60,611,229,132
12123,760,350,554123,760,518,011
13−201,341,648,072−201,342,555,836
14258,764,778,060258,768,290,312
15−259,526,848,576−259,536,667,008
16199,760,333,462199,780,185,175
17−115,384,812,402−115,413,591,260
1848,529,535,47148,558,888,146
19−14,261,258,162−14,281,668,112
202,761,839,2682,771,056,042
21−322,801,582−325,317,240
2219,673,19120,047,888
23−465,508−491,272
241,5322,024
Table 8. Matching polynomials of isomers of C50 fullerene.
Table 8. Matching polynomials of isomers of C50 fullerene.
kC50(C2)-269C50(C2v)-13C50(D3)-270C50(D5h)-271C50(D3h)-3
011111
1−75−75−75−75−75
226252625262526252625
3−56,975−56,975−56,975−56,975−56,975
4859,575859,575859,575859,575859,575
5−9,576,453−9,576,453−9,576,453−9,576,453−9,576,453
681,704,03081,704,03081,704,03081,704,03081,704,030
7−546,377,070−546,377,070−546,377,070−546,377,070−546,377,070
82,907,494,4832,907,494,4892,907,494,4812,907,494,4802,907,494,493
9−12,430,405,477−12,430,405,765−12,430,405,379−12,430,405,330−12,430,405,959
1042,930,480,51042,930,486,72442,930,478,35642,930,477,27942,930,490,950
11−120,030,365,482−120,030,445,280−120,030,337,392−120,030,323,355−120,030,500,010
12271,475,360,667271,476,041,210271,475,118,473271,474,997,635271,476,511,246
13−494,947,493,439−494,951,565,536−494,946,037,473−494,945,313,135−494,954,392,290
14722,829,954,436722,847,560,835722,823,685,011722,820,579,630722,859,812,505
15−837,705,092,102−837,760,905,963−837,685,538,979−837,675,911,840−837,799,701,683
16760,514,520,422760,644,785,556760,470,388,059760,448,826,260760,734,797,436
17−531,573,181,792−531,795,993,434−531,501,952,881−531,467,488,160−531,948,041,766
18279,547,624,962279,823,262,630279,467,433,844279,429,116,630280,007,334,171
19−107,244,065,832−107,485,053,846−107,183,756,466−107,155,471,120−107,640,515,031
2028,768,905,08528,912,538,93428,740,833,45128,728,154,28029,000,371,557
21−5,080,606,155−5,135,852,304−5,073,704,083−5,070,966,660−5,166,984,589
22539,267,259551,856,582538,762,086538,792,490558,135,165
23−29,582,483−31,053,612−29,652,450−29,762,020−31,680,375
24623,747690,021630,684642,645717,746
25−2099−2719−2136−2343−3276
Table 9. Matching polynomials of isomers of C52 fullerene.
Table 9. Matching polynomials of isomers of C52 fullerene.
KC52(D2)-433C52(C2)-434C52(C1)-436C52(T)-437
01111
1−78−78−78−78
22847284728472847
3−64,636−64,636−64,636−64,636
41,023,3991,023,3991,023,3991,023,399
5−12,009,570−12,009,570−12,009,570−12,009,570
6108,366,033108,366,033108,366,033108,366,033
7−769,906,260−769,906,260−769,906,260−769,906,260
84,374,890,4204,374,890,4194,374,890,4184,374,890,418
9−20,087,482,056−20,087,482,004−20,087,481,952−20,087,481,952
1074,993,403,69674,993,402,47574,993,401,25474,993,401,254
11−228,347,209,688−228,347,192,556−228,347,175,414−228,347,175,408
12567,290,814,788567,290,654,621567,290,494,046567,290,493,824
13−1,147,459,180,912−1,147,458,127,888−1,147,457,067,510−1,147,457,063,880
141,881,094,347,0281,881,089,339,9881,881,084,255,6921,881,084,221,076
15−2,481,558,198,016−2,481,540,762,864−2,481,522,802,116−2,481,522,588,124
162,608,075,274,1682,608,030,772,4582,607,983,836,7362,607,982,934,853
17−2,154,508,107,056−2,154,425,661,680−2,154,335,363,811−2,154,332,709,330
181,374,351,797,3841,374,243,411,0721,374,117,256,6641,374,111,754,991
19−661,324,154,776−661,227,247,200−661,102,300,879−661,094,291,988
20232,679,709,256232,625,587,423232,541,130,115232,533,025,512
21−57,362,136,248−57,347,243,516−57,310,508,266−57,304,888,024
229,326,080,6569,326,685,0349,317,305,1379,314,670,876
23−912,517,384−913,967,662−912,732,486−911,913,672
2446,175,68846,459,73146,372,43446,216,468
25−900,864−913,906−906,964−893,568
262904294128142700
Table 10. Matching polynomials of C58(C3v)-1 and heptagonal C58(Cs)-hept with 1 heptagon and 13 pentagons.
Table 10. Matching polynomials of C58(C3v)-1 and heptagonal C58(Cs)-hept with 1 heptagon and 13 pentagons.
kC58(C3v)-1C58(Cs)-hept
011
1−87−87
235673567
3−91,669−91,669
41,656,8281,656,828
5−22,400,052−22,400,051
6235,240,023235,239,954
7−1,967,080,257−1,967,078,043
813,320,537,62413,320,493,739
9−73,905,469,014−73,904,866,941
10338,630,578,458338,624,507,555
11−1,287,860,109,036−1,287,813,476,284
124,076,572,360,4084,076,293,247,395
13−10,748,037,091,998−10,746,716,517,904
1423,577,708,216,70823,572,726,102,672
15−42,911,701,168,180−42,896,647,640,354
1664,495,465,418,16364,459,007,350,761
17−79,522,974,846,489−79,452,370,129,844
1879,733,207,226,75479,624,483,458,808
19−64,270,421,735,034−64,138,493,127,215
2041,044,816,761,15040,920,303,468,711
21−20,382,041,191,170−20,292,270,811,849
227,681,900,580,2057,633,661,167,869
23−2,128,631,549,481−2,109,955,392,032
24415,408,342,364410,442,814,122
25−53,747,261,070−52,902,565,843
264,212,821,4334,130,671,443
27−172,462,371−168,752,079
282,762,9702,717,607
29−7308−7525
Table 11. Matching polynomials of buckminsterfullerene (C60(Ih)) and its isomer C60(D3)-1811.
Table 11. Matching polynomials of buckminsterfullerene (C60(Ih)) and its isomer C60(D3)-1811.
kC60(Ih)C60(D3)-1811
011
1−90−90
238253825
3−102,120−102,120
41,922,0401,922,040
5−27,130,596−27,130,596
6298,317,860298,317,860
7−2,619,980,460−2,619,980,460
818,697,786,68018,697,786,686
9−109,742,831,260−109,742,831,644
10534,162,544,380534,162,555,702
11−2,168,137,517,940−2,168,137,722,048
127,362,904,561,7307,362,907,079,705
13−20,949,286,202,160−20,949,308,744,700
1449,924,889,888,85049,925,041,449,174
15−99,463,457,244,844−99,464,238,463,876
16165,074,851,632,300165,077,976,023,361
17−227,043,126,274,260−227,052,877,002,918
18256,967,614,454,320256,991,374,424,828
19−237,135,867,688,980−237,180,889,766,676
20176,345,540,119,296176,411,295,787,590
21−104,113,567,937,140−104,186,538,219,098
2247,883,826,976,58047,944,056,256,236
23−16,742,486,291,340−16,778,325,531,438
244,310,718,227,6854,325,385,183,252
25−783,047,312,406−786,868,226,034
2694,541,532,16595,084,107,821
27−6,946,574,300−6,969,881,806
28269,272,620266,597,229
29−4,202,760−3,954,300
3012,5009622
Table 12. Matching polynomials of two isomers of C72: C72(C2v)-11188 fullerene with non-isolated pentagon structure and C72(D6d)-11190 a.
Table 12. Matching polynomials of two isomers of C72: C72(C2v)-11188 fullerene with non-isolated pentagon structure and C72(D6d)-11190 a.
kC72(C2v)-11188 (Non-ISP)C72(D6d)-11190
011
1−108−108
255625562
3−181,836−181,836
44,238,3794,238,379
5−74,997,996−74,997,996
61,047,459,3261,047,459,326
7−11,852,752,392−11,852,752,392
8110,690,579,974110,690,579,973
9−864,652,893,966−864,652,893,884
105,705,866,144,1225,705,866,140,966
11−32,043,716,552,498−32,043,716,476,716
12153,971,991,502,747153,971,990,229,848
13−635,430,828,140,544−635,430,812,245,836
142,257,883,027,813,5752,257,882,874,735,690
15−6,917,120,612,820,084−6,917,119,448,538,120
1618,275,900,215,535,84818,275,893,112,174,600
17−41,618,256,862,032,538−41,618,221,745,909,600
1881,556,423,951,149,66981,556,282,447,653,600
19−137,186,547,343,055,238−137,186,081,385,089,000
20197,391,632,599,522,833197,390,379,486,861,000
21−241,844,737,361,104,930−241,841,995,535,982,000
22250,872,705,868,808,807250,867,862,188,707,000
23−218,784,208,970,190,972−218,777,389,029,217,000
24159,029,898,793,311,758159,022,409,077,892,000
25−95,339,702,590,544,974−95,333,534,217,776,900
2646,538,823,097,489,22846,535,333,778,130,000
27−18,206,425,404,161,442−18,205,449,520,239,600
285,596,680,643,711,9505,596,999,388,641,250
29−1,318,481,294,247,250−1,318,986,435,090,070
30230,452,534,808,904230,726,853,211,188
31−28,618,179,154,208−28,704,598,608,024
322,376,922,675,7832,393,469,043,524
33−120,715,631,942−122,558,197,024
343,237,686,9913,345,162,432
35−34,480,394−37,159,200
3663,48777,400
a Results shown for C72(D6d) in the third column are from [50].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Balasubramanian, K. Matching Polynomial-Based Similarity Matrices and Descriptors for Isomers of Fullerenes. Inorganics 2023, 11, 335. https://doi.org/10.3390/inorganics11080335

AMA Style

Balasubramanian K. Matching Polynomial-Based Similarity Matrices and Descriptors for Isomers of Fullerenes. Inorganics. 2023; 11(8):335. https://doi.org/10.3390/inorganics11080335

Chicago/Turabian Style

Balasubramanian, Krishnan. 2023. "Matching Polynomial-Based Similarity Matrices and Descriptors for Isomers of Fullerenes" Inorganics 11, no. 8: 335. https://doi.org/10.3390/inorganics11080335

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