Next Article in Journal
Optimal Test Plan of Step-Stress Model of Alpha Power Weibull Lifetimes under Progressively Type-II Censored Samples
Next Article in Special Issue
Near-Miss Bi-Homogenous Symmetric Polyhedral Cages
Previous Article in Journal
Analytical Scaling Solutions for the Evolution of Cosmic Domain Walls in a Parameter-Free Velocity-Dependent One-Scale Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Digitalizing Structure–Symmetry Relations at the Formation of Endofullerenes in Terms of Information Entropy Formalism

by
Denis Sh. Sabirov
*,
Alina A. Tukhbatullina
and
Igor S. Shepelevich
Laboratory of Mathematical Chemistry, Institute of Petrochemistry and Catalysis UFRC RAS, 450075 Ufa, Russia
*
Author to whom correspondence should be addressed.
Symmetry 2022, 14(9), 1800; https://doi.org/10.3390/sym14091800
Submission received: 30 June 2022 / Revised: 8 August 2022 / Accepted: 25 August 2022 / Published: 30 August 2022
(This article belongs to the Special Issue Symmetry and Asymmetry in Nature-Inspired, Bio-Based Materials)

Abstract

:
Information entropy indices are widely used for numerical descriptions of chemical structures, though their applications to the processes are scarce. We have applied our original information entropy approach to filling fullerenes with a guest atom. The approach takes into account both the topology and geometry of the fullerene structures. We have studied all possible types of such fillings and found that information entropy (ΔhR) and symmetry changes correlate. ΔhR is negative, positive or zero if symmetry is increased, reduced or does not change, respectively. The ΔhR value and structural reorganization entropy, a contribution to ΔhR, are efficient parameters for the digital classification of the fullerenes involved into the filling process. Based on the calculated values, we have shown that, as the symmetry of the fullerene cage becomes higher, the structural changes due to the filling it with a guest atom become larger. The corresponding analytical expressions and numerical data are discussed.

1. Introduction

Information entropy (or Shannon entropy) and related quantities are widely used in structural and mathematical chemistry for numerically assessing the complexity of chemical objects [1,2,3,4,5,6,7,8,9,10,11,12]. One of the most common approaches deduces information entropy values from molecular graphs by counting atom types and their populations [13,14]. It treats a molecule as a system of subsets (atom types), whose number corresponds to the number of the signals in NMR spectra [14,15]. As previously found [13,15,16,17], such a partition of the fullerenes correlates with their symmetry point groups within a certain series of related or isomeric compounds. As the symmetry point group of the molecule becomes higher, its information entropy becomes lower, as was exemplified with the fullerene isomeric series of C60 [16] and C84 [15]. Moreover, information entropies of oligomers (C60)n congruently oscillate with the rotational symmetry numbers depending on the odd/even number of a homolog in the series [18]. Correlations between the symmetry and information entropy could be also found in the case of crystalline compounds [12,19].
In general, the information entropy characterizes the complexity of the molecules more accurately than symmetry [8,20]. Importantly, these estimates may relate to physicochemical processes [7,21,22,23,24] and, therefore, may be applicable to searching for correlations between the molecular structure, macroscopic properties, activities, performances, etc. [4,5,8,11,12,13,15,16,17,25,26,27,28,29].
The inner cavity, as an empty space which can be filled with guest atoms, is one of the attractive features of fullerenes [30,31]. The atoms can be trapped by the fullerene cages during fullerene synthesis [32], or high-pressure/high-temperature techniques can be used to introduce guest species inside the yet synthesized fullerenes [33]. In the last case, the formal chemical reaction occurs (X is a guest atom and N denotes the number of carbon atoms in the fullerene cage that is filled):
X + CN → X@CN
Such filling leads to changes in molecular/physicochemical properties and reactivity (e.g., [30,34,35,36,37,38,39,40,41]) and is accompanied with the slight extending of the fullerene cage [35,39] (with rare exceptions when the fullerene cage becomes more compact [42,43]). These changes have been scrutinized both theoretically and experimentally, and such studies usually include background discussions on the symmetry of the formed endofullerenes. However, their symmetry has never been addressed in a separate study. Encouraged by our recent advances in numerical descriptions of exohedral fullerene compounds with information entropy [13,17,18], we have decided to fill this gap.
In the present work, we consider typical cases of the formation of endofullerenes with single atoms inside in the aspect of the symmetry changes. For this purpose, we apply our original information-entropy-based formalism to the analysis of chemical processes [44]. Note that we do not look for correlations between the structural descriptors and the observed physicochemical properties. We instead focus on the interpretation of information entropy values in the context of digitalizing structural chemistry.

2. Preliminary Remarks

2.1. Mathematical Description of Fullerene Molecules

Fullerenes are very attractive models for structural chemistry studies, as their molecules are constructed in line with strict mathematical regularities. The topology of fullerenes has been discussed in a book [45] and in a comprehensive review [46], so here we make some important remarks on the relations between the topology, geometry and symmetry of the molecules.
Fullerenes are molecular objects, and their topology can be represented as molecular graphs [45,46]. The last ones in fullerene science are called Schlegel diagrams, which are introduced as the plain projections of polyhedra representing fullerene molecules (see examples in Section 4.5). Analyzing the corresponding adjacency matrices allows for the sortation of vertices over inequivalent types. For this purpose, the paths between the vertices are analyzed, and vertices are distributed over the topological orbits [45,46]. The exhaustive description of this topological approach can be also found in our studies [15,18].
There are algorithms connecting the topological parameters of Schelegel diagrams [45,46] with the expected symmetries of the fullerene molecules. Deduced from the topology, the symmetry of a certain fullerene corresponds to the highest one possible for this topological structure. In most cases, this ideal symmetry coincides with the actual one found on the level where geometry (spatial arrangement of atoms) is considered. However, on this level, reductions in symmetry can take place due to the Jahn–Teller effect (we scrutinize this case in Section 4.4). This reduction makes topologically equivalent atoms spatially inequivalent. To account for this geometrical inequivalence, one must perform a separate study of the symmetry based on the Cartesian coordinates of the atoms that make up the molecule. The coordinates for this purpose are usually obtained from high-level quantum chemical computations (e.g., [16,47]).
In brief, the symmetry of a fullerene molecule is deducible from its topology. However, it is better analyzed with the geometry of the molecule in order not to miss possible Jahn–Teller symmetry reduction. Herein, the topologies of the fullerene molecule in the high-symmetry and reduced-symmetry states are identical.

2.2. Information Entropy Formalism Applied to (Endo)Fullerenes

To apply information entropy formalism, the molecule must be presented as the set of the elements, atoms or bonds. We prefer to deal with representing molecules as sets of atoms due to two reasons [7]. First, the distribution of the molecule’s atoms over atom types corresponds to its NMR spectrum, i.e., there is a direct correspondence to experimental techniques. The second reason deals with a particular case of the chemical structure of endofullerenes. The atom placed inside has no covalent bonds with the fullerene cage. Hence, the molecular graph of endofullerene contains one isolated vertex, and considering the edges of this structure seems impractical.
Therefore, we treat (endo)fullerenes as sets of atoms which are distributed over atom types based on their geometries obtained in previous studies.

2.3. Topological Stability of the Fullerene Cages in Endofullerenes

Another note on endofullerenes deals with chemical aspects of their formation. When formed, the topology of the fullerene cage does not change. This regularity is rarely violated, and the corresponding cases are far from the objects under the present study, as they deal with further chemical transformations of the already formed endohedral complexes (e.g., the compression of endofullerenes with reactive fillings [38,48,49], the chemical reactions of metal carbide endofullerenes [50] and cage-opened fullerene derivatives [51]).

3. Computational Details

3.1. Information Entropy Indices for the Analysis of Chemical Processes

The fullerene molecules are represented as the sets of N1 atoms of the 1st type, N2 atoms of the 2nd type, … and Nn atoms of the n-th type, where n is the number of atom types, and Σ Nj = N is the total number of carbon atoms in the cage. The distribution of the atoms over atom types depends on both the topologies and geometries of the fullerene molecules that allow reflecting their symmetries [16]. The information entropy of the molecule (h) equals the sum of the logarithms associated with each atom type [16,44]:
h = j = 1 n N j j = 1 n N j log 2 N j j = 1 n N j
Herein, we use two as the base of the logarithms, which is conventional [7] and allows for expressing all information entropy values in bits. Single atoms have zero information entropy according to Equation (2) (n = 1 and Nj = 1).
Following Ugi and Gillespi [52], the chemical reaction is considered the ‘isomerization’ of one molecular ensemble (ME) to another. Hence, the change in information entropy upon the chemical process equals the difference between the hME values of two ensembles, products and reactants [14,44,53]:
Δ h R = h ME p r o d h ME r e a c t
The information entropy of each ME is calculated as:
h ME = H Ω + i = 1 m ω i h i
where ωi is the fraction of the i-th molecule in ME:
ω i = N i i = 1 m N i
and HΩ is the cooperative entropy:
H Ω = i = 1 m ω i log 2 ω i
The HΩ is an emergent parameter that arises due to the mixing of molecules when they form the ensemble. It is independent from the molecular structure of the ME members (hi) and is defined only by their sizes (ωi). HΩ = 0 in the case of monomolecular ME (when ω = 1).
A combination of Equations (3) and (6) provides the following expression for information entropy changes in a chemical process:
Δ h R = H r e d i s t r + H r e o r g
where the first term is the redistribution information entropy, reflecting the difference in the size of the molecules in MEs of products and reactants:
H r e d i s t r = H Ω p r o d H Ω r e a c t
The second term, called reorganization information entropy, depends on both hi and ωi, i.e., on structure and size, respectively:
H r e o r g = i p r o d ω i h i j r e a c t ω j h j
The last term can be further divided over two contributions:
H r e o r g s t r = i p r o d h i j r e a c t h j
H r e o r g s t r + s i z e = j r e a c t ( 1 ω j ) h j i p r o d ( 1 ω i ) h i
Using this formalism, we can separately assess the changes in information entropy that correspond to molecular size and/or molecular structure. In Equations (7)–(11), the upper indices ‘str’ and ‘size’ indicate the references to molecular structure and size, respectively. Redistribution entropy Hredistr (Equation (8)) depends only on the size of the reaction participants. The explanatory remarks to this approach can be found in our key work [44] (the corresponding chemical and mathematical justifications are presented in earlier works [14,53]).
In brief, we list the information entropies used in the work and their designations. The h and hME values characterize the complexity of the molecules and molecular ensembles, respectively. The hME value contains cooperative entropy HΩ as a part. This is an emerging parameter that reflects the effect of uniting molecules in the ensemble. The information entropies with delta signs on designated with capital letters correspond to chemical processes (ΔhR and its components Hredistr and Hreorg).

3.2. Structures of (Endo)Fullerenes for Analysis and Symmetry Determination

As mentioned in Section 2.1, the molecular graphs are necessary to obtain the information entropies of the fullerene molecules. To define the symmetries of the (endo)fullerenes, we used our previous works on (endo)fullerenes, whereby their structures were obtained with reliable density functional theory methods [16,17,38,39]. The symmetry point groups of chemical objects in this and our previous studies were determined in program Chemcraft [54].

4. Results

In this work, we consider three main cases of the process presented with Equation (1). The cases differ in what happens with the symmetry of the original (empty) fullerene cage (Figure 1). It can be served, increased or decreased depending on the chemical features of the interacting guest atom X and fullerene host CN. Herein, the topology of the fullerene cage does not change after filling. Additionally, we focus below on the analytical expressions for ΔhR and its components by providing numerical results.

4.1. Introducing X into the Fullerene Cage: Ommon Analytical Expressions

Before starting a study, we briefly demonstrate the calculations of the information entropies of fullerene molecules using C60 (Ih) and C70 (D5h) as examples (Figure 2). Their partitions are 1 × 60 and 3 × 60 + 2 × 20, respectively. The substitution of the populations of atom types to Equation (2) leads to the h values for these fullerenes, equal to 0 and 2.236 bits.
In the case of introducing X into the CN cage, we have the following input data for simplifying characteristic Equations (7)–(11): (a) the information entropy of the atom equals zero, hX = 0; and (b) the cooperative entropy of the molecular ensemble of the products is zero because the single product is formed under reaction (1), H Ω p r o d = 0. The symmetry of the cage does not change (Figure 3) because atom X takes the position in the mass center of the cage, which coincides with the intersection of symmetry elements, if any.
Considering the above, we obtain:
H r e d i s t r = N N + 1 log 2 N log 2 ( N + 1 )
H r e o r g = h X @ C N N N + 1 h C N
H r e o r g s t r + s i z e = 1 N + 1 h C N
H r e o r g s t r = h X @ C N h C N
If the symmetry does not change, the partitions over atom types for X@CN can be simply represented as 1 × 1 + {partition of CN}. This allows rewriting Equation (13) as:
H r e o r g = log 2 ( N + 1 ) N N + 1 log 2 N
Thus, H r e o r g = H r e d i s t r and Δ h R = 0 (according to Equation (7)).
We present the work on the approach on endofullerenes in Table 1. Note that Hredistr depends only on the size of the filled fullerene, so this value is insensible towards the structural differences in the cases below. As the symmetry does not change upon the encapsulation, we may compare the structural effects of trapped atom X for different fullerenes. From the chemical point of view, as the fullerene cage becomes larger, the impact of trapped atom X on the system becomes smaller. In terms of information entropy, it means that Hreorg decreases with increasing N, which is observed in the calculated values (Table 1).

4.2. Introducing X into a Zero-h Fullerene Cage While Serving Initial Symmetry

Fullerene C60 (Ih) is the most prominent fullerene among other member of the fullerene family. It is the only fullerene molecule having zero information entropy calculated via Equation (2) due to the equivalence of all atoms [16]. Therefore, the case of h C N = 0 deserves separate attention. The C60 cage is very stable, and, if there are no special interactions between the introduced guest atom X and the cage (e.g., charge transfer), the guest atom holds the position in the mass centrum of the cage. The cage is negligibly extended and serves its initial icosahedral symmetry. The mentioned situation is typical for the formation of noble gas endofullerenes [35,36,37,39].
All carbon atoms in X@C60 (Ih) remain equivalent, and its partition is 1 × 1 + 1 × 60, or 1 × 1 + 1 × N in the general case for similar X@CN. For the information entropy of X@CN, we can write:
h X @ C N = log 2 ( N + 1 ) N N + 1 log 2 N
and it is obvious that h X @ C N = H r e d i s t r , cf.: Equation (12). We obtain the same equations for Hredistr and Hreorg as in the case above; therefore, ΔhR = 0. However, the components of Hreorg differ:
H r e o r g s t r = log 2 ( N + 1 ) N N + 1 log 2 N
H r e o r g s t r + s i z e = 0 ,
i.e., there are no structural changes simultaneously depending on the molecular size and molecular structure.
In addition to zero-h fullerene C60, we consider other carbon molecules having h = 0, which can form complexes in a similar way (with placing guest atom in the central position). These are the hypothetical fullerene-like molecules C24 (Oh) and C48 (Oh), containing polygons untypical for fullerenes (tetra- and octagons) [47,55], and synthesized cyclo [18] carbon C18 has a D9h symmetry [56] (Figure 4). The numerical results obtained for C60 (Ih) and these structures are shown in Table 2.

4.3. Introducing X into Maximum-h Fullerene Cage

Symmetric fullerenes are the main products of arc-discharge fullerene synthesis [57]. Moreover, there are examples of fullerenes with no symmetry, whose molecules are attributed to the C1 symmetry point group (e.g., see [58,59,60,61]). Additionally, the majority of the hypothetical fullerene structures belong to this type [16,45]. For these fullerenes, information entropy achieves its maximal value for a given N (which corresponds to the partition of the molecule N × 1) [16]:
h C N   ( C 1 ) = log 2 N
The application of Equation (20) to expressions (7)–(11) after some simplifications affords the same equations for Hredistr and Hreorg, as in the two cases above, and, consequently, ΔhR = 0. The components of Hreorg are the following:
H r e o r g s t r = log 2 ( N + 1 ) log 2 N
H r e o r g s t r + s i z e = 1 N + 1 log 2 N
As can be seen, they differ from the case of the zero-h fullerene, as the reorganization entropy is divided over two contributions ( H r e o r g s t r + s i z e = 0   in the case of h C N = 0 ). As for the comparison with the case of Section 3.1, here, H r e o r g s t r + s i z e   achieves its maximal value due to condition (20). Some typical numerical examples of this case are collected in Table 3.

4.4. Introducing X into the Fullerene Cage with Initially Reduced Symmetry

Some of the fullerenes, especially nonconventional structures, reveal Jahn–Teller symmetry reduction [47]. We do not discuss here its molecular–orbital reasons, and we instead focus on the consequences that are important for the present study. The topology of the fullerene cage undergoing the effect remains the same, but spatially, the structure is distorted; therefore, the actual symmetry becomes lower compared to the ideal symmetry deduced from the topology [16,47]. The C20 fullerene is a typical example of such a fullerene. Its ideal symmetry is Ih, but it is reduced to the Jahn–Teller effect to Ci [62]. (Note that the actual symmetry of C20 is questionable because different quantum–chemical approximations predict different symmetries, C2, C2h, Ci, D3d and D2h [47]. All of them are lower than Ih. We use the Ci symmetry structure as obtained in our previous work [62].) Other representatives of fullerenes of this type are listed in Table 4.
Filling the C20 (Ci) cage with noble gas atoms (He or Ne) leads to expanding the cage but does not change its symmetry, i.e., He@C20 and Ne@C20 are also Ci symmetry structures [35,36,37,38,39]. However, if metal species such as Nd, U, Pm+, Np+, Sm2+, Pu2+, Eu3+, Am3+, Gd4+ or Cm4+ play the role of guest atoms, the C20 cages obtain icosahedral symmetry. A metal guest takes the place at the center of the cage, so the formed X@C20 structures are Ih-symmetric [63]. This is reflected in Figure 5.
A similar ideal symmetry restoration within the endohedral complexes may be typical for other fullerenes, which demonstrate the Jahn–Teller symmetry reduction in the empty state (e.g., C24-1, C26-1, C40-1, C76-2 and C80-7 [16], the designation after the hyphen is the number of the isomer according to Atlas [45]). The corresponding information entropy changes, and their contributions are described with Equations (12)–(16) (Table 5). Herein, the relation is typical for this type of formed endofullerene (Table 4):
h C N > h X @ C N i d e a l h X @ C N > h C N i d e a l
where the upper index ‘ideal’ indicates that the fullerene cage is in the highest symmetry state possible for the given topology. Due to the above inequality, H r e d i s t r H r e o r g ; therefore, the resulting ΔhR values are negative in contrast to the cases considered above with unchanged symmetry. The information entropy changes in such processes are calculated as:
Δ h R = h X @ C N N N + 1 ( h C N log 2 N ) log 2 ( N + 1 )   <   0

4.5. Introducing X into the Fullerene Cage with X Coordination Relative to the Cage

The last case of process (1) deals with the coordination of the guest atom inside the cage. This is possible when the guest atom and host cage strongly interact due to the charge transfer. Endofullerene Li@C60 is a typical synthesized example of such endohedral complexes (Figure 6) [64,65,66].
Following Sokolov [67], we also consider hypothetical coordination modes of the trapped atom inside the C60 cage, viz., the orientation of the X atom toward the pentagons, the atom of the cage and the 5.6 and 6.6 carbon–carbon bonds of the fullerene cage. To calculate the h values of the formed X@C60 complexes, we deduce the corresponding partitions from the Schlegel diagrams that account for the inequivalence of carbon atoms that occurs due to complexation (Figure 7).
As found, the symmetry of the fullerene system is reduced from the initial Ih to Cnv and CS depending on the coordination of X (Table 6). The symmetry goes down, and this must be reflected with increasing h values. Indeed, we observe positive ΔhR values for this case.
The above considerations are applicable to other fullerenes in the same way. Other fullerenes have lower symmetries; therefore, endo-atom X may be coordinated towards a larger number of the substructures of the fullerene cage (but their types are the same as in the described C60 case of pentagons, hexagons, atoms and 5.6- and 6.6-bonds). Note that our approach does not distinguish topological isomers X@CN and X…CN, i.e., the isomers differ in the X location, inside or outside the cage. Therefore, the obtained relations are also characteristic for exohedral fullerene complexes, another class of fullerene compounds [68].

4.6. Information Entropy Indices of Processes X + CN → X@CN within Series of Isomeric Fullerenes

A number of isomeric fullerene structures may correspond to each molecular size (N), and the task of comparing fullerene isomers often arises in fullerene science. Therefore, we explore this case for two fullerene series taken from our previous works [15,16]. The original symmetry of the fullerene cage is conserved under encapsulation. As follows from Section 4.1, ΔhR = 0 and H r e o r g = H r e d i s t r = f ( N ) are the same for the endofullerene formation within the isomeric series. Hence, we focus on the H r e o r g s t r values, which are distinctive for the isomers.
Before mathematical treatment, we consider some chemical aspects of process (1) regarding the behavior of information entropy. These processes share one common feature: the symmetry of the cage is the same in empty and encapsulated states. The difference between the processes in each isomeric series is due to the symmetry of certain fullerene molecules. At first glance, it seems that the H r e o r g s t r values (calculated via Equation (15)) should be the same for different fullerenes isomers. However, in fact, the H r e o r g s t r values for the selected C60 isomers differ (Table 7). Herein, as the symmetry of the fullerene isomer becomes higher, H r e o r g s t r becomes larger. Note that, as a preliminary example, to find a pattern, we use a series of only six isomers of C60 that have different symmetries (their total number equals 1812 [69]). As follows from Table 7, the rotational symmetry number is not the best mode to quantify the fullerene symmetry. Indeed, σ = 2 for C60 (C2v) and C60 (C2), but these isomers differently crumble over atom types. For further work, we need to use some quantity that numerically catches the differences in the partitions of the fullerene isomers.
By scrutinizing isomeric series with N = const, we introduce logarithmic partition number (χ) for convenience:
χ = j = 1 n N j log 2 N j
It characterizes the partition and correlates with σ (Figure 8), and the information entropies of (endo)fullerenes are expressed with χ (combining Equations (2) and (25)):
h C N = log 2 N χ N
h X @ C N = log 2 ( N + 1 ) χ N + 1
Importantly, minimal χ equals zero and corresponds to the lowest symmetry (C1), whereas χ max = N log 2 N corresponds to a high symmetry (Ih in the C60 series). In general, χ → 0 if the partition of the molecule includes singly populated atom types, i.e., when nN and Nj → 1. In the opposite case, when the atom types are highly populated, we have n → 1, NjN and χ → χmax.
Substituting expressions (26) and (27) into Equation (15) provides linear dependence for structural reorganization entropy on the logarithmic partition number:
H r e o r g s t r = log 2 N + 1 N + χ N ( N + 1 )
According to the last expression, H r e o r g s t r increases with χ (Figure 8). As χ is connected with a linear equation and with the information entropy of fullerenes and correlates with their symmetries, we can interpret Equation (28) in terms of both symmetry and information entropy. As the information entropy of the fullerene become smaller, its symmetry becomes higher, and the structural changes expressed with H r e o r g s t r become larger. In other words, the information entropy associated with structural changes at the endofullerene formation depends on the symmetry of the original structure.
To demonstrate that the discussed statements are valid in other cases, we have also applied them to the series of isolated pentagon isomers of C84. As seen in Figure 8, the correlations between χ and σ becomes worse in this wider series, but these values remain symbatic. In general, χ provides more diversified numerical assessing chemical structures as compared with σ. The use of the logarithmic partition number allows for generalizing the above cases (substituting χmax and χmin in Equation (28) leads to Equations (18) and (21) for zero-h and maximum-h fullerenes, respectively).

5. Discussion

5.1. General Remarks

We have applied our original information entropy approach to describe the relations between chemical structure and symmetry upon filling fullerene cages with guest atom X. The obtained numerical data are well interpretable in terms of general chemical notions. Herein, we have operated this total information entropy change both in the process and its components.
The total change ΔhR allows for discriminating the cases when symmetry is changed upon the title process; ΔhR is negative, positive or zero if symmetry is increased, reduced or does not change, respectively. Furthermore, the components of ΔhR correspond to the types of empty fullerenes and formed endofullerenes, and this correspondence is unambiguous. Thus, we state here that the type of the (endo)fullerene structure can be deduced from the information entropy estimates of process (1). It can be performed digitally using the algorithm based on comparing ΔhR and H r e o r g s t r + s i z e values (Figure 9). Information-entropy-based classifications have been elaborated to sort natural compounds [25] and interstellar [13] and isentropic molecules [70], and they have operated with numerical estimates of chemical structures. We propose a scheme (Figure 9) utilizing indices relating to the process, which involves chemical structures. Such tasks are currently not widespread in mathematical chemistry. Deciphering the chemical structure based on the information about its chemical processes is standard for classical (experimental) chemistry. Therefore, we can assume that the digital task, similar to that one, will arise in chemical sciences in the nearest future.

5.2. Dependence of the Strucutral Reorganization Entropy of Encapsulation on the Symmetry and Size of Fullerenes

The main aim of this proposed concept is developing an information entropy approach applicable to all molecular compounds. We focus on fullerenes because they are rigid and symmetric molecules that are very convenient for testing structural descriptors and linking numerical estimates with chemical notions. Information entropy and symmetry are concepts closely relating to molecular complexity. High information entropy corresponds to low symmetry and high complexity of the molecules [2,16].
The most interesting thing we have found in this work is that the information entropy of the intact fullerene h C N influences the structural reorganization entropy H r e o r g s t r + s i z e of its filling. This finding is not obvious in the context of chemical intuition, and, therefore, we discuss it below.
If the symmetry remains unchanged, we have similar substrates involved in the identical processes, resulting in similar products (Equation (1)). The similarity of the processes is reflected by the equality of their ΔhR values (ΔhR = 0), and this is well understandable. At first glance, numerical estimates associated with structural changes ( H r e o r g s t r + s i z e ) for these processes should also be equal, but, in fact, they are not the same. We rationalize this inequality using the logarithmic partition number χ as an auxiliary parameter. The χ number linearly linked with the information entropy correlates with the symmetry of the fullerene. As χ becomes larger, the number of equivalent atoms in the molecule becomes larger, and as the h value becomes lower, its symmetry and the H r e o r g s t r + s i z e of the encapsulation become higher. The latter means that, if the symmetry of the fullerene is high, the guest atom introduces more structural changes when encapsulated and vice versa. In symmetrical fullerene structures, the number of atom types is low, and they are highly populated. Filling them with X crashes this uniformity. In contrast, there are many sparsely populated atom types in low-symmetry structures, i.e., they have some degree of disorder before encapsulation, and the next additional atom in the system makes a small contribution to the structural complexity increase.
Thus, the information entropy of the empty fullerene defines the structural reorganization entropy of the encapsulation process. Taking into account the fact that information entropy reflects the complexity of the molecular system, we can propose a stronger version of this statement: the complexity of the initial structure defines the complexity change during the chemical reaction. The last sentence should be accepted as the assumption that requires further justification with other chemical compounds, chemical reactions and complexity measures.
Structural reorganization entropy H r e o r g s t r + s i z e also depends on the size of the fullerene cage expressed with N. By analyzing fullerenes with different N (Table 1, Table 2, Table 3 and Table 5), one can find that the H r e o r g s t r + s i z e decreases N. This means that the structural changes relating to introducing the endo-atom to the molecular system becomes smaller in the background of the whole. This fits into the previously found regularities for chemical reactions in homological series of hydrocarbons and their derivatives [14,44].

5.3. Prospectives

The approach used in the present work is quite new, and we are going to extrapolate it to other chemical systems and processes involving fullerenes, including chemical processes, e.g., filling fullerenes with two or more atoms, multi-atomic clusters or small molecules. Such endofullerenes are currently being synthesized (e.g., H2O@C60/70, He2@C60/70, HeN@C60/70, Sc3N@CN, M2C2@CN) (see works [71,72,73] and the references therein). We also think that the approach could be useful for the analysis of exohedral additions to fullerenes [68,74,75], where information entropy is expected to catch both the complexity of the addition pattern [74] and the symmetry of functional groups attached to the fullerene cage [75].
In a recent review, we point out that the information entropy concept is a good value for interdisciplinary studies, as it has applications in various fields [7]. In the present work, we have used it only as a structural descriptor, which could be further incorporated into the studies on planning syntheses [76], self-assembly processes [21,22,23,24] and informational [77,78,79] and geometrical [80] thermodynamics. This may be insightful for the interface between structural chemistry, physics and information sciences.

6. Conclusions

We have studied all cases of introducing one guest atom into the fullerene cage, forming endofullerenes in terms of original information entropy formalism. As has been found, the calculated total information entropy change reflects the structure and symmetry changes of the process, whereas its components can be used for the digital identification of key structural features of fullerenes participating in the process of filling.
We point out the importance of structural reorganization entropy (a part of the total information entropy change), and it depends on the structure and size of fullerenes. As the information entropy of an empty fullerene becomes lower, the H r e o r g s t r + s i z e value becomes larger. The latter means that, in terms of symmetry, as the symmetry of the fullerene cage becomes higher, the structural changes due to filling it with a guest atom become larger.
In the future, we plan to apply the developed formalism to various reactions of organic compounds (including organic derivatives of fullerenes). The processes studied in the present work are simple in the following sense: only one atom is specifically introduced to a chemical system, with minimal symmetry changes and unchanged topology of the cage. This allows for rationalizing the structural meaning of the contributions to the information entropy change, which is important for more complex cases.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym14091800/s1, Table S1: Numerical data on the formation of endofullerenes X@C84 associated with Figure 8.

Author Contributions

Conceptualization, D.S.S.; methodology, D.S.S.; validation, A.A.T. and I.S.S.; formal analysis, D.S.S.; investigation, A.A.T.; writing—original draft preparation, D.S.S.; writing—review and editing, I.S.S.; visualization, A.A.T.; project administration, D.S.S.; funding acquisition, D.S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, project “Information entropy of chemical reactions: A novel methodology for digital organic chemistry”, grant number 22-13-20095.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Stankevich, M.I.; Stankevich, I.V.; Zefirov, N.S. Topological indices in organic chemistry. Russ. Chem. Rev. 1988, 57, 191–208. [Google Scholar] [CrossRef]
  2. Bonchev, D. Kolmogorov’s information, Shannon’s entropy, and topological complexity of molecules. Bulg. Chem. Commun. 1995, 28, 567–582. [Google Scholar]
  3. Barigye, S.J.; Marrero-Ponce, Y.; Pérez-Giménez, F.; Bonchev, D. Trends in information theory-based chemical structure codification. Mol. Divers. 2014, 18, 673–686. [Google Scholar] [CrossRef] [PubMed]
  4. Basak, S.C.; Harriss, D.K.; Magnuson, V.R. Comparative study of lipophilicity versus topological molecular descriptors in biological correlations. J. Pharm. Sci. 1984, 73, 429–437. [Google Scholar] [CrossRef]
  5. Basak, S.C.; Gute, B.D.; Grunwald, G.D. Use of topostructural, topochemical, and geometric parameters in the prediction of vapor pressure:  A Hierarchical QSAR approach. J. Chem. Inf. Comput. Sci. 1997, 37, 651–655. [Google Scholar] [CrossRef]
  6. Böttcher, T. An additive definition of molecular complexity. J. Chem. Inf. Model. 2016, 56, 462–470. [Google Scholar] [CrossRef]
  7. Sabirov, D.S.; Shepelevich, I.S. Information entropy in chemistry: An overview. Entropy 2021, 23, 1240. [Google Scholar] [CrossRef]
  8. Nagaraj, N.; Balasubramanian, K. Three perspectives on complexity: Entropy, compression, subsymmetry. Eur. Phys. J. Spec. Top. 2017, 226, 3251–3272. [Google Scholar] [CrossRef]
  9. Krivovichev, S.V. Structural complexity of minerals: Information storage and processing in the mineral world. Mineral. Mag. 2013, 77, 275–326. [Google Scholar] [CrossRef]
  10. Krivovichev, S.V. Structure description, interpretation and classification in mineralogical crystallography. Crystallogr. Rev. 2017, 23, 2–71. [Google Scholar] [CrossRef]
  11. Krivovichev, S.V.; Krivovichev, V.; Hazen, R.M.; Aksenov, S.M.; Avdontceva, M.S.; Banaru, A.M.; Gorelova, L.A.; Ismagilova, R.M.; Kornyakov, I.V.; Kuporev, I.V.; et al. Structural and chemical complexity of minerals: An update. Mineral. Mag. 2022, 86, 183–204. [Google Scholar] [CrossRef]
  12. Banaru, A.M.; Aksenov, S.M. Complexity of molecular nets: Topological approach and descriptive statistics. Symmetry 2022, 14, 220. [Google Scholar] [CrossRef]
  13. Sabirov, D.S. Information entropy of interstellar and circumstellar carbon-containing molecules: Molecular size against structural complexity. Comput. Theor. Chem. 2016, 1097, 83–91. [Google Scholar] [CrossRef]
  14. Sabirov, D.S. Information entropy changes in chemical reactions. Comput. Theor. Chem. 2018, 1123, 169–179. [Google Scholar] [CrossRef]
  15. 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]
  16. Sabirov, D.S.; Osawa, E. Information entropy of fullerenes. J. Chem. Inf. Model. 2015, 55, 1576–1584. [Google Scholar] [CrossRef]
  17. Sabirov, D.S.; Terentyev, A.O.; Sokolov, V.I. Activation energies and information entropies of helium penetration through fullerene walls. Insights into the formation of endofullerenes nX@C60/70 (n = 1 and 2) from the information entropy approach. RSC Adv. 2016, 6, 72230–72237. [Google Scholar] [CrossRef]
  18. Sabirov, D.S.; Ori, O.; Tukhbatullina, A.A.; Shepelevich, I.S. Covalently bonded fullerene nano-aggregates (C60)n: Digitalizing their energy-topology-symmetry. Symmetry 2021, 13, 1899. [Google Scholar] [CrossRef]
  19. Banaru, A.; Aksenov, S.; Krivovichev, S. Complexity parameters for molecular solids. Symmetry 2021, 13, 1399. [Google Scholar] [CrossRef]
  20. Sabirov, D.S.; Ori, O.; Tukhbatullina, A.A.; Shepelevich, I.S. Structural descriptors of benzenoid hydrocarbons: A mismatch between the estimates and parity effects in helicenes. C 2022, 8, 42. [Google Scholar] [CrossRef]
  21. Aleskovskii, V.B. Chemical and Information Synthesis. The Beginnings of the Theory. Methods; Publishing House of St. Petersburg University: St. Petersburg, Russia, 1997; 72p. [Google Scholar]
  22. Talanov, V.M.; Ivanov, V.V. Structure as the source of information on the chemical organization of substance. Russ. J. Gen. Chem. 2013, 83, 2225–2236. [Google Scholar] [CrossRef]
  23. Bal’makov, M.D. Information basis of nanochemistry. Russ. J. Gen. Chem. 2002, 72, 1023–1030. [Google Scholar] [CrossRef]
  24. Kadomtsev, B.B. Dynamics and information. Phys. Uspekhi 1994, 37, 425–499. [Google Scholar] [CrossRef]
  25. Castellano, G.; González-Santander, J.L.; Lara, A.; Torrens, F. Classification of flavonoid compounds by using entropy of information theory. Phytochemistry 2013, 93, 182–191. [Google Scholar] [CrossRef]
  26. Feng, B.; Zhuang, X. Carbon-enriched meso-entropy materials: From theory to cases. Acta Chim. Sin. 2020, 78, 833–847. [Google Scholar] [CrossRef]
  27. Krivovichev, S.V. Structural complexity and configurational entropy of crystals. Acta Crystallogr. B Struct. Sci. Cryst. Eng. Mater. 2016, 72, 274–276. [Google Scholar] [CrossRef]
  28. Jiménez-Ángeles, F.; Odriozola, G.; Lozada-Cassou, M. Entropy effects in self-assembling mechanisms: Also a view from the information theory. J. Mol. Liq. 2011, 164, 87–100. [Google Scholar] [CrossRef]
  29. Champion, Y.; Thurieau, N. The sample size effect in metallic glass deformation. Sci. Rep. 2020, 10, 10801. [Google Scholar] [CrossRef]
  30. Popov, A.A.; Yang, S.; Dunsch, L. Endohedral fullerenes. Chem. Rev. 2013, 113, 5989–6113. [Google Scholar] [CrossRef]
  31. Liu, S.; Sun, S. Recent progress in the studies of endohedral metallofullerenes. J. Organomet. Chem. 2000, 599, 74–86. [Google Scholar] [CrossRef]
  32. Akhanova, N.Y.; Shchur, D.V.; Pomytkin, A.P.; Zolotarenko, A.D.; Zolotarenko, A.D.; Gavrylyuk, N.A.; Ualkhanova, M.; Bo, W.; Ang, D. Methods for the synthesis of endohedral fullerenes. J. Nanosci. Nanotechnol. 2021, 21, 2446–2459. [Google Scholar] [CrossRef]
  33. Saunders, M.; Jiménez-Vázquez, H.A.; Cross, R.J.; Poreda, R.J. Stable compounds of helium and neon: He@C60 and Ne@C60. Science 1993, 259, 1428–1430. [Google Scholar] [CrossRef] [PubMed]
  34. Guha, S.; Nakamoto, K. Electronic structures and spectral properties of endohedral fullerenes. Coord. Chem. Rev. 2005, 249, 1111–1132. [Google Scholar] [CrossRef]
  35. Levin, A.A.; Breslavskaya, N.N. Energy of compressed endoatoms and the energy capacity of small endohedral rare-gas fullerenes. Russ. Chem. Bull. 2005, 54, 1999–2002. [Google Scholar] [CrossRef]
  36. Yan, H.; Yu, S.; Wang, X.; He, Y.; Huang, W.; Yang, M. Dipole polarizabilities of noble gas endohedral fullerenes. Chem. Phys. Lett. 2008, 456, 223–226. [Google Scholar] [CrossRef]
  37. Sabirov, D.S.; Bulgakov, R.G. Polarizability exaltation of endofullerenes X@Cn (n = 20, 24, 28, 36, 50, and 60; X is a noble gas atom). JETP Lett. 2010, 92, 662–665. [Google Scholar] [CrossRef]
  38. 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]
  39. 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]
  40. Osuna, S.; Swart, M.; Sola, M. The reactivity of endohedral fullerenes. What can be learnt from computational studies? Phys. Chem. Chem. Phys. 2011, 13, 3585–3603. [Google Scholar] [CrossRef]
  41. Ma, F.; Li, Z.-R.; Zhou, Z.-J.; Wu, D.; Li, Y.; Wang, Y.-F.; Li, Z.-S. Modulated nonlinear optical responses and charge transfer transition in endohedral fullerene dimers Na@C60C60@F with n-fold covalent bond (n = 1, 2, 5, and 6) and long range ion bond. J. Phys. Chem. C 2010, 114, 11242–11247. [Google Scholar] [CrossRef]
  42. Cioslowski, J.; Fleischmann, E.D. Endohedral complexes: Atoms and ions inside the C60 cage. J. Chem. Phys. 1991, 94, 3730. [Google Scholar] [CrossRef]
  43. Tukhbatullina, A.A.; Zakirova, A.D.; Sabirov, D.S. The volume of the cage of endohedral complexes of the C60 fullerene and halogenide-ions. Vestn. Bashkir. Univ. 2021, 26, 602–604. [Google Scholar] [CrossRef]
  44. Sabirov, D.S.; Tukhbatullina, A.A.; Shepelevich, I.S. Molecular size and molecular structure: Discriminating their changes upon chemical reactions in terms of information entropy. J. Mol. Graph. Model. 2022, 110, 108052. [Google Scholar] [CrossRef] [PubMed]
  45. Fowler, P.W.; Manolopoulos, D.E. An Atlas of Fullerenes; Clarendon Press: Oxford, UK, 1995; p. 392. [Google Scholar]
  46. Schwerdtfeger, P.; Wirz, L.N.; Avery, J. The topology of fullerenes. WIREs Comput. Mol. Sci. 2015, 5, 96–145. [Google Scholar] [CrossRef]
  47. 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]
  48. Sabirov, D.S. From endohedral complexes to endohedral fullerene covalent derivatives: A density functional theory prognosis of chemical transformation of water endofullerene H2O@C60 upon its compression. J. Phys. Chem. C 2013, 117, 1178–1182. [Google Scholar] [CrossRef]
  49. Pizzagalli, L. First principles molecular dynamics calculations of the mechanical properties of endofullerenes containing noble gas atoms or small molecules. Phys. Chem. Chem. Phys. 2022, 24, 9449–9458. [Google Scholar] [CrossRef]
  50. Zhang, J.; Bowles, F.L.; Bearden, D.W.; Keith Ray, W.; Fuhrer, T.; Ye, Y.; Dixon, C.; Harich, K.; Helm, R.F.; Olmstead, M.M.; et al. A missing link in the transformation from asymmetric to symmetric metallofullerene cages implies a top-down fullerene formation mechanism. Nat. Chem. 2013, 5, 880–885. [Google Scholar] [CrossRef]
  51. Hashikawa, Y.; Murata, Y. Water-mediated thermal rearrangement of a cage-opened C60 derivative. ChemPlusChem 2021, 86, 1559–1562. [Google Scholar] [CrossRef]
  52. Ugi, I.; Gillespie, P. Representation of chemical systems and interconversions bybe matrices and their transformation properties. Angew. Chem. 1971, 10, 914–915. [Google Scholar] [CrossRef]
  53. Sabirov, D.S. Information entropy of mixing molecules and its application to molecular ensembles and chemical reactions. Comput. Theor. Chem. 2020, 1187, 112933. [Google Scholar] [CrossRef]
  54. Chemcraft. Available online: http://www.chemcraftprog.com (accessed on 28 July 2021).
  55. Silant’ev, A.V. Energy spectrum and optical absorption spectrum of fullerene C24 within the Hubbard model. Phys. Solid State 2020, 62, 542–554. [Google Scholar] [CrossRef]
  56. Kaiser, K.; Scriven, L.M.; Schulz, F.; Gawel, P.; Gross, L.; Anderson, H.L. An sp-hybridized molecular carbon allotrope, cyclo[18]carbon. Science 2019, 365, 1299–1301. [Google Scholar] [CrossRef] [PubMed]
  57. Osawa, E. Formation mechanism of C60 under nonequilibrium and irreversible conditions—An annotation. Fuller. Nanotub. Carbon Nanostruct. 2012, 20, 299–309. [Google Scholar] [CrossRef]
  58. Yang, H.; Mercado, B.Q.; Jin, H.; Wang, Z.; Jiang, A.; Liu, Z.; Beavers, C.M.; Olmstead, M.M.; Balch, A.L. Fullerenes without symmetry: Crystallographic characterization of C1(30)-C90 and C1(32)-C90. Chem. Commun. 2011, 47, 2068–2070. [Google Scholar] [CrossRef]
  59. Tamm, N.B.; Guan, R.; Yang, S.; Troyanov, S.I. New isolated-pentagon-rule isomers of fullerene C96 captured as chloro derivatives. Eur. J. Inorg. Chem. 2020, 2020, 2092–2095. [Google Scholar] [CrossRef]
  60. Yang, S.; Wang, S.; Troyanov, S.I. The most stable isomers of giant fullerenes C102 and C104 captured as chlorides, C102(603)Cl18/20 and C104(234)Cl16/18/20/22. Chem.–Eur. J. 2014, 20, 6875–6878. [Google Scholar] [CrossRef]
  61. Tamm, N.B.; Yang, S.; Wei, T.; Troyanov, S.I. Five isolated pentagon rule isomers of higher fullerene C94 captured as chlorides and CF3 derivatives: C94(34)Cl14, C94(61)Cl20, C94(133)Cl22, C94(42)(CF3)16, and C94(43)(CF3)18. Inorg. Chem. 2015, 54, 2494–2496. [Google Scholar] [CrossRef]
  62. 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]
  63. Manna, D.; Ghanty, T.K. Theoretical prediction of icosahedral U@C20 and analogous systems with high HOMO–LUMO gap. J. Phys. Chem. C 2012, 11, 16716–16725. [Google Scholar] [CrossRef]
  64. Matsuo, Y.; Okada, H.; Ueno, H. History of Li@C60. In Endohedral Lithium-Containing Fullerenes; Springer: Singapore, 2017; pp. 15–23. [Google Scholar] [CrossRef]
  65. Chandler, H.J.; Stefanou, M.; Campbell, E.E.B.; Schaub, R. Li@C60 as a multi-state molecular switch. Nat. Commun. 2019, 10, 2283. [Google Scholar] [CrossRef] [PubMed]
  66. García-Hernández, D.A.; Manchado, A.; Cataldo, F. Hydrogenation of [Li@C60]PF6: A comparison with fulleranes derived from C60. Fuller. Nanotub. Carbon Nanostruct. 2022. [Google Scholar] [CrossRef]
  67. Sokolov, V.I. Fullerene C60 as a ligand with variable hapticity. Dokl. Akad. Nauk 1992, 326, 647. [Google Scholar]
  68. Soto, D.; Salcedo, R. Coordination modes and different hapticities for fullerene organometallic complexes. Molecules 2012, 17, 7151–7168. [Google Scholar] [CrossRef] [PubMed]
  69. 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] [PubMed]
  70. Sabirov, D.; Koledina, K. Classification of isentropic molecules in terms of Shannon entropy. EPJ Web Conf. 2020, 244, 01016. [Google Scholar] [CrossRef]
  71. Zhang, R.; Murata, M.; Aharen, T.; Wakamiya, A.; Shimoaka, T.; Hasegawa, T.; Murata, Y. Synthesis of a distinct water dimer inside fullerene C70. Nat. Chem. 2016, 8, 435–441. [Google Scholar] [CrossRef] [PubMed]
  72. Khong, A.; Jiménez-Vázquez, H.A.; Saunders, M.; Cross, R.J.; Laskin, J.; Peres, T.; Lifshitz, C.; Strongin, R.; Smith, A.B. An NMR Study of He2 Inside C70. J. Am. Chem. Soc. 1998, 120, 6380–6383. [Google Scholar] [CrossRef]
  73. Morinaka, Y.; Sato, S.; Wakamiya, A.; Nikawa, H.; Mizorogi, N.; Tanabe, F.; Murata, M.; Komatsu, K.; Furukawa, K.; Kato, T.; et al. X-ray observation of a helium atom and placing a nitrogen atom inside He@C60 and He@C70. Nat. Commun. 2015, 4, 1554. [Google Scholar] [CrossRef]
  74. Sabirov, D.S.; Tukhbatullina, A.A.; Bulgakov, R.G. Dependence of static polarizabilities of C60Xn fullerene cycloadducts on the number of added groups X = CH2 and NH (n = 1–30). Comput. Theor. Chem. 2021, 993, 113–117. [Google Scholar] [CrossRef]
  75. Sabirov, D.S.; Terentyev, A.O.; Bulgakov, R.G. Counting the isomers and estimation of anisotropy of polarizability of the selected C60 and C70 bisadducts promising for organic solar cells. J. Phys. Chem. A 2015, 119, 10697–10705. [Google Scholar] [CrossRef]
  76. Bertz, S.H. Complexity of synthetic reactions. The use of complexity indices to evaluate reactions, transforms and. New J. Chem. 2003, 27, 860–869. [Google Scholar] [CrossRef]
  77. Parrondo, J.; Horowitz, J.; Sagawa, T. Thermodynamics of information. Nat. Phys. 2015, 11, 131–139. [Google Scholar] [CrossRef]
  78. Sagawa, T.; Ueda, M. Minimal energy cost for thermodynamic information processing: Measurement and information erasure. Phys. Rev. Lett. 2011, 102, 250602. [Google Scholar] [CrossRef] [PubMed]
  79. Davis, S.; González, D. Hamiltonian formalism and path entropy maximization. J. Phys. A Math. Theor. 2015, 48, 425003. [Google Scholar] [CrossRef]
  80. Parker, M.C.; Jeynes, C. Fullerene stability by geometrical thermodynamics. ChemistrySelect 2020, 5, 5–14. [Google Scholar] [CrossRef]
Figure 1. Three cases of the endofullerene formation (Equation (1)) depending on the relations between the symmetries of empty and filled fullerene cages.
Figure 1. Three cases of the endofullerene formation (Equation (1)) depending on the relations between the symmetries of empty and filled fullerene cages.
Symmetry 14 01800 g001
Figure 2. The partitions of the molecules of the two most abundant fullerenes over atom types. The C60 (Ih) fullerene’s atoms are shown in one color, as all of them belong to one atom type (left). In the case of C70 (D5h), atoms of different atom types are shown in different colors and designated with Latin letters (center). The attributions of C70′s atoms are shown in a structural formula that corresponds to the top view on the molecule in the direction of its C5 symmetry axis (right).
Figure 2. The partitions of the molecules of the two most abundant fullerenes over atom types. The C60 (Ih) fullerene’s atoms are shown in one color, as all of them belong to one atom type (left). In the case of C70 (D5h), atoms of different atom types are shown in different colors and designated with Latin letters (center). The attributions of C70′s atoms are shown in a structural formula that corresponds to the top view on the molecule in the direction of its C5 symmetry axis (right).
Symmetry 14 01800 g002
Figure 3. Formation of endohedral complexes of C70 (D5h) with conserving symmetry. The symmetry axes are shown as yellow arrows.
Figure 3. Formation of endohedral complexes of C70 (D5h) with conserving symmetry. The symmetry axes are shown as yellow arrows.
Symmetry 14 01800 g003
Figure 4. Carbon structures having all atoms equivalent and, hence, zero information entropy: synthesized cyclo[18]carbon C18 (D9h) and hypothetical fullerene-like cages C24 (Oh) and C48 (Oh).
Figure 4. Carbon structures having all atoms equivalent and, hence, zero information entropy: synthesized cyclo[18]carbon C18 (D9h) and hypothetical fullerene-like cages C24 (Oh) and C48 (Oh).
Symmetry 14 01800 g004
Figure 5. Formation of X@C20 (Ih) from C20 (Ci) and X. The attributions of carbon atoms in the fullerene cages are shown with Latin letters.
Figure 5. Formation of X@C20 (Ih) from C20 (Ci) and X. The attributions of carbon atoms in the fullerene cages are shown with Latin letters.
Symmetry 14 01800 g005
Figure 6. Formation of Li@C60 (C3v) from C60 (Ih) and Li. Coordination bonds of the Li atom inside the C60 cage are shown.
Figure 6. Formation of Li@C60 (C3v) from C60 (Ih) and Li. Coordination bonds of the Li atom inside the C60 cage are shown.
Symmetry 14 01800 g006
Figure 7. Schlegel diagrams of X@C60 with coordination X via (a) hexagon, (b) pentagon, (c) 5.6 bond, (d) 6.6 bond and (e) atom of the fullerene cage. These elements of the fullerene structure are shown in blue. The carbon atoms of different atom types are lettered.
Figure 7. Schlegel diagrams of X@C60 with coordination X via (a) hexagon, (b) pentagon, (c) 5.6 bond, (d) 6.6 bond and (e) atom of the fullerene cage. These elements of the fullerene structure are shown in blue. The carbon atoms of different atom types are lettered.
Symmetry 14 01800 g007aSymmetry 14 01800 g007b
Figure 8. Information entropy and symmetry indices of processes X + CN → X@CN within series of isomeric fullerenes C60 (a,c,e) and C84 (b,d,f). Numerical data associated with the plot can be found in Table 7 and Supplementary Materials.
Figure 8. Information entropy and symmetry indices of processes X + CN → X@CN within series of isomeric fullerenes C60 (a,c,e) and C84 (b,d,f). Numerical data associated with the plot can be found in Table 7 and Supplementary Materials.
Symmetry 14 01800 g008
Figure 9. Algorithm for deducing the type of (endo)fullerene from the numerical information entropy estimates of process X + CN → X@CN, in which it is involved. The input data contain ΔhR and its components.
Figure 9. Algorithm for deducing the type of (endo)fullerene from the numerical information entropy estimates of process X + CN → X@CN, in which it is involved. The input data contain ΔhR and its components.
Symmetry 14 01800 g009
Table 1. Information entropy indices of filling typical fullerenes (Equation (1)) with conserved symmetry (herein, all values are in bits).
Table 1. Information entropy indices of filling typical fullerenes (Equation (1)) with conserved symmetry (herein, all values are in bits).
Fullerene CN
(Partition) a
h C N h X @ C N H r e o r g s t r H r e o r g s t r + s i z e H r e o r g H r e d i s t r Δ h R
C20 (Ci) b
(10 × 2)
3.32193.43990.11800.15820.2762−0.27620
C36-15 (D6h)
(3 × 12)
1.58501.72140.13640.04280.1793−0.17930
C50-271 (D5h)
(3 × 10 + 1 × 20)
1.92192.02350.10150.03770.1392−0.13920
C70-1 (D5h)
(2 × 20 + 3 × 10)
2.23592.31120.07530.03150.1068−0.10680
C84-20 (Td)
(1 × 12 + 3 × 24)
1.95022.01950.06930.02290.0923−0.09230
C84-24 (D6h)
(3 × 12 + 2 × 24)
2.23592.30190.06600.02630.0923−0.09230
a The designations of the fullerene isomers are according to Atlas [45]. b The case of filling C20 (Ci) with enhancing symmetry is described in Section 4.4.
Table 2. Information entropy indices of zero-h carbon species forming complexes with X with conserved symmetry.
Table 2. Information entropy indices of zero-h carbon species forming complexes with X with conserved symmetry.
Carbon Molecule h C N h X @ C N H r e o r g s t r H r e o r g s t r + s i z e H r e o r g H r e d i s t r Δ h R
Fullerene C60 (Ih)00.12070.120700.1207−0.12070
Fulleroid C24 (Oh)00.24230.242300.2423−0.24230
Fulleroid C48 (Oh)00.14370.143700.1437−0.14370
Cyclocarbon C18 (D9h)00.29750.297500.2975−0.29750
Table 3. Information entropy indices of maximum-h fullerenes filled with X (systems with C1 symmetry).
Table 3. Information entropy indices of maximum-h fullerenes filled with X (systems with C1 symmetry).
Fullerene CN h C N h X @ C N H r e o r g s t r H r e o r g s t r + s i z e H r e o r g H r e d i s t r Δ h R
C605.90695.93070.02390.09680.1207−0.12070
C706.12936.14970.02050.08630.1068−0.10680
C806.32196.33990.01790.07800.0960−0.09600
C906.49196.50780.01590.07130.0873−0.08730
Table 4. Information entropy of fullerene structures with Jahn–Teller symmetry reduction and their endohedral complexes, in which the effect vanishes.
Table 4. Information entropy of fullerene structures with Jahn–Teller symmetry reduction and their endohedral complexes, in which the effect vanishes.
Fullerene CNIdealized SymmetryActual Symmetry h C N i d e a l h X @ C N i d e a l h C N
C20-1Ih (1 × 20)Ci (10 × 2)0.00000.27623.3219
C24-1D6d (2 × 12)C2 (12 × 2)1.00001.20233.5850
C26-1D3h (1 × 2 + 2 × 6 + 1 × 12)C1 (26 × 1)1.77591.93864.7004
C40-1D5d (4 × 10)Ci (20 × 2)2.00002.11664.3219
C76-2Td (1 × 4 + 2 × 12 + 2 × 24)S4 (3 × 4 + 4 × 8 + 2 × 16)2.11482.18732.9848
C80-7Ih (1 × 20 + 1 × 60)C3v (1 × 2 + 1 × 6 + 6 × 12)0.81130.89722.8766
Table 5. Information entropy indices of filling fullerenes accompanied with increasing symmetry of the fullerene cage.
Table 5. Information entropy indices of filling fullerenes accompanied with increasing symmetry of the fullerene cage.
Fullerene CN H r e o r g s t r H r e o r g s t r + s i z e H r e o r g H r e d i s t r Δ h R
C20-1 (CiIh)−3.04570.1582−2.8875−0.2762−3.1637
C24-1 (C2D6d)−2.38270.1434−2.2393−0.2423−2.4816
C26-1 (C1D3h)−2.76180.1741−2.5877−0.2285−2.8163
C40-1 (CiD5d)−2.20530.1054−2.0999−0.1654−2.2653
C76-2 (S4Td)−0.79740.0388−0.7587−0.1000−0.8587
C80-7 (C3vIh)−1.97940.0355−1.9438−0.0960−2.0398
Table 6. Information entropy indices of filling C60 (Ih) with X coordination relative to the cage.
Table 6. Information entropy indices of filling C60 (Ih) with X coordination relative to the cage.
Coordination of X Atom inside C60Symmetry of X@C60 and Partition H r e o r g = H r e o r g s t r = h X @ C 60 H r e d i s t r Δ h R
PentagonC5v (4 × 5 + 4 × 10 + 1 × 1)2.991−0.1207+2.870
HexagonC3v (8 × 6 + 4 × 3 + 1 × 1)3.585−0.1207+3.464
5.6-bondCS (5 × 1 + 28 × 2)5.013−0.1207+4.892
6.6-bondC2v (5 × 1 + 28 × 2)5.013−0.1207+4.892
AtomCS (5 × 1 + 28 × 2)5.013−0.1207+4.892
Table 7. Information entropy indices of filling C60 isomer with different symmetries.
Table 7. Information entropy indices of filling C60 isomer with different symmetries.
IsomerPartitionσχ h C N h X @ C N H r e o r g s t r
C60 (Ih)1 × 6030354.40.0000.1210.121
C60 (D5d)2 × 20 + 2 × 1010239.31.9182.0080.089
C60 (D2h)3 × 4 + 6 × 841683.1073.1770.070
C60 (C2v)6 × 2 + 12 × 421084.1074.1600.053
C60 (C2)30 × 22604.9074.9470.040
C60 (C1)60 × 1105.9075.9310.024
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sabirov, D.S.; Tukhbatullina, A.A.; Shepelevich, I.S. Digitalizing Structure–Symmetry Relations at the Formation of Endofullerenes in Terms of Information Entropy Formalism. Symmetry 2022, 14, 1800. https://doi.org/10.3390/sym14091800

AMA Style

Sabirov DS, Tukhbatullina AA, Shepelevich IS. Digitalizing Structure–Symmetry Relations at the Formation of Endofullerenes in Terms of Information Entropy Formalism. Symmetry. 2022; 14(9):1800. https://doi.org/10.3390/sym14091800

Chicago/Turabian Style

Sabirov, Denis Sh., Alina A. Tukhbatullina, and Igor S. Shepelevich. 2022. "Digitalizing Structure–Symmetry Relations at the Formation of Endofullerenes in Terms of Information Entropy Formalism" Symmetry 14, no. 9: 1800. https://doi.org/10.3390/sym14091800

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