Skip to content
Publicly Available Published by De Gruyter February 12, 2019

Hückel Molecular Orbital Quantities of {X,Y}-Cyclacene Graphs Under Next-Nearest-Neighbour Approximations in Analytical Forms

  • Tapanendu Ghosh , Swapnadeep Mondal , Sukanya Mondal and Bholanath Mandal EMAIL logo

Abstract

Hückel molecular orbital (HMO) quantities, viz., electron densities, charge densities, bond orders, free valences, total π-electron energies and highest occupied molecular orbital-lowest unoccupied molecular orbital (HOMO–LUMO) or band gaps of {X,Y}-cyclacene graphs under next-nearest-neighbour (nnn) approximations are expressed in analytical forms within a certain range of nnn approximation parameter (m). The critical values of m for {X,Y}-cyclacenes with varying X (=C, N, B) and Y (=C, N, B) are calculated. For {X,X}-cyclacenes with a π-electron on each atom, all HMO quantities except total π-electron energies for a given value of m are found to be independent of X. The cyclic dimer (CD) is constructed in obtaining the eigenvalues corresponding to the singular points of the density of states (DOS) of such {X,Y}-cyclacene. Although the HOMO–LUMO gap of the CD differs from that of the cyclacene with a large number of repeating units (i.e. n ⟶ ∞) but becomes the same for m = 0. The analytical expressions can be used for facile computer programming in obtaining the HMO quantities. Such nnn interaction approximations actually release, to some extent, the strain that results in due to the geometrical structures of such cyclacenes, which is evident from the plots of strain energy per segment vs. contribution of such interactions on the total π-electron energy, where the slopes decrease with an increase in m. The vertical absorption energy difference for singlet-triplet states bears excellent linear correlation with the HOMO–LUMO gaps for a certain m value (m = 0.3) in the case of an even n, but for an odd n, such energy difference remains invariant.

1 Introduction

A cyclacene [1], [2], [3], [4], [5], [6], [7], [8] is a cylindrical or belt-like polycyclic aromatic compound composed of only hexagonal rings and is supposed to be obtained by gluing two bonds of the two terminal hexagons of a linear polyacene. Such compounds were proposed as potentially game-changing molecules in chemistry [5], [6] even before the discovery of carbon nanotubes (CNTs) [7] in 1991, although it appeared in the literature in 1954 as a hypothetical molecule for theoretical study [8]. Cyclacenes, being the building blocks of several important carbon nanostructures, like nanotubes and nanotories, have become important materials for experimental and theoretical investigation [1], [2], [3], [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18], [19], [20]. Cyclacenes are found to appear as intervening belt in the elongated fullerenes [1], [2], [3], [18], [19], [20], [21]. Wegner presented the first synthesis of an aromatic nanobelt on the way to carbon nanotubes [22]. The synthesis of one type of cyclacenes through iterative Wittig reactions followed by a nickel-mediated aryl–aryl coupling reaction was reported, and ultimately the cylindrical shape of its belt structure was confirmed by X-ray crystallography [23]. Very recently, the difficulties for the synthesis of the cyclacenes over the linear analogues due to their radicaloid character were addressed, and the potential use of these systems as molecular templates for the growth of well-shaped carbon nanotubes were explored [24] using Finite-Temperature Density Functional Theory (FT-DFT) and restricted-active-space (RAS) spin-flip (SF) methods. Battaglia et al. [25], [26] used the complete active space self-consistent field method and second-order n-electron valence perturbation theory to investigate the radicality and the vertical singlet–triplet energy gap of cyclacenes as a function of the system size and performed a theoretical study for computing the orbital energies and eigenvectors analytically for such molecules at tight-binding level. DFT calculations on heterocyclacenes with respect to geometry and relative energies were made, and thermodynamically stable macrocycles built from linear annelation of conjugated rings to a hoop shape were predicted [27].

Figure 1: (a) {X,Y}-cyclacene with butadiene as repeating unit (portion with bold edges), (b) {X,Y}-cyclacene with next-nearest-neighbour (interactions) approximations indicting butadiene as repeating unit (portion with bold edges), (c) {X,X}-cyclacene, (d) {X,X}-cyclacene with next-nearest-neighbour (interactions) approximations.
Figure 1:

(a) {X,Y}-cyclacene with butadiene as repeating unit (portion with bold edges), (b) {X,Y}-cyclacene with next-nearest-neighbour (interactions) approximations indicting butadiene as repeating unit (portion with bold edges), (c) {X,X}-cyclacene, (d) {X,X}-cyclacene with next-nearest-neighbour (interactions) approximations.

Very recently, eigensolutions (eigenvalues and eigenvectors) [1], [2] of {X,Y}-cyclacenes with X (=C, N, B, etc.) and Y (=C, N, B, etc.) were obtained in analytical forms under the nearest neighbour (nn) as well as the next-nearest-neighbour (nnn) approximations. The weight of an edge (bond) between two vertices (X and Y) under nnn approximations is expressed as kXY=mkXY, where kXY is the weight of a bond between X and Y under nn approximations, and m is a parameter that measures nnn interactions over nn interaction between two concerned vertices with values that lie in the range 0m1. The parameter m is, thus, a measure of extra restriction imposed on the system over the nn interaction. In the present study, the Hückel quantities (like electron density, charge density, bond orders, free valences, total π-electron energy, etc.) of such cyclacenes with the nnn approximations [2], [28] are obtained in analytical forms and are discussed in the ongoing sections.

2 Methodology

Considering a butadiene unit as the recurring unit n-{X,Y}-cyclacene graph shown in Figure 1, the analytical expressions of eigenvaluues (λt) and eigenvectors or molecular orbitals (Ψ(λt)) were recently found out. Few such cyclacene graphs are shown in Figure 1. As given in [1] and [2], on application of rotational symmetry followed by plane of symmetry, the basis eigenvectors or symmetry adapted linear combinations (SALCs) of atomic orbitals on the vertices of two symmetry fragmented graphs (+ and −) of cyclacene (G+ and G) with nnn approximations [2] are found to be obtained as

(1)ψ1±4=12np=1nεpt(φp(1)±φp(4))ψ2±3=12np=1nεpt(φp(2)±φp(3))};εt=e2iπt/n

Now the eigenvectors (MOs) in the t-th representation can be expressed [1], [2] as

(2)ψ(λt+)=12np=1nεpt{c1+(φp(1)+φp(4))+c2+(φp(2)+φp(3))}ψ(λt)=12np=1nεpt{c1(φp(1)φp(4))+c2(φp(2)φp(3))}};εt=e2iπt/n

On the basis of (2), two secular equations are obtained, as given below, that can be used for expressing eigenvalues and eigenvectors or molecular orbitals (MOs) of a cyclacene graph G.

(3)(λt++hX+mkXY(εt+εt)kXY(1+m)(1+εt)kXY(1+m)(1+εt)hY+kYY+λt++mkXY(εt+εt))×(c1+c2+)=(00)(λt+hX+mkXY(εt+εt)kXY(1m)(1+εt)kXY(1m)(1+εt)hYkYY+λt+mkXY(εt+εt))×(c1c2)=(00)};εt=e2πit/n

The relations given in (3) result in the following eigenvalue equations [1], [2]

(4)(λt+)2{hX+hY+kYY+4mkXYcos(2πt/n)}λt+{2kXY2(1+m)2(1+cos(2πt/n))(hX+2mkXYcos(2πt/n))×(hY+kYY+2mkXYcos(2πt/n))}=0(λt)2{hX+hYkYY+4mkXYcos(2πt/n)}λt{2kXY2(1m)2(1+cos(2πt/n))(hX+2mkXYcos(2πt/n))×(hYkYY+2mkXYcos(2πt/n))}=0}

The solutions of (4) results in the eigenvalue expression [1], [2] as follows

(5)λt++=12{(hX+hY+kYY+4mkXYcos(2πtn))+(hXhYkYY)2+8kXY2(1+m)2{1+cos(2πtn)}}λt+=12{(hX+hY+kYY+4mkXYcos(2πtn))(hXhYkYY)2+8kXY2(1+m)2{1+cos(2πtn)}}λt+=12{(hX+hYkYY+4mkXYcos(2πtn))+(hXhY+kYY)2+8kXY2(1m)2{1+cos(2πtn)}}λt=12{(hX+hYkYY+4mkXYcos(2πtn))(hXhY+kYY)2+8kXY2(1m)2{1+cos(2πtn)}}};t=1,2,,n

Using (5) along with the required normalization conditions for such cases, (2) can be found to be written as

(6)ψ(λt±+)=12np=1nεpt{c1+(λt±+)(φp(1)+φp(4))+c2+(λt±+)(φp(2)+φp(3))}ψ(λt±)=12np=1nεpt{c1(λt±)(φp(1)φp(4))+c2(λt±)(φp(2)φp(3))}};εt=e2iπt/n

where

(7)c1+(λt±+)=kXY(1+m)(1+εt){λt±+hX2mkXYcos(2πt/n)}2+2kXY2(1+m)2{1+cos(2πt/n)}c2+(λt±+)=λt±+hX2mkXYcos(2πt/n){λt±+hX2mkXYcos(2πt/n)}2+2kXY2(1+m)2{1+cos(2πt/n)}c1+(λt++)=0,c2+(λt++)=1c1+(λt+)=1,c2+(λt+)=0}neven,t=n/2}

and

(8)c1(λt±)=kXY(1m)(1+εt){λt±hX2mkXYcos(2πt/n)}2+2kXY2(1m)2{1+cos(2πt/n)}c2(λt±)=λt±hX2mkXYcos(2πt/n){λt±hX2mkXYcos(2πt/n)}2+2kXY2(1m)2{1+cos(2πt/n)}c1(λt+)=1,c2(λt+)=0c1(λt)=0,c2(λt)=1}neven,t=n/2}

Now, to express the Hückel’s quantities [1], [2], [3], [4], [14], [15], [16], [17], [29], [30], [31], we need to find out the eigenvectors (MOs) that are occupied, i.e. to find out the half of the total eigenvalues that have larger values in the β unit than the rest. In the case of m = 0, as we see from the eigenvalue expressions given in (5) that the eigenvalues in the β unit follow the decreasing order as λt++>λt+>λt+>λt and that the eigenvalues {λt++,λt+}>{λt+,λt},t for the odd n, whereas for the even n, the same orders follow except that λt+ equals λt+ at t=n/2. Again, with an increase in m (i.e. with an increase in perturbation through the nnn interactions), the energy eigenvalues change, and beyond a critical value of m (depending on n and whether it is odd or even), the above order of energy eigenvalues, i.e. {λt++,λt+}>{λt+,λt},t is not followed on the account of the reshuffling of the energy eigenvalues as well as a faster decrease in the differences, λt+(n/2)λt(t=n) for the even n and λt+(t=n±12)λt(t=n) for the odd n with an increase in m. The critical value of m beyond which the eigenvalues of cyclacene disobey the above order can be calculated with the following eigenspectra of a given n-cyclacene with even and odd n are shown below.

2.1 Critical Values of m for {X,Y}-Cyclacenes With Even n

For an even n, λt+ at t=n/2 is minimum in the eigenvalue set {λt+}, whereas λt at t=n is maximum in the eigenvalue set {λt}. Again, with an increase in m, λt+ at t=n/2 decreases and λt at t=n increases; thus, to get the critical value of m for which the eigenvalues in the set {λt++,λt+} would be greater than that in the set {λt+,λt} for any value of n, we need to equate λt+ at t=n/2 and λt+ at t=n/2 as shown below for different cases.

Case I: If the value of hX exceeds (hY+kYY), then, the expressions of λt+ at t=n/2 and λt at t=n would be

(9)λt+=12(2hX2kYY4mkXY)fort=n/2λt=12{(hX+hYkYY+4mkXY)(hXkYYhY)2+ 16kXY2(1m)2}fort=n}

Now putting λt+(t=n/2)=λt(t=n), we have the following quadratic equation for m

(10)m2(hXhYkYY2kXY3kXY)m13=0

whose solution gives the critical value of m (taking the positive root) as

(11)m=(hXhYkYY2kXY6kXY)+(hXhYkYY2kXY6kXY)2+13

below which the eigenvalues obey the relation {λt++,λt+}>{λt+,λt}.

Case II: If (hY+kYY) exceeds hX, then, the expressions of λt+ at t=n/2 and λt at t=n take the forms as

(12)λt+=12{(hX+hYkYY4mkXY+hY+kYYhX)}=12(2hY4mkXY)λt=12{(hX+hYkYY+4mkXY)(hXkYYhY)2+16kXY2(1m)2}}

As for {B,N}-cyclacenes hX=hB=0.45, hY=hN=1.37, hYY=hNN=0.98, hXY=hBN=0.53, thus, for all values of m (0m1), from (12), we can have

λt+λt=12{(hNhB+kNN8mkBN)+(hBkNNhN)2+16kBN2(1m)2}>0

i.e. for such cyclacenes, frontier orbitals do not change with the change in m values within the specified range.

For other cyclacenes, equating λt+ and λt of the above (12), we have the following quadratic equation of m

(13)3m2+m(hXhYkYY+2kXYkXY)13=0

whose acceptable root is the positive as m is positive and is given as

(14)m=(hXhYkYY+2kXY)6kXY+{(hXhYkYY+2kXY)6kXY}2+13

We can see from (14) that the critical m does not depend on the number of rings of cyclacene. The values of m were calculated for different {X,Y}-cyclacenes with X, Y = C, N, B and are given in Table 1. The values of hX, hY, kXX, kXY=kYX, and kYY for the required calculations are taken from the work of Van-Catledge [32].

Table 1:

Critical values of m (nnn interaction parameter) for which the eigenvalues λt++ and λt+ are to be greater than the eigenvalues λt+ and λt of {X,Y}-cyclacenes.

{X,Y}-CyclaceneCritical values of m
n-evenn-odd
{C,C} hC=0.0, kCC=1.00.4340.420
{N,N} hN=1.37, kNN=0.980.4340.420
{B,B} hB=0.45, kBB=0.870.4340.420
{C,N} hC=0.0, hN=1.37, kCN=0.89, kNN=0.980.6930.452
{N,C} hN=1.37, hC=0.0, kCN=0.89, kCC=1.00.3700.415
{C,B} hC=0.0, hB=0.45, kCB=0.73, kBB=0.870.3860.416
{B,C} hB=0.45, hC=0.0, kCB=0.73, kCC=1.00.5750.436
{N,B} hN=1.37, hB=0.45, kNB=0.53, kBB=0.870.5430.432

2.2 Critical Values of m for {X,Y}-Cyclacenes With Odd n

For an odd n, two eigenvalues, λt+ at t=(n±1)/2, are degenerate and are minimum in the set {λt+}, whereas the eiegenvalue λt at t=n is maximum in the set {λt} and with an increase in m, λt+ at t=(n±1)/2 decreases and λt at t=n increases. The expressions of λt+ at t=(n±1)/2 and λt at t=n are given below.

(15)λt+=12{(hX+hYkYY4mkXYcos(πn))+(hXkYYhY)2+8kXY2(1m)2{1cos(πn)}}fort=n±12λt=12{(hX+hYkYY+4mkXY)(hXkYYhY)2+16kXY2(1m)2}fort=n}

For {B,N}-cyclacenes, hB=0.45,hN=1.37,kNB=0.53,kNN=0.98; thus, λt+λt becomes

λt+λt=12{4mkBN(1+cos(πn))+(hBkNNhN)2+8kBN2(1m)2{1cos(πn)}+(hBkNNhN)2+16kBN2(1m)2}

which is always positive for all values of m (0m1), thus, for such cyclacenes, frontier orbitals do not change with the m values within the specified range.

For all other cyclacenes, the critical value of m, below which {λt++,λt+} would be greater than that in the set {λt+,λt}for any value of n, can be obtained by just equating λt+ at t=(n±1)/2 and λt at t=n given in the above (15) that results in a quadratic equation of m (critical value for such case) as shown below.

(16)m2+2{3cos(πn)}{(1+cos(πn))2{3cos(πn)}}m(hXhYkYY2kXY)2+{3cos(πn)}{(1+cos(πn))2{3cos(πn)}}=0

that ultimately results in m in following form

(17)m={3cos(πn)}{{3cos(πn)}(1+cos(πn))2}±{{3cos(πn)}(1+cos(πn))2{3cos(πn)}}2+(hXhYkYY2kXY)2+{3cos(πn)}(1+cos(πn))2{3cos(πn)}

As for n = 3, the coefficient of m shown in (16) is negative, the value of m in (17) would be taken as

(18)m={3cos(πn)}{{3cos(πn)}(1+cos(πn))2}{{3cos(πn)}(1+cos(πn))2{3cos(πn)}}2(hXhYkYY2kXY)2+{3cos(πn)}{3cos(πn)}(1+cos(πn))2

whereas for all other values of n, the coefficient of m given in (16) is positive; thus, the value of m would be taken as

(19)m={3cos(πn)}{{3cos(πn)}(1+cos(πn))2}+{{3cos(πn)}(1+cos(πn))2{3cos(πn)}}2+(hXhYkYY2kXY)2+{3cos(πn)}(1+cos(πn))2{3cos(πn)}

From (16), it reveals that the critical value of m for odd cyclacenes depends on n, i.e. the number of rings in a given cyclacene and approximate values of m in varying n can be calculated with the help of (18) and (19).

For a large n, the value of m as given by the (19) reaches at

(20)m=1+2+(hXhYkYY8kXY)2

The value of m given in (20) is the value up to which one can ascertain about the validity of λt++>λt+>λt+>λt and {λt++,λt+}>{λt+,λt},t for any kind of cyclacene with varying n. Such values of m were calculated for different {X,Y}-cyclacenes with X, Y = C, N, B taking the values of hX, hY, kXX, kXY=kYX and kYY from the work of Van-Catledge [32] and are given in Table 1.

Thus, to calculate the Hückel MO (HMO) quantities [1], [2], [3], [4], [14], [15], [16], [17], [29], [30], [31], [32], [33], [34], [35], [36] under nnn approximations [2], [28] within the range of m given in Table 1, we need to consider the MOs with eigenvalues λt+± for odd n and MOs with eigenvalues λt+± along with λt+(t=n/2) for an even n as follows [1], [2].

(21)ψ(λt++)={(kXY(1+m)(1+εt))p=1nεpt(φp(1)+φp(4))2n{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+(λt++hX2mkXYcos(2πt/n))p=1nεpt(φp(2)+φp(3))2n{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}}ψ(λt++)=12np=1n(1)p(φp(2)+φp(3)),neven,t=n/2}
(22)ψ(λt+)={kXY(1m)(1+εt)p=1nεpt(φp(1)φp(4))2n{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}+(λt+hX2mkXYcos(2πt/n))p=1nεpt(φp(2)φp(3))2n{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}ψ(λt+)=12np=1n(1)p(φp(1)φp(4)),neven,t=n/2}
(23)ψ(λt+)={kXY(1+m)(1+εt)p=1nεpt(φp(1)+φp(4))2n{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+(λt+hX2mkXYcos(2πt/n))p=1nεpt(φp(2)+φp(3))2n{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}}ψ(λt+)=12np=1n(1)p(φp(1)+φp(4)),neven,t=n/2}
(24)ψ(λt)={kXY(1m)(1+εt)p=1nεpt(φp(1)φp(4))2n{(λthX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}+(λthX2mkXYcos(2πt/n))p=1nεpt(φp(2)φp(3))2n{(λthX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}ψ(λt)=12np=1n(1)p(φp(2)φp(3)),neven,t=n/2}

Using (5), (21)–(24), the HMO quantities, viz., electron density, charge density, bond orders, free valences, total π-electron energy, etc., were derived in analytical forms and are given in the following sections.

3 Hückel Molecular Orbital (HMO) Quantities

From the symmetry of a cyclacene graph, we can see that there are two different sets of vertices (1 and 2) and only two different sets of symmetry equivalent bonds (1–2 and 2–3); thus, for such systems, there are two sets of charge densities, two sets of bond orders, and hence, two sets of free valences that were derived in analytical forms as discussed in the following sections.

3.1 Electron Density and Charge Density

Electron density (qre) and charge density (qr) at r-th atom are defined [29], [30], [31], [33], [34], [35], [36] as

(25)qre=t=1nNjcrjcrjandqr=1qre

Thus, for the odd n with the help of (21), (22), and (25), the electron and charge densities can be expressed as

(26)q1e=1nt=1n[2kXY2(1+m)2(1+cos(2πt/n)){(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+2kXY2(1m)2(1+cos(2πt/n)){(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}]q1=1q1e}
(27)q2e=1nt=1n[(λt++hX2mkXYcos(2πt/n))2{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+(λt+hX2mkXYcos(2πt/n))2{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}]q2=1q2e}

In the case of the odd cyclacenes with X = Y, the above (26) and (27) reduce to the following equations on substitution of (5)

(28)q1e=1n×t=1n[8(1+m)2{1+cos(2πtn)}{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}+8(1m)2{1+cos(2πtn)}{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}]q1=1q1e}

and

(29)q2e=1n×t=1n[{1+1+8(1+m)2{1+cos(2πtn)}}2{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}+{1+1+8(1m)2{1+cos(2πtn)}}2{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}]q2=1q2e}

respectively.

For the even n, as we see from the expressions of the MOs that at t = n/2, the MO ψ(λt++) have no contribution, whereas both MOs ψ(λt+) and ψ(λt+) contribute equally (1/2n) to the electron densities of atoms 1 and 4. Again towards the electron densities of atoms 2 and 3, the MO ψ(λt++) at t = n/2 has a contribution of (1/2n) to each of the atoms. Thus, considering the above contributions for such {X,Y}-cyclacene graphs, the electron density and charge density at atom 1 and atom 2 can be expressed with the help of (21), (22), (23), and (25) as

(30)q1e=1n×[1+t=1,n/2n[2kXY2(1+m)2{1+cos(2πtn)}{{λt++hX2mkXYcos(2πtn)}2+2kXY2(1+m)2{1+cos(2πtn)}}+2kXY2(1m)2{1+cos(2πtn)}{{λt+hX2mkXYcos(2πtn)}2+2kXY2(1m)2{1+cos(2πtn)}}]]q1=1q1e}
(31)q2e=1n×[1+t=1,n/2n[{λt++hX2mkXYcos(2πtn)}2{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2{1+cos(2πtn)}}+{λt+hX2mkXYcos(2πtn)}2{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2{1+cos(2πtn)}}]]q2=1q2e}

Using (5), the above (30) and (31) for the {X,Y}-cyclacenes with X = Y are converted to

(32)q1e=1n[1+t=1,n/2n[8(1+m)2{1+cos(2πtn)}{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}+8(1m)2{1+cos(2πtn)}{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}]]q1=1q1e}

and

(33)q2e=1n[1+t=1,n/2n[{1+1+8(1+m)2{1+cos(2πtn)}}2{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}+{1+1+8(1m)2{1+cos(2πtn)}}2{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}]]q2=1q2e}

From (28), (29), (32), and (33), we can see that for the atom-homogeneity (X = Y) of cyclacenes, the electron and, hence, the charge densities are independent on the atom made up of the cyclacene if each atom of the molecule carries a π-electron.

From the above (26), (27), (30), and (31), we have

(34)q1e+q2e=2q1+q2=2(q1e+q2e)=0}

For {X,Y}-cyclacenes with no nnn interactions and with X = Y, i.e. m = 0, (28), (29), (32), and (33) reduce to

(35)q1e=q2e=1q1=q2=0}

which is consistent with (34). Electron densities and charge densities for {X,X}-cyclacenes with X = C, B, N are calculated using (28), (29), (32), and (33) and given in Table 2.

Table 2:

Values of electron densities (ED), charge densities (CD), bond orders (BO), free valences (FV), and total π-electron energies in β unit for {C,C}-, {N,N}-, and {B,B}-cyclacenes with m (nnn approximations parameter) ranging from 0.0 to 0.4.

{X,Y}-CyclacenenEDCDBOFVEπ in β unit
12121–22–312{C,C}/{N,N}/{B,B}
{C,C} {N,N} {B,B} m = 0.031.0001.0000.0000.0000.6220.3790.4890.11017.190/33.287/9.556
41.0001.0000.0000.0000.5760.4770.5800.10322.246/43.721/12.154
51.0001.0000.0000.0000.5990.4180.5330.11528.154/54.991/15.494
61.0001.0000.0000.0000.5880.4490.5560.10733.613/65.820/18.443
71.0001.0000.0000.0000.5940.4310.5440.11339.309/76.883/21.599
81.0001.0000.0000.0000.5910.4410.5500.10944.874/87.816/24.640
{C,C} {N,N} {B,B} m = 0.131.0320.968−0.0320.0320.6210.3810.4910.11017.200/33.296/9.564
41.0210.979−0.0210.0210.5750.4790.5820.10322.254/43.729/12.161
51.0310.969−0.0310.0310.5980.4200.5350.11528.169/55.006/15.507
61.0240.976−0.0240.0240.5870.4510.5570.10733.627/65.835/18.456
71.0290.971−0.0290.0290.5930.4330.5450.11339.330/76.903/21.617
81.0260.974−0.0260.0260.5900.4430.5520.10944.894/87.837/24.658
{C,C} {N,N} {B,B} m = 0.231.0640.936−0.0640.0640.6170.3890.4970.10917.229/33.325/9.590
41.0420.958−0.0420.0420.5730.4840.5860.10222.280/43.755/12.184
51.0620.938−0.0620.0620.5960.4260.5410.11528.217/55.052/15.549
61.0500.950−0.0500.0500.5850.4570.5620.10633.673/65.879/18.495
71.0590.941−0.0590.0590.5910.4380.5500.11239.392/76.964/21.671
81.0530.947−0.0530.0530.5880.4490.5570.10844.959/87.900/24.714
{C,C} {N,N} {B,B} m = 0.331.0990.901−0.0990.0990.6120.4010.5080.10817.281/33.375/9.634
41.0660.934−0.0660.0660.5690.4940.5940.10022.326/43.799/12.223
51.0950.905−0.0950.0950.5910.4360.5510.11428.299/55.133/15.620
61.0770.923−0.0770.0770.5800.4670.5720.10533.753/65.958/18.565
71.0900.910−0.0900.0900.5860.4480.5600.11239.502/77.072/21.766
81.0820.918−0.0820.0820.5830.4590.5660.10745072/88.011/24.813
{C,C} {N,N} {B,B} m = 0.431.1360.864−0.1360.1360.6030.4190.5260.10717.359/33.452/9.702
41.0920.908−0.0920.0920.5620.5080.6080.10022.396/43.868/12.285
51.1300.870−0.1300.1300.5830.4510.5660.11528.423/55.255/15.728
61.1070.893−0.1070.1070.5730.4820.5860.10433.876/66.078/18.672
71.1240.876−0.1240.1240.5790.4620.5740.11339.667/77.233/21.910
81.1130.887−0.1130.1130.5750.4740.5810.10745.245/88.180/24.963

3.2 Bond Orders

The bond order prs of the bond (r-s) between the r and s atoms is defined [4], [16], [17], [29], [30], [31], [33], [34], [35], [36], [37], [38], [39], [40] as

(36)prs=jOccupied MOsNjcrjcsj

where crj is the jth MO coefficient at the rth atom of cyclacene, and Nj is the population of the jth MO.

In the case of the odd n, from (5), we can see that for t = n, such cyclacene graph has nondegenerate eigenvalues, but for tn, eigenvalues are doubly degenerate as the tth and (n-t)th representations are equivalent. For this case, the expressions of bond orders given in (36) take the form

(37)prs=2{crn+(λn++)csn+(λn++)+crn(λn+)csn+(λn+)}+2t=1(n1)/2{crt+(λt++)cst+(λt++)+crt+(λt++)cst+(λt++)+crt(λt+)cst+(λt+)+crt(λt+)cst+(λt+)}

Taking r-s as 1–2 we have

(38)p12=2{c1n+(λn++)c2n+(λn++)+c1n(λn+)c2n+(λn+)}+2t=1(n1)/2{c1t+(λt++)c2t+(λt++)+c1t+(λt++)c2t+(λt++)+c1t(λt+)c2t+(λt+)+c1t(λt+)c2t+(λt+)}

Thus, putting the values of the coefficients in (38), the final form of bond order of the 1–2 bond is as follows

(39)p12=(kXYn)×t=1n[{{(1+m)(λt++hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+{(1m)(λt+hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}]

Similarly, the bond order of the 2–3 bond can be expressed as

(40)p23=2{c2n+(λn++)c3n+(λn++)+c2n(λn+)c3n+(λn+)}+2t=1(n1)/2{c2t+(λt++)c3t+(λt++)+c12+(λt++)c3t+(λt++)+c2t(λt+)c3t+(λt+)+c2t(λt+)c3t+(λt+)}

As c2t+=c2t+ and c3t+=c3t+, thus, the above (40) can be written as

(41)p23=2{c2n+(λn++)c3n+(λn++)+c2n(λn+)c3n+(λn+)}+2t=1(n2){c2t+(λt++)c3t+(λt++)+c2t(λt+)c3t+(λt+)}=2t=1n{c2t+(λt++)c3t+(λt++)+c2t(λt+)c3t+(λt+)}

Equation (41) finally takes the following form. Following substitution, the coefficients form (21) and (22)

(42)p23=1n×t=1n{(λt++hX2mkXYcos(2πt/n))2{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}(λt+hX2mkXYcos(2πt/n))2{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}

In the case of {X,X}-cyclacenes with odd number of vertices, the above (39) and (42) for bond orders get reduced to

(43)p12=(1n)×t=1n{{2(1+m){1+1+ 8(1+m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}+{2(1m){1+1+8(1m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}}
(44)p23=1n×t=1n{{1+1+8(1+m)2{1+cos(2πtn)}}2{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}{1+1+8(1m)2{1+cos(2πtn)}}2{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}}

In the case of the even cyclacene graphs, the representations corresponding to t = 0 and n/2 are nondegenerates, whereas the rest of the representations are doubly degenerate. For such cyclacene graphs, the bond orders can be expressed as

(45)prs=2{{crn+(λn++)csn+(λn++)+crn(λn+)crn+(λn+)}+{crn/2+(λn/2++)csn/2+(λn/2++)+crn/2(λn/2+)crn/2+(λn/2+)}+t=1(n2)/2{crt+(λt++)cst+(λt++)+crt+(λt++)cst+(λt++)+crt(λt+)crt+(λt+)+crt(λt+)crt+(λt+)}}

Thus,

(46)p12=2{{c1n+(λn++)c2n+(λn++)+c1n(λn+)c2n+(λn+)}+{c1n/2+(λn/2++)c2n/2+(λn/2++)+c1n/2(λn/2+)c2n/2+(λn/2+)}+t=1(n2)/2{c1t+(λt++)c2t+(λt++)+c1t+(λt++)c2t+(λt++)+c1t(λt+)c2t+(λt+)+c1t(λt+)c2t+(λt+)}}

For t = n/2 from the expressions of the MOs, we can see that the coefficients of φp(1) and φp(4) have zero values; thus, the MOs corresponding to the representations t = n/2 have no contributions to the bond orders of 1–2 and 3–4 bonds. Thus, the bond order of the 1–2 bond for such cyclacenes is expressed as

p12=2{{c1n+(λn++)c2n+(λn++)+c1n(λn+)c2n+(λn+)}+t=1(n2)/2{c1t+(λt++)c2t+(λt++)+c1t+(λt++)c2t+(λt++)+c1t(λt+)c2t+(λt+)+c1t(λt+)c2t+(λt+)}},

which takes the final form as below

(47)p12=(kXYn)×t=1,n/2n[{{(1+m)(λt++hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+{(1m)(λt+hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}]

For such cyclacenes, the bond order of the 2–3 bond can be expressed as

(48)p23=2{{c2n+(λn++)c3n+(λn++)+c2n(λn+)c3n+(λn+)}+{c2n/2+(λn/2++)c3n/2+(λn/2++)+c2n/2(λn/2+)c3n/2+(λn/2+)}+t=1(n2)/2{c2t+(λt++)c3t+(λt++)+c2t+(λt++)c3t+(λt++)+c2t(λt+)c3t+(λt+)+c2t(λt+)c3t+(λt+)}}

Putting the values of c2n/2+ and c3n/2+ from (23) as well as c2t+=c2t+ and c3t+=c3t+, the above equation can be written as

(49)p23=1n+2t=1,n/2n{c2t+(λt++)c3t+(λt++)+c2t(λt+)c3t+(λt+)}

which is finally reduced after incorporation of the values of the coefficients from (21) and (22) to the following form

(50)p23=1n×{1+t=1,n/2n{(λt++hX2mkXYcos(2πt/n))2{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}(λt+hX2mkXYcos(2πt/n))2{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}}

For {X,X}-cyclacenes with even n, (47) and (50) are reduced to

(51)p12=(kXYn)×t=1,n/2n{{2(1+m){1+1+8(1+m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}+{2(1m){1+1+8(1m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}}

and

(52)p23=1n×{1+t=1,n/2n{{1+1+8(1+m)2{1+cos(2πtn)}}2{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}{1+1+8(1m)2{1+cos(2πtn)}}2{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}}}

Thus, (43), (44), (51), and (52) show that the {X,Y}-cyclacenes with atom-homogeneity (i.e. X = Y), and the bond orders do not depend on the atom made up of the cyclacene if each atom of the molecule carries an π-electron. The bond orders of the 1–2 and 2–3 bonds for the {X, X}-cyclacenes with X = C, B, N were calculated using (43), (44), (51), and (52) and are given in Table 2.

3.3 Free Valence

Free valence [33], [34], [35], [41], [42], [43], [44], [45], [46] is considered as one of the criteria to measure the reactivity of an atom in a molecule. This concept can be considered as the quantum mechanical expression of ideas advanced at the beginning of the century by Flürscheim (affinity demand), Werner (residual affinity), Thiele (partial valence) according to which the atoms in a molecule may not have used up their capacity to combine [47]. Each atom in a conjugated system (molecule) can exert only a certain π-bonding power (real), although it has a maximum capacity (theoretical, threshold) for combining (bonding) in the molecule. The difference between the theoretical capacity and real capacity of bonding is always positive. As this difference goes up, the surplus valence increases at the position for accommodating another group or atom at that position; thus, reactivity increases as free valence increases at a position. Free valence was originally defined by Linus Pauling in terms of bond orders calculated from the resonance valence bond theory. It was shown later on that a similar conclusion of free valence can be reached from the HMO theory, and it was C.A. Coulson who first proposed the free valence (Fr) in the HMO formalism [41], [42] as

(53)Fr=Nmaxsprs=3sprs

where Nmax is the maximum bond order at atom r, and prs is the bond order of the bond between the r and s atoms. Here, Nmax is the sum of the bond orders at the central vertex (atom) of a hypothetical molecule or molecular species where the central vertex is bonded to three peripheral vertices each with one electron including the central vertex.

The free valence of vertex 1 or 4 can be expressed with the help of (53) as

(54)F1=F4=3ss1p1s=3ss4p4s=32p12=32p43

For an odd n, putting the bond orders value from (39) in (54), we have

(55)F1=3(1n)×t=1n{{2kXY(1+m)(λt++hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+{2kXY(1m)(λt+hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}

Similarly, for the even n, inserting the bond order value from (47) in (54), we have

(56)F1=3(2kXYn)×t=1,n/2n{{(1+m)(λt++hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt++hX2mkXYcos(2πt/n))2+2kXY2(1+m)2(1+cos(2πt/n))}+{(1m)(λt+hX2mkXYcos(2πt/n))(1+cos(2πt/n))}{(λt+hX2mkXYcos(2πt/n))2+2kXY2(1m)2(1+cos(2πt/n))}}

The free valence at atoms 1 or 4 for {X,X}-cyclacenes with odd and even n can be obtained as

(57)F1=3(4n)×t=1n{{(1+m){1+1+8(1+m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}+{(1m){1+1+8(1m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}}

and

(58)F1=3(4n)×t=1,n/2n{{(1+m){1+1+8(1+m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1+m)2{1+cos(2πtn)}}2+8(1+m)2{1+cos(2πtn)}}+{(1m){1+1+8(1m)2{1+cos(2πtn)}}{1+cos(2πtn)}}{{1+1+8(1m)2{1+cos(2πtn)}}2+8(1m)2{1+cos(2πtn)}}}

respectively.

The free valence of the vertex 2 or 3 for such cyclacenes can be expressed as

(59)F2=F3=3ss2p2s=3ss3p3s=32p12p23=3p232p34

Now, for the cyclacenes with an odd n, the above expression (59) is reduced in taking the values of bond orders from (39) and (42) as follows

(60)F2=31nt=1n{1+2kXY(1+m)(1+cos(2πt/n))(λt++hX2mkXYcos(2πt/n))1+2kXY2(1+m)2(1+cos(2πt/n))(λt++hX2mkXYcos(2πt/n))212kXY(1m)(1+cos(2πt/n))(λt+hX2mkXYcos(2πt/n))1+2kXY2(1m)2(1+cos(2πt/n))(λt+hX2mkXYcos(2πt/n))2}

Similarly, for an even n, using (47) and (50), (58) is reduced to

(61)F2=31n×[1+t=1,n/2n{{1+2kXY(1+m)(1+cos(2πt/n))(λt++hX2mkXYcos(2πt/n))1+2kXY2(1+m)2(1+cos(2πt/n))(λt++hX2mkXYcos(2πt/n))2}{12kXY(1m)(1+cos(2πt/n))(λt+hX2mkXYcos(2πt/n))1+2kXY2(1m)2(1+cos(2πt/n))(λt+hX2mkXYcos(2πt/n))2}}]
Table 3:

Strain energies per fragment (EStrain) in kJ mol−1 for {C,C}-cyclacenes from M06-2X and M06-2X D3(BJ) methods along with their respective values of ΔEπ(n) in β unit.

nStrain energy per fragment (EStrain) in kJ mol−1ΔEπ in β unit
M06-2XM06-2X D3(BJ)m = 0.0m = 0.1m = 0.2m = 0.3m = 0.4m = 0.43
6157.7157.60.00.01460.05990.14010.26290.31
7122.2121.60.00.02030.08290.19230.35730.4197
893.693.00.00.02090.08530.19870.37170.4376
981.280.70.00.02530.10320.23970.44620.5245
1065.565.00.00.02690.10950.25480.47580.5598
1160.259.80.00.03050.12430.28870.5380.6326

As given earlier, with the free valence at atom 2 for {X,X}-cyclacenes with odd and even n, (60) and (61) change to the following forms

(62)F2=3(1n)t=1n{1+4(1+m){1+cos(2πtn)}{1+1+8(1+m)2{1+cos(2πtn)}}1+8(1+m)2{1+cos(2πtn)}{1+1+8(1+m)2{1+cos(2πtn)}}21{4(1m){1+cos(2πtn)}}{1+1+8(1m)2{1+cos(2πtn)}}1+8(1m)2{1+cos(2πtn)}{1+1+8(1m)2{1+cos(2πtn)}}2}

and

(63)F2=3(1n)×[1+t=1,n/2n{1+4(1+m){1+cos(2πtn)}{1+1+8(1+m)2{1+cos(2πtn)}}1+8(1+m)2{1+cos(2πtn)}{1+1+8(1+m)2{1+cos(2πtn)}}21{4(1m){1+cos(2πtn)}}{1+1+8(1m)2{1+cos(2πtn)}}1+8(1m)2{1+cos(2πtn)}{1+1+8(1m)2{1+cos(2πtn)}}2}]

Similar to the cases of electron (and, hence, charge) densities and bond orders, the free valences are independent of the atom the cyclacene is made up if each atom of cyclacene with atom-homogeneity caries a π-electron as is evident from (57), (58), (62), and (63). Free valences of vertices 1 and 2 for {X,X}-cyclacenes with X = C, B, N are calculated using (62) and (63) and given in Table 2.

3.4 Total π-Electron Energy

Assuming that each atom in the {X,Y}-cyclacene molecule possesses one π-electron, then, the expression of total π-electron energy [1], [2], [3], [4], [14], [15], [17], [29], [30], [31], [34], [35], [36], [48], [49], [50] can be derived with the help of (5) as

(64)Eπ=2t=1n{λt+++λt+}Eπ=t=1n[2{hX+hY+4mkXYcos(2πt/n)}+{hX(hY+kYY)}2+16kXY2(1+m)2cos2(πt/n)+{hX(hYkYY)}2+16kXY2(1m)2cos2(πt/n)]

For the {X,Y}-cyclacenes with X = Y, i.e. hX=hY=h and kXY=kXX=kYY=k, (64) reduces to

(65)Eπ=t=1n[4{h+2mkcos(2πtn)}+k1+16(1+m)2cos2(2πtn)+k1+16(1m)2(2πtn)]

The systems that belong to such {X,Y}-cyclacene systems are {C,C}, {C,N}2n+, {N,C}2n+, {N,N}4n+, {C,B}2n−, {B,C}2n−, {B,B}4n−, etc. The systems {C,N}2n+ and {N,C}2n+ are generated on removal of one electron from each N-atom of {N,N}-cyclacene; the systems {C,B}2n− and {B,C}2n− are realized after adding one electron to each B-atom of {B,B}. Similarly, the system {N,N}4n+ comes into the picture on removal of one electron from each N-atom of {N,N}-cyclacene system, whereas the {B,B}4n− appears when one electron is added to each B-atom of {B,B}-cyclacene.

Putting (65) under the condition of non-nnn interaction (i.e. m = 0), we have

(66)Eπ=2t=1n[2h+k1+16cos2(2πtn)]

The systems that belong to such {X,Y}-cyclacene systems are {C,C}, {N,N}4n+, {B,B}4n-, etc. The total π-electron energies for the {X,X}-cyclacenes with X = C, B, N were calculated using (65) and are given in Table 2.

3.4.1 Strain Energies of {C,C}-cyclacene and Their Total π-Electron Energies Under nnn Interactions

Strain energies [51] per fragment (EStrain) for the {C,C}-cyclacenes were obtained from the TD-DFT level calculation with dispersion-uncorrected (M06-2X) and dispersion-corrected (M06-2X-D3(BJ)) methods and with the 6-31+G basis set (given in Table 3). The formula used for such calculations is,

EStrain=ΔHf0(anthracene)ΔHf0({C,C}cyclacene)+ΔHf0(naphthalene)

where ΔHf0 refers to the enthalpy of formation at 0K including vibrational energy of the different compounds at their respective optimized geometry. Such energies [51] were found to decrease with the increase in the sizes of such cyclacenes. Now, let us consider Eπ0(n) to be the total π-electron energy for a cyclacene of n hexagonal rings and Eπm(n) be the same for a cyclacene of n hexagonal rings but the only difference is that, in the later case, the nnn interaction is present. Then, the difference ΔEπm=Eπm(n)Eπ0(n) is a measure of the effect of the nnn interaction on the total π-electron energy for a cyclacene of n hexagonal rings. This energy difference ΔEπm given in Table 3 is found to correlate the strain energy per fragment for {C,C}-cyclacenes obtained from the TD-DFT level calculations. As the values of the strain energies per fragment obtained from the methods M06-2X and M06-2X-D3(BJ) are more or less the same, we used the data from method M06-2X. The plot strain energy per fragments vs. ΔEπm obtained are straight lines for the values of the nnn-interaction parameters (m) in the range 0.1–0.43 except the variation of the slopes and intercepts. The sign of ΔEπm is found to be negative from the plots, but as it is in terms of β, thus, the slopes are actually positive. Such interactions actually release, to some extent, the strain that results due to the geometrical structures of such cyclacenes, which is evident from the slopes of the plots that decrease with an increase in m. The equations that fit the plots of the strain energies per fragment (EStrain) vs. ΔEπm with m values in the ranges m = 0.0–0.43 are given below.

EStrain=(242.788±22.246)+(6327.277±940.610)×ΔEπ0.1;Corrln. Coeff.=0.95853EStrain=(243.853±22.598)+(1562.06±234.270)×ΔEπ0.2;Corrln. Coeff.=0.95784EStrain=(244.872±22.518)+(676.277±100.401)×ΔEπ0.3;Corrln. Coeff.=0.95863EStrain=(245.908±22.298)+(365.043±53.305)×ΔEπ0.4;Corrln. Coeff.=0.9599EStrain=(246.306±22.259)+(311.157±45.241)×ΔEπ0.43;Corrln. Coeff.=0.96023

One such plot with m = 0.3 is shown in Figure 2.

Figure 2: Plot of EStrain vs. ΔEπm\(\Delta E_{\pi}^{m}\)for {C,C}-cyclacenes with m = 0.43.
Figure 2:

Plot of EStrain vs. ΔEπmfor {C,C}-cyclacenes with m = 0.43.

3.5 HOMO–LUMO or Band Gap (Δ)

HOMO–LUMO or band gap (Δ) is an important quantity to predict reactivity and stability of a molecule [2], [29], [30], [31], [48]. To find out this quantity for {X,Y}-cyclacenes, we need to figure out first the HOMO and LUMO energy levels from the eigenvalue expressions given in (20). As eigenvalues follow the order λt++>λt+>λt+>λt and {λt++,λt+}>{λt+,λt},t within the range of m given in Table 1, we observe that the HOMO would be λt+ at t=n/2 for the even n and λt+ at t=(n±1)/2 for the odd n. The LUMO for all such cyclacenes with odd n is λt+ at t=(n±1)/2. For the even n, the LUMO is found to be λt+ at t=(n±2)/2 for all cyclacenes with {X,Y} = {C,C}, {B,B}, {N.N}, {C,B}, and {B,C} (Case-I), or it would be λt+ at t=n/2 for cyclacenes with {X,Y} = {B,N}, {C,N}, {N,B}, and {N,C} (Case-II) for some lower values of m depending on the respective values of h and k. Thus, the expressions of HOMO and LUMO energies along with the HOMO–LUMO or band gap (Δ) with lower values of m for the two cases (I and II) are obtained with even and odd n as follows.

Case-I: Cyclacenes with {X,Y} = {C,C}, {B,B}, {N.N}, {C,B}, and {B,C}

HOMO–LUMO expression for n-even

(67)λt+(HOMO)=12{(hX+hYkYY4mkXY)+(hXhY+kYY)2}λt+(LUMO)=12{(hX+hY+kYY4mkXYcos(2πn))(hXhYkYY)2+8kXY2(1+m)2{1cos(2πn)}}}

HOMO–LUMO expression for n-odd

(68)λt+(HOMO)=12{(hX+hYkYY4mkXYcos(πn))+(hXhY+kYY)2+8kXY2(1m)2{1cos(πn)}}λt+(LUMO)=12{(hX+hY+kYY4mkXYcos(πn))(hXhYkYY)2+8kXY2(1+m)2{1cos(πn)}}}
Table 4:

Vertical absorption energy differences ΔEST (=ES1ET1) in eV between singlet–triplet (S1–T1) states of {C,C}-cyclacenes calculated from M06-2X method with the 6-31+G basis set and the HOMO–LUMO gaps in the β unit with varying nnn interactions.

n-{C,C}-cyclacenesΔESTin eVHOMO–LUMO gap (Δ) in β unit
m = 0.0m = 0.1m = 0.2m = 0.3m = 0.4m = 0.43
61.240.620.610.600.590.200.02
70.660.340.340.350.350.370.37
80.980.410.420.430.440.200.02
90.800.280.220.220.230.240.24
100.800.300.310.320.330.200.02
110.920.150.150.150.160.170.17

Thus, the HOMO–LUMO or band gaps (Δ) for the even and odd n (some higher values) in the β unit can be expressed from (67) and (68) as

(69)Δ=12{(2kYY+4mkXY{1cos(2πn)})|(hXhY+kYY)|(hXhYkYY)2+8kXY2(1+m)2{1cos(2πn)}}

for the even n and

(70)Δ=12{2kYY(hXhYkYY)2+8kXY2(1+m)2{1cos(πn)}(hXhY+kYY)2+8kXY2(1m)2{1cos(πn)}}

for odd-n, respectively.

Case-II: Cyclacenes with {X,Y} = {B,N}, {C,N}, {N,B}, and {N,C}

For the even-n, the expressions of HOMO and LUMO energies for such cyclacenes are as follows:

(71)λt+(HOMO)=12{(hX+hYkYY4mkXY)+|(hXhY+kYY)|}λt+(LUMO)=12{(hX+hY+kYY4mkXY)|(hXhYkYY)|}}

whereas for the odd-n, the HOMO–LUMO energies are the same as given in (68) and HOMO–LUMO gap is same as given in (70) above.

Thus, the expression of the HOMO–LUMO or band gap (Δ) for such systems with the even-n in the β unit would look like

(72)Δ=12{2kYY|(hXhYkYY)||(hXhY+kYY)|}

Here, we can see that for the {X,Y}-cyclacenes under the approximations of a large number of repeating units (i.e. n → ∞), all the expressions of the HOMO–LUMO or band gap (Δ) given in (69), (70), and (72) reduce to the same expression as shown by (72).

3.5.1 Vertical Absorption Energy Differences Between Singlet–Triplet States for {C,C}-Cyclacenes and Their HOMO–LUMO Gaps (Δ)

The HOMO–LUMO gap of a molecule could be a reasonable factor in determining the singlet–triplet splitting behaviour. This led us to search the correlations of the HOMO–LUMO gaps (Δ) with that of the vertical absorption energy differences ΔEST (=ES1ET1) for {C,C}-cyclacenes calculated [51] from the M06-2X method with the use of the 6-31+G basis set (Table 4). For the odd n, the value ΔEST remains invariant with the HOMO–LUMO gap, whereas for the even n, the plot of ΔEST vs. HOMO–LUMO (Δ) bears an excellent linear relationship – with increase in m, the correlation coefficient improves, and at a certain m (m = 0.3), it reaches its optimum value (R = 0.99987). Such a plot is shown in Figure 3. The equation that ΔEST and Δ obey is as follows:

(73)ΔEST=(0.2384±0.01263)+(1.69472±0.02712)×Δ;Corrln. Coeff.=0.99987
Figure 3: Plot of vertical absorption energy differences, ΔEST\(\Delta{E_{ST}}\) calculated from M06-2X method using 6-31+G∗ basis set vs. HOMO–LUMO (Δ) for n-even {C,C}-cyclacenes with nnn interaction (m = 0.3).
Figure 3:

Plot of vertical absorption energy differences, ΔEST calculated from M06-2X method using 6-31+G basis set vs. HOMO–LUMO (Δ) for n-even {C,C}-cyclacenes with nnn interaction (m = 0.3).

3.6 Density of States (DOS)

The density of states (DOS) [11], [48], [52], [53], [54] of a system (molecular network or solid) gives the idea of the number of states per unit energy range at a given energy level that are supposed to be occupied, and it is an important theoretical quantity used in the description of electronic states of the system concerned. The cyclic dimer (CD) is a dimer obtained from a system with repeating monomers under cyclic boundary conditions, and its importance in the routine derivation of the DOS is now well recognized [11], [53], [54], [55]. From the HMO method, it has been found that almost all molecular orbital energies of the CD of a periodic polyhex network coincide [11], [53] with the singular points in its DOS, which has been reconfirmed by the extensive variable-β,γ version of PPP MO-type calculations. Thus, the secrecy of the electronic properties of the infinitely large periodic polyhex networks is shown to be hidden in their respective CD graphs that are called either polyomino or kegome graphs [11].

Figure 4: {X,Y}-cyclacene graph (G) showing next-nearest neighbour (nnn) interaction in one ring and butadiene graph (portion with bold edges) as repeating unit, repeating graph (GR), condensed graph (GC) with edge and vertex weights under action of n-fold rotational symmetry, cyclic dimmer (GCD), and its symmetry plane fragments.
Figure 4:

{X,Y}-cyclacene graph (G) showing next-nearest neighbour (nnn) interaction in one ring and butadiene graph (portion with bold edges) as repeating unit, repeating graph (GR), condensed graph (GC) with edge and vertex weights under action of n-fold rotational symmetry, cyclic dimmer (GCD), and its symmetry plane fragments.

The CD graph (GCD) of the {X,Y}-cyclacene is shown in Figure 4 whose vertex weights are shown on the loops, the edge weights on the edges marked as a broken line given below the graph GCD in Figure 4, the edge weights on the two edges (marked as solid lines) that are parallel but perpendicular to the bottom line of the page are of weight kYY each, whereas all the solid edges are of weight kXY each. The graph GCD possesses a plane of symmetry and is factored into two graphs GCD+ and GCD, and the edge and vertex weights of the fragmented graphs are obtained from the symmetry plane consideration. The vertex weights are given on the respective vertices, whereas the edge weights are shown separately above and below the factored graphs against the respective edges to avoid complexity in the figure. Both the graphs GCD+ and GCD further factorize under the action of the symmetry planes. The graph GCD+ factorizes into GCD++ and GCD+ while GCD+ factorizes into GCD+ and GCD. An illustration of the symmetry factorization is given in Figure 4 along with the vertex and edge weights of the factored graphs. The eigenvalues of the cyclic dimer (GCD) are obtained by solving four two-degree equations resulting from four 2 × 2 determinants corresponding to the four factorized graphs GCD++, GCD+, GCD+, and GCD as follows.

(74)λ(GCD++)=hX+hY+kYY+4mkXY2±12(hXhYkYY)2+16kXY2(1+m)2λ(GCD+)=hX+hY+kYY4mkXY2±12(hXhYkYY)2+16m2kXY2λ(GCD+)=hX+hYkYY+4mkXY2±12(hXhY+kYY)2+16(1m)2kXY2λ(GCD)=hX+hYkYY4mkXY2±12(hXhY+kYY)2+16m2kXY2}

The HOMO–LUMO gap or band gap (Δ) of the cyclic dimer in the β unit can be obtained from (74) as

(75)Δ=λ(GCD+)λ(GCD)+=12{2kYY(hXhYkYY)2+16m2kXY2(hXhY+kYY)2+16m2kXY2}

As we see from (72) and (75) above, in presence of the nnn interactions, the expression of the density of states obtained from the HOMO–LUMO or band gap (Δ) under the approximations of n → ∞ differs from that obtained from the CD. However, in absence of the nnn interaction (i.e. for m = 0), both the approaches end up with the same expression as given in (72).

4 Conclusion

An {X,Y}-cyclacene graph with varying X (=C, N, B) and Y (=C, N, B) as well as a number of hexagonal rings represents a variety of cyclacenes. The HMO quantities, viz., electron and charge densities, bond orders, free valences, total π-electron energies, HOMO–LUMO or band gaps, etc., under the nnn approximations for such systems were expressed in compact analytical forms that are very much useful in calculating the values of the respective quantities in a very easy and convenient manner. Although the HMO quantities for the {X,X}-cyclacenes were calculated and are shown in Table 2, the said quantities for the {X,Y}-cyclacenes can also be calculated with the use of the analytical expressions given in (26), (27), (30), (31), (39), (42), (47), (50), (55), (56), (60), (61), and (64). Facile computer programs can be made with the analytical expressions for making the calculations further simple and very less time consuming. Furthermore, the HMO quantities are very useful in judging several properties of the concerned molecule, for example, the relative charge density (electron density) on an atom indicates the vulnerability on its position for an electrophile (nucleophile), the bond order of a bond dictates its relative strength, the free valence of a atom measures how much its position is susceptible to an electrophilic attack compared to other atomic positions, the total π-electronic energy give the relative stability, the HOMO–LUMO or band gap dictates the stability/reactivity, π-electron transitions, conducting properties of the concerned cyclacenes, etc. Such nnn interaction approximations actually release, to some extent, the strain that results due to the geometrical structures of such cyclacenes. The vertical absorption energy difference for singlet–triplet states bears excellent linear correlation with the HOMO–LUMO gaps for a certain m value (m = 0.3) in case of an even n, but for an odd n, such energy difference remains invariant. The eigenvalues of the cyclic dimer of the cyclacene represents the singular point of the density of states, which is an important quantity for exploring the electronic behaviour of the material. Thus, considering all these, the present study has some important relevance.

Acknowledgement

One of the authors, Mr. Swapnadeep Mondal, thanks the Council of Scientific and Industrial Research (CSIR), New Delhi, for financial assistance under the Junior Research Fellowship. Authors are thankful to the learned Reviewers for their valuable suggestions in improving the article.

References

[1] S. Karmakar, S. Mondal, and B. Mandal, Mol. Phys. 113, 719 (2015).10.1080/00268976.2014.971898Search in Google Scholar

[2] S. Karmakar and B. Mandal, Polycyl. Aromat. Comp. DOI: https://doi.org/10.1080/10406638.2017.1293699.10.1080/10406638.2017.1293699Search in Google Scholar

[3] J. R. Dias, Mol. Phys. 116, 423 (2018).10.1080/00268976.2017.1395918Search in Google Scholar

[4] L. Türker and S. Gümüs, J. Mol. Struct (Theochem) 685, 1 (2015).10.1016/j.theochem.2004.04.021Search in Google Scholar

[5] K. Tahara and Y. Tobe, Chem. Rev. 106, 5279 (2006).10.1021/cr050556aSearch in Google Scholar

[6] D. Eisenberg, R. Shenhar, and M. Rabinovitz, Chem. Soc. Rev. 39, 2879 (2010).10.1039/b904087kSearch in Google Scholar

[7] S. Iijima, Nature 354, 56 (1991).10.1038/354056a0Search in Google Scholar

[8] E. Heilbronner, Helv. Chim. Acta 37, 921 (1954).10.1002/hlca.19540370336Search in Google Scholar

[9] E. Nakamura, K. Tahara, Y. Matsuo, and M. Sawamura, J. Am. Chem. Soc. 15, 2834 (2003).10.1021/ja029915zSearch in Google Scholar

[10] H. S. Choi and K. S. Kim, Angew. Chem. Int. Edition 38, 2256 (1999).10.1002/(SICI)1521-3773(19990802)38:15<2256::AID-ANIE2256>3.0.CO;2-BSearch in Google Scholar

[11] H. Hosoya, H. Kumazaki, K. Chida, M. Ohuchi, and Y. Gao, Pure Appl. Chem. 62, 445 (1990).10.1351/pac199062030445Search in Google Scholar

[12] N. Bej, D. C. Mukherjee, and A. K. Mukherjee, Mol. Phys. 101, 865 (2003).10.1080/0026897021000054853Search in Google Scholar

[13] P. Ghosh, D. J. Klein, and B. Mandal, Mol. Phys. 112, 2093 (2014).10.1080/00268976.2014.886737Search in Google Scholar

[14] L. Türker, Polycyl. Aromat. Comp. 4, 191 (1994).10.1080/10406639408014703Search in Google Scholar

[15] L. Türker, Polycyl. Aromat. Comp. 4, 231 (1995).10.1080/10406639508009621Search in Google Scholar

[16] L. Türker, Polycyl. Aromat. Comp. 8, 67 (1996).10.1080/10406639608048335Search in Google Scholar

[17] I. Gutman, Polycyl. Aromat. Comp. 8, 251 (1996).10.1080/10406639608048352Search in Google Scholar

[18] B. Mandal, Intl. J. Quantum Chem. 103, 140 (2005).10.1002/qua.20503Search in Google Scholar

[19] B. Mandal, J. Mol. Struct. (THEOCHEM) 757, 99 (2005).10.1016/j.theochem.2005.08.011Search in Google Scholar

[20] B. Mandal and D. J. Klein, Bull. Chem. Soc. Jpn. 87, 491 (2014).10.1246/bcsj.20130254Search in Google Scholar

[21] D. J. Klein and X. Liu, J. Math Chem. 11, 199 (1992).10.1007/BF01164204Search in Google Scholar

[22] H. A. Wegner, Angew. Chem. Int. Ed. 56, 10995 (2017).10.1002/anie.201705970Search in Google Scholar PubMed

[23] G. Povie, Y. Segawa, T. Nishihara, Y. Miyauchi, and K. Itami, Science 356, 172 (2017).10.1126/science.aam8158Search in Google Scholar PubMed

[24] A. Pérez-Guardiola, M. E. Sandoval-Salinas, D. Casanova, E. San-Fabián, A. J. Pérez-Jiménez, et al., Phys. Chem. Chem. Phys. 20, 7112 (2018).10.1039/C8CP00135ASearch in Google Scholar

[25] S. Battaglia, N. Faginas-Lago, D. Andrae, S. Evangelisti, and T. Leininger, J. Phys. Chem. A 121, 3746 (2017).10.1021/acs.jpca.7b00123Search in Google Scholar PubMed

[26] S. Battaglia, H.-A. Le, G. L. Bendazzoli, N. Faginas-Lago, T. Leininger, et al., Int. J. Quantum Chem. 118, e25569 (2018).10.1002/qua.25569Search in Google Scholar

[27] S. C. Kornmayer, B. Esser, and R. Gleiter, Org. Lett. 11, 725 (2009).10.1021/ol8028115Search in Google Scholar PubMed

[28] D. J. Klein and B. Mandal, MATCH Commun. Math. Comput. Chem. 74, 247 (2015).Search in Google Scholar

[29] N. Trinajstić, Chemical Graph Theory, CRC Press, Boca Raton, FL 1992.Search in Google Scholar

[30] J. R. Dias, Molecular Orbital Calculations Using Chemical Graph Theory, Springer-Verlag, Berlin 1993.10.1007/978-3-642-77894-0Search in Google Scholar

[31] I. Gutman and O. E. Polansky, Mathematical Concepts in Organic Chemistry, Springer-Verlag, Berlin 1986.10.1007/978-3-642-70982-1Search in Google Scholar

[32] F. A. Van-Catledge, J. Org. Chem. 45, 4801 (1980).10.1021/jo01311a060Search in Google Scholar

[33] F. A. Cotton, Chemical Applications of Group Theory, Wiley India Pvt. Ltd., New Delhi 2015.Search in Google Scholar

[34] M. J. S. Dewar, The Molecular Orbital Theory of Organic Chemistry, McGraw-Hill Book Company, New York 1969.Search in Google Scholar

[35] A. Streitwieser Jr., Molecular Orbital Theory for Organic Chemists, John Wiley & Sons, Inc., New York 1961.Search in Google Scholar

[36] T. Ghosh, S. Mondal, S. Karmakar, and B. Mandal, Mol. Phys. 114, 3307 (2016).10.1080/00268976.2016.1229872Search in Google Scholar

[37] C. A. Coulson, Proc. R. Soc. Lond. A 169, 413 (1939).10.1098/rspa.1939.0006Search in Google Scholar

[38] W. C. Herndon, J. Am. Chem. Soc. 96, 7605 (1974).10.1021/ja00832a001Search in Google Scholar

[39] M. Randić, D. Plavšića, and N. Trinajstić, Int. J. Quantum Chem. 37, 437 (1990).10.1002/qua.560370414Search in Google Scholar

[40] B. Mandal, M. Banerjee, and A. K. Mukherjee, Int. J. Quantum Chem. 99, 119 (2004).10.1002/qua.20120Search in Google Scholar

[41] C. A. Coulson, Discuss. Faraday Soc. 2, 9 (1947).10.1039/df9470200009Search in Google Scholar

[42] F. H. Burkitt, C. A. Coulson, and H. C. Longuet-Higgins, Trans. Faraday Soc. 47, 553 (1951).10.1039/tf9514700553Search in Google Scholar

[43] L. B. Kier and L. H. Hall, J. Chem. Inf. Comput. Sci. 37, 548 (1997).10.1021/ci970002bSearch in Google Scholar

[44] H. Villar, A. L. Capparelli, and A. Spina, Croat. Chem. Acta 59, 383 (1986).Search in Google Scholar

[45] O. B. Tomilin and I. V. Stankevich, Bull. Acad. Sci. USSR, Div. Chem. Sci. 23, 1547 (1974).10.1007/BF00929678Search in Google Scholar

[46] J. N. Herak and N. Trinajstić, Theoret. Chim. Acta 9, 333 (1968).10.1007/BF00529927Search in Google Scholar

[47] C. K. Ingold, Structure and Mechanism in Organic Chemistry, Cornell University Press, New York 1953.Search in Google Scholar

[48] B. Mandal, Bull. Chem. Soc. Jpn. 81, 956 (2008).10.1246/bcsj.81.956Search in Google Scholar

[49] P. Ghosh, S. Karmakar, and B. Mandal, Chem. Phys. Lett. 594, 41 (2014).10.1016/j.cplett.2014.01.020Search in Google Scholar

[50] J. R. Dias, Mol. Phys. 115, 771 (2017).10.1080/00268976.2017.1284358Search in Google Scholar

[51] A. Tiwari, P. K. Iyer, V. Kumar, and H. Swart (eds.), Advanced Magnetic and Optical Materials, Scrivener Publishing LLC, Beverly, Massachusetts 2017, pp. 165–184.10.1002/9781119241966Search in Google Scholar

[52] J. K. Burdett and S. Lee, J. Am. Chem. Soc. 107, 3050 (1985).10.1021/ja00297a010Search in Google Scholar

[53] J. Burdett, S. Lee, and W. C. Sha, Croat. Chem. Acta 57, 1193 (1984).Search in Google Scholar

[54] H. Hosoya, M. Aida, R. Kumagai, and K. Watanabe, J. Comput. Chem. 8, 358 (1987).10.1002/jcc.540080412Search in Google Scholar

[55] U. Norinder and O. Wennerstrom, Tetrahedron Lett. 25, 1397 (1984).10.1016/S0040-4039(01)80168-4Search in Google Scholar

Received: 2018-10-29
Accepted: 2019-01-15
Published Online: 2019-02-12
Published in Print: 2019-06-26

©2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 28.3.2024 from https://www.degruyter.com/document/doi/10.1515/zna-2018-0488/html
Scroll to top button