Acessibilidade / Reportar erro

RING CURRENT IN ANTHRACENE AND PHENANTHRENE: CORRECTION TO HÜCKEL PARAMETERS

Abstract

Excessive degeneracy in the ground state of the π-electron energy levels of anthracene was removed using the Hückel method by correcting the Coulomb integral of the four central carbon atoms. A further correction to the resonance integral was proposed based on the ring-current model, which describes the π-ring current flow along the molecule’s perimeter, by introducing a new parameter for the bridge carbon atoms expressed as a fraction of the resonance energy. The solution to the Hamiltonian was obtained based on the symmetry group theory, which provides the advantage of solving the determinant matrix or secular equations in a particular irreducible representation with a relatively small matrix dimension. The results were further analyzed to determine the bond order and bond length. The values of the harmonic oscillator model of aromaticity, GEO, and EN indices of aromaticity and their ratio for the central and outer rings of benzene were used to evaluate the validity of the correction. Applying the same method to phenanthrene as a topological analog of anthracene allows for an interesting comparison of the stabilities of the two molecules.

Keywords:
Hückel methods; anthracene; phenanthrene; ring current model; harmonic oscillator model of aromaticity.


INTRODUCTION

The Hückel molecular orbital (HMO) theory is the simplest theory for investigating the ground state of a conjugated organic molecule with a planar conformation that has a π-electron system (See Ref. 1 for a complete review). Despite the roughness of its approximations, the validity of this model is undisputable. The HMO theory is indispensable as it provides a basis for understanding conjugated hydrocarbons.22 Kutzelnigg, W.; J. Comput. Chem. 2006, 28, 28. HMO is related to Hückel’s aromaticity (4n + 2) rule, which indicates that cyclic conjugated hydrocarbons with (4n­ + 2)π-electrons are aromatic, while those with (4n)π-electrons are antiaromatic.11 Yates, K.; Hückel molecular orbital theory, Academic Press: New York, 2012. In many cases, a molecule’s aromaticity is related to increased stability, small bond length alternation, and unique magnetic properties associated with the ring current.33 Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12. However, the excessive degeneracy in the energy levels obtained from Hückel theory is often due to a Hamiltonian symmetry, which is higher than that of the molecule,44 Wild, U.; Keller, J.; Gonthard, Hs, H.; Theor. Chim. Acta 1969, 14, 383-395. and a simplified Hamiltonian, which ignores all non-nearby-neighbor interactions.55 Torres, E. M.; Chem. Educator 2006, 16, 1-8.

In the case of benzene (C6H6), the Hückel theory predicts a homogeneous bond order of all nearby neighbor carbon atoms, suggesting a resonance stabilization of benzene.11 Yates, K.; Hückel molecular orbital theory, Academic Press: New York, 2012.,66 McQuarrie, D. A.; Simon, J. D.; Physical chemistry: a molecular approach, Sterling Publishing Company:New York, 1997.,77 Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum (Pengantar Teori dan Contoh Aplikasinya), Karya Putra Darwati Bandung: Bandung, 2012. In this case, the energy level of benzene has two doubly degenerate states (Figure 1(a)) that are related to the high symmetry (point group D6h) of the molecule.11 Yates, K.; Hückel molecular orbital theory, Academic Press: New York, 2012.,66 McQuarrie, D. A.; Simon, J. D.; Physical chemistry: a molecular approach, Sterling Publishing Company:New York, 1997.,77 Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum (Pengantar Teori dan Contoh Aplikasinya), Karya Putra Darwati Bandung: Bandung, 2012. In the case of naphthalene (C10H10) with two fused benzene rings, lowering the symmetry of the molecule (point group D2h) gives rise to non-degenerate energy levels (Figure 1(b)).66 McQuarrie, D. A.; Simon, J. D.; Physical chemistry: a molecular approach, Sterling Publishing Company:New York, 1997.,77 Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum (Pengantar Teori dan Contoh Aplikasinya), Karya Putra Darwati Bandung: Bandung, 2012.,88 Setiawan, D.; Kraka, E.; Cremer, D.; J. Org. Chem. 2016, 81, 9669. Unexpected degeneracy also occurs in anthracene (C14H10), as shown in Figure 1(c), consisting of 3 fused benzenes that belong to the same point group as naphthalene.88 Setiawan, D.; Kraka, E.; Cremer, D.; J. Org. Chem. 2016, 81, 9669. From a theoretical point of view, non-degenerated energy levels were predicted from the irreducible representation of D2h, as confirmed from the semi-empirical self consistent field (SCF) calculation.99 Fukui, K.; Morokuma, K.; Yonezawa, T.; Bull. Chem. Soc. Jpn. 1959, 32, 853.

Figure 1
Molecular structure and the resulting energy level of (a) benzene, (b) naphthalene, and (c) anthracene from Hückel theory

An old-fashioned paradigm of aromaticity states that aromatic molecules support diatropic ring currents, whereas antiaromatic molecules support paratropic ring currents in the presence of an external magnetic field.1010 Steiner, E.; Fowler, P. W.; Chem. Commun. 2001, 2220. Delocalization of the π-electron develop into the assignment of aromaticity as a visual inspection of a map of induced (π) current density, the so-called ring current model (RCM).1111 Steiner, E.; Fowler, P. W.; Int J Quantum Chem 1996, 60, 609. Hence, an external magnetic field applied perpendicular to the ring produces a ring current that induces a counter field. Delocalization of the π-electrons of the molecule affects the long-range contributions of the induced magnetic field. On the other hand, since the σ-electrons are more localized than π-electrons, they generate a short-range magnetic response with diatropic areas along with the bonds and a paratropic response in the molecular center.1212 Heine, T.; Islas, R.; Merino,G.; J. Comput. Chem. 2006, 28, 302. Application of the RCM to benzene, naphthalene, and anthracene shows that those molecules sustain π-ring currents over all the carbon skeleton beyond the molecular periphery.1111 Steiner, E.; Fowler, P. W.; Int J Quantum Chem 1996, 60, 609.,1313 Fowler, P.W.; Steiner, E.; Havenith, R.W.A.; Jenneskens, L.W.; Magn. Reson. Chem. 2004, 42, S68.

14 Cuesta, I. G.; De Merás, A. S.; Pelloni, S.; Lazzeretti, P.; J. Comput. Chem. 2009, 30, 551.

15 Steiner, E.; Fowler, P.W.; Havenith, R. W. A.; J. Phys. Chem. A 2002, 106, 7048.
-1616 Anusooya, Y.; Chakrabarti, A.; Pati, S. K.; Ramasesha, S.; Int. J. Quantum Chem. 1998, 70, 503. Recently, the existence of circular currents in the macroscopic ring structures of benzene, naphthalene, and anthracene has been emulated experimentally by a network of macroscopic microwave resonators.1717 Stegmann, T.; Villafañe, J. A. F.; Ortiz, Y. P.; Deffner, M.; Herrmann, C.; Kuhl, U.; Mortessagne, F.; Leyvraz, F.; Seligman, T. H.; Phys. Rev. B 2020, 102, 075405.

In this study, a correction to the Hückel theory was applied to anthracene by assigning a different Coulomb integral parameter to the four central carbon atoms because of their particular environment,1818 Pritchard, H. O.; Sumner, F. H.; Trans. Faraday Soc. 1955, 51, 457. and different resonance integrals based on a phenomenological description of the π-ring current in a magnetic field normal to the molecular plane.1111 Steiner, E.; Fowler, P. W.; Int J Quantum Chem 1996, 60, 609.,1414 Cuesta, I. G.; De Merás, A. S.; Pelloni, S.; Lazzeretti, P.; J. Comput. Chem. 2009, 30, 551.,1515 Steiner, E.; Fowler, P.W.; Havenith, R. W. A.; J. Phys. Chem. A 2002, 106, 7048. Because of the high symmetry of the molecule, the solution to the Schrödinger equation will be obtained by employing the group theory. The resulting linear combination coefficients of the molecular orbitals were used to calculate the bond order and bond length. The validity of the results was examined using the harmonic oscillator model of aromaticity (HOMA) index,33 Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12.,1919 Kruszewski, J.; Krygowski, T. M.; Tetrahedron Lett. 1972, 13, 3839.

20 Cyrański, M. K.; Krygowski, T. M.; Tetrahedron, 1999, 55, 6205.
-2121 Krygowski, T. M.; Cyraliski, M.; Tetrahedron 1996, 52, 10255. which is widely accepted for measuring aromaticity based on the structure of the molecule.2222 Dobrowolski, J.Cz.; ACS Omega 2019, 4, 18699. Compared to the magnetic-based nucleus-independent chemical shift (NICS) index reported by Schleyer et al.2323 Schleyer, P. von R.; Maerker, C.; Dransfeld, A.; Jiao, H.; van Eikema Hommes, N. J. R.; J. Am. Chem. Soc. 1996, 118, 6317. and the electronically based para-delocalization index (PDI),2424 Poater, J.; Fradera, X.; Duran, M.; Sola, M.; Chem. Eur. J. 2003, 9, 400. the HOMA is much simpler, more successful, and more widely used.1919 Kruszewski, J.; Krygowski, T. M.; Tetrahedron Lett. 1972, 13, 3839.,2525 Kwapisz, J. H.; Stolarczyk, L. Z.; Struct. Chem. 2021, 32, 1393. We also applied the same method to phenanthrene, a topological analog of anthracene, albeit with different ring currents. Hence, theoretical calculations show that for anthracene, the ring currents are mainly localized on the central benzene ring, whereas for phenanthrene, they are more intense in the outer benzene ring.1616 Anusooya, Y.; Chakrabarti, A.; Pati, S. K.; Ramasesha, S.; Int. J. Quantum Chem. 1998, 70, 503.,2626 Ligabue, A.; Pincelli, U.; Lazzeretti, P.; Zanasi, R.; J. Am. Chem. Soc. 1999, 121, 5513. We compared the stability of these two molecules based on the molecular orbital energy diagram. The calculation was performed using MATLAB® software to solve the reduced Hamiltonian based on an irreducible representation (IR) and visualized using Gnuplot©. The results were then compared to those of previous advanced quantum mechanics calculations and experimental studies.

Hückel molecular orbital theory

The Hückel molecular orbital theory uses a linear combination of the atomic orbital (LCAO) method to solve the Hamiltonian of the π-electron system11 Yates, K.; Hückel molecular orbital theory, Academic Press: New York, 2012.

(1)H^ψ=Eψ

with

(2)ψn=i=1ncn,ipi

where cn,i is the coefficient of linear combination, and pi is the 2pz atomic orbital of the ith order of the C atom, and n is the total number of carbon atoms that form the molecule. In the Hückel method, the Hamiltonian matrix element is approximated as follows:

(i) All the diagonal elements of the Hamiltonian matrix are assumed to have the same value, namely, the Coulomb integral (α),

(3)Hij=α,i=j

(ii) The nearest neighbor is taken as the resonance integral (β):

(4)Hij=β,ij=±1

In addition, the overlap integral is diagonal,

(5)Sij=δij

The resulting coefficient of linear combination can be further used to calculate the bond order, as defined by

(6)ρij=vnncnicnj

where νn is the occupation number of π electrons in a particular orbital of the molecule.

The delocalization energy of a molecule is the difference between the total energy of the π electron system and the simple isolated system,

(7)Edeloc =Etotal ,πEisolated

For anthracene and phenanthrene, the isolated system consists of 7 ethylene molecules, and the total energy of one ethylene molecule is 2(α + β). In addition, the HOMO-LUMO energy gap is defined as:

(8)Egap =ELUMO EHOMO
where ELUMO and EHOMO are the energy of the lowest unoccupied molecular orbital (LUMO) and highest occupied molecular orbital (HOMO), respectively.

Symmetry group theory of anthracene and phenanthrene

The anthracene molecule is shown in Figure 2, along with the numbering of carbon atoms. Each carbon atom donates one pz atomic orbital as the basis for forming a molecular orbital.

Figure 2
C-atom numbering of Anthracene (a) and Phenanthrene (b)

The symmetry group of the anthracene is D2h88 Setiawan, D.; Kraka, E.; Cremer, D.; J. Org. Chem. 2016, 81, 9669. with a character table shown in Table 1.77 Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum (Pengantar Teori dan Contoh Aplikasinya), Karya Putra Darwati Bandung: Bandung, 2012. This group contains eight elements: the identity element (I), three rotations through angle π about the coordinate axes (C2z, C2y, C2x), the inversion (i), and three reflections (σxy, σxz, σyz). As we see in Table 1, all the irreducible representations of D2h are one-dimensional; hence, no degeneracy should be expected.

Table 1
Character table for D2h point group

On the other hand, the symmetry group of the phenanthrene is C2v88 Setiawan, D.; Kraka, E.; Cremer, D.; J. Org. Chem. 2016, 81, 9669., which contains four elements: the identity element (I), rotation through angle π about z the coordinate axes, and two vertical plane reflections (σxz, σyz). The corresponding character tables are listed in Table 2.77 Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum (Pengantar Teori dan Contoh Aplikasinya), Karya Putra Darwati Bandung: Bandung, 2012.,2727 Atkins, P.; Friedman, R.; Molecular Quantum Mechanics, 4th ed., Oxford University Press: Oxford, 2005.

Table 2
Character table for C2v point group

When a symmetry transformation R is applied to the orbital (Figure 2), we obtain

(9)Rpz(i)=jΓ(R)ijpz(j)
where Γ(R)ij is the ij element of the matrix that represents the symmetry element R. Under these symmetry operations, the 14-pz orbital transforms accordingly. The characteristics of this matrix of anthracene are:

D2h I C2z C2y C2x i σxy σxz σyz Γ 14 0 -2 0 0 -14 0 2

while the corresponding characteristics of phenanthrene are:

C2v I C2 σv(xz) σv(yz) Γ 14 0 14 0

Symmetry adapted linear combinations (SALCs) are formed using the projection operator,66 McQuarrie, D. A.; Simon, J. D.; Physical chemistry: a molecular approach, Sterling Publishing Company:New York, 1997.,77 Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum (Pengantar Teori dan Contoh Aplikasinya), Karya Putra Darwati Bandung: Bandung, 2012.,2727 Atkins, P.; Friedman, R.; Molecular Quantum Mechanics, 4th ed., Oxford University Press: Oxford, 2005.

(10)Pμ=ΣRχμ(R)OR

The results of normalized SALCs for anthracene are shown in Table 3. These imply a reduction of a 14×14 Hamiltonian matrix according to the IR representation of D2h, with doubly 3×3 matrices for B2g and Au IRs, and doubly 4×4 matrices for B3g and B1u IRs. We note that for this particular molecule, there is no SALC for Ag, B1g, B2u, and B3u IRs.

Table 3
The resulted SALC of anthracene (p refers to pz-orbital)

The representation in Table 2 can be reduced into the irreducible ones, namely

Γ=3B2g+4B3g+3Au+4B1u

The corresponding IRs and SALCs for phenanthrene are shown in Table 4.

Table 4
The resulted SALC of phenanthrene (p refers to pz-orbital)

The information shown above justify the reduction of a 14×14 Hamiltonian matrix to become a doubly 7×7 matrix for A1 and B1 IRs, according to

Γ=7A1+7B1

Bond order and bond length correlation

Pritchard and Sumner1818 Pritchard, H. O.; Sumner, F. H.; Trans. Faraday Soc. 1955, 51, 457. introduced a formula for bond length as a function of bond order for benzene, naphthalene, and anthracene as

(11)R=ssd1+k(1ρijρij)
where s = 1.54 Å and d = 1.33 Å are the lengths of single and double bonds between sp22 Kutzelnigg, W.; J. Comput. Chem. 2006, 28, 28. carbon centers, respectively, and k = 0.765 is a parameter that accounts for the larger force constant of the double bond.2828 Mataga, N.; Nishimoto, K.; Mataga, S.; Bull. Chem. Soc. Jpn. 1959, 32, 395.

Ring current model

The delocalized property of π-electrons determines the magnetic properties of the molecule. As illustrated in Figure 3, an external magnetic field (Bext) perpendicular to the ring’s molecule produces a ring current for aromatic molecules. According to Biot-Savart’s law, this ring current induces a counter magnetic field, (Bind) , with a circular orientation across the molecule’s plane.1212 Heine, T.; Islas, R.; Merino,G.; J. Comput. Chem. 2006, 28, 302.-1313 Fowler, P.W.; Steiner, E.; Havenith, R.W.A.; Jenneskens, L.W.; Magn. Reson. Chem. 2004, 42, S68.

Figure 3
Illustration of the ring current model of an aromatic molecule

Inside the molecule, Bext is in the opposite direction as Bind, while outside of the molecule Bext is in the same direction as Bind. Thus, the total magnetic field outside the ring is more significant than that inside the ring. For antiaromatic molecules, the reverse situation occurs. Hence, aromaticity is commonly associated with a diamagnetic ring current or diatropic ring current, whereas antiaromaticity is associated with a paratropic ring current.

Both diatropic and paratropic currents are experimentally detectable by their contribution to chemical shifts. Theoretically, the assignment of aromaticity is investigated by visualizing the induced π-current density map through perturbation theory.1010 Steiner, E.; Fowler, P. W.; Chem. Commun. 2001, 2220.

11 Steiner, E.; Fowler, P. W.; Int J Quantum Chem 1996, 60, 609.

12 Heine, T.; Islas, R.; Merino,G.; J. Comput. Chem. 2006, 28, 302.

13 Fowler, P.W.; Steiner, E.; Havenith, R.W.A.; Jenneskens, L.W.; Magn. Reson. Chem. 2004, 42, S68.

14 Cuesta, I. G.; De Merás, A. S.; Pelloni, S.; Lazzeretti, P.; J. Comput. Chem. 2009, 30, 551.
-1515 Steiner, E.; Fowler, P.W.; Havenith, R. W. A.; J. Phys. Chem. A 2002, 106, 7048. The results for polycyclic aromatic hydrocarbon (PAH) molecules such as benzene, naphthalene, and anthracene show the flow of π-ring currents over the molecular perimeter, with weak diamagnetic π-vortices around the center of each ring and the midpoint of the C-C bond connecting them.1111 Steiner, E.; Fowler, P. W.; Int J Quantum Chem 1996, 60, 609.,1313 Fowler, P.W.; Steiner, E.; Havenith, R.W.A.; Jenneskens, L.W.; Magn. Reson. Chem. 2004, 42, S68.-1414 Cuesta, I. G.; De Merás, A. S.; Pelloni, S.; Lazzeretti, P.; J. Comput. Chem. 2009, 30, 551. This result agrees with the circuit model proposed by Pauling2929 Linus Pauling, J. Chem. Phys. 1936, 4, 673. to account for the diamagnetic anisotropy of some aromatic molecules.

Harmonic oscillator model of aromaticity (HOMA)

Aromaticity has been the central concept in organic chemistry for a long time.1919 Kruszewski, J.; Krygowski, T. M.; Tetrahedron Lett. 1972, 13, 3839. One measure of aromaticity is structurally based-HOMA33 Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12.,1919 Kruszewski, J.; Krygowski, T. M.; Tetrahedron Lett. 1972, 13, 3839. that was originally defined as

(12)HOMA=1αni(RoptRi)2
where Ri is the ith bond length in the analyzed molecule’s ring, n is the number of C-C bonds in the ring of the molecule, and α is a normalization factor making the unitless HOMA index equal to 1 for perfectly aromatic benzene and 0 for a perfectly alternating hypothetical Kekulé cyclohexatriene ring. Optimal bond length, Ropt, is the proper bond length for fully aromatic molecules and denotes the bond length for which expansion to the single bond and compression to the double bond requires the same energy.33 Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12. For carbocyclic molecules, the HOMA index defined in Eq. (12) can also be expressed as:33 Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12.,2020 Cyrański, M. K.; Krygowski, T. M.; Tetrahedron, 1999, 55, 6205.,2121 Krygowski, T. M.; Cyraliski, M.; Tetrahedron 1996, 52, 10255.
(13a)HOMA=1GEOEN

with

(13b)GEO=αni(RiRav)2
(13c)EN=α(RavRopt)2

and

(14)Rav=1niRi
where Rav is the average bond length for a ring of the molecule. Thus, two contributions to the HOMA index have been considered to mean the decrease in aromaticity of the π-electron system arising from independent mechanisms, i.e., (i) GEO as the degree of bond length alternation; and (ii) EN as the extension of the bonds over the mean bond length.33 Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12.,2020 Cyrański, M. K.; Krygowski, T. M.; Tetrahedron, 1999, 55, 6205.,2121 Krygowski, T. M.; Cyraliski, M.; Tetrahedron 1996, 52, 10255. For the C-C bond, using 1,3 butadiene as a reference (with Rs = 1.467 Å and Rd = 1.349 Å as reference bond lengths) one can obtain Ropt = 1.388 Å and α = 257.7 Å-2.33 Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12.,2020 Cyrański, M. K.; Krygowski, T. M.; Tetrahedron, 1999, 55, 6205.,2121 Krygowski, T. M.; Cyraliski, M.; Tetrahedron 1996, 52, 10255.

Correction to Hückel parameters

Correction to the Hückel method was performed because the carbon atoms at positions 11, 12, 13, 14 for anthracene have a different environment compared to those at other positions, and they are more electronegative compared to the others.1818 Pritchard, H. O.; Sumner, F. H.; Trans. Faraday Soc. 1955, 51, 457. Similarly, for phenanthrene, correction was performed for the carbons at positions 2, 5, 10, and 11. In addition, the π-ring current in anthracene and phenanthrene describes the π-current flow through the perimeter of the molecule.1111 Steiner, E.; Fowler, P. W.; Int J Quantum Chem 1996, 60, 609.,1414 Cuesta, I. G.; De Merás, A. S.; Pelloni, S.; Lazzeretti, P.; J. Comput. Chem. 2009, 30, 551.,1515 Steiner, E.; Fowler, P.W.; Havenith, R. W. A.; J. Phys. Chem. A 2002, 106, 7048.,1616 Anusooya, Y.; Chakrabarti, A.; Pati, S. K.; Ramasesha, S.; Int. J. Quantum Chem. 1998, 70, 503.,2626 Ligabue, A.; Pincelli, U.; Lazzeretti, P.; Zanasi, R.; J. Am. Chem. Soc. 1999, 121, 5513. Two parameters were proposed to describe the change in the Coulomb energy of C11, C12, C13, and C14 of anthracene and C2, C5, C10, and C11 of phenanthrene:

(15)α=α+a|β|
as well as the resonance energy between C11-C12 and C13-C14 of anthracene and C2-C11 and C5-C10 of phenanthrene, as
(16)β=γ|β|

In this case, for anthracene, the secular determinant equations for each irreducible representation of the Hückel matrix become:

B2 g:H=|αEβββα+βE0β0α+βE|=0B3 g:H=|αEβ0ββαβE0000αE2ββ02βαβE|=0|A u:H=|αEβββαβE0β0αβE|=0B1 u:H=|αEβ0ββα+βE0000αE2ββ02βα+βE|=0
while for phenanthrene they become:
A1:H=|αβ0β000βαβ000β0βα+β0000β00αβ00000β0β00000βαβ0β000βα+β|=0B1:H=|αβ0β000βαβ000β0βαβ0000β00αβ00000β0β00000βαβ0β000βαβ|=0

The values of a and γ were varied within the range 0 < a < 1.0 and 0.1 < γ < 1.6, respectively. We note that a set value of a = 0 and γ = 1 refers to the original Hückel parameter.

The bond length was calculated based on Eq. (11) to obtain the HOMA, GEO, and EN indices (Eq. (13)), and the ratio between the HOMA, GEO, and EN values between the benzene rings at the outer and center of the molecule. We note that applying Eq. (11) to estimate the bond length of benzene and naphthalene has resulted in a HOMA index of 1.000 and 0.910 for those two molecules, respectively, in good agreement with a previous study.2525 Kwapisz, J. H.; Stolarczyk, L. Z.; Struct. Chem. 2021, 32, 1393. For the HOMA index of anthracene, the results of previous studies show that the benzene ring at the center (B) of anthracene is more aromatic than the outer (A), opposite results to those of phenanthrene.88 Setiawan, D.; Kraka, E.; Cremer, D.; J. Org. Chem. 2016, 81, 9669.,3030 Krygowski, T. M.; J. Chem. Inf. Comput. Sci. 1993, 33, 70.

31 Portella, G.; Poater, J.; Sola, M.; J. Phys. Org. Chem. 1993, 18, 785.

32 Poater, J.; Duran, M.; Solà M.; Front. Chem. 2018, 6; 561.
-3333 Kalescky, R.; Kraka, E.; Cremer, D.; J. Phys. Chem. A 2014, 118, 223.Table 5 shows the HOMA, GEO, and EN values, as well as their ratios for the two inequivalent benzene rings of anthracene and phenanthrene from previous studies.

Table 5
The values of HOMA, GEO, and EN of anthracene and phenanthrene from previous studies, and their ratios for the two inequivalent benzene rings.

Three criteria were used to determine the best values of a and γ, namely, the values of the HOMA, GEO, and EN indices; ratio of HOMA, GEO, and EN for the two non-equivalent benzene rings; and standard deviation of the bond length compared to the corresponding values from ref. 33. Figure 4 illustrates the correction results for the Hückel parameter for a variation in the Coulomb integral (α’) and resonance integral (β’) for the four central carbon atoms of anthracene, expressed in a and γ values. The left figure (i) shows the HOMA, GEO, and EN indices of the outer benzene ring (A) and the central or inner benzene ring (B) of the anthracene; and the right figure (ii) shows the ratio of HOMA, GEO, and EN indices of the two benzene rings, compared with the results of a previous study.3333 Kalescky, R.; Kraka, E.; Cremer, D.; J. Phys. Chem. A 2014, 118, 223.

Figure 4
(i) HOMA, GEO, and EN indices for the outer benzene ring (A) and central benzene ring (B) of the anthracene, and (ii) the ratio of HOMA, GEO, and EN indices of the two benzene rings, for the variation of: (a) Coulomb integral (a) for γ = 1, (b) resonance integral (γ) for a = 1, and (c) resonance integral (γ) for a = 0.6. In figs (ii) the corresponding values from Ref. [33] are also shown for comparison

Upon variation of a, we found a non-monotonous variation in the HOMA indices. A qualitative agreement with previous studies, where HOMA-B > HOMA-A was obtained for a < 0.3. For γ variation with a = 1.0, the condition of HOMA-B > HOMA-A was obtained for γ ≥ 1.3. However, in both cases, the ratio values of HOMA, GEO, and EN for the two non-equivalent benzene rings are relatively far from the values obtained from the experimental results (as most clearly observed for the GEO ratio, except for a = 1 and γ = 1.2). The best result was obtained for a = 0.6 and γ = 1.1 (Figure 4(c)).

Applying the same method to phenanthrene, Figure 5 illustrates the HOMA index and the standard deviation value of bond length upon γ variation for a = 0.6. From this figure, HOMA-A > HOMA-B for all γ values. From the three criteria mentioned above, the best result was obtained for a = 0.6 and γ = 1.2, along with the smallest standard deviation value.

Figure 5
HOMA indices for the outer benzene ring (A) and central benzene ring (B) of the phenanthrene, along with the standard deviation values of bond length compared to the corresponding values from ref. 33.

To further validate the result, the calculated bond lengths for some unique bonds of the two molecules for the best parameters, a = 0.6 and γ = 1.1 for anthracene and a = 0.6 and γ = 1.2 for phenanthrene, were compared to previous theoretical and experimental studies, as shown in Figure 6. From this figure, one can see a good agreement for the bond order and bond length pattern with the results of previous studies. For the best results in the correction of the Hückel parameters, the improvement in the standard deviation value of the bond length compared to the original Hückel method was 4.39% for anthracene and 16.71% for phenanthrene.

Figure 6
Bond lengths (a) anthracene and (b) phenanthrene for the best results in the present work compared with the results from original Hückel parameters and previous studies1818 Pritchard, H. O.; Sumner, F. H.; Trans. Faraday Soc. 1955, 51, 457.,3232 Poater, J.; Duran, M.; Solà M.; Front. Chem. 2018, 6; 561.

33 Kalescky, R.; Kraka, E.; Cremer, D.; J. Phys. Chem. A 2014, 118, 223.

34 Fedorov, I. A.; Zhuravlev, Y. N.; Berveno, V. P.; Phys. Chem. Chem. Phys. 2011, 13, 5679.

35 Kiralj, R.; Ferreira, M. M. C.; J. Chem. Inf. Comput. Sci. 2002, 42, 508.

36 McL Mathieson, A.; Robertson, J. M.; Sinclair, V. C.; Acta Cryst. 1950, 3, 245.

37 Sinclair, V. C.; Robertson, J. M.; McL Mathieson, A.; Acta Cryst. 1950, 3, 251.

38 Kiralj, R.; Kojić-Prodić, B.; Žinić, M.; Alihodžić, S.; Trinajstić, N.; Acta Cryst. 1996, B52, 823.
-3939 Kubba, R. M.; Al-ani, R. I.; Shanshal, M.; Zeitschrift fur Naturforschung A 2005, 60a, 165.

The molecular energy level diagram for the best result to correct the Hückel parameter and the associated molecular orbital obtained from a linear combination of SALC are shown in Figure 7(a) for anthracene and Figure 7(b) for phenanthrene.

Figure 7
Molecular energy level diagram of (a) anthracene and (b) phenanthrene for the best result to a correction of Hückel parameters and the associated molecular orbitals. The black and white circles indicate the positive and negative coefficients of SALC, respectively

From this figure, the non-degenerate energy levels of anthracene are in good agreement with a previous semi-empirical self-consistent field-linear combination of atomic-molecular orbital study.99 Fukui, K.; Morokuma, K.; Yonezawa, T.; Bull. Chem. Soc. Jpn. 1959, 32, 853. The lowest energy level corresponds to the delocalization of π electrons, and the number of nodal lines increases with increasing energy level. The HOMO-LUMO gap energy and the delocalization energy of anthracene calculated from this correction were 0.82β and 8.08β, respectively. The corresponding values for phenanthrene were 1.22β and 8.45β. We note that the original Hückel method resulted in HOMO-LUMO gap energy and delocalization energies of 0.82β and 5.31β for anthracene, while for phenanthrene they were 1.22β and 5.45β, respectively. We note that for the two molecules, the HOMO-LUMO gap energy is almost constant both to a and γ variations, while the delocalization energy increases linearly with an increase in a and γ. Thus, the Hückel parameter correction increased the stability of the molecule, as shown by the increase in the delocalization energy.

As also shown in Figure 6, the HOMO-LUMO gap energy and delocalization energy of phenanthrene is larger than that of anthracene. This result is in good agreement with previous quantum mechanical calculation studies based on Hartree Fock,4040 Howard, S. T.; Krygowski, T. M.; Can. J. Chem. 1997, 75, 1174. density functional theory,4141 Malloci, G.; Cappellini, G.; Mulas, G.; Mattoni, A.; Chem. Phys. 2011, 384, 19. and time-dependent density functional theory methods4141 Malloci, G.; Cappellini, G.; Mulas, G.; Mattoni, A.; Chem. Phys. 2011, 384, 19.. The commonly accepted calculated values are 3.23 - 3.64 eV for anthracene4242 Menon, A.; Dreyer, J. A. H.; Martin, J. W.; Akroyd, J.; Robertson, J.; Kraft, M.; Phys. Chem. Chem. Phys., 2019, 21, 16240.

43 Kukhta, A. V.; Kukhta, I. N.; Kukhta, N. A.; Neyra, O. L.; Meza, E.; J. Phys. B: At. Mol. Opt. Phys. 2008, 41, 205701.
-4444 Perger, W. F.; Chem. Phys. Lett. 2003, 368, 319. and 4.02 - 4.74 eV for phenanthrene.4141 Malloci, G.; Cappellini, G.; Mulas, G.; Mattoni, A.; Chem. Phys. 2011, 384, 19.-4242 Menon, A.; Dreyer, J. A. H.; Martin, J. W.; Akroyd, J.; Robertson, J.; Kraft, M.; Phys. Chem. Chem. Phys., 2019, 21, 16240.,4545 Gumuş, A.; Gumuş, S.; Maced. J. Chem. Chem. Eng. 2017, 36, 243. The experimental optical band gap values based on ultraviolet-visible spectroscopy are 3.27 eV for anthracene and 4.16 eV for phenanthrene.4242 Menon, A.; Dreyer, J. A. H.; Martin, J. W.; Akroyd, J.; Robertson, J.; Kraft, M.; Phys. Chem. Chem. Phys., 2019, 21, 16240. Thus, the results presented in this study are in good agreement with previous studies that show that kinked phenanthrene is more stable than straight anthracene.3232 Poater, J.; Duran, M.; Solà M.; Front. Chem. 2018, 6; 561.,4646 Poater, J.; Visser, R.; Sola, M.; Bickelhaupt, F. M.; J. Org. Chem. 2007, 72, 1134.

CONCLUSIONS

Anthracene with point group D2h shows excessive degeneracy in the energy levels of the π-electron system. Based on the inequivalent carbon atoms and the physical phenomenology of the ring current that flows along the perimeter of the molecule, we proposed a correction to the Hückel parameters. The solution to the Hamiltonian of the π-electron was performed using symmetry group theory to solve the determinant matrix or secular equations in a particular IR. The correction was conducted by introducing a change in the Coulomb energy (α′) for the four central carbon atoms and resonance energy (β′) for C-C bonds along the bridge of the molecule, expressed in terms of a and γ values as a fraction of resonance energy (β). The computational process was performed using MATLAB® software. The resulting linear combination coefficient of the molecular orbital was used to calculate the bond order and bond length, and the results were similar to those of previous studies. The validity of the correction was shown by the values of HOMA, GEO, and EN indices of aromaticity which showed that the central benzene ring is more aromatic than the outer benzene ring. The best result was shown for a = 0.6 and γ = 1.1. Further analysis revealed the non-degenerate energy levels and molecular orbitals obtained as a linear combination of SALCs. We applied the same method to phenanthrene (point group C2v) as the topological analog of anthracene. For this molecule, the outer benzene ring is more aromatic than the central ring, and the best result was shown for a = 0.6 and γ = 1.2. Compared to the original Hückel method, this correction resulted in higher delocalization energy, signifying that the molecule has a higher stability, while the HOMO-LUMO energy gap remains the same. From the HOMO-LUMO energy gap and delocalization energy, phenanthrene is more stable than anthracene, which is in good agreement with previous advanced calculations and experimental studies.

ACKNOWLEDGEMENTS

This study is the output of the P2MI ITB 2022 research scheme. It is dedicated to the late Professor M.O. Tjia for his memorable contributions to conjugated polymers, material physics, and chemistry at the Physics Department, Faculty of Mathematics and Natural Science (FMIPA) of the Institut Teknologi Bandung.

ReferenceS

  • 1
    Yates, K.; Hückel molecular orbital theory, Academic Press: New York, 2012.
  • 2
    Kutzelnigg, W.; J. Comput. Chem. 2006, 28, 28.
  • 3
    Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12.
  • 4
    Wild, U.; Keller, J.; Gonthard, Hs, H.; Theor. Chim. Acta 1969, 14, 383-395.
  • 5
    Torres, E. M.; Chem. Educator 2006, 16, 1-8.
  • 6
    McQuarrie, D. A.; Simon, J. D.; Physical chemistry: a molecular approach, Sterling Publishing Company:New York, 1997.
  • 7
    Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum (Pengantar Teori dan Contoh Aplikasinya), Karya Putra Darwati Bandung: Bandung, 2012.
  • 8
    Setiawan, D.; Kraka, E.; Cremer, D.; J. Org. Chem. 2016, 81, 9669.
  • 9
    Fukui, K.; Morokuma, K.; Yonezawa, T.; Bull. Chem. Soc. Jpn. 1959, 32, 853.
  • 10
    Steiner, E.; Fowler, P. W.; Chem. Commun. 2001, 2220.
  • 11
    Steiner, E.; Fowler, P. W.; Int J Quantum Chem 1996, 60, 609.
  • 12
    Heine, T.; Islas, R.; Merino,G.; J. Comput. Chem. 2006, 28, 302.
  • 13
    Fowler, P.W.; Steiner, E.; Havenith, R.W.A.; Jenneskens, L.W.; Magn. Reson. Chem. 2004, 42, S68.
  • 14
    Cuesta, I. G.; De Merás, A. S.; Pelloni, S.; Lazzeretti, P.; J. Comput. Chem. 2009, 30, 551.
  • 15
    Steiner, E.; Fowler, P.W.; Havenith, R. W. A.; J. Phys. Chem. A 2002, 106, 7048.
  • 16
    Anusooya, Y.; Chakrabarti, A.; Pati, S. K.; Ramasesha, S.; Int. J. Quantum Chem. 1998, 70, 503.
  • 17
    Stegmann, T.; Villafañe, J. A. F.; Ortiz, Y. P.; Deffner, M.; Herrmann, C.; Kuhl, U.; Mortessagne, F.; Leyvraz, F.; Seligman, T. H.; Phys. Rev. B 2020, 102, 075405.
  • 18
    Pritchard, H. O.; Sumner, F. H.; Trans. Faraday Soc. 1955, 51, 457.
  • 19
    Kruszewski, J.; Krygowski, T. M.; Tetrahedron Lett. 1972, 13, 3839.
  • 20
    Cyrański, M. K.; Krygowski, T. M.; Tetrahedron, 1999, 55, 6205.
  • 21
    Krygowski, T. M.; Cyraliski, M.; Tetrahedron 1996, 52, 10255.
  • 22
    Dobrowolski, J.Cz.; ACS Omega 2019, 4, 18699.
  • 23
    Schleyer, P. von R.; Maerker, C.; Dransfeld, A.; Jiao, H.; van Eikema Hommes, N. J. R.; J. Am. Chem. Soc. 1996, 118, 6317.
  • 24
    Poater, J.; Fradera, X.; Duran, M.; Sola, M.; Chem. Eur. J. 2003, 9, 400.
  • 25
    Kwapisz, J. H.; Stolarczyk, L. Z.; Struct. Chem. 2021, 32, 1393.
  • 26
    Ligabue, A.; Pincelli, U.; Lazzeretti, P.; Zanasi, R.; J. Am. Chem. Soc. 1999, 121, 5513.
  • 27
    Atkins, P.; Friedman, R.; Molecular Quantum Mechanics, 4th ed., Oxford University Press: Oxford, 2005.
  • 28
    Mataga, N.; Nishimoto, K.; Mataga, S.; Bull. Chem. Soc. Jpn. 1959, 32, 395.
  • 29
    Linus Pauling, J. Chem. Phys. 1936, 4, 673.
  • 30
    Krygowski, T. M.; J. Chem. Inf. Comput. Sci. 1993, 33, 70.
  • 31
    Portella, G.; Poater, J.; Sola, M.; J. Phys. Org. Chem. 1993, 18, 785.
  • 32
    Poater, J.; Duran, M.; Solà M.; Front. Chem. 2018, 6; 561.
  • 33
    Kalescky, R.; Kraka, E.; Cremer, D.; J. Phys. Chem. A 2014, 118, 223.
  • 34
    Fedorov, I. A.; Zhuravlev, Y. N.; Berveno, V. P.; Phys. Chem. Chem. Phys. 2011, 13, 5679.
  • 35
    Kiralj, R.; Ferreira, M. M. C.; J. Chem. Inf. Comput. Sci. 2002, 42, 508.
  • 36
    McL Mathieson, A.; Robertson, J. M.; Sinclair, V. C.; Acta Cryst. 1950, 3, 245.
  • 37
    Sinclair, V. C.; Robertson, J. M.; McL Mathieson, A.; Acta Cryst 1950, 3, 251.
  • 38
    Kiralj, R.; Kojić-Prodić, B.; Žinić, M.; Alihodžić, S.; Trinajstić, N.; Acta Cryst. 1996, B52, 823.
  • 39
    Kubba, R. M.; Al-ani, R. I.; Shanshal, M.; Zeitschrift fur Naturforschung A 2005, 60a, 165.
  • 40
    Howard, S. T.; Krygowski, T. M.; Can. J. Chem. 1997, 75, 1174.
  • 41
    Malloci, G.; Cappellini, G.; Mulas, G.; Mattoni, A.; Chem. Phys. 2011, 384, 19.
  • 42
    Menon, A.; Dreyer, J. A. H.; Martin, J. W.; Akroyd, J.; Robertson, J.; Kraft, M.; Phys. Chem. Chem. Phys., 2019, 21, 16240.
  • 43
    Kukhta, A. V.; Kukhta, I. N.; Kukhta, N. A.; Neyra, O. L.; Meza, E.; J. Phys. B: At. Mol. Opt. Phys. 2008, 41, 205701.
  • 44
    Perger, W. F.; Chem. Phys. Lett. 2003, 368, 319.
  • 45
    Gumuş, A.; Gumuş, S.; Maced. J. Chem. Chem. Eng. 2017, 36, 243.
  • 46
    Poater, J.; Visser, R.; Sola, M.; Bickelhaupt, F. M.; J. Org. Chem. 2007, 72, 1134.

Publication Dates

  • Publication in this collection
    29 Aug 2022
  • Date of issue
    2022

History

  • Received
    23 Sept 2021
  • Accepted
    16 Dec 2021
  • Published
    10 Mar 2022
Sociedade Brasileira de Química Secretaria Executiva, Av. Prof. Lineu Prestes, 748 - bloco 3 - Superior, 05508-000 São Paulo SP - Brazil, C.P. 26.037 - 05599-970, Tel.: +55 11 3032.2299, Fax: +55 11 3814.3602 - São Paulo - SP - Brazil
E-mail: quimicanova@sbq.org.br