ArticlePDF Available

Aromaticity of aza aromatic molecules: prediction from Hückel theory with modified parameters

Authors:

Abstract and Figures

Hückel theory is a simple and powerful method for predicting the molecular orbital and the energy of conjugated molecules. However, the presence of nitrogen atoms in aza aromatic molecules alters the Coulomb and resonance integrals owing to the difference in electronegativity between nitrogen and carbon atoms. In this study, we focus on acridine and phenazine. Further correction is implemented based on the ring current model, thus revealing the change in resonance integral for the carbon–carbon bond along the bridge of the molecule. The Hamiltonian of the π–electron system in the Hückel method is solved using the HuLiS software. Various geometry-based aromaticity indices are used to obtain the aromaticity indices of the two non-equivalent rings. For further evaluation, the results for bond lengths are used to calculate the associated bond energy. Considering the carbon–hydrogen (CH) bonds, the total molecular energy is compared with the experimental heats of formation for a number of benzenoid hydrocarbons and aza aromatics, in addition to the two studied molecules. Finally, the correlation between the nitrogen atom on the aromaticity index and the ring energy content is evaluated to determine to which extent the Hückel model agrees with previous experimental and advanced computational studies.
Content may be subject to copyright.
Quim. Nova, Vol. XY, No. 00, 1-10, 200_ http://dx.doi.org/10.21577/0100-4042.20230017
*e-mail: im_sutjahja@itb.ac.id
AROMATICITY OF AZA AROMATIC MOLECULES: PREDICTION FROM HÜCKEL THEORY WITH
MODIFIED PARAMETERS
Inge M. Sutjahjaa,*,, Yuanita P. D. Sudarsob, Sho Dhiya ‘Ulhaqa and Erik Bekti Yutomoa
aDepartment of Physics, Faculty of Mathematics and Natural Sciences, Institut Teknologi Bandung, 40132 Bandung, Indonesia
bDepartment of Physics, Faculty of Information Technology and Science, Parahyangan Catholic University, 40141 Bandung,
Indonesia
Recebido em 14/10/2022; aceito em 11/11/2022; publicado na web 27/01/2023
Hückel theory is a simple and powerful method for predicting the molecular orbital and the energy of conjugated molecules. However,
the presence of nitrogen atoms in aza aromatic molecules alters the Coulomb and resonance integrals owing to the difference in
electronegativity between nitrogen and carbon atoms. In this study, we focus on acridine and phenazine. Further correction is
implemented based on the ring current model, thus revealing the change in resonance integral for the carbon–carbon bond along
the bridge of the molecule. The Hamiltonian of the π–electron system in the Hückel method is solved using the HuLiS software.
Various geometry-based aromaticity indices are used to obtain the aromaticity indices of the two non-equivalent rings. For further
evaluation, the results for bond lengths are used to calculate the associated bond energy. Considering the carbon–hydrogen (CH)
bonds, the total molecular energy is compared with the experimental heats of formation for a number of benzenoid hydrocarbons
and aza aromatics, in addition to the two studied molecules. Finally, the correlation between the nitrogen atom on the aromaticity
index and the ring energy content is evaluated to determine to which extent the Hückel model agrees with previous experimental and
advanced computational studies.
Keywords: Hückel theory; aza aromatic molecule; HuLiS software; geometry-based aromaticity; bond energy.
INTRODUCTION
Polycyclic π-conjugated carbon-based molecules have numerous
potential applications in organic electronics; therefore, a detailed
understanding of their fundamental properties is crucial.1 Doping
heteroatoms into carbon-based molecules is an effective strategy to
adjust their physical and chemical properties of the material, thereby
improving their performance in electronic, photonic, optoelectronic,
and spintronic applications.2-11 Nitrogen is the most commonly used
dopant as it has a similar covalent radius as carbon while providing
one extra electron. A fundamental property of heteroatom-doped
carbon-based molecules is their stability with respect to the control
of energy levels, which can be measured by the energy gap (Egap).
Aromaticity is related to stability and Egap, and it has been a
central concept in organic chemistry for over a century.12 Among
the numerous aromaticity indices,13,14 a geometry-based aromaticity
index is the most simple, successful, and widely used to quantitatively
describe the π−electron delocalization in homo- and hetero-
atomic molecules. Starting from the harmonic oscillator model of
aromaticity (HOMA),15-17 the model was developed for the harmonic
oscillator model of electron delocalization (HOMED)18,19 and the
harmonic oscillator model of heterocyclic electron delocalization
(HOMHED).20 These aromaticity indices use bond length data from
molecular geometry, and the uniformity of π–electron distribution
in the molecule is associated with the equalization of bond lengths.
To calculate the aromaticity index, each of the three models uses
some reference molecules or their related hybridizations to measure
the optimal bond length from the single bond, Rs, and the double
bond, Rd.
Hückel molecular orbital (HMO) theory is a well-known, simple
theory, easily understood by undergraduate students, to predict
Egap and the behavior of the π–electron system.21-23 The HMO
theory uses only the topology of the molecule, is independent of
structural parameters, such as bond lengths and bond angles, and
ignores the strain energy. For the heteroatoms, the Coulomb and
resonance integral parameters can be adjusted according to the atom
type and coordination number. These parameterizations are due to
the different core energies of the heteroatom and the change in the
effective electronegativity, compared with carbon, of the remaining
non-bonded p orbitals at the center. Although the parameterizations
are standardized in numerous chemistry and physics textbooks, the
importance of distinguishing the Coulomb parameter of carbon atoms
adjacent to heteroatoms is rarely emphasized. This is so particular
for nitrogen-containing molecules, due to the large electronegativity
of nitrogen.24,25 This parameterization should also depend on the
molecule, which would require an additional parameter to measure
the accuracy of the parameterization.
Pauling26 established a formula relating bond length to bond
order and an empirical rule for the relation between bond order
and bond energy;27 subsequently, Krygowski developed a model to
estimate bond energy from bond length data.28 Summation of the bond
energy over all bonds of a cyclic molecule provides the ring energy
content (REC), and summation over the entire molecule provides
the molecule energy content (MEC).29 Considering the CH bond
energy, one may obtain the total molecular energy and compare it
with the experimental data for the heat of formation.28 Krygowski28
and Cyrański30 reported that the local and global aromaticity index of
the HOMA correlates well with REC and MEC, with a higher mutual
correlation for angular polyacenes compared with linear systems.
However, the analyses so far are limited to benzenoid hydrocarbon
molecules, leaving some questions regarding heteroatom molecules,
including aza aromatic molecules.
This study applied a variation in the Hückel parameters of
the aza aromatic molecules, acridine and phenazine. Previous
theoretical studies have shown that the carbon atom adjacent to
nitrogen should have a different Coulomb parameter owing to
Artigo
Sutjahja et al.
2Quim. Nova
the large electronegativity of nitrogen.24,25 However, the variation
applied is not systematic, and generalization for its dependence on
molecules is unclear.31 A further correction was performed based on
the ring current model, which describes the ow of the π−electron
system along the molecule’s perimeter due to delocalization,32,33 as
in the case of aromatic hydrocarbon molecules.34,35 The latter was
implemented by assigning different resonance integral parameters
for carbon–carbon atoms along the bridge of the molecule.36 For this
purpose, we used the HuLiS software,37 which provides a facility for
the proposed variations. The bond length data were further analyzed
to predict the geometry-based aromaticity index based on the HOMA,
HOMED, and HOMHED models. The results were used to examine
the effect of each modied Hückel parameter on the molecular energy
level and aromaticity of the two inequivalent rings. To estimate the
bond energy from the bond length and its relation to the aromaticity
index, the bond lengths of the single and double bonds were varied
based on the values proposed in the three structure-based aromaticity
models mentioned above.
Hückel molecular orbital theory
Parametrization of the Hamiltonian matrix element according to
Hückel theory is:21-23
(1)
where pi is the 2pz atomic orbital of the ith order of the C atom; α is the
Coulomb integral; and β is the resonance integral of the π-electron
system. In addition, the overlap integral was diagonal.
(2)
For a molecule with a heteroatom, the Coulomb integral (hX)
and resonance integral (kXY) of the π–electron system depends on the
core energy and the change in the effective electronegativity of the
remaining non-bonded p orbitals at the center. In this case, the values
of α and β in Equation (1) can be written as21,22
(3a)
(3b)
Table 1 lists the commonly accepted values of hX and kXY for C
and N atoms.22
According to Hückel theory, the molecular orbital ψ is formed
from atomic orbital pi through a linear combination of atomic orbitals
(LCAO) model, and the coefcients of atomic orbital {cni} can be
further used to obtain the π-electron bond order:
(4)
with νn is the number of π-electrons occupying the corresponding
energy level.
The energy gap is the difference between highest occupied
molecular orbital (HOMO) level and the lowest unoccupied molecular
orbital (LUMO) level,
(5)
In addition, the delocalization energy is the difference between the
total energies of the π–electron system and the simple isolated system.
(6)
The isolated state of acridine (C13H9N) consists of six ethylene
(C2H2) and one methane imine (CH3N), whereas phenazine (C12H8N2)
consists of ve ethylene and two methane imines.38 The energies
of one molecule of ethylene and methane imine are 2(α + β) and
2α+2.61β, respectively.
Bond order and bond length correlation
Pauling established a well-known formula relating the bond length
R(p) to its bond order p as26
(7)
where c is an empirical constant and R(1) is the standard single bond
length. The value of c can be calculated by taking the bond lengths
for the typical single (p = l) and double (p = 2) bonds as follows:
(8)
Using Equation (7), it is also possible to calculate the bond
number, p, for any bond length, R(p), as follows:
(9)
Focusing to Huckel theory that deals with π-electron system,
the bond length can be calculated from the π-electron bond order
following Gordy’s formula:39,40
(10)
where a and b are the modied Gordy values, equal to 7.33 and 2.09
for CC bonding, and 6.52 and 2.03 for CN bonding, respectively;40
and N = 1 + ρij is the total bond order.39
Geometry-based aromaticity index
The geometry-based aromaticity index of the HOMA13-17
is simple, successful, and widely used,41 both for homo- and
heteroatomic systems. Subsequently, the HOMED18,19 and HOMHED
were developed.20 For each of these models, the bond length values for
the single bond, Rs, and the double bond, Rd, were proposed according
to the reference molecules.
In general the HOMA index can be calculated as13-17,42,43
(11)
where n is the number of bonds taken into summation; and Rj,i is the
experimental or computed bond length of the system for a certain
type of bond (j). In Equation (11), the summation is also held for
Table 1. Values of Hückel parametrization22
Element hXkXY
ChC = 0.00 kCC = 1.00
N2* hN = 0.51 kCN = 1.02
N3** hN = 1.37 kCN = 0.89
*Dicoordinated. **Tricoordinated, planar geometry.
Aromaticity of aza aromatic molecules 3Vol. XY, No. 00
all types of bonds that are present in the ring of a molecule, that is,
carbon–carbon (CC), carbon–nitrogen (CN), and nitrogen–nitrogen
(NN) for the aza molecule. The optimal bond lengths Ro and a were
calculated using the following formula.
(12)
(13)
where w is the ratio of the stretching force constants for pure single
(Rs) and double (Rd) bonds, and the common value is equal to 2 for CC
and CX bonds.44 Table 2 presents the Rs, Rd, Ro, α, and c values for CC,
CN, and NN bonds, and the reference molecule for each type of bond.
In Equation (11), it is clear that the decreased aromaticity in the
π–electron system can be described by two different and independent
mechanisms, namely, an increase in the bond length alternation
(GEO) and an extension of the mean bond length (bond elongation)
(EN); the latter is due to a decrease in the resonance energy. These
two dearomatization terms, i.e., geometric GEO and energetic EN
contributions, are calculated by transforming the bond order of CX or
XY bonds. This is derived from Equation (9) considering the virtual
CC bonds according to the Pauli formula (Equation (7)),
(14)
such that
where rav represents the average value of ri.
The HOMED index was proposed by Raczyńska et al.18 in 2010,
primarily for molecules that contain heteroatoms. The HOMED index
is formally the same as that of the HOMA, Equation (11) but differs
for the reference molecules. Herein, quantum-chemical methods were
used to estimate the bond lengths using simple saturated systems for
single bonds and simple unsaturated systems for double bonds. The
optimal bond length, Ro, was chosen from various simple molecules
for which equalization of the bonds occurs. Table 3 presents the
Rs, Rd, and Ro values for the HOMED calculations, along with the
reference molecules.
The α calculation follows the following rules. For molecules
with even numbers of bonds (2i), α can be calculated using Equation
(13). In contrast, for molecules with odd numbers of bonds (2i + 1),
α can be calculated using Equations (15) and (16), each of them for
(i + 1) single bonds and (i) double bonds, and (i) single bonds, and
(i + 1) double bonds.
(15)
(16)
The HOMHED model proposed by Frizzo et al.20 in 2012 is
based on the average experimental (X-ray diffraction and neutron
diffraction) data of several reference molecules. Table 4 presents
the Rs, Rd, Ro, and α values for the HOMHED calculation, along
with the hybridization of the reference molecules for a particular
type of bonding. The HOMHED index was calculated using
Equation (11).
Table 2. Values of Rs, Rd, Ro, α and c for the HOMA index along with reference molecules15,43
Type of bond Rs (Å) Rd (Å) Ro (Å) αcReference molecule
CC 1.467 1.349 1.388 257.7 0.1702 1,3-butadiene, CH2=CH−CH=CH2
CN 1.465 1.269 1.334 93.52 0.2828 Methylamine, H2N−CH3 and methylene imine,
HN=CH2
NN 1.420 1.254 1.309 130.33 0.2395 (CH3)2C=N–N(CH3)2 and H3C–N=N–CH3
Table 3. Values of Rs, Rd, and Ro for the HOMED index along with the reference molecules18,19
Type of bond Rs (Å) Molecule reference Rd (Å) Molecule reference Ro (Å) Reference molecule
CC 1.5300 H3C−CH31.3288 H2C=CH21.3943
CN 1.4658 H3C−NH21.2670 H2C=NH 1.3342
NN 1.4742
1.2348
1.3193
Table 4. Values of Rs, Rd, Ro, and α for the HOMHED index along with the hybridization of the reference molecules20
Type of bond Rs (Å) Rd (Å) Ro (Å) αHybridization of the reference molecules
CC 1.530 1.316 1.387 78.6 Csp3−Csp3; Csp2−Csp2
CN 1.474 1.271 1.339 87.4 C−Nsp3
NN 1.454 1.240 1.311 78.6 Nsp3−Nsp3; C–N=N–C
Sutjahja et al.
4Quim. Nova
Bond energy from bond length
Bond energy is related to geometry-based aromaticity indices
and is estimated from the bond length. For benzenoid hydrocarbon
molecules, Krygowsky et al.28 has developed a model that relates
the bond energy, E(i), calculated directly from the bond length as:
(17)
where R(1) and E(1) are the bond lengths of a single bond and its
associated energy, respectively; R(i) and E(i) are the investigated bond
lengths in the molecule and its related energy, respectively; and α′ is
an empirical constant that can be calculated according to the reference
bond lengths and bond energies that correspond to the single and
double bonds. Generalization of the above formula to aza aromatic
molecules was performed by considering the CN and NN bonds in
addition to CC bonding. The REC can be obtained by summing E(i)
for all bonds in a ring of the molecule, as follows.
(18)
whereas the MEC can be obtained by summing over the molecule.
Considering the carbon and hydrogen bonds in the molecule, one can
obtain the total bond energy of the molecule:
(19)
This value can be compared with the experimental data for the
heat of formation from atoms (HtFfA),28
(20)
where Hf,molecule is the heat of formation of the corresponding
molecule; and HA is the heat of atomization; additionally, summation
(k) is performed for all atoms that make up the molecule.
METHODS
HuLiS software
HuLiS37 is a software package for molecular electronic structure
calculations based on the Hückel method. Figure 1 shows the HuLiS
software interface. The input parameter used to solve the Hamiltonian
system is the molecular drawing. This software allows for a change
in the Coulomb integral of any atom (α) or resonance integral (β) by
clicking on a certain atom or bonding.
The output consists of the energy level and the coefcients of the
linear combination of molecular orbitals. Additionally, this software
can display the electron distribution prole for each energy level.
Parametrization of Hückel parameters
For various N-doped molecules of pyridine, quinoline, iso-
quinoline, and acridine, Longuet-Higgins and Coulson in 194624
proposed the same correction to the Coulomb integrals of nitrogen
and carbon atoms that directly bonded to nitrogen, whereas Dasgupta
in 196425 proposed different parameter values that depend on the
molecular type, as presented in Table 5. In this table, α is the Coulomb
integral for all other carbon atoms and β is the resonance integral of
the CC bonding.
The data presented in Table 5 reveal a signicant difference in the
parameter values, whereas the commonly accepted values are listed
in Table 1, without the need to give the difference for the carbon
atom close to nitrogen.
Herein, we performed a systematic study for the parametrization
of the Hückel parameters of acridine and phenazine, which consist
of the Coulomb integral of carbon adjacent to nitrogen (αCN) and
the resonance integral for CC bonding (βCC) along the bridge of the
molecule, expressed in terms of resonance energy,36
βCC = γ β (21)
The values of αCN are 0, 0.025, 0.050, 0.075, 0.0825, 0.100, and
0.250; and 1.0 γ 1.5 with steps of 0.1.
Bond energy and heat of atomization
To compare the total bond energy and heat of formation from
atoms, Table 6 lists the experimental data for the heat of atomization
Figure 1. HuliS software interface
Table 5. Different Coulomb and resonance parameters of aza aromatic mo-
lecules from previous studies24,25
Molecule αNαCN Reference
Pyridine α + 2β α + 0.25β 24
α + 0.2β α + 0.025β 25
Quinoline α + 0.2β 25
Iso- Quinoline α + 0.2β 25
Acridine α + 2β α + 0.25β 24
α + 0.66β α + 0.0825β 25
Aromaticity of aza aromatic molecules 5Vol. XY, No. 00
of C, H, and N from the gas phase and the bond dissociation energies
of CH, CC, C=C, CN, and C=N.45 Table 7 lists the experimental
data for the enthalpy of formation of the investigated molecules.
RESULTS AND DISCUSSION
The results of Hückel method
The molecular structures of acridine and phenazine are shown in
Figure 2. The direct application of the Hückel method to acridine and
phenazine by applying the Coulomb integral and resonance integral
parameters listed in Table 1 gave the energy level diagram shown in
the above panel of Figure 2. This gure shows that for both molecules,
the energy levels of the HOMO-1 and LUMO+1 states consist of
degenerate states. However, group theory predicted no degeneracy
for the two molecules in the C2v or D2h point groups.51-53
As shown in Figure 2, the aromaticity indices for the two non-
equivalent rings from the previous structural data of acridine54 and
phenazine53 clearly demonstrate that the ring containing the nitrogen
atom (B) is more aromatic than the benzene ring (A).55,56 Furthermore,
for ring-B of the two molecules, the index value of EN is comparable
to GEO. We observe that, particularly for acridine, this variation in
the aromaticity index cannot be qualitatively obtained from the direct
application of the Hückel method.
Correction to Hückel parameters
Figure 3 illustrates the results of the HOMA aromaticity index
of the two non-equivalent molecular rings of (a) acridine and (b)
phenazine for the variation of αCN (i) and βCC (ii). Evidently, different
models of the aromaticity index (HOMED and HOMHED) exhibited
similar proles, although with different absolute values. As shown in
this gure, the standard deviation of the bond length was compared
with the corresponding values from Phillips54 for acridine and
Wozniak et al.53 for phenazine.
Evidently from Figure 3(i), upon varying αCN, the aromaticity
index of the ring-A is smaller than the aromaticity index of ring-B that
contains nitrogen for αCN = 0.1 for acridine, whereas for phenazine, it
occurred for all values of αCN and a minimum in the standard deviation
value of the bond length was observed for αCN = 0.1. However, using
different parameters for carbon atoms close to nitrogen removed the
degeneracy in the energy levels of acridine and phenazine, which is
in good agreement with a previous semi-empirical self-consistent
eld study.24 When varying βCC, a minimum in the standard deviation
value of bond length occurred at βCC = 1.1 for acridine and 1.2
for phenazine. Thus, we concluded that the best modied Hückel
parameters for αCN and βCC are 0.1 and 1.1 for acridine and 0.1 and
1.2 for phenazine, respectively. In addition, set values of aromaticity
indices for acridine are {H = 0.805, G = 0.088, E = 0.107} for ring-A
and {H = 0.837, G=0.044, E = 0.119} for ring-B. The same indices
for phenazine are {H = 0.791, G = 0.104, E = 0.105} for ring-A and
{H = 0.883, G=0.057, E = 0.060} for ring-B. By modifying the
Hückel parameters, it can be seen that the aromaticity indices of
the two molecules qualitatively agree with the experimental results
presented previously.
Table 6. Heat of atomization, bond dissociation energy45
Hf (kcal mol-1)Eb (kcal mol-1)
C(g) 171.2 CH98.3
H(g) 52.1 CC82.6
N(g) 112.9 C=C 144.0
CN72.8
C=N 147.0
Table 7. Enthalpy of formation of molecule
Molecule Formula Hf
(kJ mol-1)
Hf
(kcal mol-1)Reference
Benzene (l) C6H649 11.70 46
Naphthalene (s) C10H878 18.63 46
Anthracene (s) C14H10 127 30.33 46
Phenanthrene (s) C14H10 110 26.27 46
Pyridine (l) C5H5N 100.02 23.89 47
Acridine (g) C13H9N 273.9 65.42 48,49
Phenazine (g) C12H8N2338.3 80.80 50
Figure 2. Molecular structure with atomic numbering and energy level diagram of (a) acridine and (b) phenazine. The numbers in the rings represent the index
values of HOMA (H), GEO (G), and EN (E) calculated from experimental geometric data53,54
Sutjahja et al.
6Quim. Nova
Bond length and molecular orbital description
Figure 4 shows the calculated bond lengths for some unique bonds
of the two molecules for the best parameters compared with previous
theoretical and experimental studies. To study the effect of nitrogen,
we plotted anthracene data from a previous study using the same
method with the best correction to the Hückel parameters.36 Figure
4 shows that the results are in good agreement with previous studies.
For the best corrected Hückel parameters, the standard deviation
values of the bond length compared with the experimental data53,54
were 1.08 % for acridine and 1.55 % for phenazine.
Figure 4 shows that for both molecules the CC bond distance of
the benzene ring (ring A) was 1.34–1.44 Å, whereas the CN distance
of the nitrogen-containing ring (ring B) was approximately 1.34 Å,
both consistent with a delocalized scheme. The bonding properties
of the outer rings of acridine and phenazine were similar to each
other and to those of anthracene, differing only in the middle ring.
This result is in good agreement with previous studies51,57 and is
supported by the fact that the force constants of the outer rings of
the three molecules are similar, differing in the middle ring due to
the presence of the nitrogen atom.57
Further, we noted a mistake in the bonding labels in Figure 6(a)
in Sudarso et al.36 with reference to the numbering of the carbon
atoms of anthracene in Figure 2(a) of the same reference; the correct
bonding labels should be 1-2, 1-11, 9-11, 2-3, 11-12.
The molecular energy level diagram for the best modified
Figure 3. Results of aromaticity index of HOMA of the two non-equivalent rings (A and B) of (a) acridine and (b) phenazine, for the variation of (i) αCN (for
βCC= 1.0) and (ii) βCC (for αCN = 0.1). For all graphs, the right axis shows the standard deviation of bond length
Figure 4. Bond length of (a) acridine and (b) phenazine from the present work with the best result for the modied Hückel parameters, compared with the data
for anthracene36 and experimental and computational data from previous studies.53,54,57-60 Shading denotes CN bonding
Aromaticity of aza aromatic molecules 7Vol. XY, No. 00
Hückel parameter and the associated HOMO and LUMO obtained
from LCAO model are summarized in Table 8 for both acridine and
phenazine. Data for anthracene were obtained from our previous
study.36
Evident from the molecular energy level diagram, degeneracy in
the energy levels of acridine and phenazine was removed, which is
consistent with a previous semi-empirical self-consistent eld-LCAO-
MO study.24 The HOMO and LUMO plots obtained resembled those
obtained from the density functional theory (DFT) studies of Adad61
for acridine and Zendaoui59 for phenazine. The HOMO and LUMO
plots for acridine and phenazine were similar to those of anthracene.
In particular, the HOMO is localized primarily on the two C6 rings
but extends to a certain degree to the central ring. In contrast, the
LUMO is localized primarily on the central ring and presents a non-
negligible contribution from the C6 rings. For acridine and phenazine,
the essential contributor of the central ring is from the nitrogen atoms.
Table 9 compares some of the energy characteristics of the two
studied molecules with those of anthracene, compared with previous
experimental and advanced calculation studies. This correction
shows that the HOMO-LUMO gap energy was 0.84β for acridine
and 0.77β for phenazine. In addition, the delocalization energy was
5.72β for acridine and 6.07β for phenazine. The higher gap energy
of acridine relative to phenazine agrees with previous experiments
and advanced computational studies. Compared with anthracene, the
trend of the gap energy was consistent with the experimental optical
band-gap data, and the gap energy decreased with increasing nitrogen
to carbon ratio, which is in good agreement with a previous study.62
The delocalization energy of phenazine was found to be larger than
acridine, which is in good agreement with the increased molecular
stability when increasing the number of nitrogen atoms, but still
smaller than the parent molecule anthracene.
Based on the result of the HOMO energy, with β < 0, one
might expect a substantial shift to higher ionization energies from
anthracene to acridine and phenazine, which is consistent with
previous computational studies.67,68 In contrast, the opposite trend
was observed for the LUMO energy.
Bond energy and heat of formation from atoms
Figure 5 shows a comparison of the total molecular energy with
the experimental HtFfA for various benzenoid hydrocarbons (a)
and aza aromatic molecules (b). For the benzenoid hydrocarbons
anthracene and phenanthrene, we used the best modied Hückel
parameters, as reported previously.36 For both groups of molecules,
increasing the number of rings increased HtFfA, and the HtFfA value
of the angular molecule, phenantrene was larger than that of the
linear molecule, anthracene. For aza aromatic molecules, increasing
the number of nitrogen atoms from acridine to phenazine decreased
HtFfA. In addition, the dependence of HtFfA on the number of ring
molecules and nitrogen atoms is shown by the smaller corresponding
values of pyridine compared to those of benzene and the larger
corresponding values of acridine compared with those of anthracene.
In general, the variation in the total molecular energy related
to the three aromaticity indices followed the experimental HtFfA
value. However, a smaller deviation was observed for the bond
energies derived from the bond lengths based on the HOMED,
Table 8. Energy level diagram and plots of HOMO and LUMO of acridine and phenazine from the present work compared with those of anthracene. Different
colors of red and grey circles indicate the positive and negative coefcients of linear combinations, respectively
Anthracene Acridine Phenazine
MO Energy level diagram
HOMO
LUMO
Sutjahja et al.
8Quim. Nova
followed by those derived from the HOMHED and HOMA. In
particular, all models predicted a larger total molecular energy
value for phenanthrene compared with anthracene, which is in
agreement with the experimental HtFfA. Thus, the modied Hückel
parameters yielded total molecular energy values that matched the
experimental data. Compared with benzenoid hydrocarbon molecules,
a larger deviation between the calculated total energy molecular and
experimental HtFfA values was found for the aza aromatic molecules.
To further correlate the bond energy with the aromaticity index,
Table 10 presents the corresponding aromaticity HOMED index with
the REC divided by the number of CC and CN bonds (n).
Evidently, aromaticity varied with an increase in the number
of rings and the number of nitrogen atoms in the molecule. For
anthracene, the aromaticity of the central benzene ring was higher
than that of the outer benzene ring. The opposite trend was observed
for phenanthrene, which is in good agreement with our previous
study36 using the HOMA aromaticity index. With an increased
number of rings, the aromaticity index generally decreased, as
reported previously30,55. Upon replacement of carbon with nitrogen,
the aromaticity index of nitrogen-containing rings increased (for
example, 0.994 for acridine as compared with 0.972 for anthracene),
whereas the aromaticity index of the benzene ring decreased (0.952
for acridine as compared with 0.961 for anthracene). This result is
consistent with previous studies.55,56 For acridine and phenazine, the
aromaticity index of the nitrogen-containing ring was higher than that
of the benzene ring, which was consistent with previous studies using
geometric experimental data.55,56 However, the smaller aromaticity
index for the benzene- and nitrogen-containing rings in phenazine
compared with those of acridine indicates that further increase in the
number of nitrogen atoms decreased the aromaticity index.
Interestingly, for both groups of molecules, the variation in REC/n
from the present study generally followed the corresponding values
from experimental geometry data.30 The results of the present study
show that the REC/n value generally decreased as the number of
rings increased. In addition, upon replacing carbon with nitrogen, the
REC/n value generally decreased for both single-ring molecules and
for benzene- and nitrogen-containing rings. The results presented in
this study should be supported by further analysis based on advanced
calculations studies, in particular, studies on the relationship between
aromaticity and the number of nitrogen atoms and NN bonding, as
well as the topological environment of the ring.32,55
CONCLUSIONS
This study presented a correction to the Hückel parameters of
the aza aromatic molecules acridine and phenazine. The parameter
correction consists of the Coulomb integral of the carbon atom
adjacent to nitrogen (αCN) and the resonance integral of the
carbon–carbon atoms (βCC). The latter is based on the ring current
model, which describes the delocalization of π-electrons along the
perimeter of the molecule. The calculation was performed using
HuLiS software. The resulting bond order was transformed into a
bond length based on Gordy’s formula with modied Gordy values.
The validity of the results was examined using the structurally based
aromaticity indices of the HOMA, HOMED, and HOMHED. For the
three aromaticity indices, the experimental data of the two molecules
revealed that the nitrogen-containing ring is more aromatic than the
benzene ring. Additionally, both αCN and βCC parameters were found to
be responsible for the excessive degeneracy in the molecular orbitals.
The best values of αCN and βCC were 0.1 and 1.1 for acridine and 0.1 and
1.2 for phenazine, respectively. Comparison of the calculated bond
lengths with previous experimental and advanced calculation studies
revealed a delocalized scheme of CC and CN bonding. Compared with
anthracene, the bonding properties of the outer rings of acridine and
phenazine were similar to each other and also similar to anthracene,
differing only in the middle ring owing to the presence of nitrogen
Table 9. HOMO–LUMO gap energy, delocalization energy, HOMO and LUMO energies, and optical band gap of phenazine and acridine from the present work
(p.w) and previous experimental and advanced calculation studies, compared with those of anthracene.
Molecule HOMO–LUMO
gap energy Edeloc EHOMO ELUMO
HOMO–LUMO
gap energy (eV) Eg,opt (eV)
Anthracene 0.82β36 8.08β36 0.43β36 -0.39β36 3.5863 3.20*63
Acridine 0.84β (p.w) 5.72β (p.w) 0.52β (p.w) -0.32β (p.w) 3.764
3.70161 3.22*65
Phenazine 0.77β (p.w) 6.07β (p.w) 0.59β (p.w) -0.18β (p.w) 2.3759 3.0766
*Estimated from the absorption spectra by determining the wavelength of absorption onset and converting it from nm to eV.62
Figure 5. Total molecular energy based on HOMA, HOMED and HOMHED calculations compared with the experimental HtFfA for (a) benzenoid hydrocarbon
and (b) aza aromatic molecules. The numbers on (a) are Emolecule of anthracene and phenantrene calculated based on HOMED
Aromaticity of aza aromatic molecules 9Vol. XY, No. 00
atoms. The HOMO-LUMO energy gap was 0.84β for acridine and
0.77β for phenazine, which is in good agreement with the smaller
energy gap of phenazine compared to acridine from previous DFT and
experimental studies. The delocalization energy of phenazine (6.07β)
was larger than that of acridine (5.72β), which is in good agreement
with the trend of increasing molecular stability with an increasing
number of nitrogen atoms, but they are still smaller than anthracene.
To further validate the aromaticity index, we calculated the
molecular bond energy from the bond length for several benzenoid
hydrocarbons and aza aromatic molecules, including acridine and
phenazine. Considering the energy related to the CH bond, the total
molecular energy data were compared with the experimental data
for the heat of formation from atoms (HtFfA). The comparison
revealed that in general, modied Hückel parameters with calculated
bond energies derived from the three aromaticity indices based on
each single and double bond length as the reference resemble the
experimental HtFfA. The smallest deviation was observed for the
bond energy derived from the HOMED model. However, a larger
difference between the two values for aza aromatic molecules may
indicate some limitations of the existing model. A correlation study
between the HOMED index and REC divided by the number of CC
and CN bonds revealed a different role between the number of rings
and nitrogen atoms. Remarkably, the results from the simple Hückel
theory with modied parameters can resemble the experimental
results, from the viewpoints of the aromaticity index and energy
related to the bond length. However, advanced computational
studies are needed to conrm these results and to achieve a better
understanding of the relationship between the aromaticity index and
molecular energy from bond length and the bonding properties of a
particular molecule.
Table 10. Aromaticity indices of HOMED and REC/n for the individual rings. The values in the parentheses are calculated from experimental geometry data
listed in the reference
Group Molecule Molecule Ring HOMED index REC/n Reference
Benzenoid
hydrocarbon
Benzene
A 0.997 122.3 (118.8) 69
Naphthalene*
A 0.967 119.1 (116.6) 70
Anthracene
A 0.961 118.1 (116.7)
71
B 0.972 115.5 (115.3)
Phenanthrene
A 0.986 119.5 (117.0)
71
B 0.922 114.9 (111.3)
Aza aromatic
Pyridine
A 0.999 118.7 (113.8) 72
Acridine
A 0.952 115.7 (116.1)
54
B 0.994 112.7 (114.1)
Phenazine
A 0.948 115.8 (118.4)
53
B 0.970 112.1 (110.8)
*The modied Hückel parameters of a = 0.5, γ = 1.5.
ACKNOWLEDGEMENTS
This study is the output of the P2MI ITB 2023 research scheme.
It is dedicated to the late Professor Pantur Silaban for his memorable
contributions to theoretical physics at the Physics Department,
Faculty of Mathematics and Natural Sciences, Institut Teknologi
Bandung.
REFERENCES
1. Kubozono, Y.; Physics and Chemistry of Carbon-Based Materials, 1st
ed.; Springer: Singapore, 2019.
2. Anthony, J. E.; Chem. Rev. 2006, 106, 5028. [Crossref]
3. Mei, J.; Diao, Y.; Appleton, A. L.; Fang, L.; Bao, Z.; J. Am. Chem. Soc.
2013, 135, 6724. [Crossref]
4. Miao, Q.; Polycyclic Arenes and Heteroarenes: Synthesis, Properties,
and Applications, 1st ed.; Wiley-VCH: Weinheim, 2015.
5. Yutomo, E. B.; Noor, F. A.; Winata, T.; RSC Adv. 2021, 11, 18371.
[Crossref]
6. Ostroverkhova, O.; Chem. Rev. 2016, 116, 13279. [Crossref]
7. Ghosh, S.; Barg, S.; Jeong, S. M.; Ostrikov, K. (Ken); Adv. Energy
Mater. 2020, 10, 2001239. [Crossref]
8. Woo, J.; Lim, J. S.; Kim, J. H.; Joo, S. H.; Chem. Commun. 2021, 57,
7350. [Crossref]
9. Wang, X.-Y.; Yao, X.; Narita, A.; Müllen, K.; Acc. Chem. Res. 2019, 52,
2491. [Crossref]
10. Zhang, J.; Xia, Z.; Dai, L.; Sci. Adv. 2015, 1, 1. [Crossref]
11. Irham, M. A.; Anrokhi, M. S.; Anggraini, Y.; Sutjahja, I. M.; Quim. Nova
2021, 44, 1204. [Crossref]
12. Minkin, V. I.; Glukhovtsev, M. N.; Simkin, B. Y.; Aromaticity and
Sutjahja et al.
10 Quim. Nova
Antiaromaticity: Electronic and Structural Aspects, 1st ed.; Wiley-
Interscience: New York, 1994.
13. Szatylowicz, H.; Wieczorkiewicz, P. A.; Krygowski, T. M.; Sci 2022, 4,
24. [Crossref]
14. Fernandez, I.; Aromaticity (Modern Computational Methods and
Applications), 1st ed.; Elsevier: Amsterdam, 2021.
15. Krygowski, T. M.; Szatylowicz, H.; ChemTexts 2015, 1, 12. [Crossref]
16. Krygowski, T. M.; J. Chem. Inf. Comput. Sci. 1993, 33, 70. [Crossref]
17. Krygowski, T. M.; Szatylowicz, H.; Stasyuk, O. A.; Dominikowska, J.;
Palusiak, M.; Chem. Rev. 2014, 114, 6383. [Crossref]
18. Raczyńska, E. D.; Hallman, M.; Kolczyńska, K.; Stępniewski, T. M.;
Symmetry 2010, 2, 1485. [Crossref]
19. Raczyńska, E. D.; Symmetry 2019, 11, 146. [Crossref]
20. Frizzo, C. P.; Martins, M. A. P.; Struct. Chem. 2012, 23, 375. [Crossref]
21. Yates, K.; Hückel Molecular Orbital Theory; Academic Press: New
York, 2012.
22. Rauk, A.; Orbital Interaction Theory of Organic Chemistry, 2nd ed.; John
Wiley & Sons: New York, 2001.
23. Tjia, M. O.; Sutjahja, I. M.; Orbital Kuantum - Pengantar Teori dan
Contoh Aplikasinya; Karya Putra Darwati Bandung: Bandung, 2012.
24. Longuet-Higgins, H. C.; Coulson, C. A.; Trans. Faraday Soc. 1947, 43,
87. [Crossref]
25. Dasgupta, M.; Proc. Indian Natl. Sci. Acad. 1965, 31, 413. [Link]
accessed in January 5, 2023
26. Pauling, L.; J. Am. Chem. Soc. 1947, 69, 542. [Crossref]
27. Johnston, H. S.; Parr, C.; J. Am. Chem. Soc. 1963, 85, 2544. [Crossref]
28. Krygowski, T. M.; Ciesielski, A.; Bird, C. W.; Kotschy, A.; J. Chem. Inf.
Comput. Sci. 1995, 35, 203. [Crossref]
29. Krygowski, T. M.; Cyrański, M. K.; Theor. Comput. Chem. 1998, 5, 153.
[Crossref]
30. Cyrański, M. K.; Stępień, B. T.; Krygowski, T. M.; Tetrahedron 2000,
56, 9663. [Crossref]
31. Sevenster, A. J. L.; Tabner, B. J.; Magn. Reson. Chem. 1984, 22, 521.
[Crossref]
32. Abraham, R. J.; Reid, M.; J. Chem. Soc., Perkin Trans. 2 2002, 2, 1081.
[Crossref]
33. Najmidin, K.; Kerim, A.; Abdirishit, P.; Kalam, H.; Tawar, T.; J. Mol.
Model. 2013, 19, 3529. [Crossref]
34. Steiner, E.; Fowler, P. W.; Int. J. Quantum Chem. 1996, 60, 609.
[Crossref]
35. Cuesta, I. G.; De Merás, A. S.; Pelloni, S.; Lazzeretti, P.; J. Comput.
Chem. 2008, 30, 551. [Crossref]
36. Sudarso, Y. P. D.; Maulana, A. L.; Soehiani, A.; Sutjahja, I. M.; Quim.
Nova 2022, 45, 742. [Crossref]
37. Carissan, Y.; Hagebaum-Reignier, D.; Goudard N.; Humbel S.; J. Phys.
Chem. A 2008, 112, 13256. [Crossref]
38. Dascălu, D.; Isac, D.; Pahomi, A.; Isvoran, A.; Rom. J. Biophys. 2018,
28, 59. [Crossref]
39. Gordy, W.; J. Chem. Phys. 1947, 15, 305. [Crossref]
40. Paolini, J. P.; J. Comput. Chem. 1990, 11, 1160. [Crossref]
41. Kwapisz, J. H.; Stolarczyk, L. Z.; Struct. Chem. 2021, 32, 1393.
[Crossref]
42. Cyrański, M. K.; Krygowski, T. M.; Tetrahedron 1999, 55, 6205.
[Crossref]
43. Krygowski, T. M.; Cyrański, M.; Tetrahedron 1996, 52, 10255.
[Crossref]
44. Kruszewski, J.; Krygowski, T. M.; Tetrahedron Lett. 1972, 13, 3839.
[Crossref]
45. Huheey, J. E.; Keiter, E. A.; Keiter, R. L.; Inorganic Chemistry :
Principles of Structure and Reactivity, 4th ed.; HarperCollins College:
New York, 1993.
46. Roux, M. V.; Temprado, M.; J. Phys. Chem. Ref. Data 2008, 37, 1855.
[Crossref]
47. Hubbard, W. N.; Frow, F. R.; Waddington, G.; J. Phys. Chem. 1961, 65,
1326. [Crossref]
48. National Institute of Standards and Technology (NIST); Standard
Reference Database 69: Acridine; NIST Chemistry WebBook, 2021.
[Link] accessed in January 5, 2023
49. Steele, W. V.; Chirico, R. D.; Hossenlopp, I. A.; Nguyen, A.; Smith, N.
K.; Gammon, B. E.; J. Chem. Thermodyn. 1989, 21, 81. [Crossref]
50. National Institute of Standards and Technology (NIST); Standard
Reference Database 69: Phenazine; NIST Chemistry WebBook, 2021.
[Link] accessed in January 5, 2023
51. Fu, A.; Du, D.; Zhou, Z.; Spectrochim. Acta, Part A 2003, 59, 245.
[Crossref]
52. Li, W. H.; Li, X. Y.; Yu, N. T.; Chem. Phys. Lett. 2000, 327, 153.
[Crossref]
53. Woźniak, K.; Kariuki,B.; Jones, W.; Acta Crystallogr., Sect. C: Struct.
Chem. 1991, 47, 1113. [Crossref]
54. Phillips, D. C.; Acta Crystallogr. 1956, 9, 237. [Crossref]
55. Krygowski, T. M.; Cyrański, M. K.; Chem. Rev. 2001, 101, 1385.
[Crossref]
56. Cyrański, M.; Krygowski, T. M.; Tetrahedron 1996, 52, 13795.
[Crossref]
57. Bandyopadhyay, I.; Manogaran, S.; J. Mol. Struct.: THEOCHEM 2000,
507, 217. [Crossref]
58. Herbstein, F. H.; Schmidt, G. M. J.; Nature 1952, 169, 323. [Crossref]
59. Zendaoui, S. M.; Zouchoune, B.; Polyhedron 2013, 51, 123. [Crossref]
60. Hirshfeld, F. L.; Schmidt, G. M. J.; J. Chem. Phys. 1957, 26, 923.
[Crossref]
61. Adad, A.; Hmammouchi, R.; Lakhli, T.; Bouachrine, M.; J. Chem.
Pharm. Res. 2013, 5, 26. [Link] accessed in January 5, 2023
62. Kotwica, K.; Wielgus, I.; Proń, A.; Materials 2021, 14, 5155. [Crossref]
63. Costa, J. C. S.; Taveira, R. J. S.; Lima, C. F. R. A. C.; Mendes, A.;
Santos, L. M. N. B. F.; Opt. Mater. (Amsterdam, Neth.) 2016, 58, 51.
[Crossref]
64. Elangovan, A.; Chiu, H. H.; Yang, S. W.; Ho, T. I.; Org. Biomol. Chem.
2004, 2, 3113. [Crossref]
65. Zhou, C.; Zhang, T.; Zhang, S.; Liu, H.; Gao, Y.; Su, Q.; Wu, Q.; Li, W.;
Chen, J.; Yang, B.; Dyes Pigm. 2017, 146, 558. [Crossref]
66. Plasseraud, L.; Cattey, H.; Richard, P.; Ballivet-Tkatchenko, D.; J.
Organomet. Chem. 2009, 694, 2386. [Crossref]
67. Dolgounitcheva, O.; Zakrzewski, V. G.; Ortiz, J. V.; J. Phys. Chem. A
1997, 101, 8554. [Crossref]
68. Adhikari, S.; Santra, B.; Ruan, S.; Bhattarai, P.; Nepal, N. K.; Jackson,
K. A.; Ruzsinszky, A.; J. Chem. Phys. 2020, 153, 184303. [Crossref]
69. Jeffrey, G. A.; Ruble, J. R.; McMullan, R. K.; Pople, J. A.; Proc. R. Soc.
A 1987, 414, 47. [Crossref]
70. Brock, C. P.; Dunitz, J. D.; Acta Crystallogr., Sect. B: Struct. Sci., Cryst.
Eng. Mater. 1982, B38, 2218. [Crossref]
71. Kalescky, R.; Kraka, E.; Cremer, D.; J. Phys. Chem. A 2014, 118, 223.
[Crossref]
72. Herzberg, G.; Molecular Spectra and Molecular Structure. Vol. III:
Electronic Spectra and Electronic Structure of Polyatomic Molecules,
2nded.; Von Nostrand Reinhold Co.: New York, 1966.
This is an open-access article distributed under the terms of the Creative Commons Attribution License.
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Aromaticity, a very important term in organic chemistry, has never been defined unambiguously. Various ways to describe it come from different phenomena that have been experimentally observed. The most important examples related to some theoretical concepts are presented here.
Article
Full-text available
This short critical review is devoted to the synthesis and functionalization of various types of azaacenes, organic semiconducting compounds which can be considered as promising materials for the fabrication of n-channel or ambipolar field effect transistors (FETs), components of active layers in light emitting diodes (LEDs), components of organic memory devices and others. Emphasis is put on the diversity of azaacenes preparation methods and the possibility of tuning their redox and spectroscopic properties by changing the C/N ratio, modifying the nitrogen atoms distribution mode, functionalization with electroaccepting or electrodonating groups and changing their molecular shape. Processability, structural features and degradation pathways of these compounds are also discussed. A unique feature of this review concerns the listed redox potentials of all discussed compounds which were normalized vs. Fc/Fc+. This required, in frequent cases, recalculation of the originally reported data in which these potentials were determined against different types of reference electrodes. The same applied to all reported electron affinities (EAs). EA values calculated using different methods were recalculated by applying the method of Sworakowski and co-workers (Org. Electron. 2016, 33, 300-310) to yield, for the first time, a set of normalized data, which could be directly compared.
Article
Full-text available
Doping with nitrogen atom is an effective way to modify the electronic and magnetic properties of graphene. In this paper, we studied the effect of the number of dopant atoms on the electronic and magnetic properties of the two most common nitrogen bond configurations in N-doped graphene, that is, graphitic and pyridinic, using density functional theory (DFT). We found that the formation of graphitic and pyridinic configurations can initiate the transition of the electronic properties of graphene from semimetal to metal with n-type conductivity for the graphitic configuration and p-type conductivity for the pyridinic configuration. The formation of a bandgap-like structure was observed in both configurations. The bandgap increased with the increase in the number of dopant atoms. We also observed that the formation of graphitic configuration did not cause a transition to the magnetic properties of graphene even though the number of dopant atoms was increased. In the pyridinic configuration, the increase in the number of dopant atoms caused graphene to be paramagnetic, with the remarkable total magnetic moment of 0.400 μ B per cell in the pyridinic-N3 model. This study provides a deeper understanding of the modification of electronic and magnetic properties of N-doped graphene by controlling the bond configuration and the number of nitrogen dopants.
Article
Full-text available
In this study, we investigate the effects of nitrogen and boron dopants on the properties of phenalene/phenalenyl systems based on the Hückel theory by using the Hueckel Molecular Orbital software. The dopants configurations are graphitic, pyridinic, and pyrrolic. The electronic configuration of bare phenalene confirms the delocalization of π electrons and the radical properties of the molecule, which is in good agreement with the results of previous studies. Dopant types and positions strongly affect the number of π electrons in the system, molecular orbital energy, total energy, average π-electron energy, and gap energy. The molecular energy level degeneracy strongly depends on the rotational symmetry of the system, in the order of graphitic, pyridinic, and pyrrolic. A preserved radical behavior and the number of π electrons are found for the pyridinic dopant type, while closed electronic configuration is observed for graphitic and pyrrolic types. A lower gap energy is typically found for B-doped phenalene compared to that for N-doped phenalene; this opens the possibility for the enhancement of photoluminescence intensity. This study, although qualitative, confirms the effects of dopants on the chemical and physical properties of phenalene/phenalenyl systems.
Article
Full-text available
The equilibrium carbon-carbon (C-C) bond lengths in π -electron hydrocarbons are very sensitive to the electronic ground-state characteristic. In the recent two papers by Stolarczyk and Krygowski (J Phys Org Chem, 34:e4154,e4153, 2021) a simple quantum approach, the Augmented Hückel Molecular Orbital (AugHMO) model, is proposed for the qualitative, as well as quantitative, study of this phenomenon. The simplest realization of the AugHMO model is the Hückel-Su-Schrieffer-Heeger (HSSH) method, in which the resonance integral β of the HMO model is a linear function the bond length. In the present paper, the HSSH method is applied in a study of C-C bond lengths in a set of 34 selected polycyclic aromatic hydrocarbons (PAHs). This is exactly the set of molecules analyzed by Riegel and Müllen (J Phys Org Chem, 23:315, 2010) in the context of their electronic-excitation spectra. These PAHs have been obtained by chemical synthesis, but in most cases no diffraction data (by X-rays or neutrons) of sufficient quality is available to provide us with their geometry. On the other hand, these PAHs are rather big (up to 96 carbon atoms), and ab initio methods of quantum chemistry are too expensive for a reliable geometry optimization. That makes the HSSH method a very attractive alternative. Our HSSH calculations uncover a modular architecture of certain classes of PAHs. For the studied molecules (and their fragments – modules), we calculate the values of the aromaticity index HOMA.
Article
Full-text available
(Semi)-local density functional approximations (DFAs) suffer from self-interaction error (SIE). When the first ionization energy (IE) is computed as the negative of the highest-occupied orbital (HO) eigenvalue, DFAs notoriously underestimate them compared to quasi-particle calculations. The inaccuracy for the HO is attributed to SIE inherent in DFAs. We assessed the IE based on Perdew–Zunger self-interaction correction on 14 small to moderate-sized organic molecules relevant in organic electronics and polymer donor materials. Although self-interaction corrected DFAs were found to significantly improve the IE relative to the uncorrected DFAs, they overestimate. However, when the self-interaction correction is interiorly scaled using a function of the iso-orbital indicator zσ, only the regions where SIE is significantly get a correction. We discuss these approaches and show how these methods significantly improve the description of the HO eigenvalue for the organic molecules.
Article
Full-text available
Nanographenes, which are defined as nanoscale (1–100 nm) graphene cutouts, include quasi-one-dimensional graphene nanoribbons (GNRs) and quasi-zero-dimensional graphene quantum dots (GQDs). Polycyclic aromatic hydrocarbons (PAHs) larger than 1 nm can be viewed as GQDs with atomically precise molecular structures and can thus be termed nanographene molecules. As a result of quantum confinement, nanographenes are promising for next-generation semiconductor applications with finite band gaps, a significant advantage compared with gapless two-dimensional graphene. Similar to the atomic doping strategy in inorganic semiconductors, incorporation of heteroatoms into nanographenes is a viable way to tune their optical, electronic, catalytic, and magnetic properties. Such properties are highly dependent not only on the molecular size and edge structure but also on the heteroatom type, doping position, and concentration. Therefore, reliable synthetic methods are required to precisely control these structural features. In this regard, bottom-up organic synthesis provides an indispensable way to achieve structurally well-defined heteroatom-doped nanographenes.
Article
Full-text available
The geometry-based HOMA (Harmonic Oscillator Model of Aromaticity) descriptor, based on the reference compounds of different delocalizations of n- and π-electrons, can be applied to molecules possessing analogous bonds, e.g., only CC, only CN, only CO, etc. For compounds with different heteroatoms and a different number of CC, CX, XX, and XY bonds, its application leads to some discrepancies. For this reason, the structural descriptor was modified and the HOMED (Harmonic Oscillator Model of Electron Delocalization) index defined. In 2010, the HOMED index was parameterized for compounds with C, N and O atoms. For parametrization, the reference molecules of similar delocalizations of n- and π-electrons were employed. In this paper, the HOMED index was extended to compounds containing the CP, CS, NN, NP, PP, NO, NS, PO, and PS bonds. For geometrical optimization of all reference molecules and of all investigated heterocompounds, the same quantum–chemical method {B3LYP/6-311+G(d,p)} was used to eliminate errors of the HOMED estimation. For some tautomeric systems, the Gn methods were also employed to confirm tautomeric preferences. The extended HOMED index was applied to five-membered heterocycles, simple furan and thiophene, and their N and P derivatives as well as for tautomeric pyrrole and phosphole and their N and P derivatives. The effects of additional heteroatom(s) in the ring on the HOMED values for furan are parallel to those for thiophene. For pyrroles, aromaticity dictates the tautomeric preferences. An additional N atom in the ring only slightly affects the HOMED values for the favored and well delocalized NH tautomers. Significant changes take place for their rare CH forms. When intramolecular proton-transfer is considered for phosphole and its P derivatives, the PH tautomers seem to be favored only for 1,2,3-triphosphole/1,2,5-triphosphole and for 1,2,3,5-tetraphosphole. For other phospholes, the CH forms have smaller Gibbs energies than the PH isomers. For phosphazoles, the labile proton in the favored form is linked to the N atom. The PH forms have smaller HOMED indices than the NH tautomers but higher than the CH ones.
Article
Oxygen reduction reaction (ORR) plays a pivotal role in electrochemical energy conversion and commodity chemical production. Oxygen reduction involving a complete four-electron (4e-) transfer is important for the efficient operation of polymer electrolyte fuel cells, whereas the ORR with a partial 2e- transfer can serve as a versatile method for producing industrially important hydrogen peroxide (H2O2). For both the 4e- and 2e- pathway ORR, platinum-group metals (PGMs) have been materials of prevalent choice owing to their high intrinsic activity, but they are costly and scarce. Hence, the development of highly active and selective non-precious metal catalysts is of crucial importance for advancing electrocatalysis of the ORR. Heteroatom-doped carbon-based electrocatalysts have emerged as promising alternatives to PGM catalysts owing to their appreciable activity, tunable selectivity, and facile preparation. This review provides an overview of the design of heteroatom-doped carbon ORR catalysts with tailored 4e- or 2e- selectivities. We highlight catalyst design strategies that promote 4e- or 2e- ORR activity. We also summarise the major active sites and activity descriptors of the respective ORR pathways and describe the catalyst properties controlling the ORR mechanisms. We conclude the review with a summary and suggestions for future research.
Book
This book includes the fundamental science and applications of carbon-based materials, in particular fused polycyclic hydrocarbon, fullerene, diamond, carbides, graphite and graphene etc. During the past decade, these carbon-based materials have attracted much interest from many scientists and engineers because of their exciting physical properties and potential application toward electronic and energy devices. In this book, the fundamental theory referring to these materials, their syntheses and characterizations, the physical properties (physics), and the applications are fully described, which will contribute to an advancement of not only basic science in this research field but also technology using these materials. The book's targets are researchers and engineers in the field and graduate school students who specialize in physics, chemistry, and materials science. Thus, this book addresses the physics and chemistry of the principal materials in the twenty-first century.