You are on page 1of 17

Polycyclic Aromatic Compounds

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/gpol20

Vibrational Hamiltonian of Naphthalene (C10H8)


Using Dynamical U(2) Lie Algebras

Malleswara Rao Balla & Vijayasekhar Jaliparthi

To cite this article: Malleswara Rao Balla & Vijayasekhar Jaliparthi (2021): Vibrational Hamiltonian
of Naphthalene (C10H8) Using Dynamical U(2) Lie Algebras, Polycyclic Aromatic Compounds, DOI:
10.1080/10406638.2021.1901126

To link to this article: https://doi.org/10.1080/10406638.2021.1901126

Published online: 22 Mar 2021.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gpol20
POLYCYCLIC AROMATIC COMPOUNDS
https://doi.org/10.1080/10406638.2021.1901126

Vibrational Hamiltonian of Naphthalene (C10H8) Using


Dynamical U(2) Lie Algebras
Malleswara Rao Ballaa and Vijayasekhar Jaliparthib
a
Department of Physics, GITAM (Deemed to be University), Hyderabad, India; bDepartment of Mathematics,
GITAM (Deemed to be University), Hyderabad, India

ABSTRACT ARTICLE HISTORY


This article is the first to demonstrate the application of dynamical U(2) Lie Received 8 November 2020
algebras to model a vibrational Hamiltonian of naphthalene (C10H8), a Accepted 2 March 2021
prototype polycyclic aromatic hydrocarbon molecule containing a fused
KEYWORDS
pair of benzene rings, in the gas phase having considerable astrophysical
Dynamical U(2) Lie
interest. A vibrational Hamiltonian that preserves the D2h symmetry of algebras; molecular
naphthalene is modeled using 19 coupled one-dimensional Morse oscilla- vibration spectroscopy;
tors to describe the eight C-H single bonds, the eleven C-C aromatic naphthalene; polycyclic
bonds, and their dynamic interactions, making use of the Casimir and aromatic hydrocarbons
Majorana operators. Results indicate that the dynamical U(2) Lie algebraic
Hamiltonian has successfully reproduced the vibrational frequencies at the
48 fundamental excitations near to the level of spectroscopic accuracy, as
well as predicted their first three overtones. The observed degeneracy of
the stretching, bending, and combination vibrations at the fundamental
and lower overtone excitations are expected to modify the shape of the
two major bands (centered at 782 cm1 and 3065 cm1), and several minor
bands in the infrared absorption spectrum of naphthalene in the
gas phase.

1. Introduction
Molecular spectroscopy of aromatic hydrocarbons is an exotic study of analyzing and interpreting
their interaction with electromagnetic radiation in terms of rotation-vibration (rovibration) states
from the very standpoint of the underlying molecular structure and dynamics. This exhilarating
paradigm is undergoing a pragmatic change in recent years with the development of highly
sophisticated experimental techniques. For example, the advent of tunable, stable and more
powerful lasers demand newer detection techniques with sensitivities far exceeding the limits of
today’s detectors. The dramatic interplay of quantum theory and the refined experimental obser-
vations concerned with rovibrational spectra of aromatic hydrocarbons gained much attention of
both experimentalists as well as theoreticians nowadays. The limitations inherent with current
theoretical approaches and the availability of newer experimental results, in turn, demand the
development of alternative and more robust theoretical models to test the assignments and expli-
cate the features when observations are open for multiple interpretations.
There have been two conventional theoretical approaches for the analysis and interpretation of
experimental data involving rovibration spectra of polyatomic molecules. The first approach is
the Dunham expansion1 which is a series of expansion of energy levels in terms of rovibration
quantum numbers. Dunham expansion, however, is not suggestible for a rigorous modeling work

CONTACT Vijayasekhar Jaliparthi vijayjaliparthi@gmail.com Department of Mathematics, GITAM (Deemed to be


University), Hyderabad, India.
ß 2021 Taylor & Francis Group, LLC
2 M. R. BALLA AND V. JALIPARTHI

due to several limitations. First of all, this expansion lacks in correspondence to the wave func-
tions of individual states, and hence operator matrix elements cannot be calculated directly.
Secondly, the number of parameters required to account for large polyatomic molecules in the
Dunham expansion is large, and the experimental database needed to optimize these parameters
is also large, which may not be feasible always.2 The second approach, and better as compared to
Dunham expansion, is solving the Schrodinger equation with potentials by conveniently modify-
ing ab-initio method3 and the provided wave functions to calculate the matrix elements of the
required operators. Since all the manipulations are of either differentiation or integration, this
approach also encounters problems for the case of larger molecules and highly excited levels. The
third, and probably a more interesting, alternative is the symmetry-adapted Lie alge-
braic approach.4
The framework of Lie algebras was introduced by Sophus Lie more than a century ago; how-
ever its application to molecular spectroscopy, introduced as the vibron model by Iachello,5 is
quite exciting and relatively new. The Lie algebraic approach is based on the dynamical symmetry
of structures, which when applied to molecules yields an effective Hamiltonian operator that
describes in handy the rotational and vibrational degrees of freedom of the physical system under
consideration. Here, the vibrational Hamiltonian is expressed in terms of boson operators to
characterize the local and normal modes of the system. Interestingly, Lie algebraic framework is
designed in such a way that it encases the underlying physical information of both the ab-initio
and semi-empirical approaches. The vibron model in its original formulation6–8 is the U(4) vari-
ant where rotations and vibrations were accounted for simultaneously. While applying for poly-
atomic molecules, it is found to be more convenient to first isolate the rotations from vibrations
and subsequently to treat them in terms of coupled U(2) Lie algebras9–12 such that each inter-
atomic potential is assigned with a U(2) algebra. By doing so, the computation of matrix elements
will be greatly simplified, and the underlying discrete symmetries will also be well-incorporated.
This symmetry adopted vibron model is very well-suited to describe the vibrations of polyatomic
molecules having a high degree of symmetry. The U(2) Lie algebraic models have been reported
to be successful in accurately predicting the vibrational spectra of molecules evolving in a variety
of structures including tetrahedral,13,14 octahedral,11,15 fullerenes,16–19 and even the nano-bio
molecules.20–24
Aromatic hydrocarbons are the compounds that exhibit aromaticity by virtue of having one or
more benzene-like carbon-hydrogen cyclic structures with their p electrons delocalized around all
over them. Polycyclic aromatic hydrocarbons (PAHs) are considered to have prevailed in all
phases of the interstellar medium, and they are also believed to be the precursors of biogenic
molecules that hold the key to the origin of life on earth.25 They are also suggested to be the car-
riers of the unidentified emission bands at certain specific infrared frequencies (e.g., at 3050,
1610, 1300, 1160 and 890 cm1) and the diffuse interstellar bands.26,27 Naphthalene (C10H8) is
the simplest and prototype PAH containing a fused pair of benzene rings. It is also one among
the 200 plus PAHs identified in stellar environments and hence is of considerable interest in
astrophysics.26–31
Several researchers have reported the theoretically computed vibrational spectra of naphtha-
lene.26,27,32–34 However, these earlier works were carried out using either ab-initio theory or
semi-empirical methods, and were mostly focused on deriving frequencies and assignments for
the fundamental mode vibrations only. Mackie et al35 presented the anharmonic theoretical spec-
tra of naphthalene, computed with a locally modified version of the SPECTRO program using
Cartesian derivatives transformed from Gaussian 09 normal coordinate force constants. They
have observed a fairly good agreement of their calculated vibrational frequencies with the NIST
experiment, except for few deviations due to the level of electronic structure theory and lack of
higher order overtone spectra. Chakraborty et al36 reported the gas phase infrared spectra of
naphthalene in the 3200 to 500 cm1 range and compared their experimental results with
POLYCYCLIC AROMATIC COMPOUNDS 3

Figure 1. Structure of an isolated naphthalene (C10H8) molecule in the gas phase. Also indicated are the assigned Lie algebras
U1 ð2Þ to U8 ð2Þ for C-H and U1 ð2Þ to U11

ð2Þ for C-C bond species.

electronic structure calculations performed following two theoretical frameworks (MP2 and
B3LYP) and three different methods (VCCM, VSCF, and VPT2). They have observed that results
vary significantly from method to method, between the theoretical frameworks, as well as across
the spectral region. The present work, on the other hand, is anticipated at modeling a vibrational
Hamiltonian of naphthalene using the one-dimensional framework of dynamical U(2) Lie alge-
bras to derive the vibrational frequencies more accurately at the fundamental as well as higher
overtones. As compared to the other works reported earlier, the present work is anticipated at
predicting the vibrational frequencies more accurately at a significantly lower computational cost.

2. Structure of naphthalene
Naphthalene (C10H8) is the simplest and prototype PAH molecule containing a fused pair of ben-
zene rings, as depicted in Figure 1. It belongs to the symmetry point group D2h with the sym-
metry species (irreducible representations) A1g, B1g, B2g, B3g, A1u, B1u, B2u, and B3u. Naphthalene
is characterized by 48 normal vibrational modes, out of which 33 are symmetric, and 15 are
asymmetric. As per the structure shown, there are four C ¼ C double bonds, seven C  C single
bonds, and eight C  H single bonds present in the molecule. However, due to the presence of
delocalized p electrons resonating over the pair of benzene-like structures, the C ¼ C and C  C
bonds will become indistinguishable. Hence, all these eleven bonds will be treated as identical. A
one-dimensional Morse oscillator and a U(2) Lie algebra are assigned to each of these bond spe-
cies to model their interactions appropriately in a vibrational Hamiltonian.
4 M. R. BALLA AND V. JALIPARTHI

3. U(2) Lie algebraic Hamiltonian of naphthalene


As per the structure of an isolated molecule of naphthalene in the gas phase (Figure 1), we have
introduced eight U(2) Lie algebras to describe the eight C-H stretching bonds:
 
U1 ð2Þ, U2 ð2Þ, U3 ð2Þ, U4 ð2Þ, U5 ð2Þ, U6 ð2Þ, U7 ð2Þ, U8 ð2Þ

The interactions among these vibrations can be effectively described by two chains of dynam-
ical U(2) Lie algebras as:

U1 ð2Þ  U2 ð2Þ  U3 ð2Þ  U4 ð2Þ  U5 ð2Þ  U6 ð2Þ  U7 ð2Þ  U8 ð2Þ
Chain I :
 O1 ð2Þ  O2 ð2Þ  O3 ð2Þ  O4 ð2Þ  O5 ð2Þ  O6 ð2Þ  O7 ð2Þ  O8 ð2Þ

U1 ð2Þ  U2 ð2Þ  U3 ð2Þ  U4 ð2Þ  U5 ð2Þ  U6 ð2Þ  U7 ð2Þ  U8 ð2Þ
Chain II :
 Uð2Þ  Oð2Þ
Here, the Chain I and Chain II describe local mode and normal mode couplings, respectively.
Similarly, to describe the eleven identical C-C stretching bonds, we have introduced eleven
U(2) Lie algebras:
  
U1 ð2Þ, U2 ð2Þ, U3 ð2Þ, U4 ð2Þ, U5 ð2Þ, U6 ð2Þ, U7 ð2Þ, U8 ð2Þ, U9 ð2Þ, U10 ð2Þ, U11 ð2Þ

The interactions among the C-H bond stretching vibrations in naphthalene are:
Interaction I: First–neighbour couplings
 
U1 ð2Þ  U2 ð2Þ, U2 ð2Þ  U3 ð2Þ, U3 ð2Þ  U4 ð2Þ, U4 ð2Þ  U5 ð2Þ,
U5 ð2Þ  U6 ð2Þ, U6 ð2Þ  U7 ð2Þ, U7 ð2Þ  U8 ð2Þ, U1 ð2Þ  U8 ð2Þ
Interaction II: Second–neighbour couplings
 
U1 ð2Þ  U3 ð2Þ, U2 ð2Þ  U4 ð2Þ, U3 ð2Þ  U5 ð2Þ, U4 ð2Þ  U6 ð2Þ,
U5 ð2Þ  U7 ð2Þ, U6 ð2Þ  U8 ð2Þ, U1 ð2Þ  U7 ð2Þ, U2 ð2Þ  U8 ð2Þ
Interaction III: Third–neighbour couplings
 
U1 ð2Þ  U4 ð2Þ, U2 ð2Þ  U5 ð2Þ, U3 ð2Þ  U6 ð2Þ, U4 ð2Þ  U7 ð2Þ,
U5 ð2Þ  U8 ð2Þ, U1 ð2Þ  U6 ð2Þ, U2 ð2Þ  U7 ð2Þ, U3 ð2Þ  U8 ð2Þ
Interaction IV: Fourth–neighbour couplings
 
U1 ð2Þ  U5 ð2Þ, U2 ð2Þ  U6 ð2Þ, U3 ð2Þ  U7 ð2Þ, U4 ð2Þ  U8 ð2Þ

The interactions among the C-C bond stretching vibrations in naphthalene are:

Interaction I: First–neighbour couplings


8  9
< U1 ð2Þ  U2 ð2Þ, U2 ð2Þ  U3 ð2Þ, U3 ð2Þ  U4 ð2Þ, U4 ð2Þ  U5 ð2Þ, U5 ð2Þ  U6 ð2Þ, =
U ð2Þ  U7 ð2Þ, U7 ð2Þ  U8 ð2Þ, U8 ð2Þ  U9 ð2Þ, U9 ð2Þ  U10 ð2Þ, U10 ð2Þ  U11 ð2Þ,
: 6 ;
U1 ð2Þ  U11 ð2Þ

Interaction II: Second–neighbour couplings


8  9
< U1 ð2Þ  U3 ð2Þ, U2 ð2Þ  U4 ð2Þ, U3 ð2Þ  U5 ð2Þ, U4 ð2Þ  U6 ð2Þ, U5 ð2Þ  U7 ð2Þ, =
U ð2Þ  U8 ð2Þ, U7 ð2Þ  U9 ð2Þ, U8 ð2Þ  U10 ð2Þ, U9 ð2Þ  U11 ð2Þ, U1 ð2Þ  U10 ð2Þ,
: 6 ;
U2 ð2Þ  U11 ð2Þ
POLYCYCLIC AROMATIC COMPOUNDS 5

Interaction III: Third–neighbour couplings


8  9
< U1 ð2Þ  U4 ð2Þ, U2 ð2Þ  U5 ð2Þ, U3 ð2Þ  U6 ð2Þ, U4 ð2Þ  U7 ð2Þ, U5 ð2Þ  U8 ð2Þ, =
U ð2Þ  U9 ð2Þ, U7 ð2Þ  U10 ð2Þ, U8 ð2Þ  U11 ð2Þ, U1 ð2Þ  U9 ð2Þ, U2 ð2Þ  U10 ð2Þ,
: 6 ;
U3 ð2Þ  U11 ð2Þ

8 Interaction IV: Fourth–neighbour couplings 9


< U1 ð2Þ  U5 ð2Þ, U2 ð2Þ  U6 ð2Þ, U3 ð2Þ  U7 ð2Þ, U4 ð2Þ  U8 ð2Þ, U5 ð2Þ  U9 ð2Þ, =
U ð2Þ  U10 ð2Þ, U7 ð2Þ  U11 ð2Þ, U1 ð2Þ  U8 ð2Þ, U2 ð2Þ  U9 ð2Þ, U3 ð2Þ  U10 ð2Þ,
: 6 ;
U4 ð2Þ  U11 ð2Þ

8 Interaction V: Fifth–neighbour couplings 9


< U1 ð2Þ  U6 ð2Þ, U2 ð2Þ  U7 ð2Þ, U3 ð2Þ  U8 ð2Þ, U4 ð2Þ  U9 ð2Þ, U5 ð2Þ  U10 ð2Þ, =
U ð2Þ  U11 ð2Þ, U1 ð2Þ  U7 ð2Þ, U2 ð2Þ  U8 ð2Þ, U3 ð2Þ  U9 ð2Þ, U4 ð2Þ  U10 ð2Þ,
: 6 ;
U5 ð2Þ  U11 ð2Þ
The Hamiltonian operator that preserves the D2h symmetry of the molecule is now constructed
as follows. Since all the C-H bonds in the molecule are equivalent, the effective Hamiltonian for
C-H stretching vibrations of naphthalene is given by
9
X
n¼8 X
n¼8 X
n¼8
>
H ¼ E0 þ Ai Ci þ Aij Cij þ ðkij þ kij þ kij þ kij Þkij Mij >
1 2 3 4 >
>
>
>
>
i¼1 i<j i<j >
>
with the symmetry adapted 1st, 2nd, 3rd and 4th neighbour > >
=
coupling coefficients as (1a)
k112 , :::, k118 ¼ 1; k113 , :::, k128 ¼ 0; k114 , :::, k138 ¼ 0; k115 , :::, k148 ¼ 0 >
>
>
>
k212 , :::, k218 ¼ 0; k213 , :::, k228 ¼ 1; k214 , :::, k238 ¼ 0; k215 , :::, k248 ¼ 0 >
>
>
>
k312 , :::, k318 ¼ 0; k313 , :::, k328 ¼ 0; k314 , :::, k338 ¼ 1; k315 , :::, k348 ¼ 0 >
>
>
;
k412 , :::, k418 ¼ 0; k413 , :::, k428 ¼ 0; k414 , :::, k438 ¼ 0; k415 , :::, k448 ¼ 1
Similarly, the effective Hamiltonian for the C-C stretching vibrations of naphthalene is given
by
nX
¼11 nX
¼11 nX
¼11
9
>
>
H ¼ E0 þ 0
A i Ci þ 0
A ij Cij þ ðlij1 þ lij2 þ lij3 þ lij4 þ lij5 Þk0 ij Mij >
>
>
>
>
i¼1 i<j i<j >
>
with the symmetry adapted 1st, 2nd, 3rd, 4th and 5th neighbour >
>
>
>
coupling coefficients as =
1
l12 , :::, l111
1
¼ 1; l13
1
, :::, l211
1
¼ 0; l14
1
, :::, l311
1
¼ 0; l15
1
, :::, l411
1
¼ 0; l16
1
, :::, l511
1
¼ 0> (1b)
>
l12 , :::, l111 ¼ 0; l13 , :::, l211 ¼ 1; l14 , :::, l311 ¼ 0; l15 , :::, l411 ¼ 0; l16 , :::, l511
2 2 2 2 2 2 2 2 2 2
¼ 0>>
>
>
3
l12 , :::, l111
3
¼ 0; l13
3
, :::, l211
3
¼ 0; l14
3
, :::, l311
3
¼ 1; l15
3
, :::, l411
3
¼ 0; l16
3
, :::, l511
3
¼ 0>>
>
>
4
l12 , :::, l111
4
¼ 0; l13
4
, :::, l211
4
¼ 0; l14
4
, :::, l311
4
¼ 0; l15
4
, :::, l411
4
¼ 1; l16
4
, :::, l511
4
¼ 0>>
>
;
5
l12 , :::, l111
5
¼ 0; l13
5
, :::, l211
5
¼ 0; l14
5
, :::, l311
5
¼ 0; l15
5
, :::, l411
5
¼ 0; l16
5
, :::, l511
5
¼ 1
Here, E0 is the term representing the eigenvalue of the Schrodinger equation associated with
the electronic ground state of the bond vibrations, which will be taken as the zero reference for
all the excitations; Ai , A0 i , kij , k0 ij are the algebraic parameters; Ci , Cij are the Casimir (invariant)
operators of the applicable Lie algebras; and Mij are the Majorana (invariant) operators connected
with coupling schemes involving Lie algebras of the n interacting one-dimensional Morse oscilla-
tors. The Casimir operator Cij is diagonal in vi and vj, which means that Cij does not change the
numbers of vibrational quanta in the bonds i and j. Hence, Cij is interpreted as the static inter-
action between the bonds i and j. The Majorana operator, Mij , in contrast, transfers one vibra-
tional quantum from bond i to the bond j; and hence is regarded as the dynamic interaction
between bonds i and j through exchanging vibrational quanta where the configurations kij are
6 M. R. BALLA AND V. JALIPARTHI

P
n
mixed up by these dynamic interactions. The summation Ai Ci is devoted to the description of
i¼1
n independent, anharmonic sequence of vibrational levels associated with n independent, local
Pn
oscillators; the summation Aij Cij leads to cross-anharmonicities between pairs of distinct local
i<j
Pn 1 Pn 1
i<j ðkij þ kij þ kij þ kij Þkij Mij and i<j ðlij þ lij þ lij þ lij þ
2 3 4 2 3 4
oscillators; and the summations
lij5 Þk0 ij Mij describe anharmonic, non-diagonal interactions involving pairs of local oscillators.
For naphthalene in the gas phase, the algebraic parameters are determined from the spectro-
scopic data,37 and the algebraic operators are determined from the following expressions:38

hCi i ¼ 4ðNi vi  v2i Þ (2)


  h i
hNi , vi ; Nj , vj Cij Ni , vi ; Nj , vj i ¼ 4 ðvi þ vj Þ2  ðvi þ vj ÞðNi þ Nj Þ (3)
  9
hNi , vi ; Nj , vj Mij Ni , vi ; Nj , vj i ¼ ðNi vj þ Nj vi – 2vi vj Þ >
   1=2 =
hNi , vi þ 1; Nj , vj  1Mij Ni , vi ; Nj , vj i ¼  vj ðvi þ 1ÞðNi –vi ÞðNj –vj þ 1Þ (4)
   1=2 >
;
hNi , vi  1; Nj , vj þ 1Mij Ni , vi ; Nj , vj i ¼  vi ðvj þ 1ÞðNj –vj ÞðNi –vi þ 1Þ

Here vi , vj denotes the local stretching vibrations of the molecule.


The working equations for the matrix elements of hCi i, hCij i, hMij i takes the following values
for the fundamental vibrations
9
hCi i ¼ 4ðN  1Þ >
>
hCij i ¼4ð2N  1Þ =
(5)
Nði 6¼ jÞ >>
hMij i ¼ ;
Nði ¼ jÞ

The vibron number N, which correspond to the maximum number of bound states of the
Morse potential in each vibrating bond species, is calculated from the relation39
xe
N¼ 1 (6)
xe xe
where xe and xexe are the spectroscopic constants of the corresponding bonds obtained from
experimental data40,41 of diatomic molecules. Considering the experimental values of these con-
stants for C-H (xe ¼ 2860.7508 cm1, xexe ¼ 64.4387 cm1) and C-C (xe ¼ 1855.0663 cm1,
xexe ¼ 13.6007 cm1), their N values for naphthalene are fixed as 43 and 135, respectively.
The initial guess for Ai is obtained from the energy expression for single-oscillator fundamen-
tal mode39,42
)
 CH 
E i ðv ¼ 1Þ ¼ 4A i ðN1  1Þ (7)
 CC 0
E i ðv ¼ 1Þ ¼ 4A i ðN2  1Þ

as
9
 CH
E >
>
i ¼
A i >
=
4ð1  N1 Þ
(8)
 CC
E >
>
 i0 ¼
A i >
;
4ð1  N2 Þ
 i and E
where A  i are the average values of the corresponding parameters Ai 0 s and E0 s:
POLYCYCLIC AROMATIC COMPOUNDS 7

Table 1. Optimized fitting parameters of the U(2) Lie algebraic Hamiltonian for naphthalene molecule.
U(2) Lie algebraic model parameter Optimized value
N1 ðCHÞ 43
N2 ðCCÞ 135
Ai ðCHÞ 1.0157
A0 i ðCCÞ 1.9854
kij ðCHÞ 0.8052
k0 ij ðCCÞ 1.9016
Aij ðCHÞ 0.1105
A0 ij ðCCÞ 0.1773

The initial guesses for kij , k0 ij are obtained by the relations39


 CH  9
E  ECH  >
kij ¼ 1 2 >
=
 CC2N CC  (9)
E  E  >
k0 ij ¼ 1 2 >
;
6N
Here, ECH CC CH CC
1 (E1 ) and E2 (E2 ) are the different energies corresponding to symmetric and anti-
symmetric combination of the two local modes. Their role is to split the initial degenerate local
modes, placed here at the common value E  used in eq. (8). The values of Ai , A0 i , kij , k0 ij , NCH , NCC
are optimized by least-square regression fitting, starting from the initial guesses as given by equa-
tions (8) and (9). The initial guesses for Aij , A0 ij are taken as zero.

4. Results and discussion


In this work, we have applied the symmetry-adopted dynamical U(2) Lie algebras to model a
vibrational Hamiltonian for naphthalene (C10H8), an important PAH in the gas phase having
astrophysical interest. Starting from the initial guesses as obtained from standard spectroscopic
procedures and the subsequent numerical fitting aided by least-square regression analysis resulted
in the optimized values of Ai , A0 i , Aij , A0 ij , kij , k0 ij : The resulted fitting coefficients and the vibron
numbers (N1, N2) for the C-H and C-C bond species for naphthalene are listed in Table 1. All
the parameters are expressed in cm1, except N1 and N2, which are dimensionless.
Vibrational frequencies (x  calc , in units of cm1) of naphthalene (C10H8) at the fundamental
mode, as predicted the dynamical U(2) Lie algebraic Hamiltonian and their experimentally
observed counterparts (x  exp ), along with vibrational mode assignments, symmetry species, and
percentage deviations from the experiment, are given in Table 2. The dynamical U(2) Lie alge-
braic Hamiltonian modeled in this work is successful in realizing all the 48 fundamental vibra-
tional frequencies (labelled v1 through v48) of naphthalene molecule in the gas phase observed
experimentally. These 48 fundamental vibrations are appropriately assigned for their vibrational
modes and symmetry species. Also shown in Table 2 are the available vibrational spectra (x  exp )
of gas phase naphthalene from experiment,37 and the calculated percentage deviations (Dx  calc ) of
the dynamical U(2) Lie algebraic Hamiltonian predicted vibrational frequencies from that of the
experiment. It is observed that the lowest deviation of the predicted frequencies stands within
0.01% of the experiment, which occurs for the v31 (CC stretch þ HCC i-p bend) combination
vibration with the symmetry B2u. The highest deviation, on the other hand, stands within 0.78%
of the experiment, which occurs for the v48 (ring torsion) vibration with the symmetry B3u. The
root-mean-squared deviation (Dx  rms ), which reflects the overall accuracy of the dynamical U(2)
Lie algebraic Hamiltonian in deriving frequencies over the 48 fundamental vibrations, is found to
be 1.2834 cm1. Such a low root-mean-squared deviation over a large number of modes indicates
that the dynamical U(2) Lie algebraic Hamiltonian modeled in this work is successful in deriving
the fundamental vibration frequencies near to the spectroscopic level of accuracy.
8 M. R. BALLA AND V. JALIPARTHI

Table 2. Fundamental vibration spectra of naphthalene (C10H8) as predicted by our U(2) Lie algebraic Hamiltonian (x  calc ) and
their experimentally observed counterparts (x exp ), along with vibrational mode assignments, symmetry species (irreducible
representations), and percentage deviations from the experiment. Root-mean-squared deviation (Dx  rms ) of our U(2) Lie alge-
braic Hamiltonian predicted values (fundamental mode) stands within 1.2834 cm1 of the experiment.
Vibrational Mode Symmetry  obs (cm1)
x  calc (cm1)
x  calc
% Deviation Dx
v1 Aromatic CH stretch Ag 3060 3058.19 0.06
v2 Aromatic CH stretch Ag 3031 3029.87 0.04
v3 CC stretch (major) Ag 1577 1578.05 0.07
v4 HCC i-p bend (major) Ag 1460 1462.21 0.15
v5 CC stretch (major) Ag 1376 1377.15 0.08
v6 HCC i-p bend (major) Ag 1145 1146.98 0.17
v7 CC stretch þ HCC i-p bend Ag 1025 1023.03 0.19
v8 CC stretch þ HCC i-p bend þ CCC bend Ag 758 759.44 0.19
v9 CCC ring bend Ag 512 511.77 0.04
v10 HCC o-o-p bend Au 970 968.18 0.19
v11 HCC o-o-p bend Au 841 843.68 0.32
v12 ring torsion Au 581 583.23 0.38
v13 ring torsion Au 195 194.73 0.14
v14 HCC o-o-p bend B1g 943 941.78 0.13
v15 HCC o-o-p bend B1g 717 715.55 0.2
v16 ring torsion B1g 386 385.03 0.25
v17 Aromatic CH stretch B1u 3065 3066.01 0.03
v18 Aromatic CH stretch B1u 3058 3057.17 0.03
v19 CC stretch þ HCC i-p bend B1u 1595 1594.12 0.06
v20 HCC i-p bend B1u 1389 1390.34 0.1
v21 HCC i-p bend B1u 1265 1266.14 0.09
v22 HCC i-p bend B1u 1125 1126 0.09
v23 CCC ring bend þ HCC i-p bend B1u 753 752.59 0.05
v24 CCC ring bend B1u 359 358.05 0.26
v25 HCC o-o-p bend B2g 980 982 0.2
v26 HCC o-o-p bend B2g 876 875.09 0.1
v27 ring torsion B2g 770 768.95 0.14
v28 ring torsion B2g 461 459.9 0.24
v29 Aromatic CH stretch B2u 3090 3091.76 0.06
v30 Aromatic CH stretch B2u 3027 3028.1 0.04
v31 CC stretch þ HCC i-p bend B2u 1506 1505.87 0.01
v32 HCC i-p bend B2u 1361 1360.29 0.05
v33 HCC i-p bend B2u 1209 1208.81 0.02
v34 HCC i-p bend B2u 1138 1139 0.09
v35 CC stretch þ HCC i-p bend B2u 1008 1007 0.1
v36 ring torsion B2u 618 617.84 0.03
v37 Aromatic CH stretch B3g 3092 3093.02 0.03
v38 Aromatic CH stretch B3g 3060 3061.23 0.04
v39 CC stretch þ HCC i-p bend B3g 1624 1622.93 0.07
v40 HCC i-p bend B3g 1438 1437.15 0.06
v41 HCC i-p bend B3g 1239 1240.73 0.14
v42 CC stretch þ HCC i-p bend B3g 1158 1159.84 0.16
v43 CCC ring bend þ HCC i-p bend B3g 935 935.88 0.09
v44 CCC ring bend B3g 506 505.62 0.08
v45 HCC o-o-p bend B3u 958 957.81 0.02
v46 HCC o-o-p bend B3u 782 780.87 0.14
v47 ring torsion B3u 476 475.22 0.16
v48 ring torsion B3u 176 177.37 0.78
Note: i-p: in-plane, o-o-p: out-of-plane.
 x
ðx a Þ
 calc ¼ obsx obs calc  100:
Dx
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P ffi
 rms ¼ Nmodes
Dx 1
Nmodes ðx  obs  x  calc Þ2 ¼ 1.2834 cm1.

As the present work is anticipated at predicting the vibrational frequencies more accurately at
a significantly lower computational cost, we wish to mention the results obtained by few recent
studies on the vibrational spectra of naphthalene35,36 for comparison. Mackie et al.35 presented
the anharmonic theoretical spectra of naphthalene, computed with a locally modified version of
POLYCYCLIC AROMATIC COMPOUNDS 9

the SPECTRO program using Cartesian derivatives transformed from Gaussian 09 normal coord-
inate force constants. They have reported the theoretical anharmonic calculations for the gas
phase naphthalene deviated 6.5 cm1 (average) and 28 cm1 (maximum) from the experiment.
Their calculations with harmonic approximation differed by 10 cm1 at lower frequencies and
100 cm1 at higher frequencies. Chakraborty et al.36 reported the gas-phase infrared spectra of
naphthalene in the 3200 to 500 cm1 range and compared their experimental results with elec-
tronic structure calculations performed following two theoretical frameworks (MP2 and B3LYP)
and three different methods (VCCM, VSCF, and VPT2). They have observed that results vary sig-
nificantly from method to method, between the theoretical frameworks and across the spectral
region. They have estimated that the theoretically computed anharmonic vibrational frequencies
(fundamental) of naphthalene across different theoretical frameworks and methods mentioned
above are within a range of as low as 11.4 cm1 (for VCCM with MP2/6-311G(2d,2p)) to as high
as 37 cm1 (for VSCF with MP2/6-311G(2d,2p)) from the experiment. In the realm of ab initio
theory electronic structure calculations the vibrational frequencies reported in the above studies
are often considered highly accurate. However, as compared to those results, our U(2) Lie alge-
braic Hamiltonian predicted vibrational frequencies over the 48 fundamental modes deviated only
below 1.2834 cm1 from the experiment.
Though the current formulation is completely different from that of the ab initio method and
requires the experimental data of one-dimensional oscillators to compute the vibrational wave-
numbers, the U(2) Lie algebraic method deserves its accreditation for predicting the higher over-
tone frequencies at a much less computational effort. As far as the ab initio theoretical methods
are concerned, the computational cost is primarily determined by the number of basis functions.
Usually if one wishes to double the number of basis functions and perform the calculations, the
computational cost will be 16 times larger. As expressed by Chakraborty et al36 the highly accur-
ate VCCM calculations demand a huge memory space to store the effective Hamiltonian matrix
and the left and right eigenvectors. Apart from which diagonalization of such a matrix requires
large CPU time. Though the effective Hamiltonian matrix in their calculations is block diagonal-
ized for each symmetry irreducible representation, the size of matrix for naphthalene turns out to
be large and usually beyond the desktop computational resources.36 As compared to the above
studies, our U(2) Lie algebraic vibrational Hamiltonian for naphthalene claims minimal computa-
tional cost since no calculation of differential or integral terms of the potential surface is required,
and all the operations are in terms of algebraic manipulations only.
Vibrational spectra containing overtones and combinations at higher quantum numbers often
help to retrieve precise spectral parameters used to describe the intercoupling among the vibra-
tions. In this work, we have derived the excited vibrational frequencies of naphthalene molecule
up to the third overtone using the dynamical U(2) Lie algebraic Hamiltonian. The predicted over-
tone vibrational spectra (x  calc ) along with the mode assignments and symmetry species are given
in Table 3. It is seen that the predicted third overtone vibrations extend reaching energy up to
12,351.97 cm1, which occurs for the 4v37 (aromatic CH stretch) vibration with the symmetry
B3g. Up to the third overtone, various stretching and bending vibrations and their interactions in
naphthalene contributed about 192 energy level transitions in the infrared band which are distrib-
uted among the A1g, B1g, B2g, B3g, A1u, B1u, B2u, and B3u symmetries, and over the vibrational
quantum numbers 1, 2, 3, and 4. Based on their vibrational mode these 192 frequencies are segre-
gated into the following 9 vibrational groups: VG1 (aromatic CH stretch comprised of v1, v2, v17,
v18, v29, v30, v37, and v38), VG2 (CC stretch comprised of v3 and v5), VG3 (HCC i-p bend com-
prised of v4, v6, v20, v21, v22, v32, v33, v34, v40, and v41), VG4 (CC stretch þ HCC i-p bend com-
prised of v7, v19, v31, v35, v39, and v42), VG5 (CC stretch þ HCC i-p bend þ CCC bend comprised
of v8), VG6 (CCC ring bend comprised of v9, v24, and v44), VG7 (HCC o-o-p bend comprised of
v10, v11, v14, v15, v25, v26, v45, and v46), VG8 (ring torsion comprised of v12, v13, v16, v27, v28, v36,
v47, and v48), and VG9 (CCC ring bend þ HCC i-p bend comprised of v23, and v43). The very
10 M. R. BALLA AND V. JALIPARTHI

 calc ), along
Table 3. Overtone vibrational spectra of naphthalene (C10H8) as predicted by the U(2) Lie algebraic Hamiltonian (x
with vibrational mode assignments and symmetry species (irreducible representations).
Vibrational Mode Symmetry  calc (cm1)
x
2v1 Aromatic CH stretch Ag 5997.37
2v2 Aromatic CH stretch Ag 5983.26
2v3 CC stretch (major) Ag 2972.51
2v4 HCC i-p bend (major) Ag 2899.35
2v5 CC stretch (major) Ag 2678.77
2v6 HCC i-p bend (major) Ag 2271.93
2v7 CC stretch þ HCC i-p bend Ag 2007.25
2v8 CC stretch þ HCC i-p bend þ CCC bend Ag 1495.08
2v9 CCC ring bend Ag 989.88
2v10 HCC o-o-p bend Au 1924.18
2v11 HCC o-o-p bend Au 1626.36
2v12 ring torsion Au 1145.78
2v13 ring torsion Au 388.93
2v14 HCC o-o-p bend B1g 1848.94
2v15 HCC o-o-p bend B1g 1419.79
2v16 ring torsion B1g 745.12
2v17 Aromatic CH stretch B1u 6121.94
2v18 Aromatic CH stretch B1u 6102.44
2v19 CC stretch þ HCC i-p bend B1u 3186.07
2v20 HCC i-p bend B1u 2762.48
2v21 HCC i-p bend B1u 2525.08
2v22 HCC i-p bend B1u 2220.84
2v23 CCC ring bend þ HCC i-p bend B1u 1484.50
2v24 CCC ring bend B1u 707.96
2v25 HCC o-o-p bend B2g 1947.76
2v26 HCC o-o-p bend B2g 1732.86
2v27 ring torsion B2g 1529.05
2v28 ring torsion B2g 914.89
2v29 Aromatic CH stretch B2u 6177.92
2v30 Aromatic CH stretch B2u 6041.25
2v31 CC stretch þ HCC i-p bend B2u 3008.91
2v32 HCC i-p bend B2u 2714.94
2v33 HCC i-p bend B2u 2410.22
2v34 HCC i-p bend B2u 2259.74
2v35 CC stretch þ HCC i-p bend B2u 2009.59
2v36 ring torsion B2u 1221.58
2v37 Aromatic CH stretch B3g 6167.85
2v38 Aromatic CH stretch B3g 6106.58
2v39 CC stretch þ HCC i-p bend B3g 3211.74
2v40 HCC i-p bend B3g 2861.00
2v41 HCC i-p bend B3g 2458.66
2v42 CC stretch þ HCC i-p bend B3g 2304.07
2v43 CCC ring bend þ HCC i-p bend B3g 1868.97
2v44 CCC ring bend B3g 997.05
2v45 HCC o-o-p bend B3u 1907.77
2v46 HCC o-o-p bend B3u 1550.96
2v47 ring torsion B3u 934.75
2v48 ring torsion B3u 341.85
3v1 Aromatic CH stretch Ag 9025.08
3v2 Aromatic CH stretch Ag 8974.97
3v3 CC stretch (major) Ag 4683.08
3v4 HCC i-p bend (major) Ag 4313.96
3v5 CC stretch (major) Ag 3904.71
3v6 HCC i-p bend (major) Ag 3411.06
3v7 CC stretch þ HCC i-p bend Ag 2983.93
3v8 CC stretch þ HCC i-p bend þ CCC bend Ag 2236.67
3v9 CCC ring bend Ag 1518.91
3v10 HCC o-o-p bend Au 2879.89
3v11 HCC o-o-p bend Au 2506.02
3v12 ring torsion Au 1718.69
3v13 ring torsion Au 569.01
3v14 HCC o-o-p bend B1g 2797.61
(continued)
POLYCYCLIC AROMATIC COMPOUNDS 11

Table 3. Continued.
Vibrational Mode Symmetry  calc (cm1)
x
3v15 HCC o-o-p bend B1g 2138.05
3v16 ring torsion B1g 1142.95
3v17 Aromatic CH stretch B1u 9169.35
3v18 Aromatic CH stretch B1u 9144.09
3v19 CC stretch þ HCC i-p bend B1u 4773.00
3v20 HCC i-p bend B1u 4153.17
3v21 HCC i-p bend B1u 3742.02
3v22 HCC i-p bend B1u 3345.03
3v23 CCC ring bend þ HCC i-p bend B1u 2197.84
3v24 CCC ring bend B1u 1055.89
3v25 HCC o-o-p bend B2g 2925.22
3v26 HCC o-o-p bend B2g 2612.54
3v27 ring torsion B2g 2287.95
3v28 ring torsion B2g 1341.01
3v29 Aromatic CH stretch B2u 9260.08
3v30 Aromatic CH stretch B2u 9070.47
3v31 CC stretch þ HCC i-p bend B2u 4504.11
3v32 HCC i-p bend B2u 4065.61
3v33 HCC i-p bend B2u 3608.83
3v34 HCC i-p bend B2u 3407.03
3v35 CC stretch þ HCC i-p bend B2u 3013.99
3v36 ring torsion B2u 1843.07
3v37 Aromatic CH stretch B3g 9261.07
3v38 Aromatic CH stretch B3g 9169.88
3v39 CC stretch þ HCC i-p bend B3g 4864.24
3v40 HCC i-p bend B3g 4297.85
3v41 HCC i-p bend B3g 3703.61
3v42 CC stretch þ HCC i-p bend B3g 3455.59
3v43 CCC ring bend þ HCC i-p bend B3g 2791.52
3v44 CCC ring bend B3g 1497.83
3v45 HCC o-o-p bend B3u 2857.93
3v46 HCC o-o-p bend B3u 2317.48
3v47 ring torsion B3u 1401.02
3v48 ring torsion B3u 511.00
4v1 Aromatic CH stretch Ag 12070.07
4v2 Aromatic CH stretch Ag 12000.99
4v3 CC stretch (major) Ag 6275.31
4v4 HCC i-p bend (major) Ag 5800.82
4v5 CC stretch (major) Ag 5496.72
4v6 HCC i-p bend (major) Ag 4499.61
4v7 CC stretch þ HCC i-p bend Ag 4097.08
4v8 CC stretch þ HCC i-p bend þ CCC bend Ag 2892.79
4v9 CCC ring bend Ag 1988.71
4v10 HCC o-o-p bend Au 3868.46
4v11 HCC o-o-p bend Au 3315.47
4v12 ring torsion Au 2289.46
4v13 ring torsion Au 749.791
4v14 HCC o-o-p bend B1g 3739.89
4v15 HCC o-o-p bend B1g 2847.80
4v16 ring torsion B1g 1526.97
4v17 Aromatic CH stretch B1u 12192.74
4v18 Aromatic CH stretch B1u 12185.96
4v19 CC stretch þ HCC i-p bend B1u 6353.58
4v20 HCC i-p bend B1u 5524.96
4v21 HCC i-p bend B1u 5042.06
4v22 HCC i-p bend B1u 4484.49
4v23 CCC ring bend þ HCC i-p bend B1u 2986.17
4v24 CCC ring bend B1u 1424.92
4v25 HCC o-o-p bend B2g 3904.03
4v26 HCC o-o-p bend B2g 3489.92
4v27 ring torsion B2g 3060.39
4v28 ring torsion B2g 1817.97
4v29 Aromatic CH stretch B2u 12332.94
(continued)
12 M. R. BALLA AND V. JALIPARTHI

Table 3. Continued.
Vibrational Mode Symmetry  calc (cm1)
x
4v30 Aromatic CH stretch B2u 12102.54
4v31 CC stretch þ HCC i-p bend B2u 6018.00
4v32 HCC i-p bend B2u 5427.73
4v33 HCC i-p bend B2u 4811.61
4v34 HCC i-p bend B2u 4546.96
4v35 CC stretch þ HCC i-p bend B2u 4011.08
4v36 ring torsion B2u 2469.93
4v37 Aromatic CH stretch B3g 12351.97
4v38 Aromatic CH stretch B3g 12220.74
4v39 CC stretch þ HCC i-p bend B3g 6469.17
4v40 HCC i-p bend B3g 5743.27
4v41 HCC i-p bend B3g 4937.32
4v42 CC stretch þ HCC i-p bend B3g 4621.97
4v43 CCC ring bend þ HCC i-p bend B3g 3732.90
4v44 CCC ring bend B3g 2005.00
4v45 HCC o-o-p bend B3u 3813.90
4v46 HCC o-o-p bend B3u 3087.77
4v47 ring torsion B3u 1891.43
4v48 ring torsion B3u 698.06

purpose of such a segregation is to analyze the occupation of mode-wise vibrational levels over
the energy space.
Figure 2 shows the distribution of the dynamical U(2) Lie algebraic Hamiltonian predicted
vibrational frequencies of naphthalene which were segregated into the 9 vibrational groups (VG1
through VG9) and plotted as a function of their vibrational quantum number. The vibrational
groups are also color coded, according to whether the vibrations are aromatic CH stretch
(magenta), CC stretch (orange), bending (blue), or stretching-bending combination (olive), to
analyze their distribution over the energy space. It is clearly seen from the plotted data that vari-
ous bending vibrations (VG3, VG6, VG7, VG8 and VG9) occupy the lower end of the energy
space while the aromatic CH stretching vibrations (VG1) occupy the higher end. The stretching-
bending combination vibrations (VG4 and VG5) occupy the middle portion of the energy space.
Further, the aromatic CH stretching vibrations are well separated in energy space from those of
the CC stretching vibrations. All of these observations are highly consistent with (i) the general
principle of molecular spectroscopy that bending vibrations take less energy to excite while
stretching vibrations take a lot more, and (ii) the properties of the CH and CC bonds in naphtha-
lene (C10H8) molecule, due to the fact that the higher force constant and lower effective mass of
the CH species makes it to vibrate at higher frequencies; while the lower force constant and
higher effective mass of the CC species makes it to vibrate at lower frequencies.
The strength of the infrared transitions and band shapes in a molecule are usually determined
by the distribution of dipole moments along the principal vibrational coordinate axes and the
line positions. In naphthalene the near degeneracy of the CH and CC bond vibrational states also
appear to affect the band shapes. An important observation from the plotted data in Figure 2 is
the near degeneracy of vibrational energy levels within the same class of irreducible representa-
tions which would modify the shape of absorption bands in the infrared spectrum of naphthalene.
For example, the nearly degenerate bending vibrations belonging to the representations B1u (v23)
and B3u (v46) produce a major absorption band centered at 782 cm1 as observed in the infrared
spectrum of naphthalene.43 Similarly, the four aromatic CH stretching fundamental vibrations
belonging to the same class of representations B1u (v17, v18) and B2u (v29, v30) piled up in the
energy space to produce another major absorption band centered at 3065 cm1. The band centers
of these two major absorption bands, as observed in the experimental infrared spectrum,43 are
marked as P1 (782 cm1) and P2 (3065 cm1) in the plotted data (Figure 2). Apart from these
two major bands, near degeneracy of the overtone excitations are also expected to affect the band
POLYCYCLIC AROMATIC COMPOUNDS 13

Figure 2. U(2) Lie algebraic Hamiltonian predicted vibrational energy levels of naphthalene (C10H8) as a function of vibrational
quantum number (V)(right panel). The vibrational frequencies are grouped (VG1 through VG9) and color coded for analysis of
their distribution on the energy space (please see section 5 for details). The left panel shows infrared transmission spectrum of
naphthalene in the gas phase plotted from the experimental standard reference database 69 (Courtesy of NIST Chemistry
WebBook). Also marked are the band centers of two major absorption bands P1 (752 cm1) and P2 (3065 cm1), as observed in
the measured spectrum.43

shapes in several absorption bands of medium to small intensity below the wavenumbers up to
4000 cm1. However, this effect fades away with higher vibrational quanta, due to the dilution of
degeneracy as can be seen from the plotted data in Figure 2, and the reduced dipole moment
with respect to the overtone and/combinations are expected to result in several very weak absorp-
tion bands in a wide range of spectral coverage. Thus, it is clear from the above discussion that
the distribution of dipole moments along the principal vibrational coordinate axes as well as the
observed near degeneracy of vibrational energy states from the lower quanta excitations basically
characterize the infrared spectrum of naphthalene in the gas phase.

5. Conclusions
The following conclusions may be drawn from the content presented in this paper and the dis-
cussion thereof:

1. This article is the first to demonstrate the application of dynamical U(2) Lie algebras to
model a vibrational Hamiltonian of naphthalene (C10H8), a prototype polycyclic aromatic
hydrocarbon molecule containing a fused pair of benzene rings, in the gas phase having con-
siderable astrophysical interest.
2. A vibrational Hamiltonian that preserves the D2h symmetry of naphthalene is modeled using
19 coupled one-dimensional Morse oscillators to describe the eight C-H single bonds, the
eleven C-C aromatic bonds, and their dynamic interactions, making use of the Casimir and
Majorana operators.
14 M. R. BALLA AND V. JALIPARTHI

3. Results indicate that the dynamical U(2) Lie algebraic Hamiltonian has successfully repro-
duced the vibrational frequencies at the 48 fundamental excitations near to the level of spec-
troscopic accuracy, as well as predicted their first three overtones.
4. The distribution of dipole moments along the principal vibrational coordinate axes as well as
the observed degeneracy of the stretching, bending, and combination vibrations at the funda-
mental and overtone excitation levels are expected to produce two major bands (centered at
782 cm1 and 3065 cm1), and several minor bands in the infrared absorption spectrum of
the gas phase naphthalene.
5. The degeneracy effect, however, fades away with higher vibrational quanta and the reduced
dipole moments with respect to the overtone and/combinations are expected to result in sev-
eral fragile absorption bands in a wide range of spectral coverage.

Acknowledgements
The authors thank the topical editor Prof. Philippe Garrigues and the anonymous reviewers for valuable sugges-
tions. Acknowledgments are due to National Institute of Science and Technology (https://webbook.nist.gov) for
providing the experimental reference spectrum database 69.

Disclosure statement
There are no conflicts of interest.

ORCID
Malleswara Rao Balla http://orcid.org/0000-0003-2700-4215
Vijayasekhar Jaliparthi http://orcid.org/0000-0002-2745-7401

References
1. J. L. Dunham, “The Energy Levels of a Rotating Vibrator,” Physical Review 41, no. 6 (1932): 721–31..
2. M. Molski, “Spectral Expansion of the Rovibrational Energy of Diatomic Molecules Described by Morse
Potential,” Acta Physica Polonica A 84, no. 6 (1993): 1041–8.
3. M.A. Mroginski, “QM/MM Calculations of Vibrational Spectra,” in Encyclopaedia of Biophysics, edited by
G. C. K. Roberts, 2013 ed. (Berlin: Springer, 2013), 56–73.
4. R. Bijker, A. Frank, and R. Lemus, J.M. Arias, and F. Perez-Bernal, “A Comparison Between Algebraic
Models of Molecular Spectroscopy,” in Symmetries in Science X, edited by B. Gruber and M. Ramek.
(Boston: Springer, 1998), 37–46.
5. F. Iachello, “Algebraic Methods for Molecular Rotation-Vibration Spectra,” Chemical Physics Letters 78, no.
3 (1981): 581–5.
6. F. Iachello, and R. D. Levine, “Algebraic Approach to Molecular Rotation-Vibration Spectra. I. Diatomic
Molecules,” The Journal of Chemical Physics 77, no. 6 (1982): 3046–55.
7. O. S. van Roosmalen, A. E. L. Dieperink, and F. Iachello, “A Dynamic Algebra for Rotation-Vibration
Spectra of Complex Molecules,” Chemical Physics Letters 85, no. 1 (1982): 32–6.
8. O. S. van Roosmalen, F. Iachello, R. D. Levine, and A. E. L. Dieperink, “A Dynamic Algebra for Rotation-
Vibration Spectra of Complex Molecules,” The Journal of Chemical Physics 79, no. 6 (1983): 2515–36.
9. O. S. van Roosmalen, R. D. Levine, and A. E. L. Dieperink, “The Geometrical-Classical Limit of Algebraic
Hamiltonians for Molecular Vibrotational Spectra,” Chemical Physics Letters 101, no. 6 (1983): 512–7.
10. O. S. van Roosmalen, I. Benjamin, and R. D. Levine, “A Unified Algebraic Model Description for Interacting
Vibrational Modes in ABA Molecules,” The Journal of Chemical Physics 81, no. 12 (1984): 5986–97.
11. F. Iachello, and S. Oss, “Model of n Coupled Anharmonic Oscillators and Applications to Octahedral
Molecules,” Physical Review Letters 66, no. 23 (1991): 2976–97.
12. A. Frank, and R. Lemus, “Comment on "Model of n coupled anharmonic oscillators and applications to
octahedral molecules",” Physical Review Letters 68, no. 3 (1992): 413.
13. L. Wiesenfeld, “The Vibron Model for Methane: Stretch–Bend Interactions,” Journal of Molecular
Spectroscopy 184, no. 2 (1997): 277–87.
POLYCYCLIC AROMATIC COMPOUNDS 15

14. X. W. Hou, Y. Z. Ding, and Z. Q. Ma, “Algebraic Model for Stretching and Bending Vibrations of Bent
Triatomic Molecules,” International Journal of Theoretical Physics 38, no. 3 (1999): 985–91. no.
15. J. Chen, F. Iachello, and J. Ping, “The Method of Symmetrized Bosons with Applications to Vibrations of
Octahedral Molecules,” The Journal of Chemical Physics 104, no. 3 (1996): 815–25.
16. R. Sen, A. Kalyan, R. S. Paul, N. K. Sarkar, and R. Bhattacharjee, “A Study of Vibrational Spectra of
Fullerene C70 and C80: An Algebraic Approach,” Acta Physica Polonica A 120, no. 3 (2011): 407–11.
17. R. Sen, A. Kalyan, R. Das, N. K. Sarkar, and R. Bhattacharjee, “Vibrational Frequencies of
Buckminsterfullerene: An Algebraic Study,” Spectroscopy Letters 45, no. 4 (2012): 273–9.
18. K. S. Rao, V. U. M. Rao, and J. Vijayasekhar, “Spectroscopic Studies of Distorted Structure Bio-Nano
Molecules,” Procedia Materials Science 10, (2015): 737–47.
19. R. Sen, A. Kalyan, N. K. Sarkar, and R. Bhattacharjee, “Application of U(2) Algebraic Model in the Study of
Stretching Vibrational Spectra of Large Fullerenes C180 and C240,” Fullerenes, Nanotubes and Carbon
Nanostructures 21, no. 8 (2013): 725–79.
20. Srinivasa Rao Karumuri, J. Vijaya Sekhar, V. Sreeram, V. Uma Maheswara Rao, and M. V. Basaveswara Rao,
“Spectroscopic Studies on Distorted Structure Molecules by Using U(2) Lie Algebraic Method,” Journal of
Molecular Spectroscopy 269, no. 1 (2011): 119–23.
21. Srinivasa Rao Karumuri, Vijayasakhar Jallaparthi, and Sreeram Venigalla, “Spectroscopic Studies of Distorted
Structure Systems in the Vibron Model: Application to Porphyrin and Its Isotopomers,” International
Journal of Spectroscopy 2011, (2011): 1–5.
22. S. R. Karumuri, “A Study of Vibrational Spectra of Metallotetraphenyl Porphryins: An Algebraic Approach,”
Indian Journal of Physics 86, no. 12 (2012): 1147–53.
23. K. S. Rao, G. Srinivas, J. Vijayasekhar, V. U. M. Rao, Y. Srinivas, K. S. Babu, V. S. S. Kumar, and A.
Hanumaiah, “Analysis of Vibrational Spectra of Nano-Bio Molecules: Application to Metalloporphyrins,”
Chinese Physics B 22, no. 9 (2013): 090304.
24. S. Rao Karumuri, K. Girija Sravani, J. Vijayshekar, and L. S. S. Reddy, “Spectroscopic Studies on Distorted
Structure Nano Molecules by Lie Algebraic Model,” Acta Physica Polonica A 122, no. 1 (2012): 49–52.
25. S. A. Sandford, M. P. Bernstein, and L. J. Allamandola, “The Mid-Infrared Laboratory Spectra of
Naphthalene (C10H8) in Solid H2O,” The Astrophysical Journal 607, no. 1 (2004): 346–60.
26. F. Pauzat, D. Talbi, M. D. Miller, D. J. DeFrees, and Y. Ellinger, “Theoretical IR Spectra of Ionized
Naphthalene,” The Journal of Physical Chemistry 96, no. 20 (1992): 7882–6.
27. S. R. Langhoff, “Theoretical Infrared Spectra for Polycyclic Aromatic Hydrocarbon Neutrals, Cations, and
Anions,” The Journal of Physical Chemistry 100, no. 8 (1996): 2819–41.
28. W. J. Hehre, L. Radom, P. v. R. Schleyer, and J. A. Pople, Ab Initio Molecular Orbital Theory (New York:
John Wiley, 1986), 548.
29. J. Szczepanski, and M. Vala, “Laboratory Evidence for Ionized Polycyclic Aromatic Hydrocarbons in the
Interstellar Medium,” Nature 363, no. 6431 (1993): 699–701.
30. B. H. Foing, and P. Ehrenfreund, “Detection of Two Interstellar Absorption Bands Coincident with Spectral
Features of C60,” Nature 369, no. 6478 (1994): 296–8.
31. D. M. Hudgins, S. A. Sandford, and L. J. Allamandola, “Infrared Spectroscopy of Polycyclic Aromatic
Hydrocarbon Cations. 1. Matrix-Isolated Naphthalene and Perdeuterated Naphthalene,” The Journal of
Physical Chemistry 98, no. 16 (1994): 4243–53.
32. N. Rougeau, J. P. Flament, P. Youkharibache, H. P. Gervais, and G. Berthier, “Vibrational Modelling in
Large Polycyclic Aromatic Hydrocarbons,” Journal of Molecular Structure: THEOCHEM 254, no. 12 (1992):
405–28.
33. M. Ibrahim, A. Nadamm, and D. Eldin, “Density Functional Theory and FTIR Spectroscopic Study of
Carboxyl Group,” Indian Journal of Pure and Applied Physics 43, no. 12 (2005): 911–7. no. http://hdl.han-
dle.net/123456789/8906
34. A. Srivatsava, and V. B. Singh, “Theoretical and Experimental Studies of Vibrational Spectra of Naphthalene
and Its Cation,” Indian Journal of Pure and Applied Physics 45, (2017): 714–20. http://hdl.handle.net/
123456789/2661
35. Cameron J. Mackie, Alessandra Candian, Xinchuan Huang, Elena Maltseva, Annemieke Petrignani, Jos
Oomens, Wybren Jan Buma, Timothy J. Lee, and Alexander G. G. M. Tielens, “The Anharmonic Quartic
Force Field Infrared Spectra of Three Polycyclic Aromatic Hydrocarbons: Naphthalene, Anthracene, and
Tetracene,”The Journal of Chemical Physics 143, no. 22 (2015): 224314
36. S. Chakraborty, S. Banik, and P. K. Das, “Anharmonicity in the Vibrational Spectra of Naphthalene and
Naphthalene-d8: Experiment and Theory,” The Journal of Physical Chemistry. A 120, no. 49 (2016): 9707–18.
37. J. M. L. Martin, J. El-Yazal, and J.-P. Francois, “Structure and Vibrational Spectrum of Some Polycyclic
Aromatic Compounds Studied by Density Functional Theory. 1. Naphthalene, Phenanthrene, and
Anthracene,” The Journal of Physical Chemistry 100, no. 38 (1996): 15358–67. no.
38. F. Iachello and R. D. Levine, Algebraic theory of molecules (Oxford University Press, Oxford, 1995).
16 M. R. BALLA AND V. JALIPARTHI

39. S. Oss, “Algebraic Models in Molecular Spectroscopy,” in Advances in Chemical Physics, edited by I.
Prigogine and Stuart A. Rice (Oxford, UK: John Wiley Sons, Inc., 1996), 455–649.
40. K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Compounds: Part A: Theory and
Applications in Inorganic Chemistry, 6th ed. (New York: Wiley, 2009).
41. K. P. Huber, and G. Herzberg, Molecular Spectra and Molecular Structure. IV: Constants of Diatomic
Molecules (New York: Van Nostrand Reinhold, 1979).
42. M. R. Balla, and V. S. Jaliparthi, “Vibrational Hamiltonian of Methylene Chloride Using U(2) Lie Algebra,”
Molecular Physics 119, no. 5 (2021): e1828634.
43. P. J. Linstorm, and W. G. Mallard (Eds.), NIST Chemistry WebBook, NIST Standard Reference Database
Number 69, National Institute of Standards and Technology, Gaithersburg MD, 20899 (retrieved October 29,
2020).

You might also like