You are on page 1of 15

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965

Contents lists available at ScienceDirect

Spectrochimica Acta Part A: Molecular and


Biomolecular Spectroscopy
journal homepage: www.elsevier.com/locate/saa

Structure–activity relations of 2-(methylthio)benzimidazole by FTIR,


FT-Raman, NMR, DFT and conceptual DFT methods
V. Arjunan a,⇑, Arushma Raj a, P. Ravindran b, S. Mohan c
a
Department of Chemistry, Kanchi Mamunivar Centre for Post-Graduate Studies, Puducherry 605 008, India
b
Department of Chemistry, Mahatma Gandhi Government Arts College, Mahe 673 310, India
c
School of Sciences and Humanities, Vel Tech University, Avadi, Chennai 600 062, India

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 The vibrational fundamentals of 2- The vibrational fundamental modes of 2-(methylthio)benzimidazole have been analysed by combining
(methylthio)benzimidazole (2MTBI) FTIR, FT-Raman and quantum chemical calculations. The structural parameters of the compound were
were analysed. determined from the optimised geometry by B3LYP with 6-31G⁄⁄, 6-311++G⁄⁄ and cc-pVTZ basis sets
1 13
 H and C NMR chemical shifts were and giving energies, harmonic vibrational frequencies, depolarisation ratios, IR intensities and Raman
calculated using GIAO method and activities. The 1H and 13C nuclear magnetic resonance chemical shifts of the molecule were analysed.
interpreted. The chemical reactivity and site selectivity of the molecule has been determined with the help of global
 The stable molecular geometry was and local reactivity descriptors.
found by conformational analysis.
 The pC8–C9 ? pC4—C5 interaction is
strongly stabilized by
261.47 kJ mol1.
 The structure–activity relationships
were investigated by conceptual DFT
methods.

a r t i c l e i n f o a b s t r a c t

Article history: The vibrational fundamental modes of 2-(methylthio)benzimidazole (2MTBI) have been analysed by
Received 30 July 2013 combining FTIR, FT-Raman and quantum chemical calculations. The structural parameters of the com-
Received in revised form 11 September pound are determined from the optimised geometry by B3LYP with 6-31G⁄⁄, 6-311++G⁄⁄ and cc-pVTZ
2013
basis sets and giving energies, harmonic vibrational frequencies, depolarisation ratios, IR intensities
Accepted 26 September 2013
Available online 8 October 2013
and Raman activities. 1H and 13C NMR spectra have been analysed and 1H and 13C nuclear magnetic res-
onance chemical shifts are calculated using the gauge independent atomic orbital (GIAO) method. The
structure–activity relationship of the compound is also investigated by conceptual DFT methods. The
Keywords:
FTIR
chemical reactivity and site selectivity of the molecule has been determined with the help of global
FT-Raman and local reactivity descriptors.
2-(Methylthio)benzimidazole Ó 2013 Elsevier B.V. All rights reserved.
DFT
NMR
Reactivity descriptors

⇑ Corresponding author. Tel.: +91 413 2211111, mobile: +91 9442992223; fax: +91 413 2251613.
E-mail address: varjunftir@yahoo.com (V. Arjunan).

1386-1425/$ - see front matter Ó 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.saa.2013.09.100
952 V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965

Introduction

Metal complexes of biologically important ligands are some-


times more effective than free ligands. Benzimidazole acts as a
ligand towards transition metal ions with a variety of biological
molecules including nonheme systems, Vitamin B12 and its deriv-
atives and several metallo-proteins [1,2]. Benzimidazole is also of
interest as a corrosion inhibitor for metals and alloys. Rao and Ven-
kataraghavan [3] reported the infrared spectra of five membered
heterocyclic derivatives such as triazoles, thiazoles, and thiadiz-
oles. They have also attempted empirical assignments for benzo-
thiazole derivatives. The infrared and Raman spectra of
benzimidazole have been reported by Suwaiyan et al. [4].
2-Aminothiazoles involved in numerous applications, including
human and veterinary medicine [5,6]. The scaled quantum chemi-
cal studies of the structure and vibrational spectra of 2-(methyl-
thio)benzimidazole have been reported by Krishnakumar and
Ramasamy [7]. But several points were wrongly reported and dis-
cussed in the manuscript. Few of them are (i) the conformers of the
molecule (Fig. 1 of the reported manuscript [7]) were wrongly re-
ported as s-cis and s-trans, (ii) the unbelievable theoretical spectra
were found in Fig. 2 of the reported manuscript [7], (iii) the for-
mula used to determine the root–mean square deviation is wrong,
(iv) the fundamental modes of methyl group (as well as for any AB3
and AB2 systems) are usually named as asymmetric stretching,
symmetric stretching, asymmetric deformation, symmetric defor-
mation, rocking, wagging and twisting but the nomenclature for
the methyl group vibrations were wrongly reported as out-of-
plane stretching (CH3 ops), in-plane stretching (CH3 ips), in-plane
bending (CH3 ipb), symmetric bending (CH3 sb), out-of-plane rock-
ing (CH3 opr) etc, (v) most of the nomenclature used in Table 5 of
the reported manuscript namely (ring symmetrical deformation
(Rsymd), ring trigonal deformation (Rtrig), ring asymmetric torsion
(tRasy), wagging CAS (xCS), etc. are also wrong. (vi) The CAS
stretching vibration of the compound is wrongly assigned at
1267/1268 cm1 (This region belongs to C@S stretching mode
but not to CAS stretch) and (vii) In Table 5 of the reported manu-
script, they quoted that ‘‘Relative Raman intensities calculated by
Eq. (1) and normalised to 100’’ but no mode has the intensity (Ii)
as 100. Thus the adopted procedure for normalisation is wrong.
Consideration of all these factors and owing to the biological
significance of the compound motivated us to undertake afresh
vibrational spectroscopic and structure–activity relationship

Fig. 2. Potential energy profile and the possible conformations of 2-


(methylthio)benzimidazole.

studies of 2-(methylthio)benzimidazole (2MTBI). Thus, in the pres-


ent manuscript all the above said discrepancies are rectified and
more studies like detailed conformational analysis, natural bond
orbital analysis, determination of the molecular electrostatic po-
tential, various types of donor–acceptor interactions and their sta-
bilisation, NMR investigations, the global and local reactivity
descriptors were included for the benefit of the researchers.

Experimental

The compound 2-(methylthio)benzimidazole (2MTBI) was pur-


chased from Aldrich chemicals, USA, and is used as such to record
the FTIR, FT-Raman and NMR spectra. The FTIR spectrum of the
compound was recorded by KBr pellet method on a Bruker IFS
66 V spectrometer equipped with a Globar source, Ge/KBr beam
splitter and a TGS detector in the range of 4000–400 cm1. The
spectral resolution is 2 cm1. The FT-Raman spectrum of the com-
Fig. 1. Potential energy surface of 2-(methylthio)benzimidazole. pound was also recorded in the range 4000–100 cm1 using the
V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965 953

where ZA is the charge on nucleus A, located at RA and q(r0 ) is the


electronic density function for the molecule. The first and second
terms represent the contributions to the potential due to nuclei
and electrons, respectively. V(r) is the resultant electrical potential
at each point r, which is the net electrostatic effect produced at
the point r by both the electrons and nuclei of the molecule. The
molecular electrostatic potential (MEP) serves as a useful quantity
to explain hydrogen bonding, reactivity and structure–activity rela-
tionship of molecules including biomolecules and drugs [20]. The
geometry resulting from the plot of electron density surface
mapped with electrostatic potential surface depict the shape, size,
charge density distribution and the sites of chemical reactivity of
a molecule. GaussView 5.0.8 visualisation program [21] has been
utilised to construct the MEP surface, the shape of highest occupied
molecular orbital (HOMO) and lowest unoccupied molecular orbital
Fig. 3. The optimised geometry of 2-(methylthio)benzimidazole with scheme of
(LUMO) orbitals.
atom numbering obtained by B3LYP/6-311++G⁄⁄ method.
The 1H and 13C NMR isotropic shielding constants are calcu-
lated using the GIAO method [22] with the optimised parameters
same instrument with FRA 106 Raman module equipped with obtained from B3LYP/6-311++G⁄⁄ method. The effect of solvent
Nd:YAG laser source operating at 1.064 lm with 200 mW powers. on the theoretical NMR parameters is included using the PCM
A liquid nitrogen cooled–Ge detector was used. The frequencies of model. The isotropic shielding constant values are used to calcu-
all sharp bands are accurate to 2 cm1. The 1H (400 MHz; CDCl3) late the isotropic chemical shifts d with respect to tetramethylsil-
and 13C (100 MHz; CDCl3) nuclear magnetic resonance (NMR) ane (TMS) using the relation diso(X) = risoTMS(X)  riso(X), where
spectra were recorded on a Bruker HC400 instrument. Chemical diso – isotropic chemical shift and riso – isotropic shielding
shifts for protons are reported in parts per million scales (d scale) constant.
downfield from tetramethylsilane. Various reactivity and selectivity descriptors such as chemical
hardness, chemical potential, softness, electrophilicity, nucleophi-
licity and the appropriate local quantities employing natural pop-
Computational details
ulation analysis (NPA) are calculated. Both the global and local
reactivity descriptors are determined using finite difference
The molecular structure of 2MTBI is fully optimised with DFT
approximation to reveal the intramolecular reactivity of the mole-
method by using the standard 6-31G⁄⁄, 6-311++G⁄⁄ and high level
cule. The vertical ionisation potential (I), electron affinity (A) and
cc-pVTZ basis sets. The effects of electron correlation on the geom-
the electron populations were determined on the basis of B3LYP/
etry optimisation have been taken into account. Theoretically the
cc-pVTZ method. The energy of the N-electron species is deter-
vibrational wavenumbers and their intensities are calculated in or-
mined by using restricted DFT method while the energies of the
der to provide more information for the vibrational assignments of
N1 and N+1 electronic species are calculated using open shell re-
2MTBI. The quantum chemical calculations have been performed
stricted DFT method using the B3LYP/cc-pVTZ optimised geometry
with Gaussian-09 [8] program, invoking gradient geometry optimi-
of the N-electron species. The site-selectivity of a chemical system,
sation [9]. The density functional theory (DFT) [10] with the three
can be determined by using Fukui function [23,24] which can be
parameter hybrid functional (B3) [11,12] for the exchange part and
interpreted either as the change of electron density q(r) at each
the Lee–Yang–Parr (LYP) correlation functional [13] have been uti-
point r when the total number of electrons is changed or as the
lised for the computation of molecular structure optimisation,
sensitivity of chemical potential (l) of a system to an external per-
vibrational frequencies, thermodynamic properties and energies
turbation at a particular point r.
of the optimised structures. The force constants obtained from
the 6-311++G⁄⁄ basis set are utilised in the normal coordinate anal-
   
@ qðrÞ dl
ysis using the Wilson’s FG matrix method [14–16] with the pertur- f ðrÞ ¼ ¼
@N v ðrÞ dv ðrÞ N
bation program written by Furher et al. [17].
The Raman scattering activities (Si) calculated by Gaussian 09W Yang and Parr in 1985 introduced local softness s(r) to predict the
program were suitably converted to relative Raman intensities (Ii) reactivity [25]. The s(r) describes the sensitivity of the chemical po-
using the following relationship derived from the basic theory of tential of the system to the local external perturbation and is ob-
Raman scattering [18]. tained by simply multiplying Fukui function f(r) with global
softness S. Though s(r) and f(r) contain the same informations, the
f ðm0  mi Þ4 Si
Ii ¼ ð1Þ local softness values are generally used in predicting the reactivities
mi ½1  expðhcmi =kTÞ such as electrophilic, nucleophilic and free radical reactions, regi-
where v0 is the exciting frequency (cm1), vi is the vibrational wave- oselectivity etc.
number of the ith normal mode, h, c and k are universal constants,  
@ qðrÞ
and f is the suitably chosen common scaling factor for all the peak sðrÞ ¼ @l
v ðrÞ
intensities. sðrÞ ¼ f ðrÞS
Isoelectronic molecular electrostatic potential surfaces (MEP)
and electron density surfaces [19] are developed by using 6- where S is the global softness which is inversely related to global
311++G⁄⁄ basis set. The molecular electrostatic potential (MEP) at hardness (g).
a point ‘r’ in the space around a molecule (in atomic units) can
be expressed as:
Conformational analysis
X Z 0
ZA qð~
r 0 Þdr
VðrÞ ¼  Molecular geometry is a sensitive indicator of intra and inter-
A j~
RA  ~
rj j~
r0  ~
rj
molecular interactions. Conformational analysis of the compound
954 V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965

Table 1
Structural parameters of 2-(methylthio)benzimidazole calculated by B3LYP/6-31G⁄⁄, B3LYP/6-311G++⁄⁄ and B3LYP/cc-pVTZ methods.

Structural Parameters B3LYP/6-31G⁄⁄ B3LYP/6-311G++⁄⁄ B3LYP/cc-pVTZ Experimentala


Internuclear distance (Å)
N1AC2 1.38 1.38 1.38 1.35
N1AC9 1.39 1.39 1.38 1.40
N1AH14 1.01 1.01 1.00 0.90
C2AN3 1.31 1.31 1.30 1.31
C2AS15 1.76 1.76 1.76
N3AC8 1.39 1.39 1.39 1.37
C4AC5 1.39 1.39 1.39 1.38
C4AC8 1.40 1.40 1.39 1.39
C4AH10 1.09 1.08 1.08 0.94
C5AC6 1.41 1.41 1.40 1.40
C5AH11 1.09 1.08 1.08 1.02
C6AC7 1.39 1.39 1.39 1.39
C6AH12 1.09 1.08 1.08 1.07
C7AC9 1.39 1.39 1.39 1.40
C7AH13 1.09 1.08 1.08 0.98
C8AC9 1.42 1.41 1.41 1.39
S15AC16 1.83 1.83 1.82
C16AH17 1.09 1.09 1.09
C16AH17 1.09 1.09 1.09
C16AH19 1.09 1.09 1.09
Bond Angle (°)
C2AN1AC9 106.7 106.7 106.8 106.6
C2AN1AH14 126.3 126.2 126.1
C9AN1AH14 127.0 127.0 127.0 121.0
N1AC2AN3 113.6 113.4 113.3 114.0
N1AC2AS15 120.0 120.0 119.9
N3AC2AS15 126.3 126.6 126.7
C2AN3AC8 104.7 105.0 105.0 104.2
C5AC4AC8 118.0 118.0 118.0 117.8
C5AC4AH10 121.7 121.6 121.6
C8AC4AH10 120.2 120.3 120.4 117.0
C4AC5AC6 121.4 121.4 121.5 120.9
C4AC5AH11 119.5 119.5 119.5 117.0
C6AC5AH11 119.1 119.1 119.0
C5AC6AC7 121.5 121.4 121.4 122.3
C5AC6AH12 119.4 119.4 119.3 111.0
C7AC6AH12 119.2 119.2 119.3
C6AC7AC9 116.7 116.8 116.8 116.1
C6AC7AH13 121.3 121.2 121.2 121.0
C9AC7AH13 122.0 122.0 122.0
N3AC8AC4 129.8 129.9 130.0
N3AC8AC9 110.5 110.3 110.3 109.5
C4AC8AC9 119.7 119.8 119.8 120.6
N1AC9AC7 132.9 132.9 132.9
N1AC9AC8 104.5 104.6 104.6 105.8
C7AC9AC8 122.7 122.6 122.6 122.4
C2AS15AC16 99.1 99.5 99.5
S15AC16AH17 110.1 110.2 110.1
S15AC16AH18 105.9 105.5 105.7
S15AC16AH19 110.1 110.2 110.1
H17AC16AH18 110.5 110.4 110.5
H17AC16AH19 109.7 110.1 109.9
H18AC16AH19 110.5 110.4 110.5
a
Values taken from Ref. [26].

2MTBI is carried out by using B3LYP/6-31G⁄⁄ method. All possible Hartree. The conformer ‘a’ is more stable than the conformer ‘b’ by
geometry of the conformers are optimised to find out the 3.54 kcal mol1. The transition state of the compound having en-
energetically and thermodynamically most stable configuration ergy 3.65 kcal mol1 is the least stable. The energy difference be-
of the compound. The initial value of the dihedral angle tween the conformer ‘b’ and the third conformer ‘c’ is only
N1AC2AS15AC16 is set to zero and for every 15° the energies of 0.11 kcal mol1.
the conformers are determined. The potential energy surface dia-
gram of 2MTBI obtained by the rotation of the methylthio group
along the dihedral angle N1AC2AS15AC16 is presented in Fig. 1 Results and discussion
and the possible conformations of the compound are shown in
Fig. 2. The compound 2MTBI has three different conformers. The Structural properties
stability of the conformer is in the order a > b > c. The points corre-
sponds to geometry ‘b’ is the ‘saddle point’ while the geometry cor- The optimised geometry and the scheme of numbering the
responds to the geometry ‘c’ is the transition state. Thus, the atoms of 2MTBI are shown in Fig. 3. The optimised structural
molecule has only one stable configuration. The energy of the con- parameters bond length and bond angle for the thermodynami-
former ‘a’ determined by the B3LYP/6-31G⁄⁄ method is 817.4031 cally preferred geometry determined at B3LYP method with
V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965 955

Table 2
The thermodynamic parameters calculated by B3LYP method with 6-31G⁄⁄, 6-311++G⁄⁄ and cc-pVTZ basis sets, energies of frontier molecular orbitals and global reactivity
descriptors of 2-(methylthio)benzimidazole.

Thermodynamic parameters (298 K) B3LYP/6-31G⁄⁄ B3LYP/6-311++G⁄⁄ B3LYP/cc-pVTZ


SCF Energy (a.u) 817.2244429 817.5099813 817.556294
Total energy (thermal), Etotal (kcal mol1) 98.30 98.07 97.92
Heat capacity at const. volume, Cv (cal mol1 K1) 36.51 36.63 36.46
Entropy, S (cal mol1 K1) 97.92 97.61 97.13
Vibrational energy, Evib (kcal mol1) 96.52 96.30 96.14
Zero-point vibrational Energy, E0 (kcal mol1) 92.25 92.02 91.91
Rotational constants (GHz)
A 3.04 3.04 3.07
B 0.56 0.56 0.56
C 0.47 0.47 0.48
Dipolemoment (Debye)
lx 0.63 0.76 0.86
ly 1.60 1.68 1.68
lz 0.00 0.00 0.00
ltotal 1.72 1.85 1.89
ELUMO+1 (eV) 0.4863
ELUMO (eV) 0.6710
EHOMO (eV) 5.9022
EHOMO1 (eV) 6.4127
ELUMO  EHOMO (eV) 5.2312
Ionisation potential, I (eV) 7.6192
Electron affinity, A (eV) 0.3218
Electronegativity (v) 3.6487
Chemical potential (l) 3.6487
Electrophilicity (x) 1.6765
Hardness (g) 3.9705
Softness (S) 0.1259
Electrofugality (DEe) 9.2956
Nucleofugality (DEn) 1.9983

Fig. 4. The total electron density mapped with electrostatic potential surface of 2-
(methylthio)benzimidazole.

6-31G⁄⁄, 6-311++G⁄⁄ and cc-pVTZ basis sets are summarised in Ta- Fig. 5. The contour map of electrostatic potential of 2-(methylthio)benzimidazole.
ble 1. For correlation of the theoretical values with that of the
experimental data, only B3LYP/cc-pVTZ method has been consid-
ered, because the energy of the molecule determined by this meth- of double bond. The bond angles C8AC4AC5 and C4AC8AC9 are
od is small and thus the geometry is considered to be more stable. less than 120° while the other CCC bond angles are more than
The optimised parameters of 2MTBI are compared with X-ray dif- 120°. The bond angles in the hetero ring show more distortions
fraction data of benzimidazole [26] to show the substituent effects. and are due to the ring strain and the presence of substitutions
Closer examination of the bond lengths and angles of 2MTBI in Ta- in the C2 carbon atom. The bond angle at C2 carbon (N1AC2AN3)
ble 1 relative to the XRD values shows that larger deviations in the is more (113.3°) than the other bond angles of the hetero ring and
bond lengths of NAH and CAH are observed. There is no significant is due to the electron donating substitutions.
deviations are observed between the CAC bond lengths of the aro- The thermodynamic parameters of the compound total thermal
matic ring and the experimental
0
values. The S15AC16 bond length energy, vibrational energy contribution to the total energy, the
of 2MTBI is longer 1.82 ÅA, where the ACH3 group is attached. The rotational constants and the dipole moment values determined
shorter bond length of C2AN3 than N1AC2 is due to the presence by B3LYP method with 6-31G⁄⁄, 6-311++G⁄⁄ and cc-pVTZ basis sets
956 V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965

Fig. 6. The electrostatic potential surface of 2-(methylthio)benzimidazole.

Fig. 8. (a) Experimental FTIR, (b) Theoretical B3LYP/cc-pVTZ and (c) B3LYP/6-
311++G⁄⁄ infrared spectra of 2-(methylthio)benzimidazole.

Intermediary colours represent intermediary electrostatic potentials.


Fig. 7. The Frontier molecular orbitals of 2-(methylthio)benzimidazole.
Areas of low potential, red, are characterised by an abundance of
are presented in Table 2. The energy of the compound determined electrons. Areas of high potential, blue, are characterised by a rela-
by B3LYP/cc-pVTZ method is 817.5563 Hatrees. The total dipole tive absence of electrons. Figs. 4 and 5 indicates that the region
moment of the molecule determined by B3LYP/cc-pVTZ method around the nitrogen atom (N3) linked with carbon through the dou-
is 1.89 D reveals the high polarity of the molecule. The energies ble bond is the most negative potential region (red) and the hydro-
of HOMO and LUMO are given in Table 2 signify the possibility of gen atom linked with nitrogen (N1) is the most bang of positive
the electronic transition between the molecular orbitals and the region (blue). The predominance of light green region in the MEP
stability of the molecule. surfaces of the aromatic ring corresponds to a potential halfway be-
tween the two extremes red and dark blue colour. The isosurface
clearly reveals the presence of positive charge where the methyl
Analysis of molecular electrostatic potential group resides. The red region around the nitrogen atom (N3) con-
firms the delocalisation of electrons towards it. The electrostatic po-
The total electron density and MEP surfaces of the molecules tential surface of the molecule is shown in Fig. 6. The total electron
under investigation are constructed by using B3LYP/6-311++G⁄⁄ density has the limits 6.318e  102 to +6.318e  102 while the
method. Molecular electrostatic potential (MEP) mapping is very electrostatic potential has the value from 1.930e  102 to
useful in the investigation of the molecular structure with its phys- +1.930e  102. Thus, a high electrostatic potential indicates the rel-
iochemical property relationships [19,27–29]. The total electron ative absence of electrons and a low electrostatic potential indicates
density mapped with electrostatic potential surface, the contour an abundance of electrons. The electrostatic potential maps also
map of electrostatic potential and molecular electrostatic potential illustrate the reactive sites of molecules whose polarity is not easy
surface of 2MTBI are shown in Figs. 4 and 5. to intuitively deduce.
The red1 colour indicates the lowest electrostatic potential en-
ergy, and blue indicates the highest electrostatic potential energy. Analysis of frontier molecular orbitals

1
For interpretation of color in Figs. 4 and 5, the reader is referred to the web The energies of HOMO, LUMO, LUMO+1 and HOMO–1 and
version of this article. the energy gap of LUMO–HOMO are calculated using B3LYP/
V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965 957

NAH vibrations

Normally, in all the heterocyclic compounds, the NAH stretch-


ing vibration occurs in the region 3500–3000 cm1 [30]. The band
appear at 3430 cm1 in FTIR spectrum is assigned to the NAH
stretching mode of 2MTBI. The NAH in-plane bending vibration
is found at 1145 cm1 in Raman spectrum. The out of plane bend-
ing mode is assigned to the calculated scaled frequency 394 cm1.
There is no possibility of observation of the N1AH  S15
0
type
hydrogen bond due to the longer H  S15 distance (2.94 Å
A) in the
molecule.

C@N and CAN vibrations

In general, imidazoles have three or four bands in the regions


1660–1450 cm1 due to C@N and C@C stretching vibrations. The
intensities of these bands depend on the nature and position of
the substituents. Saxena et al. [31] reported a value 1608 cm1 for
polybenzodithazole and Klots and Collier [32] reported a value of
1517 cm1 for benzoxazole as C@N stretching mode. Yang et al.
[33] reported an IR band at 1626 cm1 for the oxazole ring. The very
strong band occurring at 1531 cm1 in IR and at 1521 cm1 Raman
spectra is assigned to the C@N stretching vibration of 2-amino-4-
methylbenzothiaole [34]. Benzimidazoles absorbs at 1560–
1520 cm1 due to C@N stretching vibration. But there is no
absorption found in the infrared and Raman spectra of 2MTBI. Thus,
the band observed at 1487 cm1 is assigned to C@N stretching
vibration. The red shift of the C@N stretching frequency of 2MTBI
is due to the presents of methylthio group at C2 carbon. The funda-
mental modes observed at 1398, 1249, 1215 cm1 and 1259,
1227 cm1 in the IR and Raman spectra, respectively are assigned
to CAN stretching modes. The CNC in-plane bending mode is calcu-
lated at 957 cm1. The CNC out of plane bending mode is assigned
Fig. 9. (a) Experimental FTRaman, (b) Theoretical B3LYP/cc–pVTZ and (c) B3LYP/6- at 628 cm1 in the IR spectrum. The NCN out of plane bending mode
311++G⁄⁄ Raman spectra of 2-(methylthio)benzimidazole. is found at 598 cm1 in IR spectrum. The CCN out of plane bending
mode is attributed to 329 cm1 in Raman spectrum.

6-311++G⁄⁄ method and the pictorial illustration of the frontier CAH vibrations
molecular orbitals and their respective positive and negative re-
gions are shown in Fig. 7. The positive and negative phases are rep- The strong to medium intensity bands occur in the region
resented in red and green colour, respectively. Visual analysis of 3100–3000 cm1 is common for aromatic structure. In the present
the molecular orbitals allows concepts of structural symmetry to study, the aromatic CAH stretching vibrations are observed at
be extended to frontier electron symmetry. It is actually possible 3047 cm1 in IR and 3068, 3053 cm1 in the Raman spectra. The
to predict reactivity by examining the molecular orbitals [28,29]. bands due to CAH in-plane bending vibrations occur in the region
The LUMO orbital feature a larger number of nodes than the HOMO 1290–950 cm1. The bands observed at 1154, 1124, 1008 and
orbital. The HOMO speared out over the entire part of the molecule 1015 cm1 in IR and Raman spectra are assigned to CAH in-plane
while the LUMO does not spread over the methyl group. This bending vibrations of the molecule. The CAH out of plane bending
clearly indicates the n ? p⁄ and p ? p⁄ transitions are the most modes is usually medium intensity and is observed in the region
probable. The hardness and softness of the molecule depends on 950–600 cm1 [35–37]. The out of plane CAH bending vibrations
the frontier molecular orbital energies. The calculated energy gap are observed at 931 and 851 cm1 in IR spectrum.
of LUMO–HOMO’s explains the ultimate charge transfer interface
within the molecule. The LUMO–HOMO energy gap of 2MTBI CAC vibrations
determined by B3LYP/6-311++G⁄⁄ method is 5.2312 eV.
The ring C@C stretching vibrations occur in the region 1650–
1430 cm1 [38]. In the present study, the C@C stretching vibrations
Vibrational analysis of 2MTBI are observed at 1629 and 1596 cm1 in the infrared spec-
trum while the corresponding Raman wavenumbers are 1637 and
The experimental FTIR and FT-Raman spectra and the theoreti- 1611 cm1. The CAC stretching vibrations are assigned to the
cal spectra of 2MTBI are shown in Figs. 8 and 9. The observed and modes 1353 and 1314 cm1 in the IR and 1346 and 1298 cm1 in
calculated frequencies using B3LYP method with 6-311++G⁄⁄ and Raman spectra. The ring CCC in-plane bending vibrations are gen-
cc-pVTZ basis sets along with their relative intensities, probable erally weak often being masked by other stronger absorptions due
assignments are summarised in Table 3. The geometry of the to the substituent groups. The in-plane bending vibrations of
2MTBI molecule possesses Cs point group symmetry. The 51 funda- 2MTBI are theoretically determined at 891, 808, 582 cm1 and
mental modes of vibrations are distributed into 34 in-plane vibra- the calculated out of plane vibrations are 431, 392, 249, 139 and
tions under A0 species and 17 out of plane vibrations under A00 101 cm1. These assignments are in good agreement with litera-
species. All vibrations are active in both IR and Raman. ture [39].
Table 3

958
The observed FTIR, FT-Raman and calculated frequencies using B3LYP/cc-pVTZ and B3LYP/6-311++G⁄⁄ methods with their relative intensities, probable assignments and potential energy distribution (PED) of 2-
(methylthio)benzimidazolea.

Species Observed B3LYP/cc-pVTZ calculated wavenumber B3LYP/6-311++G⁄⁄ calculated wavenumber Depolarisation Assignment %PED
wavenumber ratio
(cm1)
FTIR FTR Unscaled Scaled IR Raman Unscaled Scaled IR Raman
(cm1) (cm1) intensity intensity (cm1) (cm1) intensity intensity
A0 3430 3440 3658 3439 6.41 34.19 3668 3411 7.10 39.04 0.75 mNAH 92mNH
vw vw
A0 3068 m 3197 3069 0.68 100.00 3213 3068 0.86 100.00 0.75 mCAH 94mCH

V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965
A0 3189 3061 0.00 49.75 3204 3060 0.01 54.53 0.75 mCAH 91mCH
A0 3177 3050 3.02 48.54 3192 3048 3.05 51.70 0.57 mCAH 90mCH
A0 3047 w 3053 w 3168 3041 3.39 17.74 3183 3040 2.98 19.39 0.75 mCAH 94mCH
A0 2955 3159 2954 2.44 16.33 3179 2956 2.57 19.27 0.30 maCH3 93mCH
vw
A00 2919 2924 w 3152 2947 0.13 33.19 3171 2949 0.31 37.73 0.75 maCH3 91mCH
vw
A0 2869 w 3063 2864 6.94 75.11 3074 2859 75.87 75.51 0.17 msCH3 90mCH
A0 1629 w 1637 1658 1625 68.96 4.89 1665 1632 6.86 7.45 0.75 mC@C 92mCC
vw
A0 1596 w 1611 w 1627 1594 12.42 4.72 1631 1598 11.94 7.25 0.75 mC@C 91mCC
A0 1487 m 1494 1529 1498 1.83 19.06 1530 1499 2.10 32.24 0.62 mC@N 94mCN
vw
A0 1504 1474 1.31 21.90 1510 1480 1.70 31.67 0.75 mC@C 89mCC
A0 1468 s 1479 1449 2.68 10.56 1483 1453 2.63 13.21 0.69 daCH3 77dCH3 + 16bCS
A0 1467 1438 0.17 55.79 1474 1445 0.34 49.51 0.75 mCAC 89mCC
A00 1420 w 1458 1429 4.49 3.50 1462 1433 5.53 5.22 0.75 daCH3 78dCH3 + 16bCS
A0 1398 vs 1426 1397 2.92 5.04 1432 1403 3.64 6.75 0.22 mCAN 90mCN
A0 1353 vs 1346 1371 1344 63.24 1.92 1381 1353 60.04 2.55 0.75 mCAC 87mCC
vw
A0 1332 s 1323 w 1356 1329 3.85 4.82 1369 1342 20.42 7.83 0.75 dsCH3 79dCH3 + 15bCS
A0 1314 m 1298 vs 1317 1291 1.57 7.03 1315 1289 1.34 7.55 0.08 mCAC 86mCC
A0 1249 s 1259 m 1281 1255 0.41 47.05 1288 1262 0.37 55.46 0.75 mCAN 89mCN
A0 1215 s 1227 1246 1221 2.17 15.60 1247 1222 1.93 12.81 0.51 mCAN 90mCN
vw
A0 1154 1175 1152 2.44 2.11 1174 1151 2.22 2.71 0.75 bCAH 67bCH + 18bCCC
vw
A0 1145 w 1165 1142 5.98 15.22 1170 1147 6.63 15.43 0.14 bNAH 69bCH + 20bCCC
A0 1124 1134 1111 3.78 1.20 1132 1109 0.03 1.28 0.75 bCAH 65bCH + 22bCCC
vw
A0 1008 m 1015 1033 1012 0.13 10.57 1032 1011 3.77 13.84 0.75 bCAH 68bCH + 16bCCC
vw
A00 979 m 1002 982 5.18 1.14 1008 988 6.27 1.93 0.74 xCH3 62xCH3 + 21cCS
A0 986 966 4.78 0.02 989 969 6.09 0.75 0.09 bCAH 70bCH + 14bCCC
A00 983 963 1.76 0.66 977 957 2.17 0.06 0.57 cCAH 65cCH + 15cCCC
A0 977 957 25.02 1.42 975 956 22.08 1.21 0.22 bCNC 61bCNC + 21bNH
A00 931 w 945 926 1.05 0.02 937 918 1.16 0.04 0.28 cCAH 52cCH + 16cCCC
A0 909 891 14.80 0.47 903 885 9.60 0.49 0.27 bCCC 58bCCC + 21bCH
A00 851 w 852 w 868 851 62.11 0.19 858 841 55.63 0.13 0.25 cCAH 50cCH + 18cCCC
A0 824 808 25.43 10.96 824 808 25.50 13.78 0.39 bCCC 56bCCC + 22bCH
A0 761 w 786 770 15.31 0.05 764 749 14.19 0.24 0.37 qCH3 66qCH3 + 15dCH3
A00 756 741 53.36 0.00 750 735 68.62 0.26 0.66 cCAH 57cCH + 21cCCC
A0 714 vs 715 vw 706 692 59.87 3.31 709 695 64.10 4.34 0.65 mSAC(H3) 86mCS
A0 675 w 665 vw 694 680 8.44 0.03 679 665 10.16 0.01 0.75 mCAS 82mCS
A00 628 vw 632 619 105.52 2.62 629 616 81.47 3.20 0.34 cCNC 64cCNC + 15cCH
A00 598 w 605 593 62.18 0.92 603 591 63.80 0.97 0.39 cNCN 61cNCN + 12cCS
V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965 959

65bCS + 15 bNCN

52cCS + 24xCH3
60cNH + 21cCNC
61bCCN + 18bCH

51cCCN + 20cCH
59bCCC + 18bCH

56cCCC + 16cCH

56cCCC + 18cCH

52cCCC + 18cCH

51cCCC + 22cCH
50cCCC + 24cCH
69bCS + 18dCH3
bSAC(H3)

cSAC(H3)
cNAH
bCCN

bCAS
cCCN
bCCC

cCCC

cCCC

cCCC

cCCC
cCCC
0.25
0.38
0.73

0.72
0.75
0.66
0.75
0.31
0.14
0.24
0.60
0.06
1.66

1.97
1.05

0.24
0.19

0.13

0.12
0.39
0.15
0.03

0.01

0.30
46.35

11.44
21.59
12.38
63.58
67.30

18.20
9.57

1.99
2.02

3.10

0.04

m – stretching; b – in-plane bending; d – deformation; q – rocking; c – out of plane bending; x – wagging and s – twisting.
573
493
428
392
376

277
246
142
139
305

100
41
585

437

384
311
283
251
145
142
503

102
400

42

Fig. 10. The linear regression between the experimental and scaled theoretical
wavenumbers (a) B3LYP/cc-pVTZ, (b) B3LYP/6-311++G⁄⁄ and methods of
2-(methylthio)benzimidazole.
1.42

1.75
0.31

0.24
0.41
0.16
0.34
0.80

0.50

0.07
0.00

0.00

Table 4
The Experimental and calculated 1H and 13C isotropic chemical shifts (diso, ppm)
34.49
42.51

59.22
16.70

10.59
21.04
12.90
8.47

2.71
2.47
1.90

0.13

with respect to TMS and isotropic magnetic shielding tensors (riso) of 2-


(methylthio)benzimidazole.

Assignment riso Cal. Expt. Assignment riso Cal. Expt.


(1H) (diso) (d) (13C) (diso) (d)
H10 24.12 7.85 7.26 C2 22.10 162.43 151.54
582
494
431
394
392

277
249
143
139
307

101
47

H11 24.57 7.40 7.20 C4 59.97 124.56 115.43


H12 24.60 7.37 7.20 C5 56.87 127.66 122.43
H13 24.41 7.56 7.26 C6 56.30 128.23 122.43
H14 23.68 8.29 7.50 C7 69.72 114.81 115.43
H17 29.01 2.96 2.79 C8 32.34 152.19 140.50
H18 29.71 2.26 2.79 C9 41.18 143.35 140.50
H19 29.01 2.96 2.79 C16 163.16 21.37 14.84
594

313
283
254
146
142
504
440
402

103
400

48
449 vw

329 vw

CAS Vibrations
177 s
152 s
133 s

Compounds that contain a thiocarbonyl group (C@S) show


absorption in 1250–1020 cm1 region [40–42]. The CAS stretch-
515 w
447 w

418 w

ing frequency is quite consistent for many different types of com-


pounds that contain the CAS moiety. The stretching vibration of
H3CAS linkage usually occurs as a polarised strong band in the
region of 750–590 cm1 [40–42]. In the present study, the
A00
A00
A00
A00

A00

A00
A00
A00
A0
A0

A0

A0

SAC(H3) stretching vibration is observed at 714 and 715 cm1


a
960 V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965

frequencies are established at 2955, 2919 and 2924 cm1 in the


infrared and Raman spectra, respectively while ms(CH3) is attrib-
uted to 2869 cm1 in the infrared spectrum. The asymmetric
methyl deformation modes, da(CH3) are observed at 1468 and
1420 cm1 in Raman spectrum. The strong mode observed at
1332 in IR and the weak one at 1323 cm1 in Raman spectra is as-
signed to the symmetrical methyl deformation mode. These
assignments are substantiated by the reported literatures [43–
46]. The methyl rocking and wagging modes are also presented
in Table 3.

7. Scale factors

Computed harmonic frequencies typically overestimate vibra-


tional fundamentals due to basis set truncation and neglect of elec-
tron correlation and mechanical anharmonicity [47]. To
compensate these shortcomings and better agreement between
the computed and experimental frequencies can be obtained by
Fig. 11. 1H NMR spectrum of 2-(methylthio)benzimidazole. using different scale factors. To determine the scale factors, the
procedure used previously [44,46,48–54] has been followed that
minimises the deviations between the experimental and theoreti-
in IR and Raman spectra. The other CAS stretching vibration is ob- cally predicted vibrational frequencies. The scale factors for vibra-
served at 675 cm1 in IR and 665 cm1 in Raman spectra. The CAS tional frequencies were determined by minimising the residual
stretching vibration is wrongly assigned to 1267/1268 cm1 by
N 
X 2
Krishnakumar and Ramasamy [7]. The in-plane SAC(H3) bending
D¼ kxTheor  mExpt
vibrations are observed at 177 cm1 in Raman spectrum. i i
i

Methyl group vibrations where xTheo


i and mExpt
i are the ith theoretical harmonic frequency
and ith experimental fundamental frequency (in cm1), respectively
Considering the assignments of ACH3 group frequencies, which and N is the total number of fundamental modes which leads to
make a significant contribution to 2MTBI spectra, one can expect rffiffiffiffi
D
that nine fundamentals can be associated to each ACH3 group, RMS ¼
N
namely symmetric, ms(CH3), and asymmetric, ma(CH3), stretching
modes; the symmetric, ds(CH3), and asymmetric, da(CH3), hydrogen In the present investigation a scale factor of 0.93 for NAH and ACH3
deformation modes; the in-plane methyl rocking vibrations; and group stretching, 0.955 for CAH stretching, and 0.98 for all other
the out of plane ACH3 wagging and twisting modes. The ma(CH3) vibrational modes are used in B3LYP/6-311++G⁄⁄ method. The scale

13
Fig. 12. C NMR spectrum of 2-(methylthio)benzimidazole.
V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965 961

Fig. 13. The linear regression between the experimental and theoretical (a) 1H and Fig. 14. The linear regression between the experimental (a) 1H and (b) 13C NMR
(b) 13C NMR Chemical shifts (d) of 2-(methylthio)benzimidazole. Chemical shifts (d) and the isotropic shielding (r) of 2-(methylthio)benzimidazole.

factor of 0.94 for NAH stretching, 0.96 for CAH stretching, 0.935 for
ACH3 stretching and 0.98 for all other vibrational modes are used in
B3LYP/cc-pVTZ method. All the correction factors are very much
closer to unity and the frequencies are much closer to the experi-
mental values and more reliable. The determined RMS deviation
is only 9–10. The correlation diagram for the theoretical and the
experimental frequencies of 2MTBI is shown in Fig. 10.

NMR spectral studies

The 1H and 13C theoretical and experimental chemical shifts


[55], isotropic shielding constants and the assignments of 2MTBI
are presented in Table 4. The observed 1H and 13C NMR spectra
of the compound 2MTBI in CDCl3 solvent are given in Figs. 11
and 12 respectively. Aromatic carbons give signals with chemical
shift values in the range 100–200 ppm [56,57]. The observed
experimental chemical shift positions of ring carbons of 2MTBI
lie in the range 115.43–151.54 ppm. The carbon atom C2 is highly
deshielded due to the cumulative AI effect of nitrogen in the hetero
ring, thus its NMR signal is found in the very downfield at
151.54 ppm. Due to the deshielding effect of electronegative N1 Fig. 15. Correlation of the experimental and theoretical 1H NMR chemical shifts of
and N3 atoms the chemical shift value of C8 and C9 are also 2-(methylthio)benzimidazole.
962 V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965

upfiled at 2.79 ppm they are under high magnetic shielding. The
proton in the heterocyclic ring H14 is realised high deshielding ef-
fect (7.50 ppm) than the aromatic protons.
The calculated and experimental chemical shift values given in
Table 4 shows a good agreement with each other. The linear
regression between the experimental and theoretical 1H and 13C
NMR Chemical shifts are represented in Fig. 13. The linear regres-
sion between the experimental 1H and 13C chemical shift and the
isotropic shielding constants are represented in Fig. 14. Correlation
of the experimental and theoretical 1H and 13C NMR chemical
shifts are shown in Figs. 15 and 16, respectively. The agreements
between the experimental and calculated shifts for 1H and 13C
atoms are good [58].

Natural bond orbital (NBO) analysis

The NBO method demonstrates the bonding concepts like atom-


ic charge, Lewis structure, bond type, hybridisation, bond order,
Fig. 16. Correlation of the experimental and theoretical 13C NMR chemical shifts of charge transfer and resonance possibility. Table 5 depicts the type
2-(methylthio)benzimidazole.
of bond orbital, their occupancies, the natural atomic hybrids of
which the NBO is composed, giving the percentage of the NBO on
each hybrid, the atom label, and a hybrid label showing the hybrid
attributed to the downfield NMR signals 140.50 ppm, respectively. orbital (spx) composition (the amount of s-character, p-character,
The other benzene ring carbon atoms show signals at 122.43 and etc.) of 2MTBI determined by B3LYP/6-311++G⁄⁄ method. The
115.43 ppm. occupancies of NBOs reflect their exquisite dependence on the
1
H chemical shifts of 2MTBI were obtained and interpreted crit- chemical environment. The NBO energy values show the corre-
ically to quantify the possible different effects acting on the shield- sponding spatial symmetry breaking in the direction of unpaired
ing constant of protons. The hydrogen atoms H10, H11, H12 and spin. For example, the bonding orbital for C2AN3 with 1.9852 elec-
H13 present in the benzene ring shows NMR peaks in the normal trons has 41.72% C2 character in a sp1.70 hybrid and has 58.28% N3
range of aromatic hydrogen atoms and are assigned to 7.26, 7.20, character in a sp1.80 hybrid orbital of 2MTBI. The bonding orbital for
7.20 and 7.26 ppm, respectively. The methyl group protons are N1AC2 with 1.9873 electrons has 62.46% N1 character in a sp1.88

Table 5
Bond orbital analysis of 2-(methylthio)benzimidazole.

Bond orbital Occupancy Atom Contribution from Parent NBO (%) Co-efficiency Atomic hybrid contributions (%)
N1AC2 1.9873 N1 62.46 0.7903 s (34.65) + p1.88 (65.35)
C2 37.54 0.6127 s (31.92) + p2.13 (67.97)
N1AC9 1.9820 N1 62.04 0.7877 s (35.66) + p1.80 (64.30)
C9 37.96 0.6161 s (26.44) + p2.78 (73.46)
N1AH14 1.9893 N1 71.17 0.8436 s (29.55) + p2.38 (70.41)
H14 28.83 0.5369 s (99.93) + p0.00 (0.07)
C2AN3 1.9852 C2 41.72 0.6459 s (37.02) + p1.70 (62.90)
N3 58.28 0.7634 s (35.66) + p1.80 (64.24)
C2AN3 1.8830 C2 40.25 0.6344 s (0.00) + p1.00 (99.83)
N3 59.75 0.7730 s (0.00) + p1.00 (99.83)
C2AS15 1.9804 C2 54.70 0.7396 s (31.06) + p2.21 (68.79)
S15 45.30 0.6730 s (16.78) + p4.92 (82.56)
N3AC8 1.9692 N3 58.36 0.7639 s (31.91) + p2.13 (68.00)
C8 41.64 0.6453 s (28.97) + p2.45 (70.94)
C4AC5 1.9764 C4 50.42 0.7100 s (36.26) + p1.76 (63.70)
C5 49.58 0.7042 s (35.94) + p1.78 (64.02)
C4AC5 1.7109 C4 49.03 0.7002 s (0.00) + p1.00 (99.95)
C5 50.97 0.7140 s (0.00) + p1.00 (99.96)
C4AC8 1.9762 C4 48.64 0.6974 s (34.49) + p1.90 (65.46)
C8 51.36 0.7167 s (38.45) + p1.60 (61.53)
C5AC6 1.9791 C5 49.88 0.7063 s (35.67) + p1.80 (64.28)
C6 50.12 0.7079 s (35.97) + p1.78 (63.98)
C6AC7 1.9748 C6 49.15 0.7011 s (35.65) + p1.80 (64.31)
C7 50.85 0.7131 s (36.55) + p1.74 (63.41)
C6AC7 1.7223 C6 48.94 0.6996 s (0.00) + p1.00 (99.96)
C7 51.06 0.7146 s (0.00) + p1.00 (99.96)
C7AC9 1.9757 C7 48.25 0.6946 s (34.10) + p1.93 (65.85)
C9 51.75 0.7194 s (40.14) + p1.49 (59.83)
C8AC9 1.9643 C8 48.88 0.6991 s (32.32) + p2.09 (67.63)
C9 51.12 0.7150 s (33.20) + p2.01 (66.75)
C8AC9 1.5947 C8 49.65 0.7047 s (0.00) + p1.00 (99.97)
C9 50.35 0.7096 s (0.00) + p1.00 (99.97)
S15AC16 1.9845 S15 50.01 0.7071 s (16.60) + p4.99 (82.87)
C16 49.99 0.7071 s (21.53) + p3.64 (78.29)
V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965 963

Table 6 Donor–acceptor interactions


Second order perturbation analysis of Fock matrix of 2-(methylthio)benzimidazole
using NBO analysis.
Natural bond orbital (NBO) analysis is useful for understanding
Donor (i)–Acceptor (j) E(2)a E(j)  E(i)b F(i, j)c the delocalisation of electron density from Lewis occupied donor
interaction (kJ mol1) (a.u.) (a.u.) NBOs to unoccupied non-Lewis acceptor NBOs within the mole-
r(C2AS15) ? r⁄(N1AC9) 17.22 0.35 0.076 cule. The stabilisation of orbital interaction is proportional to the
r(N3AC8) ? r⁄(C4AC5) 8.18 0.95 0.079 energy difference between interacting orbitals. Therefore, the
r(C4AC5) ? r⁄(C4AC8) 5.51 1.16 0.071
p(C4AC5) ? p⁄(C8AC9) 20.2 0.28 0.068
interaction having strongest stabilisation takes place between
p(C4AC5) ? p⁄(C8AC9) 18.91 0.27 0.067 effective donors and effective acceptors. These bonding–antibond-
r(C6AC7) ? r⁄(N1AC9) 6.83 1.13 0.078 ing interactions can be quantitatively described by means of sec-
r(C6AC7) ? r⁄(N1AC9) 6.83 1.13 0.078 ond-order perturbation interaction energy E(2) [59–62]. This
p(C6AC7) ? p⁄(C4AC5) 17.59 0.29 0.065
energy represents the estimate of the off-diagonal NBO Fock ma-
p(C6AC7) ? p⁄(C8AC9) 20.08 0.28 0.070
p(C8AC9) ? p⁄(C2AN3) 14.66 0.24 0.053 trix element. The stabilisation energy E(2) associated with
p(C8AC9) ? p⁄(C4AC5) 18.62 0.29 0.067 i(donor) ? j(acceptor) delocalisation is estimated as
p(C8AC9) ? p⁄(C6AC7) 19.26 0.28 0.067
n[LP(1)] (N1) ? r⁄(C2AN3) 47.48 0.27 0.103 F 2 ði; jÞ
n[LP(1)] (N1) ? r⁄(C8AC9) 30.55 0.30 0.089 Eð2Þ ¼ qi
ej  ei
n[LP(1)] (N3) ? r⁄(N1AC2) 9.33 0.78 0.077
n[LP(1)] (N3) ? r⁄(C8AC9) 5.7 0.91 0.065
where qi is the donor orbital occupancy, ei and ej are diagonal ele-
n[LP(1)] (S15) ? r⁄(C2AN3) 5.71 1.2 0.074
n[LP(2)] (S15) ? p⁄(C2AN3) 25.81 0.25 0.076 ments (orbital energies) and F(i, j) is the off-diagonal Fock matrix
p(C2AN3) ? p⁄(C8AC9) 57.48 0.03 0.058 element.
p(C8AC9) ? p⁄(C4AC5) 261.47 0.01 0.082 The stabilisation energies of the donor–acceptor interactions
LP – Lone pair. with more than 5 kJ mol1 determined by second order perturba-
a
Stabilisation (delocalisation) energy. tion analysis of Fock matrix of 2MTBI are summarised in Table 6.
b
Energy difference between i(donor) and j(acceptor) NBO orbitals. In 2MTBI molecule, pC8–C9 ? pC4—C5 interaction is strongly stabi-
c
Fock matrix element i and j NBO orbitals. lized by 261.47 kJ mol1. The stabilisation energy of pC2–
1
N3 ? pC8—C9 is 57.48 kJ mol

. The lone pair donor orbital,
n(N1) ? rC2—N3

interaction has stabilisation energy of
hybrid and 37.54% C2 character in a sp2.13 hybrid orbital. A bonding 47.48 kJ mol1. The stabilisation energy of lone pair donor orbital,
orbital for N1AH14 with 1.9893 electrons has 71.17% N1 character n(S15) ? rC2—N3 is 25.81 kJ mol1.
in a sp2.38 hybrid and has 28.83% H14 character in a sp hybrid orbi-
tal. The CAC bonds of the aromatic ring possess more p character Analysis of structure–activity descriptors
than s character. This clearly indicates the delocalisation of p elec-
trons among all the carbon atoms. Similarly the bond C2AN3 has The atomic charges of the neutral, cationic and anionic species
also possessed more p character. of 2MTBI calculated by natural population analysis using B3LYP/

Table 7
Calculated local reactivity properties of 2-(methylthio)benzimidazole using B3LYP/6-311++G⁄⁄ method for natural population analysis (NPA) derived charges.

Atom Neutral Anion Cation fkþ fk fk0 Df(k)

N1 0.5717 0.5969 0.5597 0.0252 0.0120 0.0186 0.0132


C2 0.2537 0.2540 0.2730 0.0003 0.0193 0.0095 0.0196
N3 0.5364 0.5581 0.4581 0.0217 0.0783 0.0500 0.0566
C4 0.1977 0.2134 0.1646 0.0157 0.0331 0.0244 0.0174
C5 0.2190 0.2337 0.1663 0.0147 0.0527 0.0337 0.0380
C6 0.2060 0.2234 0.0654 0.0174 0.1406 0.0790 0.1232
C7 0.2377 0.2464 0.2379 0.0087 0.0002 0.0043 0.0089
C8 0.1144 0.1150 0.1875 0.0006 0.0731 0.0363 0.0737
C9 0.1264 0.1248 0.2072 0.0016 0.0808 0.0412 0.0792
S15 0.2520 0.2009 0.5368 0.0511 0.2848 0.1680 0.2337
C16 0.6807 0.7276 0.6935 0.0469 0.0128 0.0171 0.0597

Table 8
Calculated local reactivity properties of 2-(methylthio)benzimidazole using B3LYP/6-311++G⁄⁄ method for natural population analysis (NPA) derived charges.

Atom sþ
k
s
k s0k Dsk xþk xk x0k Dxk Electro philicity Nucleo philicity

N1 0.0032 0.0015 0.0023 0.0017 0.0422 0.0201 0.0312 0.0221 2.1000 0.4762
C2 0.0000 0.0024 0.0012 0.0025 0.0005 0.0324 0.0159 0.0329 0.0155 64.3333
N3 0.0027 0.0099 0.0063 0.0071 0.0364 0.1313 0.0838 0.0949 0.2771 3.6083
C4 0.0020 0.0042 0.0031 0.0022 0.0263 0.0555 0.0409 0.0292 0.4743 2.1083
C5 0.0019 0.0066 0.0042 0.0048 0.0246 0.0884 0.0565 0.0637 0.2789 3.5850
C6 0.0022 0.0177 0.0099 0.0155 0.0292 0.2357 0.1324 0.2065 0.1238 8.0805
C7 0.0011 0.0000 0.0005 0.0011 0.0146 0.0003 0.0071 0.0149 43.5000 0.0230
C8 0.0001 0.0092 0.0046 0.0093 0.0010 0.1226 0.0608 0.1236 0.0082 121.8333
C9 0.0002 0.0102 0.0052 0.0100 0.0027 0.1355 0.0691 0.1328 0.0198 50.5000
S15 0.0064 0.0359 0.0211 0.0294 0.0857 0.4775 0.2816 0.3918 0.1794 5.5734
C16 0.0059 0.0016 0.0021 0.0075 0.0786 0.0215 0.0286 0.1001 3.6641 0.2729
964 V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965

6–311++G⁄⁄ method are presented in Table 7. Among the benzene The chemical reactivity and site selectivity of the molecule has
ring carbon atoms, a positive charge resides on C8 and C9 while been determined with the help of global and local reactivity
others have negative charge. The cationic species has more positive descriptors. The dual descriptors Dsk and Dfk and the multiphilicity
charges at these carbon atoms. The high positive charge at C2 in descriptor (Dxk) reveals that the atoms N1, C7 and C16 are favour-
the hetero ring is due to the AI effect of the surrounding electro- able for electrophilic attack while other atoms are favourable for
negative nitrogen and sulphur atoms. The methyl group carbon nucleophilic attack. The S15 and C6 atoms are more prone to
atom C16 has high negative charge because of no bond resonance. nucleophilic attack. Thus, the present investigation provides a
All other carbon atoms also possess negative charges. Both the complete and reliable structural, vibrational and structure–actitvi-
nitrogen atoms contain more negative charge due to the high elec- ty relations of the compound.
tronegativity. The positive charge on S15 reveals the delocalisation
of lone pair of electrons towards the hetero ring. The high positive
charge of H14 is due to the attachment of electronegative nitrogen
References
atom N1 while the other hydrogen atoms present in the aromatic
ring and methyl group have less positive charges. [1] I.S. Ahuja, I. Prasad, Inorg. Nucl. Chem. Lett. 12 (1976) 777–784.
The understanding of chemical reactivity and site selectivity of [2] S.O. Podunavac–Kuzmonovic, L.M. Leovac, N.U. Perisicjanjic, J. Rogan, J. Balaz, J.
the molecular systems has been effectively handled by the concep- Serb. Chem. Soc. 64 (1999) 381–388.
[3] C.N.R. Rao, R. Venkataraghavan, Can. J. Chem. 42 (1964) 43–49.
tual density functional theory (DFT) [63]. Chemical potential, glo- [4] A. Suwaiyan, R. Zwarich, N. Baig, J. Raman Spectrosc. 21 (1990) 243–249.
bal hardness, global softness, electronegativity and [5] D.E. Lynch, L.J. Nicholls, G. Smith, K.A. Byriel, C.H.L. Kennard, Acta Cryst. B 55
electrophilicity are global reactivity descriptors, highly successful (1999) 758–766.
[6] J.V.N. Vara–Prasad, A. Panapoulous, J.R. Rubin, Tetrahedron Lett. 41 (2000)
in predicting global chemical reactivity trends. The global parame- 4065.
ters ionisation potential (I), electron affinity (A), electrophilicity [7] V. Krishnakumar, R. Ramasamy, Spectrochim. Acta 62A (2005) 570–577.
(x), electronegativity (v), hardness (g), and softness (S) of the mol- [8] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman,
G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato,
ecule are determined and displayed in Table 2. The site-selectivity X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M.
of a chemical system, cannot, however, be studied using the global Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y.
descriptors of reactivity. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery, Jr., J.E. Peralta, F. Ogliaro,
M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J.
Fukui functions and local softness are extensively applied to Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M.
probe the local reactivity and site selectivity. The formal defini- Cossi, N. Rega, J.M. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo,
tions of all these descriptors and working equations for their com- J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C.
Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, G.A. Voth,
putation have been described [64–66]. The Fukui functions of the
P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, O. Farkas, J.B. Foresman,
individual atoms of the neutral, cationic and anionic species of J.V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian Inc., Wallingford CT, 2009.
2MTBI calculated by B3LYP/6-311++G⁄⁄ method are presented in [9] H.B. Schlegel, J. Comput. Chem. 3 (1982) 214–218.
Table 7. The molecule under investigation mainly gives substitu- [10] P. Hohenberg, W. Kohn, Phys. Rev. B 136 (1964) 864–871.
[11] A.D. Becke, J. Chem. Phys. 98 (1993) 5648–5652.
tion reactions. It is clearly understood that the atoms N1, C7 and [12] A.D. Becke, Phys. Rev. A 38 (1988) 3098–3100.
C16 are favourable for electrophilic attack. The other atoms are [13] C. Lee, W. Yang, R.G. Parr, Phys. Rev. B 37 (1988) 785–789.
favourable for nucleophilic attack. The S15 and C6 are more prone [14] E.B. Wilson Jr., J. Chem. Phys. 7 (1939) 1047–1052.
[15] E.B. Wilson Jr., J. Chem. Phys. 9 (1941) 76–84.
to nucleophilic attack. [16] E.B. Wilson Jr., J.C. Decius, P.C. Cross, Molecular Vibrations, McGraw Hill, New
The local softness, relative electrophilicity (sþk
=s
k ) and relative York, 1955.
 þ
nucleophilicity (sk =sk ) indices, the dual local softness Dsk and the [17] H. Fuhrer, V.B. Kartha, K.L. Kidd, P.J. Kruger, H.H. Mantsch, Computer Program
for Infrared and Spectrometry, Normal Coordinate Analysis, vol. 5, National
multiphilicity descriptors (Dxk) have also been determined to pre- Research Council, Ottawa, Canada, 1976.
dict the reactive sites of the molecule and are summarised in Ta- [18] V. Krishnakumar, G. Keresztury, T. Sundius, R. Ramasamy, J. Mol. Struct. 702
ble 8. From the dual local softness Dsk and the multiphilicity (2004) 9–21.
[19] J.S. Murray, K. Sen, Molecular Electrostatic Potentials, Concepts and
descriptors (Dxk) one can understand that the atoms N1, C7 and Applications, Elsevier, Amsterdam, 1996.
C16 are favourable for electrophilic attack while other atoms are [20] S. Chidangil, M.K. Shukla, P.C. Mishra, J. Mol. Model. 4 (1998) 250–258.
favourable for nucleophilic attack. The S15 and C6 atoms have [21] R.I. Dennington, T. Keith, J. Millam, GaussView, Version 5.0.8, Semichem. Inc,
Shawnee Mission, KS, 2008.
more prone to nucleophilic attack. The local reactivity descriptors
[22] R. Duchfield, J. Chem. Phys. 56 (1972) 5688–5691.
of the individual atoms of the molecule sak ¼ fka S, xak ¼ xfka and fka [23] K. Wolinski, J.F. Hinton, P. Pulay, J. Am. Chem. Soc. 112 (1990) 8251–8260.
where, a = +,  and 0 represents local philicity quantities describ- [24] R.G. Parr, W. Yang, J. Am. Chem. Soc. 106 (1984) 4049–4050.
ing nucleophilic, electrophilic and free radical attack, respectively [25] W. Yang, R.G. Parr, Proc. Natl. Acad. Sci. USA 82 (1985) 6723–6726.
[26] C.J. Dik–Edixhoven, H. Schenk, H. Van der Meer, Cryst. Struct. Commun. 2
and presented in Tables 7 and 8 where the nature of activity of (1973) 23–24.
the individual atoms can be determined. [27] I. Fleming, Frontier Orbitals and Organic Chemical Reactions, John Wiley and
Sons, New York, 1976. pp. 5–27.
[28] J.M. Seminario, Recent Developments and Applications of Modern Density
Functional Theory, vol. 4, Elsevier, 1996. pp. 800-806.
Conclusions [29] T. Yesilkaynak, G. Binzet, F. Mehmet Emen, U. Florke, N. Kulcu, H. Arslan, Eur. J.
Chem. 1 (2010) 1–5.
[30] G. Socrates, Infrared and Raman characteristic group frequencies, tables and
The structure of 2MTBI is optimised with B3LYP method using charts, third ed., Wiley, Chichester, 2001.
6-31G⁄⁄, 6-311++G⁄⁄ and cc-pVTZ basis sets. The most stable con- [31] R. Saxena, L.D. Kandpal, G.N. Mathur, J. Polym. Sci. A: Polym. Chem. 40 (2002)
former is determined by conformational analysis and is more sta- 3959–3966.
[32] T.D. Klots, W.B. Collier, Spectrochim. Acta 51A (1995) 1291–1316.
bilised by 3.54 kcal mol1 than the other conformers. The [33] G. Yang, S. Matsuzono, E. Koyama, H. Tokuhisa, K. Hiratani, Macromolecules 34
complete molecular structural parameters and thermodynamic (2001) 6545–6552.
properties of the optimised geometry have been determined. The [34] V. Arjunan, T. Rani, S. Sakiladevi, C.V. Mythili, S. Mohan, Spectrochim. Acta 88A
(2012) 220–231.
charge transfer depends on the LUMO–HOMO energy gap. The en-
[35] Y. Wang, S. Saebo, C.U. Pittman Jr., J. Mol. Struct.: Theochem. 281 (1993) 91–
ergy gap between LUMO and HOMO in 2MTBI is found to be 98.
5.2312 eV by B3LYP/6-311++G⁄⁄ method. The vibrational frequen- [36] V. Arjunan, N. Puviarasan, S. Mohan, Spectrochim. Acta 64A (2006) 233–239.
cies of the fundamental modes of the compound have been pre- [37] A. Altun, K. Golcuk, M. Kumru, J. Mol. Struct.: Theochem. 637 (2003) 155–169.
[38] D.N. Sathyanarayana, Vibrational Spectroscopy—Theory and Applications,
cisely assigned and analysed. 1H and 13C NMR isotropic chemical second ed., New Age International (P) Limited Publishers, New Delhi, 2004.
shifts are determined and compared with the experimental values. [39] V. Krishna kumar, R. John Xavier, Indian J. Pure Appl. Phys. 41 (2003) 95–99.
V. Arjunan et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 118 (2014) 951–965 965

[40] R.M. Silverstein, G.C. Bassler, T.C. Morrill, Spectrometric Identification of [54] V. Arjunan, P. Ravindran, T. Rani, S. Mohan, J. Mol. Struct. 988 (2011) 91–101.
Organic Compounds, fifth ed., 1991. [55] I. Kowalczyk, J. Mol. Struct. 973 (2010) 163–172.
[41] D. Lin–Vien, N.B. Colthup, V.G. Fateley, J.G. Grasselli, The Handbook of Infrared [56] H.O. Kalinowski, S. Berger, S. Braun, Carbon–13 NMR spectroscopy, John Wiley
and Raman Characteristic Frequencies of Organic Molecules, Academic Press, and Sons, Chichester, 1988.
Boston, 1991. [57] K. Pihlaja, E. Kleinpeter (Eds.), Carbon–13 chemical shifts in Structural and
[42] W. Qian, S. Krimm, Biopolym. 32 (1992) 1503–1518. Sterochemical Analysis, VCH Publishers, Deerfield Beach, 1994.
[43] A. Altun, K. Golcuk, M. Kumru, J. Mol. Struct. (Theochem) 625 (2003) 17–24. [58] B. Osmiałowski, E. Kolehmainen, R. Gawinecki, Magn. Res. Chem. 39 (2001)
[44] V. Arjunan, S. Mohan, Spectrochim. Acta 72A (2009) 436–444. 334–340.
[45] J.R. Durig, M.M. Bergana, H.V. Phan, J. Raman Spectrosc. 22 (1991) 141–154. [59] A.E. Reed, F. Weinhold, J. Chem. Phys. 83 (1985) 1736–1740.
[46] V. Arjunan, S. Mohan, J. Mol. Struct. 892 (2008) 289–299. [60] A.E. Reed, R.B. Weinstock, F. Weinhold, J. Chem. Phys. 83 (1985) 735–746.
[47] J.A. Pople, H.B. Schlegel, R. Krishnan, J.S. Defrees, J.S. Binkley, M.J. Frisch, R.A. [61] A.E. Reed, F. Weinhold, J. Chem. Phys. 78 (1983) 4066–4073.
Whiteside, Int. J. Quant. Chem. Quant. Chem. Symp. 15 (1981) 269. [62] J.P. Foster, F. Weinhold, J. Am. Chem. Soc. 102 (1980) 7211–7218.
[48] H.F. Hameka, J.O. Jensen, J. Mol. Struct. (Theochem) 362 (1996) 325–330. [63] R.G. Parr, W. Yang, Density Functional Theory of Atoms and Molecules, Oxford
[49] J.O. Jensen, A. Banerjee, C.N. Merrow, D. Zeroka, J.M. Lochner, J. Mol. Struct. University Press, Oxford, 1989.
(Theochem) 531 (2000) 323–331. [64] R.G. Pearson, Chemical Hardness – Applications from Molecules to Solids, VCH
[50] A.P. Scott, L. Radom, J. Phys. Chem. 100 (1996) 16502–16513. Wiley, Weinheim, 1997.
[51] M.P. Andersson, P. Uvdal, J. Chem. Phys. A 109 (2005) 2937–2941. [65] P. Geerlings, F. De Proft, W. Langenaeker, Chem. Rev. 103 (2003) 1793–1873.
[52] M. Alcolea Palafox, M. Gill, N.J. Nunez, V.K. Rastogi, Lalit Mittal, Int. J. Quant. [66] P.K. Chattaraj (Ed.) Special Issue of J. Chem. Sci. on Chemical Reactivity, Vol.
Chem. 103 (2005) 394–421. 117, 2005.
[53] M. Alcolea Palafox, Int. J. Quant. Chem. 77 (2000) 661–684.

You might also like