Sei sulla pagina 1di 10

Journal of Molecular Structure 1005 (2011) 192201

Contents lists available at SciVerse ScienceDirect

Journal of Molecular Structure


journal homepage: www.elsevier.com/locate/molstruc

Spectroscopic studies, HOMOLUMO and NBO calculations on monomer and dimer conformer of 5-nitrosalicylic acid
T. Karthick a, V. Balachandran b,, S. Perumal c, A. Nataraj d
a

Department of Physics, Vivekanandha College for Women, Tiruchengode 637 205, India Department of Physics, AA Government Arts College, Musiri, Tiruchirappalli 621 201, India c Department of Physics, S.T. Hindu College, Nagarcoil 629 002, India d Department of Physics, Thanthai Hans Roever College, Perambalur 621 212, India
b

a r t i c l e

i n f o

a b s t r a c t
In this work, FT-IR and FT-Raman spectra are recorded on the solid phase of 5-nitrosalicylic acid (abbreviated as 5-NSA). The energies of all possible conformers obtained from DFT theory with 6-311++G(d,p) basis set identied the most stable conformer of 5-NSA as C6 form. Optimized geometrical parameters, vibrational assignments, HOMOLUMO energies and NBO calculations are performed on the stable monomer and dimer structure of 5-NSA using the same level of theory. Second order perturbation energies and electron density (ED) transfer from lled lone pairs of Lewis base to unlled Lewis acid sites of 5-NSA are discussed on the basis of NBO analysis. Inter- and intramolecular hydrogen bonds exist between ACOOH and AOH group, give the evidence for the formation of dimer entities in the title molecule. The variations in bond lengths, ED and vibrational wavenumbers of 5-NSA dimer are being discussed. The spectroscopic and theoretical results are compared to the corresponding properties for 5-NSA monomer and dimer of C6 conformer. Reliable vibrational modes associated with 5-NSA are made on the basis of total energy distribution (TED) results obtained from scaled quantum mechanical (SQM) method. Crown Copyright 2011 Published by Elsevier B.V. All rights reserved.

Article history: Received 17 July 2011 Received in revised form 26 August 2011 Accepted 26 August 2011 Available online 16 September 2011 Keywords: Vibrational spectra 5-Nitrosalicylic acid DFT HOMOLUMO Dimer NBO analysis

1. Introduction Owing to their dimeric nature, salicylic acid and its derivatives have recently become attractive to spectroscopic researchers. Boczar et al. have been reported the optimized dimer geometry and vibrational assignments of salicylic acid [1]. Experimental FT-Raman, FT-IR and theoretical dimer conformer of 5-bromosalicylic acid [2] and 5-uro, 5-chlorosalicylic acid [3] have been studied by Karabacak et al. Apart from the dimeric characteristics, the monomer geometry of salicylic acid and its analogs play a vital role in the vibrational analysis. Chen and Shyu investigated the conformers and intramolecular hydrogen bonding of salicylic acid and its anions [4]. Goulet and Aroca [5] presented the infrared and Raman spectra of solid salicylic acid and performed DFT/ B3LYP/6-311+G(d,p) calculations. Likewise, the infrared, Raman spectra, ab initio calculations and the vibrational assignments of 4-aminosalicylic acid were studied by Akkaya and Akuyz [6]. They obtained seven stable conformers for 4-aminosalicylic acid. Varghese et al. elucidated the vibrational assignments of 3,5-dinitrosalicylic acid on the basis of experimental IR and Raman spectra [7]. Recently FT-Raman, FT-IR spectra and the structure of benzoic and
Corresponding author. Tel.: +91 431 2432454; fax: +91 4326 262630.
E-mail address: brsbala66@gmail.com (V. Balachandran).

3,5-dichlorosalicylic acid were discussed by Krishnakumar and Mathamal [8]. The simplest aromatic carboxylic acid known as salicylic acid is found in nature both in the free state (owers of wild chamomile and a species of the willow) and in the bound state in many essential oils (clove, tuberose, etc.). It is widely used as a plant growth regulator, antimicrobial and antifungal agent in food product industry [9,10]. In medicine, salicylic acid and its derivatives are employed as an intestinal antiseptic agent for treating rheumatic fever [11]. Upto now, in the spectroscopic eld, salicylic acid and its derivatives have been taken for elucidating their vibrational assignments and optimized molecular structures. Besides, a few of inter- and intramolecular hydrogen bonding studies have been performed by modern spectroscopists [1,3]. To the extent of our knowledge, neither FT-IR and FT-Raman spectra nor dimer geometries of 5-nitrosalicylic acid (5-NSA) have been performed so far. The main objective of this work is to illustrate vibrational assignments, optimized geometrical parameters, HOMOLUMO energies and natural bond orbital analysis of stable conformer of 5-NSA monomer and dimer conformer. In the present investigation, the delocalization of electron density (ED) from the lled lone pairs of Lewis base to the unlled Lewis acid sites of 5-NSA dimer and the corresponding changes in bond lengths, ED, vibrational wavenumbers upon dimerization are being reported.

0022-2860/$ - see front matter Crown Copyright 2011 Published by Elsevier B.V. All rights reserved. doi:10.1016/j.molstruc.2011.08.050

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201

193

2. Experimental The commercial crystalline sample of 5-NSA (99% purity) obtained from Lancaster Chemical Company (UK) is directly (without further purication) used for spectral measurements. The Fourier transform infrared spectrum (FT-IR) of the title compound is recorded in the frequency region 4004000 cm1 on a NEXUS 670 spectrophotometer equipped with an MCT detector, a KBr pellet technique. The FT-Raman spectrum of 5-NSA is recorded in the frequency region 04000 cm1 on a NEXUS 670 spectrophotometer equipped with Raman module accessory operating at 1.5 W power with Nd:YAG laser of wavelength 1064 nm as an excitation source. 3. Computational details The entire vibrational assignments and optimized monomer and dimer geometrical parameters of stable conformer of 5-NSA are predicted by means of density functional B3LYP method with internally stored standard 6-311++G(d,p) basis set level in Gaussian 03W software package [12]. B3LYP represents Beckes threeparameter hybrid functional method [13] with LeeYangParrs correlation functional (LYP) [14,15]. In order to obtain the stable conformer, the self consistent eld (SCF) energy calculation is performed on eight possible conformers of 5-NSA using B3LYP/6-311++G(d,p) method. From SCF energy calculation we deduce that, the conformer C6 acquire dominant stability among others as shown in Table 1. The electronic structure of 5-NSA monomer and dimer of C6 conformer in the ground state were optimized by assuming C1 point group symmetry. The optimized geometrical parameters and vibrational wavenumbers of monomer and dimer of C6 conformer are obtained from the same level of DFT theory. In the calculations, the charge of each point is taken as zero and the spin multiplicity is taken as one. Due to the neglect of anharmonicity effect at the beginning of calculation, initially the predicted vibrational wavenumbers by B3LYP/6-311++G(d,p) are found to be in disagreement with experimental wavenumbers. Further, these discrepancies are overcome by applying the selective scaling factors in the wavenumber region ranging from 4000 to 1700 cm1 and below 1700 cm1. Vibrational mode assignments made in this work are performed on the basis of total energy distribution (TED) results obtained from MOLVIB program (version V7.0-G77) written by Sundius [1618]. The molecular orbital (MO) calculations such as HOMOLUMO and NBO calculations are also performed on the monomer and dimer conformer of 5-NSA with the same level of DFT theory. 4. Results and discussion The present compound under investigation has become a greater interest because it has three different substituents namely hydroxyl group (AOH), carboxyl group (ACOOH), and Nitro group
Table 1 Calculated energies and energy difference for eight conformers of 5-NSA by B3LYP/6311++G(d,p) method. Conformers C1 C2 C3 C4 C5 C6 C7 C8 Energy (a.u) 700.65161 700.65617 700.65005 700.55947 700.64835 700.66584 700.63449 700.65287 Energy differencesa (a.u) 0.01423 0.00967 0.01579 0.10637 0.01749 0.00000 0.03135 0.01297

(ANO2) that are attached to the benzene ring. The molecular energies of eight possible conformers of the title molecule are calculated using density functional theory with 6-311++G(d,p) basis set. From the calculations, the most stable conformer is identied as C6 (EC6 = 700.66584 a.u) and it is also found that, the conformer C4 (EC4 = 700.63449 a.u) is the least stable conformer among others as shown in Table 1. The stable conformer structure C6 shown in Fig. 1 contains intramolecular (known as hydrogen bond) O H bonding between the hydroxyl group and O@COH. The intramolecular hydrogen bonding arising in the conformer C6 causes dominant stability of molecule among the others. 4.1. Dimer entity The Fig. 2 presents the geometry of 5-NSA dimer optimized at B3LYP/6-31++G(d,p) level using Gaussian 03W program package. The dimer entities in 5-NSA can be proved by shaping the structure by joining high-frequency OAH stretching and low-frequency O O stretching mode. The basic mechanism by coupling high-frequency OAH and low-frequency O O band is known as anharmonic-type coupling [1]. The dimer structure shown in Fig. 2 contains two intermolecular and two intramolecular hydrogen bonds which are similar as in the model given by Boczar [1], Marechal and Witkowski [19]. The energy of the dimer structure calculated at B3LYP/6-31+G(d,p) is about (EDim = 1401.18601 a.u) and it is also found that, the energy of 5-NSA dimer is found to be twice that of its stable monomer structure. 4.2. Geometrical parameters The predicted geometrical parameters such as, bond lengths and bond angles for the most stable C6 conformer and the dimer of C6 conformer of 5-NSA calculated at B3LYP/6-311++G(d,p) method are presented in Table 2 in accordance with atom numbering schemes given in Figs. 1 and 2 respectively. For comparative purpose, the experimental X-ray diffraction data [20] of 5-NSA is also presented. The intermolecular interaction through ACOOH group of each 5-NSA molecule forms a dimer and intramolecular interaction through OAH O connecting the adjacent dimers may lead to innite chains in random directions. When comparing experimental values, the computed bond lengths, bond angles and vibrational frequencies are slightly larger, because theoretical calculations are performed upon the isolated molecule in the gaseous state and the experimental results are performed on the solid phase of the molecule [21]. The experimental CAC bond lengths of the aromatic ring fall in the range from 1.3749 to 1.4059 , while the results obtained from B3LYP/6311++G(d,p) fall in the range 1.37891.4063 for the monomer structure and 1.37851.4068 for the dimer structure. In contrast, carbon atom C7 in carboxylic group attached to ring C1 makes bond length C1AC7 longer than that of ring CAC. The experimentally observed C1AC7 bond length (1.4772 ) is approximately 0.1 greater than that of ring CAC which is in good agreement with the calculated value. The CAH bond lengths such as C3AH13, C4AH14, and C6AH18 calculated at B3LYP/6-311++G(d,p) are too long in comparison with the experimental values. It is well known that DFT method predict bond lengths that are systematically too long, especially in CAH bond lengths [2]. This may be due to the fact that, the low scattering factor of hydrogen atoms involved in the X-ray diffraction experiment produces large deviation from the theoretical CAH bond lengths. According to international crystallography table [22] the C@O and CAO bond lengths in the aromatic carboxylic group conform to an average value of 1.2260 and 1.3050 respectively. The experimental C@O and CAO bond lengths of the title molecule are 1.2376 and 1.3105 [20] respectively. The calculated value

a Energies of the other seven conformers relative to the most stable C6 conformer.

194

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201

C1

C2

C3

C4

C5

C6

C7

C8

Fig. 1. The possible theoretical conformers of 5-NSA.

Fig. 2. C6 dimer conformer of 5-NSA.

of CAO (1.3135 ) in 5-NSA dimer reported in Table 2 is in good agreement with experimental data. In contrast, the calculated C@O bond length of dimer is slightly larger than that of experimental. This is because of the fact that, upon dimerization one can nd that the electron density of atoms delocalized from a lled lone pair of lewis base to an unlled lewis acid. The similar effect was obtained on the angle C7AO8AH9 and C1AC7AO10. The experimental and the calculated values for the nitro group bond lengths such as N15AO16 and N15AO17 shows considerable double bond character type. Inter- (H27 O10, O28 H9) and intramolecular (O10 H12, O28 H30) bond distances which causes stabilization of the monomer and dimer structure of 5-NSA are also reported in Table 2. The optimized hexagonal CACAC bond angles fall in the range 119121 except C2AC1AC7 (124) and C6AC1AC7 (116). The hexagonal ring CACAH angles are found to be 120. The CACAO bond

angles of 5-NSA are greatly affected by H-bonding interactions as shown in Table 2. 4.3. NBO analysis In order to explain the hyperconjugative interactions, inter-, intramolecular hydrogen bonding, intermolecular charge transfer (ICT), electron density transfer (EDT) and cooperative effect due to delocalization of electron density from the lled lone pairs n(Y) of Lewis base Y into the unlled antibond r(XH) of Lewis acid XAH in XAH Y hydrogen bonding systems, one can nd natural bond orbital (NBO) analysis as an effective tool [23]. In the present work, NBO analysis has been performed on 5-NSA monomer and dimer to study the intermolecular hydrogen bonding, intermolecular charge transfer (ICT), delocalization of electron density and cooperative effect due to n(O) ?r(OAH) using NBO 3.1 program

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201 Table 2 Comparison of geometrical parameters, bond lengths (), and bond angles (), for the monomer and dimer C6 conformer of 5-NSA calculated by the B3LYP/6-311++G(d,p) method. Parametersa Experimental X-rayb B3LYP/6-311++G(d,p) Monomer Bond lengths () C1AC2 C1AC6 C1AC7 C2AC3 C2AO11 C3AC4 C3AH13 C4AC5 C4AH14 C5AC6 C5AN15 C6AH18 C7AO8 C7AO10 O8AH9 O11AH12 N15AO16 N15AO17 1.4059 1.3996 1.4772 1.4052 1.3466 1.3749 0.9500 1.3944 0.9500 1.3794 1.4625 0.9500 1.3105 1.2376 0.8400 0.8400 1.2320 1.2261 1.4063 1.3993 1.4695 1.4057 1.3335 1.3789 1.0824 1.4007 1.0816 1.3823 1.4704 1.0806 1.3423 1.2243 0.9692 0.9841 1.2258 1.2256 1.7461 119.6 118.9 121.5 119.4 122.9 117.7 120.4 118.3 121.3 119.6 121.2 119.3 121.5 119.3 119.2 119.5 120.5 120.0 121.6 114.6 123.8 107.6 108.3 117.5 117.8 124.7 145.1 145.1 Dimer 1.4068 1.4009 1.4682 1.4058 1.3340 1.3785 1.0824 1.4010 1.0816 1.3809 1.4716 1.0806 1.3135 1.2454 0.9972 0.9812 1.2257 1.2250 1.7544 1.6798 1.6807 119.5 120.0 120.5 119.5 123.2 117.4 120.5 118.2 121.3 119.5 121.3 119.3 121.5 119.3 119.2 119.6 120.3 120.1 121.8 115.8 122.4 110.7 108.3 117.5 117.8 124.8 144.5 144.5 179.9 179.9

195

Inter- and intra-molecular H bond lengths O11AH12 O10 1.9100 O26AH27 O10 O8AH9 O28 Bond angle () C2AC1AC6 C2AC1AC7 C6AC1AC7 C1AC2AC3 C1AC2AO11 C3AC2AO11 C2AC3AC4 C2AC3AH13 C4AC3AH13 C3AC4AC5 C3AC4AH14 C5AC4AH14 C4AC5AC6 C4AC5AN15 C6AC5AN15 C1AC6AC5 C1AC6AH18 C5AC6AH18 C1AC7AO8 C1AC7AO10 O8AC7AO10 C7AO8AH9 C2AO11AH12 C5AN15AO16 C5AN15AO17 O16AN15AO17 119.6 120.5 119.9 119.9 123.6 116.6 120.3 119.9 119.9 119.1 120.4 120.4 122.1 119.3 118.6 119.0 120.5 120.5 121.6 114.6 123.8 109.5 109.5 118.4 118.3 123.4

The NBO analysis of 5-NSA monomer and dimer clearly give the evidences for the formation of two strong H-bonded interactions between oxygen lone electron pairs and r(OAH) antibonding orbitals. The occupancies and their respective energies of oxygen lone pairs and antibonding orbitals which are responsible for the stabilization of H-bonded monomer and dimer entities of 5-NSA are given in Table 3. The magnitude of charges transferred from the lone pairs of Lewis base sites n1(O10), n1(O28) into antibonding Lewis acid sites r(O11AH12), r(O29AH30) respectively are signicantly changed upon dimerization as given in Table 3. The change in occupancies and their respective change in energies of Lewis base and Lewis acid sites upon dimerization directly give the evidence for the existence of intramolecular interactions within the molecule. In addition to occupancy numbers and energies of various Lewis base-acid sites, the stabilization energies E(2) associated with the hyperconjugative interactions are given in Table 4. The stabilization energies corresponding to hyperconjugative interactions n1(O10) ? r(O11AH12), n2(O10) ? r(O11AH12), n1(O28) ? r(O29AH30), n2(O28) ? r(O29AH30) which are responsible for the stabilization of the molecule are obtained as 5.57, 9.89, 5.54, 9.90 kcal mol1, respectively. The differences in E(2) energies are reasonably due to the fact that the accumulation of electron density in the OAH bond is not only drawn from the n(O) of hydrogen-acceptor but also from the entire molecule. Furthermore, the composition of H-bonded NBOs of 5-NSA is presented in Table 5. It is also observed that, the s-character of C7AO10 hybrid orbital decreases (8.46%) from sp2.21 to sp2.11 upon dimerization. This leads to a conspicuous weakening of C7AO10 bond and its elongation. This shows the existence of a mesomeric structure characterized by delocalization of electron density from r(C7AO10) antibonding orbital to the remaining part of the molecule. This is quite possible because the energy of r(C7AO10) antibonding orbital (0.48102 a.u) is higher than the energy of r(O11AH12) antibonding orbital (0.36845 a.u) which supports the likelihood of the delocalization of ED from C7AO10 to O11AH12 region. This is clearly reected in the geometry as bond C7AO10 elongates to an amount of 0.0211 with respect to the theoretical monomer upon dimerization. Further, the second order perturbation theory analysis of Fock matrix in NBO basis shows that n1(O10) can readily interact with r(C7AO8) and r(O11AH12) antibonding orbitals. It is also observed that, the s-character of C7O8 increases (0.74%) from sp2.56 to sp2.47 upon dimerization. This results in the strengthening of C7O8 bond and its contraction. On the other hand, the delocalization of ED from C7O8 to O8H9 increases the bond length of C7O8 by an amount 0.0288 with respect to the monomer. 4.4. HOMOLUMO energy gap An analysis of the electron density of highest occupied molecular orbitals (HOMO) and lowest unoccupied molecular orbitals (LUMO) of 5-NSA can give us some idea about the ground and excited state proton transfer processes. Both HOMO and LUMO of 5-NSA monomer and dimer are of p type, but their phases are quite different as shown in Figs. 3 and 4, respectively. The energies corresponding to various HOMO and LUMO levels of 5-NSA are performed by density functional B3LYP/6-311++G(d,p) method. The HOMOLUMO energy calculation reveals that, there are 47 occupied and 48 unoccupied molecular orbitals associated with 5-NSA monomer. The energies corresponding to the highest occupied and lowest unoccupied molecular orbitals of 5-NSA monomer are found to be 0.2835 and 0.1125 a.u respectively as shown in Table 6. The energy gap between various occupied and unoccupied molecular orbitals of 5-NSA calculated at B3LYP/6-311++G(d,p) level is given in Table 6 reects the chemical reactivity of the molecule. LUMO as an electron acceptor represents the ability to obtain

Inter- and intra-molecular H bond angles O11AH12 O10 144.0 O29AH30 O28 144.0 O8AH9 O28 O26AH27 O10
a b

For numbering of atom see Fig. 1. Taken from Ref. [21].

as implemented in Gaussian 03 W package. The intramolecular OAH O hydrogen bonding is formed by the orbital overlap between n(O) and r(OAH) which results ICT causing stabilization of the H-bonded systems. Hence hydrogen bonding interaction leads to an increase in electron density (ED) of OAH antibonding orbital. The increase in population of OAH antibonding orbital weakens the OAH bond. Thus the nature and strength of the intramolecular hydrogen bonding can be explored by studying the changes in electron densities in vicinity of O H hydrogen bonds.

196

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201 Table 5 Composition of hydrogen bonded NBOs in terms of natural atomic hybrids. H-bonded NBOs Dimer 0.780 0.368 0.780 0.368 0.346 0.353 0.482 0.070 0.342 0.368 Monomer sp 31.08% 37.49% 62.51% 0.805 0.678 sp2.56 28.01% 32.27% 67.73% 0.805 0.667
2.21

Table 3 Occupancies and energies of interacting Lewis base and Lewis acid sites. Parameters Occupancy (e) Monomer n1(O10) r(O11AH12) n1(O28)a r(O29AH30)a r(C20AO29)a r(C25AO26)a r(C25AO28)a p(C25AO28)a r(O26AH27)a r(O29AH30)a
a

Energy (a.u)

Dimer sp 22.62% 34.59% 65.41% 0.829 0.717 sp2.47 28.75% 33.10% 66.90% 0.829 0.663
2.11

DNBO
s 8.46 2.9 +2.9% +0.024 0.039 s +0.74 +0.83 0.83% +0.024 0.004

Dimer 1.946 0.042 1.946 0.042 0.019 0.061 0.026 0.350 0.063 0.042

Docc.
0.034 0.008 0.034 0.008 0.006 0.002 0.002 0.046 0.057 0.008

Monomer 0.817 0.439 0.817 0.439 0.232 0.179 0.227 0.117 0.369 0.439

DE
0.037 0.071 0.037 0.071 0.114 0.174 0.255 0.047 0.027 0.071

1.980 0.034 1.980 0.034 0.025 0.063 0.028 0.304 0.006 0.034

sp (C7AO10) %s char. %P char. of %P char. of q(C7)/e q(O10)/e spn(C7AO8) %s char. %P char. of %P char. of q(C7)/e q(O8)/e

C7 O10

C7 O8

Values for monomer are taken from the identical NBOs of other unit.

Table 4 Second order perturbation theory analysis of Fock Matrix in NBO basis. Donor (i) Within Unit n1(O10) n2(O10) n1(O10) n2(O10) Acceptor (j) 1 E(2)a (kcal/mol) 1.89 21.75 5.57 9.89 11.08 14.77 11.04 14.73 1.88 21.73 5.54 9.90 E(j)E(i)b (a.u) 1.13 0.71 1.15 0.72 1.12 0.70 1.12 0.70 1.13 0.71 1.15 0.72 F(i, j)c (a.u) 0.041 0.113 0.072 0.078 0.100 0.092 0.100 0.092 0.041 0.113 0.071 0.078

r(C7AO8) r(C7AO8) r(O11AH12) r(O11AH12)

From Unit 1 to Unit 2 n1(O10) r(O26AH27) n2(O10) r(O26AH27) From Unit 2 to Unit 1 n1(O28) r(O8AH9) n2(O28) r(O8AH9) Within Unit n1(O28) n2(O28) n1(O28) n2(O28)
a b c

r(C25AO26) r(C25AO26) r(O29AH30) r(O29AH30)

E(2) means energy of hyperconjugative interactions. Energy difference between donor and acceptor i and j NBO orbitals. F(i, j) is the Fock matrix element between i and j NBO orbitals.

an electron. HOMO represents the ability to donate an electron. HOMO orbital on the aromatic ring of 5-NSA (Fig. 3a) is primarily of anti-bonding character type over C1, C2, C3 atoms, whereas C4, C5, C6 and O11, H12 show bonding character. Both the hydroxyl and carbonyl oxygen having bonding character type, with a larger electron density over hydroxyl oxygen. A HOMO 1 orbital shown in Fig. 3c has no amplitude [24] over the aromatic ring, whereas the orbital overlap on the nitro group of 5-NSA shows considerable double bond character type. The HOMO 2 orbital (Fig. 3e) on the aromatic ring show that the atoms C6, C1 having considerable bonding character, whereas the atoms C3, C4, C5 having anti-bonding character. In contrast, all the three LUMO surfaces shown in Fig. 3b, d and f are p in nature. Because if we look into the electronic distribution of LUMO within the ring, the C1, C2, C3 position have bonding character, whereas the C4, C6 position have anti-bonding character. The orbital overlapping on the 5-NSA dimer is shown in Fig. 4. It is worth mentioning that, bonding character positions on the left side of dimer structure is anti-bonding character in nature on its right side and vice versa.

between calculated and observed vibrational wavenumbers. To overcome discrepancies between observed and calculated wavenumbers, the scale factor of 0.958 and 0.983 are introduced in the region from 4000 to 1700 cm1 and lower than 1700 cm1, respectively [25]. The present molecule 5-NSA consists of 18 atoms, so it has 48 normal vibrational modes. On the basis of C1 symmetry the 48 fundamental vibrations of the title molecule can be distributed into 33 in-plane and 15 out-of-plane vibrations of same species. Figs. 5 and 6 present the experimental and calculated IR and Raman spectra for comparative purposes, where the calculated IR intensities and Raman activities are plotted against harmonic vibrational wavenumbers. The experimental wavenumbers are depicted in Table. 7 together with the calculated wavenumbers of monomer and dimer C6 conformer of 5-NSA molecule. The resulting vibrational wavenumbers for the optimized geometry and the proposed vibrational assignments as well as IR intensities (IIR) and Raman scattering activities (SRaman) are also given in Table 7. The complete vibrational assignments provided in this study are based on the total energy distribution (TED) results obtained from MOLVIB program [1618]. It is observed that, in solid 5-NSA the ACOOH and AOH groups are involved in inter and intramolecular hydrogen bonding interactions. In order to simulate hydrogen bonding through the ACOOH group, we also calculate vibrational wavenumbers of 5-NSA dimer of C6 conformer. 4.5.1. CAH vibrations In the present study, the three adjacent hydrogen atoms left around the benzene ring of 5-NSA give rise to three CAH stretching modes (m44Am46), three CAH in-plane bending (m30, m32, m33) and three CAH out-of-plane bending (m24, m26, m27) modes. The heteroaromatic organic molecule shows the presence of the CAH stretching vibrations in the 30003100 cm1 range which is the characteristic region for the identication of CAH stretching vibrations [26]. Accordingly, the CAH stretching modes of 5-NSA is assigned to 3095 and 3060 cm1 in FT-IR and 3054 cm1 in FT-Raman. These modes are calculated from 3099 to 3068 cm1 for the most stable C6 conformer. They are very pure modes since their TED contributions are almost 100%. For 3,5-dinitrosalicylic acid, the CAH stretching modes are assigned to 3106 cm1 in FT-IR and 3085 cm1 in FT-Raman [7]. For 5-BrSA, these modes are assigned to 3075 cm1 in FT-Raman and 3061 cm1 in FT-IR [2]. In aromatic compounds, the CAH in-plane and out-of-plane bending vibrations appear in the range 10001300 cm1 and 7501000 cm1 [27,28] respectively. Hence the CAH in-plane bending modes of 5-NSA are assigned to 1147, 1235 and 1300 cm1 in FT-IR and 1148, 1235 and 1305 cm1 in FT-Raman. The calculated values of this mode are 1151, 1247 and 1311 cm1 which shows better agreement with the experimental values. For 3,5-dichlorosalicylic acid, these bands are observed at

4.5. Vibrational spectra In order to obtain the spectroscopic signature of 5-NSA molecule, a frequency calculation is performed on the gaseous phase of the molecule, while experimental FT-IR and FT-Raman are performed on the solid phase of the molecule. Hence there are disagreements

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201

197

(a) HOMO

(b) LUMO

(c) HOMO1

(d) LUMO+1

(e) HOMO2

(f) LUMO+2

Fig. 3. Electron density plot for the frontier molecular orbitals of stable 5-NSA monomer.

1222, 1312 cm1 in FT-IR and 1152, 1226, 1318 cm1 in FT-Raman [8]. The TED contribution results at the last column of Table 7 show that, OAH in-plane bending vibrations interacting considerably with CAH in-plane bending mode. The CAH out-of-plane bending vibrations of 5-NSA are attributed to 863, 990 cm1 in FT-IR. FTRaman bands corresponding to these vibrations are not observed.

In the present study, the scaled theoretical values of CAH out-ofplane bending modes calculated at B3LYP/6-311++G(d,p) show good agreement with the experimental values of 5-NSA as well as with that of similar kind of molecules [2,8]. The TED contribution to these modes indicates that, CAH out-of-plane bending modes are also highly pure modes like CAH stretching modes.

198

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201

(a) HOMO

(b) LUMO

(c) HOMO1

(d) LUMO+1

(e) HOMO1

(f) LUMO+1

Fig. 4. Electron density plot for the frontier molecular orbitals of stable 5-NSA dimer.

4.5.2. Carboxylic acid group vibrations Generally carboxylic acid group containing molecules possesses dimeric character. The carboxylic acid dimer is formed by strong hydrogen bonding in the solid and liquid state. Hence the derivatives of carboxylic acids are best characterized by the carbonyl and hydroxyl groups. The presence of carbonyl group is the most important in the infrared spectrum because of its strong intensity of absorption and high sensitivity toward relatively minor changes in its environment. Intra- and intermolecular hydrogen bonding factors affect the carbonyl and absorptions in common organic compounds due to inductive, mesomeric, eld and conjugation effects [8]. The characteristic infrared absorption wavenumber of
Table 6 Selected occupied and unoccupied molecular orbital energies and energy gap of 5-NSA. Molecular orbitals MonomerB3LYP/6-311++G(d,p) Energy E (a.u) 0.2835 0.2839 0.3156 0.1125 0.1124 0.1071 Possible molecular orbital energy transition HOMO ? LUMO HOMO 1 ? LUMO HOMO 2 ? LUMO HOMO ? LUMO + 1 HOMO 1 ? LUMO + 1 HOMO 2 ? LUMO + 1 HOMO ? LUMO + 2 HOMO 1 ? LUMO + 2 HOMO 2 ? LUMO + 2

C@O in acids are normally strong in intensity and found in the region 16901800 cm1 [26,2]. In the present study, the strong band at 1665 cm1 in FT-IR and the band at 1657 cm1 in FT-Raman are assigned to C@O stretching. The calculated value of C@O stretching mode at B3LYP/6-311++G(d,p) shows good agreement with the experimental values. The CAO stretching of the carboxylic acid group of 5-NSA is highly coupled with the vibrations of adjacent group i.e OAH inplane bending. The band observed at 1507 cm1 in FT-IR and 1512 cm1 in FT-Raman are assigned to CAO stretching (m39) mode. The wavenumber of this mode calculated by DFT is in excellent agreement with the experimental FT-IR and FT-Raman

DimerB3LYP/6-311++G(d,p) Energy gap DE (a.u) 0.1710 0.1714 0.2031 0.1711 0.1715 0.2032 0.1764 0.1768 0.2085 Energy E (a.u) 0.2836 0.2840 0.3157 0.1125 0.1124 0.1070 Possible molecular orbital energy transition HOMO ? LUMO HOMO 1 ? LUMO HOMO 2 ? LUMO HOMO ? LUMO + 1 HOMO 1 ? LUMO + 1 HOMO 2 ? LUMO + 1 HOMO ? LUMO + 2 HOMO 1 ? LUMO + 2 HOMO 2 ? LUMO + 2 Energy gap DE (a.u) 0.1711 0.1715 0.2032 0.1712 0.1716 0.2033 0.1765 0.1769 0.2086

Occupied HOMO HOMO 1 HOMO 2 Unoccupied LUMO LUMO + 1 LUMO + 2

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201

199

100

Transmittance (%)

80 60 40 20 0 4000

3000

2000

1500

1000

400

Wavenumber

(cm-1)

Transmittance (%)
4000

3500

3000

2500

2000

1500

1000

500

Wavenumber

(cm-1)

Fig. 5. Experimental (FT-IR), calculated IR spectrum of 5-NSA.

wavenumbers. The OAH stretching vibrations are characterized by a very broad band appearing in the region 34003600 cm1 [2]. In the present study, the band for this mode is not active in FT-Raman spectrum. Hence the band observed at 3610 cm1 in FT-IR is assigned to OAH stretching of carboxylic acid group of 5-NSA. The scaled theoretical value 3613 cm1 by B3LYP/6-311++G(d,p) is in good agreement with OAH stretching of similar kind of molecules [2,8]. For example in the case of 5-BrSA, the OAH stretching of carboxylic acid group is assigned to 3551 cm1 [2]. The scaled theoretical value of this mode reported by Mehmet Karabacak et al. is also in good agreement with that of this title molecule. In the case of carboxylic acid containing dimer structure, the OAH in-plane bending and CAO stretching bands involve some interaction be100

tween them. Hence these are referred to as coupled OAH in-plane and CAO stretching vibrations [29]. This is also conrmed by TED output results. The strong band observed at 1470 cm1 in FT-IR and 1475 cm1 in FT-Raman are assigned to OAH in-plane bending of carboxylic acid group. In this mode, the TED contribution of CAO stretching is signicant. 4.5.3. Hydroxyl group vibrations The CAO stretching and OAH bending modes of the hydroxyl group of 5-NSA are coupled with the vibrations of adjacent groups. Intramolecular hydrogen bonding between the hydroxyl and the carboxylic acid group of the title molecule alters the wavenumbers of CAO stretching and OAH bending vibrations. In the present

Raman Intensity

80 60 40 20 0 4000

3000

2000

1500

1000

500

Wavenumber (cm-1)

Transmittance (%)
4000

3500

3000

2500

2000

1500

1000

500

Wavenumber (cm-1)
Fig. 6. Experimental (FT-Raman), calculated Raman spectrum of 5-NSA.

200

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201

Table 7 Comparison of the experimental (FT-IR, FT-Raman wavenumbers (cm1)) and theoretical harmonic scaled wavenumbers (cm1), infrared intensities (IIR), Raman scattering activities (SRaman) of C6 monomer and dimer conformer of 5-NSA calculated by B3LYP/6-311++G(d,p) method. Mode no Experimental wavenumber Theoretically calculated at B3LYP/6-311++G(d,p) Monomer FT-IR FT-Raman 45 s 78 vs 115 ms 150 vw 161 vw 285 w 310 w 367 vw 450 w 512 w 645 ms 680 w 720 w 760 vs 780 s 795 w 829 w 928 vs 1077 ms 1130 ms 1148 s 1198 vw 1235 ms 1305 w 1340 vs 1475 vw 1512 w 1578 s 1625 w 1657 w 3054 s Wave number 47 77 126 145 158 276 303 350 358 424 441 505 507 570 572 651 678 686 722 765 780 791 839 855 932 952 997 1079 1134 1151 1195 1247 1311 1333 1364 1405 1427 1472 1507 1575 1592 1631 1663 3068 3085 3099 3342 3613 IIR 0.02 2.22 1.65 0.15 3.61 0.33 1.99 1.31 8.06 9.50 2.95 4.75 8.99 4.23 90.28 20.16 63.61 14.19 23.37 70.69 22.65 61.75 3.24 17.91 23.66 12.96 0.05 52.76 157.12 17.92 312.98 39.92 77.28 76.00 471.70 98.89 114.23 11.83 143.12 136.63 98.07 144.15 500.61 0.27 3.76 10.53 400.14 138.29 SRaman 0.85 0.35 0.13 0.96 0.69 1.85 0.61 1.81 1.32 0.20 3.61 0.23 2.52 1.20 2.19 3.04 3.81 0.25 0.99 0.53 33.06 0.01 0.22 0.04 10.91 0.07 0.02 5.04 41.78 4.40 29.94 12.48 3.70 75.73 323.44 27.11 28.28 21.11 7.00 25.81 71.69 27.70 76.17 86.55 99.12 33.11 121.10 136.31 46, 55 70, 71 120, 121 161, 172 187, 203 309 315, 317 351, 361 372 413 447, 448 524, 523 529, 525 587 596 650 652 685, 687 698, 702 748, 749 787 796, 798 832, 833 863, 865 931, 932 970, 971 1001, 1002 1098, 1099 1124, 1126 1147 1165, 1166 1239, 1240 1287, 1288 1321, 1330 1353, 1354 1377, 1378 1406, 1412 1508, 1514 1522, 1523 1596, 1598 1630, 1631 1671, 1675 1707, 1749 2979, 3080 3224, 3225 3248, 3249 3256, 3259 3589, 3595 Dimer wavenumber Assignmentsa/TED (%)

m1 m2 m3 m4 m5 m6 m7 m8 m9 m10 m11 m12 m13 m14 m15 m16 m17 m18 m19 m20 m21 m22 m23 m24 m25 m26 m27 m28 m29 m30 m31 m32 m33 m34 m35 m36 m37 m38 m39 m40 m41 m42 m43 m44 m45 m46 m47 m48
a

502 vw 570 s 696 vs 757 w 788 w 835 vs 863 w 990 s 1075 w 1135 vw 1147 w 1195 s 1235 s 1300 w 1355 w 1412 w 1470 ms 1507 vw 1580 ms 1622 s 1665 s 3060 w 3095 w 3610 vw

sNO2 (65), cring(22) cCO(52), cring(19) cCO(53), cring(22) cCCC(47), cring(30) cCANO2(49) cring(41), cCO(17) cring(43), cCO(27), cOH(14)
bCANO2(45), bring(12) bring(42), bCO(22), bOH(18) bCC(56), bOH(24), bring(13) cring(52), cCANO2(14) qNO2(49), bring(26), bCO(18) cCO(45), cCH(24) bring(62), bCN(18) cOH(76) bCO(58), bring(25), bOH(12) cOH(89) bring(52), bOH(17), NO2 sci.(11) bCCO(71), bOH(12), bring(10) xNO2(48), cCH(26), cCN(11) tCC(65), bring(22) bCO(63), bOH(12) dNO2(55), bring(13), tCO(12), tCN(11) cCH(78) tCN(42), bring(22), tCC(18) cCH(88) cCH(89) tCC(52) tCC(41), bOH(35), bCH(15) bCH(47), bOH(25) tCC (45), bOH(17), bCH(13) bCH(52), bOH(26) bCH(60), tCC(13) tsNO2(41), tCN(32), bCH(11) tCO(48), bOH(30) bOH(48), tCO(18), tCC(12) tCC(43), bCH(22), bOH(13) bOH(47), tCO(18), bCH(10) tCO(58), bOH(17), bCH(12) tasNO2(58), bOH(15), bCH(10) tCC(50), bCH(18), bCN(10) tCC(55), bOH(15), bCN(12) tC@O(72), bOH(19) tCH(98) tCH(98) tCH(100) tOH(100) tOH(100)

ts, symmetry stretching; tas, asymmetry stretching; b, in-plane bending; c, out-of-plane bending; d, scissoring; x, wagging; @, twisting; q, rocking; s, torsion.

study, the OAH in-plane bending mode appears as weak band at 1412 cm1 in FT-IR spectrum. For 5-BrSA, the OAH in-plane bending mode of the hydroxyl group is assigned to 1411 cm1 in FTRaman [2]. For 3-, 4-, 5-aminosalicylic acid, the title vibration is observed at 1461, 1450 and 1454 cm1, respectively [10]. The weak band at 1355 cm1 in FT-IR spectrum of 5-NSA is assigned to CAO stretching mode. The TED contribution for this mode is about 48%. It is observed that, the OAH in-plane bending mode is highly coupled with this mode is evident from TED result. The calculated value of 1364 cm1 at B3LYP/6-311++G(d,p) level of theory is in good agreement with the experimental value of CAO stretching mode. 4.5.4. NO2 vibrations Generally the asymmetric and symmetric stretching vibrations of NO2 give rise to bands in the regions 15001570 cm1 and 13001370 cm1 in nitro benzenes and substituted nitro benzenes

[30] respectively. In the title molecule, the medium band at 1580 cm1 in FT-IR and 1578 cm1 in FT-Raman is assigned to NO2 asymmetric stretching mode. The very strong band at 1340 cm1 in FT-Raman is attributed to NO2 symmetric stretching. The theoretically scaled values at 1575 and 1333 cm1 by B3LYP/6311++G(d,p) method exactly correlate with the experimental observations. Aromatic nitro compounds have a band of weak-to-medium intensity in the region of 500590 cm1 belongs to NO2 bending vibrations [26]. The nitro group of the compound is capable of different bending vibrations such as scissoring, wagging, rocking and twisting. These vibrations give rise to several variable intensity bands at lower wavenumber region. In the present study, the prominent band at 835 cm1 in FT-IR and weak band at 829 cm1 in FTRaman is assigned to NO2 scissoring (dNO2) mode. In the case of 3,5-dinitrosalicylic acid, the bands at 800 and 915 cm1 in the IR and at 798 cm1 in Raman spectra are assigned to dNO2 modes.

T. Karthick et al. / Journal of Molecular Structure 1005 (2011) 192201

201

In aromatic compounds, the wagging modes (xNO2) are expected in the region 690790 cm1 with a moderate to strong intensity. In this region, cCH is also active [31]. For 5-NSA, the weak band at 757 cm1 in FT-IR and a very strong band at 760 cm1 in FT-Raman is assigned to NO2 wagging (xNO2). The cCH contribution to this mode is about 26%. The NO2 rocking (qNO2) and NO2 twisting modes of the title molecule are not observed in FT-IR spectrum. Hence the weak band at 512 cm1 and a strong band at 45 cm1 in FT-Raman are assigned to NO2 rocking and twisting modes. The calculated values of all the NO2 bending modes at B3LYP/6-311++G(d,p) method are in excellent agreement with the experimental observations. 4.5.5. CAN vibrations NO2 group attached to ring carbon atom C5 of 5-NSA gives rise to three vibrational modes such as CAN stretching (tCN), CAN inplane bending (bCANO2) and CAN out-of-plane bending (cCANO2). The identication of CAN bands in the vibrational spectrum is rather a difcult task since these bands are mixed with the other vibrational modes. Varghese et al. [7], assigned the CAN stretching mode of 3,5-dinitrosalicylic acid at 939 cm1 in IR and 936 cm1 in Raman spectrum. In the present study, the prominent band at 928 cm1 in FT-Raman is assigned to tCN mode. On the other hand, the assignments of in-plane and out-of-plane CAN bending modes are made on the basis of TED results. These bands are calculated at 350 and 158 cm1 respectively. 4.5.6. Ring vibrations In case of 5-NSA, the benzene ring possesses six ring carbon carbon stretching vibrations in the region 14601660 cm1 and 10701150 cm1. As revealed by TED, the ring CAC stretching modes are observed at 1622, 1195, 1135 and 1075 cm1 in FT-IR and 1625, 1198, 1130 and 1077 cm1 in FT-Raman for 5-NSA. The in-plane and out-of-plane bending vibrations of the benzene ring are generally observed below 1000 cm1 [32] and these modes are not pure but contain a signicant contribution from other modes and are substituent-sensitive. In the title molecule, ring in-plane and out-of-plane bending modes are observed at 570, 696 cm1 in FT-IR and 367 cm1 in FT-Raman. From TED results, the bands at 285, 310, 450 cm1 in FT-Raman are assigned to ring out-of-plane bending vibrations. The peaks for these modes are not observed in FT-IR spectrum since these modes are possible to appear only in far IR spectrum. The scaled theoretical wavenumbers corresponding to ring vibrations are found to have a good correlation with the experimental observations. 5. Conclusion In this study, we have recorded FT-IR and FT-Raman spectra on the solid phase of 5-NSA. The comparative results between the experimental and theoretical wavenumbers gave us a full description of the vibrational properties of this molecule. The most stable conformer of 5-NSA was determined and according to these results the dimer conformations were analyzed with DFT/B3LYP/6311++G(d,p) level of theory. Inter- and intramolecular hydrogen bonding between the carboxylic acid and hydroxyl group of the title molecule were discussed. The signicant changes in bond lengths and vibrational wavenumbers of OAH group involving

hydrogen bond were discussed. The delocalization of ED and the changes in energy upon dimerization were analyzed with the aid of NBO analysis. HOMOLUMO calculations were performed on the stable monomer and the dimer of 5-NSA. The energy difference between various HOMO and LUMO levels were reported. The calculated geometrical parameters and vibrational frequencies obtained from density functional theory calculations are in good agreement with the experimental values obtained for the investigated molecule. References
[1] Marek Boczar, Lukasz Boda, Marek J. Wojcik, Spectrochim. Acta A 64 (2006) 757. [2] Mehmet Karabacak, J. Mol. Struct. 919 (2009) 215. [3] Mehmet Karabacak, Etem Kose, Mustafa Kurt, J. Raman Spectrosc. 41 (2010) 1085. [4] C. Chen, S.F. Shyu, J. Mol. Struct. (Theochem) 536 (2001) 25. [5] P.J.G. Goulet, R.F. Aroca, Can. J. Chem. 82 (2004) 987. [6] Y. Akkaya, S. Akuyz, Vib. Spectrosc. 42 (2006) 292. [7] H.T. Varghese, C.Y. Panicker, D. Philip, J. Chowdhury, M. Ghosh, J. Raman Spectrosc. 38 (2007) 323. [8] V. Krishnakumar, R. Mathamal, J. Raman Spectrosc. 40 (2009) 264. [9] P. Wen, J. Chen, S. Wan, W. Kong, P. Zhang, W. Wang, J. Zhan, Q. Pan, W. Huang, Plant Growth Regul. 55 (2008) 1. [10] Y.P. Singh, R. Das, R.A. Sigh, Afr. J. Biochem. Res. 1 (2) (2007) 19. [11] M.K. Jain, S.C. Sharma, Organic Chemistry, Shoban Lal Nagin Chand & Company, Educational Publishers, New Delhi, 1980. [12] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, J.A. Montgomery Jr., T. Vreven, K.N. Kudin, J.C. Burant, J.M. Millam, S.S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G.A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J.E. Knox, H.P. Hratchian, J.B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austn, R. Cammi, C. Pomelli, J.W. Ochterski, P.Y. Ayala, K. Morokuma, G.A. Voth, P. Salvador, J.J. Dannenberg, V.G. Zakrzewski, S. Dapprich, A.D. Daniels, M.C. Strain, O. Farkas, D.K. Malick, A.D. Rabuck, K. Raghavachari, J.B. Foresman, J.V. Ortiz, Q. Cui, A.G. Baboul, S. Clifford, J. Cioslowski, B.B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R.L. Martin, D.J. Fox, T. Keith, M.A. Al-Laham, C.Y. Peng, A. Nanayakkara, M. Challacombe, P.M.W. Gill, B. Johnson, W. Chen, M.W. Wong, C. Gonzalez, J.A. Pople, Gaussian, Inc., Gaussian 03, Revision B.01, Pittsburgh, PA, New York, 2003. [13] A.D. Becke, J. Chem. Phys. 98 (1993) 5648. [14] C. Lee, W. Yang, G.R. Parr, Phys. Rev. B 37 (1998) 785. [15] B. Miehlich, A. Savin, H. Stoll, H. Preuss, Chem. Phys. Lett. 157 (1989) 200. [16] T. Sundius, J. Mol. Struct. 218 (1990) 321. [17] T. Sundius, Vib. Spectrosc. 29 (2002) 89. [18] Molvib (V.7.0), QCPE Program No. 807, 2002. [19] Y. Marechal, A. Witkowski, J. Chem. Phys. 48 (1968) 3697. [20] National Crystallography Service. <http://www.ncs.chem.soton.ac.uk/>. [21] Mehmet Karabacak, Mehmet Cinar, Sahin Ermec, Mustafa Kurt, J. Raman Spectrosc. 41 (2010) 98. [22] F.H. Allen, Acta Crystallogr. B58 (2002) 380. [23] C. James, C. Ravikumar, Tom Sundius, V. Krishnakumar, R. Kesavamoorthy, V.S. Jayakumar, I. Hubert Joe, Vib. Spectrosc. 47 (2008) 10. [24] Mama Nsangou, Nejm-Eddine Jaidane, Zohra Ben Lakhdar, Int. Elect. J. Mol. Des. 5 (2006) 89. [25] N. Sundaraganesan, S. Ilakiamani, H. Saleem, P.M. Wojciechowski, D. Michalska, Spectrochim. Acta A 61 (2005) 2995. [26] M. Silverstein, G. Clayton Basseler, C. Morill, Spectrometric Identication of Organic Compounds, Wiley, New York, 1981. [27] G. Socrates, Infrared Characteristic Group of Frequencies, Wiley, New York, 1980. [28] G. Varasanyi, Assignments of Vibrational Spectra of Seven Hundred Benzene Derivatives, Wiley, New York, 1974. [29] Y.R. Sharma, Elementary Organic Spectroscopy, Shoban Lal Nagin Chand & Company, Educational Publishers, New Delhi, 1980. [30] D.N. Sathyanarayana, Vibrational Spectroscopy Theory and Applications, second ed., New Age International (P) Limited Publishers, New Delhi, 2004. [31] E.F. Mooney, Spectrochim. Acta 20 (1964) 1021. [32] N.P.G. Roeges, A Guide to the Complete Interpretation of Infrared Spectra of Organic Structures, Wiley, New York, 1994.

Potrebbero piacerti anche