10.8
CiteScore
 
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
10.8
CiteScore
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original article
10 (
2_suppl
); S2281-S2294
doi:
10.1016/j.arabjc.2013.08.004

Spectroscopic (FT-IR, FT-Raman), first order hyperpolarizability, NBO analysis, HOMO and LUMO analysis of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide by density functional methods

Department of Physics, St. Xavier’s College, Vaikom, Kothavara, Kottayam, Kerala, India
Department of Physics, Fatima Mata National College, Kollam, Kerala, India
Department of Physics, TKM College of Arts and Science, Kollam, Kerala, India
Department of Physics, Karpagam University, Pollachi Main Road, Eachanari, Coimbatore, Tamilnadu, India
Faculty of Pharmacy in Hradec Kralove, Charles University in Prague, Heyrovskeho 1203, Hradec Kralove 500 05, Czech Republic
Department of Chemistry, University of Antwerp, B2610 Antwerp, Belgium

⁎Corressponding author. Tel.: +91 9895370968. cyphyp@rediffmail.com (C. Yohannan Panicker)

Disclaimer:
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.

Peer review under responsibility of King Saud University.

Abstract

The optimized molecular structure, vibrational frequencies, corresponding vibrational assignments of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide have been investigated experimentally and theoretically using Gaussian09 software package. Potential energy distribution of normal modes of vibrations was done using GAR2PED program. The HOMO and LUMO analysis are used to determine the charge transfer within the molecule. The stability of the molecule arising from hyper-conjugative interaction and charge delocalization has been analyzed using NBO analysis. From the NBO analysis it is evident that the increased electron density at the nitrogen and carbon atoms leads to the elongation of their respective bond lengths and a lowering of their corresponding stretching wavenumbers. The calculated geometrical parameters are in agreement with that of similar derivatives. The calculated first hyperpolarizability is high and the calculated data suggest an extended π-electron delocalization over the pyrazine ring and carboxamide moiety which is responsible for the nonlinearity of the molecule.

Keywords

FT-IR
FT-Raman
Pyrazine
Carboxamide
PED
1

1 Introduction

Pyrazinecarboxamide is among the first line drugs for the treatment of tuberculosis and it kills or stops the growth of certain bacteria that cause tuberculosis (Opletalova et al., 2002). Recent years have seen increased incidence of tuberculosis in both developing and industrialized countries, the wide spread emergence of drug-resistant strains and a deadly synergy with human immunodeficiency virus (Raviglione et al., 1997; Houston and Fanning, 1996). Pyrazine and its derivatives form an important class of compounds present in several natural flavors and complex organic molecules (Endredi et al., 2003). 2-Chloropyrazine and 2,6-dichloropyrazine are mainly found as medical and agricultural drug intermediates (Endredi et al., 2003). The resistance of Pyrazinamide arises by the absence of the enzyme, Pmc A. The major side effect of Pyrazinamide is dose-related hepatotoxicity. Pyrazinoic acid disrupts membrane energetics and inhibits membrane transport function in Mycobacterium tuberculosis (Zhang et al., 2003). A different analog of Pyrazinamide, 5-chloropyrazine-2-carboxamide, has previously been shown to inhibit mycobacterial fatty acid synthase I (Ngo et al., 2007). Pyrazinamide is a member of the pyrazine family and it is known as a very effective antimycobacterial agent, with a well established role in tuberculosis treatment (Somoskovi et al., 2004). Pyrazinamide is bactericidal to semidormant mycobacteria and reduces total treatment time (Mitchison, 1996). The finding that pyrazinamide-resistant strains lose amidase (pyrazinamidase or nicotinamidase) activity and the hypothesis that amidase is required to convert pyrazinamide to pyrazinoic acid intracellularly led to the synthesis and study of various prodrugs of pyrazinoic acid (Bergmann et al., 1996). The activity of Pyrazinamide appears to be pH dependent, since it is bactericidal at pH 5.5, but inactive at neutral pH. Pyrazinamide is used in combinations with INH and rifampicin. It is especially effective against semi-dormant mycobacteria. Its mechanism of action appears to involve its hydrolysis to pyrazinoic acid via the bacterial enzyme pmcA (Speirs et al., 1995). Pyrazinamide can also be metabolized by hepatic microsomal deamidase to pyrazinioic acid, which is a substrate for xanthine oxidase, affording 5-hydroxypyrazinoic acid. The acid is believed to act as an antimetabolite of nicotinamide and interferes with NAD biosynthesis. Spleen tyrosine kinase activities of a series of aminopyrazine were reported by Forns et al. (2012).

Akyuz et al. (2007) reported the vibrational spectroscopic study of two dimensional polymer compounds of pyrazinamide. It has been reported that the amides of substituted pyridine-4-carboxylic acids (Miletin et al., 2000) as well as anilides of the substituted pyrazine-2-carboxylic acids (Dolezal et al., 2000, 2002, 2006) inhibited oxygen evolution rate in spinach chloroplasts and they showed some antialgal properties. Pyrazine derivatives are important drugs with antibacterial, diuretic, hypolipidemic, antidiabetic, hypnotic, anticancer and antiviral activities. Pyrazinamide, another first-line TB drug, was discovered through an effort to find antitubercular nicotinamide derivatives (Dolezal, 2006). Although the exact biochemical basis of pyrazinamide activity in vivo is not known, under acidic conditions it is thought to be a prodrug of pyrazinoic acid, a compound with antimycobacterial activity (Cynamon et al., 1992). Wang et al. (2011) reported the synthesis of a series of pyrimidine carboxamides and evaluated as cholecystokinin I receptor agonists.

The dynamical pattern of the 2-aminopyrazine-3-carboxylic acid molecule by inelastic and incoherent neutron scattering, Raman spectroscopy and ab initio calculations was reported by Pawlukojc et al. (2000). Billes et al. (1998) calculated the vibrational frequencies of the three parent diazines (pyrazine, pyridazine and pyrimidine) applying ab initio quantum chemical methods, Moller–Pleassett perturbation and local density function methods. Various compounds possessing -NHCO- groups, e.g. substituted amides, acyl and thioacyl anilides, benzanilides, phenyl carbamates, etc., were found to inhibit photo synthetic electron transport (Kralova et al., 2002, 2008). Therefore, the vibrational spectroscopic studies of the amides of pyrazine-2-carboxylic acids are added areas of interest.

In the present work, FT-IR and FT-Raman spectra of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide are reported both experimentally and theoretically. The HOMO and LUMO analysis have been used to elucidate information regarding charge transfer within the molecule. The stability of the molecule arising from hyper-conjugative interaction and charge delocalization has also been analyzed using NBO analysis.

2

2 Experimental

All the chemicals are purchased from Sigma Aldrich, Germany. Melting point was determined using a SMP 3 melting point apparatus (BIBBYB Stuart Scientific, UK) and is uncorrected. The reactions were monitored and the purity was checked by TLC (Merck UV 254 TLC plates, Darmstadt, Germany) using petroleum ether/EtOAc (9:1). Purification of compounds was made using a Flash Master Personal chromatography system from Argonaut Chromatography (Argonaut Technologies, Redwood City, CA, USA) with gradient of elution from 0% to 20% ethyl-acetate in hexane. As a sorbent, Merck Silica Gel 60 (0.040–0.063 mm) was used (Merck). Elemental analysis was performed on an automatic microanalyser CHNS-O CE instrument (FISONS EA 1110, Milano, Italy). 1H and 13C NMR spectra were recorded (at 300 MHz for 1H and 75 MHz for 13C) in CDCl3 solutions at ambient temperature on a Varian Mercury-Vx BB 300 instrument (Varian, Palo Alto, CA, USA). The chemical shifts were recorded as δ values in ppm and were indirectly referenced to tetramethylsilane (TMS) via the solvent signal (7.26 for 1H and 77.0 for 13C in CDCl3).

The pyrazinecarboxylic acid (50.0 mmol) and thionyl chloride (5.5 mL, 75.0 mmol) in dry toluene (20 mL) was refluxed for about 1 h (Dolezal et al., 2009, 2007). Excess thionyl chloride was removed by repeated evaporation with dry toluene in vacuo. The crude acyl chloride dissolved in dry acetone (50 mL) was added drop wise to a stirred solution of the 4-trifluoromethylaniline (50.0 mmol) and pyridine (50.0 mmol) in dry acetone (50 mL) kept at room temperature. After the addition was complete, stirring was continued for 30 min., then the reaction mixture was poured into cold water (100 mL) and the crude amide was collected and purified by column chromatography.

N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide. Yield 71%. Anal. Calcd. for C12H8F3N3O (267.2), 53.94% C, 3.02% H, 15.73% N; Found: 54.08% C, 3.02% H, 15.98% N; MP 176.0–177.5 °C; Log P: 1.51; Clog P: 2.40270.

The FT-IR spectrum (Fig. 1) was recorded using KBr pellets on a DR/Jasco FT-IR 6300 spectrometer in KBr pellets. The spectral resolution was 2 cm−1. The FT-Raman spectrum (Fig. 2) was obtained on a Bruker RFS 100/s, Germany. For excitation of the spectrum the emission of Nd:YAG laser was used, excitation wavelength 1064 nm, maximal power 150 mW, measurement on solid sample. The spectral resolution after apodization was 2 cm−1.

FT-IR spectrum of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide: (a) experimental and (b) theoretical (SDD).
Figure 1
FT-IR spectrum of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide: (a) experimental and (b) theoretical (SDD).
FT-Raman spectrum of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide: (a) experimental and (b) theoretical (SDD).
Figure 2
FT-Raman spectrum of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide: (a) experimental and (b) theoretical (SDD).

3

3 Computational details

Calculations of the title compound were carried out with Gaussian09 software (Gaussian 09, Revision B.01, Frisch et al., 2010) program using B3LYP/6-31G, B3PW91/6-31G and B3LYP/SDD basis sets to predict the molecular structure and vibrational wavenumbers. Calculations were carried out with Becke’s three parameter hybrid model using the Lee–Yang–Parr correlation functional (B3LYP) method. Molecular geometries were fully optimized by Berny’s optimization algorithm using redundant internal coordinates. Harmonic vibrational wavenumbers were calculated using analytic second derivatives to confirm the convergence to minima on the potential surface. The Stuttagard/Dresden effective core potential basis set (SDD) (Hay and Wadt, 1985) was chosen particularly because of its advantage of using faster calculations with relatively better accuracy and structures (Zhao et al., 2004). Then frequency calculations were employed to confirm the structure as minimum energy points. At the optimized structure (Fig. 3) of the examined species, no imaginary wavenumber modes were obtained, proving that a true minimum on the potential surface was found. The DFT method tends to over estimate the fundamental modes; therefore a scaling factor of 0.9613 has to be used for obtaining a considerably better agreement with experimental data (Foresman and Frisch, 1996). The observed disagreement between theory and experiment could be a consequence of the anharmonicity and of the general tendency of the quantum chemical methods to over estimate the force constants at the exact equilibrium geometry. The optimized geometrical parameters (B3LYP/SDD) are given in Table 1. The assignments of the calculated wavenumbers are aided by the animation option of GAUSSVIEW program, which gives a visual presentation of the vibrational modes (Dennington et al., 2009). The potential energy distribution (PED) is calculated with the help of GAR2PED software package (Martin and Van Alsenoy, 2007).

Optimized geometry (SDD) of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide.
Figure 3
Optimized geometry (SDD) of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide.
Table 1 Optimized geometrical parameters (SDD) of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide, atom labeling according to Fig. 3.
Bond lengths (Å) Bond angles (°) Dihedral angles (°)
C1–C2 1.4121 C2–C1–N6 121.0 C6–C1–C2–N3 0.0
C1–N6 1.3517 C2–C1–H27 121.5 C6–C1–C2–H7 −180.0
C1–H27 1.0850 N6–C1–H27 117.5 H27–C1–C2–N3 −180.0
C2–N3 1.3577 C1–C2–N3 121.8 H27–C1–C2–H7 0.0
C2–H7 1.0853 C1–C2–H7 121.3 C2–C1–N6–C5 0.0
C4–N3 1.3577 N3–C2–H7 117.0 H27–C1–N6–C5 −180.0
C4–C5 1.4104 C2–N3–C4 117.1 C1–C2–N3–C4 −0.0
C4–H8 1.0834 N3–C4–C5 121.2 H7–C2–N3–C4 180.0
C5–N6 1.3608 N3–C4–H8 118.2 C2–N3–C4–C5 0.0
C5–C9 1.5079 C5–C4–H8 120.6 C2–N3–C4–H8 −180.0
C9–N19 1.3782 C4–C5–N6 121.4 N3–C4–C5–N6 0.0
C9–O20 1.2580 C4–C5–C9 120.6 N3–C4–C5–C9 −180.0
C10–C11 1.4157 N6–C5–C9 118.0 H8–C4–C5–N6 −180.0
C10–C15 1.4171 C1–N6–C5 117.5 H8–C4–C5–C9 0.0
C10–N19 1.4120 C5–C9–N19 112.8 C4–C5–N6–C1 −0.0
C11–C12 1.4021 C5–C9–O20 121.2 C9–C5–N6–C1 180.0
C11–H16 1.0819 N19–C9–O20 126.1 C4–C5–C9–N19 −180.0
C12–C13 1.4069 C11–C10–C15 119.6 C4–C5–C9–O20 0.0
C12–H17 1.0856 C11–C10–N19 123.0 N6–C5–C9–N19 0.0
C13–C14 1.4098 C15–C10–N19 117.3 N6–C5–C9–O20 −180.0
C13–C23 1.4956 C10–C11–C12 119.5 C5–C9–N19–C10 −180.0
C14–C15 1.3980 C10–C11–H16 119.7 C5–C9–N19–H21 −0.0
C14–H18 1.0858 C12–C11–H16 120.8 O20–C9–N19–C10 0.0
C15–H22 1.0878 C11–C12–C13 120.7 O20–C9–N19–H21 −180.0
N19–H21 1.0209 C11–C12–H17 119.4 C15–C10–C11–C12 −0.3
C23–F24 1.4025 C13–C12–H17 119.9 C15–C10–C11–H16 180.0
C23–F25 1.4069 C12–C13–C14 120.0 N19–C10–C11–C12 180.0
C23–F26 1.4163 C12–C13–C23 120.4 N19–C10–C11–H16 −0.1
C14–C13–C23 119.6 C11–C10–C15–C14 0.2
C13–C14–C15 119.8 C11–C10–C15–H22 −179.7
C13–C14–H18 120.3 N19–C10–C15–C14 −180.0
C15–C14–H18 120.0 N19–C10–C15–H22 0.1
C10–C15–C14 120.5 C11–C10–N19–C9 −0.0
C10–C15–H22 119.8 C11–C10–N19–H21 180.0
C14–C15–H22 119.7 C15–C10–N19–C9 −179.8
C9–N19–C10 128.6 C15–C10–N19–H21 0.2
C9–N19–H21 113.5 C10–C11–C12–C13 0.2
C10–N19–H21 118.0 C10–C11–C12–H17 179.6
C13–C23–F24 113.0 H16–C11–C12–C13 −179.7
C13–C23–F25 112.9 H16–C11–C12–H17 −0.3
C13–C23–F26 113.1 C11–C12–C13–C14 −0.2
F24–C23–F25 106.7 C11–C12–C13–C23 −178.0
F24–C23–F26 105.5 H17–C12–C13–C14 −179.6
F25–C23–F26 105.1 H17–C12–C13–C23 2.6
C12–C13–C14–C15 0.2
C12–C13–C14–H18 179.5
C23–C13–C14–C15 178.0
C23–C13–C14–H18 2.7
C12–C13–C23–F24 24.6
C12–C13–C23–F25 −145.7
C12–C13–C23–F26 95.2
C14–C13–C23–F24 157.6
C14–C13–C23–F25 36.5
C14–C13–C23–F26 −82.6
C13–C14–C15–C10 −0.2
C13–C14–C15–H22 179.8
H18–C14–C15–C10 −179.6
H18–C14–C15–H22 0.4

4

4 Results and discussion

4.1

4.1 IR and Raman spectra

The observed IR, Raman bands and calculated wavenumbers (scaled) and assignments are given in Table 2. The phenyl CH stretching modes occur above 3000 cm−1 and are typically exhibited as multiplicity of weak to moderate bands compared with the aliphatic CH stretching (Coates, 2000). In the present case, the SDD calculations give υCH modes of the phenyl rings in the range 3149–3076 cm−1 and the bands are observed at 3023 cm−1 in the IR spectrum and 3116 and 3039 cm−1 in the Raman spectrum The benzene ring possesses six ring stretching vibrations, of which the four with the highest wavenumbers (occurring near 1600, 1580, 1490 and 1440 cm−1) are good group vibrations (Roeges, 1994). With heavy substituents, the bands tend to shift to somewhat lower wavenumbers. In the absence of ring conjugation, the band at 1580 cm−1 is usually weaker than that at 1600 cm−1. In the case of C⚌O substitution, the band near 1490 cm−1 can be very weak. The fifth ring stretching vibration is active near 1315 ± 65 cm−1, a region that overlaps strongly with that of the CH in-plane deformation. The sixth ring stretching vibration, or the ring breathing mode, appears as a weak band near 1000 cm−1, in mono-, 1,3-di- and 1,3,5-trisubstituted benzenes. In the otherwise substituted benzenes, however, this vibration is substituent sensitive and difficult to distinguish from the ring in-plane deformation (Roeges, 1994; Varsanyi, 1974). The υPh modes are expected in the region 1280–1630 cm−1 for para substituted benzene rings (Roeges, 1994). The SDD calculations give the Ph stretching modes in the range 1334–1598 cm−1 for the title compound. The bands observed at 1599, 1580, 1401, and 1328 cm−1 in the IR spectrum and 1580, 1387, and 1323 cm−1 in the Raman spectrum are assigned as the phenyl ring stretching modes for the title compound. Most of the modes are not pure but contain significant contributions from other modes also. The ring breathing mode of the para-substituted benzenes with entirely different substituents (Varsanyi, 1974) has been reported in the interval 780–880 cm−1. For the title compound, the ring breathing mode Ph is observed at 771 cm−1 in the IR spectrum, and 765 cm−1 in the Raman spectrum, which is supported by the computational result at 771 cm−1. For para-substituted benzene, the ring breathing mode was reported at 804 and 792 cm−1 experimentally and 782 and 795 cm−1 theoretically (Mary et al., 2008; Ambujakshan et al., 2007). The in-plane and out-of-plane CH deformations of the phenyl ring are expected above and below 1000 cm−1, respectively (Roeges, 1994). The in-plane CH deformation bands are assigned at 1308, 1189, 1137, and 1010 cm−1 for the phenyl ring theoretically (SDD) and bands are observed at 1188 and 1153 cm−1 in the Raman spectrum. Generally, the CH out-of-plane deformations with the highest wavenumbers have a weaker intensity than those absorbing at lower wavenumbers (Roeges, 1994). The strong γCH occurring at 840 ± 50 cm−1 is typical for 1,4-disubstituted benzenes and the band observed at 839 cm−1 in the IR spectrum is assigned to this mode which finds supports from the computation value 845 cm−1 (Roeges, 1994). Again according to the literature (Roeges, 1994; Varsanyi, 1974) a lower γCH absorbs in the neighborhood 820 ± 45 cm−1, but is much weaker or infrared inactive. The SDD calculations give a γCH at 816 cm−1 and no band is experimentally observed for this mode. The out-of-plane CH deformation bands of the phenyl ring are assigned at 976, 959, 845, and 816 cm−1 for Phenyl ring theoretically. Most of the modes are not pure according to PED calculations, but contain significant contributions from other modes also. The substituent sensitive modes of the phenyl rings and other modes are also identified and assigned.

Table 2 Calculated (scaled) wavenumbers, IR, Raman bands and assignments of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide.
B3PW91/6-31G B3LYP/6-31G B3LYP/SDD IR Raman Assignmentsa
υ (cm−1) IRI RA υ (cm−1) IRI RA υ (cm−1) IRI RA υ (cm−1) υ (cm−1)
3396 95.58 238.84 3391 85.89 231.72 3388 101.81 261.27 3343 3349 υNH(99)
3158 5.93 44.65 3155 5.18 41.85 3149 4.83 46.88 υCH(97)
3141 0.46 96.23 3133 0.17 88.55 3133 1.57 94.81 3127 3139 υCHPz(97)
3128 26.00 223.20 3116 28.54 237.78 3122 27.75 187.91 υCHPz(97)
3123 1.78 113.84 3113 2.29 113.03 3111 1.02 118.29 3116 υCH(98)
3120 1.79 75.65 3109 2.36 80.04 3109 3.20 49.52 υCH(96)
3110 1.15 82.52 3098 1.45 83.81 3099 0.16 64.48 3090 3080 υCHPz(100)
3087 8.77 47.47 3076 9.97 48.74 3076 9.56 40.21 3023 3039 υCH(99)
1653 133.52 250.07 1636 123.20 307.31 1610 121.67 691.92 1617 1616 υCO(67), υPh(13)
1625 75.19 301.01 1613 47.66 255.01 1598 105.09 40.43 1599 υPh(64), υCO(20)
1597 192.62 183.40 1586 224.87 183.95 1573 150.42 79.92 1580 1580 δNH(17), υPh(68)
1555 16.98 474.67 1542 25.76 500.25 1526 32.13 617.66 1533 1528 υPz(52), δCHPz(24)
1538 510.21 584.70 1527 469.63 433.72 1511 486.26 454.42 δNH(55), υPh(21)
1518 32.60 0.62 1514 22.38 5.20 1495 10.95 36.60 δCH(13), υPh(73)
1517 2.45 63.72 1501 1.84 75.57 1489 80.05 35.16 1475 υPz(71), υPh(13)
1447 14.35 30.11 1443 12.54 27.47 1424 16.95 28.88 1416 1417 δCHPz(45), υPz(48)
1417 80.12 71.15 1412 88.38 75.97 1397 84.56 60.95 1401 1387 υPh(69), δCH(15)
1391 35.11 35.00 1386 33.92 30.07 1371 31.41 36.19 1365 1367 δCHPz(42), υPz(29)
1355 95.52 209.09 1335 104.45 272.06 1334 116.90 249.50 1328 1323 υPh(67), υPz(20)
1322 27.22 62.39 1324 7.57 19.30 1308 18.56 56.55 δCH(74), υPh(13)
1305 526.18 109.35 1295 523.19 103.64 1275 21.16 3.51 1273 1276 δNH(74), υCC(19)
1286 0.44 4.62 1284 2.46 2.19 1269 497.50 154.21 δCHPz(92)
1272 7.57 30.07 1262 11.21 184.84 1249 0.80 16.32 υCF(59), υPh(14), υCN(16)
1268 11.28 389.48 1260 23.98 304.22 1246 22.47 461.77 1244 1244 υPh(24), υCN(56)
1215 13.31 38.47 1193 55.64 125.64 1189 27.89 55.20 1188 δCH(56), υCN(19)
1193 31.81 75.48 1191 1.53 1.44 1183 35.70 93.73 1186 1188 υPz(81)
1156 37.68 1.68 1152 27.31 1.96 1137 15.34 1.89 1153 υCF(19), δCH(54)
1154 25.58 1.10 1149 17.45 1.27 1128 0.78 0.59 1116 1124 υPz(27), υCF(20), δCHPz(17)
1125 18.87 1.42 1117 20.44 1.58 1097 18.30 2.19 1071 1103 υCN(66), υPz(13)
1102 132.11 3.24 1094 137.67 1.54 1028 187.57 1.78 1052 υCF(51), υPz(13)
1099 236.50 17.15 1085 237.97 19.41 1016 81.88 22.59 1020 1022 υCF(71), δCF(22)
1042 109.88 19.09 1036 110.29 19.53 1010 14.21 45.89 υPh(18), δCH(53), υCF(11), υPz(10)
1037 22.18 37.24 1030 17.57 36.44 1002 40.11 4.34 1000 υPz(58), δCH(11)
1007 61.10 6.47 1006 71.24 7.58 989 132.04 18.51 982 δPh(58), υPh(18)
996 43.56 29.35 994 47.14 31.55 977 71.93 15.99 δPz(27), υPz(50)
995 1.44 1.37 991 1.13 1.31 976 148.32 31.28 γCH(84)
983 0.08 0.39 978 0.10 0.47 967 0.49 0.30 γCHPz(90)
964 1.37 2.27 961 1.26 2.56 966 16.10 2.73 γCHPz(79), τPz(15)
961 0.94 1.71 956 0.39 1.79 959 1.54 0.41 951 951 γCH(82), τPh(11)
881 18.72 8.17 877 18.73 8.50 870 112.55 1.70 870 873 δNH(26), δCO(58)
872 82.50 1.86 867 73.54 1.96 862 9.69 2.29 γCH(40), γCHPz(25)
863 59.17 0.36 860 37.38 0.27 859 16.58 11.30 γCH(32), γCHPz(50)
847 68.98 4.34 843 71.76 5.11 845 95.69 0.21 839 846 γCH(67), γNH(10)
837 6.81 29.87 832 6.32 29.14 816 8.44 31.23 γCH(69), υPh(17), δPh(18)
799 28.12 0.77 792 40.58 1.07 795 35.85 0.47 795 γNH(53), γCH(26)
774 3.13 4.01 770 1.49 4.07 771 9.81 2.13 771 765 τPz(11), δCHPz(10), υPh(77)
745 8.08 18.28 743 9.25 18.71 739 4.17 1.32 733 δPz(46), δPh(21)
737 0.93 1.59 734 1.18 1.47 727 12.33 24.25 730 τPh(64), γCN(13), γCC(11)
698 0.51 2.35 694 0.26 2.40 696 0.01 0.64 689 γCO(49), δPz(36)
696 13.31 5.62 692 11.50 6.50 674 6.42 10.93 675 664 δCC(19), δCF(45)
639 0.19 7.72 641 0.22 7.90 625 0.44 10.04 631 629 δPh(82)
614 2.53 9.22 617 2.23 9.33 603 2.68 10.30 592 δPz(84)
601 2.43 0.61 599 2.78 0.68 585 2.25 1.26 576 δPh(22), δCF(52), δCO(11)
577 22.35 0.88 576 18.45 0.91 569 26.68 0.93 554 568 γCN(26),τPh(22), γCC(14), δCF(11)
534 0.14 1.27 531 0.10 1.39 515 0.15 1.55 515 529 δCF(51), δCC(11)
492 0.11 1.32 490 0.09 1.39 483 0.34 1.02 495 496 τPh(31), γCN(17), δCF(17)
481 2.14 8.90 479 2.17 8.71 472 1.98 8.88 479 476 δCO(38), δCC(29)
454 21.52 0.22 455 21.59 0.23 454 19.07 0.25 456 τPz(51), γCC(17)
413 0.51 0.10 415 0.30 0.10 409 0.55 0.06 418 419 τPz(81)
399 7.41 1.00 399 6.90 0.84 394 7.01 0.63 δCF(47), τPz(35)
389 12.20 2.58 389 12.61 2.66 379 11.17 2.55 δCF(59), δCN(16)
374 0.35 1.24 375 0.32 1.41 370 0.98 1.86 τPz(37), γCC(11), δCF(27)
366 17.79 2.10 365 17.42 1.99 355 20.19 2.86 351 δCF(54), δCN(12), τPz(15)
328 1.62 1.53 329 1.63 1.51 321 1.60 1.29 δCO(35), δCN(19), δNH(14)
60 13.87 0.60 261 13.95 0.61 255 12.15 0.76 255 δCC(36), δCN(18), δCF(17)
250 1.97 0.40 249 1.82 0.43 247 1.99 0.45 τPh(36), δCF(17),γCC(12)
187 0.46 1.49 188 0.50 1.53 182 0.43 1.73 177 δCC(20), δCF(14),δPh(14), δNH(14)
166 0.54 8.67 167 0.65 8.78 165 0.39 7.55 τPz(37), γCC(27)
145 3.12 0.93 146 2.99 0.95 143 3.44 0.94 δCC(31), δCO(18),δCF(11)
123 1.29 0.36 122 1.25 0.38 121 1.44 0.31 τCO(30), γCC(16),τNH(25)
76 2.37 0.55 76 2.44 0.55 76 2.89 0.43 87 τCO(40), τPh(12),γCC(11), τNH(10)
63 0.08 0.22 64 0.10 0.23 62 0.12 0.20 δNH(26), δCO(25),δCN(15), δCC(21)
51 0.40 0.57 51 0.39 0.63 51 0.47 0.11 τNH(57), τCO(20), τCF(12)
30 0.15 1.50 30 0.18 1.53 30 0.14 1.27 τNH(55), τPh(18)
15 0.93 3.95 17 0.89 3.99 13 1.04 3.66 τCF(63), τPh(13)
abbreviations: υ –stretching; δ – in-plane deformation; γ – out-of-plane deformation; τ –torsion; Ph-phenyl ring; Pz – Pyrazine ring; potential energy distribution is given in brackets with the assignments in the last column; IRI – IR intensity; RA – Raman activity.

The CF3 group possesses as many normal vibrations as methyl and as, halogen atoms are much heavier than hydrogen, CF3 modes are expected at considerably lower values. According to Roeges (1994) the absorption regions of –CF3 in substituted benzene are: υCF 1300–1100; δCF 720–440; ρCF 470–260 and τCF below 100 cm−1. In the present case the bands at 1052 and 1020 cm−1 in the IR spectrum, 1022 cm−1 in the Raman spectrum and 1249, 1028, and 1016 cm−1 (SDD) are assigned as the stretching modes of CF3 group. The deformation bands are assigned at 675, 576, and 515 cm−1 in the IR spectrum, 664, 529, and 351 cm−1 in the Raman spectrum and 674, 585, 515, 394, 379, and 355 cm−1 theoretically. Iriarte et al. (2008) reported the CF3 stretching values at 1244, 1213, 1140, and 1128, (DFT), 1218, 1176, 1136, and 1127 (experimental), deformation bands of CF3 at 729 cm−1 (experimental), 743 cm−1 (DFT). The CF modes are reported at 1317, 1270, 1143, 643, 632, and 574 cm−1 in the IR spectrum, 1275, 1150, 640, 582, 416, 379, 336, 331, and 234 cm−1 in the Raman spectrum and 1270, 1135, 643, 590, 588, 410, 372, 368, 326, 228, and 118 cm−1 theoretically (Anbarasu and Arivazhagan, 2011; Krishnakumar and Xavier, 2005).

The carbonyl group vibrations give rise to characteristic bands in the vibrational spectra. The intensity of these bands can increase owing to conjugation or formation of hydrogen bonds. In the present case the stretching mode of C⚌O is assigned at 1616 cm−1 in the Raman spectrum, 1617 cm−1 in the IR spectrum and the SDD calculations give this mode at 1610 cm−1. The deformation bands are observed at 870 cm−1 in the IR spectrum, 873 and 689 cm−1 in the Raman spectrum and 870 and 696 cm−1 theoretically. In the case of 2-aminopyrazine-3-carboxylic acid (Pawlukojc et al., 2000) the vibrational modes of carboxylic group were assigned as follows: torsional C⚌O 101 cm−1 (Raman), 111 (HF); rocking C⚌O 537 (Raman), 571 (HF); wagging C⚌O 723 (Raman), 736 (HF); bending C⚌O 912 (Raman), 886 (HF); stretching mode C⚌O 1718 (Raman), 1786 cm−1 (HF).

The N–H stretching vibrations generally give rise to bands (Spire et al., 2000; Bellamy, 1975) at 3500–3300 cm−1. In the present study, the NH stretching band is observed at 3343 cm−1 in the IR spectrum and at 3349 cm−1 in the Raman spectrum. The SDD calculations give this mode at 3388 cm−1. In mono-substituted amides, the in-plane bending frequency and the resonance stiffened CN band stretching frequency fall close together and therefore interact. The CNH vibration where the nitrogen and the hydrogen move in opposite directions relative to the carbon atom involves both NH bend and CN stretching and absorbs (Colthup et al., 1975) near 1500 cm−1. The CNH vibration where N and H atoms move in the same direction relative to the carbon atom gives rise to a weaker band (Colthup et al., 1975) near 1250 cm−1. In the present case the bands observed at 1276 cm−1 in the Raman spectrum, 1273 cm−1 in the IR spectrum and 1511 and 1275 cm−1 (SDD) are assigned as CNH bending modes. The out-of-plane NH wag absorbs at 795 cm−1 in the Raman spectrum and 795 cm−1 theoretically. Mary et al. (2008) reported the NH bands at 1537, 1250, and 650 cm−1 in the IR spectrum and 1580, 1227, and 652 cm−1 theoretically, for a similar derivative.

Primary aromatic amines with nitrogen directly on the ring absorb at 1330–1260 cm−1 because of the stretching of the phenyl C–N bond (Colthup et al., 1975). For the title compound, the υC10–N19 mode is observed at 1244 experimentally and 1246 cm−1 theoretically. The C9–N19 stretching band is reported in the range 1050–1150 cm−1 (Kundoo et al., 2003). In the present case υC9–N19 is observed at 1071 cm−1 in the IR spectrum, 1103 cm−1 in the Raman spectrum and 1097 cm−1 theoretically.

The CH stretching modes of pyrazinecarboxamide are reported in the range 3050–3088 cm−1 and the in-plane CH deformations at 1316, 1188, and 1062 cm−1 and out-of-plane CH deformations at 970, 922, and 792 cm−1 (Kalkar et al., 1989; Schettino et al., 1972). The ring stretching modes of pyrazinecarboxamide are reported at 1589, 1530, 1480, 1443, 1175, and 1035 cm−1 (Kalkar et al., 1989), 1581, 1525, 1479, 1437, 1166, and 1025 cm−1 (Akyuz, 2003). For the title compound the pyrazine CH stretching modes are assigned at 3133, 3122, and 3099 cm−1 theoretically and experimentally bands are observed at 3127 and 3090 cm−1 in the IR spectrum and 3139 and 3080 cm−1 in the Raman spectrum. The CH stretching band of pyrazine was reported in the range 3100–3000 cm−1 (Schettino et al., 1972; Sbrana et al., 1973; Arenas et al., 1985). These modes are observed at 3057, 3070, and 3086 cm−1 (IR), 3060, 3070, and 3087 (Raman) and 3061, 3074, and 3079 cm−1 (theoretical) for 2-chloropyrazine and 3099 and 3104 (IR), 3078 and 3103 (Raman), 3096 and 3100 cm−1 (calculated) for 2,6-dichloropyrazine (Ngo et al., 2007). For 2-aminopyrazine-3-carboxylic acid (Pawlukojc et al., 2000) the pyrazine ring stretching modes are observed at 1564, 1536, 1468, 1458, 1360, 1258, and 1083 cm−1 (Raman) and 1567, 1565, 1457, 1415, 1373, 1231, and 1068 cm−1 (ab initio). For pyrazine, the ring stretching modes are reported at 1411, 1485, 1128, and 1065 (IR), 1579, 1522, and 1015 (Raman) and 1597, 1543, 1482, 1412, 1140, 1063, and 1009 cm−1 theoretically (Endredi et al., 2003). For 2-chloropyrazine the pyrazine ring stretching modes are reported at 1561, 1515, 1198, 1176, 1133, and 1047 (IR), 1562, 1518, 1197, 1177, 1135, and 1049 (Raman) and 1563, 1533, 1210, 1167, 1128, and 1042 (DFT) and for 2,6-dichloropyrazine these values are reported at 1550, 1540, 1412, 1230, 1189, 1151, and 1136 (IR), 1551, 1407, 1172, 1152, and 1138 (Raman) and 1547, 1541, 1407, 1229, 1170, 1167, and 1143 cm−1 (DFT) (Endredi et al., 2003). For the title compound, the pyrazine ring stretching modes are observed at 1533, 1475, and 1416 cm−1 in the IR spectrum, 1528, 1417, 1188, and 1000 cm−1 in the Raman spectrum and 1526, 1489, 1424, 1183, 1002, and 977 cm−1 theoretically. The ring breathing mode of pyrazine is reported at 1015 cm−1 (Endredi et al., 2003). In the Raman spectrum of 2,6-dichloropyrazine and 2-chloropyrazine the ring breathing mode is reported at 1131 and 1049 cm−1 (Endredi et al., 2003). The band at 1002 cm−1 (DFT) is assigned as the ring breathing mode of the pyrazine ring, for the title compound. The δCH bending modes in 2,6-dichloropyrazine are reported at 1189 and 1151 cm−1 in the IR spectrum (Endredi et al., 2003). For 2-chloropyrazine these modes are reported at 1455, 1380, 1280, and 1179 cm−1 in the IR spectrum and 1457, 1380, 1289, and 1167 cm−1 theoretically (Endredi et al., 2003). For pyrazine, the δCH modes are observed at 1485, 1413, and 1065 cm−1 in the IR spectrum and 1482, 1412, 1227, and 1063 cm−1 theoretically (Endredi et al., 2003). In the present case the CH in-plane bending modes are observed at 1417 and 1367 cm−1 in the Raman spectrum, 1416 and 1365 in the IR spectrum and 1424, 1371, and 1269 cm−1 theoretically (SDD). For the title compound the out-of-plane CH modes are assigned at 967, 966, and 859 cm−1 theoretically. For pyrazine γCH modes are observed at 790 cm−1 in the IR spectrum, 976 and 925 cm−1 in the Raman spectrum and 985, 972, 930, and 790 cm−1 by DFT calculations (Endredi et al., 2003). For 2-chloropyrazine the γCH modes are reported at 954, 929, and 844 cm−1 in the IR spectrum, 960, 928, and 847 cm−1 in the Raman spectrum and 960, 923, and 837 cm−1 theoretically (Endredi et al., 2003). In the case of 2,6-dichloropyrazine these bands are reported at 897 and 875 cm−1 in the IR spectrum, 896 cm−1 in the Raman spectrum and 919 and 869 cm−1 theoretically (Endredi et al., 2003). For 2-aminopyrazine-3-carboxylic acid γCH modes are reported at 1006 and 850 cm−1 in the Raman spectrum and 987 and 852 cm−1 by HF calculations (Pawlukojc et al., 2000). The substituent sensitive modes and other deformation bands of the phenyl and pyrazine ring are also identified and assigned (Table 2). Most of the modes are not pure, but contain significant contributions from other modes.

In order to investigate the performance of vibrational wavenumbers of the title compound, the root mean square (RMS) value between the calculated and observed wavenumbers were calculated. The RMS values of wavenumbers were calculated using the following expression (Joseph et al., 2012). RMS = 1 n - 1 i n ( υ i calc - υ i exp ) 2 .

The RMS error of the observed IR and Raman bands are found to be 30.70 and 25.28 for B3PW91/6-31G, 24.46 and 19.35 for B3LYP/6-31G and 14.93 and 11.46 for B3LYP/SDD methods, respectively. The small differences between experimental and calculated vibrational modes are observed. This is due to the fact that experimental results belong to solid phase and theoretical calculations belong to gaseous phase.

4.2

4.2 Geometrical parameters

To the best of our knowledge, no X-ray crystallographic data of the title compound has yet been established. However, the theoretical results obtained are almost comparable with the reported structural parameters of similar molecules. For pyrazine ring (Bormans et al., 1997) the CN bond lengths are 1.339 and 1.331 Ǻ. For 2-chloropyrazine (Endredi et al., 2003) CN bond lengths are in the range 1.335–1.312 Ǻ. For the title compound the pyrazine bond lengths C1–C2, C1–N6, C5–N6, C5–C4, C4–N3 and C2–N3 are 1.4121, 1.3517, 1.3608, 1.4104, 1.3577, and 1.3577 Ǻ, respectively. For a similar derivative Mary et al. (2008) reported the corresponding values as 1.3917, 1.2996, 1.3229, 1.3840, 1.322 and 1.3116 Ǻ. For 2-aminopyrazine-3-carboxylic acid (Pawlukojc et al., 2000; Ptasiewicz-Bak and Leciejewicz, 1997), the bond lengths C5–C9, C9–O20, C5–N6 and C5–C4 are1.479, 1.212, 1.333, and 1.4079 (XRD) and 1.492, 1.186, 1.315, and 1.4092 Ǻ (ab initio calculations), respectively. In the present case, the corresponding values are 1.5079, 1.2580, 1.3608 and 1.4104 Ǻ. Endredi et al. (2004) reported the bond lengths C2–C1, C1–N6, C5–N6, C5–C4 and C4–N3 as 1.391, 1.331, 1.331, 1.331, and 1.331 for pyrazine, 1.4, 1.327, 1.333, 1.387, and 1.334 for 2-methylpyrazine, 1.41, 1.331, 1.33, 1.385, and 1.331 Ǻ for 2,3-dimthylpyrazine, 1.396, 1.327, 1.335, 1.396, and 1.335 for 2,5-dimethylpyrazine, and 1.399, 1.326, 1.332, 1.399, and 1.332 Ǻ for 2,6-dimethylpyrazine. For pyrazinamide, Chis et al. (2005) reported bond lengths, C5–N6, C4–C5, C4–N3, C2–N3, C1–C2, C1–N6, C5–C9, C9–N19, C9–O20, C4–H8, and N19–H21 as 1.341, 1.4, 1.337, 1.338, 1.397, 1.336, 1.510, 1.357, 1.226, 1.086, and 1.010 Ǻ and these results are in agreement with the present study.

The CN bond length in the pyrazine ring of the title compound C1–N6 = 1.3517, C5–N6 = 1.3608, C2–N3 = 1.3577 and C4–N3 = 1.3577 Ǻ is much shorter than the normal C–N single bond that is referred to 1.49 Ǻ. The same results are shown for the bond lengths of the two C–C bonds, C2–C1 = 1.4121 and C5–C4 = 1.4104 Ǻ in the pyrazine ring and are also smaller than that of the normal C–C single bond of 1.54 (He et al., 2004). The C–N bond lengths C9–N19 = 1.3782 and C10–N19 = 1.4120 are also shorter than the normal C–N single bond of 1.49, which confirms this bond to have some character of a double or conjugated bond (Bakalova et al., 2004). The CF bond lengths are reported as 1.4068, 1.3284, 1.3251, and 1.3284 Ǻ theoretically (Anbarasu and Arivazhagan, 2011; Krishnakumar and Xavier, 2005) and for the title compound the CF lengths are in the range 1.4025–1.4163 Ǻ. For the title compound, the CCF angles lie in the range 112.9–113.1° and the FCF angles in the range 105.1–106.7°, which are in agreement with the reported literature (Iriarte et al., 2008).

For benzamide derivatives, Noveron et al. (2003)) reported the bond lengths C10–N19, C9–O20, C9–N19, C9–C5 and N19–H21 as 1.3953, 1.2253, 1.3703, 1.4943, and 0.733, whereas the corresponding values for the title compound are 1.4120, 1.2580, 1.3789, 1.5079, and 1.0209Ǻ. The C⚌O and CN bond lengths (Takeuchi et al., 1999) in benzamide, acetamide and formamide are respectively, 1.2253, 1.2203, and 1.2123 and 1.3801, 1.3804, and 1.3683 Ǻ. According to the literature (Stevens, 1978; Gao et al., 1991) the changes in bond lengths in C⚌O and CN are consistent with the following interpretation: that is, hydrogen bond decreases the double bond character of C⚌O bond and increases the double bond character of C–N bond.

At N19 position, the angles C9–N19–H21 is 113.5°, C10–N19–H21 is 118.9° and C9–N19–C10 is 128.6°. This asymmetry of angles at N19 position indicates the weakening of N19–H21 bond resulting in proton transfer to the oxygen atom O20 (Barthes et al., 1996). At C10 position the angel C11–C10–N19 is increased by 3.0° and C15–C10–N19 is reduced by 2.7° from 120° and this asymmetry reveals the interaction between the amide moiety and the phenyl ring. The CC bond lengths in the phenyl ring lie between 1.3980–1.4171 Ǻ and the CH bond lengths lie between 1.0819–1.0878 Ǻ. The CC bond length of benzene (Tamagawa et al., 1976) is 1.3993 Ǻ and benzaldehyde (Borizenko et al., 1996) is 1.3973 Ǻ.

4.3

4.3 First hyperpolarizability

Non-linear optical (NLO) is at the forefront of current research because of its importance in providing key functions of optical modulation, optical switching, optical logic and optical memory for the emerging technologies in areas such as telecommunications, optical interconnections and signal processing (Andraud et al., 1994). The first hyperpolarizability (β0) of this novel molecular system is calculated using the B3LYP/6-31G method, based on the finite field approach. In the presence of an applied electric field, the energy of a system is a function of the electric field. First hyperpolarizability is a third rank tensor that can be described by a 3 × 3 × 3 matrix. The 27 components of the 3D matrix can be reduced to 10 components due to the Kleinman symmetry (Kleinman, 1962). The components of β are defined as the coefficients in the Taylor series expansion of the energy in the external electric field. When the electric field is weak and homogeneous, this expansion becomes. E = E 0 - i μ i F i - 1 2 ij α ij F i F j - 1 6 ijk β ijk F i F j F k - 1 24 ijkl γ ijkl F i F j F k F l + . . .

where E 0 is the energy of the unperturbed molecule, F i is the field at the origin, μ i , α ij , β ijk and γ ijkl are the components of dipole moment, polarizability, the first hyper polarizabilities, and second hyperpolarizibilites, respectively. The calculated first hyperpolarizability of the title compound is 9.77 × 10−30 esu (Table 3). The C–N distances in the calculated molecular structure vary from 1.3517–1.4120 Ǻ which are intermediate between those of a C–N single bond (1.48 Ǻ) and a C⚌N double bond (1.28 Ǻ). Therefore, the calculated data suggest an extended π-electron delocalization over the pyrazine ring and carboxamide moiety (Tain et al., 2002) which is responsible for the nonlinearity of the molecule. We conclude that the title compound is an attractive object for future studies of non linear optical properties.

Table 3 β components and first order hyperpolarizability β0 (in esu) value.
Parameters B3LYP/SDD
βxxx −972.546
βxxy 297.886,
βxyy −123.413
βyyy −49.475
βxxz −3.2761
βxyz 0.1546
βyyz 5.698,
βxzz −8.079
βyzz −2.415
βzzz −6.807
β0 esu) 9.77 × 10−30

4.4

4.4 NBO analysis

The natural bond orbitals’ (NBO) calculations were performed using a NBO 3.1 program (Glendening et al., 2003) as implemented in the Gaussian 09 package at the DFT/B3LYP level in order to understand various second-order interactions between the filled orbitals of one subsystem and vacant orbitals of another subsystem, which is a measure of the intermolecular delocalization or hyper conjugation. NBO analysis provides the most accurate possible ‘natural Lewis structure’ picture of ‘j’ because all orbital details are mathematically chosen to include the highest possible percentage of the electron density. A useful aspect of the NBO method is that it gives information about interactions of both filled and virtual orbital spaces that could enhance the analysis of intra and inter molecular interactions. The second-order Fock- matrix was carried out to evaluate the donor–acceptor interactions in the NBO basis. The interactions result in a loss of occupancy from the localized NBO of the idealized Lewis structure into an empty non-Lewis orbital. For each donor (i) and acceptor (j) the stabilization energy (E2) associated with the delocalization i → j is determined as. E ( 2 ) = Δ E ij = q i ( F i , j ) 2 ( E j - E i )

qi → donor orbital occupancy

Ei, Ej → diagonal elements

Fij → the off diagonal NBO Fock matrix element

In NBO analysis a large E(2) value shows the intensive interaction between electron-donors and electron-acceptors and greater the extent of conjugation of the whole system, the possible intensive interactions are given in Table 4. The second-order perturbation theory analysis of Fock matrix in NBO basis shows strong intra-molecular hyper-conjugative interactions of π electrons. The intra-molecular hyper-conjugative interactions are formed by the orbital overlap between n (N) and σ (C–O) bond orbitals which results in ICT causing stabilization of the system. The strong intra-molecular hyper-conjugative interaction of C9–O20 from n1(N19) → π (C9–O20) which increases ED (0.32646e) that weakens the respective bonds leading to stabilization of 68.38 kcal mol−1. Also the strong intra-molecular hyper-conjugative interaction of C9–N19 from n2(O20) → σ (C9–N19) which increases ED (0.06813 e) that weakens the respective bonds leading to stabilization of 22.08 kcal mol−1. These interactions are observed as an increase in electron density (ED) in C–O and C–N anti-bonding orbitals that weaken the respective bonds. The increased electron density at the nitrogen and carbon atoms leads to the elongation of their respective bond lengths and a lowering of the corresponding stretching wavenumbers. The electron density (ED) is transferred from the n(O), and n(N) to the anti-bonding π orbital of the C–N, and C–O explaining both the elongation and the red shift (Choo et al., 2002). The –NH, and –C⚌O stretching modes can be used as a good probe for evaluating the bonding configuration around the corresponding atoms and the electronic distribution of the molecule. Hence the above structure is stabilized by these orbital interactions.

Table 4 Second-order perturbation theory analysis of Fock matrix in NBO basis corresponding to the intra-molecular bonds of the title compound.
Donor (i) Type ED/e Acceptor (j) Type ED/e E(2)a E(j) − E(i)b F(i,j)c
C1–C2 π 1.5771 C5–N6 π 0.42322 21.20 0.25 0.065
C1–N6 σ 1.98466 C5–C9 σ 0.07003 3.78 1.23 0.062
N3–C4 σ 1.98576 C5–C9 σ 0.07003 3.12 1.21 0.056
N3–C4 π 1.67080 C1–C2 π 0.27670 23.95 0.31 0.078
C5–N6 σ 1.98487 C1–H27 σ 0.01941 2.34 1.35 0.050
C5–N6 π 1.69921 C1–C2 π 0.27670 22.11 0.32 0.076
C5–C9 σ 1.96787 C10–N19 σ 0.03248 6.21 1.06 0.072
C9–N19 σ 1.98584 C4–C5 σ 0.04081 2.28 1.32 0.049
C9–O20 π 1.96970 C5–N6 π 0.42322 6.64 0.33 0.047
C10–C11 σ 1.97240 C10–C15 σ 0.02211 2.77 1.20 0.052
C10–C11 π 1.61214 C12–C13 π 0.38340 25.43 0.28 0.075
C10–C15 σ 1.97303 C9–N19 σ 0.06813 5.12 1.11 0.068
C10–N19 σ 1.98386 C14–C15 σ 0.01136 2.16 1.35 0.048
C11–C12 σ 1.97466 C10–N19 σ 0.03248 5.09 1.08 0.066
C14–C15 σ 1.97550 C10–N19 σ 0.03248 4.22 1.08 0.061
C14–C15 π 1.68520 C10–C11 π 0.37411 22.32 0.28 0.072
N19–H21 σ 1.98036 C9–O20 σ 0.01071 5.18 1.14 0.069
C23–F24 σ 1.98903 C13–C14 σ 0.02392 2.55 1.46 0.055
C23–F25 σ 1.98875 C12–C13 σ 0.02410 2.28 1.46 0.052
C23–F26 σ 1.98776 C23–F25 σ 0.10336 1.70 1.03 0.038
LP(1)N3 σ 1.93202 C4–C5 σ 0.04081 10.58 0.84 0.085
LP(1)N6 n 1.92067 C4–C5 σ 0.04081 10.13 0.86 0.084
LP(1)N19 σ 1.62822 C9–O20 π 0.32646 68.38 0.25 0.118
LP(1)O20 σ 1.97577 C5–C9 σ 0.07003 2.68 1.06 0.048
LP(2)O20 n 1.88796 C9–N19 σ 0.06813 22.08 0.67 0.110
LP(1)F24 n 1.98878 C13–C23 σ 0.05094 0.87 1.44 0.032
LP(2)F24 n 1.96168 C23–F26 σ 0.12111 5.00 0.53 0.047
LP(3)F24 n 1.94301 C23–F25 σ 0.10336 10.45 0.53 0.067
LP(1)F25 σ 1.98884 C13–C23 σ 0.05094 0.83 1.44 0.031
LP(2)F25 n 1.96256 C23–F26 σ 0.12111 4.90 0.53 0.046
LP(3)F25 n 1.94342 C23–F24 σ 0.09748 10.26 0.53 0.067
LP(1)F26 σ 1.98905 C13–C23 σ 0.05094 0.75 1.44 0.030
LP(2)F26 n 1.96233 C13–C23 σ 0.05094 4.60 0.77 0.053
LP(3)F26 n 1.94329 C23–F24 σ 0.09748 9.19 0.53 0.063
E(2) means energy of hyper-conjugative interactions (stabilization energy in kcal mol−1).
Energy difference between donor and acceptor i and j NBO orbitals in a.u.
F(i,j) is the Fock matrix elements between i and j NBO orbitals in a.u.

The NBO analysis also describes the bonding in terms of the natural hybrid orbital n2(F24), which occupies a higher energy orbital (−0.43933 a.u.) with considerable p-character (99.98%) and low occupation number (1.96168 a.u.) and the other n1(F24) occupies a lower energy orbital (−1.10579) with p-character (21.75%) and high occupation number (1.98878 a.u.). Also n2(F25), which occupies a higher energy orbital (−0.43722 a.u.) with considerable p-character (100.0%) and low occupation number (1.96256 a.u.) and the other n1(F25) occupies a lower energy orbital (−1.10735 a.u.) with p-character (21.43%) and high occupation number (1.98884 a.u.). Again n2(F26), which occupies a higher energy orbital (−0.43166 a.u.) with considerable p-character (99.74%) and low occupation number (1.96233 a.u.) and the other n1(F26) occupies a lower energy orbital (−1.10750 a.u.) with p-character (20.75%) and high occupation number (1.98905 a.u.). Also n2(O20), which occupies a higher energy orbital (−0.28291 a.u.) with considerable p-character (100.0%) and low occupation number (1.88796a.u.) and the other n1(O20) occupies a lower energy orbital (−0.71130 a.u.) with p-character (35.25%) and high occupation number (1.97577 a.u.).Thus, a very close to pure p-type lone pair orbital participates in the electron donation to the σ(C–N), π(C–O), and σ(C–F) orbital interactions in the compound. The results are tabulated in Table 5.

Table 5 NBO results showing the formation of Lewis and non-Lewis orbitals.
Bond(A–B) ED/energya EDA% EDB% NBO s% p%
πC1–C2 1.5771 50.43 49.57 0.7001(sp1.00)C 0.00 100.0
−0.31174 +0.7041(sp1.00)C 0.00 100.0
σC1–N6 1.98466 38.84 61.16 0.6232(sp2.23)C 30.91 69.09
−0.87749 +0.7820(sp1.81)N 35.55 64.45
σN3–C4 1.98576 60.01 39.99 0.7747(sp1.91)N 34.35 65.65
−0.86150 +0.6324(sp1.00)C 30.87 69.13
σC5–N6 1.67080 58.22 41.78 0.7630(sp1.00)C 0.00 100.0
−0.33949 +0.6463(sp1.00)N 0.00 100.0
σC5–N6 1.98487 39.87 60.13 0.6314(sp2.28)C 30.53 69.47
−0.87834 +0.7041(sp1.86)N 35.00 65.00
πC5–N6 1.69921 43.06 56.94 0.6562(sp1.00)C 0.00 100.0
−0.35345 +0.7546(sp1.00)N 0.00 100.0
σC5–C9 1.96787 51.71 48.29 0.7191(sp2.08)C 32.47 67.53
−0.69377 +0.6949(sp1.87)C 34.50 65.50
σC9–N19 1.98584 36.72 63.28 0.6060(sp2.10)C 32.23 67.77
−0.85230 +0.7955(sp1.87)N 34.80 65.20
πC9–O20 1.96970 30.99 69.01 0.5567(sp1.00)C 0.00 100.0
−0.38789 +0.8307(sp1.00)O 0.00 100.0
σC10–C11 1.97240 51.21 48.79 0.7156(sp1.71)C 36.91 63.09
−0.71438 +0.6985(sp1.95)C 33.85 66.15
σC10–N19 1.98386 37.37 62.63 0.6113(sp2.61)C 27.70 72.30
−0.82652 +0.7914(sp1.66)N 37.54 62.46
σC23–F24 1.98903 26.65 73.35 0.5162(sp3.81)C 20.77 79.23
−0.95881 +0.8565(sp3.60)F 21.72 78.28
σC23–F25 1.98875 26.64 73.36 0.5162(sp3.86)C 20.59 79.41
−0.95118 +0.8565(sp3.67)F 21.41 78.59
σC23–F26 1.98776 26.82 73.18 0.5179(sp3.89)C 20.44 79.56
−0.93132 +0.8555(sp3.88)F 20.49 79.51
n1N3 1.93202 sp2.17 31.52 68.48
−0.37525
n1N19 1.92067 sp2.39 29.48 70.52
−0.39470
n1N19 1.62822 sp1.00 0.00 100.0
−0.29177
n1O20 1.97577 sp0.54 64.75 35.25
−0.71130
n2O20 1.88796 sp1.00 0.00 100.0
−0.28291
n1F24 1.98878 sp0.28 78.25 21.75
−1.10579
n2F24 1.96168 sp99.99 0.02 99.98
−0.43933
n3F24 1.94301 sp99.99 0.01 99.99
−0.43826
n1F25 1.98884 sp0.27 78.57 21.43
−1.10735
n2F25 1.96256 sp1.00 0.00 100.0
−0.43722
n3F25 1.94342 sp99.99 0.02 99.98
−0.43681
n1F26 1.98905 sp0.26 79.25 20.75
−1.10750
n2F26 1.96233 sp99.99 0.26 99.74
−0.43166
n3F26 1.94329 sp1.00 0.00 100.0
−0.42938
ED/energy is expressed in a.u.

4.5

4.5 Frontier molecular orbitals

To explain several types of reactions and for predicting the most reactive position in conjugated systems, molecular orbitals and their properties such as energy are used (Choudhary et al., 2013). The highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) are the most important orbitals in a molecule. The eigen values of HOMO and LUMO and their energy gap reflect the biological activity of the molecule. A molecule having a small frontier orbitals gap is more polarizable and is generally associated with a high chemical reactivity and low kinetic stability (Sinha et al., 2011; Lewis et al., 1994; Kosar and Albayrak, 2011). HOMO, which can be thought as the outer orbital containing electrons, tends to give these electrons as an electron donor and hence the ionization potential is directly related to the energy of the HOMO. On the other hand LUMO can accept electrons and the LUMO energy is directly related to electron affinity (Gece, 2008; Fukui, 1982). Two important molecular orbitals (MO) were examined for the title compound, the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) which are given in Figs. 4 and 5. In the title compound, the HOMO of π nature is delocalized over the whole C–C bonds of the phenyl ring, and CONH and CF3 groups. By contrast, LUMO is located over the pyrazine ring and CONH group. Accordingly, the HOMO–LUMO transition implies an electron density transfer from the phenyl ring to the pyrazine ring through the CONH group. For understanding various aspects of pharmacological sciences including drug design and the possible ecotoxicological characteristics of the drug molecules, several new chemical reactivity descriptors have been proposed. Conceptual DFT based descriptors have helped in many ways to understand the structure of molecules and their reactivity by calculating the chemical potential, global hardness and electrophilicity. Using HOMO and LUMO orbital energies, the ionization energy and electron affinity can be expressed as: I = −EHOMO, A = −ELUMO, η = (−EHOMO + ELUMO)/2 and μ = 1/2(EHOMO + ELUMO) (Koopmans, 1993). Parr et al. (1999) proposed the global electrophilicity power of a ligand as ω = μ2/2η. This index measures the stabilization in energy when the system acquires an additional electronic charge from the environment. Eelctrophilicity encompasses both the ability of an electrophile to acquire additional electronic charge and the resistance of the system to exchange electronic charge with the environment. It contains information about both electron transfer (chemical potential) and stability (hardness) and is a better descriptor of global chemical reactivity. The hardness η and chemical potential μ are given by the following relations η = (I − A)/2 and μ = −(I + A)/2, where I and A are the first ionization potential and electron affinity of the chemical species (Parr and Pearson, 1983).

HOMO plot of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide.
Figure 4
HOMO plot of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide.
LUMO plot of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide.
Figure 5
LUMO plot of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide.

For the title compound, EHOMO = −6.984, ELUMO = −2.925, Energy gap = HOMO–LUMO = 4.059, Ionization potential I = 6.984, Electron affinity A = 2.925, global hardness η = 2.0295, chemical potential μ = −4.9545, and global electrophiliciy = μ2/2η = 6.05. It is seen that the chemical potential of the title compound is negative and it means that the compound is stable. They do not decompose spontaneously into the elements they are made up of. The hardness signifies the resistance toward the deformation of electron cloud of chemical systems under small perturbation encountered during the chemical process. The principle of hardness works in Chemistry and Physics but it is not physically observable. Soft systems are large and highly polarizable, while hard systems are relatively small and much less polarizable.

4.6

4.6 1H and 13C NMR analysis

The calculated values of 13C and 1H NMR chemical shifts by the B3LYP/GIAO model level of theory in the DMSO environment for the title compound are summarized in Table 6. Density functional theory shielding calculations are rapid and applicable to large systems, but the paramagnetic contribution to the shielding tends to be over estimated. In this sense, theoretical calculations of the chemical shifts may be used as an aid for the assignment of the experimental data. The observed NMR data are given as supporting information. The absolute isotropitc chemical shielding of N-[(4-(trifluoromethyl)phenyl]pyrazine-2-carboxamide was calculated by the B3LYP/GIAO model (Wolinski et al., 1990). Relative chemical shifts were then estimated by using the corresponding TMS shielding: σcalc (TMS) is calculated in advance at the same theoretical level. Numerical values of chemical shift σpred = σcalc(TMS) − σcalc together with calculated values of σcalc(TMS), are given in Table 6. It could be seen from Table 6 that the chemical shift was in agreement with the experimental data. Thus, the results showed that the predicted chemical shifts were in good agreement with the experimental data for the title compound.

Table 6 The observed (in DMSO) and predicted 1H and 13C NMR isotropic chemical shifts (with respect to TMS, all values in ppm).
Atom Experimental Calculated
H7 8.63 8.65
H8 9.52 9.58
H16 7.70 7.76
H17 7.93 8.00
H18 7.85 7.90
H21 9.80 9.83
H22 7.62 7.68
H27 8.84 8.86
C1 126.4 127.0
C2 126.3 126.9
C4 124.0 124.7
C5 119.5 120.0
C10 147.9 148.2
C11 144.8 145.0
C12 143.8 144.2
C13 142.4 142.9
C14 140.2 141.0
C15 141.3 142.0
C23 160.9 161.2

5

5 Conclusion

The FT-IR and FT-Raman spectra of N-(4-trifluoromethylphenyl)pyrazine-2-carboxamide were studied. The molecular geometry and wavenumbers were calculated using DFT methods and the optimized geometrical parameters (SDD) are in agreement with that of similar derivatives. The small differences between experimental and calculated vibrational modes are observed. This is due to the fact that experimental results belong to solid phase and theoretical calculations belong to gaseous phase. The calculated first hyperpolarizability is high and the title compound is an attractive object for future studies of non linear optics.

Acknowledgment

Hema Tresa Varghese would like to thank the University Grants Commission, India for a research Grant.

References

  1. , . J. Mol. Struct.. 2003;651–653:541-545.
  2. , , , , . J. Mol. Struct.. 2007;834–836:399-402.
  3. , , , , , , , . Spectrochim. Acta. 2007;69:782-788.
  4. , , . Indian J. Pure Appl. Phys.. 2011;49:227-233.
  5. , , , , , , , . J. Am. Chem. Soc.
  6. , , , , , . J. Chem. Soc. Faraday Trans.. 1985;2(81):405-415.
  7. , , , , , , , . J. Mol. Struct. Theochem. 2004;710:229-234.
  8. , , , . Synth. Met.. 1996;76:337-340.
  9. , . The Infrared Spectrum of Complex Molecules (third ed.). London: Chapman and Hall; .
  10. , , , . J. Med. Chem.. 1996;39:3394-3400.
  11. , , , . J. Mol. Struct. Theochem. 1998;423:225-234.
  12. , , , . J. Phys. Chem.. 1996;100:7426-7434.
  13. , , , . J. Mol. Struct.. 1997;42:121-128.
  14. , , , , , , , . Chem. Phys.. 2005;316:153-163.
  15. , , , , . J. Mol. Struct. (Theochem). 2002;587:1-8.
  16. , , , , . Comp. Theor. Chem.. 2013;1016:8-21.
  17. , . , ed. Encyclopedia of Analytical Chemistry: Interpretation of Infrared Spectra, A Practical Approach. Chichester: John Wiley and Sons; .
  18. , , , . Introduction to Infrared and Raman Spectroscopy. New York: Academic Press; .
  19. , , , , , . J. Med. Chem.. 1992;35:1212-1215.
  20. , , , . Gaussview, Version 5. Shawnee Mission KS: Semichem Inc.; .
  21. , . Chem. Listy. 2006;100:959-966.
  22. , , , , . Molecules. 2002;7:363-373.
  23. , , , , , , . Molecules. 2006;11:242-256.
  24. , , , , , . Molecules. 2007;12:2589-2598.
  25. , , , , . Chem. Pap.. 2000;54:245-248.
  26. , , , , , . Molecules. 2009;14:4180-4189.
  27. , , , . J. Mol. Struct. Theochem. 2003;633:73-82.
  28. , , , . J. Mol. Struct. Theochem. 2004;677:211-225.
  29. Foresman, J.B., Frisch, E. (Ed.), 1996. Exploring Chemistry with Electronic Structure Methods: A Guide to using Gaussian. Gaussian Inc., Pittsburg, PA.
  30. , , , , , , , , , , , . Bioorg. Med. Chem. Lett.. 2012;22:2784-2788.
  31. , . Science. 1982;218:747-754.
  32. , , , , . Acta Cryst.. 1991;47B:742-745.
  33. , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , , . Gaussian 09, Revision B.01. Wallingford CT: Gaussian, Inc.; .
  34. , . Corros. Sci.. 2008;50:2981-2992.
  35. , , , , . NBO Version 3.1. Pittsburgh, PA: Gaussian Inc.; .
  36. , , . J. Chem. Phys.. 1985;82:270-283.
  37. , , , , . J. Mol. Struct. Theochem. 2004;668:201-208.
  38. , , . Drugs. 1996;48:689-708.
  39. , , , , , , . Vib. Spectrosc.. 2008;46:107-114.
  40. , , , , , , , . Global J. Anal. Chem.. 2012;3:1-12.
  41. , , , . Spectrochim. Acta. 1989;45A:635-641.
  42. , . Phys. Rev.. 1962;126:1977-1979.
  43. , . Physica. 1993;1:104-113.
  44. , , . Spectrochim. Acta. 2011;78A:160-167.
  45. , , , , . Chem. Pap.. 2008;62:358-363.
  46. , , , , . Chem. Pap.. 2002;56:214-217.
  47. , , . Spectrochim. Acta. 2005;61:253-260.
  48. , , , , . Mater. Lett.. 2003;57:2193-2197.
  49. , , , . Xenobiotica. 1994;24:401-408.
  50. , , . GAR2PED, A Program to obtain a Potential Energy Distribution from a Gaussian Archive Record. Belgium: Unviersity of Antwerp; .
  51. , , , , . Spectrochim. Acta. 2008;71:725-730.
  52. , , , , , , . Spectrochim. Acta. 2008;71:566-571.
  53. , , , , , , . Molecules. 2000;5:208-218.
  54. , . Natur. Med.. 1996;2:635-636.
  55. , , , , , , . Antimicrob. Agents Chemother.. 2007;51:2430-2435.
  56. , , , . Chem. Mater.. 2003;15:372-374.
  57. , , , , , . Farmaco. 2002;57:135-144.
  58. , , . J. Am. Chem. Soc.. 1983;105:7512-7516.
  59. , , , . J. Am. Chem. Soc.. 1999;121:1922-1924.
  60. , , , , . J. Mol. Struct.. 2000;516:7-14.
  61. , , . Pol. J. Chem.. 1997;71:1350-1358.
  62. , , , , . Lancet. 1997;350:624-629.
  63. , . A Guide to the Complete Interpretation of the Infrared Spectra of Organic Structures. New York: Wiley; .
  64. , , , . J. Chem. Phys.. 1973;59:2441-2450.
  65. , , , . Chem. Phys. Lett.. 1972;13:284-285.
  66. , , , , . J. Mol. Simul.. 2011;37:153-163.
  67. , , , , . J. Antimicrob. Chemother.. 2004;53:192-196.
  68. , , , . Antimicrob. Agents Chemother.. 1995;39:1269-1271.
  69. , , , , . Physics. 2000;D137:392-396.
  70. , . Acta Cryst.. 1978;34B:544-551.
  71. , , , , , , . Polyhedron. 2002;21:1217-1222.
  72. , , , , , , . J. Mol. Struct.. 1999;485–486:175-181.
  73. , , , . J. Mol. Struct.. 1976;30:243-253.
  74. , . Assignments of Vibrational Spectra of Seven Hundred Benzene Derivatives. New York: Wiley; .
  75. , , , , , , , , , , , . Bioorg. Med. Chem. Lett.. 2011;21:2911-2915.
  76. , , , . J. Am. Chem. Soc.. 1990;112:8251-8260.
  77. , , , , , . J. Antimicrob. Chemother.. 2003;52:790-795.
  78. , , , . J. Mol. Struct. Theochem. 2004;671:179-187.

Appendix A

Supplementary data

Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.arabjc.2013.08.004.

Appendix A

Supplementary data

Supplementary Fig. 1
Supplementary Fig. 2

Show Sections