SciELO - Scientific Electronic Library Online

 
vol.63 issue1CRYSTAL AND MOLECULAR STRUCTURE OF 14α-ACETOXY-13α-HYDROXYMULIN-11-EN-20-OIC ACID MONOHYDRATEPRELIMINARY RESULTS OF THE REACTION OF CYCLOTRIMERIZATION OF PHENYLACETYLENE [2+2+2] CATALYZED BY [(Cp*)Co(Indene)] COMPLEX author indexsubject indexarticles search
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • On index processCited by Google
  • Have no similar articlesSimilars in SciELO
  • On index processSimilars in Google

Share


Journal of the Chilean Chemical Society

On-line version ISSN 0717-9707

J. Chil. Chem. Soc. vol.63 no.1 Concepción Mar. 2018

http://dx.doi.org/10.4067/s0717-97072018000103887 

Article

A THEORETICAL QUANTUM STUDY OF THE ELECTRONIC PROPERTIES OF MENTOXY DICHLORO PHOSPHOROUS (C10H19OPCl2)

Amir Lashgaria 

Shahriar Ghamamia 

M. Govindarajanb 

Guillermo Salgado-Moránc 

Paola  Montes Romeroc 

Lorena  Gerli Candiad* 

aDepartment of Chemistry, Faculty of Science, Imam Khomeini International University, Qazvin, 34148-96818

bDepartment of Physics, Avvaiyar Government College for Women (AGCW), Karaikal, Puducherry 609602, India

cDepartamento de Ciencias Químicas, Facultad de Ciencias Exactas, Universidad Andrés Bello, Sede Concepción, Concepción, Chile

dDepartamento de Química Ambiental, Facultad de Ciencias, Universidad Católica de la Santísima Concepción, Concepción, Chile

ABSTRACT

A theoretical quantum study of the organophosphorus compound with formula C10H19OPCl2 (MEPCL2) was carried out. The results of the calculations show excellent agreement between experimental and computed frequencies evaluated at the B3LYP/6-311++G(d,p) level of theory. A study of the electronic properties, such as excitation energies and wavelengths were performed employing the time-dependent DFT (TD-DFT) method. Global a chemical reactivity of MEPCL2 was analyzed through global reactivity descriptors, while its local reactivity was analyzed by mean maps of the electrostatic potential. Also, the orbital energies values suggest that a charge transfer is occurring within the molecule.

Keywords: Phosphorous compound; DFT; HF; HOMO-LUMO

1. INTRODUCTION

Organophosphorus compounds are of great interest because many biological processes such as energy transfer, bone synthesis, amino acid synthesis, among others are related to them1. Consequently, it is not so strange that many researchers are studying them to understand their role in biological systems24. Moreover, phosphorus compounds are employed in numerous synthetic procedures to manufacture detergents, fertilizers, pesticides, toxic industrial phosphate esters, natural products, among others5,6.

Recently, mentoxy dichloro phosphorous (C10H19OPCl2) (MEPCL2) was synthesized from the reaction of 2-(2-propyl)-5-methyl-1-cyclohexanol (menthol) with PCl3. This new organophosphorus compound exhibits a maximum absorbance at 227 nm, this value indicates that MEPCL2 has important absorption properties in the visible region of the electromagnetic spectra, by which may useful in the fabrication of photovoltaic devices. However, to the best of our knowledge, there is no information about the electronic properties and the chemical behavior of MEPCL2. Thus, it is necessary to get a deeper knowledge about this compound, to identify its potential uses and applications. In this sense, the evaluation of these properties may result expensive and complicated from the experimental point of view. However, the accepted theories of the quantum chemistry provide advantages to analyze the electronic properties and reactivity of molecules reliably7. Therefore, in the present work, we perform a computational and theoretical quantum study about the electronic properties and chemical reactivity exhibited by MEPCL2. We consider that this kind of study will contribute to getting a better understanding of the chemical behavior of this new organophosphorus compound.

2. METHODOLOGY

The entire quantum chemical calculations have been performed at HF and DFT (B3LYP) methods with 6-311++G(d,p) basis sets using the Gaussian 03W program8. The optimized structural parameters have been evaluated for the calculations of vibrational frequencies by assuming Cs point group symmetry. At the optimized geometry for the title molecule no imaginary frequency modes were obtained, therefore there is a true minimum on the potential energy surface was found. As a result, the unscaled calculated frequencies, reduced masses, force constants, infrared intensities, and depolarization ratios are obtained. In order to fit the theoretical wavenumbers to the experimental, the scaling factors have been introduced by using a least square optimization of the computed to the experimental data. Vibrational frequencies are scaled as 0.9067 for HF and 0.961 for B3LYP9 to account for systematic errors caused by basis set incompleteness, neglect of electron correlation and vibrational anharmonicity. The assignments of the compound are calculated by using VEDA program10.

The electronic absorption spectra for optimized molecule calculated with the time-dependent DFT (TD-DFT) at B3LYP/6-311++G(d,p) level in gas phase and solvent (Acetonitrile, chloroform, and water). The 13C nuclear magnetic resonance (NMR) chemical shifts of the molecule were calculated by the gauge-independent atomic orbital (GIAO) method in CDCl3 and compared with experimental results.

3. RESULTS AND DISCUSSION

3.1. Molecular geometry

The geometry optimization of MEPCL2, see Fig. 1, was carried out at the HF/6-311++G(d,p) and B3LYP//6-311++G(d,p) levels of theory.

Fig. 1 Optimized molecular structure of MEPCL2 with atom numbering. 

No imaginary frequencies were obtained, which ensures us that each gradient optimization located corresponds to true minimum energy on the energy potential surface. The selected optimized structural parameters, bond lengths and bond angles of the title compound are reported in Table 1. From this table, it is possible to observe that the bond lengths and bond angles values calculated at the HF/6-311++G(d,p) and B3LYP/6-311++G(d,p) levels of theory compare favorably well with those reported in the literature. A linear fit of the optimized data, see Fig. 2 suggests that the bond lengths values calculated at the HF/6-311++G(d,p) level is slightly better than the geometry obtained at the B3LYP/6-311++G(d,p) level. However, this difference is in the order of the error associated with the resolution of the wavefunction when numerical algorithms are employed. Thus, no significant differences were obtained, which is indicative that the electronic correlation has a little effect on the molecular geometry of the MEPCL2.

Table 1 Selected optimized parameters of MEPCL2 

Bond length HF/6-311++ G(d,p) (Å) B3LYP/6-311++ G(d,p) (Å) Experimental Value (Å)7 Bond angle HF/6-311++ G(d,p) (°) B3LYP/ 6-311++ G(d,p) (°) Experimental Value (°)7
C1-H12 1.083 1.093 H12-C1-O11 107.0 106.4
C2-H13 1.085 1.095 0.970 H13-C2-C3 105.6 105.8
C3-H14 1.086 1.094 0.970 H14-C3-H15 105.8 105.7 108.1
C3-H15 1.088 1.096 0.970 H15-C3-C4 108.9 109.0 109.5
C4-H16 1.086 1.094 0.970 H16-C4-C5 109.2 109.3 109.5
C4-H17 1.087 1.095 0.970 H17-C4-H16 106.2 106.1 108.1
C5-H18 1.088 1.096 0.970 H18-C5-C7 106.6 106.9 107.9
C6-H19 1.082 1.090 0.970 H19-C6-C1 109.2 109.2 109.3
C6-H20 1.085 1.094 0.970 H20-C6-H19 106.8 106.9 107.9
C7-H21 1.086 1.094 H21-C7-H23 107.4 107.4
C7-H22 1.087 1.094 H22-C7-H21 107.7 107.8
C7-H23 1.086 1.093 H23-C7-H22 107.4 107.3
C8-H24 1.088 1.098 H24-C8-C10 106.0 106.2
C9-H25 1.087 1.094 H25-C9-H27 107.1 107.1
C9-H26 1.084 1.093 H26-C9-H25 107.5 107.3
C9-H27 1.083 1.091 H27-C9-H26 108.5 108.4
C10-H28 1.086 1.094 H28-C10-H30 107.7 107.7
C10-H29 1.084 1.092 H29-C10-H28 108.0 108.0
C10-H30 1.087 1.094 H30-C10-H29 107.6 107.7
C1-C6 1.526 1.528 1.534 C4-C3-C2 114.6 114.5 110.5
C4-C3 1.531 1.535 1.521 C2-C1-C6 115.7 115.9 110.8
C2-C1 1.537 1.540 1.532 C7-C5-C6 112.6 112.5
C7-C5 1.537 1.538 C9-C8-C2 118.4 118.1
C9-C8 1.534 1.539 C10-C8-C9 110.1 110.0
C5-C4 1.535 1.543 1.523 C6-C5-C4 109.1 109.2 111.7
C10-C8 1.539 1.540 C8-C2-C3 107.7 107.6
C6-C5 1.536 1.546 1.525 O11-C1-C6 113.9 113.8
C3-C2 1.546 1.552 1.513 P31-O11-C1 132.3 130.1
C8-C2 1.447 1.567 Cl32-P31-O11 102.4 102.8
O11-C1 1.563 1.576 Cl33-P31-Cl32 98.4 98.5
P31-O11 1.568 1.600
Cl32-P31 2.093 2.141
Cl33-P31 2.093 2.141

Fig. 2 Linear fit bond length and bond angle. 

3.2. Vibrational analysis

The number of potential active fundamental vibrations of a non-linear molecule which contains N atoms is equal to (3N-6), apart from three translational and three rotational degrees of freedom. Therefore, MEPCL2 molecule has 33 atoms with 93 normal modes of vibrations. In the present work, we have done a detailed vibrational assignment of the experimental wavenumbers reported in a previous work, through the comparison with theoretically scaled wavenumbers evaluated at the HF/6-311++G(d,p) and B3LYP/6-311++G(d,p) levels of theory. In order to fit the theoretical wavenumbers to the experimental data, the scaling factors have been introduced by using a least square optimization of the computed to the experimental data. Vibrational frequencies are scaled as 0.9067 for HF and 0.961 for B3LYP10 to take into account the systematic errors caused by basis set incompleteness, neglect of electron correlation, in the case of HF calculations, and vibrational anharmonicity. The assignments of the compound are calculated by using VEDA program10. The calculated experimental and scaled frequencies using HF and DFT (B3LYP) with the 6-311++G(d,p) basis set are listed in Table 2.

Table 2 Detailed vibrational assignments of observed and computed wavenumbers of MEPCL2. 

Modes species Expa HF/6-311++G(d,p) B3LYP/6-311++G(d,p) Mode description b
Infrared7 unscaled scaled IIR unscaled scaled
1. A′ 3266 2962 31.7 3111 2989 γCH cyclohexane
2. A′ 3255 2951 21.4 3109 2988 γCH cyclohexane
3. A′ 3243 2941 24.1 3104 2983 γCH cyclohexane
4. A′ 3237 2935 37.4 3097 2976 γCH cyclohexane
5. A′ 3227 2926 44.0 3086 2966 γCH cyclohexane
6. A′ 3224 2923 52.6 3085 2965 γCH cyclohexane
7. A′ 2957 vs 3219 2918 16.5 3081 2961 γCH cyclohexane
8. A′ 3216 2916 64.0 3064 2944 γCH cyclohexane
9. A′ 3208 2908 9.3 3056 2937 γCH cyclohexane
10. A′ 3205 2906 15.4 3053 2934 γCH cyclohexane
11. A′ 2928 vs 3198 2899 33.4 3052 2933 γCH3 sym
12. A′ 3189 2892 11.1 3034 2916 γCH3 sym
13. A′ 3175 2879 52.9 3031 2913 γCH3 sym
14. A′ 3173 2877 35.7 3030 2912 γCH3 sym
15. A′ 3163 2867 33.0 3026 2908 γCH3 sym
16. A′ 3162 2867 15.4 3024 2906 γCH3 sym
17. A′ 3159 2864 6.7 3017 2899 γCH3 sym
18. A′ 3155 2860 16.6 3012 2894 γCH3 sym
19. A′ 3145 2851 22.0 2994 2877 γCH3 sym
20. A′ 1456 m 1644 1490 9.0 1519 1459 γC-C ring
21. A′ 1637 1484 10.5 1515 1456 γC-C ring
22. A′ 1630 1478 3.9 1512 1453 γC-C ring
23. A′ 1627 1475 19.4 1506 1447 βCH
24. A′ 1621 1470 3.3 1502 1444 γC-C
25. A′ 1620 1469 2.2 1500 1441 γC-C ring
26. A′ 1617 1466 9.9 1494 1436 γC-C
27. A′ 1610 1459 1.4 1488 1430 γC-C ring
28. A′ 1607 1457 0.5 1485 1427 γC-C ring
29. A′ 1388 m 1560 1415 8.6 1431 1375 βCH
30. A′ 1370 m 1555 1410 13.8 1422 1366 γC-O
31. A′ 1546 1402 1.9 1418 1363 γC-C
32. A′ 1538 1394 1.6 1409 1354 γC-C
33. A′ 1348 w 1532 1389 8.5 1401 1347 βCH
34. A′ 1521 1379 1.0 1396 1341 βCH
35. A′ 1330 w 1512 1371 1.9 1384 1330 βCH
36. A′ 1500 1360 2.2 1375 1322 βCH
37. A′ 1493 1354 0.5 1368 1315 βCH
38. A′ 1480 1342 2.3 1361 1308 βCH
39. A′ 1478 1341 3.4 1358 1305 βCH
40. A′ 1457 1321 1.6 1330 1278 βCH
41. A′ 1447 1312 7.3 1325 1274 βCH
42. A′ 1432 1298 4.3 1303 1253 βCH
43. A′ 1224 m 1381 1252 4.2 1272 1222 βCH m
44. A′ 1358 1231 0.9 1258 1209 βCH m
45. A′ 1180 m 1305 1183 3.0 1205 1158 βCH m
46. A′ 1283 1164 0.5 1188 1142 βCH
47. A′ 1277 1158 1.9 1173 1128 βCH m
48. A′ 1255 1138 3.7 1155 1110 βCH m
49. A′ 1060 m 1193 1082 4.0 1109 1066 βCH m
50. A′ 1188 1077 0.8 1100 1057 βCH m
51. A′ 1174 1065 2.4 1089 1046 βCH m
52. A′ 1140 1034 4.7 1054 1013 βCH m
53. A′ 992 s 1128 1023 5.3 1042 1001 γP-O
54. A′ 974 s 1120 1015 38.5 990 951 βCCC ring
55. A′ 1067 968 3.2 988 950 βCCC
56. A′ 934 m 1066 967 455.3 973 935 βCCC ring
57. A′ 1043 946 30.6 967 929 βCCC
58. A′ 1032 935 13.2 959 921 βCCC ring
59. A′ 1021 926 2.6 948 911 βCCC
60. A′ 997 904 0.7 929 893 βCCO
61. A′ 879 w 979 887 2.8 905 870 βPOC
62. A′ 937 850 3.2 872 838 βPOCl
63. A′ 825 w 929 842 18.6 862 829 βClPCl
64. A″ 770 vw 857 777 15.6 792 761 φCH
65. A″ 829 752 4.3 774 744 φCH
66. A″ 794 720 1.6 745 716 φCH
67. A″ 723 656 3.1 670 644 φCH
68. A″ 586 vw 646 585 2.8 604 581 φCH
69. A″ 495 w 582 527 1.6 543 522 φCH
70. A″ 447 vw 535 485 64.7 482 463 γΡ-Cl
71. A″ 499 452 18.9 440 423 γΡ-Cl
72. A″ 469 425 108.2 428 411 φCH m
73. A″ 457 414 40.4 414 398 φCH m
74. A″ 454 411 15.4 411 395 φCH m
75. A″ 424 384 3.5 390 374 φCH m
76. A″ 416 377 3.3 381 366 φCH m
77. A″ 387 351 3.7 349 335 βClPCl
78. A″ 368 334 2.8 338 325 φCH m
79. A″ 347 314 1.9 319 306 φCH m
80. A″ 318 289 0.2 288 277 φCH m
81. A″ 308 280 0.3 280 270 φCCC ring
82. A″ 247 224 0.4 231 222 φCCC ring
83. A″ 234 212 0.0 221 212 φCCC
84. A″ 233 211 0.2 214 206 βClPCl
85. A″ 222 201 0.4 196 188 φPOCC
86. A″ 204 185 0.5 185 178 φClPOC
87. A″ 180 163 1.7 163 157 φCCC
88. A″ 142 129 1.6 129 124 φCCC
89. A″ 90 82 1.1 86 83 φClOlP
90. A″ 79 72 0.5 70 67 φCCC
91 A″ 60 55 0.2 50 48 φCCC
92 A″ 57 52 0.2 46 44 φCCC
93 A″ 34 31 0.2 26 25 φPCCl

as: strong; vs: very strong; m: medium; w: weak; vw: very weak.

bγ: stretching; β: in-plane bending; φ: out-of-plane bending; IIR: IR intensity.

3.2.1 Cyclohexane ring vibrations

Cyclohexyl ring in MEPCL2 contains three methylene (CH2) groups, each group has six modes of vibration namely asymmetric and symmetric stretching, scissoring, rocking, wagging and twisting modes. In general, in cyclohexane, the CH2 stretching vibrations are usually observed below 3000 cm-1 11. The asymmetric CH2 stretching vibration generally observed in the region is 3000-2900 cm−1 while the CH2 symmetric stretch is between 2900 and 2800 cm-1 12,13. In MEPCL2, the calculated wavenumbers at 2934, 2933, 2916, 2913, 2912, 2908 and 2906 cm−1 are attributed to asymmetric CH2 stretching vibrations, while the symmetric CH2 stretching wave numbers are calculated as 2899, 2894 and 2877 cm−1 at the B3LYP/6-311++G(d,p) level of theory. One stretching vibration of cyclohexyl ring is observed in MEPCL2, as very strong band at 2928 cm−1 in FT-IR spectrum. The vibrations due to aromatic C-H in-plane bending are observed in the region 1000-1300 cm-1 14. For this compound, the C-H in-plane bending vibrations were observed at 1224, 1180 and 1060 cm−1 in FT-IR. The theoretically scaled vibrations predicted at the y B3LYP/6-311++G(d,p) level are obtained at 1305, 1274, 1253, 1222, 1209, 1158, 1142 1110 and 1066 cm−1. The C-H out-of-plane bending vibrations are appearing within the region 900-675 cm-1 15. The vibrations identified at 770, 586, 4985 and 447 cm−1 in FT-IR are assigned to C-H out-of-plane bending for MEPCL2.

3.2.2. C-C vibrations

The ring stretching vibrations are useful to identify characteristic of the ring itself. For the title compound, the C=C stretching vibrations are recorded at 1456 and 1388 cm−1 in FT-IR with medium intensities. All bands are appearing in the expected range, except first band. Most of the bands are observed with medium and strong intensities. The computed values are at 1459, 1456, 1453, 1441, 1430 and 1427 cm−1 at the B3LYP/6-311++G(d,p) level of theory. The other C-C vibrations are computed at 1447, 1444, 1436, 1375 and 1366 cm−1. These vibrations are downshifted when they are compared with those exhibited by aromatic compounds and other C-C vibrations are more shifted with ring C-C vibrations. Only two bands at 992 and 974 cm−1 are assigned to C-C-C in-plane bending vibrations of MEPCL2. The two bands are in infrared region with very strong intensities. The computed vibrations are tabulated to C-C-C in-plane and out-of-plane bending vibrations at 1001, 951, 921 and 270, 222 and 266 cm−1.

3.2.3. Methyl group vibrations

For the assignment of CH3 group frequencies, nine fundamental vibrations can be associated with each CH3 group. Three stretching, three bending, two rocking modes and single torsional mode describe the motion of the methyl group. In the experimental FT-IR band is observed at 2928 cm−1 for MEPCL2 have been assigned to CH3 symmetric stretching vibration. The CH3 stretching vibrations are calculated as 2893, 2916, 2913, 2912, 2908, 2906, 2899, 2894 and 2877 cm−1 at the B3LYP/6-311G++(d,p) level of theory. The FT-IR band observed at 1388 cm−1 have been assigned to CH3 in-plane bending vibration for MEPCL2.

3.2.4. P-Cl and P-O vibrations

The experimental P-Cl stretching vibrations are observed in the interval 587-435 cm−1. Also, a band at 447 cm−1 is assigned as P-Cl vibration. The calculated P-Cl in-plane and out-of-plane bending vibrations are observed at 387, 233 and 34 cm−1. The P-O phenyl linkage gives rise to two bands. A strong band at 1260-1160 cm−1 is mainly due to the stretching of the C-O bond of the phenyl group. Also, the band at 992 cm−1 is related to a C-O stretching vibration.

The simulated infrared spectra of MEPCL2 obtained at the HF/ 6-311G++(d,p) and B3LYP/6-311G++(d,p) levels of theory are shown in Fig. 3.

Fig. 3 Computed FT-IR spectra for MEPCL2 

3.3. Frontier molecular orbitals (FMOs)

The highest occupied molecular orbital (HOMO) and the lowest-lying unoccupied molecular orbital (LUMO) may describe the electronic transition, non-linear optic properties, and UV-Vis spectra of a molecular system16. Also, the energy gap between HOMO and LUMO determines the kinetic stability, chemical reactivity and, optical polarizability and chemical hardness- softness of a molecule17,18. The hard molecules are less polarizable than the soft ones because they need big energy to excitation. The electronic calculated through the TD-DFT method. In order to evaluate the energetic behavior of MEPCL2 in the solvent, we carried out calculations considering acetonitrile, water, chloroform and gas phases. The energies of the four molecular orbitals of MEPCL2: the second highest and highest occupied MO's (HOMO and HOMO-1), the lowest and the second lowest unoccupied MO's (LUMO and LUMO+1) were calculated at the TD-DFT/B3LYP/6-311++G(d,p) level of theory and they are reported in Table 3. Also, the 3D plots of the HOMO- 1, HOMO, LUMO and LUMO+1 orbitals computed at TD-DFT/B3LYP/6-311++G(d,p) level of theory for MEPCL2 molecule are depicted in Fig. 4. It is clear from this figure that, while the HOMO is localized on almost the whole molecule, LUMO is especially localized on the ring. Also, note that both, the HOMOs and the LUMOs are mostly anti-bonding type orbitals.

Table 3 Computed energy values of MEPCL2 in acetonitrile, water, chloroform and gas. 

TD-DFT/B3LYP/6-311++G(d,p) Acetonitrile Water Chloroform Gas
Etotal (Hartree) -1729.77 -1729.77 -1729.76 -1729.76
EHOMO (eV) -7.90 -7.91 -7.89 -7.85
ELUMO (eV) -1.60 -1.61 -1.58 -1.54
ΔEHOMO-LUMO gap (eV) 6.30 6.30 6.31 6.32
EHOMO-1 (eV) -8.05 -8.05 -8.08 -8.19
ELUMO+1 (eV) -1.18 -1.18 -1.58 -1.14
ΔEHOMO-1-LUMO+1 gap (eV) 6.87 6.86 6.50 7.05
EHOMO-2 (eV) -8.18 -8.18 -8.21 -8.34
ELUMO+2 (eV) -0.24 -0.24 -0.28 -0.43
ΔEHOMO-2-LUMO+2 gap (eV) 7.94 7.94 7.93 7.92

Fig. 4 3D view of HOMO and LUMO diagram with energy gaps. 

The calculated energy values of the HOMO and LUMO energy gaps are 6.3017, 6.3008, 6.3085 and 6.3183 eV in acetonitrile, water, chloroform and gas phases, respectively. Thus, it is clear that the highest energy gap is obtained when chloroform solvent is employed which suggest that MEPCL2 is more chemically stable in such solvent. In view of calculated absorption spectra, the maximum absorption wavelength corresponds to the electronic transition from the HOMO to LUMO with 92% and from the HOMO to LUMO+1 with 9% contribution, see Table 4. The other wavelength, excitation energies, oscillator strength and calculated counterparts with major contributions are listed in Table 4.

Table 4 Theoretical electronic absorption spectra of MEPCL2 excitation energies E (eV), (absorption wavelength λ (nm), and oscillator strengths (f) using TD-DFT/B3LYP/6-311++G(d,p) method. 

Solvent Energy (eV) Wavelength Oscillator strength Major contribution
Acetonitrile 5.4418 227.8 0.0794 HOMO->LUMO (96%)
5.6102 220.9 0.0005 H-2->LUMO (10%), H-1->LUMO (80%)
5.8609 211.5 0.0014 H-2->LUMO (84%), H-1->LUMO (11%)
5.9338 208.9 0.0332 H-5->LUMO (10%), H-4->LUMO (15%), HOMO->L+1 (69%)
6.1502 201.6 0.0558 H-1->L+1 (87%)
Water 5.4427 227.8 0.0789 HOMO->LUMO (96%)
5.6072 221.1 0.0005 H-2->LUMO (10%), H-1->LUMO (80%)
5.8559 211.7 0.0014 H-2->LUMO (84%), H-1->LUMO (11%)
5.9353 208.9 0.0330 H-5->LUMO (13%), H-4->LUMO (11%), HOMO->L+1 (70%)
6.1475 201.7 0.0554 H-1->L+1 (87%)
Chloroform 5.4331 228.2 0.0839 HOMO->LUMO (97%)
5.6409 219.8 0.0005 H-1->LUMO (81%)
5.9147 209.6 0.0095 H-2->LUMO (67%), HOMO->L+1 (18%)
5.9202 209.4 0.0260 H-4->LUMO (19%), H-2->LUMO (18%), HOMO->L+1 (50%)
6.1769 200.7 0.0579 H-1->L+1 (84%)
Gas 5.4788 226.2 0.0579 HOMO->LUMO (96%)
5.7101 217.1 0.0612 H-1->LUMO (81%)
5.9263 209.2 0.0002 H-3->LUMO (24%), HOMO->L+1 (55%)
6.0786 203.9 0.0146 H-2->LUMO (89%)
6.2299 199.0 0.0003 H-1->L+1 (80%)

Note that the calculations of the molecular orbital show that the visible absorption maxima of MEPCL2 corresponds to the electron transition between frontier orbitals such as translation from HOMO to LUMO, see Table 4. The calculated absorption spectra showed five bands at 227.8, 220.9, 211.5, 208.9 and 201.6 nm for acetonitrile and at 227.2 nm in the experimental UV spectrum with maximum absorbance in the same solvent. In chloroform, water and gas phases, the theoretical maximum absorption bands are predicted at 228.2, 227.8 and 226.2 nm, respectively. All the maximum absorption bands are coming from HOMO to LUMO transition with energy contribution about 96 to 97 %. The next maximum peaks are predicated on HOMO-1 to LUMO+1 in all UV spectra with 80 to 90% around 200 nm. In Fig. 5 are shown the theoretical UV spectra obtained at the TD-DFT/6-311++G(d,p) level of theory, in all cases, it is possible to observe that the maximum absorbance is in the range 226-228 nm.

Fig. 5 Computed UV-Vis graphs in different solvent for MEPCL2. 

3.4. Ή and 13C NMR spectra from quantum calculations

The theoretical values for 1H and 13C NMR of MEPCL2 are given in Table 5. The theoretical 1H and 13C NMR chemical shifts of MEPCL2 have been compared with the experimental data measured in water and CDCl3 solvents.

Table 5 Experimental and theoretical probable 1H and 13C NMR isotropic chemical shifts (with respect to TMS and in Water and CDCl3 solution) of MEPCL2 compound. 

Atom Experimental (ppm)7 Theoretical (B3LYP) (ppm)
Water CDCl3
H12 5.123 5.085
H19 2.288 2.277 2.238
H13 2.092 2.207 2.206
H15 2.047 2.100 2.056
H18 2.038 2.080 2.053
H16 1.988 1.998 1.970
H20 1.826 1.814 1.813
H24 1.788 1.786 1.754
H27 1.511 1.591 1.529
H23 1.466 1.480 1.459
H14 1.439 1.468 1.441
H26 1.390 1.214 1.250
H29 1.353 1.192 1.237
H17 1.331 1.189 1.174
H28 1.310 1.087 1.103
H25 1.077 1.082 1.082
H21 0.975 0.958 0.995
H30 0.947 0.958 0.950
H22 0.809 0.814 0.799
C1 129.88 91.018 107.714
C2 75.14 49.268 66.567
C8 63.21 41.677 59.452
C6 48.24 38.348 55.768
C5 48.26 34.568 52.099
C3 48.15 32.526 50.223
C4 48.01 28.808 46.545
C10 43.19 27.558 45.318
C9 40.47 22.992 40.696
C7 34.74 18.220 35.978

A comparison, between the experimental and computed 13C NMR spectra of MEPCL2, indicates an increase in the value of the chemical shifts of the carbon atoms C1 and C2 of cyclohexane, due to heavy substitutions, which is caused by the electronic charge distribution around of these carbon atoms.

The hydrogen peaks in the cyclohexane are observed experimentally from 2.228 to 1.331 ppm, while that the evaluated at the B3LYP/6-311+G (d,p) level of theory are in the range 2.238 – 1.174 ppm.

The methyl hydrogen atom H22 peak identified at 0.809 ppm is the lowest chemical shift among the entire hydrogen atoms. Probably, it is because of its electronic interaction with other atoms is lesser. The correlation graphs of the experimental and theoretical 1H, and 13C NMR chemical shift are presented as supplementary material in Fig 6. A good correlation between predicted and observed 13C and 1H NMR chemical shifts is found. Moreover, the slope and intercept of the least-square, correlation lines were used to scale the GIAO isotropic absolute shielding constants. From Fig. 6 it is clear that the solvent water shows more deviation that the observed in the CDCl3 solvent. The relation is usually linear and described by the following equations (1-4):

Fig. 6 Correlation Proton and Carbon NMR between experimental and computed data. 

Proton NMR

vcal.=0.1286+1.06991vexp(R2=0.96503)in solventwater (1)
vcal.=0.0923+1.03958vexp(R2=0.97219)insolventCDCl3 (2)

Carbon NMR:

vcal.=3.86793+0.73135vexp(R2=0.97407)insolvent water (3)
vcal.=14.31258+0.072025vexp(R2=0.97442)insolventCDCl3 (4)

These results indicate that the calculations performed at the B3LYP/6-311++G(d,p) predicted adequately the experimental behavior of the Ή and 13C NMR spectra of MEPCL2.

3.5. Global Chemical Reactivity of MEPCL2

The reactivity of a molecular system can be analyzed employing the global reactivity descriptors derived from the DFT theory, which are evaluated through the total energies of the neutral, anionic and cationic systems. Thus, the ionization potential is determined from the energy difference between the energy of the compound derived from electron-transfer (radical cation) and the respective neutral compound; IP = Ecation- En; while the electron affinity is evaluated as the energy difference between the neutral molecule and the anion molecule: EA = En - Eanion. According to the Koopmans’ theorem, in HF calculations, ionization potentials (IP) and electron affinities (EA) may be approximated to the HOMO and LUMO's energies respectively. On the other hand, the validity of the Koopmans’ theorem within the DFT approximation is controversial, nonetheless, it has been shown that the Khon-Sham orbitals produce DFT reactivity descriptors that correlate quite well with the reactivity descriptors obtained from Hartree-Fock calculations19. Thus, in the present work, we decided to employ the second approximation. Additionally, from IP and EA values it is possible to evaluate the electronegativity (χ), hardness (η), softness (ζ), and electrophilicity index (ψ) of the molecular system, through the equations (5-8)20:

Electronegativity(χ):μχ=IP+EA2 (5)
Chemical hardness(η)IPEA2 (6)
Softness(ζ)=12η (7)
Electrophili city index(ψ)=μ22η (8)

The values of electronegativity, chemical hardness, softness, electrophilicity index and dipolar moment are reported in Table 6, in the solvents acetonitrile, water, chloroform and gas phases.

Table 6 Computed electronegativity, chemical hardness-softness, and dipolar moment of MEPCL2 in acetonitrile, water, chloroform and gas phases. 

B3LYP/6-311++G(d,p) Acetonitrile Water Chloroform Gas
Electronegativity χ (eV) 4.75 4.76 4.74 4.70
Chemical hardness η (eV) 3.15 3.15 3.15 3.16
Softness ζ (eV)−1 0.16 0.16 0.16 0.16
Electrophilicity index ψ (eV) 3.59 3.59 3.55 3.49
Dipolar moment (Debye) 3.87 3.89 3.67 3.15

From these values, observe that the values of χ, η, ζ, and ψ are similar, this means that MEPCL2 is showing the same global chemical behavior in the different solvents, but it is clear that the dipolar moment has different values which depend on the solvent in where the molecule is immersed. The above mentioned is indicative of the presence of different intramolecular interactions between the solvent and MEPCL2.

3.6. Local reactivity of MEPCL2 from Molecular Electrostatic Potential

In the present study, the molecular electrostatic potential (MESP) of MEPCL2 is plotted in order to analyze its local reactivity, see Fig. 7. The MESP is a plot of the electrostatic potential mapped onto an electron density isosurface and is useful to identify electrophilic and nucleophilic regions around the molecule. In the majority of the MESP, while the maximum negative region, which preferred site for electrophilic attack indications as red color, the maximum positive region which preferred site for nucleophilic attack symptoms as blue color. The electrophilic attack is more around phosphorous and carbon atoms. It is extended around chlorine atoms. Similarly, nucleophilic attacks are shown in hydrogen atoms and more in methyl group substitutions. The different values of the electrostatic potential at the MESP surface are represented by different colors; red, blue and green represent the regions of most negative, most positive and zero electrostatic potential, respectively. The color code of these maps is in the range between −0.02335 e. (deepest red) to 0.02335 e (deepest blue) in the compound, where blue indicates the strongest attraction and red indicates the strongest repulsion2126.

Fig. 7 Molecular electrostatic potential (MESP) of MEPCL2 

CONCLUSIONS

In this work, the geometric parameters and vibrational frequencies of the mentoxy dichloro phosphorous (C10H19OPCl2) were evaluated at the HF/6-311++G(d,p) and B3LYP/6-311++G(d,p) levels of theory. In comparison to the experimental results, the computed vibrational frequencies obtained by the B3LYP method are better than those obtained by the Hartree-Fock method. Fundamental vibrations with their mode were fully discussed in order to give a better understanding of the electronic structure of MEPCL2. The HOMO-LUMO gaps and implications of the electronic transitions were examined. The kinetic stability, chemical reactivity, optical polarizability, and chemical hardness-softness were discussed by frontier molecular orbital gaps. With orbital analysis, it has been suggested that MEPCL2 is more stable in chloroform than in other solvents. The 1H, and 13C NMR recorded and isotropic chemical shifts were calculated and they compare favorably well with the experimental results. The evaluation of the global reactivity descriptors suggests a similar chemical behavior of MEPCL2 in the different solvents analyzed. The molecular electrostatic potential isosurface provides a visual performance of the chemically active sites and comparative reactivity of atoms. Thus the present investigation provides complete vibrational assignments, structural information, chemical shifts and electronic properties of the MEPCL2.

ACKNOWLEDGEMENTS

The authors gratefully acknowledge the financial support from the Research Council of Imam Khomeini International University. Also, the authors greatly appreciate Prof. M. Govindarajan for valuable comments and discussions. LHMH expresses his gratitude to the Mexican National Council for Science and Technology (CONACYT) for financing this work through the Research Project Grant 257823 and to the Universidad Autónoma del Estado de Hidalgo. The authors gratefully acknowledge too the financial support of research Direction UCSC.

REFERENCES

1. Sengupta, S. K., Pandey, O. P., Rao, G. P., Vishen, P.: Metal Based Drugs. 8, 293 (2002) [ Links ]

2. Suarez, P., Dullius, J.E. L., Einloft, S., De Souza, R., Dupont, R.: J. Inorg. Chim. Acta. 255, 207-209 (1997) [ Links ]

3. Boon, A., Levisky, J. L., Pflug, L., Wilkes, J. S.: J. Inorg. Chem. 51, 480483 (1986) [ Links ]

4. Yanes, E. G., Gratz, S. R., Baldwin, M. J., Robinson, S.E., Stalcup, A. M.: Anal. Chem. 73, 3838 (2001) [ Links ]

5. Racke, K.D.: Degradation of organophosphorus insecticides in environmental matrices. in: Chambers, J.E., Levi, P.E. (eds.), Organophosphates: Chemistry, Fate, and Effects. Academic Press, San Diego. 47-73 (1992). [ Links ]

6. Quin, L. D.: A Guide to Organophosphorus Chemistry; John Wiley & Sons, (2000). [ Links ]

7. Mendoza-Huizar, L.H., J. Mex. Chem. Soc, 5(4) 416 (2014) [ Links ]

8. Gaussian 03, Revision C.02, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A.; Gaussian, Inc., Wallingford CT, 2004. [ Links ]

9. Karabacak, M., Kurt, M.: Spectrochim. Acta A. 71, 876-883 (2008). [ Links ]

10. Jamróz, M.H.: Vibrational Energy Distribution Analysis, VEDA 4, Warsaw, (2004). [ Links ]

11. Kemp, W.: Organic Spectroscopy, Mac Millan Press Ltd., London, (1991). [ Links ]

12. Shishkov, I.F., Sadova, N.I., Novikov, V.P., Vilkov, L.V., Strukt. Khim, Zh.: 25 (1984) 98-102. [ Links ]

13. Sinha, L., Karabacak, M., Narayan, V., Cinar, M., Prasad, O.: Spectrochim. Acta [ Links ]

14. 109A, 298-307 (2013). [ Links ]

15. Socrates, G.: Infrared and Raman Characteristic Group Frequencies - Tables and Charts. third ed., Wiley, New York, (2001). [ Links ]

16. Karabacak, M., Karagoz, D., Kurt, M.: J. Mol. Struct., 892, 25-31 (2008). [ Links ]

17. Rani, A.U., Sundaraganesan, N., Kurt, M., Cinar, M., Karabacak, M.: Spectrochim. Acta A 75, 1523-1529 (2010). [ Links ]

18. Balachandran, V., Lakshmi, A., Janaki, A.: Spectrochimica Acta A 81, 1-7 (2011) [ Links ]

19. Bevan Ott, J., Boerio-Goates, J. : Calculations from Statistical Thermodaynamics, Academic Press, (2000). [ Links ]

20. Martínez-Araya, J.I., Salgado-Morán, G., Glossman-Mitnik, D. J. Chem., 2013, 1 (2013) [ Links ]

21. Krishnakumar, V., John Xavier, R.: Indian J. Pure Appl. Phys., 41, 597 (2003). [ Links ]

22. Lashgari, A., Ghamami, S., Salgado-Morán, G., Ranirez-Tagle, R., Lorena Gerli - Candia, Journal of the Chilean Chemical Society, 61(1), 2821-2827 (2016). [ Links ]

23. Lashgari, A., Ghammamy, S., Ramirez-Tagle, R., Salgado-Moran, G., Journal of Structural Chemistry, 56(8), 1505-1513, (2015). [ Links ]

24. Ghamami, S., Shahbazkhany, S., Lashgari, A., International Journal of Pure and Applied Chemistry, 9(1-2), 57-62 (2014). [ Links ]

25. Ghamami, S., Lashgari, A., Open Access Library Journal, 1: e422 (2014). [ Links ]

26. Lashgari, A., Ghammamy, S., Shahsavari, M., Asian Journal of Research in Chemistry, 7(7), 677-680 (2014). [ Links ]

Creative Commons License All the contents of this journal, except where otherwise noted, is licensed under a Creative Commons Attribution License