Journal of Molecular Structure 1162 (2018) 125e139
Contents lists available at ScienceDirect
Journal of Molecular Structure journal homepage: http://www.elsevier.com/locate/molstruc
A theoretical and experimental study on isonitrosoacetophenone nicotinoyl hydrazone: Crystal structure, spectroscopic properties, NBO, NPA and NLMO analyses and the investigation of interaction with some transition metals lu a, *, Hümeyra Batı b, Necmi Dege c Ays¸in Zülfikarog a
Department of Chemistry, Faculty of Arts & Sciences, Amasya University, 05000 Amasya, Turkey Department of Chemistry, Faculty of Arts & Sciences, Ondokuz Mayıs University, 55139 Samsun, Turkey c Department of Physics, Faculty of Arts & Sciences, Ondokuz Mayıs University, 55139 Samsun, Turkey b
a r t i c l e i n f o
a b s t r a c t
Article history: Received 9 December 2017 Received in revised form 18 February 2018 Accepted 20 February 2018 Available online 23 February 2018
A new hydrazone oxime compound, isonitrosoacetophenone nicotinoyl hydrazone (inapNH2), was synthesized and characterized by spectroscopic techniques (FT-IR, 1H-NMR and 13C-NMR) and singlecrystal X-ray diffraction. The molecular geometry, NMR chemical shift values and vibrational frequencies of the inapNH2 in the ground state have been calculated by using the Density Functional Method (DFT/B3LYP) with 6-31G(d) and 6-311þþG(d,p) basis sets. The computational results obtained were in agreement with the experimental results. The thermodynamic parameters of the inapNH2 were calculated at different temperatures, and the changes in thermodynamic properties were studied with increasing temperature. The molecular stability originating from charge transfer and hyperconjugative interactions in the title compound was analyzed using Natural Bond Orbital (NBO) and Natural Localized Molecular Orbital (NLMO) analyzes. The Natural Population Analysis (NPA) charges obtained from NBO analysis were used in order to find out the possible coordination modes of the inapNH2 compound with metal ions. To predict the chemical reactivity of the molecule, the molecular electrostatic potential (MEP) surface map of inapNH2 was investigated and some of its global reactivity descriptors (chemical potential m, electronegativity c, hardness h and electrophilicity index u) were calculated using DFT. Furthermore, the strength of metaleligand interaction between chlorides of Co(II), Ni(II), Cu(II), Zn(II) and inapNH2, in both aqueous and ethanol phases, was elucidated by using the values of Charge Transfer (DN) and Energy Lowering (DE). The results indicated that the best interaction in both solvents is between CuCl2 and inapNH2. © 2018 Elsevier B.V. All rights reserved.
Keywords: Hydrazone oxime X-ray diffraction Density functional theory NBO NLMO Metaleligand interactions
1. Introduction Hydrazone and oxime derivatives occupy a special place among organic molecules because of their wide applications in analytical chemistry, medicine, bioorganic systems, electrochemistry, catalysis and industry [1e4]. Hydrazones possess antimicrobial, anticonvulsant, analgesic, anti inflammatory, antiplatelet, antipyretic, antitubercular and antitumoral properties [5e7]. Furthermore, they behave as nematicides, insecticides, herbicides, rodenticides and plant growth regulators. In industry, hydrazones have been widely
* Corresponding author. lu). E-mail address: aysin.zulfi
[email protected] (A. Zülfikarog https://doi.org/10.1016/j.molstruc.2018.02.079 0022-2860/© 2018 Elsevier B.V. All rights reserved.
studied as plasticizers and stabilizers for polymers, polymerization initiators, antioxidants, etc. In published literature, there is also information that these compounds are used as synthetic intermediates in various reactions [8]. In a similar way, oximes also have common areas of usage. These compounds exhibit a wide range of biological and pharmaceutical activities, such as fungicide [9,10] bactericide [9e11], analgesic, anti-inflammatory [12], antioxidant [9] and antitumor [4,13e15] properties. In industrial practices and engineering applications, oximes have been used as corrosion inhibitors for metal or metal alloys, as flotation collectors in mineral processing [16], and in the synthesis of materials with interesting properties such as single-molecule magnetism and single-chain magnetism [17]. Hydrazones and oximes bearing both nitrogen and oxygen
126
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
donors that exhibit both hard and soft base characters are versatile ligands. Owing to their unique chemical reactivity, these compounds can form more stable complexes with metal ions. Therefore, hydrazones and oximes have been widely used in the detection, determination and extraction of metal ions in analytical chemistry for a long time [3,4,8,18e23]. In recent years, computational studies on hydrazones and oximes have been carried out to investigate their molecular properties, including chemical reactivity, charge distribution [24e41], their equilibrium structure, tautomerism and isomerism [27,29,41e45], their UVeVisible, FT-IR, Raman and NMR spectra [24e26,28,31e33,35,41,45e62] and their thermosalient effect [63]. However, to the best of our knowledge, the theoretical studies in published literature are very limited on hydrazone oxime [42,64e68]. Hydrazone oxime compounds include both hydrazone and oxime groups in their structure and are generally formed from condensation reactions of hydrazides with ketoximes [3,17,42,64e83]. In this work, a new hydrazone oxime compound, isonitrosoacetophenone nicotinoyl hydrazone (inapNH2), was synthesized and then characterized using FT-IR, 1H-NMR and 13CNMR, single-crystal X-ray diffraction technique as well as quantum chemical methods. Inspection of the molecular structure, vibrational and NMR spectra of inapNH2 were carried out using density functional theory (DFT/B3LYP) at 6-31G(d) and 6-311þþG(d,p) basis set levels. The bond parameters (bond lengths and angles) of the optimized molecular structures, the computed vibrational frequencies and the NMR chemical shift values are consistent with the experimental results. The thermodynamic properties (thermal energy, heat capacity, entropy, etc.) of inapNH2 were calculated at different temperatures in the gas phase. The delocalization effect from the lone pairs (donor) to the anti-bonding orbitals (acceptor) and the stabilization energies (E(2)) derived from the interactions between the donor and acceptor were investigated to predict molecular stability by means of Natural Bond Orbital (NBO) analysis. In order to shed light on the detailed information about the reactivity of inapNH2, the NPA atomic charges and MEP surface map were obtained and its chemical reactivity descriptors, e.g. chemical potential m, electronegativity c, hardness h, electrophilicity index u, were examined. Finally, the stabilities of metaleinapNH2 interaction were evaluated by energy lowering (DE) and charge Transfer (DN) values in both aqueous and ethyl alcohol phases. 2. Experimental methods 2.1. Material and spectroscopic measurements Nicotinic acid hydrazide, anhydrous ethanol and glacial acetic acid were obtained commercially and used without further purification. Isonitrosoacetophenone was synthesized according to procedures described in published literature [84]. 1H-NMR (400 MHz) and 13C-NMR (500 MHz) spectra of the title compound were recorded on Bruker spectrometers in DMSO-d6 using tetramethylsilane (TMS) as the internal standard. The infrared (IR) spectrum was performed on a MATTSON 1000 FT-IR spectrophotometer in the range of 4000 to 400 cm1 using the KBr pellet technique. 2.2. Synthesis of the title compound Nicotinic acid hydrazide (0.14 g, 1 mmol) and isonitrosoacetephenone (0.15 g, 1 mmol) were dissolved in hot anhydrous ethanol (20 ml). A few drops of glacial acetic acid were added to this solution. The reaction mixture was refluxed for 8 h. The white product formed was separated by filtration and washed with ethanol. The solid product was recrystallized from ethanol and acetone (1:1) to give crystals of the title compound. Scheme 1
shows the synthesis of inapNH2 (yield: 74%, melting point: 183 C). 2.3. X-ray diffraction analysis Single-crystal X-ray data were collected with a Stoe IPDS II [85] single-crystal diffractometer using monochromated MoKa radiation at 296 K. X-AREA [85] and X-RED [85] programs were used for cell refinement and data reduction, respectively. SHELXS-97 [86] and SHELXL-97 [86] programs were used to solve and refine the structure, respectively. ORTEP-3 for Windows [87] was used to prepare the figures. WinGX [88] and PLATON [89] software were used to prepare material for publication. The H atoms of the water molecule were located in a difference map and freely refined. H atoms attached to C, N and O atoms were positioned geometrically [CeH ¼ 0.930 Å, NeH ¼ 0.860 Å and OeH ¼ 0.820 Å] and treated as riding with Uiso(H) ¼ 1.2Ueq(C, N) and Uiso(H) ¼ 1.5Ueq(O). 3. Computational methods All theoretical calculations for the title compound were carried out with the Gaussian 09 quantum chemistry program [90] on a personal computer. Geometric optimization for inapNH2 in the ground state (in vacuo) was performed using molecular geometry taken from X-ray diffraction (XRD) experimental results without any constraints. Initial calculations were carried out using the Density Functional Theory (DFT) method with the restricted B3LYP [91e93] level of theory, using 6-31G(d) and 6-311þþG(d,p) basis sets, for all atoms. The harmonic vibrational frequencies were computed at the same theoretical levels for the molecule in the gas phase. The frequency values computed at these levels contain known systematic errors and, therefore, these values were scaled by 0.983 up to 1700 cm1 and by 0.958 for greater than 1700 cm1 at B3LYP/6311þþG(d,p) [94], while an identical scaling factor of 0.961 was applied for all B3LYP/6-31G(d) frequencies [95]. The vibrational frequency assignments were made using a GaussView molecular visualization program [96]. The vibrational frequency calculations, which gave positive frequency values, indicated that the optimized structures of inapNH2 at these levels were stable. The temperature dependent thermodynamic parameters (heat capacity ðC 0p;m Þ, entropy ðS 0m Þ and enthalpy ðH 0m Þ were calculated in the range from 200 K to 500 K from the vibrational frequency calculations of the title compound in the gas phase using the B3LYP/6-311þþG(d,p) level. The 1H and 13C-NMR chemical shifts for the optimized molecular geometry of the molecule were calculated at B3LYP/6-31G(d) and B3LYP/6-311þþG(d,p) levels by using GIAO (the Gauge-Independent Atomic Orbital) [97] approach in dimethyl sulfoxide (DMSO) solvent. The NMR chemical shifts were converted into the TMS scale by subtracting the calculated absolute chemical shielding of TMSewhose value was 32.16 ppm for the B3LYP/6-31G(d) level and 31.98 ppm for the B3LYP/6311þþG(d,p) level. In order to investigate the molecular stability and hyperconjugative interactions, Natural Bond Orbital (NBO) analysis [98] was carried out at B3LYP/6-311þþG(d,p) by means of the NBO 3.1 program as implemented in the Gaussian 09 W package. The atomic charge values were determined using the Natural Population Analysis (NPA) at B3LYP/6-311þþG(d,p) in the gas phase to predict donor atoms that may be coordinated to the metal ion. The reactive sites of the molecules were studied using molecular electrostatic potential (MEP), which was evaluated using the B3LYP/6-311þþG(d,p) method. Based on the frontier molecular orbitals, the various reactivity descriptors such as the energy gap (D(H-L)), the electronegativity (c), the chemical hardness (h), the chemical potential (m) and the electrophilicity index (u) were calculated in both aqueous and ethanol phases according to
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
127
Scheme 1. The synthesis schematic for inapNH2.
following equations [97e101]:
DðH LÞ ¼ ELUMO EHOMO c ¼ ðEHOMO ELUMO Þ=2 m ¼ c h ¼ ðELUMO EHOMO Þ =2 u¼
m2 2h
To investigate the interactions of inapNH2 as a ligand with metal chlorides, the electronegativity, chemical hardness and chemical potential for CoCl2, NiCl2, CuCl2 and ZnCl2 were also calculated using the B3LYP/6-311þþG(d,p) method in the same solvents. Then the amount of Charge Transfer (DN) and Energy Lowering (DE) [98] which accompany the formation of metal complexes with inapNH2 (donor) and chlorides of Co(II), Ni(II), Cu(II), Zn(II) (acceptor) were computed using the same method.
DN ¼ ðcA cB Þ=2ðhA þhB Þ DE ¼ ðcA cB Þ2 =4ðhA þhB Þ
4. Results and discussion 4.1. Structural properties Isonitrosoacetophenone nicotinoyl hydrazone (inapNH2) crystallizes in the monoclinic, space group P21/c with Z ¼ 4 in the unit cell. Details of crystal parameters, data collection, structure solution and refinement are given in Table 1. The molecular structure and ORTEP [87] view of inapNH2, which contains pyridine and phenyl rings together with hydrazone and oxime groups, is shown Fig. 1. In the molecule, the dihedral angle between the benzene and pyridine ring is 7.48 (9) . As can be seen in Fig. 1, these rings are almost planar (r.m.s. deviation ¼ 0.0016 Å and 0.0037 Å). The optimized structural parameters of inapNH2 were calculated using the DFT/B3LYP method with 6-31G(d) and 6311þþG(d,p) basis sets. Some of the selected bond lengths and angles of the title compound are listed in Table 2 with the experimental data. The optimized molecular geometries of the title compound are shown in Fig. S1a and S1b (See Supplementary materials). The N2eN3 bond distance of 1.364 (17) Å is appreciably shorter than a typical NeN single bond, such as that found in free 2,4-dinitrophenylhydrazone at 1.405 (6) Å [102], n-nonanoic acid
hydrazide at 1.415 Å [103] and 1-benzoyl-2-(propan-2-ylidene) hydrazone at 1.3931 (13) Å [104]; this suggests the existence of a delocalized double bond [62,104,105]. This bond length has been found to be 1.3432 Å for B3LYP/6-31G(d) and 1.3423 Å for the B3LYP/6-311þþG(d,p). The title compound exists in keto form with the C9eO1 bond length of 1.209 (2) Å, typical of a carbon oxygen double bond [56,104,106], which was calculated at 1.2171 Å and 1.2116 Å for DFT using B3LYP/6-31G(d) and the B3LYP/6311þþG(d,p), respectively. The bond distances of the azomethine (C2eN2) and oxime groups (C1eN1) in the title compound were measured as 1.297 (2) Å and 1.269 (2) Å, respectively. These values, which are in good agreement with the data from similar hydrazone oxime compounds [42,46,65,66,107], show that these bonds are typical of a double bond localized between the C and N atoms. These bond distances were calculated to be 1.3022 Å and 1.2892 Å for the B3LYP/6-31G(d), and 1.2976 Å and 1.2836 Å using the B3LYP/6311þþG(d,p). The NeO distance of the free oxime is close to those commonly found in oxime derivatives [N1eO2 ¼ 1.377 (2) Å] [105,107,108] and were calculated to be 1.3954 Å and 1.3962 Å using B3LYP/6-31G(d) and B3LYP/6-311þþG(d,p), respectively. The dihedral angles of oxime and hydrazone groups are O2eN1¼C1eC2 ¼ 177.60 (17) and N3eN2¼C2eC3 ¼ 174.26 (15) . The values found in this study are in agreement with values in published literature, such as 177.6 (3) [66], 178.2 (15) [25], 179.69 (18) [65] and 173.92 (12) [28], 175.4 (3) [66] and 174.73 (12) [104]. These angles have been found to be 176.81 and 179.37 for the B3LYP/6-31G(d) level and 177.33 and 179.12 for the B3LYP/6-311þþG(d,p) level. As shown, both the oxime and hydrazone groups in the title compound are planar and have an E configuration [66,109,110]. The other selected parameters of the title compound are listed in Table 2. Furthermore, the crystal packing of isonitrosoacetophenone nicotinoyl hydrazone (inapNH2) forms an intra- and four intermolecular hydrogen bonds. These hydrogen bonds are shown in Table 3. The intra molecular NeH/N hydrogen bond is shown in Fig. 1. Also, the intermolecular OeH/O and OeH/N hydrogen bonds are shown in Fig. 2a and b. As can be seen in Fig. 1, the intramolecular NeH/N hydrogen bond forms with the N atom of the oxime group acting as an acceptor in the formation of an S(6) motif [42]. In inter-molecular hydrogen bondings, the O3 atom of a water molecule at (x, y, z) acts as a hydrogen-bond bifurcated donor, via atom H3A, to atoms O1 and N2 of the carbonyl and hydrazone groups at (1- x, -y, 1-z). Also, atom O3 at (x, y, z) acts as a hydrogenbond donor, via atom H3B, to atom N4 of the pyridine ring at (1 þ x, y, 1 þ z). As can be seen in Fig. 2a and b, the solvent water molecules form dimers by linking two molecules with different symmetry codes. These dimers proceed along the [101] direction. In addition to these hydrogen bonds, the O2 atom of the oxime group at (x, y, z) acts as a hydrogen-bond donor, via the H2 atom, to atom O3 at (1-x, ½þy,
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
128 Table 1 Crystal data and experimental details for inapNH2. Chemical formula Molecular weight Temperature, K Wavelength, Å Crystal system Space group Unit cell parameters: a, b, c, Å; b, deg Volume, A3; Z Calculated density, g/cm3 Absorption coefficient, mm1 F(000) Crystal size, mm q range, deg h, k, l limits Measured/independent reflections Data completeness on q ¼ 26.5 , % Max. and min. transmission Reflections/restraints/parameters S factor on F2 R factor [I > 2s(I)] R factor (al data) Max. and min. residual electron density, e/A3
C14H12N4O2$H2O 286.29 296 0.71073 Monoclinic P21/c 11.8434(6), 13.3105(9), 9.2772(5), 110.187(4) 1372.63(14), 4 1.385 0.101 600 0.67 0.31 0.11 1.83e27.56 14 h 14, 12 k 16, 11 l 11 7889/2820 [Rint ¼ 0.031] 98.9 0.963 and 0.989 2820/0/198 1.01 R1 ¼ 0.0471, wR2 ¼ 0.097 R1 ¼ 0.0802, wR2 ¼ 0.108 0.144 and 0.195
½-z). All of these hydrogen bonds are effective in crystal packing. As seen in Table 2, the calculated geometrical parameters are generally in good agreement with the experimental structural parameters. There are minor differences between them. The reason for the differences is that the experimental results belong to the solid phase in which inter-molecular interactions occur, while the theoretical calculations are for the gas phase in which the molecules are isolated [111,112]. In order to compare the theoretical results with the experimental values, the root mean square error (RMSE) and maximum difference given in Table 2 are used. The maximum difference between experimental and calculated bond lengths is about 0.0635 Å for 6-31G(d) basis set and 0.0293 Å for 6311þþG(d,p). A global comparison was carried out by superimposing the molecular skeletons obtained from the theoretical calculations and the X-ray diffraction (Fig. 3), giving RMSE values of 0.340 Å and
0.147 Å for B3LYP/6-31G(d) and B3LYP/6-311þþG(d,p) levels, respectively. According to the magnitudes of RMSE, it can be concluded that the 6-311þþG(d,p) basis set correlates well for geometrical parameters when compared to the 6-31G(d) basis set.
4.2. NMR spectral analysis Nuclear Magnetic Resonance Spectroscopy (NMR) is the most powerful and versatile spectroscopic technique for understanding the relationship between the electronic properties and the molecular structure. The combined use of experimental measurements and theoretical calculations help us to accurately interpret and predict the structure of organic molecules and their complexes [26,113,114]. In this study, experimental 1H and 13C-NMR spectra of the title compound's powder were measured in DMSO-d6 (Figs. S2 and S3 Supplementary materials). GIAO 1H and 13C chemical shift values were calculated using the B3LYP method with the 631G(d) and 6311þþG(d,p) basis sets for the optimized geometry with respect to TMS in DMSO (dimethylsulfoxide) solvent. The experimental and calculated chemical shift values are given in Table 1S (See Supplementary materials). In the 1H-NMR spectra for the hydrazone oxime compound, the singlets observed at 13.06 ppm and 12.65 ppm are assigned to the characteristic oxime OH proton and NH proton, respectively. These chemical shifts, which are characteristic for oxime and hydrazone [3,74,77,79], were calculated to be 7.34 ppm (631G), 7.69 ppm (6311þþG(d,p)), and 11.62 ppm (631G) and 13.06 ppm (6311þþG(d,p)), respectively. While the NH proton signal shows good agreement with calculated values, the experimental shift of the oxime OH is significantly larger than calculated values. This may be explained by the inter-molecular hydrogen bonding neglected in the calculations [3]. The other values obtained for 1H-NMR chemical shifts of inapNH2 were (eCH]NeOH) at 9.08 ppm, PyH at 8.8e77.6 ppm and ArH at 7.77 to 7.6 ppm. These data are in agreement with those both previously reported for similar compounds and the calculated values given in Table 1S (See Supplementary materials) [3,113]. In the 13C-NMR spectra, the peak observed at 162.20 ppm is
Fig. 1. ORTEP drawing of the basic crystallographic unit of inapNH2, showing the atom-numbering scheme. Displacement ellipsoids are drawn at the 50% probability level and all H atoms are shown as small spheres of arbitrary radii.
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139 Table 2 Selected experimental and optimized geometric parameters of inapNH2. Parameter N1eC1 N1eO2 O1eC9 N2eC2 N2eN3 C9eN3 C9eC10 C3eC2 N4eC13 N4eC14 RMSEa Max. Differencea Bond angles ( ) C1eN1eO2 C2eN2eN3 O1eC9eN3 O1eC9eC10 N3eC9eC10 C9eN3eN2 N1eC1eC2 N2eC2eC1 N2eC2eC3 C1eC2eC3 RMSEa Max. Differencea Torsion angles ( ) C2eN2eN3eC9 O2eN1eC1eC2 N3eN2eC2eC3 O1eC9eN3eN2 C10eC9eN3eN2 C14eC10eC9eO1 C11eC12eC13eN4 C4eC3eC2eN2 N3eC9eC10eC11 N3eC9eC10eC14 C8eC3eC2eN2 C14eN4eC13eC12
Experimental
DFT/B3LYP/6-31G(d)
DFT/B3LYP/6-311 þþG(d,p)
1.269 (2) 1.377(2) 1.209 (2) 1.297(2) 1.364(17) 1.364(2) 1.499(2) 1.493(2) 1.327(3) 1.332(2)
1.2892 1.3954 1.2171 1.3022 1.3432 1.3945 1.4995 1.4890 1.3401 1.3955 0.0253 0.0635
1.2836 1.3962 1.2116 1.2976 1.3423 1.3933 1.5003 1.4891 1.3377 1.3327 0.0143 0.0293
113.65 117.31 123.11 122.46 114.42 119.42 121.59 126.62 115.30 117.84
111.14 119.98 123.49 122.62 113.89 119.63 121.99 127.08 115.53 117.31 1209 2,67
111.45 120.16 123.69 122.52 113.78 119.76 121.87 126.93 115.73 117.28 1204 2,85
173.61 176.81 179.37 5.28 175.47 22.28 0.52 151.16 24.77 156.97 27.61 0.115
173.22 177.33 179.12 5.81 174.96 26.61 0.25 149.93 29.56 152.64 28.73 0.065
(14) (15) (15) (15) (15) (15) (15) (14) (15) (14)
174.83 (17) 177.60 (17) 174.26 (15) 0.8 (3) 178.64 (15) 9.1 (3) 0.3 (4) 162.61 (18) 10.5 (3) 171.46 (18) 20.1 (3) 0.8 (4)
a RMSE and maximum differences between the bond lengths and angles computed using theoretical methods and those obtained from X-ray diffraction.
assigned to the carbonyl carbon C(9) in the molecule and was calculated to be 155.40 ppm (631G(d)) and 170.96 ppm (6311þþG(d,p)). This signal appeared downfield due to the conjugative effect of the NeNeCeO core of hydrazone [3,115]. The pyridine ring carbons (C13 and C14) attached to the N atom have bigger chemical shifts than other ring carbon atoms because the nitrogen atom has a greater electronegative property, which polarizes the electron distribution in its bond with adjacent carbon atoms [115,116]. The signals in the 136.02 to 127.44 ppm region belong to carbon atoms in a phenyl ring. As seen in Table 1S (See Supplementary materials), the theoretical 13C shift results for the title compound are generally closer to the experimental 13C shift data, which are in agreement with the data from other oximes and hydrazones [3,17,74,77,79,113,116].
4.3. IR spectral analysis The theoretical vibrational spectra were calculated at the B3LYP/ 6-31G(d) and B3LYP/6311þþG(d,p) levels in the gas phase, then scaled by 0.983 up to 1700 cm1 and 0.958 for greater than 1700 cm1 at B3LYP/6311þþG(d,p) [94], while an identical scaling factor of 0.961 was applied for all B3LYP/6-31G(d) frequencies [117]. The experimental and theoretical IR spectra are given in Figs. S4
129
Table 3 Hydrogen-bond geometry (Å, ). DdH/A
DdH
H/A
D/A
N3eH3/N1 O2eH2/O3i O3eH3B/N4ii O3eH3A$$$O1iii O3eH3A$$$N2iii
0.86 0.82 0.82 (2) 0.81 (2) 0.81 (2)
1.92 1.93 2.04 (2) 2.45 (2) 2.53 (3)
2.593 2.667 2.859 3.056 3.301
DdH/A (2) (19) (2) (2) (2)
135 149 174 (3) 133 (2) 160 (3)
Symmetry codes: (i) x þ 1, y þ ½, z þ ½; (ii) x þ 1, y, z þ 1; (iii) x þ 1, y, z þ 1.
and S5 (See Supplementary materials). The vibrational band assignments were made using the Gaussian-View molecular visualization program [96]. The experimental frequencies and the corresponding computed results are summarized in Table 2S (See Supplementary materials). The experimental spectrum of the title compound showed a strong and broad band at 3352 cm1, which is attributed to the characteristic (OeH) oxime absorption. This band was calculated to be 3679 cm1 for the B3LYP/631G(d) level and 3663 cm1 for B3LYP/6311þþG(d,p) level. The band observed at 3135 cm1 indicates NeH stretching in the COeNHeN]CH azomethine group. This band was found to be 3302 cm1 and 3272 cm1 using B3LYP/ 631G(d) and B3LYP/6311þþG(d,p) levels, respectively. The differences between the experimental and theoretical values of these vibrations can be explained by the intra-molecular and intermolecular hydrogen bonding in the solid phase [118,119]. The hydrazone oxime compounds with (C]O), (C]N) and (NeO) groups as donor sites can coordinate to metal ions through the oxygen and nitrogen atoms [3,17,42,74,77,79]. The values of the vibration frequencies of these groups in the free hydrazone oxime compounds are very important because they give information about the coordination mode of the hydrazone oximes with the metal ions. In the experimental spectrum of our compound, y(C] O) stretching vibration produced a band at 1670 cm1, while the calculations for the B3LYP/631G(d) and B3LYP/6311þþG(d,p) levels reproduced this stretching vibration at 1723 cm1 and 1686 cm1, respectively. The oxime and azomethine C]N stretching vibrations give rise to a band at 1594 cm1 and 1543 cm1. The theoretically computed values of C¼Noxime and C¼Nazomethine stretching vibrations were found to be 1622 cm1 and 1535 cm1 for the B3LYP/ 631G(d) level and 1643 cm1 and 1549 cm1 for the B3LYP/ 6311þþG(d,p) level. The band observed at 1005 cm1 attributed to NeO stretching vibrations, which are characteristic for oximes, were calculated to be 997 cm1 for B3LYP/631G(d) level and 1008 cm1 for the B3LYP/6311þþG(d,p) level. As seen from Table 2S (See Supplementary materials), the B3LYP/631G(d) and B3LYP/6311þþG(d,p) levels used in this work produce results close to both each other and to experimental data. 4.4. Thermodynamic properties Several thermodynamic parameters (such as thermal energy, heat capacity, entropy, zero-point vibrational energies (ZPVE), rotational constants and rotational temperatures) were calculated using B3LYP with 6-31G(d) and 6-311þþG(d,p) basis sets at 298.150 K and 1 atm of pressure. Table 4 indicates the values of some thermodynamic parameters without experimental determinations. These calculated results are consistent with reported works [113,116], which are important to test the reliability of our results. Furthermore, based on the vibrational analysis and statistical thermodynamics, the standard thermodynamic functions of heat capacity ðC 0p;m Þ, entropy ðS 0m Þ and enthalpy ðH 0m Þ were obtained at the B3LYP/6-311þþG(d,p) level and listed in Table 5. It can be
130
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
Fig. 2. a ORTEP drawing of the hydrogen bonds of inapNH2 with symmetry codes. 2b PLATON drawing of the hydrogen bonds of inapNH2 in the unit cell.
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
131
Fig. 3. Atom-by-atom superimposition of the structures calculated (blue) by B3LYP/6-31G(d) and B3LYP/6-311þþG(d,p) over the X-ray structure (black) for inapNH2. Hydrogen atoms omitted for clarity. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)
observed that these thermodynamic functions are increasing with temperature ranging from 200 to 500 K due to the fact that molecular vibrational intensities increase with temperature [116]. For inapNH2, the correlation equations between these thermodynamic properties and temperatures are as follows:
C 0p;m ¼ 7:6207 þ 0:2621T 1:0591 104 T 2 S 0m ¼ 71:7976 þ 0:3204T þ 9:5013 105 T 2 H 0m ¼ 0:9149 þ 0:0213T þ 9:4993 105 T 2
R2 ¼ 0:9998
R2 ¼ 0:9999
R2 ¼ 0:9999
All the thermodynamic data can be used for further studies of the title compound and supply helpful information for the synthesis of some new oxime and hydrazone compounds. 4.5. NBO and NLMO analysis Nature Bond Orbital (NBO) analysis provides an efficient method for studying inter-molecular bonding and interaction among bonds and also provides a convenient basis for investigating charge transfer or conjugative interaction in molecular systems. Delocalization of electron density between occupied (bond or lone pairs) NBOs and formally unoccupied (antibond or Rydberg) NBOs correspond to a stabilizing donoreacceptor interaction [120e122]. This donoreacceptor interaction can be quantitatively described in terms of the NBO approach, which is expressed by means of the second-order perturbation interaction energy (E(2)). This energy represents the prediction of the off-diagonal NBO Fock matrix elements. It can be deduced from the second-order perturbation approach [123,124]:
Eð2Þ ¼ DEij ¼ qi
Fði; jÞ2 εj εi
where qi is the ith donor orbital occupancy, εi, εj are diagonal elements (orbital energies) and F(i,j) is the off-diagonal NBO Fock matrix element. The larger the E(2) value, the more intensive is the interaction between electron donors and electron acceptors, i.e. the greater tendency for electron donors to donate to electron acceptors, and the greater the extent of conjugation of the whole system [121]. The Natural Bond Orbital (NBO) analysis has been performed on the title compound at the DFT/B3LYP/6-311þþG(d,p) level. A summary of electron donor orbitals, acceptor orbitals and the interaction stabilization energy that resulted from the secondorder perturbation theory analysis of the Fock matrix in the NBO
basis is presented in Table 6. For this table, the stabilization energies larger than 2 kcal/mol have been chosen. Intra-molecular interactions are formed by the orbital overlap between bonding (CeC), (CeO), (CeN), (CeH), (NeO), (NeN), (NeH) and (OeH), antibonding (CeC), (CeO), (CeN), (CeH), (NeO) and (NeN) orbitals, which results in intra-molecular charge transfer (ICT) causing system stabilization. These interactions are observed as an increase in electron density (ED) in antibonding orbitals that weaken the respective bonds [125]. The electron density of six single bonds of pyridine and phenyl rings (~1.977e) demonstrates delocalization within these rings leading to an average stabilization energy of 3.032 kcal/mol. Similarly, the ED at the conjugated p bonds (1.705e1.626e) and p* bonds (0.374e0.296e) of the pyridine and phenyl rings indicate strong p electron delocalization within rings. The hyperconjugative interactions of p (C3eC4) / p*(C7eC8)/p*(C5eC6), p (C7eC8) / p*(C3eC4)/p*(C5eC6), p (C5eC6) / p*(C3eC4)/ p*(C7eC8) are responsible for conjugation of the respective p bonds in the phenyl ring. These interactions result in an average stabilization energy of 19.95 kcal/mol. The interactions of the conjugate system in the pyridine ring: p (C10eC11) / p*(N4eC14)/ p*(C12eC13), p (N4eC14) / p*(C10eC11)/p*(C12eC13), p * (C12eC13) / p (C10eC11)/p*(N4eC14) also stabilized the pyridine ring up to 27.68 to 13.16 kcal/mol. In addition, LP(1) N4 in the pyridine ring is conjugated to the antibonding orbital of s*(C10eC14) and s*(C12eC13) contributing energy of 9.24 and 9.06 kcal/mol, respectively. The energies for interaction LP(1) N1 / s*(C1eH1) and LP(2) O2 / p* N1eC1 are 6.35 and 17.28 kcal/mol, respectively, indicate the intra-molecular hyperconjugative interactions between the oxime (eHC]NeOH) group. The energies for interaction s (N3eH3) / s*(O1eC9), LP(1) N2 / s*(N3eH3), LP(2) * * O1 / s (C9eN3) and LP(1) N3 / p (O1eC9) are 3.26, 9.11, 28.60 and 43.86 kcal/mol in the azomethine (eCOeNHeN]Ce) group. These interactions demonstrate the single-double bond arrangement that leads to keto-enol tautomerism. The intra-molecular hyperconjugative interactions of s (C1eH1) / s*(N2eC2) [ E(2) ¼ 4.51 kcal/mol], s (C1eC2) * (2) / s (N1eO2) [ E ¼ 4.91 kcal/mol], s (C7eC8) / s*(C3eC2) [ E(2) ¼ 3.24 kcal/mol], s (C4eC5) / s*(C2eC3) [ E(2) ¼ 3.36 kcal/ mol], s (C2eC3) / s*(N2eN3) [ E(2) ¼ 5.20 kcal/mol], s * (2) (N2eN3) / s (C2eC3) [ E ¼ 2.74 kcal/mol], p (N1eC1) / p*(N2eC2) [ E(2) ¼ 9.99 kcal/mol], p (N2eC2) / p*(N1eC1) [ E(2) ¼ 12.16 kcal/mol], p (C3eC4) / p*(N2eC2) [ E(2) ¼ 12.77 kcal/ mol] and p (N2eC2) / p*(C3eC4) [ E(2) ¼ 7.35 kcal/mol] indicate that strong hyperconjugative interactions of the s and p electrons among the oxime (eHC]NeOH) group, phenyl ring and azomethine (eCOeNHeN]Ce) group. However, the p (O1eC9) / p*(C10eC11) [ E(2) ¼ 3.65 kcal/mol] and p (C10eC11) / p*(O1eC9) [ E(2) ¼ 15.40 kcal/mol] bring about
132
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
Table 4 The calculated thermodynamic parameters of inapNH2. Parameters
B3LYP/6-31G(d)
B3LYP/6-311 þþG(d,p)
Thermal energy, E (kcal/mol) Rotational Translational Vibrational Total
0.899 0.899 179.696 181.474
0.889 0.889 178.707 180.484
Heat capacity, Cv (cal/mol K) Rotational Translational Vibrational Total
2.981 2.981 70.206 76.168
2.981 2.981 70.206 76.168
Entropy, S (cal/mol K) Rotational Translational Vibrational Total
34.807 42.851 78.806 156.465
34.799 42.851 81.267 158.918
Rotational constants (GHZ) A B C
0.5812 0.1379 0.1153
0.58225 0.13783 0.11615
Rotational temperature (Kelvin) A B C
0.02789 0.00662 0.00554
0.02794 0.00661 0.00557
Thermal properties (Hartree/particle) Zero-point correction Thermal correction to Energy Thermal correction to Enthalpy Thermal correction to Gibss Free Energy Sum of electronic and zero-point Energies Sum of electronic and thermal Energies Sum of electronic and thermal Enthalpies Sum of electronic and thermal Free Energies Zero-point vibrational energy (kcal/mol)
0.268355 0.289197 0.290141 0.215800 986.226832 986.205990 986.205046 986.279388 168.39525
0.266488 0.287620 0.288564 0.213057 986.528003 986.506871 986.505926 986.581433 167.22343
stabilization between the azomethine (eCOeNHeN]Ce) group and pyridine ring. From the azomethine moiety, three important interactions also lead to the stability of the title molecule. These include the donation of lone pairs of LP(2) O1, LP(1) N2 and LP(1) N3 of the azomethine group to antibonding s*(C9eC10), s*(C1eC2) and s*(N2eC2) that correspond to the stabilization energies of 19.31, 12.10 and 34.39 kcal/mol. The intra-molecular NeH/N hydrogen bond in the NBO analysis result is caused by the interaction between the nitrogen lone pair LP(1) N1 and the antibonding orbital s*(N3eH3). Despite the fact that the energy contribution (9.11 kcal/mol) of the hyperconjugative interaction is weak, this E(2) value is chemically significant and can be used as a measure of the intra-molecular delocalization [117]. In addition, the p*(C10eC11) NBO conjugates with the respective bond of p*(O1eC9) resulting in an enormous stabilization energy of 131.73 kcal/mol. Similarly, the interactions of p*(N1eC1) / p*(N2eC2) [ E(2) ¼ 99.81 kcal/mol] and p*(N2eC2) / p*(C3eC4) [ E(2) ¼ 53.06 kcal/mol] are adding stability to the structure because of their higher E(2) values. Table 7 gives the occupancy and percentage of significant natural atomic hybrids of the Natural Bond Orbital (NBO) of the title compound calculated using the B3LYP/6-311þþG(d,p) method. The NBO hybrid analysis shows that all the CeN/NeH bond orbitals are polarized towards the nitrogen atom (ED, % ¼ 54.43 to 72.58 for N) and CeO/OeH/NeO bond orbitals towards the oxygen atom (ED,
Table 5 Thermodynamic properties of inapNH2 at different temperatures at the B3LYP/6311þþG(d,p). T(K)
H0m (kcal mol1)
ðC0p;m Þ (cal mol1 K1)
S0m (cal mol1 K1)
200.00 250.00 298.15 300.00 350.00 400.00 450.00 500.00
7.170 10.327 13.853 13.998 18.175 22.832 27.935 33.444
56.040 66.294 76.168 76.543 86.458 95.753 104.268 111.957
131.997 146.045 158.918 159.400 172.261 184.684 196.694 208.296
T: temperature; ðC0p;m Þ, heat capacity; ðS0m Þ entropy; (H0m Þ enthalpy.
% ¼ 58.33 to 73.69 for O). The hydrogen atoms have almost 0% p character in the OeH, NeH and CeH (eHC]NeOH) bonds. The first lone pair of N9 and second lone pairs of O2 and O3 have almost 100% p character. The natural localized molecular orbitals (NLMOs) show how bonding in a molecule is composed of orbitals localized around different atoms. The derivation of NLMOs from NBOs gives direct insight into the nature of the localized molecular orbital's “delocalization tails” [120,126]. Table 8 shows occupancy of significant NLMOs, the percentage of parent NBO and the atomic hybrid contributions of the title compound calculated at the B3LYP level using the 6-311þþG(d,p) basis set. As seen from Table 8, the most delocalized NLMO is p (C12eC13) for inapNH2. It has only about 80.5% contribution from the localized p (C12eC13) parent NBO, with the delocalized tail composed of contributions (~19.5) from hybrids of O1, C9, C10, C11, N4 and C14. Similarly, the NLMO of p(C7eC8) has a delocalized tail (~19) composed of hybrids of C3, C4, C5 and C6. The other most delocalized NLMOs in the title compound are p(C10eC11), p(C3eC4), p(N4eC14), p(C5eC6) and the NLMO due to the first lone pair of the nitrogen atom N3. They have nearly 15e19% delocalized tails. These larger amounts of delocalization indicate that the donor (C]C)/C]N bonds in the rings and the first lone pair of the nitrogen atom N9 are strongly delocalized into the neighboring regions. These interactions, which have rather larger E(2) values that can also be observed from the perturbation theory energy analysis are given in Table 6. 4.6. Atomic charges The calculation of atomic charges plays an important role in the application of quantum mechanical calculations to molecular systems because atomic charges affect dipole moment, molecular polarizability, electronic structure, acidityebasicity behavior and a lot more properties of molecular systems [127,128]. Natural Population Analysis (NPA) calculates atomic charges and orbital populations of molecular wave functions for general atomic orbital basis sets [129]. For the title molecule, the net charges are calculated by Natural Population Analysis (NPA). The calculated natural atomic charge values are obtained from the Natural Bond Orbital Analysis (NBO) [98] and are listed in Table 9. As can be seen in Table 9, the magnitudes of the carbon atomic charges change from 0.24228 to 0.6613. The carbonyl carbon atom (C9 ¼ 0.66113) is the most positive carbon atom due to the highly electronegative oxygen atom (O1 ¼ 0.56375) and nitrogen atom (N3 ¼ 0.44340) present in the adjacent position. In the phenyl and pyridine rings, because C atoms have a higher electronegativity than H atoms, seven C atoms bonded to H atoms have negative atomic charges. On the other hand, the C14 (0.10218) and C13 (0.06708) atoms in the pyridine ring have positive atomic charges. This is due to the attachment of the negatively charged nitrogen (N4) atom. Similarly, the carbon atoms (C1 ¼ 0.01338 and
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
133
Table 6 Significant donoreacceptor interactions of inapNH2 and their second order perturbation energies. Type
Donor NBO (i)
ED (i)/e
Type
Acceptor NBO (j)
ED (i)/e
E(2) (kcal/mol)
E(j)-E(i) a.u.
F(i,j) a.u.
s p p s s s
N1eO2 N1eC1 O1eC9 O2eH2 N2 eN3 N2eC2
1.98782 1.94433 1.97805 1.99214 1.98652 1.98552
p
N2eC2
1.87699
s
C9eC10
1.97361
s s
C9eN3 C10eC11
1.98835 1.97484
p
C10eC11
1.64301
s s
C10eC14 C3eC14
1.97763 1.97335
p
C3eC4
1.64893
s
C3eC8
1.97154
s
C2eC3
1.96691
s s s
N3eH3 C1eH1 C1eC2
1.98434 1.97821 1.97148
s
C4eC5
1.97833
s
C7eC8
1.97915
p
C7eC8
1.65778
s
C11eC12
1.97899
p
N4eC14
1.70497
s
C5eC6
1.97965
p
C5eC6
1.66467
s
C6eC7
1.97990
s p
C12eC13 C12eC13
1.98522 1.62616
LP(1)
N1
1.93641
LP(1) LP(2)
O1 O1
1.98036 1.85756
O2 N2
1.89767 1.92114
LP(1)
N3
1.62488
LP(1)
N4
1.91792
LP(2)
O3
1.99034
C1eC2 N2eC2 C10eC11 N1eC1 C2eC3 C9eN3 C2eC3 N1eC1 C3eC4 N2eN3 C10eC11 C11eC12 N4eC14 N2eC2 C10eC14 C11eC12 O1eC9 N4eC14 C12eC13 C10eC11 C3eC8 C4eC5 N2eC2 C7eC8 C5eC6 C3eC4 C2eC3 C7eC8 N2eN3 N2eC2 C3eC8 C4eC5 C7eC8 O1eC9 N2eC2 N1eO2 C3eC8 C3eC4 C2eC3 C5eC6 C3eC8 C2eC3 C6eC7 C3eC4 C5eC6 C9eC10 C10eC11 C12eC13 C10eC11 C12eC13 C4eC5 C6eC7 C3eC4 C7eC8 C7eC8 C5eC6 C11eC12 C10eC11 N4eC14 N3eH3 C1eH1 C9eC10 C9eC10 C9eN3 N1eC11 C3eC2 N3eH3 C1eC2 O1eC9 N2eC2 C10eC14 C12eC13 C1eH1
0.03996 0.25907 0.34781 0.00999 0.03301 0.08914 0.03301 0.17435 0.37359 0.02107 0.02321 0.01603 0.01338 0.01709 0.03139 0.01603 0.26068 0.33356 0.29654 0.02321 0.02348 0.01554 0.25907 0.29705 0.33530 0.02373 0.03301 0.01495 0.02107 0.01709 0.02348 0.01554 0.01495 0.01407 0.01709 0.02679 0.02348 0.02373 0.03301 0.01634 0.02348 0.03301 0.01656 0.37359 0.33530 0.06809 0.02321 0.02599 0.34781 0.29654 0.01554 0.01656 0.37359 0.29705 0.01495 0.01634 0.01603 0.34781 0.33356 0.05963 0.02818 0.06809
LP(2) LP(1)
s* p* p* s* s* s* s* p* p* s* s* s* s* s* s* s* p* p* p* s* s* s* p* p* p* s* s* s* s* s* s* s* s* s* s* s* s* s* s* s* s* s* s* p* p* s* s* s* p* p* s* s* p* p* s* s* s* p* p* s* s* s* s* s* p* s* s* s* p* p* s* s* s*
3.39 9.99 3.65 2.85 2.74 2.29 2.04 12.16 7.35 3.76 2.13 2.00 2.78 2.39 3.56 2.87 15.40 26.40 16.65 3.71 3.85 2.98 12.77 18.71 20.31 3.94 2.12 2.81 5.20 2.55 2.03 2.18 2.02 3.26 4.51 4.91 2.08 3.38 3.36 2.76 3.18 3.24 2.72 21.31 21.27 3.56 3.17 2.35 13.16 27.68 2.80 2.61 20.40 18.69 2.70 2.65 2.54 23.95 15.77 11.53 6.35 2.23 19.31 28.60 17.28 2.26 9.11 12.10 43.86 34.39 9.24 9.06 2.71
1.29 0.36 0.40 1.34 1.33 1.31 1.35 0.32 0.36 1.07 1.22 1.23 1.21 1.40 1.27 1.28 0.30 0.28 0.29 1.27 1.27 1.27 0.26 0.29 0.28 1.25 1.14 1.28 1.05 1.23 1.22 1.23 1.24 1.29 1.12 0.90 1.25 1.27 1.15 1.28 1.27 1.15 1.27 0.28 0.28 1.13 1.28 1.28 0.32 0.32 1.28 1.28 0.28 0.29 1.29 1.27 1.29 0.28 0.27 0.85 0.88 1.10 0.66 0.66 0.33 0.83 0.77 0.82 0.30 0.28 0.89 0.90 0.94
0.059 0.057 0.037 0.055 0.054 0.050 0.047 0.057 0.049 0.057 0.045 0.044 0.052 0.052 0.060 0.054 0.062 0.077 0.063 0.061 0.062 0.055 0.053 0.067 0.068 0.063 0.044 0.054 0.066 0.050 0.045 0.046 0.045 0.058 0.064 0.059 0.046 0.058 0.056 0.053 0.057 0.055 0.053 0.068 0.069 0.057 0.057 0.049 0.058 0.084 0.053 0.052 0.069 0.066 0.053 0.052 0.051 0.074 0.059 0.089 0.067 0.045 0.103 0.124 0.069 0.039 0.075 0.090 0.106 0.091 0.082 0.082 0.045
0.08914 0.17435 0.03301 0.05963 0.03996 0.26068 0.25907 0.03139 0.02599 0.02818
(continued on next page)
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
134 Table 6 (continued ) Type
Donor NBO (i)
ED (i)/e
Type
Acceptor NBO (j)
ED (i)/e
E(2) (kcal/mol)
E(j)-E(i) a.u.
F(i,j) a.u.
p* p* p*
N1eC1 N2eC2 C10eC11
0.17736 0.24961 0.34781
p* p* p*
N2eC2 C3eC4 O1eC9
0.25907 0.37359 0.26068
99.81 53.06 131.73
0.02 0.02 0.01
0.075 0.056 0.066
C2 ¼ 0.14347) of the oxime and azomethine groups also have positive charges. All hydrogen atoms are positively charged. The hydrogen atoms (H2 ¼ 0.46901 and H3 ¼ 0.40834) attached to the oxime oxygen (O3) and amide nitrogen (N9) bear more positive charge compared to others. Four nitrogen atoms of the title molecule exhibit negative charges, but the N3 atom joined in the formation of the intra-molecular hydrogen bond and the N4 atom in the pyridine ring have a higher negative charge value than the
other nitrogen atoms. However, the maximum negative atomic charges are around the oxygen atoms (O1 ¼ 0.56378 and O2 ¼ 0.53511) of the carbonyl and oxime groups. On the whole, considering the atomic charge distributions and the steric effect, the O2, O3, N1, N5 and N9 atoms in which the negative charge is delocalized can serve as coordinating sites for metal ions and thus the title compound can act as a multidentate ligand through the nitrogen and oxygen atoms.
Table 7 Selected Lewis orbitals (occupied bond or lone pair) of inapNH2 with their valence hybrids performed at B3LYP/6-311þþG(d,p): Lpeloan pair, ED(e)eelectron density and (ED, %)epercentage electron density. Bond
Type
ED(e)
Atom
ED, %
%s
%p
%d
NBO Hybrid Orbital
N1eO2
s
1.98782
N1eC1
s
1.99181
p
1.94433
s
1.99302
p
1.97805
O2eH2
s
1.99214
N2eN3
s
1.98652
N2eC2
s
1.98552
p
1.87699
C9eC10
s
1.97361
C9eN3
s
1.98835
C10eC14
s
1.97763
C3eC4
s
1.97335
p
1.64893
C3eC8
s
1.97154
C3eC2
s
1.96691
N3eH3
s
1.98434
C1eH1
s
1.97821
C1eC2
s
1.97148
N4eC14
s
1.98662
p
1.70497
N4eC13
s
1.98661
LP(1)N1 LP(1)O1 LP(2)O1 LP(1)O2 LP(2)O2 LP(1)N2 LP(1)N3 LP(1)N4
e e e e e e e e
1.93641 1.98036 1.85756 1.99201 1.89767 1.92114 1.62488 1.91792
N1 O2 N1 C1 N1 C1 O1 C9 O1 C9 O2 H2 N2 N3 N2 C2 N2 C2 C9 C10 C9 N3 C10 C14 C3 C4 C3 C4 C3 C8 C3 C2 N3 H3 C1 H1 C1 C2 N4 C14 N4 C14 N4 C13 N1 O1 O1 O2 O2 N2 N3 N4
41.67 58.33 59.40 40.60 58.37 41.63 64.29 35.71 68.36 31.64 73.69 26.31 47.11 52.89 59.07 40.93 54.43 45.57 48.05 51.95 37.07 62.93 52.05 47.95 51.01 48.99 50.37 49.63 51.39 48.61 49.79 50.21 72.58 27.42 62.26 37.74 49.69 50.31 59.11 40.89 60.09 39.91 59.14 40.86 e e e e e e e e
16.54 21.12 42.37 32.37 0.01 0.03 40.34 32.79 0.52 0.49 21.46 99.84 25.12 30.56 41.86 31.91 0.00 0.01 36.41 30.52 30.13 36.72 33.94 37.93 34.56 35.52 0.02 0.02 34.51 34.97 30.89 33.74 32.39 99.91 33.02 99.94 34.84 34.26 35.83 32.10 0.00 0.00 35.40 31.82 41.50 59.17 0.01 57.50 0.00 33.08 0.36 28.87
83.22 78.77 57.54 67.53 99.84 99.77 59.54 67.05 99.36 99.11 78.45 0.16 74.75 69.37 58.05 68.00 99.84 99.86 63.55 69.44 69.76 63.24 66.02 62.03 65.40 64.45 99.95 99.94 65.46 64.99 69.07 66.23 67.56 0.09 66.93 0.06 65.13 65.71 64.08 67.81 99.85 99.85 64.51 68.08 58.46 40.81 99.92 42.48 99.95 66.85 99.64 71.05
0.24 0.11 0.09 0.10 0.14 0.20 0.12 0.15 0.12 0.41 0.09 e 0.13 0.07 0.08 0.09 0.16 0.13 0.04 0.05 0.11 0.04 0.04 0.04 0.04 0.04 0.03 0.04 0.04 0.04 0.04 0.03 0.04 e 0.05 e 0.04 0.04 0.09 0.10 0.15 0.15 0.09 0.10 0.03 0.02 0.08 0.02 0.05 0.07 0.01 0.08
sp5.03 sp3.73 sp1.36 sp2.09 p p sp1.48 sp2.04 p p sp3.66 s sp2.98 sp2.27 sp1.39 sp2.13 p p sp1.75 sp2.28 sp2.32 sp1.72 sp1.94 sp1.64 sp1.89 sp1.81 p p sp1.90 sp1.86 sp2.24 sp1.96 sp2.09 s sp2.03 s sp1.87 sp1.92 sp1.79 sp2.11 p p sp1.82 sp2.14 sp1.41 sp0.69 p sp0.74 p sp2.02 p sp2.46
O1eC9
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139 Table 8 The occupacy of significant NLMOs, percentage from parent NBO and atomic hybrid contributions of inapNH2. Bond
Occupany Percentage from parent NBO Hybrid contributions
p(C10eC11) 2.00000
81.2067
p(C3eC4)
2.00000
82.3483
p(C7eC8)
2.00000
80.9905
p(N4eC14)
2.00000
84.6481
p(C5eC6)
2.00000
82.4490
p(C12eC13) 2.00000
80.5351
LP(1)N3
2.00000
80.6170
Atom
%
O1 C9 C10 N3 C11 N4 C14 C12 C13 N2 C9 C3 C4 C8 C2 C5 C6 C7 C3 C4 C8 C5 C6 C7 C10 C11 N4 C14 C12 C13 N2 C8 C4 C8 C2 C5 C6 C7 O1 C9 C10 C11 N4 C14 C12 C13 N1 O1 N2 C9 C10 C3 N3 C1 C2
0.744 1.500 46.456 0.206 34.860 2.034 8.070 3.071 2.891 0.825 0.100 41.584 40.766 6.087 1.049 6.069 1.808 1.366 4.550 4.865 39.666 4.574 4.847 41.326 2.712 2.965 52.871 31.826 1.992 7.557 0.212 3.643 4.257 2.054 0.276 39.704 42.790 6.910 0.141 0.287 2.466 9.417 2.987 4.016 45.570 35.026 0.499 3.114 4.280 6.600 0.129 0.244 80.631 0.658 3.257
4.7. Molecular electrostatic potential The molecular electrostatic potential (MEP), created in the space around a molecule by its nuclei and electrons, is very helpful in interpreting and predicting the reactive behavior of a wide variety of chemical systems in both electrophilic and nucleophilic reactions, which are the study of biological recognition processes, including drugereceptor, enzymeesubstrate interactions and hydrogen bonding interactions [130,131]. ! The electrostatic potential, V( r ) at any point r is given by the following equation:
135
!0 ! Z r d r0 r X Z ! A Vð r Þ ¼ ! ! R A r/ r0 r/ A where ZA is the charge on nucleus A, located at RA, and p(r) is the electronic density function of the molecule obtained from its ! computed wave function. V( r ) is a real physical property that can be determined experimentally as well as computationally [130,132]. To investigate the reactive sites of the title compound, the molecular electrostatic potentials were evaluated using the B3LYP/6311þþG(d,p) method. The different values of the electrostatic potential at the surface are defined by different colors. Potential increases in the order red < orange < yellow < green < light blue < blue. Fig. 4 shows the electrostatic potential contour map where the negative (red and yellow) regions of MEP are related to electrophilic attacks and the positive (blue) regions are related to nucleophilic reactivity. As can be seen in Fig. 4, the negative regions in the studied molecule are mainly localized over the O1, N4, N2, N3, O2 and N1 atoms with values around 0.063 a.u., 0.045 a.u., 0.038 a.u., 0.035 a.u., 0.028 a.u. and 0.002 a.u., respectively. Thus, it can be predicted that the most preferred region for electrophilic attack is around the oxygen atom (O1) of the carbonyl group. By contrast, a maximum value of 0.069 a.u. is found on the oxime hydrogen (H2) atom in the positive regions of V(r) indicating that this site is probably involved in nucleophilic processes. The MEP is also used to identify sites for intra- and intermolecular interactions [133]. When an intra-molecular interaction occurs, the electrostatic potential of the negative atom becomes less negative and the positive region of the other atom becomes less positive [133,134]. For the MEP surface in the studied molecule, the weak negative region associated with the N1 atom, and also the weak positive region at the nearby H3 atom, are indicative of intra-molecular (N1/H3eN3) hydrogen bonding [134,135]. 4.8. Frontier molecular orbital energies and chemical reactivity The Density Functional Theory (DFT) is one of the important tools to define and justify chemical concepts that give information about the reactivity of molecules [136e138]. For the title compound, the chemical reactivity descriptors were obtained by the DFT method employing the B3LYP/6-311þþ(d,p) level. Table 10 shows the values of the calculated frontier orbital energies (EHOMO, ELUMO), total energies (Etotal), energy gap (D(H-L)) and global reactivity descriptors (electronegativity (c) chemical potential (m), chemical hardness (h) and electrophilicity index (u)) for inapNH2 in aqueous and ethyl alcohol phases. The frontier orbitals, such as HOMO (Highest Occupied Molecular Orbital) and LUMO (Lowest Unoccupied Molecular Orbital), are very important quantum chemical parameters because they determine the way molecules interacts with each other. The HOMO orbital can act as an electron donor, while the LUMO orbital can act as an electron acceptor. The molecules with a high energy HOMO orbital tend to donate electrons to lower energy unoccupied orbitals of an acceptor such as a metal ion [16,137]. The HOMOeLUMO gap and total energy are important stability indexes. A low total energy and large band gap indicate high stability and low chemical reactivity. In other words, a smaller band gap and higher total energy indicate less stability and higher reactivity of a molecule, and the easier the electron transition will be [16]. In addition to this, the lower the value of chemical potential (m), chemical hardness (h)
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
136
Table 9 Calculated atomic chargesa by Natural Population Analysis (NPA). Atom N1 O1 O2 H2 N2 C9 C10 C3 N3 H3 C1 H1 C4 H4 C8 H8 C2 C11 a
B3LYP/6-311þþG(d,p) (NPA) natural charges
Atom
0.18445 0.56375 0.53511 0.46901 0.20032 0.66113 0.17247 0.07011 0.44340 0.40834 0.01338 0.23232 0.20612 0.21334 0.16648 0.23014 0.14347 0.15634
H11 N4 C5 H5 C6 H6 7C H7 C14 H14 C12 H12 C13 H13 O3 H3A H 3B
B3LYP/6-311þþG(d,p) (NPA) natural charges 0.21152 0.44564 0.21109 0.20805 0.19931 0.20633 0.19674 0.20803 0.10218 0.20957 0.24228 0.21373 0.06708 0.18926 0.92718 0.46988 0.46401
Atomic charge unit (a.u.).
Table 10 The calculated quantum chemical parameters of inapNH2. Quantum chemical parameter
Water (Ɛ ¼ 78.39)
Ethanol (Ɛ ¼ 24.55)
EHOMO (a.u) ELUMO (a.u) Etotal (a.u) D(H-L) (a.u) (HOMO-LUMO) c (a.u) m (a.u) h (a.u) u (a.u)
0.2408 0.0969 986.8176 0.14390 0.16888 0.16888 0.07195 0.198196
0.2406 0.0967 986.8164 0.14388 0.16865 0.16865 0.07194 0.197684
Table 11 Interaction of donor (inapNH2) with CoCl2, NiCl2, CuCl2 and ZnCl2. Acceptor
hA
Aqueous phase 0.08946 CoCl2 NiCl2 0.07999 CuCl2 0.09041 ZnCl2 0.06269 Ethanol phase 0.08817 CoCl2 NiCl2 0.09022 CuCl2 0.06246 ZnCl2 0.12439
hD
cA
cD
DN
DE
0.07195 0.07195 0.07195 0.07195
0.17678 0.17866 0.22475 0.17996
0.16888 0.16888 0.16888 0.16888
0.7074 0.8196 5.6458 0.7957
0.0027 0.0038 0.1497 0.0042
0.07194 0.07194 0.07194 0.07194
0.19222 0.17781 0.22443 0.17871
0.16865 0.16865 0.16865 0.16865
2.0026 0.7686 5.6465 0.6968
0.0236 0.0035 0.1575 0.0035
and global electrophilicity index (u), the better the donor properties will be [16,138]. The energy gap, EHOMO, total energy, chemical potential (m), chemical hardness (h) and global electrophilicity index (u) values given in Table 10 for inapNH2 in both the aqueous and ethanol phases show that the title compound has slightly stronger electron-donating ability in the ethanol phase.
D: donor (inapNH2) A: acceptor (metal(II) chlorides).
4.9. Metal (acceptor)eLigand (donor) interaction
5. Conclusions
The metaleligand bond strength between the interaction of acceptor and donor has recently been evaluated with the help of quantum chemical parameters, such as the energy lowering DE and charge transfer DN values [136,138]. The values of DE and DN for the interaction of donor (inapNH2 ligand) and acceptor (CoCl2, NiCl2, CuCl2 and ZnCl2) are evaluated and listed in Table 11. The metal bond strength between the interaction of acceptor (metal (II) chloride) and donor (inapNH2 ligand) increases as the charge transfer (DN) increases and the energy lowering (DE) decreases. The bond strength of inapNH2 with metal chlorides based on the values DN and DE follow the sequence CuCl2 > NiCl2 z ZnCl2 z CoCl2 in the aqueous phase and CuCl2 > CoCl2 > NiCl2 z ZnCl2 in the ethanol phase. The values of DN and DE indicate that the most stable complex formation in both is between
In this study, a new hydrazone oxime compound was synthesized and characterized by X-ray crystallography, IR, 1H-NMR and 13 C-NMR spectral techniques. In addition to the experimental studies, most of the properties of the molecular structure were calculated using B3LYP/6-31G(d) and B3LYP/6-311þþG(d,p) methods. The X-ray structure is slightly different from its optimized molecular structures because the crystal structure in the solid phase is stabilized by the inter-molecular OeH/O and OeH/N hydrogen bonds that supply the leading contribution to stability. However, the theoretical calculations were conducted in the gas phase. The B3LYP/6-311þþG(d,p) method can reproduce the title compound better than the B3LYP/6-31G(d) method. In order to study vibrational, NMR and thermodynamic properties of inapNH2, the theoretical calculations were successfully performed by using DFT with the B3LYP/6-31G(d) and B3LYP/6-311þþG(d,p) levels. The spectroscopic results obtained are in good agreement with the experimental values. The correlations between the thermodynamic properties (heat capacity, entropy, enthalpy) and temperatures show that thermodynamic parameters increase with increasing temperature. The intra- and inter-molecular contacts and the stability of the title compound have been interpreted using NBO and NLMO analysis with B3LYP/6-311þþG(d,p). NBO/NPA atomic charges were determined to predict the possible electron donating centers on the inapNH2 that can likely interact with metal ions. The results indicated that the inapNH2 may coordinate with carbonyl, oxime group oxygen and with azomethine, oxime and amide nitrogen atoms as a multidentate ligand to the metal ions. By using HOMO and LUMO energy values for the molecule in the aqueous and ethanol phases, quantum chemical parameters have been calculated at the B3LYP/6-311þþG(d,p) level. The results show that the reactivity of the compound is slightly higher in the ethanol phase than in the water phase. Finally, the metal bond strength between the interaction of metal(II) chlorides (acceptor) and inapNH2 (donor) were investigated with the quantum chemical
Fig. 4. Molecular Electrostatic Potential map calculated with the B3LYP/6311þþG(d,p) level.
inapNH2 and CuCl2.
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
parameters of energy lowering DE and charge transfer DN in both solvents. The values of the DE and DN parameters for interaction between the inapNH2 and CoCl2, NiCl2, CuCl2 and ZnCl2 indicate that the inapNH2 can form complexes with these metal(II) chlorides, and the most stable complex formation is between inapNH2 and the Cu(II) ion. Based on these results, it can be said that inapNH2 can be used as a chelating ligand in the detection, determination and extraction of metal ions, especially Cu (II) ions, in both aqueous and ethanol phases. Acknowledgements € r for his help The authors thank Professor Dr. Orhan Büyükgüngo with the data collection and acknowledge the Ondokuz Mayıs University Research Projects Unit for financial support of project F446 and the Amasya University Research Projects Unit for financial support of project FMB-BAP-018.
[16]
[17]
[18]
[19]
[20] [21]
[22]
Appendix A. Supplementary data
[23]
Supplementary data associated with this article can be found in the online version, at https://doi.org/10.1016/j.molstruc.2018.02. 079. These data include MOL files and InChiKeys of the most important compounds described in this article.
[24]
References
[25]
basan, Crystal structure of N-Benzlidene[1] H. Tezcan, T. Tunç, E. S¸ahin, R. Yag N0 -(2-carboxyphenyl)hydrazine, Anal. Sci. 20 (2004) 137e138. [2] A.S. El-Tabla, M.M.A. El-Waheedb, M.M.E. Shakdofac, N.A.A. El-Fadla, Synthesis, spectroscopic characterization and antitumor activity of new metal complexes of isonicotinoylhydrazide oxime, Main Group Chem. 12 (2013) 153e168. lu, Crystal Structure of N-Benzlidene-N0 -(2-carboxyphenyl) [3] R. Gup, E. Gizirog Metal complexes and solvent extraction properties of isonitrosoacetophenone 2-aminobenzoylhydrazonehydrazine, Spectrochım. Acta A 65 (2006) 719e726. [4] A.S. El-Tabl, M.M.A. El-wahed, S.E.A. El-Razek, Preparation, spectroscopic investigation and antiproliferative capacity of new metal complexes of (3E)2-(hydroxyimino)-N-P-Tolyl-3-(P-tolylimino) butanamide, Spectrochım. Acta A 105 (2013) 600e611. [5] S. Rollas, S¸.G. Küçükgüzel, Biological activities of hydrazone derivatives, Molecule 12 (2007) 1910e1939. [6] H.J.C. Bezerra-Netto, D.I. Lacerda, A.L.P. Miranda, H.M. Alves, E.J. Barreiroa, C.A.M. Fragaa, Design and synthesis of 3,4-methylenedioxy-6nitrophenoxyacetylhydrazone derivatives obtained from naturalsafrole: new lead-agents with analgesic and antipyretic properties, Bioorgan. Med. Chem. 14 (2006) 7924e7935. it Kaymakçıog lu, S. Rollas, Synthesis, characterization and evalua[7] B. Koçyig tion of antituberculosis activity of some hydrazones, II Farmaco 57 (2002) 595e599. [8] O. Katyal, Y. Dutt, Analytical applications of hydrazones, Talanta 22 (1975) 151e166. [9] A. Colak, U. Terzi, M. Col, S.A. Karaoglu, S. Karabocek, A. Kucukdumlu, F.A. Ayaz, DNA binding, antioxidant and antimicrobial activities of homoand heteronuclear copper(II) and nickel(II) complexes with new oxime-type ligands, Eur. J. Med. Chem. 45 (2010) 5169e5175. [10] A.S. El-Tabl, W. Plass, A. Buchholz, M.M.E. Shakdofa, Synthesis, spectroscopic investigation and biological activity of metal(II) complexes with N2O4 ligands, J. Chem. Res. 9 (2009) 582e587. [11] A. Yilmaz, B. Taner, P. Deveci, A.Y. Obali, U. Arslan, E. Sahin, H.I. Ucan, E. Ozcan, Novel bioactive vic-dioxime ligand containing piperazine moiety: synthesis, X-ray crystallographic studies, 2D NMR applications and complexation with Ni (II), Polyhedron 29 (2010) 2991e2998. [12] M.I. El-Gamal, S.M. Bayomi, S.M. El-Ashry, S.A. Said, A.A.M. Abdel-Aziz, N.I. Abdel-Aziz, Synthesis and anti-inflammatory activity of novel (substituted)benzylidene acetone oxime ether derivatives: molecular modeling study, Eur. J. Med. Chem. 45 (2010) 1403e1414. [13] N.M. Krstie, M.S. Bjelakovie, Z. Zizak, M.D. Pavlovic, Z.D. Juranic, V.D. Vladimir, Synthesis of some steroidal oximes, lactams, thiolactams and their antitumor activities, Steroids 72 (2007) 406e414. [14] H.J. Park, K. Lee, S.J. Park, B. Ahn, J.C. Lee, H. Cho, K.I. Lee, Identification of antitumor activity of pyrazole oxime ethers, Bioorg. Med. Chem. Lett 15 (2005) 3307e3312. [15] N.I. Dodof, M. Kubiak, J. Kuduk-Jaworska, A. Mastalarz, A. Kochel, V. Vassilievia, N. Vassilev, N. Trendaflova, I. Georgieva, M. Lalia-Kantouri,
[26]
[27] [28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
137
M. Apostolova, Structure, NMR spectra and cytotoxic effect of palladium(II) and platinum(II) complexes of glyoxylic acid oxime, Chemija 20 (2009) 208e217. X. Yang, S. Liu, G. Liu, H. Zhong, A DFT study on the structureereactivity relationship of aliphatic oxime derivatives as copper chelating agents and malachite flotation collectors, J. Ind. Eng. Chem. 46 (2017) 404e415. M. Sutradhar, T.R. Barman, J. Klanke, M.G.B. Drew, E. Rentschler, A novel Cu(II) dimer containing oxime-hydrazone Schiff base ligands with an unusual mode of coordination: study of magnetic, autoreduction and solution properties, Polyhedron 53 (2013) 48e55. K.H. Reddy, N.B.L. Prasad, T.S. Reddy, Analytical properties of 1-phenyl-1,2propanedione-2-oxime thiosemicarbazone: simultaneous spectrophotometric determination of copper(II) and nickel(II) in edible oils and seeds, Talanta 59 (2003) 425e433. F. Chen, K. Li, H. Li, Synthesis, structural, and electrocatalytic properties of copper (II) complex of ONO donor acylhydrazone and 2-Methylimidazole, Synth. React. Inorg. Me. 44 (2014) 1029e1033. M.S. Ray, A. Ghosh, S. Chaudhuri, M.C.B. Drew, I. Ribas, Eur. J. Inorg. Chem. 15 (2004) 3110e3117. S. Chattopadhay, M.S. Ray, S. Chaudhuri, G. Mukhopahyay, G. Bocelli, A. Cantoni, A. Ghosh, Nickel(II) and copper(II) complexes of tetradentate unsymmetrical Schiff base ligands: first evidence of positional isomerism in such system, Inorg. Chem. Acta 359 (2006) 1367e1375. K. Szabo, N. Marek, Investigation of catalytic reaction with spectrofluorimetric method, J. Biochem. Biophys. Methods 53 (2002) 189e191. N. Galic, M. Rub cic, K. Magdic, M. Cindric, V. Tomisic, Solution and solid-state studies of complexation of transition-metal cations and Al(III) by aroylhydrazones derived from nicotinic acid hydrazide, Inorg. Chim. Acta 366 (2011) 98e104. S. Adhikari, N.R. Palepu, D. Sutradhar, S.L. Shepherd, R.M. Phillips, W. Kaminsky, A.K. Chandra, M.R. Kollipara, Neutral and cationic halfsandwich arene ruthenium, Cp*Rh and Cp*Ir oximato and oxime complexes: synthesis, structural, DFT and biological studies, J. Organomet. Chem. 820 (2016) 70e81. M. Arivazhagan, S. Jeyavijayan, J. Geethapriya, Conformational stability, vibrational spectra, molecular structure, NBO and HOMOeLUMO analysis of 5-nitro-2-furaldehyde oxime based on DFT calculations, Spectrochim. Acta A 104 (2013) 14e25. S. Ramalingam, M. Karabacak, S. Periandy, N. Puviarasan, D. Tanuja, Spectroscopic (infrared, Raman, UV and NMR) analysis, Gaussian hybrid computational investigation (MEP maps/HOMO and LUMO) on cyclohexanone oxime, Spectrochim. Acta A 96 (2012) 207e220. G.R. Silva, I. Borges Jr., J.D. Figueroa-Villar, DFT conformational studies of the HI-6 molecule, Int. J. Quant. Chem. 105 (2005) 260e269. K. Malek, H. Kozłowski, L.M. Proniewicz, Interaction of Na(I), Ni(II) and Cu(II) with 2-cyano- 2-(hydroxyimino)acetic acid: spectroscopic and theoretical studies, Polyhedron 24 (2005) 1175e1184. A.T. Castro, J.D. Figueroa-Villar, Molecular structure, conformational analysis and charge distribution of pralidoxime: ab initio and DFT studies, Int. J. Quant. Chem. 89 (2002) 135e146. T. Shia, S. Hea, X. Wanga, Y. Wangb, Z. Tanga, Z. Wen, Electronic structure calculations of copper (II) complexes of [Cu(C10H8N2O4)(H2O)2] and [Cu(C10H8N2O4)(CH3OH)(H2O)], J. Mol. Struc. Theochem. 732 (2005) 201e209. R.R. Pillai, V.V. Menon, Y.S. Mary, S. Armakovic, S.J. Armakovic, C.Y. Panicker, Vibrational spectroscopic investigations, molecular dynamic simulations and molecular docking studies of N’-diphenylmethylidene-5-methyl-1H-pyrazole-3-carbohydrazide, J. Mol. Struct. 1130 (2017) 208e222. M.S. Alam, D.U. Lee, Syntheses, crystal structure, Hirshfeld surfaces, fluorescence properties, and DFT analysis of benzoic acid hydrazone Schiff bases, Spectrochim. Acta A 145 (2015) 563e574. A.K. Singh, R.K. Singh, Synthesis, crystal structure, spectroscopic analysis and computationalstudy of (Z)-1-(2,4-dinitrophenyl)-2-((E)-3-(4methoxyphenyl)-1- (thiophen-2-yl) allylidene) hydrazine by DFT and AIM approach, J. Mol. Struct. 1089 (2015) 191e205. A. Das, H. Agarwala, T. Kundu, P. Ghosh, S. Mondal, S.M. Mobinb, G.K. Lahiri, Electronic structures and selective fluoride sensing features of Os(bpy) 2(HL2) and [{Os(bpy)2}2(m-HL2)]2þ(H3L:5-(1H-benzo[d]imidazole-2yl)-1H-imidazole-4-carboxylic acid), Dalton Trans. 43 (2014) 13932e13947. O. Tamer, D. Avcı, Y. Atalay, Quantum chemical characterization of N-(2hydroxybenzylidene)acetohydrazide (HBAH): a detailed vibrational and NLO analysis, Spectrochim. Acta A 117 (2014) 78e86. L. Liu, M.S. Alam, D.U. Lee, Synthesis, antioxidant activity and fluorescence properties of novel europium complexes with (E)-2- or 4-hydroxy-N'-[(2hydroxynaphthalen-1-yl)methylene]benzohydrazide schiff base, Bull. Kor. Chem. Soc. 33 (10) (2012) 3361e3367. S.M. Soliman, J.H. Albering, Synthesis, X-ray structure and topology (AIM and Hirshfeld) analyses of the new square planar [Ag(pyridine-2-aldoxime)2] ClO4 complex; A comparative study with its nitrato analogue, J. Mol. Struct. 1151 (2018), 204e2017. L.Q. Chai, L.J. Tang, K.Y. Zhang, J.Y. Zhang, H.S. Zhang, Two two-dimensional supramolecular copper(II) and cobalt(III) complexes derived from a new quinazoline-type ligand: syntheses, structures, and spectral, thermal, electrochemical and antimicrobial activity studies, Appl. Organomet. Chem. 31 (2017) e3786, 1e13.
138
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139
[39] S.G. Patra, N. Mandal, A. Datta, D. Datta, On bonding in bis(dimethylglyoximato)nickel(II), Comput. Theor. Chem. 1114 (2017) 118e124. [40] S.G. Patra, N.K. Shee, P. Mitra, M.G.B. Drew, D. Datta, Concomitant polymorphism of an octahedral, homoleptic zinc(II) biscomplex of an N, N,O donor hydrazone, J. Mol. Struct. 1155 (2018) 469e476. dınç, Molecular structures, spectroscopic (FTeIR, NMR, UV) [41] A. Es¸me, S.G. Sag studies, NBO analysis and NLO properties for tautomeric forms of 1,3dimethyl-5-(phenylazo)-6-aminouracil by density functional method, Spectrochim. Acta A. 188 (2018) 443e455. [42] Y. Kaya, C. Icsel, V.T. Yilmaz, O. Buyukgungor, Structural, spectroscopic and quantum chemical studies of acetylhydrazone oxime and its palladium(II) and platinum(II) complexes, J. Mol. Struct. 1095 (2015) 51e60. [43] J.E. Rode, J. Frelek, Circular dichroism spectroscopy and DFT calculations in determining absolute configuration and E/Z isomers of conjugated oximes, Chirality 29 (2017) 653e662. [44] D. Tzeli, P.G. Tsoungas, I.D. Petsalakis, P. Kozielewicz, Intramolecular single H bonding vs bifurcation in tuning the conformation of 2,20 -dihydroxybenzophenone and its derivatives: a DFT insight, Struct. Chem. 28 (2017) 925e943. [45] M. Sheikhi, S. Shahab, L. Filippovich, H. Yahyaei, E. Dikusar, M. Khaleghian, New derivatives of (E,E)-azomethines: design, quantum chemical modeling, spectroscopic (FT-IR, UV/Vis, polarization) studies, synthesis and their applications: experimental and theoretical investigations, J. Mol. Struct. 1152 (2018) 368e385. [46] Y. Kaya, C. Icsel, V.T. Yilmaz, O. Buyukgungor, A combined experimental and theoretical investigation of a new imineoxime and its palladium(II) and platinum(II) complexes: synthesis, structural characterization and spectroscopic properties, Spectrochim. Acta A. 133 (2014) 93e101. lez, C. Kremer, [47] L. Martínez, J.S. Gancheff, F.E. Hahn, R.A. Burrow, R. Gonza R. Chiozzone, Nickel(II) complexes with methyl(2-pyridyl)ketone oxime: synthesis, crystal structures and DFT calculations, Spectrochim. Acta A 105 (2013) 439e445. [48] T. Stepanenko, L. Lapinski, M.J. Nowak, L. Adamowicz, Infrared spectra of ryn and anti isomers of benzaldoxime and pyridine-4-aldoxime: an experimental matrix isolation and theoretical density functional theory study, Vib. Spectrosc. 26 (2001) 65e82. [49] K. Malek, M. Vala, H. Kozlowski, L.M. Proniewicz, Experimental and theoretical NMR study of selected oxocarboxylic acid oximes, Magn. Reson. Chem. 42 (2004) 23e29. [50] B. Golec, Z. Mielke, Formaldoxime dimers: infrared matrix isolation and theoretical study, J. Mol. Struct. 844e845 (2007) 242e249. [51] B. Golec, J. Grzegorzek, Z. Mielke, Complexation of formaldoxime and acetaldoxime with nitrogen, Chem. Phys. 353 (2008) 13e18. [52] T. Irshaidat, Some physical organic aspects of salicylaldehydes oximes,a theoretical study, Tetrahedron Lett. 49 (2008) 631e635. [53] V. Arjunan, C.V. Mythili, K. Mageswari, S. Mohan, Experimental and theoretical investigations of benzamide oxime, Spectrochim. Acta A 79 (2011) 245e253. [54] N.R. Gonewar, V.B. Jadhav, K.D. Jadhav, R.G. Sarawadekar, Theoretical calculations of infrared, NMR and electronic spectra of 2-nitroso-1, naphthol or 1-2 naphthoquinine-2 oxime and comparison with experimental data, Res. Pharm. 2 (2012) 18e25. [55] R. Maheswari, J. Manjula, Vibrational spectroscopic analysis and molecular docking studies of (E)-4-methoxy-N´-(4-methylbenzylidene) benzohydrazide by DFT, J. Mol. Struct. 1115 (2016) 144e155. rez-Alvarez, [56] M.S.C. Henriques, R. DelAmparo, D. Pe B.A. Nogueira, M.C. Rodríguez-Argüelles, J.A. Paix~ ao, Synthesis, structural and spectroscopic studies of 2-oxoacenaphthylen-1(2H)-ylidene nicotinohydrazide, Spectrochim. Acta A 172 (2017) 189e198. [57] A. Arokiasamy, G. Manikandan, V. Thanikachalam, K. Gokula Krishnan, Synthesis, spectroscopic investigation and theoretical studies of 2-((E)-(2-(2cyanoacetyl)hydrazono)methyl)-4-((E)-phenyldiazenyl)phenyl methyl carbonate, J. Mol. Struct. 1134 (2017) 563e575. [58] P. Rawat, R.N. Singh, A. Ranjan, S. Ahmad, R. Saxena, Antimycobacterial, antimicrobial activity, experimental (FT-IR, FT-Raman, NMR, UV-Vis, DSC) and DFT (transition state, chemical reactivity, NBO, NLO) studies on pyrroleisonicotinyl hydrazone, Spectrochim. Acta A 179 (2017) 1e10. [59] I. Warad, M. Abdoh, A.A. Ali, N. Shivalingegowda, K. Kumara, A. Zarrouk, N.K. Lokanath, Synthesis, spectra and X-ray crystallography of dipyridin-2ylmethanone oxime and its CuX2(oxime)2 complexes: thermal, Hirshfeld surface and DFT analysis, J. Mol. Struct. 1154 (2018) 619e625. [60] K. Sayin, D. Karakas¸, Computational investigations of trans platinum(II) oxime complexes used as anticancer drug, Spectrochim. Acta A 188 (2018) 537e546. [61] S.M. Patil, R.S. Vadavi, U. Kendur, G.H. Chimmalagi, G.H. Pujar, S.D. Kulkarni, M. Nethaji, S. Nembenna, S.R. Inamdar, K.B. Gudasi, Synthesis, characterization and photophysical studies of zinc(II) complexes derived from a hydralazine hydrazone, J. Photochem. Photobiol. A 351 (2018) 225e230. [62] S. Naseem, M. Khalid, M.N. Tahir, M.A. Halim, A.A.C. Braga, M.M. Naseer, Z. Shafiq, Synthesis, structural, DFT studies, docking and antibacterial activity of a xanthene based hydrazone ligand, J. Mol. Struct. 1143 (2017) 235e244. [63] I. Loncaric, J. Popovic, V. Despoja, S. Burazer, I. Grgicevic, D. Popovic, Z. Skoko, Reversible thermosalient effect of N0 -2-Propylidene-4hydroxybenzohydrazide accompanied by an immense negative compressibility: structural and theoretical arguments aiming toward the elucidation
of jumping phenomenon, Cryst. Growth Des. 17 (2017) 4445e4453. [64] A. El-Asmy, B. Jeragh, M. Ali, Spectral, thermal, molecular modeling and biological studies on Mono and binuclear complexes derived from oxalo bis(2,3-butanedionehydrazone), Chem. Cent. J. 9 (69) (2015) 2e12. [65] F. Guntepe, M. Cinarli, C. Kazak, H. Bati, Spectroscopic and quantum chemical studies of (Z)-N_-(3-(hidroksiimino)butan-2-ylidene)-4- metilbenzensulfonohidrazide ligand, Mol. Cryst. Liq. Cryst. 616 (2015) 213e225. [66] P. Kurnaz, C. Yuksektepe Ataol, H. Bati, O. Buyukgungor, [3-(hydroxyimino) butan-2-ylidene]furan-2’-carbohydrazide, Mol. Cryst. Liq. Cryst. 634 (2016) 61e72. [67] N.M.H. Salem, L. El Sayed, W. Haase, M.F. Iskander, Metal complexes derived from hydrazoneoxime ligands: V. Spectral and structural studies on diacetylmonoxime n-alkanoylhydrazones and their nickel(II) and copper(II) complexes, Spectrochim. Acta A 134 (2015) 257e266. [68] V. Kamat, K. Naik, V.K. Revankar, Effect of acetate and nitrate anions on the molecular structure of 3-(hydroxyimino)-2-butanone-2-(1H-benzimidazol2-yl)hydrazone, J. Mol. Struct. 1113 (2017) 546e556. [69] M.M. Mostafa, M.A. Khattab, K.M. Ibrahim, Synthesis and spectroscopic studies of some metal complexes derived from hydrazone oximes, Transit. Met. Chem. 8 (1983) 282e285. [70] A.M. Shallaby, M.M. Mostafa, K.M. Ibrahım, M.N.H. Moussa, New uranyl(VI) complexes with hydrazone-oximes derived from aromatic acid hydrazides and biacetylmonoxime, Spectrochim. Acta A 40 (10) (1984) 999e1002. [71] Aicha Yacouta-Nour, Abobaker K.T. Maki, Mohsen M. Mostafa, K.M. Ibrahim, A.A. El-Bindary, Dioxouranium(VI) complexes hydrazone-oximes, Transit. Met. Chem. 16 (1991) 23e27. [72] M. Sutradhar, T.R. Barman, S. Ghosha, M.G.B. Drew, Synthesis and characterization of mixed-ligand complexes using a precursor mononuclear oxidovanadium(V) complex derived from a tridentate salicylhydrazone oxime ligand, J. Mol. Struct. 1037 (2013) 276e282. [73] N.M.H. Salem, L. El-Sayed, S. Foro, W. Haase, M.F. Iskander, Metal complexes derived from hydrazoneoxime ligands: III e synthesis, characterization and electrospray ionization mass spectra of some nickel(II) complexes with aroylhydrazoneoximes, Polyhedron 26 (2007) 4161e4172. [74] M.F. Iskander, L. El Sayed, N.M.H. Salem, R. Werner, W. Haase, Metal complexes derived from hydrazoneoxime ligands: I. Synthesis, characterization and magnetochemical studıes of (acylhydrazoneoxime) copper(II) complexes, Coord. Chem. 56 (12) (2003) 1075e1084. [75] M.M. Mostafa, M.A. Khattab, K.M. Ibrahim, Metal complexes of Schiff's Base derived from Salicoylhydrazine and biacetylmonoxime, Polyhedron 2 (7) (1983) 583e585. [76] S. Ghosh, T.K. Bandyopadhyay, The ligating behaviour of 3-Hydroxyimino-2butanone-1-benzoylhydrazone towards Ni II and Cu H ions, Transit. Met. Chem. 10 (1985) 57e60. [77] N.M.H. Salem, L. El Sayed, W. Haase, M.F. Iskander, Metal complexes derived from hydrazoneoxime ligands: II. Synthesis, electrospray mass spectra and magnetochemical studies of some copper(II) complexes derived from psubstituted aroylhydrazoneoximes, J. Coord. Chem. 58 (15) (2005) 1327e1343. [78] M.W. Drover, S.S. Tandon, M.U. Anwar, K.V. Shuvaev, L.N. Dawe, J.L. Collins, L.K. Thompson, Polynuclear complexes of a series of hydrazone and hydrazoneeoxime ligands e M2 (Fe), M4 (Mn, Ni, Cu), and Mn (Cu) examples, Polyhedron 68 (2014) 94e102. [79] M.M. Al-Ne’aimi, M.M. Al-Khuder, Synthesis, characterization and extraction studies of some metal (II)complexes containing (hydrazoneoxime and bisacylhydrazone) moieties, Spectrochim. Acta A 105 (2013) 365e373. [80] M.U. Anwar, L.N. Dawe, M.S. Alam, L.K. Thompson, m-O bridged Mn10 assemblies with open O6 sites for binding extra guests: structural, magnetic, and surface studies, Inorg. Chem. 51 (21) (2012) 11241e11250. [81] S. Naskar, S. Naskar, S. Mondal, P.K. Majhi, M.G.B. Drew, S.K. Chattopadhyay, Synthesis and spectroscopic properties of cobalt(III) complexes of some aroyl hydrazones: X-ray crystal structures of one cobalt(III) complex and two aroyl hydrazone ligands, Inorg. Chim. Acta. 371 (2011) 100e106. [82] M.U. Anwar, A.S. Elliott, L.K. Thompson, L.N. Dawe, Complexes of functionally modified hydrazone ligands-square, rectangular (M4), trigonal bipyramidal (M5) and pin-wheel (M6) motifs, Dalton Trans. 40 (2011) 4623e4635. [83] J.P. Naskar, B. Guhathakurta, P. Basu, N. Bandyopadhyay, G.S. Kumar, M. Zhu, Synthesis, characterization, structure and DNA binding aspects of a trinuclear copper(II) complex having a Cu3O core, Inorg. Chim. Acta. 462 (2017) 158e166. [84] J.J. Norman, R.M. Heggie, J.B. Larose, Oximes: I. The synthesis of some substituted 2-oximinoacetophenones, Can. J. Chem. 40 (8) (1962) 1547e1553. [85] Stoe & Cie, X-AREA and X-RED32. Stoe & Cie, Darmstadt, Germany, 2002. [86] G.M. Sheldrick, A short history of SHELX, Acta cryst. A64 (2008) 112e122. [87] L.J. Farrugia, ORTEP-3 for Windows - a version of ORTEP-III with a graphical user interface (GUI), J. Appl. Crystallogr. 30 (1997) 565. [88] L.J. Farrugia, WinGX suite for small-molecule single-crystal crystallography, J. Appl. Crystallogr. 32 (1999) 837e838. [89] A.L. Spek, Structure validation in chemical crystallography, Acta Crystallogr. D65 (2009) 148e155. [90] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, et al., Gaussian 09, Revision C.01, Gaussian, Inc., Wallingford, CT, 2009. [91] W. Kohn, A.D. Becke, R.G. Parr, Density functional theory of electronic
A. Zülfikaroglu et al. / Journal of Molecular Structure 1162 (2018) 125e139 structure, J. Phys. Chem. 100 (1996) 12974e12980. [92] A.D. Becke, Density-functional thermochemistry. III. The role of exact Exchange, J. Chem. Phys. 98 (7) (1993) 5648e5652. [93] A.D. Becke, Density-functional exchange-energy approximation with correct asymptotic behavior, Phys. Rev. A 38 (1988), 3098e3010. [94] N. Sundaraganesan, S. Ilakiamani, H. Saleema, P.M. Wojciechowski, D. Michalska, FT-Raman and FT-IR spectra, vibrational assignments and density functional studies of 5-bromo-2-nitropyridine, Spectrochim. Acta A 61 (2005) 2995e3001. [95] J.P. Merrick, D. Moran, L. Radom, An evaluation of harmonic vibrational frequency scale factors, J. Phys. Chem. A 111 (2007) 11683e11700. [96] R. Dennington, T. Keith, J. Millam, GaussView, Semichem Inc., Shawnee Mission, KS, 2009, Version 5. . [97] R.G. Pearson, Absolute electronegativity and hardness: applications to organic chemistry, J. Org. Chem. 54 (1989) 1423e1430. [98] R.G. Parr, R.G. Pearson, Absolute hardness: companion parameter to absolute electronegativity, J. Am. Chem. Soc. 105 (1983) 7512e7516. [99] P. Geerlings, F. De Proft, W. Langenaeker, Conceptual density functional theory, Chem. Rev. 103 (2003) 1793e1874. ly, S. Liu, Electrophilicity index, J. Am. Chem. Soc. 121 [100] R.G. Parr, L.V. Szentpa (1999) 1922e1924. [101] P.K. Chattaraj, S. Giri, Stability, reactivity, and aromaticity of compounds of a Multivalent superatom, J. Phys. Chem. A 111 (2007) 11116e11121. [102] N. Okabe, T. Nakamura, H. Fukuda, Structure of 2,4-dinitrophenylhydrazine, Acta Crystallogr. C49 (1993) 1678e1680. [103] L.H. Jensen, E.C. Lingafelter, Refinement of the structure of n-nonanoic acid hydrazide, Acta Crystallogr. 14 (1961) 507e520. [104] Q.C. Ton, M. Bolte, E. Egert, Structural similarities among eight benzoylhydrazones, Acta Crystallogr. C70 (2014) 912e919. lu, Ç. Yüksektepe, H. Bati, N. Çalıs¸kan, O. Büyükgüngo €r, Crystal [105] A. Zülfikarog structure and properties of (Z)-N’-((E)-2-(Hydroxyimino)-1Phenylethylidene) isonicotinohydrazide, J. Struct. Chem. 50 (2009) 1166e1170. [106] M. Kuriakose, M.R.P. Kurup, E. Suresh, Six coordinate Ni(II) complexes of ONN donor aroylhydrazone ligands: synthesis, spectral studies, and crystal structures, Struct. Chem. 18 (2007) 579e584. [107] S. Naskar, S. Naskar, S. Mondal, P.K. Majhi, M.G.B. Drew, S.K. Chattopadhyay, Synthesis and spectroscopic properties of cobalt(III) complexes of some aroyl hydrazones: X-ray crystal structures of one cobalt(III) complex and two aroyl hydrazone ligands, Inorg. Chim. Acta. 371 (2011) 100e106. € Andaç, H. Bati, N. Çaliskan, 3-[(2,4-Di[108] H. Saraçoglu, C. Davran, S. Soylu, O. nitro-phenyl)-hydrazono]-butan-2-one oxime, Acta Crystallogr. E60 (2004) o1307eo1309. [109] M. Mishra, K. Tiwari, S. Shukla, R. Mishra, V.P. Singh, Synthesis, structural investigation, DNA and protein binding study of some 3d-metal complexes with N´-(phenyl-pyridin-2-yl-methylene)- thiophene-2-carboxylic acid hydrazide, Spectrochim. Acta A 132 (2014) 452e464. [110] P. Sathyadevi, P. Krishnamoorthy, M. Alagesan, K. Thanigaimani, P.T. Muthiah, N. Dharmaraj, Synthesis, crystal structure, electrochemistry and studies on protein binding, antioxidant and biocidal activities of Ni(II) and Co(II) hydrazone complexes, Polyhedron 31 (2012) 294e306. € ar, H. Eserci, Crystal structure, spectroscopic [111] G. Ozdemir Tarı, Ü. Ceylan, E. Ag investigations and quantum chemical computational study of 5-(diethylamino)-2-((3-nitrophenylimino)methyl)phenol 1126 (2016) 83e93. [112] S. Rajkhowa, R.C. Deka, DFT based QSAR/QSPR Models in the development of novel anti-tuberculosis drugs targeting Mycobacterium tuberculosis, Curr. Pharmaceut. Des. 20 (2014) 4325e4343. € Tamer, Y. Atalay, The [113] N. Dege, N. S¸enyüz, H. Batı, N. Günay, D. Avcı, O. synthesis, characterization and theoretical study on nicotinic acid [1-(2,3dihydroxyphenyl)methylidene]hydrazide, Spectrochım. Acta A 120 (2014) 323e331. [114] V. Deval, A. Kumara, V. Gupta, A. Sharma, A. Gupta, P. Tandon, K.K. Kunimotoc, Molecular structure (monomeric and dimeric) and hydrogen bonds in 5- Benzyl 2- Thiohydantoin studied by FT-IR and FT-Raman spectroscopy and DFT calculations, Spectrochim Acta A 132 (2014) 15e26. [115] K.O. Ogunniran, M.A. Mesubi, K.V.S.N. Raju, T. Narender, Structural and in vitro anti-tubercular activity study of (E)-N’-(2,6-dihydroxybenzylidene) nicotinohydrazide and some transition metal complexes, J. Iran. Chem. Soc. 12 (2015) 815e829. [116] A. Suvitha, S. Periandy, S. Boomadevi, M. Govindarajan, Vibrational frequency analysis, FT-IR, FT-Raman, ab initio, HF and DFT studies, NBO, HOMOLUMO and electronic structure calculations on pycolinaldehyde oxime,
139
Spectrochim Acta A 117 (2014) 216e224. [117] C. James, A.A. Raj, R. Reghunanthan, V.S. Jayakumar, I.H. Joe, Structural conformation and vibrational spectroscopic studies of 2,6-bis(p-N,Ndimethyl benzylidene)cyclohexanone using density functional theory, J. Raman Spectrosc. 37 (2006) 1381e1392. lu, N. Çalıs¸kan, I. Yilmaz, A. Cukurovalı, Molecular [118] Ç. Yüksektepe, H. Saraçog and crystal structure, SpectroscopicProperties of N-[4-(3-Methyl-3-PhenylCyclobutyl)-Thiazol-2-yl]-N0 -(1H-Pyrrol-2-Ylmethylene)-Hydrazine by experimental method and QuantumChemical calculation, Mol. Cryst. Liq. Cryst. 533 (2010) 126e140. lu, N. Çalıs¸kan, Ç. Yüksektepe, A. Çukurovalı, Synthesis, [119] F. Güntepe, H. Saraçog molecular and crystal structure analysis of2-bromo-4-chloro-6-{[4-(3methyl-3-phenyl-cyclobutyl)-thiazol-2-yl]-hydrazonomethyl}-phenol by experimental methods and theoretical calculations, J. Mol. Struct. 1022 (2012) 204e210. [120] E.R. Alan, A.C. Larry, W. Frank, Intermolecular interaction from a natural bond orbital donor-acceptor viewpoint, Chem. Rev. 88 (6) (1988) 899e926. [121] J.N. Liu, Z.R. Chen, S.F. Yuan, Study on the prediction of visible absorption maxima of azobenzene compounds, J. Zhejiang Univ. - Sci. B 6 (6) (2005) 584e589. [122] J.E. Carpenter, F. Weinhold, Analysis of the geometry of the hydroxymethyl radical by the “different hybrids for different spins natural bond orbıtal procedure, J. Mol. Struc. Theochem. 169 (1988) 41e62. [123] L. Xiao-Hong, C. Qing-Dong, Z. Xian-Zhou, Natural bond orbital population analysis of para-substituted O-nitrosyl carboxylate compounds, Struct. Chem. 20 (2009) 1043e1048. [124] J. Chochollousova, V.V. Spirko, P. Hobza, First local minimum of the formic acid dimer exhibits simultaneously red-shifted O-H…O and improper blueshifted C-H…O hydrogen bonds, Phys. Chem. Chem. Phys. 6 (2004) 37e41. [125] S. Sebastian, N. Sundaraganesan, The spectroscopic (FT-IR, FT-IR gas phase, FT-Raman and UV) and NBO analysis of 4-Hydroxypiperidine by density functional method, Spectrochim Acta A A75 (2010) 941e952. [126] A.E. Reed, F. Weinhold, Natural localized molecular orbitals, J. Chem. Phys. 83 (1985) 1736e1740. [127] S. Gunasekaran, S. Kumaresan, R. Arunbalaji, G. Anand, S. Srinivasan, Density functional theory study of vibrational spectra, and assignment of fundamental modes of dacarbazine, J. Chem. Sci. 120 (3) (2008) 315e324. [128] I. Sıdır, Y.G. Sıdır, M. Kumalar, E. Tas¸al, Ab initio HartreeeFock and density functional theory investigations on the conformational stability, molecular structure and vibrational spectra of 7-acetoxy-6-(2,3-dibromopropyl)-4,8dimethylcoumarin molecule, J. Mol. Struct. 964 (2010) 134e151. [129] A.E. Reed, R.B. Weinstock, F. Weinhold, Natural population analysis, J. Chem. Phys. 83 (1985) 735e746. [130] P. Politzer, P.R. Laurence, K. Jayasuriya, Molecular electrostatic potentials: an effective tool for the elucidation of biochemical phenomena, Environ. Health Perspect. 61 (1985) 191e202. [131] N. Okulik, A.H. Jubert, Theoretical analysis of the reactive sites of nonesteroidal antieinflammatory drugs, internet electron, J. Mol. Des. 4 (2005) 17e30. [132] P. Politzer, P. Lane, A computational study of some nitrofluoromethanes, Struct. Chem. 1 (1990) 159e164. [133] P. Politzer, M.C. Concha, J.S. Murray, Density functional study of dimers of dimethylnitramine, Int. J. Quant. Chem. 80 (2000) 184e192. [134] W. Hussein, C.G. Walker, Z. Peralta-Inga, J.S. Murray, Computed electrostatic potentials and average local ionization energies on the molecular surfaces of some tetracyclines, Int. J. Quant. Chem. 82 (2001) 160e169. ar, M. Yavuz, Combined experimental and computational [135] H. Tanak, A. Ag modeling studies on 4-[(2-hydroxy-3 methylbenzylidene) amino]-1,5dimethyl-2-phenyl- 1,2-dihydro-3H-pyrazol-3-one, Int. J. Quant. Chem. 111 (2011) 2123e2136. [136] F. Boukli-Hacene, M. Merad, S. Ghalem, W. Soufi, DFT study of the interaction of Cu(II), Zn(II), Sn(II) with carbohydrates in aqueous solution, J. Chem. Chem. Eng. 8 (2014) 1009e1017. [137] T.A. Yousef, G.M. Abu El-Reash, R.M. El Morshedy, Quantum chemical calculations, experimental investigations and DNA studies on (E)-2-((3hydroxynaphthalen-2-yl)methylene)-N-(pyridin-2-yl)hydrazinecarbothioamide and its Mn(II), Ni(II), Cu(II), Zn(II) and Cd(II) complexes, Polyhedron 45 (2012) 71e85. [138] D. Singh, S. Ahmad, P.P. Singh, DFT based calculation of interaction energy between metal halides and organic bases, J. Mol. Struc. Theochem. 905 (2009) 13e23.