Next Article in Journal
Thermal Performance of Alginate Concrete Reinforced with Basalt Fiber
Next Article in Special Issue
Dinuclear Silver(I) Nitrate Complexes with Bridging Bisphosphinomethanes: Argentophilicity and Luminescence
Previous Article in Journal
Effect of Annealing Process on Microstructure, Texture, and Mechanical Properties of a Fe-Si-Cr-Mo-C Deep Drawing Dual-Phase Steel
Previous Article in Special Issue
Non-Covalent Interactions in Coordination and Organometallic Chemistry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Green Synthesis, SC-XRD, Non-Covalent Interactive Potential and Electronic Communication via DFT Exploration of Pyridine-Based Hydrazone

1
Department of Chemistry, Chemical Engineering and Biotechnology, Donghua University, Shanghai 201620, China
2
Department of Chemistry, University of Malakand, Chakdara, Lower Dir, Khyber Pakhtunkhwa 18800, Pakistan
3
Department of Chemistry, University of Sargodha, Sargodha 40100, Pakistan
4
Department of Chemistry, Khwaja Fareed University of Engineering & Information Technology, Rahim Yar Khan 64200, Pakistan
5
Department of Physics, University of Sargodha, Sargodha, Punjab 40100, Pakistan
6
Department of Chemistry, University of Agriculture, Faisalabad 38000, Pakistan
7
Med-X Research Institute, School of Biomedical Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
*
Authors to whom correspondence should be addressed.
Both authors contributed equally to this work.
Crystals 2020, 10(9), 778; https://doi.org/10.3390/cryst10090778
Submission received: 19 July 2020 / Revised: 17 August 2020 / Accepted: 28 August 2020 / Published: 2 September 2020

Abstract

:
Ultrasound-based synthesis at room temperature produces valuable compounds greener and safer than most other methods. This study presents the sonochemical fabrication and characterization of a pyridine-based halogenated hydrazone, (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene) acetohydrazide (HBPAH). The NMR spectroscopic technique was used to determine the structure, while SC-XRD confirmed its crystalline nature. Our structural studies revealed that strong, inter-molecular attractive forces stabilize this crystalline organic compound. Moreover, the compound was optimized at the B3LYP/6-311G(d,p) level using the Crystallographic Information File (CIF). Natural bonding orbital (NBO) and natural population analysis (NPA) were performed at the same level using optimized geometry. Time-dependent density functional theory (DFT) was performed at the B3LYP/6-311G (d,p) method to calculate the frontier molecular orbitals (FMOs) and molecular electrostatic potential (MEP). The global reactivity descriptors were determined using HOMO and LUMO energy gaps. Theoretical calculations based on the Quantum Theory of Atoms in Molecules (QT-AIM) and Hirshfeld analyses identified the non-covalent and covalent interactions of the HBPAH compound. Consequently, QT-AIM and Hirshfeld analyses agree with experimental results.

1. Introduction

Humanity faces increasing health, shelter, and economic problems as we consume more resources to pollute, urbanize, and deforest our environment. Fatal diseases have not only taken many lives but also severely harmed the global economy. To combat global pandemics and other diseases, synthetic organic chemists need to synthesize novel and potent chemicals by safe and green methods. One such chemical, hydrazones, plays a substantial role in the bio-medicinal applications due to its versatility [1,2,3]. It has many effortlessly reachable binding sites for the medicinal applications [4], such as antimicrobial [5], cardioprotective [6], anti-HIV [7], anti-inflammatory [8], anticancer [9], antihypertensive [10], antitubercular [11], antimalarial [12], antidepressant [13], antioxidant [14], and anticonvulsant [15]. For example, pyridine-based hydrazone derivatives ubiquitously displayed antifungal properties [16,17,18]. Acylpyridine derivative, 2-benzoxazolylhydrazon, suppresses leukemia, colon and ovarian cancer cell lines [19]. Acylhydrazone introduction in 1, 2, 4-triazolo [4, 3-a] pyridine derivatives by a microwave-assisted method leads to herbicidal and pesticidal lead compounds [1]. Hydrazone derivatives also possess unique physical and chemical properties including fluorescence emission [20], corrosion inhibitory properties and passivation [21], and iron chelation in iron toxicity [22]. In addition, lone electron pairs and pi-electrons play a key role in medicinal applications due to their ability of non-covalent interaction, such as van der Waals interactions, hydrophobic bonds, ionic bonds, and hydrogen bonds [23,24,25,26]. In particular, these non-covalent interactions facilitate crystals packing, proton transfer reactions, the stability of molecules, enzymatic catalysis [25,27,28,29]. Several molecules with hydrogen bonding capacity are important for catalysis in organic transformation such as diols, bisphenols, hydroxy acids, urea, guanidinium and amidinium ions, thioureas, lactams, thioureas, cinchona alkaloids, and phosphoric acids [30,31,32,33,34]. Amongst them, hydrazones have a unique chemical architecture (Figure 1), allowing its significant ability to form non-covalent interactions [35].
Studies show that microwave-assisted synthesis accelerates chemical synthesis with a better yield and higher purity in comparison to conventional methods [18,36,37]. Microwave (MW) radiations assist in non-thermal polarizing radiation, dipolar polarization, ionic conduction reactions [38]., This study reports the ultrasound-based synthesis, SC-XRD exploration, and density functional theory (DFT) analysis of the pyridine-based novel crystalline hydrazones, i.e., (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene) acetohydrazide.

2. Materials and Methods

2.1. General

Analytical grade solvents and pure reagents were used without any further purification. TLC (Thin layer chromatography) cards, coated with silica gel (0.25 mm thickness), were used to monitor the reaction progress. For the NMR spectra measurement, Bruker-Avance, A-V spectrometer, was used. For the single crystal analysis, Bruker Kappa APEX-II diffractometer was used where the data correction and data reduction were made by APEX-II and SAINT, respectively [39]. For the structure solution, SHELXS97 software [40,41] and for refinement, SHELXL2014/6 was used to minimize the structural errors [42]. For the graphical representation of the asymmetric unit, ORTEP was used while for the hydrogen bonding, PLATON was used [43].

2.2. Synthesis of 2-(6′-Chloroazin-2′-yl) oxy-aceto-hydrazide (A)

The precursor A was manufactured according to the procedure described elsewhere [35,44]. Accordingly, a mixture of ethyl 2-(6′-chloroazin-2-yl)-ox-ethanoate (131 mg, 0.61 mmol) and N2H4.H2O (0.09 mL, 1.83 mmol) in ethanol was refluxed for 3 h. The reaction on completion (monitored by TLC) was cooled to room temperature and concentrated under reduced pressure. The targeted hydrazide was purified by column chromatography yielding 89 mg of the isolated A (73%).

2.3. General Procedure for the Synthesis of (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene)acetohydrazide

A mixture of A (2-(6′-chloroazin-2′-yl) oxy-aceto-hydrazide) (0.48 mmol) and Salicylaldehyde (0.54 mmol) was dissolved in ethanol separately to make clear solutions. The solutions were mixed at room temperature, and the mixture was sonicated for 5 to 10 min. The targeted compound was precipitated that was filtered through standard filtration and recrystallized in ethanol (Scheme 1).
1H NMR (400 MHz, DMSO) δ 11.84 (s, 1H), 10.99 (s, 1H), 8.48 (s, 1H), 7.81 (dt, J = 15.3, 7.8 Hz, 2H), 7.72–7.67 (m, 1H), 7.32–7.21 (m, 2H), 7.15 (d, J = 7.5 Hz, 1H), 7.00–6.94 (m, 1H), 5.31 (s, 2H). 13C NMR (100 MHz, DMSO) δ 168.2, 163.8, 162.4, 156.4, 147.6, 142.1, 141.3, 131.2, 126.2, 119.3, 116.8, 116.1, 109.6, 62.9 (Figure S1).

2.4. Computational Studies

The simulation study for the entitled compound, (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene) acetohydrazide (HBPAH), was performed through DFT [45,46,47] employing Gaussian 09 program package [48]. By the use of GaussView 5.0. [49] all input files were organized. Finally, Chem craft [50], Avogadro [51] and Gauss Sum [52], AIM-All Professional [53], and Crystal Explorer [54] programs were used for the interpretation of output files. The Structure of HBPAH was optimized using SC-XRD-based geometry at the B3LYP/6-311G(d,p) level. The natural bond orbital (NBO) [55,56] and natural population analysis (NPA) were analyzed similarly, while the frontier molecular orbital (FMO) and molecular electrostatic potential (MEP) were calculated by the TD-DFT/B3LYP /6-311G(d,p) level [57,58]. Moreover, the Hirshfeld surface (HS) analysis [59,60] was carried out to determine the non-covalent interactions. The Quantum Theory of Atoms in Molecules (QT-AIM) [61] analysis was employed to explore the non-covalent interactions. The electron affinity (EA), electronegativity (X) [62], global electrophilicity index (ω) [62,63,64], ionization potential (IP) [65], global hardness (η) [66,67], global softness (S) [68] and chemical potential (μ) [69] were known as global reactivity parameters, and their values can be calculated by HOMO-LUMO energies. These parameters were also reported as biological activity descriptors having numerous optoelectronic applications and are helpful in determining stability, reactivity, and selectivity of the molecules [62,70,71,72]. They were calculated through Equations (1)–(7):
I P = E HOMO
E A = E LUMO
where IP = ionization potential (eV), EA = electron affinity (eV).
Koopmans’s theorem [73] was usually used to calculate the chemical potential (μ), electronegativity (x) and chemical hardness (η) and was equated as:
μ = E HOMO + E LUMO 2
x = [ IP + EA ] 2 = [ E LUMO + E HOMO ] 2
η = [ IP EA ] 2 = [ E LUMO E HOMO ] 2
The following equation was used for global softness (σ):
σ = 1 2 η
The calculation of electrophilicity index (ω) was reported by Parr et al. as:
ω = μ 2 2 η

3. Results and Discussion

The hydrazone, (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene) acetohydrazide (HBPAH), was synthesized with a yield of 85% and its structures were determined by NMR spectroscopy and SC-XRD analysis. The 1H- and 13C-NMR of the title compound showed the presence of each signal in duplication that indicates that the title compound exists in two isomeric forms; a minor isomer A (E) that is 45.87% and a major isomer B (Z) that is 54.12% (Scheme 2). The ratio of the E and Z isomers was calculated from the 1H-NMR analysis, where the methylenic signals of both isomers were integrated into the 1H NMR spectra (Figure S2 in Supplementary Materials).
The DFT calculation of HBPAH was performed by DFT/B3LYP/6-311G (d, p). Table 1 shows the single-crystal analysis details, Hirshfeld surface, and computational details.
HBPAH (Table 1, Figure 2) crystals contain two crystallographically independent molecules of (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene)acetohydrazide and one water molecule. In the first molecule (C1-C14/N1-N3/O1-O3/CL1) (red in overlay plot), the 6-chloropyridin-2-ol moiety A (C1-C5/N1/CL1), acetohydrazide group B (C6/C7/N2/N3/O2) and O-cresol moiety C (C8-C14/O3) are planar with an r.m.s deviation of 0.0058, 0.0067 and 0.0133 Å, respectively, whereas, in the second molecule (C15-C28/N4-N6/O4-O6/CL2) (blue in overlay plot), the similar moieties D (C15-C19/N4/O4/CL2), E (C20/C21/N5/N6/O5) an F (C22-C28/O6) are planar with r.m.s deviation of 0.0065, 0.0034, and 0.0108 Å, respectively. The dihedral angles between moieties in the first molecule A/B, A/C, and B/C are 14.15 (1)°, 9.82 (1)°, and 23.13 (1)°, respectively whereas the dihedral angle between similar moieties in second molecule D/E, D/F, and E/F is 12.37 (1)°, 11.87 (1)° and 3.7 (1)°, respectively. The two crystallographic independent molecules differ in terms of geometric parameters, as shown in Figure 3. The second molecule is inverted and then made to overlap with the first molecule. This analysis shows that the root mean square deviation between the first molecule and the second molecule is 0.2376 Å.
In both molecules within the lattice, the NH of acetohydrazide group interacts with the O-atom of 6-chloropyridin-2-ol moiety through intra N-H⋯O bonding to form S(5) loop, and the hydroxyl group of o-cresol moiety interacts with N-atom of acetohydrazide group through intra O-H⋯N bonding to form S(6) loop. The first molecule connects with the second molecule through N-H⋯O bonding, where NH is from acetohydrazide group E, and O-atom is from the acetohydrazide group B. Water molecule is engaged in two types of classical H-bonding named as O-H⋯O and N-H⋯O. Water acts as a donor in O-H⋯O (carbonyl O-atom of acetohydrazide group B) and O-H⋯O (carbonyl O-atom of acetohydrazide group E) to connect molecule of the first type with a molecule of the second type whereas it acts as an acceptor in N-H⋯O bonding where NH is from acetohydrazide group E. Water molecule is also engaged in one weak non-classical C-H⋯O (CH is from O-cresol moiety C) bonding with C-O distance of 3.271 Å and angle of 139.06° [74,75]. R 2 1 ( 6 ) loop is formed through classical N-H⋯O and non-classical C-H⋯O bonding in which water acts as an acceptor. The carbonyl O-atom of acetohydrazide group E is also engaged in weak non-classical C-H⋯O(CH is from o-cresol moiety F) bonding to connect molecules of the second type with each other with a C-O distance of 3.433 Å and angle of 162.40° [76].
All the above-mentioned loops and H-bonding are shown in Figure 4, Figure S3 and Table 2. Both molecules and water are connected to form an infinite 2-D network in the crystallographic plane (0 0 1) with base vector (1 0 0) and (0 1 0). Along with the intra and intermolecular H-bonding, a cyclic face-to-face stacking between different rings assists in further strengthening crystal packing. The pyridine ring (C1-C5/N1) at the asymmetric position stacks with two symmetry mates’ phenyl rings (C23-C28) located at (x, 1 + y, z) and (1 + x, 1 + y, 1 + z) with inter-centroid separation of 3.671 Å and 3.810 Å as displayed in Figure 5 and Table 3. Similarly, the phenyl ring (C9-C14) at the asymmetric position stacks with two symmetry-related pyridine rings (C15-C19/N4) located at (1 + x, 1 + y, 1 + z) and (1 + x, 1 + y, z) with inter-centroid separation of 3.874 Å and 3.876 Å, respectively. Cg(1), Cg(2), Cg(3), and Cg(4) are the centroids of pyridine ring (C1-C5/N1), phenyl ring (C9-C14), pyridine ring (C15-C19/N4), phenyl ring (C23-C28), respectively. Dde, DAde, De (f) and Df (e), respectively, show the distance between centroids of rings, the dihedral angle between the planes of rings, perpendicular distance of Cg(e) to Cg(f), perpendicular distance of Cg(f) to Cg(e).

3.1. Comparative Structural Study

The SC-XRD-based structure of HBPAH was used for geometry optimization in bond length and bond angle calculations. For HBPAH, an atom numbering scheme was presented in Figure S4 (Supplementary Information), and the aforementioned geometrical parameter results were shown in Table S1 (Supplementary Information). DFT-calculated and SC-XRD-driven parameters agree with each other with an overall variation of 0.039 ± 0.028 Å. Similarly, bond angles in HBPAH deviate around 3.0 ± 3.3°.

3.1.1. Hirshfeld Surface Analysis

The crystal structure of HBPAH contains many N-H/O, C-H/O, O-H/O, C-O⋯π, C-H⋯π, and π⋯π interactions. The HS analysis calculates the percentage of significant non-covalent interactions contributions [77,78,79,80,81,82]. The HS mapped with properties like dnorm, de, di, shape index, curvedness, and the 2D fingerprint plots of HBPAH are shown in Figure 5, Figures S5 and S6 (Supplementary Information). Red and white in the HS analysis represent the strongest and intermediate interactions, whereas blue illustrates weaker intermolecular interactions. As de and di are external and internal distance from a surface to the nearest nuclei, respectively, dnorm can be defined by Equation (8) [83]:
d n o r m =   d i   r i v d w r i v d w +   d e +   r e v d w r e v d w
In HBPAH, the dnorm surfaces with dark red spots demonstrate hydrogen bonding interactions [84,85]. The oxygen of –C=O aceto group, the nitrogen of hydrazide -NH, the hydroxyl group of N’-(2-hydroxybenzildene), and other hydrazide nitrogen near the N’-(2-hydroxybenzildene) participate the strong interactions as shown in Figure 6. The HS analysis of HBPAH has mapped the distances dnorm (−0.5807 to 1.0525 a.u.), shape index (−1.000 to 1.000 a.u.), and curvedness (−4.000 to 0.4000 a.u.), as shown in Figure 6.
In the curvedness diagram, the broader green areas separated by blue outlines show the stacking interactions. Figure 6 shows the shape index that explains the π–π stacking interactions with blue humps and red hollows.
We then used two-dimensional fingerprint plots to explain the intermolecular interactions within the molecular structure [86,87,88]. The strongest interaction among hydrogen atoms in the compound is 33.2%, as shown in Figure 7, alongside percentage contribution for all interatomic contacts. Figure S5 shows the two-dimensional fingerprint plots. The most dominant contributions within the crystal packing are as follows: H-H (33.20%), C-H (13.00%), O-H (17.20%), Cl-H (15.60%), C-C (7.50%) and C-N (2.70%). Our HS analysis shows that C-H⋯π interactions dominate the stability within the molecular structure of HBPAH.
Our HS analysis also reports secondary interactions between molecules [78,87,89], such as carbon atom attached with –NH of hydrazide part bonded with the hydrogen atom of the O=C-H group [90]. Figure 8 shows the intermolecular hydrogen bonds (dashed green lines between the hydrazide –NH and the hydroxybenzylide O-H) and intermolecular hydrogen bond with the water molecule (solvent interaction).

3.1.2. QT-AIM Analysis

Next, we used the Quantum Theory of Atoms in Molecules (QT-AIM) [61,91,92] to analyze non-covalent inter and intramolecular interactions, such as hydrogen bonding (HBs) (Table S2, Supplementary Information). The AIM analysis revealed that the crystal is stabilized through intra- and inter-molecular interactions [93,94], as shown by the dashed bond paths (BPs) (Figure 9). We calculated the non-covalent interactions (NCI) by calculating real-space regions where non-covalent interactions are essential and based entirely on ρ and its gradient [94]. HBPAH displayed intermolecular interactions that stabilized the molecules within the crystal. The O-H ρ values at BCPs (Bond critical points), H16-O36, H8-O36, H23-O36 and H19-O37 were +0.0029 e/a3, +0.0127 e/a3, +0.0091 e/a3 and +0.0073 e/a3, respectively. The N-H ρ values at BCPs, H5-N9, and H38-N42 were +0.0436 e/a3 and +0.0421 e/a3, respectively (Table 4). Other intermolecular interactions, O2-O37 and H33-H53 were +0.0066 e/a3 and +0.0016 e/a3.
HBPAH shows two different sets of HBs, intramolecular and intermolecular, with the water molecule (solvent interaction). The intramolecular HB was displayed between oxygen next to pyridine moiety and the hydrazide hydrogen, with the O-H ρ value (O2-H8 = +0.0181 e/a3 and O35-H41 = +0.0179 e/a3). The solvent-based HBs measure weaker than the intramolecular HB with O-H ρ values at BCPs, H49-O67, H56-O67, and H41-O67 were +0.0062 e/a3, +0.0114 e/a3, and +0.0148 e/a3, respectively (Table 4 and Table S2).

3.1.3. Natural Bonding Orbital (NBO) Analysis

We next used NBO analysis to interpret charge transformation, different types of HB (inter- and intra-molecular), and hyper conjugative interactions [95,96,97]. For all orbitals, second-order perturbation energy E(2) could be calculated from Equation (9).
E ( 2 ) = q i ( F i , j ) 2 ε j ε i
qi is donor orbital occupancy, εj and εi are diagonal elements, and F(i,j) is off-diagonal NBO Fock matrix element. For HBPAH, all E(2) values are displayed in Table S3, while the imperative E(2) values are arranged in Table 5.
Among probable electronic π→π* transitions of the highest magnitude, π(C46-C48)→π*(N39-C50) corresponds to stabilization energy of 30.35, kcal/mol in HBPAH. The transitions such as ∂(C13-H14)→∂*(C11-C13) show the lowest stabilization energy of 0.51 kcal/mol for HBPAH, corresponding to weak interactions between the electron donor and acceptor. Other π→π* interactions, such as π(C13-C15)→π*(N6-C17), π(N39-C50)→π*(C43-C44), π(C43-C44)→π*(C45-C48), and π(C46-C48)→π*(N39-C50), yield 29.62, 28.32, 22.45 and 30.35 kcal/mol stabilization energies, respectively (Table 5).
Moreover, the most prominent interactions in LP→π* manifested as LP1(N40)→π*(O36-C54), LP1(N7)→π*(O3-C21), LP2(O35)→π*(N39-C50), and LP2(Cl1)→π*(N6-C10) showed stabilization energies of 62.83, 56.48, 35.80, and 5.79 kcal/mol, respectively (Table 5).
For HBPAH, two additional interactions, i.e., LP1(N40)→π*(O36-C54) and LP1(N7)→π*(O3-C21) with respective high stabilization energy values of 62.83 and 56.48 kcal/mol, indicated the strong HB between lone-pair to anti-bonding orbitals in our HS and QT-AIM analyses. We conclude that these interactions directly stabilize HBPAH in its solid-state.

3.1.4. Natural Population Analysis (NPA)

For HBPAH, the natural population-based analysis on NBO was determined by B3LYP/6-311G(d,p) (Figure S7). The phenomenon correlates to charge transformation, and the electronegativity equalization process occurs in reaction to access the electrostatic ability on the external surfaces of the structure [66,98,99]. The charges of atoms play a crucial role within the molecular conformation and bonding capability in HBPAH. The electronegative atoms such as Cl, O, and N made unequal redistribution of the electron density over the pyridine or aromatic rings. Atomic charge of oxygen atoms was O2 (−0.38655e) and O36 (−0.39665e), and for hydrogen atoms, charges were H8(0.28089e) and H38(0.281296e), respectively, due to the involvement of these atoms in the intermolecular hydrogen bonding interactions.
Furthermore, the NPA of HBPAH showed that carbon atoms, namely C10, C17, C18, C22, C25, C43, C50, C51, C54, C55, and C58, were positively charged, while C11, C13, C15, C24, C25, C28, C30, C44, C46, C48, C57, C59, C61, C63, and C65 were negatively charged (Figure S7). Moreover, all oxygen, nitrogen, and chlorine atoms were negatively charged. All hydrogens are positively charged in HBPAH.

3.1.5. Frontier Molecular Orbital (FMO) Analysis

The FMOs evaluate chemical bond strength and molecule stability [100]. In HBPAH, the energy of HOMO, LUMO, and its two upper and lower orbitals (HOMO-1, HOMO-2, LUMO+1, LUMO+2) were calculated by the TD-DFT/B3LYP/6-311G (d, p) and displayed in Figure 10. The energy difference between HOMO-LUMO is assumed to be a significant key factor to illustrate the chemical reactivity, optical properties, kinetic stability, and electronic character of the compounds [101,102]. Table 6 shows the energy data with their energy gap (∆E) for six MO (molecular orbitals).
Figure 10 shows that HBPAH contained an energy gap of 3.634 eV, which exposed the effective intra-molecular charge transfer (ICT) within the compound. For HBPAH, the HOMO was populated on the first part of molecule (Z)-N′-(2-hydroxybenzylidene) acetohydrazide moiety and a small effect exists on the hydroxyl group. LUMO was populated on the second part of the molecule, i.e., (Z)-N′-(2-hydroxybenzylidene) propionohydrazide moiety (Figure 10). HOMO and LUMO energy gap values for two antifungal 1,2,4-triazolo[4,3-a]pyridine derivatives were found to be 4.318 and 3.705 eV, where high energy gap was associated with the more potent antifungal compound [16].
HBPAH contained an IP value of 5.6 eV and an EA value of 1.966 eV. Its electron loss and the electron gain capacity were defined by the ionization potential and the electron affinity values, which correlate to the HOMO-LUMO energy difference. Consequently, the IP value shows a lower magnitude than the EA value, indicating that HBPAH contained excellent electron-donating capability. This supports the findings of global electrophilicity (ω) (Table S4). In HBPAH, the calculated global softness (σ) values obtained were lower than the global hardness (η) values, making HBPAH stable and relatively unreactive. Additionally, the chemical potential (μ) value (−3.783 (eV)) revealed that HBPAH was chemically hard with the affective electron-donating ability and highest kinetic stability (Table S4).

3.1.6. Molecular Electrostatic Potential (MEP)

The MEP significance shows the size and configuration of the molecule, along with neutral (white), negative (red), and positive (blue) electrostatic potential regions comparable to shading assessing scheme. MEP explores the connection between molecular structural insights and physicochemical properties [103]. We Analyzed HBPAH’s MEP surface through the B3LYP/6-311G(d,p) level of theory, as shown in Figure 11. The negative red indicates the electrophilic sites at the oxygen atoms. Therefore, the oxygen atoms are the most effective target for nucleophilic attack, along with the most suitable sphere to attack the molecules’ positive zones. The negative potential magnitude of HBPAH is −1.00 × 10−2 to 1.00 × 10−2 a.u. Green areas represent the region of zero potential. The blue areas of the HBPAH molecule situate over the hydrogen atoms. They show a combination of positive charges, demonstrating the nucleophilic localities.

4. Conclusions

In conclusion, we used a room-temperature sonochemical approach to synthesize crystalline (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene)acetohydrazide. The SC-XRD study revealed the presence of attractive intermolecular forces for the structural stabilization in this Triclinic crystal system with P 1 ¯   space group. The QT-AIM and Hirshfeld analysis revealed the presence of non-covalent interactions (NCIs); Scheme H5-N9, H38-N42, H16-O36, H8-O36, H19-O37, and H33-H53 that stabilize the structure of the compound. The NBO study showed that HBPAH has molecular stability of hyper-conjugation due to the intramolecular charge transfer (62.83, kcal/mol for LP1(N40) →π*(O36-C54)). The HOMO/LUMO energy band gap value describes the possible charge-transfer interactions, which occur inside the molecule. The calculated FMO energy bandgap of HBPAH is 3.634 eV, which illustrates it has intra-molecular charge-transferability and good NLO properties. The global reactivity descriptors calculation illustrates less reactivity and good stability. The MEP map displayed the negative red areas indicating the electrophilic sites at the oxygen atoms. All computational and experimental findings determined that HBPAH exists in stabilized crystal form because of non-covalent interactions (NCIs) and intra- and inter-molecular H-bonding interactions.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/10/9/778/s1, Figure S1: 1H and 13C NMR spectra of HBPAH, Figure S2: The 1H-NMR of the title compound in (CD3)2SO-d6 showing the integration of methylenic (CH2) 1Hs, which has been used as a tool to calculate the ratio of two isomers (A and B), Figure S3: The geometrical parameters (bond lengths (Å) and bond angles (°) of entitled compound calculated through XRD and at DFT/ B3LYP/6-311G (d,p) level of theory, Figure S4: ORTEP diagram of HBPAH compound, Figure S5: Hirshfeld surfaces of the entitled compound mapped over, de and di for HBPAH (1 a.u. of electron density = 6.748 e.Å−3), Figure S6: 2-D Fingerprint plots for individual contributions in HBPAH, Figure S7: Natural population analysis (NPA) of entitled compound, Table S1: Comparison of XRD and DFT values of bond length (Å) and bond angle (°) of HBPAH, Table S2: AIM properties of HBPAH; Electronic density (ρ), Laplacian of density (∇2ρ), ellipticity (ε) and density of potential energy (V), Table S3: NBO analysis of HBPAH at DFT/ B3LYP/6-311G (d,p) level of theory, Table S4: Ionization potential (IP), electron affinity (EA), electronegativity (X), global hardness (η), chemical potential (μ), global electrophilicity index (ω) and global softness (σ).

Author Contributions

Conceptualization, A.A., C.L., and M.F.u.R.; methodology, A.A., and F.K.; software, S.A., and M.F.u.R.; validation, M.N.T., M.A.; formal analysis, C.L., A.A.; investigation, M.K., S.A.; resources, J.I., and F.K.; data curation, M.K.; writing—original draft preparation, A.A., C.L., and M.F.u.R.; writing—review and editing, F.K. and M.F.u.R.; visualization, M.N.T.; supervision, A.A., C.L.; project administration, A.A. and M.K.; funding acquisition, C.L. All authors have read and agreed to the published version of the manuscript.

Funding

Funding in the Lu lab was provided by the Shanghai Science and Technology Committee (19ZR1471100), Fundamental Research Funds for the Central Universities (19D210501, 19D310517).

Acknowledgments

The authors thank Muhammad Mustaqeem and research students in Muhammad Khalid’s group: Wajeeha Anwer, Aamina Khalid, and Muhammad Zahid for their participation in terms of careful reading.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rollas, S.; Küçükgüzel, S.G. Biological activities of hydrazone derivatives. Molecules 2007, 12, 1910–1939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Verma, G.; Marella, A.; Shaquiquzzaman, M.; Akhtar, M.; Ali, M.R.; Alam, M.M. A review exploring biological activities of hydrazones. J. Pharm. Bioallied Sci. 2014, 6, 69. [Google Scholar] [PubMed]
  3. Hussain, I.; Ali, A. Exploring the Pharmacological Activities of Hydrazone Derivatives: A Review. J. Phytochem. Biochem. 2017, 1, 1–5. [Google Scholar]
  4. Su, X.; Aprahamian, I. Hydrazone-based switches, metallo-assemblies and sensors. Chem. Soc. Rev. 2014, 43, 1963–1981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Rane, R.A.; Telvekar, V.N. Synthesis and evaluation of novel chloropyrrole molecules designed by molecular hybridization of common pharmacophores as potential antimicrobial agents. Bioorg. Med. Chem. Lett. 2010, 20, 5681–5685. [Google Scholar] [CrossRef] [PubMed]
  6. El-Sabbagh, O.; Shabaan, M.A.; Kadry, H.H.; Al-Din, E.S. New octahydroquinazoline derivatives: Synthesis and hypotensive activity. Eur. J. Med. Chem. 2010, 45, 5390–5396. [Google Scholar] [CrossRef] [PubMed]
  7. Ma, X.D.; Yang, S.Q.; Gu, S.X.; He, Q.Q.; Chen, F.E.; De Clercq, E.; Balzarini, J.; Pannecouque, C. Synthesis and anti-HIV activity of Aryl-2-[(4-cyanophenyl) amino]-4-pyrimidinone hydrazones as potent non-nucleoside reverse transcriptase inhibitors. ChemMedChem 2011, 6, 2225–2232. [Google Scholar] [CrossRef]
  8. Tributino, J.L.; Duarte, C.D.; Corrêa, R.S.; Doriguetto, A.C.; Ellena, J.; Romeiro, N.C.; Castro, N.G.; Miranda, A.L.P.; Barreiro, E.J.; Fraga, C.A. Novel 6-methanesulfonamide-3, 4-methylenedioxyphenyl-N-acylhydrazones: Orally effective anti-inflammatory drug candidates. Bioorg. Med. Chem. 2009, 17, 1125–1131. [Google Scholar] [CrossRef]
  9. Liu, W.-Y.; Li, H.-Y.; Zhao, B.-X.; Shin, D.-S.; Lian, S.; Miao, J.-Y. Synthesis of novel ribavirin hydrazone derivatives and anti-proliferative activity against A549 lung cancer cells. Carbohydr. Res. 2009, 344, 1270–1275. [Google Scholar] [CrossRef]
  10. Gil-Longo, J.; Laguna, M.D.L.R.; Verde, I.; Castro, M.E.; Orallo, F.; Fontenla, J.A.; Calleja, J.M.; Ravina, E.; Teran, C. Pyridazine derivatives. XI: Antihypertensive activity of 3-hydrazinocycloheptyl [1, 2-c] pyridazine and its hydrazone derivatives. J. Pharm. Sci. 1993, 82, 286–290. [Google Scholar] [CrossRef] [Green Version]
  11. Mahajan, A.; Kremer, L.; Louw, S.; Guéradel, Y.; Chibale, K.; Biot, C. Synthesis and in vitro antitubercular activity of ferrocene-based hydrazones. Bioorg. Med. Chem. Lett. 2011, 21, 2866–2868. [Google Scholar] [CrossRef] [PubMed]
  12. Fattorusso, C.; Campiani, G.; Kukreja, G.; Persico, M.; Butini, S.; Romano, M.P.; Altarelli, M.; Ros, S.; Brindisi, M.; Savini, L. Design, synthesis, and structure–activity relationship studies of 4-quinolinyl-and 9-acrydinylhydrazones as potent antimalarial agents. J. Med. Chem. 2008, 51, 1333–1343. [Google Scholar] [CrossRef] [PubMed]
  13. De Oliveira, K.N.; Costa, P.; Santin, J.R.; Mazzambani, L.; Bürger, C.; Mora, C.; Nunes, R.J.; De Souza, M.M. Synthesis and antidepressant-like activity evaluation of sulphonamides and sulphonyl-hydrazones. Bioorg. Med. Chem. 2011, 19, 4295–4306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Musad, E.A.; Mohamed, R.; Saeed, B.A.; Vishwanath, B.S.; Rai, K.L. Synthesis and evaluation of antioxidant and antibacterial activities of new substituted bis (1, 3, 4-oxadiazoles), 3, 5-bis (substituted) pyrazoles and isoxazoles. Bioorg. Med. Chem. Lett. 2011, 21, 3536–3540. [Google Scholar] [CrossRef]
  15. Jain, J.; Kumar, Y.; Sinha, R.; Kumar, R.; Stables, J. Menthone aryl acid hydrazones: A new class of anticonvulsants. Med. Chem. 2011, 7, 56–61. [Google Scholar] [CrossRef]
  16. Mu, J.-X.; Shi, Y.-X.; Wu, H.-K.; Sun, Z.-H.; Yang, M.-Y.; Liu, X.-H.; Li, B.-J. Microwave assisted synthesis, antifungal activity, DFT and SAR study of 1,2,4-triazolo[4,3-a]pyridine derivatives containing hydrazone moieties. Chem. Cent. J. 2016, 10, 50. [Google Scholar] [CrossRef] [Green Version]
  17. Özdemir, A.; Turan-Zitouni, G.; Asim kaplancikli, Z.; Demirci, F.; Iscan, G. Studies on hydrazone derivatives as antifungal agents. J. Enzym. Inhib. Med. Chem. 2008, 23, 470–475. [Google Scholar] [CrossRef]
  18. Lidström, P.; Tierney, J.; Watheyb, B.; Westmana, J. Microwave assisted organic synthesis: A review. Tetrahedron 2001, 57, 9225–9283. [Google Scholar] [CrossRef]
  19. Easmon, J.; Pürstinger, G.; Thies, K.-S.; Heinisch, G.; Hofmann, J. Synthesis, structure−activity relationships, and antitumor studies of 2-benzoxazolyl hydrazones derived from alpha-(N)-acyl heteroaromatics. J. Med. Chem. 2006, 49, 6343–6350. [Google Scholar] [CrossRef]
  20. Mukherjee, S.; Chowdhury, S.; Paul, A.K.; Banerjee, R. Selective extraction of palladium (II) using hydrazone ligand: A novel fluorescent sensor. J. Lumin. 2011, 131, 2342–2346. [Google Scholar] [CrossRef]
  21. Chaouiki, A.; Chafiq, M.; Lgaz, H.; Al-Hadeethi, M.R.; Ali, I.H.; Masroor, S.; Chung, I.-M. Green Corrosion Inhibition of Mild Steel by Hydrazone Derivatives in 1.0 M HCl. Coatings 2020, 10, 640. [Google Scholar] [CrossRef]
  22. Chaston, T.B.; Richardson, D.R. Interactions of the pyridine-2-carboxaldehyde isonicotinoyl hydrazone class of chelators with iron and DNA: Implications for toxicity in the treatment of iron overload disease. JBIC J. Biol. Inorg. Chem. 2003, 8, 427–438. [Google Scholar] [CrossRef] [PubMed]
  23. Steiner, T. The whole palette of hydrogen bonds. Angew. Chem. Int. Ed. 2002, 41, 14–76. [Google Scholar]
  24. Gerlt, J.A.; Kreevoy, M.M.; Cleland, W.; Frey, P.A. Understanding enzymic catalysis: The importance of short, strong hydrogen bonds. Chem. Biol. 1997, 4, 259–267. [Google Scholar] [CrossRef] [Green Version]
  25. Perrin, C.L.; Nielson, J.B. “Strong” hydrogen bonds in chemistry and biology. Annu. Rev. Phys. Chem. 1997, 48, 511–544. [Google Scholar] [CrossRef]
  26. Emamian, S.; Lu, T.; Kruse, H.; Emamian, H. Exploring Nature and Predicting Strength of Hydrogen Bonds: A Correlation Analysis Between Atoms-in-Molecules Descriptors, Binding Energies, and Energy Components of Symmetry-Adapted Perturbation Theory. J. Comput. Chem. 2019, 40, 2868–2881. [Google Scholar] [CrossRef]
  27. Zundel, G. Hydrogen bonds with large proton polarizability and proton transfer processes in electrochemistry and biology. Adv. Chem. Phys. 1999, 111, 1–217. [Google Scholar] [CrossRef]
  28. Bernstein, J.; Etter, M.C.; Leiserowitz, L. The role of hydrogen bonding in molecular assemblies. Struct. Correl. 1994, 431–507. [Google Scholar] [CrossRef]
  29. Pokorna, P.; Krepl, M.; Kruse, H.; Sponer, J. MD and QM/MM Study of the Quaternary HutP Homohexamer Complex with mRNA, L-Histidine Ligand, and Mg2+. J. Chem. Theory Comput. 2017, 13, 5658–5670. [Google Scholar] [CrossRef]
  30. Zhang, Z.; Schreiner, P.R. (Thio) urea organocatalysis—What can be learnt from anion recognition? Chem. Soc. Rev. 2009, 38, 1187–1198. [Google Scholar] [CrossRef]
  31. Doyle, A.G.; Jacobsen, E.N. Small-molecule H-bond donors in asymmetric catalysis. Chem. Rev. 2007, 107, 5713–5743. [Google Scholar] [CrossRef] [PubMed]
  32. Taylor, M.S.; Jacobsen, E.N. Asymmetric catalysis by chiral hydrogen-bond donors. Angew. Chem. Int. Ed. 2006, 45, 1520–1543. [Google Scholar] [CrossRef] [PubMed]
  33. Schreiner, P.R. Metal-free organocatalysis through explicit hydrogen bonding interactions. Chem. Soc. Rev. 2003, 32, 289–296. [Google Scholar] [CrossRef]
  34. Pihko, P.M. Activation of carbonyl compounds by double hydrogen bonding: An emerging tool in asymmetric catalysis. Angew. Chem. Int. Ed. 2004, 43, 2062–2064. [Google Scholar] [CrossRef] [PubMed]
  35. Ali, A.; Khalid, M.; Abid, S.; Iqbal, J.; Tahir, M.N.; Rauf Raza, A.; Zukerman-Schpector, J.; Paixão, M.W. Facile synthesis, crystal growth, characterization and computational study of new pyridine-based halogenated hydrazones: Unveiling the stabilization behavior in terms of noncovalent interactions. Appl. Organomet. Chem. 2020, 34, e5399. [Google Scholar] [CrossRef]
  36. Zhang, L.; Yang, M.; Hu, B.; Sun, Z.; Liu, X.; Weng, J.; Tan, C. Microwave-assisted synthesis of novel 8-chloro-[1, 2, 4] triazolo [4, 3- alpha ] pyridine derivatives. Turk. J. Chem. 2015, 39, 867–873. [Google Scholar] [CrossRef]
  37. De la Hoz, A.; Diaz-Ortiz, A.; Moreno, A. Microwaves in organic synthesis. Thermal and non-thermal microwave effects. Chem. Soc. Rev. 2005, 34, 164–178. [Google Scholar] [CrossRef] [PubMed]
  38. Amariucai-Mantu, D.; Mangalagiu, V.; Danac, R.; Mangalagiu, I.I. Microwave Assisted Reactions of Azaheterocycles Formedicinal Chemistry Applications. Molecules 2020, 25, 716. [Google Scholar] [CrossRef] [Green Version]
  39. APEX2 (Version 1.22) and SAINT-Plus (Version 7.06a); Bruker: Billica, MA, USA, 2009.
  40. Higashino, T.; Akiyama, Y.; Kojima, H.; Kawamoto, T.; Mori, T. Organic semiconductors and conductors with tert-butyl substituents. Crystals 2012, 2, 1222–1238. [Google Scholar] [CrossRef]
  41. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  42. Delgado, G.; Henao, J.; Quintana, J.; Al-Maqtari, H.; Jamalis, J.; Sirat, H. Structural Characterization of a New Chalcone Compound Containing a Thiophene Moiety:(E)-3-(5-Bromothiophen-2-YL)-1-(2, 5-Dichlorothiophen-3-YL)-2-Propen-1-One. J. Struct. Chem. 2018, 59, 1440–1445. [Google Scholar] [CrossRef]
  43. Thomassen, I.K.; McCormick, L.J.; Ghosh, A. Molecular Structure of a Free-Base β-Octaiodo-meso-tetraarylporphyrin. A Rational Route to cis Porphyrin Tautomers? Cryst. Growth Des. 2018, 18, 4257–4259. [Google Scholar] [CrossRef]
  44. Ali, A.; Badawy, M.; Shah, R.; Rehman, W.; El kilany, Y.; El Ashry, E.S.H.; Tahir, N. Synthesis, characterization and in-silico admet screening of mono-and dicarbomethoxylated 6, 6′-methylenebis (2-cyclohexyl-4-methylphenol) and their hydrazides and hydrazones. Chim. Sin. 2017, 8, 446–460. [Google Scholar]
  45. Braga, A.A.C.; Ujaque, G.; Maseras, F. A DFT Study of the Full Catalytic Cycle of the Suzuki−Miyaura Cross-Coupling on a Model System. Organometallics 2006, 25, 3647–3658. [Google Scholar] [CrossRef]
  46. García-Melchor, M.; Braga, A.A.C.; Lledós, A.; Ujaque, G.; Maseras, F. Computational Perspective on Pd-Catalyzed C–C Cross-Coupling Reaction Mechanisms. Acc. Chem. Res. 2013, 46, 2626–2634. [Google Scholar] [CrossRef]
  47. Braga, A.A.; Morgon, N.H.; Ujaque, G.; Maseras, F. Computational characterization of the role of the base in the Suzuki−Miyaura cross-coupling reaction. J. Am. Chem. Soc. 2005, 127, 9298–9307. [Google Scholar] [CrossRef]
  48. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.; et al. D. 0109, Revision D. 01; Gaussian Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  49. Dennington, R.; Keith, T.; Millam, J. GaussView; Version 5; Semichem Inc.: Shawnee Mission, KS, USA, 2009. [Google Scholar]
  50. Chemcraft. Graphical Software for Visualization of Quantum Chemistry Computations. Available online: https://www.chemcraftprog.com (accessed on 31 March 2020).
  51. Hanwell, M.D.; Curtis, D.E.; Lonie, D.C.; Vandermeersch, T.; Zurek, E.; Hutchison, G.R. Avogadro: An advanced semantic chemical editor, visualization, and analysis platform. J. Cheminform. 2012, 4, 17. [Google Scholar] [CrossRef] [Green Version]
  52. O’boyle, N.M.; Tenderholt, A.L.; Langner, K.M. Cclib: A library for package-independent computational chemistry algorithms. J. Comput. Chem. 2008, 29, 839–845. [Google Scholar] [CrossRef]
  53. Keith, T.A. AIMAll, TK Gristmill Software; AIMAll: Overland Park, KS, USA, 2012. [Google Scholar]
  54. Wolff, S.; Grimwood, D.; McKinnon, J.; Turner, M.; Jayatilaka, D.; Spackman, M. CrystalExplorer; Version 3.1; University of Western Australia: Crawley, Australia, 2012. [Google Scholar]
  55. Glendenning, E.; Reed, A.; Carpenter, J.; Weinhold, F. NBO; Version 3.1; 2001; Available online: http://www.ccl.net/cca/software/MS-WIN95-NT/mopac6/nbo.htm (accessed on 31 August 2020).
  56. Weinhold, F.; Glendening, E.D. NBO 5.0 Program Manual; Theoretical Chemistry Institute and Department of Chemistry, University of Wisconsin: Madison, WI, USA, 2001; Volume 53706, p. 101. [Google Scholar]
  57. Gross, E.K.U.; Kohn, W. Time-dependent density-functional theory. In Advances in Quantum Chemistry; Löwdin, P.-O., Ed.; Academic Press: Cambridge, MA, USA, 1990; Volume 21, pp. 255–291. [Google Scholar]
  58. Burke, K.; Werschnik, J.; Gross, E. Time-dependent density functional theory: Past, present, and future. J. Chem. Phys. 2005, 123, 062206. [Google Scholar] [CrossRef] [Green Version]
  59. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 2009, 11, 19–32. [Google Scholar] [CrossRef]
  60. McKinnon, J.J.; Jayatilaka, D.; Spackman, M.A. Towards quantitative analysis of intermolecular interactions with Hirshfeld surfaces. Chem. Commun. 2007, 3814–3816. [Google Scholar] [CrossRef] [PubMed]
  61. Kumar, P.S.V.; Raghavendra, V.; Subramanian, V. Bader’s theory of atoms in molecules (AIM) and its applications to chemical bonding. J. Chem. Sci. 2016, 128, 1527–1536. [Google Scholar] [CrossRef]
  62. Parr, R.G.; Donnelly, R.A.; Levy, M.; Palke, W.E. Electronegativity: The density functional viewpoint. J. Chem. Phys. 1978, 68, 3801–3807. [Google Scholar] [CrossRef]
  63. Chattaraj, P.K.; Sarkar, U.; Roy, D.R. Electrophilicity Index. Chem. Rev. 2006, 106, 2065–2091. [Google Scholar] [CrossRef] [PubMed]
  64. Chattaraj, P.K.; Roy, D.R. Update 1 of: Electrophilicity Index. Chem. Rev. 2007, 107, PR46–PR74. [Google Scholar] [CrossRef]
  65. Saravanan, S.; Balachandran, V. Quantum chemical studies, natural bond orbital analysis and thermodynamic function of 2, 5-dichlorophenylisocyanate. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 120, 351–364. [Google Scholar] [CrossRef]
  66. Parr, R.G.; Pearson, R.G. Absolute hardness: Companion parameter to absolute electronegativity. J. Am. Chem. Soc. 1983, 105, 7512–7516. [Google Scholar] [CrossRef]
  67. Sheela, N.; Muthu, S.; Sampathkrishnan, S. Molecular orbital studies (hardness, chemical potential and electrophilicity), vibrational investigation and theoretical NBO analysis of 4-4′-(1H-1, 2, 4-triazol-1-yl methylene) dibenzonitrile based on abinitio and DFT methods. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 120, 237–251. [Google Scholar] [CrossRef]
  68. Parthasarathi, R.; Padmanabhan, J.; Elango, M.; Subramanian, V.; Chattaraj, P. Intermolecular reactivity through the generalized philicity concept. Chem. Phys. Lett. 2004, 394, 225–230. [Google Scholar] [CrossRef]
  69. Politzer, P.; Truhlar, D.G. Introduction: The role of the electrostatic potential in chemistry. In Chemical Applications of Atomic and Molecular Electrostatic Potentials; Springer: Boston, MA, USA, 1981; pp. 1–6. [Google Scholar] [CrossRef]
  70. Parr, R.G.; Szentpaly, L.v.; Liu, S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. [Google Scholar] [CrossRef]
  71. Roy, D.; Sarkar, U.; Chattaraj, P.; Mitra, A.; Padmanabhan, J.; Parthasarathi, R.; Subramanian, V.; Van Damme, S.; Bultinck, P. Analyzing toxicity through electrophilicity. Mol. Divers. 2006, 10, 119–131. [Google Scholar] [CrossRef] [PubMed]
  72. Lesar, A.; Milošev, I. Density functional study of the corrosion inhibition properties of 1, 2, 4-triazole and its amino derivatives. Chem. Phys. Lett. 2009, 483, 198–203. [Google Scholar] [CrossRef]
  73. Koopmans, T. Über die Zuordnung von Wellenfunktionen und Eigenwerten zu den einzelnen Elektronen eines Atoms. Physica 1934, 1, 104–113. [Google Scholar] [CrossRef]
  74. Desiraju, G.R. The CH. cntdot. cntdot. cntdot. O hydrogen bond in crystals: What is it? Acc. Chem. Res. 1991, 24, 290–296. [Google Scholar] [CrossRef]
  75. Desiraju, G.R.; Kishan, K.R. Crystal chemistry of some (alkoxyphenyl) propiolic acids. The role of oxygen and hydrogen atoms in determining stack structures of planar aromatic compounds. J. Am. Chem. Soc. 1989, 111, 4838–4843. [Google Scholar] [CrossRef]
  76. Çakmak, O.; Ökten, S.; Alımlı, D.; Ersanlı, C.C.; Taslimi, P.; Koçyiğit, Ü.M. Novel piperazine and morpholine substituted quinolines: Selective synthesis through activation of 3, 6, 8-tribromoquinoline, characterization and their some metabolic enzymes inhibition potentials. J. Mol. Struct. 2020, 128666. [Google Scholar] [CrossRef]
  77. Spackman, M.A.; Byrom, P.G. A novel definition of a molecule in a crystal. Chem. Phys. Lett. 1997, 267, 215–220. [Google Scholar] [CrossRef]
  78. Zukerman-Schpector, J.; Dallasta Pedroso, S.; Sousa Madureira, L.; Weber Paixão, M.; Ali, A.; Tiekink, E.R. 4-Benzyl-1-(4-nitrophenyl)-1H-1, 2, 3-triazole: Crystal structure and Hirshfeld analysis. Acta Crystallogr. Sect. E Crystallogr. Commun. 2017, 73, 1716–1720. [Google Scholar] [CrossRef] [Green Version]
  79. McKinnon, J.J.; Spackman, M.A.; Mitchell, A.S. Novel tools for visualizing and exploring intermolecular interactions in molecular crystals. Acta Crystallogr. Sect. B Struct. Sci. 2004, 60, 627–668. [Google Scholar] [CrossRef]
  80. Wang, H.; Yin, Z. Crystal structure and Hirshfeld surface analysis of dibutyl 5, 5′-(pentane-3, 3-diyl) bis (1H-pyrrole-5-carboxylate). Acta Crystallogr. Sect. E Crystallogr. Commun. 2019, 75, 711. [Google Scholar] [CrossRef] [Green Version]
  81. Hirshfeld, F.L. Bonded-atom fragments for describing molecular charge densities. Theor. Chim. Acta 1977, 44, 129–138. [Google Scholar] [CrossRef]
  82. Pook, N.-P.; Adam, A.; Gjikaj, M. Crystal structure and Hirshfeld surface analysis of (μ-2-{4-[(carboxylatomethyl) carbamoyl] benzamido} acetato-κ2O: O′) bis [bis (1, 10-phenanthroline-κ2N, N′) copper (II)] dinitrate N, N′-(1, 4-phenylenedicarbonyl) diglycine monosolvate octahydrate. Acta Crystallogr. Sect. E Crystallogr. Commun. 2019, 75, 667–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Sangeetha, K.; Rajina, S.; Marchewka, M.; Binoy, J. The study of inter and intramolecular hydrogen bonds of NLO crystal melaminium hydrogen malonate using DFT simulation, AIM analysis and Hirshfeld surface analysis. Mater. Today Proc. 2020, 25, 307–315. [Google Scholar] [CrossRef]
  84. Toronto, S.T. Artificial Intelligence for Engineering Design, Analysis and Manufacturing; AIE, Cambridge University Press: Cambridge, UK, 1989; Volume 3, pp. B1–B8. [Google Scholar] [CrossRef] [Green Version]
  85. Koenderink, J.J.; Van Doorn, A.J. The singularities of the visual mapping. Biol. Cybern. 1976, 24, 51–59. [Google Scholar] [CrossRef] [PubMed]
  86. Ali, A.; Zukerman-Schpector, J.; Weber Paixão, M.; Jotani, M.M.; Tiekink, E.R. 7-Methyl-5-[(4-methylbenzene) sulfonyl]-2H, 5H-[1, 3] dioxolo [4, 5-f] indole: Crystal structure and Hirshfeld analysis. Acta Crystallogr. Sect. E Crystallogr. Commun. 2018, 74, 184–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Sebbar, N.K.; Hni, B.; Hökelek, T.; Jaouhar, A.; Labd Taha, M.; Mague, J.T.; Essassi, E.M. Crystal structure, Hirshfeld surface analysis and interaction energy and DFT studies of 3-{(2Z)-2-[(2, 4-dichlorophenyl) methylidene]-3-oxo-3, 4-dihydro-2H-1, 4-benzothiazin-4-yl} propanenitrile. Acta Crystallogr. Sect. E Crystallogr. Commun. 2019, 75, 721–727. [Google Scholar] [CrossRef] [Green Version]
  88. Tahir, M.N.; Ashfaq, M.; Alexander, F.; Caballero, J.; Hernández-Rodríguez, E.W.; Ali, A. Rationalizing the stability and interactions of 2, 4-diamino-5-(4-chlorophenyl)-6-ethylpyrimidin-1-ium 2-hydroxy-3, 5-dinitrobenzoate salt. J. Mol. Struct. 2019, 1193, 185–194. [Google Scholar] [CrossRef]
  89. Khalid, M.; Ali, A.; Rehman, M.F.U.; Mustaqeem, M.; Ali, S.; Khan, M.U.; Asim, S.; Ahmad, N.; Saleem, M. Exploration of Noncovalent Interactions, Chemical Reactivity, and Nonlinear Optical Properties of Piperidone Derivatives: A Concise Theoretical Approach. ACS Omega 2020, 5, 13236–13249. [Google Scholar] [CrossRef]
  90. Thakur, T.S.; Dubey, R.; Desiraju, G.R. Intermolecular atom–atom bonds in crystals–a chemical perspective. IUCrJ 2015, 2, 159–160. [Google Scholar] [CrossRef] [Green Version]
  91. Bader, R.F. Atoms in molecules. Acc. Chem. Res. 1985, 18, 9–15. [Google Scholar] [CrossRef]
  92. Bader, R.; Nguyen-Dang, T. Quantum theory of atoms in molecules–Dalton revisited. In Advances in Quantum Chemistry; Academic Press: Cambridge, MA, USA; Elsevier: Amsterdam, The Netherlands, 1981; Volume 14, pp. 63–124. [Google Scholar] [CrossRef]
  93. Desiraju, G.R. Crystal engineering: A brief overview. J. Chem. Sci. 2010, 122, 667–675. [Google Scholar] [CrossRef]
  94. Hernández-Paredes, J.; Carrillo-Torres, R.C.; López-Zavala, A.A.; Sotelo-Mundo, R.R.; Hernández-Negrete, O.; Ramírez, J.Z.; Alvarez-Ramos, M.E. Molecular structure, hydrogen-bonding patterns and topological analysis (QTAIM and NCI) of 5-methoxy-2-nitroaniline and 5-methoxy-2-nitroaniline with 2-amino-5-nitropyridine (1: 1) co-crystal. J. Mol. Struct. 2016, 1119, 505–516. [Google Scholar] [CrossRef]
  95. Tamer, Ö.; Avcı, D.; Atalay, Y. Quantum chemical characterization of N-(2-hydroxybenzylidene) acetohydrazide (HBAH): A detailed vibrational and NLO analysis. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 117, 78–86. [Google Scholar] [CrossRef]
  96. Szafran, M.; Komasa, A.; Bartoszak-Adamska, E. Crystal and molecular structure of 4-carboxypiperidinium chloride (4-piperidinecarboxylic acid hydrochloride). J. Mol. Struct. 2007, 827, 101–107. [Google Scholar] [CrossRef]
  97. James, C.; Raj, A.A.; Reghunathan, R.; Jayakumar, V.; Joe, I.H. Structural conformation and vibrational spectroscopic studies of 2, 6-bis (p-N, N-dimethyl benzylidene) cyclohexanone using density functional theory. J. Raman Spectrosc. Int. J. Orig. Work All Asp. Raman Spectrosc. Incl. High. Order Process. Also Brillouin Rayleigh Scatt. 2006, 37, 1381–1392. [Google Scholar] [CrossRef]
  98. Pauling, L. The Nature of the Chemical Bond; Cornell University Press: Ithaca, NY, USA, 1960; Volume 260. [Google Scholar]
  99. Mulliken, R. Overlap populations, bond orders and covalent bond energies. J. Chem. Phys. 1955, 23, 1841–1846. [Google Scholar] [CrossRef]
  100. Javed, I.; Khurshid, A.; Arshad, M.N.; Wang, Y. Photophysical and electrochemical properties and temperature dependent geometrical isomerism in alkyl quinacridonediimines. New J. Chem. 2014, 38, 752–761. [Google Scholar] [CrossRef] [Green Version]
  101. Khalid, M.; Ali, A.; Jawaria, R.; Asghar, M.A.; Asim, S.; Khan, M.U.; Hussain, R.; ur Rehman, M.F.; Ennis, C.J.; Akram, M.S. First principles study of electronic and nonlinear optical properties of A–D–π–A and D–A–D–π–A configured compounds containing novel quinoline–carbazole derivatives. RSC Adv. 2020, 10, 22273–22283. [Google Scholar] [CrossRef]
  102. Ali, A.; Khalid, M.; Rehman, M.F.u.; Haq, S.; Ali, A.; Tahir, M.N.; Ashfaq, M.; Rasool, F.; Braga, A.A.C. Efficient Synthesis, SC-XRD, and Theoretical Studies of O-Benzenesulfonylated Pyrimidines: Role of Noncovalent Interaction Influence in Their Supramolecular Network. ACS Omega 2020, 5, 15115–15128. [Google Scholar] [CrossRef]
  103. Luque, F.J.; López, J.M.; Orozco, M. Perspective on “Electrostatic interactions of a solute with a continuum. A direct utilization of ab initio molecular potentials for the prevision of solvent effects”. Theor. Chem. Acc. 2000, 103, 343–345. [Google Scholar] [CrossRef]
Figure 1. The functional diversity of the hydrazone skeleton.
Figure 1. The functional diversity of the hydrazone skeleton.
Crystals 10 00778 g001
Scheme 1. Ultrasonic-based Synthesis of (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene)acetohydrazide.
Scheme 1. Ultrasonic-based Synthesis of (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene)acetohydrazide.
Crystals 10 00778 sch001
Scheme 2. Isomeric existence of the title compound. The highlighted region shows the double-bond conferring E and Z isomers.
Scheme 2. Isomeric existence of the title compound. The highlighted region shows the double-bond conferring E and Z isomers.
Crystals 10 00778 sch002
Figure 2. ORTEP diagram of HBPAH drawn at a probability level of 30% with H-atoms are displayed by tiny circles of arbitrary radii. Red color shows oxygens, blue nitrogen, green chlorine, white is for hydrogen, and black/white contours show carbon atoms.
Figure 2. ORTEP diagram of HBPAH drawn at a probability level of 30% with H-atoms are displayed by tiny circles of arbitrary radii. Red color shows oxygens, blue nitrogen, green chlorine, white is for hydrogen, and black/white contours show carbon atoms.
Crystals 10 00778 g002
Figure 3. Molecular overlay of two crystallographically independent molecules: first molecule (red) and second molecule (blue).
Figure 3. Molecular overlay of two crystallographically independent molecules: first molecule (red) and second molecule (blue).
Crystals 10 00778 g003
Figure 4. Packing diagram of HBPAH showing H-bonded connection of first, 2nd type of molecules and water with H-atoms not engaged in H-bonding are omitted for clearness. Red color shows oxygens, blue nitrogen, green chlorine, and black/white contours shows carbon atoms.
Figure 4. Packing diagram of HBPAH showing H-bonded connection of first, 2nd type of molecules and water with H-atoms not engaged in H-bonding are omitted for clearness. Red color shows oxygens, blue nitrogen, green chlorine, and black/white contours shows carbon atoms.
Crystals 10 00778 g004
Figure 5. Cyclic Face-to-face stacking interaction between various rings in the crystal packing. Distances shown are in Å with H-atoms omitted for clarity.
Figure 5. Cyclic Face-to-face stacking interaction between various rings in the crystal packing. Distances shown are in Å with H-atoms omitted for clarity.
Crystals 10 00778 g005
Figure 6. Hirshfeld surfaces of the entitled compound mapped over (a) dnorm, (b) shape index, and (c) curvedness, respectively, for HBPAH (1 a.u. of electron density = 6.748 e.Å−3).
Figure 6. Hirshfeld surfaces of the entitled compound mapped over (a) dnorm, (b) shape index, and (c) curvedness, respectively, for HBPAH (1 a.u. of electron density = 6.748 e.Å−3).
Crystals 10 00778 g006
Figure 7. Percentage contributions of all interatomic contacts for an entitled compound.
Figure 7. Percentage contributions of all interatomic contacts for an entitled compound.
Crystals 10 00778 g007
Figure 8. Hydrogen bonds in HBPAH.
Figure 8. Hydrogen bonds in HBPAH.
Crystals 10 00778 g008
Figure 9. AIM based Schematic structure of HBPAH.
Figure 9. AIM based Schematic structure of HBPAH.
Crystals 10 00778 g009
Figure 10. Frontier molecular orbitals (FMOs) for HBPAH. (a) EHOMO, ELUMO, (b) EHOMO-1, ELUMO+1, and (c) EHOMO-2, ELUMO+2, molecular orbitals. d energy gap (∆E) in shown in eV of the entitled compound at the DFT/ B3LYP/6-311G (d,p) level of theory.
Figure 10. Frontier molecular orbitals (FMOs) for HBPAH. (a) EHOMO, ELUMO, (b) EHOMO-1, ELUMO+1, and (c) EHOMO-2, ELUMO+2, molecular orbitals. d energy gap (∆E) in shown in eV of the entitled compound at the DFT/ B3LYP/6-311G (d,p) level of theory.
Crystals 10 00778 g010
Figure 11. Molecular electrostatic potential and color scheme of HBPAH.
Figure 11. Molecular electrostatic potential and color scheme of HBPAH.
Crystals 10 00778 g011
Table 1. Experimental details of the compound (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene) acetohydrazide (HBPAH).
Table 1. Experimental details of the compound (E)-2-((6-chloropyridin-2-yl)oxy)-N′-(2-hydroxybenzylidene) acetohydrazide (HBPAH).
Crystal DataHBPAH
CCDC* number2012169
Chemical formula2(C14H12ClN3O3)·H2O
Mr629.45
Crystal system, space groupTriclinic, P 1 ¯
Temperature (K)296
a, b, c (Å)6.6987 (8), 7.3628 (9), 31.513 (4)
α, β, γ (°)90.978 (7), 93.508 (6), 113.954 (7)
V3)1416.2 (3)
Z2
Density (calculated)1.476 Mg/m3
F(000)652
Radiation typeMo Kα
Wavelength (λ)0.71073 Å
µ (mm−1)0.288
Crystal shapeNeedle
Crystal ColorColorless
Crystal size (mm)0.38 × 0.22 × 0.18
Data CollectionHBPAH
DiffractometerBruker APEXII CCD diffractometer
Absorption correctionmulti-scan (SADABS; Bruker, 2007)
No. of measured, independent and observed [I> 2s(I)] reflections15,051, 5452, 2920
Rint0.070
Theta range for data collection0.648 to 26.000°
Index ranges−8 ≤ h ≤ 7, -9 ≤ k ≤ 9, −37 ≤ l ≤ 38
(sin θ/λ)max−1)0.617
RefinementHBPAH
R[F2 > 2σ(F2)], wR(F2), S0.083, 0.195, 1.05
No. of reflections5452
No. of parameters396
H-atom treatmentH atoms treated by a mixture of independent and constrained refinement
Δρmax, Δρmin (e Å−3)0.24, −0.30
*CCDC (Cambridge Crystallographic Data Centre).
Table 2. Geometrical parameters of potential Hydrogen-bonds (Å, º) for HBPAH.
Table 2. Geometrical parameters of potential Hydrogen-bonds (Å, º) for HBPAH.
D–H⋯AD–HH⋯AD⋯AD–H⋯A
O3–H3A⋯N30.821.952.659 (5)145
N2–H2A⋯O10.862.2352.655105.34
N2–H2A⋯O70.862.082.902 (5)160
O6–H6⋯N60.821.882.587 (5)145
N5–H5⋯O40.862.1702.571108.18
N5–H5⋯O2i0.862.483.029 (5)122
O7–H7A⋯O5ii0.91 (6)2.03 (6)2.877 (5)155 (6)
O7–H7B⋯O2ii0.89 (6)1.95 (7)2.826 (6)168 (6)
Symmetry codes: i x − 1, y − 1, z; ii x − 1, y, z.
Table 3. Geometry-related parameters of cyclic face-to-face stacking interactions for HBPAH with distance given in Å.
Table 3. Geometry-related parameters of cyclic face-to-face stacking interactions for HBPAH with distance given in Å.
Cg(e)–Cg(f)DefDAefDe (f)Df (e)Ring Off-Set
Cg(1)–Cg(4)iii3.6711.2(2)3.4556(19)3.460(2)-
Cg(1)–Cg(4)iv3.8101.2(2)3.4495(19)3.479(2)-
Cg(2)–Cg(3)iv3.8742.7(2)3.4350(19)3.3581(19)-
Cg(2)–Cg(3)v3.8762.7(2)3.4495(19)3.4066(19)-
Symmetry codes: iii x, 1 + y, z; iv 1 + x, 1 + y, 1 + z; v 1 + x, 1 + y, z.
Table 4. AIM properties of selected values for the main interactions for HBPAH; electronic density (ρ), Laplacian of density ( 2 ρ ), ellipticity (ε) and density of potential energy (V), Laplacian of density ( 2 ρ ), ellipticity (ε) and density of potential energy (V).
Table 4. AIM properties of selected values for the main interactions for HBPAH; electronic density (ρ), Laplacian of density ( 2 ρ ), ellipticity (ε) and density of potential energy (V), Laplacian of density ( 2 ρ ), ellipticity (ε) and density of potential energy (V).
BCPBondsρ (e/a3) 2 ρ   ( e / a 5 ) εVa
4O2–H8+0.0181+0.0878+0.5847−0.0144
5H8–O36+0.0127+0.0487+0.0339−0.0081
8H5–N9+0.0436+0.1157+0.0397−0.0392
18H16–O36+0.0029+0.0107+0.0863−0.0017
26H23–O36+0.0091+0.0279+0.0808−0.0054
42O2–O37+0.0066+0.0267+0.2101−0.0052
43H19–O37+0.0073+0.0263+0.1561−0.0047
49O35–H41+0.0179+0.0868+0.6171−0.0143
51H38–N42+0.0421+0.1147+0.0433−0.0375
55H41–O67+0.0148+0.0553+0.1317−0.0097
65H49–O67+0.0062+0.0181+0.0638−0.0035
66H33–H53+0.0016+0.0052+0.1869−0.0007
70H56–O67+0.0114+0.0324+0.0980−0.0068
Va (hartree.e/a3).
Table 5. Natural bonding orbital (NBO) analysis for HBPAH using the B3LYP/6-311G(d,p) level.
Table 5. Natural bonding orbital (NBO) analysis for HBPAH using the B3LYP/6-311G(d,p) level.
CompoundDonor(i)TypeAcceptor(j)TypeEa(2)E(j)E(i)b(a.u)F(i,j)c(a.u)
C13-H14C11-C13∂ *0.511.090.021
C46-C48πN39-C50π *30.350.260.082
C13-C15πN6-C17π *29.620.260.082
N39-C50πC43-C44π *28.320.330.088
C43-C44πC45-C48π *22.450.300.074
C30-C32πC26-C28π *21.460.290.071
HBPAHC13-C15πC10-C11π *16.090.270.060
C43-C44πN39-C50π *14.780.270.058
N40LP(1)O36-C54π *62.830.290.121
N7LP(1)O3-C21π *56.480.290.117
O35LP(2)N39-C50π *35.800.320.103
Cl1LP(2)N6-C10∂ *5.790.850.063
N9LP(1)N7-H8∂ *7.040.810.069
O2LP(1)N6-C17∂ *6.611.080.076
O36LP(1)N7-H8∂ *2.221.130.045
* antibonding energetic orbitals; aE(2) is the energy of hyper conjugative interaction (stabilization energy in kcal mol−1); bE(j)E(i) is the energy difference between donor and acceptor i and j NBO orbitals; cF(i;j) is the Fock matrix element between i and j NBO orbitals.
Table 6. The EHOMO, ELUMO, EHOMO-1, ELUMO+1, EHOMO-2, ELUMO+2, and energy gap (∆E) in eV of the entitled compound at the DFT/ B3LYP/6-311G (d,p) level of theory.
Table 6. The EHOMO, ELUMO, EHOMO-1, ELUMO+1, EHOMO-2, ELUMO+2, and energy gap (∆E) in eV of the entitled compound at the DFT/ B3LYP/6-311G (d,p) level of theory.
HBPAH
MO(s)EnergyE(eV)
HOMO−5.6003.634
LUMO−1.966
HOMO-1−6.2664.65
LUMO+1−1.616
HOMO-2−6.2745.034
LUMO+2−1.240
HOMO = highest occupied molecular orbital; LUMO = lowest unoccupied molecular orbital, MO = molecular orbital.

Share and Cite

MDPI and ACS Style

Ali, A.; Khalid, M.; Abid, S.; Tahir, M.N.; Iqbal, J.; Ashfaq, M.; Kanwal, F.; Lu, C.; Rehman, M.F.u. Green Synthesis, SC-XRD, Non-Covalent Interactive Potential and Electronic Communication via DFT Exploration of Pyridine-Based Hydrazone. Crystals 2020, 10, 778. https://doi.org/10.3390/cryst10090778

AMA Style

Ali A, Khalid M, Abid S, Tahir MN, Iqbal J, Ashfaq M, Kanwal F, Lu C, Rehman MFu. Green Synthesis, SC-XRD, Non-Covalent Interactive Potential and Electronic Communication via DFT Exploration of Pyridine-Based Hydrazone. Crystals. 2020; 10(9):778. https://doi.org/10.3390/cryst10090778

Chicago/Turabian Style

Ali, Akbar, Muhammad Khalid, Saba Abid, Muhammad Nawaz Tahir, Javed Iqbal, Muhammad Ashfaq, Fariha Kanwal, Changrui Lu, and Muhammad Fayyaz ur Rehman. 2020. "Green Synthesis, SC-XRD, Non-Covalent Interactive Potential and Electronic Communication via DFT Exploration of Pyridine-Based Hydrazone" Crystals 10, no. 9: 778. https://doi.org/10.3390/cryst10090778

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop