Next Article in Journal
Mathematical Analysis of Reaction–Diffusion Equations Modeling the Michaelis–Menten Kinetics in a Micro-Disk Biosensor
Next Article in Special Issue
Selenoxides as Excellent Chalcogen Bond Donors: Effect of Metal Coordination
Previous Article in Journal
The Reserve/Maximum Capacity of Melatonin’s Synthetic Function for the Potential Dimorphism of Melatonin Production and Its Biological Significance in Mammals
Previous Article in Special Issue
Comparative Study of Phosgene Gas Sensing Using Carbon and Boron Nitride Nanomaterials—A DFT Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis of 5-Aryl-N-(pyrazin-2-yl)thiophene-2-carboxamides via Suzuki Cross-Coupling Reactions, Their Electronic and Nonlinear Optical Properties through DFT Calculations

1
Department of Chemistry, Government College, University Faisalabad, Faisalabad 38000, Pakistan
2
Department of Chemistry, University of Education Lahore, Attock Campus, Attock 43600, Pakistan
3
Department of Chemistry, Faculty of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
4
School of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Korea
5
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(23), 7309; https://doi.org/10.3390/molecules26237309
Submission received: 16 August 2021 / Revised: 5 November 2021 / Accepted: 7 November 2021 / Published: 2 December 2021
(This article belongs to the Special Issue Fundamental Aspects of Chemical Bonding)

Abstract

:
Synthesis of 5-aryl-N-(pyrazin-2-yl)thiophene-2-carboxamides (4a4n) by a Suzuki cross-coupling reaction of 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3) with various aryl/heteroaryl boronic acids/pinacol esters was observed in this article. The intermediate compound 3 was prepared by condensation of pyrazin-2-amine (1) with 5-bromothiophene-2-carboxylic acid (2) mediated by TiCl4. The target pyrazine analogs (4a4n) were confirmed by NMR and mass spectrometry. In DFT calculation of target molecules, several reactivity parameters like FMOs (EHOMO, ELUMO), HOMO–LUMO energy gap, electron affinity (A), ionization energy (I), electrophilicity index (ω), chemical softness (σ) and chemical hardness (η) were considered and discussed. Effect of various substituents was observed on values of the HOMO–LUMO energy gap and hyperpolarizability. The p-electronic delocalization extended over pyrazine, benzene and thiophene was examined in studying the NLO behavior. The chemical shifts of 1H NMR of all the synthesized compounds 4a4n were calculated and compared with the experimental values.

Graphical Abstract

1. Introduction

Pyrazine derivatives are a vital group of heterocyclic compounds present in nature and have also been synthesized in laboratories since 1876 [1,2]. Pyrazines and their derivatives play a significant role as intermediates for pharmaceuticals, agricultural chemicals [1], etc. In particular, pyrazine derivatives display a large number of pharmaceutical activities: anticancer [3], diuretic [4], antidiabetic [5] and anti-inflammatory [6]. Various synthetic procedures have been implemented to synthesize these biologically active pyrazine derivatives [7,8,9]. Naturally, these are found in tobacco, roasted and nut-like flavors and also the taste and fragrance of many foodstuffs [10].
Pyrazine carboxamide analogs like T-705 and T-1105 (Figure 1) are innovative broad-spectrum viral polymerase inhibitors against various RNA viruses [11,12,13]. T-705 is also studied in terms of treatment of several viral infections, including SARS-CoV-2 [14]. Pyrazinamide analogs were also studied for their high antimicrobial activity [15]. Pyrazinamide (PZA) is a vital first-line anti-tuberculosis agent and revealed a distinct inhibiting effect [16,17]. It contributes an exclusive role in shortening the therapy duration from more than nine months previously to six months as it kills semi-dormant tubercle bacilli in acidic environments compared to other TB drugs [18].
In recent years, organic NLO (nonlinear optical) materials have attained great interest due to their remarkable uses in photonics and optoelectronics comprising optical signal processing, optical communications, optical data storage and optoelectronic transfer [19,20,21]. To display prominent second-order NLO properties, a molecule should be non-centrosymmetric along with intramolecular charge transfer transitions, a large transition dipole moment and a huge difference in the molecular dipole moment at the ground and excited state [20,22]. It can be attained in linear organic compounds by substituting electron-donating and withdrawing groups as it happens in conventional organic dipolar push–pull systems [23].
Pyrazine compounds are a rigid planar conjugated structure, have electron-deficient character, and so the pyrazinyl moiety can be used as an electron-withdrawing group in electron push–pull systems [24]. Thus, pyrazine entities incorporation in luminescent materials has been studied recently [25]. Moreover, the electron push–pull pyrazine analogs exhibited fascinating second-order [26,27,28] and third-order [29,30,31] NLO properties. Therefore, in this manuscript, we elaborated on the production of a new sequence of 5-aryl-N-(pyrazin-2-yl)thiophene-2-carboxamides (4a4n) and exploration of their NLO properties. We also focused on the structural reactivity parameters and 1H NMR spectra.

2. Results and Discussion

2.1. Chemistry

The reaction of pyrazine-2-amine (1) with 5-bromothiophene-2-carboxylic acid (2) in the presence of pyridine and titanium tetrachloride was used to produce 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3) with 75% yield, as shown in Scheme 1. Thereafter, compound 3 was treated with different aryl boronic acids/pinacol esters by using catalyst Pd(PPh3)4, potassium phosphate and solvent 1,4-dioxane. The derivatives of 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (4a4n) were obtained with moderate and good yields (37–72%). The structural representation of these molecules and their yields are provided in Figure 2. Compound 4a showed the maximum yield (72%), while compound 4n exhibited the lowest yield (37%). The compounds having substitution 3,4-dichloro (4b), 3-methylcarbonyl (4c), 4-chloro (4d), 3-cholro4-fluoro (4e), 3,5-dimethyl (4f), 4-methoxy (4g), 4-methylthio (4i) and 3,5-difluoro (4j) were also synthesized with good yields. It was noted that the synthesized molecules obtained from thiophene boronic acid pinacol esters and bulky group substituted phenylboronic acid exhibited a low yield as compared to other molecules. It is noteworthy that the boronic acids and the pinacolate esters having electron-donating moieties provided good yields while the electron-withdrawing moieties provided lower yields [32]. Exceptionally, compound 4a provided a good yield.

2.2. Computational Details

In this study, the geometric optimization of the synthesized molecules (4a4n) followed by the computation of several quantum chemical parameters was studied through density functional theory (DFT) calculations. DFT is a cheap and widely used method for modeling the ground state of molecules. From the computational point of view, these methods have become very popular in recent years because they can attain similar accuracy to the ab initio methods in less time and at a lower computational cost [33]. The optimization of the molecules (4a4n) was accomplished with the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory [34] which is widely used for providing accurate geometries and electronic properties for a large number of molecules. The solvent model density (SMD) is a polarizable continuum model that includes the full solute electron density without defining partial atomic charges. It is a universal solvation model, which means that it can be applied to any kind of molecules. [35]. Frequency calculations were performed on the optimized structures to confirm the obtained geometries as true minima by the absence of imaginary frequencies. The optimized structures were further used for molecular electrostatic potential (MESP) and frontier molecular orbital (FMO) analysis on the same level of theory. Chemical reactivity is defined as the tendency of a substance to react chemically with another substance. Reactivity parameters such as electron affinity, ionization energy, chemical softness and hardness, electrophilicity and chemical potential were also calculated. Modeling of the synthesized molecules (4a4n) was performed using GaussView 6 [36] and the calculations were performed using the Gaussian 09 [37] program.

2.2.1. Optimized Geometries

In the field of computational chemistry, geometries optimization is a crucial process to find the ground state geometry of molecules that can be used to compute other properties [38]. In the case of conformationally flexible compounds, a conformational search is crucial to find out the minimum energy conformers that must be used to compute all the other properties. A relaxed potential energy scan was performed considering the important dihedrals, and low-energy conformers were selected from the resultant potential energy scan (PES) followed by geometry optimization and removal of duplicates. These conformers were then subjected to the properties calculations like NMR and Boltzmann averaging of the results was performed afterwards. The Boltzmann-averaged values are provided in the respective tables that show a very good agreement with the experimental data. The optimized geometries of the most stable conformers of the molecules (4a4n) calculated at the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory are provided in Figure 3. All the PES figures are provided in the Supplementary Materials.

2.2.2. Nuclear Magnetic Resonance Spectroscopy (NMR)

NMR is the most important technique for determining the accurate structure of organic compounds. Quantum calculations are found to be sufficiently accurate in calculating NMR spectra and studying the relationship between the molecular structure and its chemical shifts. Therefore, the use of theoretical methods is very useful for augmenting the confidence in explaining the structure of molecules [39,40,41]. In this study, the experimental 1H spectra (4a4n) were recorded in 1,4-dioxane while the calculation of the 1H and 13C NMR spectra was performed using the same level of theory as for the optimizations. The comparison of the experimental and theoretical calculated 1H NMR data of compound 4a is provided in Table 1 while the 1H NMR data of the remaining compounds 4b4n are provided in the Supplementary Materials (Tables S1–S13). It can be seen that the performance of the NMR calculations was very good and the mean absolute error (MAE) is only 0.25 ppm.

2.2.3. FMO and NLO Analysis

The frontier molecular orbitals (FMO) of a compound are the highest occupied molecular orbitals (HOMO) and the lowest unoccupied molecular orbitals (LUMO). Logically, the HOMO are regarded as electron-donating or nucleophilic while the LUMO are regarded as electron-accepting or electrophilic. In addition, chemical reactions and resonances on one or more molecules can be described by overlapping between the filled HOMOs and empty LUMOs. These basic ideas are used in the FMO theory to elucidate the structure and reactivity of compounds. To determine the chemical stability and reactivity of the molecules, the HOMO–LUMO and their energy gap are very important quantum chemical parameters. These MO also have the key role in determining optical and electrical properties.
The molecular orbitals of all the compounds 4a4n have almost the same pattern but some compounds have a different one. In most of the compounds, the density of molecular orbitals is mostly spread over the thiophene and benzene rings, but there are some compounds in which the density of molecular orbitals also spreads over the pyrazine ring. A plot of these surfaces is displayed in Figure 4.
The highest HOMO–LUMO energy gap shows that the compound is the most stable and the least reactive and the lowest energy gap shows that the compound is the least stable and the most reactive [42,43,44].
In this series (4a4n), it was observed that compound 4m consisting of three rings (pyrazine, benzene and thiophene) with two CF3 substituents and compound 4j consisting of three rings (pyrazine, benzene and thiophene) with two fluoride substituents have the highest HOMO–LUMO energy gap of 4.93 and 4.89 eV, respectively, which shows that both these compounds are the most stable and the least reactive in this series. Compound 4i consisting of three rings (pyrazine, benzene and thiophene) with an SH substituent has the lowest HOMO–LUMO energy gap of 4.21 eV which shows that it is the least stable one with the highest reactivity. All the other compounds have a HOMO–LUMO energy gap in the range of 4.21–4.93 eV. The values of HOMO and LUMO energies and their energy gap (ΔE) along with their polarizability (α) and hyperpolarizability (β) of all the compounds 4a4n are provided in Table 2.
Through experimental and theoretical methods, organic nonlinear optical (NLO) materials have been widely studied to explore their potential applications in optical signal processing, telecommunications and optical data storage. Theoretical research may contribute the key role in understanding the origin of molecular NLO properties and predicting the relationship between the structure and the NLO properties, which is the basis for the design and production of new materials [45]. Most organic compounds with pi-conjugated systems are mainly suitable for the expansion of the NLO material because, among other features, they show large nonlinearity which is due to electronic polarizability.
In this series (4a4n), compounds 4i and 4l have the highest hyperpolarizability values of 9139.57 a.u. and 9048.28 a.u., respectively, showing the highest NLO responses, while the remaining compounds show very small hyperpolarizability values. Our research group reported N-heterocyclic derivatives having smaller values of hyperpolarizability as compared to the compounds in this study (4a4n) [34].

2.2.4. Molecular Electrostatic Potential (MESP) Analysis

In recent years, the maps of molecular electrostatic potential have been extensively used to recognize the reactive sites for nucleophilic and electrophilic attacks during chemical reactions, the study of H-bond interactions and the biological recognition process. MESP is related to the distribution of the total charge of a compound and provides information about some physical properties of the molecules such as partial charges of the atoms, electronegativity, chemical reactivity and dipole moment. In the MESP diagram, different electrostatic potential values on the surface are shown with different colors. The most positive electrostatic potential area is shown in blue, the most negative electrostatic potential area is shown in red, and green is the zero potential area [46,47,48].
In the maps of the MESP of compounds 4a4n (Figure 5), the red color shows that the most negative ESP regions are mainly located over the oxygen of the carbonyl group, which is a suitable position for the attack of electrophiles while the blue color shows that the most positive ESP regions are mainly located over the hydrogen of the amide group, which is a suitable position for the attack of nucleophiles. The maps of the MESP of compounds 4a4n are provided in Figure 6.

2.2.5. Reactivity Descriptors

Chemical reactivity is the key concept because it is closely related to the reaction mechanism so chemical reactions can be understood and synthetic procedures can be improved to obtain new materials. The ionization potential (I = −EHOMO) and electron affinity (A = −ELUMO) were calculated using Koopman’s theorem [49] as well as directly and a comparison is provided in Table 3.
The electronic chemical potential (μ) and molecular hardness (η) have been proposed to measure global reactivity [50,51]. Chemical potential (μ) describes the charge transfer at the ground state within the compound. The electrophilicity index (ω) is a thermodynamic characteristic that calculates the energy changes of a saturated chemical system after the addition of electrons. It plays an important role in determining the chemical reactivity of a system. The values of η, σ, μ and ω were calculated using the following Equations (1)–(4) [52] and are mentioned in Table 4.
Chemical hardness (η) = (I − A)/2
Chemical softness (σ) = 1/η
Chemical potential (μ) = −(I + A)/2
Electrophilicity index (ω) = μ2/2η
The lowest values of ionization energy (5.67 eV) and electron affinity (0.06 eV) of compounds 4i and 4c mentioned in Table 4 endorse their high reactivity and lowest stability in series 4a4n.
Among all the derivatives, 4f has the lowest value of η (2.46 eV) and the highest value of σ (0.40); thus, it is chemically soft (more reactive), whereas 4c has the highest value of η (3.56 eV) and the lowest value of σ (0.28 eV) and is a chemically hardest compound (less reactive). These results correlate with the HOMO–LUMO energy gaps of all the synthesized compounds.
Compound 4i has the highest electronic chemical potential value (−3.02 eV) while 4j has the lowest chemical potential value (−3.90 eV). The results indicate that 4g has the lowest electrophilicity index value of 1.69 eV and has nucleophilic nature, whereas 4j has the highest value, 2.25 eV, and has a strongly electrophilic nature.

3. Materials and Methods

3.1. General Information

All the chemicals were purchased from Sigma-Aldrich (Burlington, MA, USA), and commercial-grade solvents were used. Melting points of the synthesized compounds were checked with a Buchi B-540 melting point apparatus (New Castle, DE, USA). A Bruker NMR spectrophotometer (Billerica, MA, USA) was used to attain NMR spectra by using DMSO-d6 and CDCl3 solvents. Mass spectra were generated on a JEOL spectrometer JMS-HX-110 (USA). To monitor the reaction’s progress, TLC on silica gel 60 PF254 cards (Merck, Kenilworth, NJ, USA) was used. UV light (254–365 nm) was used to visualize the product on TLC. For the purification of compounds, silica gel (70–230 mesh) was used in columns.

3.2. Synthetic Procedure for 5-Bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3)

TiCl4 (3.0 eq., 43.47 mmol, 4.78 mL) and pyrazin-2-amine (1) (1 eq., 14.49 mmol, 1.378 g) were mixed in a solution of 5-bromothiophene-2-carboxylic acid (2) (1 eq., 14.49 mmol, 3 g) and pyridine (150 mL). The reaction mixture in a sealed schlenk flask was continuously stirred for 2 h at 85 °C. After cooling, pyridine was separated by coevaporation with toluene. Then, an aqueous solution of 1N HCl (150 mL) was added and removed with DCM (3 × 150 mL). A saturated aqueous solution of NaHCO3 (3 × 150 mL) was used to wash the combined organic layers, dried by anhydrous sodium sulfate and evaporated on a rotary evaporator. The final product was obtained after column chromatography (n-hexane/ethyl acetate, 90:10) with 75% yield [53,54].

3.3. Synthetic Procedure for 5-Aryl-N-(pyrazin-2-yl)thiophene-2-carboxamide (4a4n)

In an oven-dried schlenk flask, a 1,4-dioxane (8 mL) solvent, 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3) (1 eq., 0.704 mmol, 0.2 g) with tetrakis(triphenylphosphine)-palladium (7 mol%, 0.057 g) were added at room temperature, and inert atmosphere was applied. After stirring the reaction for half an hour, boronic acid/pinacol esters (1.1 eq., 0.774 mmol), potassium phosphate (2 eq., 1.408 mmol, 0.297 g) and water (2 mL) were added [55,56,57]. The mixture was kept on stirring at reflux for more than 20 h. After cooling to normal temperature, the reaction mixture was filtered with solvent ethyl acetate and then concentrated under reduced pressure. The resulted products were purified with column chromatography (n-hexane/ethyl acetate, 75:25) [58,59,60,61,62,63]. The final products (4a4n) were dried, recrystallized and finally analyzed with NMR and mass spectrometry. The NMR spectra of the synthesized compounds are provided in the Supplementary Materials.

3.4. Characterization Data

Compound 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3). Off-white solid, MP = 224–226 °C (75% yield, 3.3 g). 1H NMR (400 MHz, CDCl3): δ 9.62 (s, 1H), 8.40 (s, 1H), 8.29 (d, J = 11.3 Hz, 2H), 7.46 (d, J = 3.9 Hz, 1H), 7.14 (d, J = 3.8 Hz, 1H); 13C NMR (100 MHz, CDCl3): δ 159.0, 151.1, 150.6, 147.0, 140.5, 130.9, 128.8, 121.4, 119.8, 115.2, 21.5. EI/MS m/z (%): 285.1 [M + H]+; 286.1 [M + 2]; [M‒Br] = 204.0. Analitically calculated for C9H6BrN3OS: C, 38.04; H, 2.13; N, 14.79. Found: C, 38.11; H, 2.15; N, 14.74.
Compound 5-(3-chlorophenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4a). Light yellow solid, MP = 187–189 °C (72% yield, 160 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.27 (s, 1H), 9.39 (s, 1H), 8.45 (d, J = 35.5 Hz, 2H), 8.26 (d, J = 3.4 Hz, 1H), 7.84 (s, 1H), 7.72–7.70 (m, 2H), 7.50–7.45 (m, 2H); 13C NMR (151 MHz, DMSO-d6): δ 160.14, 148.72, 147.66, 142.53, 139.98, 138.25, 137.31, 134.81, 134.04, 131.86, 131.12, 128.55, 125.90, 125.32, 124.60. EI/MS m/z (%): 316.8 [M + H]+; [M‒Cl] = 280.1. Analitically calculated for C15H10ClN3OS: C, 57.05; H, 3.19; N, 13.31. Found: C, 57.09; H, 3.22; N, 13.26.
Compound 5-(3,4-dichlorophenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4b). Off-white solid, MP = 191–193 °C (59% yield, 146 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.27 (s, 1H), 9.38 (s, 1H), 8.47 (s, 1H), 8.42 (d, J = 2.2 Hz, 1H), 8.25 (d, J = 3.9 Hz, 1H), 8.03 (d, J = 1.5 Hz, 1H), 7.74–7.69 (m, 3H); 13C NMR (151 MHz, DMSO-d6): δ 160.05, 148.68, 146.47, 142.51, 140.00, 138.63, 137.29, 133.38, 132.07, 131.84, 131.31, 131.18, 127.32, 126.36, 125.99. EI/MS m/z (%): 351.2 [M+H]+; [M‒2Cl] = 279.3. Analitically calculated for C15H9Cl2N3OS: C, 51.44; H, 2.59; N, 12.00. Found: C, 51.47; H, 2.62; N, 11.92.
Compound 5-(3-acetylphenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4c). Light brown solid, MP = 171–173 °C (60% yield, 137 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.25 (s, 1H), 9.42 (d, J = 1.2 Hz, 1H), 8.47 (d, J = 31.2 Hz, 2H), 8.17 (dd, J = 3.6, 1.2 Hz, 1H), 7.90–7.86 (m, 2H), 7.76 (t, J = 8.4 Hz, 1H) 7.62–7.57 (m, 2H), 2.43 (s, 3H); 13C NMR (151 MHz, DMSO-d6): δ 197.95, 160.17, 147.56, 142.59, 139.93, 138.31, 136.96, 134.86, 133.87, 132.41, 130.89, 129.42, 124.20, 27.26. EI/MS m/z (%): 324.4 [M + H]+; [M‒COCH3] = 280.1. Analitically calculated for C17H13N3O2S: C, 63.14; H, 4.05; N, 12.99. Found: C, 63.19; H, 4.08; N, 12.97.
Compound 5-(4-chlorophenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4d). Off-white solid, MP = 249–251 °C (68% yield, 150 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.25 (s, 1H), 9.39 (s, 1H), 8.47 (s, 1H), 8.41 (d, J = 2.1 Hz, 1H), 8.25 (d, J = 3.8 Hz, 1H), 7.77 (d, J = 8.3 Hz, 2H), 7.64 (d, J = 3.8 Hz, 1H), 7.51 (d, J = 8.3 Hz, 2H); 13C NMR (151 MHz, DMSO-d6): δ 160.16, 148.74, 148.15, 142.50, 139.94, 139.89, 137.84, 137.32, 133.40, 131.94, 130.77, 129.23, 128.38, 127.53, 125.36. EI/MS m/z (%): 316.8 [M + H]+; [M‒Cl] = 280.1. Analitically calculated for C15H10ClN3OS: C, 57.05; H, 3.19; N, 13.31. Found: C, 57.09; H, 3.23; N, 13.28.
Compound 5-(3-chloro-4-fluorophenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4e). Light yellow solid, MP = 229–231 °C (50% yield, 116 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.27 (s, 1H), 9.38 (s, 1H), 8.48 (d, J = 1.6 Hz, 1H), 8.42 (d, J = 2.4 Hz, 1H), 8.25 (d, J = 4.0 Hz, 1H), 8.01 (dd, J = 6.9, 2.1 Hz, 1H), 7.76 (ddd, J = 8.4, 4.4, 2.3 Hz, 1H), 7.68 (d, J = 4.0 Hz, 1H), 7.51 (t, J = 8.9 Hz, 1H); 13C NMR (151 MHz, DMSO-d6): δ 160.12, 158.13, 156.47, 148.71, 146.81, 142.53, 140.00, 137.30, 131.86, 130.78, 127.75, 126.69, 125.93, 120.62, 117.83. EI/MS m/z (%): 334.8 [M + H]+; [M‒F, Cl] = 279.3. Analitically calculated for C15H9ClFN3OS: C, 53.98; H, 2.72; N, 12.59. Found: C, 54.05; H, 2.73; N, 12.56.
Compound 5-(3,5-dimethylphenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4f). Light blue solid, MP = 154–156 °C (62% yield, 135 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.18 (s, 1H), 9.40 (s, 1H), 8.44 (s, 1H), 8.39 (s, 1H), 8.23 (d, J = 3.8 Hz, 1H), 7.52 (d, J = 3.8 Hz, 1H), 7.32 (s, 2H), 6.97 (s, 1H), 2.28 (s, 6H); 13C NMR (151 MHz, DMSO-d6): δ 160.26, 150.11, 148.83, 142.40, 139.76, 138.33, 137.24, 136.96, 136.05, 135.85, 132.66, 131.89, 130.29, 124.33, 123.55, 20.75. EI/MS m/z (%): 310.4 [M + H]+; [M‒2CH3] = 279.1. Analitically calculated for C17H15N3OS: C, 66.00; H, 4.89; N, 13.58. Found: C, 66.06; H, 4.92; N, 13.55.
Compound 5-(4-methoxyphenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4g). Light brown solid, MP = 110–112 °C (57% yield, 125 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.19 (s, 1H), 9.33 (d, J = 1.8 Hz, 1H), 8.46 (d, J = 1.8 Hz, 1H), 8.41 (d, J = 2.4 Hz, 1H), 8.24–8.20 (m, 2H), 7.66 (d, J = 8.4 Hz, 2H), 7.33 (d, J = 8.4 Hz, 2H), 2.50 (s, 3H); 13C NMR (151 MHz, DMSO-d6): δ 160.23, 158.85, 148.89, 147.99, 142.38, 139.81, 138.32, 137.18, 135.49, 132.18, 130.39, 127.47, 113.17, 53.93. EI/MS m/z (%): 312.4 [M + H]+; [M‒OCH3] = 280.1. Analitically calculated for C16H13N3O2S: C, 61.72; H, 4.21; N, 13.50. Found: C, 61.78; H, 4.24; N, 13.48.
Compound methyl 4-(5-(pyrazin-2-ylcarbamoyl)thiophen-2-yl)benzoate (4h). Brown solid, MP = 144–146 °C (46% yield, 110 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.21 (s, 1H), 9.38 (d, J = 1.2 Hz, 1H), 8.48 (s, 1H), 8.41 (d, J = 3.0 Hz, 1H), 8.26 (d, J = 4.8 Hz, 1H), 7.94 (d, J = 4.8 Hz, 1H), 7.78 (d, J = 8.4 Hz, 2H), 7.24 (d, J = 8.4 Hz, 2H), 3.89 (s, 3H); 13C NMR (151 MHz, DMSO-d6): δ 161.56, 160.45, 148.79, 142.51, 139.91, 138.55, 137.32, 133.23, 131.47, 130.78, 129.99, 128.40, 127.72, 115.08, 53.81. EI/MS m/z (%): 340.4 [M + H]+; [M‒COOCH3] = 280.1. Analitically calculated for C17H13N3O3S: C, 60.17; H, 3.86; N, 12.38. Found: C, 60.24; H, 3.90; N, 12.35.
Compound 5-(4-(methylthio)phenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4i). Light brown solid, MP = 229–231 °C (52% yield, 120 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.22 (s, 1H), 9.38 (d, J = 1.1 Hz, 1H), 8.48–8.47 (m, 1H), 8.41 (d, J = 2.4 Hz, 1H), 8.24 (d, J = 4.0 Hz, 1H), 7.70 (d, J = 8.4 Hz, 2H), 7.60 (d, J = 4.0 Hz, 1H), 7.33 (d, J = 8.4 Hz, 2H), 2.52 (s, 3H); 13C NMR (151 MHz, DMSO-d6): δ 160.25, 149.40, 148.82, 142.52, 139.87, 139.54, 137.29, 136.78, 132.00, 129.27, 126.28, 126.23, 124.33, 14.43. EI/MS m/z (%): 328.4 [M + H]+; [M‒SCH3] = 280.1. Analitically calculated for C16H13N3OS2: C, 58.69; H, 4.00; N, 12.83. Found: C, 58.75; H, 4.02; N, 12.80.
Compound 5-(3,5-difluorophenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4j). Brown solid, MP = 239–241 °C (58% yield, 130 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.28 (s, 1H), 9.37 (d, J = 1.4 Hz, 1H), 8.47 (dd, J = 2.4, 1.6 Hz, 1H), 8.26 (d, J = 4.0 Hz, 1H), 7.76 (d, J = 4.0 Hz, 1H), 7.63–7.60 (m, 1H), 7.52 (dd, J = 8.4, 1.9 Hz, 2H), 7.26 (tt, J = 9.2 Hz, 2.1 Hz, 1H); 13C NMR (151 MHz, DMSO-d6): δ 163.70, 162.07, 160.06, 148.67, 146.58, 142.53, 140.03, 138.90, 137.29, 136.12, 131.70, 128.67, 126.74, 109.13, 103.93. EI/MS m/z (%): 318.4 [M + H]+; [M‒2F] = 279.1. Analitically calculated for C15H9F2N3OS: C, 56.78; H, 2.86; N, 13.24. Found: C, 56.88; H, 2.89; N, 13.20.
Compound 5′-chloro-N-(pyrazin-2-yl)-2,2′-bithiophene-5-carboxamide (4k). Off-white solid, MP = 180–182 °C (39% yield, 90 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.26 (s, 1H), 9.42 (d, J = 3.0 Hz, 1H), 8.49 (d, J = 1.8 Hz, 1H), 8.44 (d, J = 2.4 Hz, 1H), 8.28 (d, J = 4.2 Hz, 1H), 7.52 (d, J = 4.2 Hz, 1H), 7.30 (d, J = 6.0 Hz, 1H), 7.12 (d, J = 6.0 Hz, 1H), 2.47 (s, 3H); 13C NMR (151 MHz, DMSO-d6): δ 160.22, 148.76, 147.63, 143.45, 141.95, 141.14, 138.07, 136.97, 134.24, 129.84, 128.46, 126.32, 124.84. EI/MS m/z (%): 322.8 [M + H]+; [M‒Cl] = 286.0. Analitically calculated for C13H8ClN3OS2: C, 48.52; H, 2.51; N, 13.06. Found: C, 48.58; H, 2.53; N, 13.00.
Compound 5′-methyl-N-(pyrazin-2-yl)-2,2′-bithiophene-5-carboxamide (4l). Brown solid, MP = 200–202 °C (43% yield, 96 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.21 (s, 1H), 9.37 (d, J = 1.4 Hz, 1H), 8.47 (dd, J = 2.4, 1.6 Hz, 1H), 8.41 (d, J = 2.5 Hz, 1H), 8.18 (d, J = 4.0 Hz, 1H), 7.30 (dd, J = 9.1, 3.8 Hz, 2H), 6.86–6.82 (m, 1H), 2.47 (s, 3H); 13C NMR (151 MHz, DMSO-d6): δ 160.11, 148.77, 143.29, 142.51, 141.00, 139.87, 137.25, 135.99, 133.15, 131.86, 127.00, 125.82, 123.98, 14.97. EI/MS m/z (%): 302.4 [M + H]+; [M‒SCH3] = 286.0. Analitically calculated for C14H11N3OS2: C, 55.79; H, 3.68; N, 13.94. Found: C, 55.86; H, 3.63; N, 13.96.
Compound 5-(3,5-bis(trifluoromethyl)phenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4m). Brown solid, MP = 203–205 °C (48% yield, 140 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.28 (s, 1H), 9.09 (s, 1H), 8.61 (s, 1H), 8.49–8.45 (m, 2H), 8.27 8.02 (d, J = 1.8 Hz, 2H), 8.13 7.89 (d, J = 8.4 Hz, 1H), 8.10 7.71 (d, J = 8.4 Hz, 1H); 13C NMR (151 MHz, DMSO-d6): δ 160.18, 148.73, 147.05, 142.49, 139.93, 138.29, 134.47, 134.01, 133.86, 132.41, 130.19, 129.03, 125.93, 119.97. EI/MS m/z (%): 418.4 [M + H]+; [M‒6F] = 303.1. Analitically calculated for C17H9F6N3OS: C, 48.93; H, 2.17; N, 10.07. Found: C, 48.99; H, 2.19; N, 10.11.
Compound 5-(3-chloro-5-(trifluoromethyl)phenyl)-N-(pyrazin-2-yl)thiophene-2-carboxamide (4n). Brown solid, MP = 170–172 °C (37% yield, 100 mg). 1H NMR (600 MHz, DMSO-d6): δ 11.24 (s, 1H), 9.13 (s, 1H), 8.54–8.48 (m, 3H), 8.04–8.00 (m, 2H), 7.85 (d, J = 7.8 Hz, 1H), 8.01 7.66 (d, J = 7.8 Hz, 1H); 13C NMR (151 MHz, DMSO-d6): δ 160.16, 148.77, 147.76, 142.51, 139.89, 138.31, 137.29, 135.41, 134.89, 134.06, 132.21, 130.43, 129.23, 124.87, 119.93. EI/MS m/z (%): 384.8 [M + H]+; [M‒3F] = 326.0, [M‒Cl] = 348.0. Analitically calculated for C16H9ClF3N3OS: C, 50.07; H, 2.36; N, 10.95. Found: C, 50.15; H, 2.38; N, 10.89.

4. Conclusions

In this study, 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3) was synthesized by means of TiCl4-mediated one-pot condensation of pyrazin-2-amine (1) and 5-bromothiophene-2-carboxylic acid (2) with a good 75% yield followed by a Suzuki cross-coupling reaction with various aryl/heteroaryl boronic acids/pinacol esters to synthesize 5-aryl-N-(pyrazin-2-yl)thiophene-2-carboxamides (4a4n) obtained with moderate and good yields (37–72%). The target pyrazine analogs (4a4n) were confirmed by NMR and mass spectrometry. The computational studies of these final compounds were also performed to get optimized geometries and thermodynamic parameters such as FMOs (EHOMO, ELUMO), HOMO–LUMO energy gap, electron affinity (A), ionization energy (I), electrophilicity index (ω), chemical softness (σ) and chemical hardness (η). The stability and NLO behavior of all the compounds 4a4n were studied with the help of the HOMO–LUMO energy gap and hyperpolarizability calculations. Compound 4i was found to be the most reactive compound with the highest electronic chemical potential value (−3.88 eV) and compound 4l had the highest hyperpolarizability value of 8583.80 Hartrees. The theoretical calculated chemical shifts of 1H NMR of all compounds 4a4n match very well the experimental values.

Supplementary Materials

The following are available online. Theoratical 1HNMR (Tables S1–S13), the 1HNMR and 13CNMR (Figures S1–S17), Potential energy Scan (Figures S18–S30) and lowest energy conformers are available online.

Author Contributions

Conceptualization, N.R. and G.A.; methodology, G.A., M.S.A. and N.R.; experiments, G.A.; software, A.M. and M.A.H.; formal analysis, G.A., A.M., M.A.H., M.S.A. and M.B.; investigation, G.A., N.R., S.H. and M.Z.; writing—original draft preparation M.B., S.H., G.A. and A.M.; writing—review and editing, G.A., N.R., M.H., M.A.H., S.H. and M.B.; supervision, N.R.; project administration, N.R. and A.F.Z.; funding acquisition, M.S.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors sincerely appreciate funding from the Researchers Supporting Project number (RSP-2021/399), King Saud University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Dubuisson, M.; Rees, J.-F.; Marchand-Brynaert, J. Discovery and validation of a new family of antioxidants: The aminopyrazine derivatives. Mini Rev. Med. Chem. 2004, 4, 421–435. [Google Scholar] [CrossRef] [Green Version]
  2. Higasio, Y.S.; Shoji, T. Heterocyclic compounds such as pyrroles, pyridines, pyrollidins, piperdines, indoles, imidazol and pyrazins. Appl. Catal. A 2001, 221, 197–207. [Google Scholar] [CrossRef]
  3. Kamal, A.; Reddy, J.S.; Ramaiah, M.J.; Dastagiri, D.; Bharathi, E.V.; Sagar, M.V.P.; Pushpavalli, S.; Ray, P.; Pal-Bhadra, M. Design, synthesis and biological evaluation of imidazopyridine/pyrimidine-chalcone derivatives as potential anticancer agents. MedChemComm 2010, 1, 355–360. [Google Scholar] [CrossRef]
  4. Ghosh, P.; Rasul, M.; Chakraborty, M.; Mandal, A.; Saha, A. Microwave assisted one-pot synthesis of pyrazine derivatives of pentacyclic triterpenoids and their biological activity. Indian J. Chem. 2011, 50B, 1519–1523. [Google Scholar] [CrossRef]
  5. Meher, C.; Rao, A.; Omar, M. Piperazine-pyrazine and their multiple biological activities. Asian J. Pharm. Sci. 2013, 3, 43–60. [Google Scholar]
  6. Bonde, C.G.; Gaikwad, N.J. Synthesis and preliminary evaluation of some pyrazine containing thiazolines and thiazolidinones as antimicrobial agents. Bioorg. Med. Chem. 2004, 12, 2151–2161. [Google Scholar] [CrossRef] [PubMed]
  7. Sriram, D.; Yogeeswari, P.; Reddy, S.P. Synthesis of pyrazinamide Mannich bases and its antitubercular properties. Bioorg. Med. Chem. Lett. 2006, 16, 2113–2116. [Google Scholar] [CrossRef]
  8. Vergara, F.M.; Lima, C.H.d.S.; das Graças M. de O. Henriques, M.; Candéa, A.L.; Lourenço, M.C.; Ferreira, M.d.L.; Kaiser, C.R.; de Souza, M.V. Synthesis and antimycobacterial activity of N′-[(E)-(monosubstituted-benzylidene)]-2-pyrazinecarbohydrazide derivatives. Eur. J. Med. Chem. 2009, 44, 4954–4959. [Google Scholar] [CrossRef] [PubMed]
  9. Abdel-Aziz, M.; Abdel-Rahman, H.M. Synthesis and anti-mycobacterial evaluation of some pyrazine-2-carboxylic acid hydrazide derivatives. Eur. J. Med. Chem. 2010, 45, 3384–3388. [Google Scholar] [CrossRef] [PubMed]
  10. Czerny, M.; Mayer, F.; Grosch, W. Sensory study on the character impact odorants of roasted Arabica coffee. J. Agric. Food Chem. 1999, 47, 695–699. [Google Scholar] [CrossRef]
  11. Cai, L.; Sun, Y.; Song, Y.; Xu, L.; Bei, Z.; Zhang, D.; Dou, Y.; Wang, H. Viral polymerase inhibitors T-705 and T-1105 are potential inhibitors of Zika virus replication. Arch. Virol. 2017, 162, 2847–2853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Furuta, Y.; Takahashi, K.; Shiraki, K.; Sakamoto, K.; Smee, D.F.; Barnard, D.L.; Gowen, B.B.; Julander, J.G.; Morrey, J.D. T-705 (favipiravir) and related compounds: Novel broad-spectrum inhibitors of RNA viral infections. Antivir. Res. 2009, 82, 95–102. [Google Scholar] [CrossRef]
  13. Baranovich, T.; Wong, S.-S.; Armstrong, J.; Marjuki, H.; Webby, R.J.; Webster, R.G.; Govorkova, E.A. T-705 (favipiravir) induces lethal mutagenesis in influenza A H1N1 viruses in vitro. J. Virol. 2013, 87, 3741–3751. [Google Scholar] [CrossRef] [Green Version]
  14. Du, Y.X.; Chen, X.P. Favipiravir: Pharmacokinetics and concerns about clinical trials for 2019-nCoV infection. Clin. Pharmacol. Ther. 2020, 108, 242–247. [Google Scholar] [CrossRef] [Green Version]
  15. Mallesha, L.; Mohana, K.N. Synthesis, antimicrobial and antioxidant activities of 1-(1,4-benzodioxane-2-carbonyl) piperazine derivatives. Eur. J. Chem. 2011, 2, 193–199. [Google Scholar] [CrossRef] [Green Version]
  16. Dolezal, M.; Zitko, J.; Osicka, Z.; Kunes, J.; Vejsova, M.; Buchta, V.; Dohnal, J.; Jampilek, J.; Kralova, K. Synthesis, antimycobacterial, antifungal and photosynthesis-inhibiting activity of chlorinated N-phenylpyrazine-2-carboxamides. Molecules 2010, 15, 8567–8581. [Google Scholar] [CrossRef]
  17. Satoskar, R.S.; Bhandarkar, S. Pharmacology and Pharmacotherapeutics; Elsevier India: Kolkata, India, 2020. [Google Scholar]
  18. Zhang, Y.; Wade, M.M.; Scorpio, A.; Zhang, H.; Sun, Z. Mode of action of pyrazinamide: Disruption of Mycobacterium tuberculosis membrane transport and energetics by pyrazinoic acid. J. Antimicrob. Chemother. 2003, 52, 790–795. [Google Scholar] [CrossRef]
  19. Dorn, R.; Baums, D.; Kersten, P.; Regener, R. Nonlinear optical materials for integrated optics: Telecommunications and sensors. Adv. Mater. 1992, 4, 464–473. [Google Scholar] [CrossRef]
  20. Marder, S.R.; Perry, J.W. Molecular materials for second-order nonlinear optical applications. Adv. Mater. 1993, 5, 804–815. [Google Scholar] [CrossRef]
  21. Dini, D.; Calvete, M.J.; Hanack, M. Nonlinear optical materials for the smart filtering of optical radiation. Chem. Rev. 2016, 116, 13043–13233. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, J.; Ouyang, C.; Huo, F.; He, W.; Cao, A. Progress in the enhancement of electro-optic coefficients and orientation stability for organic second-order nonlinear optical materials. Dyes Pigments 2020, 181, 108509. [Google Scholar] [CrossRef]
  23. Bureš, F. Fundamental aspects of property tuning in push–pull molecules. RSC Adv. 2014, 4, 58826–58851. [Google Scholar] [CrossRef] [Green Version]
  24. Meti, P.; Park, H.-H.; Gong, Y.-D. Recent developments in pyrazine functionalized π-conjugated materials for optoelectronic applications. J. Mater. Chem. C 2020, 8, 352–379. [Google Scholar] [CrossRef]
  25. Achelle, S.; Baudequin, C.; Plé, N. Luminescent materials incorporating pyrazine or quinoxaline moieties. Dyes Pigments 2013, 98, 575–600. [Google Scholar] [CrossRef] [Green Version]
  26. Hebbar, N.; Ramondenc, Y.; Plé, G.; Dupas, G.; Plé, N. Push–pull structures with a pyrazine core and hexatriene chain: Synthesis and light-emitting properties. Tetrahedron 2009, 65, 4190–4200. [Google Scholar] [CrossRef]
  27. Coe, B.J.; Fielden, J.; Foxon, S.P.; Asselberghs, I.; Clays, K.; Brunschwig, B.S. Two-dimensional, pyrazine-based nonlinear optical chromophores with ruthenium (II) ammine electron donors. Inorg. Chem. 2010, 49, 10718–10726. [Google Scholar] [CrossRef]
  28. Dokládalová, L.; Bureš, F.; Kuznik, W.; Kityk, I.V.; Wojciechowski, A.; Mikysek, T.; Almonasy, N.; Ramaiyan, M.; Padělková, Z.; Kulhánek, J. Dicyanobenzene and dicyanopyrazine derived X-shaped charge-transfer chromophores: Comparative and structure–property relationship study. Org. Biomol. Chem. 2014, 12, 5517–5527. [Google Scholar] [CrossRef] [Green Version]
  29. Sun, Y.; Liu, X.T.; Guo, J.F.; Ren, A.M.; Wang, D. Theoretical investigation of two-photon absorption and fluorescence properties of cypridina luciferin-based derivatives: 2, 3, 5-trisubstituted pyrazine compounds. J. Phys. Org. Chem. 2013, 26, 822–833. [Google Scholar] [CrossRef]
  30. Li, H.; Zhang, Y.; Bi, Z.; Xu, R.; Li, M.; Shen, X.; Tang, G.; Han, K. Theoretical study on the spectroscopic and third-order nonlinear optical properties of two-dimensional charge-transfer pyrazine derivatives. Mol. Phys. 2017, 115, 3164–3171. [Google Scholar] [CrossRef]
  31. Chen, M.; Nie, H.; Song, B.; Li, L.; Sun, J.Z.; Qin, A.; Tang, B.Z. Triphenylamine-functionalized tetraphenylpyrazine: Facile preparation and multifaceted functionalities. J. Mater. Chem. C 2016, 4, 2901–2908. [Google Scholar] [CrossRef] [Green Version]
  32. Imran, H.M.; Rasool, N.; Kanwal, I.; Hashmi, M.A.; Altaf, A.A.; Ahmed, G.; Malik, A.; Kausar, S.; Khan, S.U.-D.; Ahmad, A. Synthesis of halogenated [1,1′-biphenyl]-4-yl benzoate and [1,1′:3′,1″-terphenyl]-4′-yl benzoate by palladium catalyzed cascade C–C coupling and structural analysis through computational approach. J. Mol. Struct. 2020, 1222, 1–9. [Google Scholar] [CrossRef]
  33. Lewars, E. Introduction to the theory and applications of molecular and quantum mechanics. In Computational Chemistry; Springer: Berlin, Germany, 2003; Volume 318. [Google Scholar]
  34. Malik, A.; Rasool, N.; Kanwal, I.; Hashmi, M.A.; Zahoor, A.F.; Ahmad, G.; Altaf, A.A.; Shah, S.A.A.; Sultan, S.; Zakaria, Z.A. Suzuki–miyaura reactions of (4-bromophenyl)-4,6-dichloropyrimidine through commercially available palladium catalyst: Synthesis, optimization and their structural aspects identification through computational studies. Processes 2020, 8, 1342. [Google Scholar] [CrossRef]
  35. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef] [PubMed]
  36. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView 6.0.16; Semichem Inc.: Shawnee Mission, KS, USA, 2016. [Google Scholar]
  37. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09 Revision D. 01; Gaussian Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  38. Schlegel, H.B. Geometry optimization. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 790–809. [Google Scholar] [CrossRef]
  39. Singh, R.K.; Singh, A.K.; Siddiqui, S.; Arshad, M.; Jafri, A. Synthesis, molecular structure, spectral analysis and cytotoxic activity of two new aroylhydrazones. J. Mol. Struct. 2017, 1135, 82–97. [Google Scholar] [CrossRef]
  40. Gökce, H.; Öztürk, N.; Kazıcı, M.; Göreci, Ç.Y.; Güneş, S. Structural, spectroscopic, electronic, nonlinear optical and thermodynamic properties of a synthesized Schiff base compound: A combined experimental and theoretical approach. J. Mol. Struct. 2017, 1136, 288–302. [Google Scholar] [CrossRef]
  41. Kanaani, A.; Ajloo, D.; Grivani, G.; Ghavami, A.; Vakili, M. Tautomeric stability, molecular structure, NBO, electronic and NMR analyses of salicylideneimino-ethylimino-pentan-2-one. J. Mol. Struct. 2016, 1112, 87–96. [Google Scholar] [CrossRef]
  42. O’Brien, S.E.; Browne, H.L.; Bradshaw, T.D.; Westwell, A.D.; Stevens, M.F.; Laughton, C.A. Antitumor benzothiazoles. Frontier molecular orbital analysis predicts bioactivation of 2-(4-aminophenyl) benzothiazoles to reactive intermediates by cytochrome P4501A1. Org. Biomol. Chem. 2003, 1, 493–497. [Google Scholar] [CrossRef]
  43. Arshad, M.N.; Bibi, A.; Mahmood, T.; Asiri, A.M.; Ayub, K. Synthesis, crystal structures and spectroscopic properties of triazine-based hydrazone derivatives; a comparative experimental-theoretical study. Molecules 2015, 20, 5851–5874. [Google Scholar] [CrossRef] [Green Version]
  44. Belletête, M.; Beaupré, S.; Bouchard, J.; Blondin, P.; Leclerc, M.; Durocher, G. Theoretical and experimental investigations of the spectroscopic and photophysical properties of fluorene-phenylene and fluorene-thiophene derivatives: Precursors of light-emitting polymers. J. Phys. Chem. B 2000, 104, 9118–9125. [Google Scholar] [CrossRef]
  45. Peng, W.; Peiwang, Z.; Chuanguang, W.; Cheng, Y. Theoretical investigation and molecular design of pyrazine derivatives with large hyperpolarizabilities (β). J. Mol. Struct. Theor. Chem 1999, 459, 155–162. [Google Scholar] [CrossRef]
  46. Ceylan, Ü.; Tarı, G.Ö.; Gökce, H.; Ağar, E. Spectroscopic (FT–IR and UV–Vis) and theoretical (HF and DFT) investigation of 2-Ethyl-N-[(5-nitrothiophene-2-yl) methylidene] aniline. J. Mol. Struct. 2016, 1110, 1–10. [Google Scholar] [CrossRef]
  47. Mathammal, R.; Sangeetha, K.; Sangeetha, M.; Mekala, R.; Gadheeja, S. Molecular structure, vibrational, UV, NMR, HOMO-LUMO, MEP, NLO, NBO analysis of 3, 5 di tert butyl 4 hydroxy benzoic acid. J. Mol. Struct. 2016, 1120, 1–14. [Google Scholar] [CrossRef]
  48. Politzer, P.; Murray, J.S. The fundamental nature and role of the electrostatic potential in atoms and molecules. Theor. Chem. Acc. 2002, 108, 134–142. [Google Scholar] [CrossRef]
  49. Tsuneda, T.; Song, J.-W.; Suzuki, S.; Hirao, K. On Koopmans’ theorem in density functional theory. J. Chem. Phys. 2010, 133, 174101. [Google Scholar] [CrossRef]
  50. Guo, L.; Safi, Z.S.; Kaya, S.; Shi, W.; Tüzün, B.; Altunay, N.; Kaya, C. Anticorrosive effects of some thiophene derivatives against the corrosion of iron: A computational study. Front. Chem. 2018, 6, 155. [Google Scholar] [CrossRef] [PubMed]
  51. Martínez, J. Local reactivity descriptors from degenerate frontier molecular orbitals. Chem. Phys. Lett. 2009, 478, 310–322. [Google Scholar] [CrossRef]
  52. Mubarik, A.; Rasool, N.; Hashmi, M.A.; Mansha, A.; Zubair, M.; Shaik, M.R.; Sharaf, M.A.; Awwad, E.M.; Abdelgawad, A. Computational Study of Structural, Molecular Orbitals, Optical and Thermodynamic Parameters of Thiophene Sulfonamide Derivatives. Crystals 2021, 11, 211. [Google Scholar] [CrossRef]
  53. Leggio, A.; Bagalà, J.; Belsito, E.L.; Comandè, A.; Greco, M.; Liguori, A. Formation of amides: One-pot condensation of carboxylic acids and amines mediated by TiCl4. Chem. Cent. J. 2017, 11, 87–98. [Google Scholar] [CrossRef] [Green Version]
  54. Ahmad, G.; Rasool, N.; Rizwan, K.; Imran, I.; Zahoor, A.F.; Zubair, M.; Sadiq, A.; Rashid, U. Synthesis, in-vitro cholinesterase inhibition, in-vivo anticonvulsant activity and in-silico exploration of N-(4-methylpyridin-2-yl) thiophene-2-carboxamide analogs. Bioorg. Chem. 2019, 92, 103216. [Google Scholar] [CrossRef]
  55. Miyaura, N.; Suzuki, A. Palladium-catalyzed cross-coupling reactions of organoboron compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef] [Green Version]
  56. Ahmad, G.; Rasool, N.; Ikram, H.M.; Gul Khan, S.; Mahmood, T.; Ayub, K.; Zubair, M.; Al-Zahrani, E.; Ali Rana, U.; Akhtar, M.N. Efficient synthesis of novel pyridine-based derivatives via Suzuki cross-coupling reaction of commercially available 5-bromo-2-methylpyridin-3-amine: Quantum mechanical investigations and biological activities. Molecules 2017, 22, 190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ikram, H.; Rasool, N.; Ahmad, G.; Chotana, G.; Musharraf, S.; Zubair, M.; Rana, U.; Haq, M.Z.U.; Jaafar, H. Selective C-arylation of 2, 5-dibromo-3-hexylthiophene via suzuki cross coupling reaction and their pharmacological aspects. Molecules 2015, 20, 5202–5214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Ahmad, G.; Rasool, N.; Rizwan, K.; Altaf, A.A.; Rashid, U.; Mahmood, T.; Ayub, K. Role of Pyridine Nitrogen in Palladium-Catalyzed Imine Hydrolysis: A Case Study of (E)-1-(3-bromothiophen-2-yl)-N-(4-methylpyridin-2-yl) methanimine. Molecules 2019, 24, 2609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ahmad, G.; Rasool, N.; Qamar, M.U.; Alam, M.M.; Kosar, N.; Mahmood, T.; Imran, M. Facile synthesis of 4-aryl-N-(5-methyl-1H-pyrazol-3-yl) benzamides via Suzuki Miyaura reaction: Antibacterial activity against clinically isolated NDM-1-positive bacteria and their Docking Studies. Arab. J. Chem. 2021, 14, 1–11. [Google Scholar] [CrossRef]
  60. Mahmood, N.; Rasool, N.; Ikram, H.M.; Hashmi, M.A.; Mahmood, T.; Zubair, M.; Ahmad, G.; Rizwan, K.; Rashid, T.; Rashid, U. Synthesis of 3, 4-Biaryl-2, 5-Dichlorothiophene through Suzuki Cross-Coupling and Theoretical Exploration of Their Potential Applications as Nonlinear Optical Materials. Symmetry 2018, 10, 766. [Google Scholar] [CrossRef] [Green Version]
  61. Rizwan, K.; Rasool, N.; Rehman, R.; Mahmood, T.; Ayub, K.; Rasheed, T.; Ahmad, G.; Malik, A.; Khan, S.A.; Akhtar, M.N. Facile synthesis of N-(4-bromophenyl)-1-(3-bromothiophen-2-yl) methanimine derivatives via Suzuki cross-coupling reaction: Their characterization and DFT studies. Chem. Cent. J. 2018, 12, 84–92. [Google Scholar] [CrossRef] [PubMed]
  62. Ikram, H.M.; Rasool, N.; Hashmi, M.A.; Anjum, M.A.; Ali, K.G.; Zubair, M.; Ahmad, G.; Mahmood, T. Density functional theory-supported studies of structural and electronic properties of substituted-phenol derivatives synthesized by efficient O-or C-arylation via Chan--Lam or Suzuki cross-coupling reactions. Turk. J. Chem. 2019, 43, 1306–1321. [Google Scholar] [CrossRef]
  63. Abbas, M.; Rizwan, K.; Rasool, N.; Hashmi, M.A.; Ahmad, G.; Rashid, U.; Shah, S.A.A. Palladium (0) catalyzed synthesis of thiophene based 1, 3, 4-oxadiazoles their reactivities and potential nonlinear optical properties. Chiang Mai J. Sci. 2020, 47, 1255–1264. [Google Scholar]
Figure 1. RNA viral polymerase inhibitors T-705 and T-1105.
Figure 1. RNA viral polymerase inhibitors T-705 and T-1105.
Molecules 26 07309 g001
Scheme 1. Synthesis and arylation of 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3) to form analogs (4a4n).
Scheme 1. Synthesis and arylation of 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide (3) to form analogs (4a4n).
Molecules 26 07309 sch001
Figure 2. An overview of 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide derivatives via a Suzuki cross-coupling reaction.
Figure 2. An overview of 5-bromo-N-(pyrazin-2-yl)thiophene-2-carboxamide derivatives via a Suzuki cross-coupling reaction.
Molecules 26 07309 g002
Figure 3. Optimized structures of all the final derivatives (4a4n). In the 3D models, the yellow color indicates S, the red color represents oxygen, the grey color indicates C, the green color is for Cl, the light blue color shows F and the white color symbolizes H.
Figure 3. Optimized structures of all the final derivatives (4a4n). In the 3D models, the yellow color indicates S, the red color represents oxygen, the grey color indicates C, the green color is for Cl, the light blue color shows F and the white color symbolizes H.
Molecules 26 07309 g003
Figure 4. 1H NMR spectra of compound 4a at 600 MHz.
Figure 4. 1H NMR spectra of compound 4a at 600 MHz.
Molecules 26 07309 g004
Figure 5. Molecular electrostatic potential surfaces of the final compounds 4a4n computed at the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory. The units for the scale are atomic units.
Figure 5. Molecular electrostatic potential surfaces of the final compounds 4a4n computed at the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory. The units for the scale are atomic units.
Molecules 26 07309 g005
Figure 6. FMO surfaces of compounds 4a4n computed at the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory.
Figure 6. FMO surfaces of compounds 4a4n computed at the PBE0-D3BJ/def2-TZVP/SMD1,4-dioxane level of theory.
Molecules 26 07309 g006
Table 1. Comparison of the experimental and theoretical 1H NMR data of compound 4a. The data for the rest of the compounds is provided in the Supplementary Materials. The mean absolute deviation is presented as the mean absolute error and, similarly, the root-mean-square deviation is denoted as the root-mean-square error.
Table 1. Comparison of the experimental and theoretical 1H NMR data of compound 4a. The data for the rest of the compounds is provided in the Supplementary Materials. The mean absolute deviation is presented as the mean absolute error and, similarly, the root-mean-square deviation is denoted as the root-mean-square error.
Compound 4a
Carbon No.Carbon Type1H NMR (δ, ppm), Experimental1H NMR (δ, ppm), ComputedΔδ, ppm
2C
3CH9.3910.55−1.16
4N
5CH8.458.430.02
6CH8.458.420.03
2′C
3′CH7.847.240.60
4′CH7.726.800.92
5′C
1″C
2″CH8.267.450.81
3″C
4″CH7.507.370.13
5″CH7.457.390.06
6″CH7.707.250.45
Mean absolute error (MAE) = 0.17; root-mean-square error (RMSE) = 0.37.
Table 2. Energies of the HOMO, LUMO, HOMO–LUMO energy gap, polarizability (α) and hyperpolarizability (βo) values of compounds 4a4n.
Table 2. Energies of the HOMO, LUMO, HOMO–LUMO energy gap, polarizability (α) and hyperpolarizability (βo) values of compounds 4a4n.
CompoundsEHOMO (eV)ELUMO (eV)ELUMO−EHOMO (eV)α (a.u.)βo (a.u.)
4a−6.781−1.9584.823310.962700.54
4b−6.770−2.0064.763325.224139.08
4c−6.771−2.0084.763338.092571.25
4d−6.702−1.9144.788311.833958.52
4e−6.756−1.9734.784310.604004.42
4f−6.345−2.0054.340354.106120.41
4g−6.229−1.7704.459330.848398.11
4h−6.818−2.0824.736345.571363.79
4i−6.027−1.8154.212349.439139.57
4j−6.870−1.9774.893293.781926.82
4k−6.475−2.1034.373310.216520.24
4l−6.231−2.0004.231325.489048.28
4m−6.985−2.0484.937315.571103.20
4n−6.919−2.1134.806322.841416.47
Table 3. Ionization potential (I) and electron affinity (A) of compounds 4a4n calculated through Koopman’s theorem and directly, i.e., through optimization of charged molecules.
Table 3. Ionization potential (I) and electron affinity (A) of compounds 4a4n calculated through Koopman’s theorem and directly, i.e., through optimization of charged molecules.
CompoundsCalculated through Koopman’s TheoremCalculated Directly from Charged Molecules
I (eV)A (eV)I (eV)A (eV)
4a6.781.966.160.47
4b6.772.016.230.50
4c6.772.017.190.06
4d6.701.916.160.44
4e6.761.976.230.50
4f6.352.005.750.83
4g6.231.775.880.28
4h6.822.087.200.42
4i6.031.815.670.37
4j6.871.987.290.52
4k6.482.106.210.75
4l6.232.005.790.73
4m6.992.056.610.58
4n6.922.116.360.60
Table 4. Reactivity descriptors η, σ, μ and ω values of the final compounds 4a4n.
Table 4. Reactivity descriptors η, σ, μ and ω values of the final compounds 4a4n.
Compoundsη (eV)σ (eV−1)μ (eV)ω (eV)
4a2.840.35−3.311.93
4b2.860.34−3.361.97
4c3.560.28−3.621.84
4d2.860.34−3.301.90
4e2.860.34−3.361.97
4f2.460.40−3.292.20
4g2.800.35−3.081.69
4h3.390.29−3.812.14
4i2.650.37−3.021.72
4j3.380.29−3.902.25
4k2.730.36−3.482.21
4l2.530.39−3.262.10
4m3.010.33−3.592.14
4n2.880.34−3.482.10
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ahmad, G.; Rasool, N.; Mubarik, A.; Zahoor, A.F.; Hashmi, M.A.; Zubair, M.; Bilal, M.; Hussien, M.; Akhtar, M.S.; Haider, S. Facile Synthesis of 5-Aryl-N-(pyrazin-2-yl)thiophene-2-carboxamides via Suzuki Cross-Coupling Reactions, Their Electronic and Nonlinear Optical Properties through DFT Calculations. Molecules 2021, 26, 7309. https://doi.org/10.3390/molecules26237309

AMA Style

Ahmad G, Rasool N, Mubarik A, Zahoor AF, Hashmi MA, Zubair M, Bilal M, Hussien M, Akhtar MS, Haider S. Facile Synthesis of 5-Aryl-N-(pyrazin-2-yl)thiophene-2-carboxamides via Suzuki Cross-Coupling Reactions, Their Electronic and Nonlinear Optical Properties through DFT Calculations. Molecules. 2021; 26(23):7309. https://doi.org/10.3390/molecules26237309

Chicago/Turabian Style

Ahmad, Gulraiz, Nasir Rasool, Adeel Mubarik, Ameer Fawad Zahoor, Muhammad Ali Hashmi, Muhammad Zubair, Muhammad Bilal, Mohamed Hussien, Muhammad Saeed Akhtar, and Sajjad Haider. 2021. "Facile Synthesis of 5-Aryl-N-(pyrazin-2-yl)thiophene-2-carboxamides via Suzuki Cross-Coupling Reactions, Their Electronic and Nonlinear Optical Properties through DFT Calculations" Molecules 26, no. 23: 7309. https://doi.org/10.3390/molecules26237309

Article Metrics

Back to TopTop