Skip to main content
ACS Omega logoLink to ACS Omega
. 2020 May 27;5(22):13236–13249. doi: 10.1021/acsomega.0c01273

Exploration of Noncovalent Interactions, Chemical Reactivity, and Nonlinear Optical Properties of Piperidone Derivatives: A Concise Theoretical Approach

Muhammad Khalid , Akbar Ali ‡,*, Muhammad Fayyaz Ur Rehman ‡,*, Muhammad Mustaqeem §, Shehbaz Ali , Muhammad Usman Khan , Sumreen Asim , Naseeb Ahmad , Muhammad Saleem
PMCID: PMC7288701  PMID: 32548510

Abstract

graphic file with name ao0c01273_0009.jpg

The organic compounds with a π-bond system lead to electric charge delocalization which enables them to reveal fascinating nonlinear optical properties. Mono-carbonyl curcuminoids also have an appealing skeleton from the conjugation view point. Interesting chemical structures of the 3,5-bis(arylidene)-N-benzenesulfonyl-4-piperidone derivatives motivated us to perform density functional theory (DFT)-based studies. Therefore, computations using the B3LYP/6-311G(d,p) functional of DFT were executed to explore geometric parameters, highest occupied molecular orbital (HOMO)–lowest unoccupied molecular orbital (LUMO) energies, and natural bond orbital (NBO) analyses. Moreover, three different functionals such as HF, B3LYP, and M06 with the 6-311G(d,p) basis set were used to investigate the average polarizability ⟨α⟩ and first hyperpolarizability (βtot)-based properties of all compounds. A good concurrence among calculated and experimental parameters was obtained through root mean square error calculations. The molecular stability of piperidone derivatives was examined using the Hirshfeld surface and NBO analyses. Natural population analysis was also performed to obtain insights about atomic charges. Calculated HOMO–LUMO energies showed that charge transfer interactions take place within the molecules. Moreover, global reactivity parameters including electronegativity, chemical hardness, softness, ionization potential, and electrophilicity were calculated using the HOMO and LUMO energies. The average polarizability ⟨α⟩ and first hyperpolarizability (βtot) values of all compounds were observed to be larger in magnitude at the aforesaid functional than the standard compound.

Introduction

Synthetic organic chemistry is a promising field to generate numerous exciting molecules by applying various protocols, in quick time with complete structural investigation, high purity, and mostly with an attractive yield. The beauty of this chemistry is the generation of highly demanding functional groups from easily manageable precursors. Sometimes, a simple condensation reaction ends in a valuable functional group with stimulating structural features indispensable from the biological activity point of view.1 Monocarbonyl curcuminoids are also among those functionalities that could be generated by the simple condensation reaction of aliphatic ketones and aromatic aldehydes. A curcuminoid is a linear diarylpentanoid having one carbonyl or a linear diarylheptanoid having two carbonyl groups.2 Traditionally, monocarbonyl curcuminoids have been investigated for their assembly and molecular possessions.3 Additionally, α,β-unsaturated carbonyl compounds like curcumin (Curcuma longa-based natural product) are vital from the medicinal standpoint with their potential to be antitumor, antioxidant, antibacterial, and anti-inflammatory agents. For instance, α,β-unsaturated ketone like functionalized 3,5-bis(arylidene)-4-piperidone have been reported to have promising antitumor and anti-inflammatory capabilities.4 Moreover, further tuning through N-alkylation of the skeleton of 3,5-bis(arylidene)-4-piperidone by incorporation of various synthons with different electronic nature can generate chemical architectures with improved biological activities. The N-cyclopropyl monocarbonyl analogs of curcumin MAC 7a and N-benzoyl BAP c6 are the potential anti-inflammatory agents. The 3,5-bis(arylidene)-4-piperidone analog BAP 6d is the potential antitumor mediator, while BAP c6 and BAP d6 in combination are outstanding anti-tumors and anti-inflammatory agents.4 Moreover, EF24 (3,5-bis(2-fluorobenzylidene)-4-piperidone) is also a potential antitumor agent and restrain the tumor expansion as well as the metastasis.5 Anti-cancer 4-[3,5-bis(2-chlorobenzylidene)-4-oxo-piperidine-1-yl]-4-oxo-2-butenoic acid (CLEFMA) has the ability to induce apoptosis.6 In addition to these, KalmanHideg and co-workers synthesized the N-substituted 3,5-bis(arylidene)-4-piperidone derivatives and explored their anti-cancer potential of cytotoxicity against cancer cells and linked it with molecular docking simulations.7

In this scenario, Li. et al.4 reported the four 3,5-bis(arylidene)-N-benzenesulfonyl-4-piperidone derivatives. All the four compounds were crystallized by slow evaporation of dichloromethane/methanol solutions at room temperature, and the structures of the final compounds were confirmed by NMR, FT-IR, HRMS, as well as by the single-crystal analysis. The single-crystal analysis revealed that compound (a) is monoclinic with space group P21/n, while the remaining three compounds (i.e., b, c, and d) were found to be triclinic in nature with space group P1. Interestingly, all the four compounds presented similar stereochemistry that is mainly because of the linear olefinic double bond generated at either side of the 3,5-bis(arylidene)-4-piperidone ring where both the 4-(trifluoromethyl)phenyl rings adopted the E stereochemistry presenting the E,E isomer. Moreover, regarding the conformation, compound a and b, the arylsulfonyl group showed pseudo-axial conformation, while in the case of c and d, the arylsulfonyl group adopted a pseudo-equatorial conformation with respect to the 4-pyridone ring. Accordingly, the four compounds were tested for biological activity revealing that these compounds are of significant anti-inflammatory nature.

The starring role of density functional theory (DFT) in understanding the modern chemistry has gained enormous attention. These DFT-based calculations play a pivotal role in collecting key structural information as it helps in finding the insight of the reaction mechanism, stability, reacting sites, electronic properties, and so forth.8 Moreover, nonlinear optical properties (NLOs) can also be accomplished using DFT.913 NLO features can be explored at different ranges of wavelengths from the deep infrared to extreme UV region that provide valuable information which can be used in the field of information technology and telecommunication.14 Currently, many of the valuable phenomena such as quantum computing, quantum optics, particle accelerators, and ultra-cold atoms, and so forth are associated with NLO.15,16 NLO plays a key role in the diversification of lasers together with coherent light as the NLO practice may change the laser light. Interestingly, it might control the frequency and/or spatial features of the laser output. Material interaction is another important feature of NLO that has valuable applications in machining, spectroscopy, and in analysis tools.17,18 Information technology is also based on NLO features in all areas like telecom, data storage, sensors, and signal processing.19 Our research group is interested to find the NLO properties of the promising crystalline organic compounds (piperidone derivatives) reported by Li and co-workers.4 According to the best of our knowledge, neither electronic, hyper, conjugative interactions nor NLO properties of 3,5-bis(arylidene)-N-benzenesulfonyl-4-piperidone derivatives have been reported so far. Herein, DFT calculations are used to provide insights into electronic and NLO properties and structural parameters, which have been compared to experimental data for better understanding of chemical structures.

Computational Methods and Materials

The inclusive quantum chemical calculations for piperidin-4-one derivatives: (3Z,5Z)-1-tosyl-3,5-bis (4-(trifluoromethyl)benzylidene)piperidin-4-one (TBTBP), (3Z,5Z)-1-((4-fluorophenyl)sulfonyl)-3,5-bis(4-(trifluoromethyl)benzylidene)piperidin-4-one (FSTBP), (3Z,5Z)-1-((4-nitrophenyl)sulfonyl)-3,5-bis(4-(trifluoromethyl)benzylidene)piperidin-4-one (NSTBP), 4-(((3Z,5Z)-4-oxo-3,5-bis(4-(trifluoromethyl)benzylidene)piperidin-1-yl)sulfonyl)-benzonitrile (OTBPS) were executed by utilizing Gaussian 09 program package20 and making use of DFT approach with B3LYP which is abbreviated of the Becke, three-parameter, Lee–Yang–Parr exchange–correlation functional,21 and 6-311G(d,p) basis set.22,23

Initial geometries of TBTBP, FSTBP, NSTBP, and OTBPS derivatives were attained through crystal structures using the Crystallographic Information File (CIF). Complete geometry optimization followed by vibrational analysis in the gas phase exclusive of symmetry restrictions was executed at the B3LYP/6-311G(d,p) functional. For estimation of FMO and NBO analyses, the B3LYP/6-311G(d,p) level of theory was utilized. Three levels of theory as HF, B3LYP, and M06 with the 6-311G(d,p) basis set were utilized to evaluate the NLO characteristics of BTBP, FSTBP, NSTBP, and OTBPS derivatives.

Lowest unoccupied molecular orbital (LUMO)/highest occupied molecular orbital (HOMO) energies and their energy gap values were utilized with the aid of following given eqs 16 to estimate the global reactivity parameters (GRP).2427 We found out the electronic affinity (A) and ionization potential (I) values using eqs 1 and 2.

graphic file with name ao0c01273_m001.jpg 1
graphic file with name ao0c01273_m002.jpg 2

In above equations, “I” represents ionization potential, while “A” denoted the electron affinity.

Hardness and electronegativity values were accomplished using eqs 3 and 4, respectively.

graphic file with name ao0c01273_m003.jpg 3
graphic file with name ao0c01273_m004.jpg 4

The electrophilicity value was calculated using eq 5.

graphic file with name ao0c01273_m005.jpg 5

The value of softness was calculated using eq 6.

graphic file with name ao0c01273_m006.jpg 6

Equation 7 was used for estimating the average polarizability ⟨α⟩.28

graphic file with name ao0c01273_m007.jpg 7

First hyperpolarizability (βtot) was worked out from eq 8.29

graphic file with name ao0c01273_m008.jpg 8

All input as well as output files were treated with the assistance of Chemcraft,30 GaussSum,31 GaussView5.0,32 and Avogadro33 softwares.

Results and Discussion

Molecular Geometric Parameters

The geometries of the TBTBP, FSTBP, NSTBP, and OTBPS were optimized using the B3LYP/6-311G(d,p) functional. The structural parameters regarding experimental and theoretical study are organized in Tables S1–S4 (Supporting Information). The graphical illustration of bond lengths is shown in Figure 1, while graphical representation about bond angles is shown in Figure S1.

Figure 1.

Figure 1

Comparison between experimental and simulated bond lengths (Å) of the entitled compounds.

Considering bond lengths, the identical DFT and XRD determined values were found to be 1.462 Å for C14–C15 in FSTBP and 1.502 and 1.222 Å for C2–C3 and N35–O40 in NSTBP, respectively. The least deviated DFT and XRD compared value were found to be 0.001 Å for C4–C14 in NSTBP and 0.001 Å for C1–N35, C6–C7, and C14–C15 in OTBPS. However, the most deviated DFT and XRD compared values were 0.072 Å for C22–F33 in TBTBP, 0.105 Å for C13–F39 in FSTBP, 0.046 Å for N34–S41 in NSTBP, and 0.052 Å for N35–S40 in OTBPS (Figure 1).

While taking in account of bond angles, the identical DFT and XRD determined values were found to be 120.8° for C24–C23–C28 in TBTBP, 112.3° for C4–C5–N33 in FSTBP, 121.3° for C15–C16–C17 in NSTBP, and 117.9° for C8–C7–C12 in OTBPS. The least deviated DFT and XRD compared value was found to be 0.1° for C9–C10–C11 in TBTBP, C7–C8–C9 in FSTBP, C17–C18–C19 in NSTBP, and C4–C3–O37 in OTBPS. Nevertheless, the most deviated DFT and XRD compared values were 23.4° for F34–C22–F39 in TBTBP, 41.7° for F28–C13–F38 in FSTBP, 2.8° for C22–S41–O37 in NSTBP, and 2.6° for O38–S40–O39 in OTBPS (Figure S1). Overall, entitled compounds are diverged in the range of 0.0105–0.0 Å for bond lengths and 0.0–41.7° for bond angles.

The relative analysis exposed that DFT values of bond lengths (Figure 1) and bond angles (Figure S1) were elevated as compared to XRD values. However, in a few cases, the opposite happened because of the medium effect.

To further strengthen the associated among DFT study and experimental findings, eqs 912 were utilized to calculate the error.

graphic file with name ao0c01273_m009.jpg 9

The mean absolute deviation (MAD), RMSE, and mean absolute percentage error (MAPE) can be estimated using eqs 1012

graphic file with name ao0c01273_m010.jpg 10
graphic file with name ao0c01273_m011.jpg 11
graphic file with name ao0c01273_m012.jpg 12

In the above equations, DFT and EXP represent the bond angle and bond length values obtained through DFT and experimental calculations, respectively. “t” indicates the number of bond angle or bond length values, and “n” describes the total number of the considered bond angle or lengths values. The error calculation results obtained from eqs 912 had been collected in Table S5.

In structural calculations, the RMSE approach was frequently utilized. Therefore, our investigated systems were also processed using the RMSE approach. RMSE values were observed less in magnitude for bond lengths of the title compounds. However, for bond angles, the RMSE values in TBTBP and FSTBP were found to be larger because of higher deviation of the CF values.

Hirshfeld Surface Analysis

Hirshfeld surface analysis for TBTBP, FSTBP, NSTBP, and OTBPS was performed to evaluate the intermolecular interactions.3436 The Hirshfeld surface configurations for molecules such as TBTBP, FSTBP, NSTBP, and OTBPS were shown in Figure 2. Red and white colors are used in the Hirshfeld surface plot to indicate the strongest and intermediate interactions, respectively, while the blue color demonstrates negligible intermolecular interactions.

Figure 2.

Figure 2

(a–d) show Hirshfeld surfaces mapped over dnorm in the range from −0.200 to 1.307, −0.262 to 1.638, −0.291 to 1.265, and −0.185 to 1.453 au for TBTBP, FSTBP, NSTBP, and OTBPS, respectively. 1 au of electron density = 6.748 e Å–3.

Two-dimensional fingerprint plots were also used to explore the intermolecular interactions. Further disintegration of this plot enumerates the individual contributions of every intermolecular interaction involved in the molecular structure.3739Figures S2–S5 were drawn to portray two-dimensional fingerprint plots of investigated compounds.

Various interatomic contacts in terms of percentage intermolecular contributions to the Hirshfeld surface for TBTBP, FSTBP, NSTBP, and OTBPS are shown in Tables S6–S9, respectively, and each atom percentage contributions in the Hirshfeld surface along with other atoms existing outside the Hirshfeld surface to identify its part in crystal packingare shown in Tables S10–S13 and Figure 3. In crystal packing of TBTBP, FSTBP, NSTBP, and OTBPS, Tables S6–S9 show that the F···H contact emerge to be the main contributor and its contribution for TBTBP, FSTBP, NSTBP, and OTBPS was 27.6, 35.2, 22.5, and 23.8%, correspondingly, as could be seen in Figure 3. The function of hydrogen atoms is extremely important in stabilization of TBTBP, FSTBP, NSTBP, and OTBPS structures, and its percentage involvement to intermolecular contact was observed in larger magnitude (Figure 3). Tables S10–S13 shows that most interacted atoms are hydrogen atoms and play a prominent role in crystal packing, and the percentage contribution of intermolecular contacts of H-atoms with all other atoms existing outside the Hirshfeld surface are 49.4, 42.6, 47.2, and 41.2% for TBTBP, FSTBP, NSTBP, and OTBPS, respectively (Figure 3). The smallest contribution to intermolecular interaction is of the sulphur atom; its role in crystal packing is found to be 0% in TBTBP, FSTBP, and OTBPS, while in NSTBP its contribution in packing is obtained to be 0.1% (Tables S10–S13). The voids in the unit of each compound were revealed in Figure 4:

Figure 3.

Figure 3

Percentage contributions of all interatomic contacts for entitled compounds.

Figure 4.

Figure 4

Show view of the voids (Wolff et al., 2012) in the crystal structure of TBTBP, FSTBP, NSTBP, and OTBPS.

These voids were calculated on the basis of sum of spherical atomic electron densities at suitable nuclear positions (pro-crystal electron density).40 The crystal-void calculation (results under 0.002 au iso-value) shows the void volumes of TBTBP, FSTBP, NSTBP, and OTBPS are found to be 378.12, 164.27 160.11, and 202.04 Å3 respectively.

Natural Bond Orbital Analysis

Using NBO investigation, the significant explanation of orbital interaction, orbital hybridization, and atomic charge can be obtained.41 NBO analysis offers elucidation regarding inter and intra-molecular interfaces which occur in empty and filled orbitals. The strong interactions between donors and acceptors produce large stabilization energy. The stabilization energy E(2) coupled with the delocalization for each donor (i) and acceptor (j) was evaluated using eq 13.

graphic file with name ao0c01273_m013.jpg 13

where εj and εi are off-diagonal, and F(i,j) is the diagonal NBO Fock matrix elements; qi is the donor orbital occupancy and E(2) is stabilization energy.

The NBO analysis has been completed for compounds (TBTBP, FSTBP, NSTBP, and OTBPS). The NBO analysis for TBTBP, FSTBP, NSTBP, and OTBPS has been elaborated in Tables S6–S9. Some selected values are given in Table 1. The numbering scheme for entitled compounds was represented in Figures S6–S9.

Table 1. NBO Representative Values for TBTBP, FSTBP, NSTBP, and OTBPS.

derivative donor (i) type acceptor (j) type E(2)a E(j) – E(i)b [au] F(i;j)c [au]
TBTBP C37–C45 π C38–C40 π* 230.96 0.01 0.080
  C37–C43 π C38–C40 π* 217.94 0.01 0.080
  C5–C11 π C13–C21 π* 10.41 0.30 0.053
  C18–C19 σ C16–C18 σ* 5.08 1.27 0.072
  C31–C32 σ C29–C31 σ* 5.08 1.27 0.072
  F57 LP(2) C42–C43 π* 20.05 0.43 0.089
  F47 LP(2) C23–F49 σ* 5.03 0.66 0.052
FSTBP C37–C45 π C38–C40 π* 230.96 0.01 0.080
  C42–C43 π C38–C40 π* 217.94 0.01 0.080
  C16–C18 π C18–C19 π* 5.07 1.27 0.072
  C31–C32 σ C36–F55 σ* 6.90 0.50 0.056
  C31–C32 σ C36–F54 σ* 1.51 0.50 0.026
  F57 LP(2) C42–C43 π* 20.05 0.43 0.089
  O54 LP(2) C36–F55 σ* 5.12 0.66 0.053
  F56 LP(2) C36–F55 σ* 5.12 0.65 0.053
NSTBP C26–C34 π C31–C32 π* 22.37 0.28 0.071
  C13–C21 π C18–C19 π* 22.20 0.28 0.071
  C37–C45 π C38–C40 π* 18.71 0.30 0.068
  N57–O58 π N57–O58 π* 7.82 0.34 0.055
  C29–C31 σ C31–C32 σ* 5.07 1.27 0.072
  C45–C46 σ C37–C38 σ* 4.93 1.08 0.065
  O59 LP(3) N57–O58 π* 175.14 0.15 0.147
  F47 LP(2) C23–F49 σ* 5.03 0.66 0.052
OTBPS C37–C45 π C38–C40 π* 168.68 0.01 0.079
  C13–C21 π C5–C11 π* 65.07 0.02 0.069
  C57–N58 π C42–C43 π* 5.70 0.34 0.043
  C6–C7 σ C24–C26 σ* 5.13 1.13 0.068
  O1 LP(2) C6–C7 σ* 19.32 0.68 0.104
  F54 LP(2) C36–F55 σ* 5.10 0.66 0.053
  F48 LP(2) C23–F49 σ* 5.10 0.66 0.053
a

E(2)means energy of the hyper conjugative interaction (stabilization energy in kcal/mol).

b

Energy difference between the donor and acceptor i and j NBO orbitals.

c

F(i;j) is the Fock matrix element between i and j NBO orbitals.

The most probable transitions comprising gigantic stabilization energies such as π(C37–C45) → π*(C38–C40), π(C37–C45) → π*(C38–C40), π(C37–C45) → π*(C38–C40) and π(C37–C45) → π*(C38–C40) contained 230.96, 230.96, 18.71, and 168.68 kcal/mol for TBTBP, FSTBP, NSTBP, and OTBPS, respectively. The transitions indicating the conjugation in chemical structures as π(C37–C43) → π*(C38–C40) in TBTBP, π(C42–C43) → π*(C38–C40) and π(C16–C18) → π*(C18–C19) in FSTBP, π(C26–C34) → π*(C31–C32) and π(C13–C21) → π*(C18–C19) in NSTBP, and π(C13–C21) → π*(C5–C11) in OTBPS with stabilization energy values as 217.94 kcal/mol for TBTBP, 217.94 and 5.07 kcal/mol for FSTBP, 22.37 and 22.20 kcal/mol for NSTBP, and 65.07 kcal/mol for OTBPS, respectively. However, transitions such as π(C37–C45) → π*(C38–C40), π(C37–C45) → π*(C38–C40), π(C26–C34) → π*(C31–C32), and π(C37–C45) → π*(C38–C40) demonstrated highest stabilization energies of 230.96, 230.96, 22.37, and 168.68 kcal/mol, respectively, in all compounds: TBTBP, FSTBP, NSTBP, and OTBPS (see Tables S14–S17).

Moreover, transitions such as π(C5–C11) → π*(C13–C21), π(C16–C18) → π*(C18–C19), π(N57–O58) → π*(N57–O58), and π(C57–N58) → π*(C42–C43) consisting of 10.41,5.07, 7.82, and 5.70 kcal/mol stabilization energies with smallest magnitudes in TBTBP, FSTBP, NSTBP, and OTBPS, respectively. The σ → σ* transitions in contrast to π → π* transitions are originated owing to weak donor (σ), acceptor (σ*) interactions which are imperative for studying transitions with lesser stabilization energy values. Transitions such as σ(C18–C19) → σ*(C16–C18), σ(C31–C32) → σ*(C29–C31), σ(C31–C32) → σ*(C36–F55), σ(C29–C31) → σ*(C31–C32), and σ(C6–C7) → σ*(C24–C26) contained 5.08, 5.08, 6.90,5.07, and 5.13 kcal/mol energy values in TBTBP, FSTBP, NSTBP, and OTBPS, respectively, presenting larger energies amongst all σ → σ* interactions. While transitions σ(C18–C19) → σ*(C16–C18), σ(C31–C32) → σ*(C29–C31), σ(C31–C32) → σ*(C36–F54), σ(C45–C46) → σ*(C37–C38),and σ(C6–C7) → σ*(C24–C26) were having least stabilization energy values 5.08, 5.08, 1.51, 4.93, and 5.13 kcal/mol in TBTBP, FSTBP, NSTBP, and OTBPS, respectively (Table 1).

Similar sort of interactions was noted in accordance to the resonance process. For example, LP2(F57) → π*(C42–C43), LP2(F57) → π*(C42–C43), LP3(O59) → π*(N57–O58), and LP2(O1) → π*(C6–C7) produced 20.05, 20.05, 175.14, and 19.32 kcal/mol in TBTBP, FSTBP, NSTBP, and OTBPS correspondingly. While interactions as LP2(F47) → σ*(C23–F49), LP2(O54) → π*(C36–F55), LP2(F56) → σ*(C36–F55), LP2(F47) → π*(C23–F49), LP2(F54) → π*(C36–F55), and LP2(F48) → π*(C23–F49) produced 5.03, 5.12, 5.12, 5.03, 5.10, and 5.10 kcal/mol stabilization energies in case of resonance for TBTBP, FSTBP, NSTBP, and OTBPS, respectively (Table 1).

Vibrational Analysis

In order to have better perception of vibrational modes linked with TBTBP, FSTBP, NSTBP, and OTBPS, DFT studies were conducted at the B3LYP/6-311G(d,p) level of theory under solvent-free conditions (gas phase). The number of atoms in TBTBP, FSTBP, NSTBP, and OTBPS are 60, 57, 59, and 58 atoms correspondingly with the C1 point group symmetry.

The experimental vibrational bands observed at 1673–1675 cm–1, which are associated to the C=O functional group stretching vibration of TBTBP, FSTBP, NSTBP, and OTBPS, respectively, which are found in good correspondence with simulated values as 1627, 1627, 1633 and 1633 cm–1,4 respectively. The experimental wave numbers in the range of 1615–1613 cm–1 existed because of strong bands of the C–C group in the α,β-unsaturated ketone of TBTBP, FSTBP, NSTBP, and OTBPS, which matched with simulated values as 1602, 1603, 1605, and 1605 cm–1, respectively. All the assignments were obtained in fine concurrence to the earlier reported similar structured piperidones values. The sulfonamide group (−SO2N−) in TBTBP, FSTBP, NSTBP, and OTBPS are attributed with the strong bands in the range of 1169–1161 cm–1 (Experimental Section), which are also found in correlation of simulated values as 1175, 1174, 1169, and 1168 cm–1, respectively. The simulated assignments are also observed in good harmony to the IR values of reported piperidones.4 The stretching vibration of −CN is obtained at 2343 cm–1, which confirms the presence of a cyano group (−CN) in OTBPS.4

Natural Population Analysis

The natural charges of TBTBP, FSTBP, NSTBP, and OTBPS derivatives were measured by NBO analysis using the B3LYP/6-311G(d,p) functional, and results were presented in Figure 5. The natural charge examination was used frequently to evaluate the charge transformation process which originates in reactions, the phenomenon associating the electronegativity equalization and electrostatic potential on external surfaces of systems.42 The electronic charges of atoms play a vital role in the bonding capability and molecular conformation.43 Natural charges data of investigated molecules reveals that the electron density is unequally redistributed over the benzene rings in the attendance of extra electronegative atoms as N and O.44

Figure 5.

Figure 5

Natural population analysis of the title compounds.

Our concern is to appraise the reactivity of figured charges and describe the distribution of electron density over investigated compounds.45 Furthermore, no discrepancy in charge distribution was observed above all H-atoms indicated by natural population examination. Because of negative charge of carbon atoms, H-atoms bear positive charges. The high negative charges were contained by fluorine, nitrogen, and oxygen atoms in the entitled compounds. Because of the resonance process, some carbon atoms are enforced with large negative charges using oxygen and nitrogen atoms. However, the sulfur atoms in entitled compounds contained high positive charges.

Frontier Molecular Orbitals

The frontier molecular orbitals (FMOs) were not only used to describe the electric and optical properties, but chemical stability of investigated systems is also interpreted with this analysis.46 The HOMO abbreviation is used for the highest occupied molecular orbital, the LUMO is used for the lowest unoccupied molecular orbital which happens to be the most substantial orbitals of FMOs. The HOMO exhibits the capability of donating an electron, whereas the LUMO shows capability of accepting an electron. The frontier orbital energy gap is an appreciated constraint in order to get knowledge concerning the dynamic stability and chemical reactivity of species (Table 2).4755

Table 2. Frontier Molecular Orbitals Energies for TBTBP, FSTBP, NSTBP, and OTBPSa.

compounds TBTBP
FSTBP
NSTBP
OTBPS
MO(s) E (eV) ΔE (eV) E (eV) ΔE (eV) E (eV) ΔE (eV) E (eV) ΔE (eV)
HOMO –7.077 3.982 –7.163 3.988 –7.249 3.852 –7.238 4.021
LUMO –3.095   –3.175   –3.397   –3.216  
HOMO – 1 –7.077 5.074 –7.213 5.137 –7.336 4.140 –7.324 4.628
LUMO + 1 –2.003   –2.076   –3.196   –2.696  
HOMO – 2 –7.223 5.859 –7.356 5.881 –7.638 5.526 –7.624 5.527
LUMO + 2 –1.364   –1.474   –2.112   –2.097  
HOMO – 3 –7.482 6.312 –7.568 6.254 –7.968 6.083 –7.869 6.038
LUMO + 3 –1.170   –1.314   –1.886   –1.831  
a

HOMO = highest occupied molecular orbital, LUMO = lowest unoccupied molecular orbital.

The HOMO–LUMO energy values were calculated to be −7.077/–3.095, 7.163/–3.175, −7.249/–3.397, and −7.238/–3.216 eV for TBTBP, FSTBP, NSTBP, and OTBPS, respectively. The band gaps of TBTBP, FSTBP, NSTBP, and OTBPS were obtained to be 3.982, 3.988, 3.852, and 4.021 eV, respectively. The band gap of NSTBP was obtained slightly lesser than its analogues because of the electron-withdrawing ability of −NO2.

The electron density in the HOMO of TBTBP, FSTBP, NSTBP, and OTBPS were dispersed on the 3-methylene-5-(4-(trifluoromethyl)benzylidene)piperidin-4-one fragment except for the 1-methyl-4-(methylsulfonyl)benzene, phenyl, and trifluoromethane groups. The electron density in the LUMO for TBTBP, FSTBP, NSTBP, and OTBPS concentrated on 3,5-bis((E)-benzylidene)piperidin-4-one except for 1-methyl-4-(methylsulfonyl)benzene and trifluoromethane groups (Figure 6).

Figure 6.

Figure 6

Frontier molecular orbitals of TBTBP, FSTBP, NSTBP, and OTBPS.

Global Reactivity Parameters

The global softness (σ), global electrophilicity (ω), global hardness (η),electronegativity (X), chemical potential (μ), electron affinity (A), and ionization potential (I) were also calculated using the HOMO and LUMO energies.25,48,49,5658 The findings of GRPs were obtained from eqs 16, which were tabulated in Table 3. These chemical quantities revealed the chemical reactivity of the former mentioned derivatives.

Table 3. Calculated Global Reactivity Parameters of Entitled Compounds, Units in eV.

compounds I A X η μ ω σ
TBTBP 7.077 3.096 5.086 1.991 –5.086 6.498 0.251
FSTBP 7.163 3.175 5.169 1.994 –5.169 6.699 0.251
NSTBP 7.249 3.397 5.323 1.926 –5.323 7.356 0.260
OTBPS 7.238 3.216 5.227 2.011 –5.227 6.795 0.249

In broad context, A and I values were used to describe the electron-accepting and donating aptitude of investigated molecules, respectively. The I values of TBTBP, FSTBP, NSTBP, and OTBPS were found to be larger in magnitude as compared to A values. The chemically hardness values were observed as 1.991, 1.994, 1.926, and 2.011 eV for TBTBP, FSTBP, NSTBP, and OTBPS, respectively. The chemically softness values were computed as 0.251, 0.251, 0.260, and 0.249 eV for TBTBP, FSTBP, NSTBP, and OTBPS, respectively (Table 3). These results indicated that studied derivatives have enormous kinetic stability and exhibit fine concurrence with SC-XRD,4 and NBO findings.

Nonlinear Optical Properties

The examination of NLO effects in organic compounds established a fascinating topic which was initiated by Champagne and Bishop.59 The organic polymeric and heterocyclic compounds having large hyperpolarizability amplitudes have drawn attention owing to their applicability in the interest of NLO materials.60,61 Recently, many researches are enthusiastic to establish the favorable synthetic procedures of organic chromophores for their distinct properties in the NLO field.62 Furthermore, numerous researcher groups are doing extensive work on organic compounds because of their relative facile synthesis as well as greater efficacy in NLO response properties.13 DFT-based studies have delivered an alluring type role in understanding of experimental findings, especially in NLO responses.63 In this context, entitled compounds such as TBTBP, FSTBP, NSTBP, and OTBPS have not been investigated for their potential NLO characteristics by both computational and experimental researchers. The entitled compounds contain huge interest owing to their heterocyclic ring’s involvement in the electron-withdrawing groups. The electronic and optical properties might be expected significantly in magnitude because of their greater conjugated electronic systems.6469 Moreover, conjugated electronic systems could be the premeditated primary concept for the intramolecular charge transfer (ICT) process.70 Indeed, the larger charge transfer may lead to larger amplitudes for mean polarizabilities ⟨α⟩ and first hyperpolarizabilities (βtot). Herein, we have the selected functional: HF, B3LYP, and M06 with the 6-311G(d,p) basis set for determination of average polarizabilities ⟨α⟩ and second-order polarizabilities for TBTBP, FSTBP, NSTBP, and OTBPS. Equations 7 and 8 were utilized for the calculation of mean polarizabilities ⟨α⟩ and first-order hyperpolarizabilities (βtot) using x, y, and z components which are listed in Tables 4 and 5, respectively.

Table 4. Dipole Polarizabilities and Major Contributing Tensors (au) of the Studied Compounds.

func comp. αxx αyy αzz ⟨α⟩
M06 TBTBP 594.1 251.6 288.9 378.2
  FSTBP 581.1 237.7 275.3 364.7
  NSTBP 600.6 246.0 297.9 381.5
  OTBPS 283.6 155.2 678.1 372.3
B3LYP TBTBP 598.0 251.3 287.1 378.8
  FSTBP 584.9 236.9 274.2 365.3
  NSTBP 604.0 250.3 292.27 382.2
  OTBPS 605.4 254.8 290.8 383.7
HF TBTBP 512.9 233.6 273.4 339.9
  FSTBP 500.2 220.6 260.5 327.1
  NSTBP 517.6 226.3 282.7 342.2
  OTBPS 519.8 233.9 278.6 344.1

Table 5. Computed Second-Order Polarizabilities (βtot) and Major Contributing Tensors (au) of the Studied Compoundsa.

func. comp. βxxx βxxy βxyy βyyy βxxz βtot
M06 TBTBP 326.5 –505.4 89.5 –67.6 588.6 951.6
  FSTBP 280.6 –525.1 84.9 –76.4 583.7 914.2
  NSTBP –383.1 –223.9 –310.5 147.1 –807.2 1170.3
  OTBPS 283.6 –400.2 155.2 10.2 678.1 904.9
B3LYP TBTBP –427.2 –538.8 –99.1 –92.9 –647.9 1102.6
  FSTBP –369.7 –552.2 –94.2 –104.3 –646.5 1051.4
  NSTBP –536.1 –349.6 –302.2 172.2 –837.3 1382.2
  OTBPS –429.5 –470.7 –170.6 17.9 –717.7 1076.0
HF TBTBP 115.1 –312.5 48.7 –22.4 –69.3 392.4
  FSTBP 77.1 –332.6 43.5 –46.3 297.5 502.3
  NSTBP –72.6 –207.4 –144.1 19.5 –444.0 538.9
  OTBPS 79.4 –274.5 79.7 –28.6 359.7 500.2
a

Func = functionals; comp. = compounds.

The mean polarizabilities ⟨α⟩ are found to be larger at the exchange correlation functional B3LYP level as compared to magnitudes of ⟨α⟩ observed with the meta-hybrid GGA method M06 and HF method for TBTBP, FSTBP, NSTBP, and OTBPS. However, the mean polarizabilities ⟨α⟩ are obtained to be lesser at the HF level as compared to magnitudes of ⟨α⟩ at B3LYP and M06 levels for all investigated molecules. It can be concluded that mean polarizabilities ⟨α⟩ values of all studied compounds at the M06 level are found to be in between the values noted at B3LYP and HF levels. For TBTBP, FSTBP, NSTBP, and OTBPS, B3LYP and M06 methods exhibited almost similar mean polarizabilities ⟨α⟩ values (Figure 7). The average polarizability values of investigated compounds are found to be in following order as B3LYP > M06 > HF. Moreover, average polarizability ⟨α⟩ values of TBTBP, FSTBP, NSTBP, and OTBPS were found to be larger than the standard urea molecule (⟨α⟩ =32.918 au at M06).

Figure 7.

Figure 7

Calculated average polarizabilities at different levels of the studied compounds.

The literature discloses that HOMO–LUMO energy gap possess influence on the molecular polarizability.71 The HOMO–LUMO energy gap has an inverse relationship with linear polarizability and nonlinear polarizability (Figures 6 and 7). The compounds comprising small HOMO–LUMO energy gap supports large nonlinear and linear polarizabilities. In our investigation, NSTBP is observed with a narrow energy gap in contrast to TBTBP, FSTBP, and OTBPS energy gap values. Consequently, it exhibits larger linear and nonlinear polarizability values (Figures 6 and 7).

The entitled compounds: TBTBP, FSTBP, NSTBP, and OTBPS contained substituents like −CH3, −F, −NO2, and −CN on the N-benzenesulfonyl moiety, respectively. The −CH3 substituent in TBTBP displayed electron-donating effect, while −F, −NO2, and −CN substituents of entitled compounds FSTBP, NSTBP, and OTBPS respectively, showed electron-withdrawing effects. The order of substitutes in terms of the electron-withdrawing effect can be NO2 > CN > F. Interestingly, NSTBP and OTBPS having strong electron-withdrawing substitutes, −NO2 and −CN, respectively, led to a remarkable effect on NLO as compared to TBTBP containing an electron-donating group (−CH3) and FSTBP with a weak electron-withdrawing substituent (−F). Therefore, first hyperpolarizability (βtot) values of NSTBP–NO2 were found to be larger as 1170.3, 1382.2, and 538.9 au at HF, B3LYP, and M06 levels, respectively, than TBTBP, FSTBP, and OTBPS.

The effect of electron-donating substituent −CH3 in TBTBP is evident with observed βtot values 951.6 (B3LYP) and 1102.6 (M06) which are found to be smaller than NSTBP–NO2 and larger than FSTBP and OTBPS βtot values at M06, B3LYP levels of theory. The first hyperpolarizability (βtot) values of OTBPS–CN were obtained to be lesser as 904.9, 1076.0, and 500.2 au than all studied compounds at the M06 level, smaller than NSTBP, TBTBP, and larger than FSTBP at the B3LYP level of theory. The substituents on TBTBP and FSTBP compounds do not play a significant role in first-order hyperpolarizability (βtot) value patterns. There might be an influence of conformation about second-order polarizabilities of TBTBP and FSTBP compounds. The literature revealed that TBTBP and FSTBP having pseudo-axial conformation for the arylsulfonyl group with the reference of the 4-pyridone ring, while NSTBP and OTBPS having pseudo-equatorial conformation for the arylsulfonyl group.4 Furthermore, it has been observed that the HF method showed the least values of βtot for all compounds, while the B3LYP level was at the top one having the highest βtot for compounds as compared to other utilized levels HF and M06. The βtot values of studied compounds were found to be in the following order at different methods: B3LYP > M06 > HF (Figure 8). The literature review revealed that urea is used as the standard molecule for a comparative study of NLO.7275 The calculated βtot values of TBTBP, FSTBP, NSTBP, and OTBPS at all aforementioned levels were obtained to be larger than urea (βtot = 66.847 au,) at the M06 level.

Figure 8.

Figure 8

Computed first-order hyperpolarizabilities of different levels of the studied compounds.

Conclusions

The current study discloses that computed geometrical parameters (bond lengths and bond angles) are in excellent agreement with the SC-XRD data. The experimental FT-IR spectra results were found in line with simulated vibrational measurements. NBO analysis revealed that the intramolecular charge transfer exists in entitled compounds. Furthermore, NBO-based hyper conjugative interactions values were observed with larger values which endorse highest stability of investigated molecules. Moreover, Hirshfeld surface analysis also endorses the highest molecular stability. The energy gap of the HOMO and LUMO were found to be 3.982, 3.988, 3.852, and 4.021 eV for TBTBP, FSTBP, NSTBP, and OTBPS, respectively. The band gap of OTBPS was obtained slightly greater than its analogues because of the electron-withdrawing ability of −CN. The chemically hardness values for TBTBP, FSTBP, NSTBP, and OTBPS were observed larger in contrast to corresponding softness values which revealed that studied derivatives have enormous kinetic stability and exhibit fine concurrence with NBO findings. The entitled compounds have average polarizabilities ⟨α⟩ in the span of 365.3–383.7 (au) at B3LYP, 364.7–381.5 (au) at M06, and 327.1–344.1 (au) at the HF level. Furthermore, the entitled compounds have conspicuous large NLO response (βtot) in the span of 1051.45–1382.2 (au) at B3LYP, 904.9–1170.3 (au) at M06, and 392.4–538.9 (au) at the HF level. Among entitled compounds, NSTBP consists of highest ⟨α⟩ and βtot values. The βtot values of TBTBP, FSTBP, NSTBP, and OTBPS were 25.64, 24.45, 27.83, and 25.02, respectively, times larger as compared to the βtot value of the standard molecule (urea). It is anticipative that the quantum chemical-based investigation of the TBTBP, FSTBP, NSTBP, and OTBPS could be fruitful for granting a most favorable model for second-order NLO responses.

Acknowledgments

Authors are thankful to Departamento de Química Fundamental, Instituto de Quimica, Universidade de São Paulo, Av. Prof. LineuPrestes, 748, São Paulo, 05508-000, Brazil for providing theoretical laboratory facilities for DFT study. Moreover, M.K. (grant #1314/2017) is thankful to HEC Islamabad Pakistan for the financial support. Furthermore, A.A. also gratefully acknowledges the financial support of HEC Pakistan (grant # 21-2037/SRGP/R&D/HEC/2018).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.0c01273.

  • Tables of comparison of bond lengths (Å) and angles (deg) using B3LYP/6-311G(d,p), error values, percentage contributions of interatomic contacts to the Hirshfeld surface, percentage contributions of an atom in Hirshfeld surface with all other atoms present outside the Hirshfeld surface, natural bond orbital (NBO) analysis of the reported compounds usingB3LYP/6-311G(d,p). Also, the computed second-order polarizabilities (βtot) and major contributing tensors (au) of the studied compounds, comparison between experimental and simulated bond angles (deg) for the entitled compounds, two-dimensional fingerprint plot for studied compounds, and the NBO numbering scheme of the investigated compounds (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao0c01273_si_001.pdf (2.2MB, pdf)

References

  1. Carey F. A.; Sundberg R. J.. Advanced Organic Chemistry: Part B: Reaction and Synthesis; Springer Science & Business Media, 2007. [Google Scholar]
  2. Din Z. U.; Santos A. d.; Trapp M. A.; Lazarin-Bidóia D.; Garcia F. P.; Peron F.; Nakamura C. V.; Rodrigues-Filho E. Curcumin inspired synthesis of unsymmetrical diarylpentanoids with highly potent anti-parasitic activities: in silico studies and DFT-based stereochemical calculation. MedChemComm 2016, 7, 820–831. 10.1039/c5md00599j. [DOI] [Google Scholar]
  3. Zhao C.; Liu Z.; Liang G. Promising curcumin-based drug design: mono-carbonyl analogues of curcumin (MACs). Curr. Pharm. Des. 2013, 19, 2114–2135. 10.2174/1381612811319110012. [DOI] [PubMed] [Google Scholar]
  4. Li N.; Bai X.; Zhang L.; Hou Y. Synthesis, crystal structures and anti-inflammatory activity of four 3, 5-bis (arylidene)-N-benzenesulfonyl-4-piperidone derivatives. Acta Crystallogr., Sect. C: Struct. Chem. 2018, 74, 1171–1179. 10.1107/s2053229618013232. [DOI] [PubMed] [Google Scholar]
  5. Yin D.-l.; Liang Y.-j.; Zheng T.-s.; Song R.-p.; Wang J.-b.; Sun B.-s.; Pan S.-h.; Qu L.-d.; Liu J.-r.; Jiang H.-c. EF24 inhibits tumor growth and metastasis via suppressing NF-kappaB dependent pathways in human cholangiocarcinoma. Sci. Rep. 2016, 6, 32167. 10.1038/srep32167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Yadav V. R.; Sahoo K.; Awasthi V. Preclinical evaluation of 4-[3, 5-bis (2-chlorobenzylidene)-4-oxo-piperidine-1-yl]-4-oxo-2-butenoic acid, in a mouse model of lung cancer xenograft. Br. J. Pharmacol. 2013, 170, 1436–1448. 10.1111/bph.12406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Kálai T.; Kuppusamy M. L.; Balog M.; Selvendiran K.; Rivera B. K.; Kuppusamy P.; Hideg K. Synthesis of N-substituted 3, 5-bis (arylidene)-4-piperidones with high antitumor and antioxidant activity. J. Med. Chem. 2011, 54, 5414–5421. 10.1021/jm200353f. [DOI] [PubMed] [Google Scholar]
  8. Domingo L.; Ríos-Gutiérrez M.; Pérez P. Applications of the conceptual density functional theory indices to organic chemistry reactivity. Molecules 2016, 21, 748. 10.3390/molecules21060748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Hussain A.; Khan M. U.; Ibrahim M.; Khalid M.; Ali A.; Hussain S.; Saleem M.; Ahmad N.; Muhammad S.; Al-Sehemi A. G.; Sultan A. Structural parameters, electronic, linear and nonlinear optical exploration of thiopyrimidine derivatives: A comparison between DFT/TDDFT and experimental study. J. Mol. Struct. 2020, 1201, 127183. 10.1016/j.molstruc.2019.127183. [DOI] [Google Scholar]
  10. Khalid M.; Ali A.; Adeel M.; Din Z. U.; Tahir M. N.; Rodrigues-Filho E.; Iqbal J.; Khan M. U. Facile preparation, characterization, SC-XRD and DFT/DTDFT study of diversely functionalized unsymmetrical bis-aryl-α, β-unsaturated ketone derivatives. J. Mol. Struct. 2020, 1206, 127755. 10.1016/j.molstruc.2020.127755. [DOI] [Google Scholar]
  11. 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 10.1002/aoc.5399. [DOI] [Google Scholar]
  12. Ali A.; Khalid M.; Marrugo K. P.; Kamal G. M.; Saleem M.; Khan M. U.; Concepción O.; de la Torre A. F. Spectroscopic and DFT/TDDFT insights of the novel phosphonate imine compounds. J. Mol. Struct. 2020, 1207, 127838. 10.1016/j.molstruc.2020.127838. [DOI] [Google Scholar]
  13. Khalid M.; Ali A.; De la Torre A. F.; Marrugo K. P.; Concepcion O.; Kamal G. M.; Muhammad S.; Al-Sehemi A. G. Facile Synthesis, Spectral (IR, Mass, UV– Vis, NMR), Linear and Nonlinear Investigation of the Novel Phosphonate Compounds: A Combined Experimental and Simulation Study. ChemistrySelect 2020, 5, 2994–3006. 10.1002/slct.201904224. [DOI] [Google Scholar]
  14. Garmire E. Nonlinear optics in daily life. Opt. Express 2013, 21, 30532–30544. 10.1364/oe.21.030532. [DOI] [PubMed] [Google Scholar]
  15. Peng Z.; Yu L. Second-order nonlinear optical polyimide with high-temperature stability. Macromolecules 1994, 27, 2638–2640. 10.1021/ma00087a039. [DOI] [Google Scholar]
  16. Tsutsumi N.; Morishima M.; Sakai W. Nonlinear optical (NLO) polymers. 3. NLO polyimide with dipole moments aligned transverse to the imide linkage. Macromolecules 1998, 31, 7764–7769. 10.1021/ma9803436. [DOI] [Google Scholar]
  17. Papadopoulos M. G.; Sadlej A. J.; Leszczynski J.. Non-Linear Optical Properties of Matter; Springer, 2006. [Google Scholar]
  18. Akram M.; Adeel M.; Khalid M.; Tahir M. N.; Khan M. U.; Asghar M. A.; Ullah M. A.; Iqbal M. A combined experimental and computational study of 3-bromo-5-(2, 5-difluorophenyl) pyridine and 3, 5-bis (naphthalen-1-yl) pyridine: Insight into the synthesis, spectroscopic, single crystal XRD, electronic, nonlinear optical and biological properties. J. Mol. Struct. 2018, 1160, 129–141. 10.1016/j.molstruc.2018.01.100. [DOI] [Google Scholar]
  19. Breitung E. M.; Shu C.-F.; McMahon R. J. Thiazole and thiophene analogues of donor– acceptor stilbenes: molecular hyperpolarizabilities and structure– property relationships. J. Am. Chem. Soc. 2000, 122, 1154–1160. 10.1021/ja9930364. [DOI] [Google Scholar]
  20. Frisch M.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Mennucci B.; Petersson G. e.. Gaussian 09, Revision D. 01, 2014.
  21. Civalleri B.; Zicovich-Wilson C. M.; Valenzano L.; Ugliengo P. B3LYP augmented with an empirical dispersion term (B3LYP-D*) as applied to molecular crystals. CrystEngComm 2008, 10, 405–410. 10.1039/b715018k. [DOI] [Google Scholar]
  22. Poirier R.; Kari R.; Csizmadia I. G.. Handbook of Gaussian Basis Sets; Elsevier, 1985. [Google Scholar]
  23. Frisch M. J.; Pople J. A.; Binkley J. S. Self-consistent molecular orbital methods 25. Supplementary functions for Gaussian basis sets. J. Chem. Phys. 1984, 80, 3265–3269. 10.1063/1.447079. [DOI] [Google Scholar]
  24. Parr R. G.; Szentpály L. v.; Liu S. Electrophilicity index. J. Am. Chem. Soc. 1999, 121, 1922–1924. 10.1021/ja983494x. [DOI] [Google Scholar]
  25. Parr R. G.; Donnelly R. A.; Levy M.; Palke W. E. Electronegativity: the density functional viewpoint. J. Chem. Phys. 1978, 68, 3801–3807. 10.1063/1.436185. [DOI] [Google Scholar]
  26. Chattaraj P. K.; Sarkar U.; Roy D. R. Electrophilicity index. Chem. Rev. 2006, 106, 2065–2091. 10.1021/cr040109f. [DOI] [PubMed] [Google Scholar]
  27. 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. 10.1016/j.cplett.2009.10.082. [DOI] [Google Scholar]
  28. Sajan D.; Joe H.; Jayakumar V. S.; Zaleski J. Structural and electronic contributions to hyperpolarizability in methyl p-hydroxy benzoate. J. Mol. Struct. 2006, 785, 43–53. 10.1016/j.molstruc.2005.09.041. [DOI] [Google Scholar]
  29. Wazzan N.; Safi Z. DFT calculations of the tautomerization and NLO properties of 5-amino-7-(pyrrolidin-1-yl)-2, 4, 4-trimethyl-1, 4-dihydro-1, 6-naphthyridine-8-carbonitrile (APNC). J. Mol. Struct. 2017, 1143, 397–404. 10.1016/j.molstruc.2017.04.101. [DOI] [Google Scholar]
  30. Glendening E.; Reed A.; Carpenter J.; Weinhold F.. NBO, version 3.1.; b Paul B. K.; Mahanta S.; Singh R. B.; Guchhait N. J. Phys. Chem. A 2010, 114, 2618. 10.1021/jp909029c. [DOI] [PubMed] [Google Scholar]
  31. 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. 10.1002/jcc.20823. [DOI] [PubMed] [Google Scholar]
  32. Dennington R.; Keith T.; Millam J.. GaussView, version 5, 2009.
  33. 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. Cheminf. 2012, 4, 17. 10.1186/1758-2946-4-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Zukerman-Schpector J.; Dallasta Pedroso S.; Sousa Madureira L.; Weber Paixão M.; Ali A.; Tiekink E. R. T. 4-Benzyl-1-(4-nitrophenyl)-1H-1, 2, 3-triazole: crystal structure and Hirshfeld analysis. Acta Crystallogr., Sect. E: Struct. Rep. Online 2017, 73, 1716–1720. 10.1107/s2056989017014748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. 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: Struct. Rep. Online 2019, 75, 711. 10.1107/s205698901900567x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. 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: Struct. Rep. Online 2019, 75, 667–674. 10.1107/s2056989019005164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Ali A.; Zukerman-Schpector J.; Weber Paixão M.; Jotani M. M.; Tiekink E. R. T. 7-Methyl-5-[(4-methylbenzene) sulfonyl]-2H, 5H-[1, 3] dioxolo [4, 5-f] indole: crystal structure and Hirshfeld analysis. Acta Crystallogr., Sect. E: Struct. Rep. Online 2018, 74, 184–188. 10.1107/s2056989018000889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. 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: Struct. Rep. Online 2019, 75, 721–727. 10.1107/s2056989019005966. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Tahir M. N.; Ashfaq M.; de la Torre A. 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. 10.1016/j.molstruc.2019.05.003. [DOI] [Google Scholar]
  40. Chebbi H.; Mezrigui S.; Ben Jomaa M.; Zid M. F. Crystal structure, Hirshfeld surface analysis and energy framework calculation of the first oxoanion salt containing 1, 3-cyclohexanebis (methylammonium):[3-(azaniumylmethyl) cyclohexyl] methanaminium dinitrate. Acta Crystallogr., Sect. E: Struct. Rep. Online 2018, 74, 949–954. 10.1107/s2056989018008381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Reed A. E.; Weinhold F.; Weiss R.; Macheleid J. Nature of the contact ion pair trichloromethyl-chloride (CCl3+ Cl-). A theoretical study. J. Phys. Chem. 1985, 89, 2688–2694. 10.1021/j100258a051. [DOI] [Google Scholar]
  42. Mulliken R. S. Electronic Population Analysis on LCAO-MO Molecular Wave Functions. I. J. Chem. Phys. 1955, 23, 1833–1840. 10.1063/1.1740588. [DOI] [Google Scholar]
  43. Parr R. G.; Pearson R. G. Absolute hardness: companion parameter to absolute electronegativity. J. Am. Chem. Soc. 1983, 105, 7512–7516. 10.1021/ja00364a005. [DOI] [Google Scholar]
  44. Li L.; Wu C.; Wang Z.; Zhao L.; Li Z.; Sun C.; Sun T. Density functional theory (DFT) and natural bond orbital (NBO) study of vibrational spectra and intramolecular hydrogen bond interaction of l-ornithine–l-aspartate. Spectrochim. Acta, Part A 2015, 136, 338–346. 10.1016/j.saa.2014.08.153. [DOI] [PubMed] [Google Scholar]
  45. Burrows J. A.; Pauling L. The nature of the chemical bond and the structure of molecules aid crystals. Ithaca: The Cornell University Press, 1939. 430 p. Sci. Educ. 1941, 25, 120. 10.1002/sce.3730250275. [DOI] [Google Scholar]
  46. Khan M. U.; Iqbal J.; Khalid M.; Hussain R.; Braga A. A. C.; Hussain M.; Muhammad S. Designing triazatruxene-based donor materials with promising photovoltaic parameters for organic solar cells. RSC Adv. 2019, 9, 26402–26418. 10.1039/c9ra03856f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Ghiasuddin; Akram M.; Adeel M.; Khalid M.; Tahir M. N.; Khan M. U.; Asghar M. A.; Ullah M. A.; Iqbal M. A combined experimental and computational study of 3-bromo-5-(2, 5-difluorophenyl) pyridine and 3, 5-bis (naphthalen-1-yl) pyridine: Insight into the synthesis, spectroscopic, single crystal XRD, electronic, nonlinear optical and biological properties. J. Mol. Struct. 2018, 1160, 129–141. 10.1016/j.molstruc.2018.01.100. [DOI] [Google Scholar]
  48. Ahmad M. S.; Khalid M.; Shaheen M. A.; Tahir M. N.; Khan M. U.; Braga A. A. C.; Shad H. A. Synthesis and XRD, FT-IR vibrational, UV–vis, and nonlinear optical exploration of novel tetra substituted imidazole derivatives: A synergistic experimental-computational analysis. J. Phys. Chem. Solids 2018, 115, 265–276. 10.1016/j.jpcs.2017.12.054. [DOI] [Google Scholar]
  49. Shahid M.; Salim M.; Khalid M.; Tahir M. N.; Khan M. U.; Braga A. A. C. Synthetic, XRD, non-covalent interactions and solvent dependent nonlinear optical studies of Sulfadiazine-Ortho-Vanillin Schiff base:(E)-4-((2-hydroxy-3-methoxy-benzylidene) amino)-N-(pyrimidin-2-yl) benzene-sulfonamide. J. Mol. Struct. 2018, 1161, 66–75. 10.1016/j.molstruc.2018.02.043. [DOI] [Google Scholar]
  50. Jawaria R.; Hussain M.; Khalid M.; Khan M. U.; Tahir M. N.; Naseer M. M.; Braga A. A. C.; Shafiq Z. Synthesis, crystal structure analysis, spectral characterization and nonlinear optical exploration of potent thiosemicarbazones based compounds: A DFT refine experimental study. Inorg. Chim. Acta 2019, 486, 162–171. 10.1016/j.ica.2018.10.035. [DOI] [Google Scholar]
  51. Khalid M.; Ullah M. A.; Adeel M.; Usman Khan M.; Tahir M. N.; Braga A. A. C. Synthesis, crystal structure analysis, spectral IR, UV–Vis, NMR assessments, electronic and nonlinear optical properties of potent quinoline based derivatives: Interplay of experimental and DFT study. J. Saudi Chem. Soc. 2019, 23, 546–560. 10.1016/j.jscs.2018.09.006. [DOI] [Google Scholar]
  52. Tahir M. N.; Mirza S. H.; Khalid M.; Ali A.; Khan M. U.; Braga A. A. C. Synthesis, single crystal analysis and DFT based computational studies of 2, 4-diamino-5-(4-chlorophenyl)-6-ethylpyrim idin-1-ium 3, 4, 5-trihydroxybenzoate-methanol (DETM). J. Mol. Struct. 2019, 1180, 119–126. 10.1016/j.molstruc.2018.11.089. [DOI] [Google Scholar]
  53. Haroon M.; Khalid M.; Akhtar T.; Tahir M. N.; Khan M. U.; Saleem M.; Jawaria R. Synthesis, spectroscopic, SC-XRD characterizations and DFT based studies of ethyl2-(substituted-(2-benzylidenehydrazinyl)) thiazole-4-carboxylate derivatives. J. Mol. Struct. 2019, 1187, 164–171. 10.1016/j.molstruc.2019.03.075. [DOI] [Google Scholar]
  54. Rafiq M.; Khalid M.; Tahir M. N.; Ahmad M. U.; Khan M. U.; Naseer M. M.; Braga A. A. C.; Muhammad S.; Shafiq Z. Synthesis, XRD, spectral (IR, UV–Vis, NMR) characterization and quantum chemical exploration of benzoimidazole-based hydrazones: A synergistic experimental-computational analysis. Appl. Organomet. Chem. 2019, 33, e5182 10.1002/aoc.5182. [DOI] [Google Scholar]
  55. Khan B.; Khalid M.; Shah M. R.; Tahir M. N.; Khan M. U.; Ali A.; Muhammad S. Efficient Synthesis by Mono-Carboxy Methylation of 4, 4′-Biphenol, X-ray Diffraction, Spectroscopic Characterization and Computational Study of the Crystal Packing of Ethyl 2-((4′-hydroxy-[1, 1′-biphenyl]-4-yl) oxy) acetate. ChemistrySelect 2019, 4, 9274–9284. 10.1002/slct.201901422. [DOI] [Google Scholar]
  56. Geerlings P.; De Proft F.; Langenaeker W. Conceptual density functional theory. Chem. Rev. 2003, 103, 1793–1874. 10.1021/cr990029p. [DOI] [PubMed] [Google Scholar]
  57. Chakraborty A.; Pan S.; Chattaraj P. K.. Biological activity and toxicity: A conceptual DFT approach. Applications of Density Functional Theory to Biological and Bioinorganic Chemistry; Springer, 2013; pp 143–179. [Google Scholar]
  58. Chattaraj P. K.; Schleyer P. V. R. An ab initio study resulting in a greater understanding of the HSAB principle. J. Am. Chem. Soc. 1994, 116, 1067–1071. 10.1021/ja00082a031. [DOI] [Google Scholar]
  59. Champagne B.; Bishop D. M.. Calculations of nonlinear optical properties for the solid state. Advances in Chemical Physics; John Wiley & Sons, 2003; Vol. 126, pp 41–92. [Google Scholar]
  60. Li Z. a.; Kim H.; Chi S.-H.; Hales J. M.; Jang S.-H.; Perry J. W.; Jen A. K.-Y. Effects of counterions with multiple charges on the linear and nonlinear optical properties of polymethine salts. Chem. Mater. 2016, 28, 3115–3121. 10.1021/acs.chemmater.6b00641. [DOI] [Google Scholar]
  61. Recatalá D.; Llusar R.; Barlow A.; Wang G.; Samoc M.; Humphrey M. G.; Guschin A. L. Synthesis and optical power limiting properties of heteroleptic Mo 3 S 7 clusters. Dalton Trans. 2015, 44, 13163–13172. 10.1039/c5dt01244a. [DOI] [PubMed] [Google Scholar]
  62. Indira J.; Karat P. P.; Sarojini B. K. Growth, characterization and nonlinear optical property of chalcone derivative. J. Cryst. Growth 2002, 242, 209–214. 10.1016/s0022-0248(02)01306-4. [DOI] [Google Scholar]
  63. Yeung M.; Ng A. C. H.; Drew M. G. B.; Vorpagel E.; Breitung E. M.; McMahon R. J.; Ng D. K. P. Facile Synthesis and Nonlinear Optical Properties of Push– Pull 5, 15-Diphenylporphyrins. J. Org. Chem. 1998, 63, 7143–7150. 10.1021/jo971597c. [DOI] [PubMed] [Google Scholar]
  64. Khan M. U.; Khalid M.; Ibrahim M.; Braga A. A. C.; Safdar M.; Al-Saadi A. A.; Janjua M. R. S. A. First Theoretical Framework of Triphenylamine–Dicyanovinylene-Based Nonlinear Optical Dyes: Structural Modification of π-Linkers. J. Phys. Chem. C 2018, 122, 4009–4018. 10.1021/acs.jpcc.7b12293. [DOI] [Google Scholar]
  65. Janjua M. R. S. A.; Khan M. U.; Bashir B.; Iqbal M. A.; Song Y.; Naqvi S. A. R.; Khan Z. A. Effect of π-conjugation spacer (C C) on the first hyperpolarizabilities of polymeric chain containing polyoxometalate cluster as a side-chain pendant: A DFT study. Comput. Theor. Chem. 2012, 994, 34–40. 10.1016/j.comptc.2012.06.011. [DOI] [Google Scholar]
  66. Janjua M. R. S. A.; Amin M.; Ali M.; Bashir B.; Khan M. U.; Iqbal M. A.; Guan W.; Yan L.; Su Z.-M. A DFT Study on The Two-Dimensional Second-Order Nonlinear Optical (NLO) Response of Terpyridine-Substituted Hexamolybdates: Physical Insight on 2D Inorganic–Organic Hybrid Functional Materials. Eur. J. Inorg. Chem. 2012, 2012, 705–711. 10.1002/ejic.201101092. [DOI] [Google Scholar]
  67. Khan M. U.; Ibrahim M.; Khalid M.; Qureshi M. S.; Gulzar T.; Zia K. M.; Al-Saadi A. A.; Janjua M. R. S. A. First theoretical probe for efficient enhancement of nonlinear optical properties of quinacridone based compounds through various modifications. Chem. Phys. Lett. 2019, 715, 222–230. 10.1016/j.cplett.2018.11.051. [DOI] [Google Scholar]
  68. Khan M. U.; Ibrahim M.; Khalid M.; Braga A. A. C.; Ahmed S.; Sultan A. Prediction of Second-Order Nonlinear Optical Properties of D–p–A Compounds Containing Novel Fluorene Derivatives: A Promising Route to Giant Hyperpolarizabilities. J. Cluster Sci. 2019, 30, 415–430. 10.1007/s10876-018-01489-1. [DOI] [Google Scholar]
  69. Khan M. U.; Ibrahim M.; Khalid M.; Jamil S.; Al-Saadi A. A.; Janjua M. R. S. A. Quantum Chemical Designing of Indolo [3, 2, 1-jk] carbazole-based Dyes for Highly Efficient Nonlinear Optical Properties. Chem. Phys. Lett. 2019, 719, 59–66. 10.1016/j.cplett.2019.01.043. [DOI] [Google Scholar]
  70. Parusel A. B. J.; Köhler G.; Grimme S. Density functional study of excited charge transfer state formation in 4-(N, N-dimethylamino) benzonitrile. J. Phys. Chem. A 1998, 102, 6297–6306. 10.1021/jp9800867. [DOI] [Google Scholar]
  71. Khalid M.; Hussain R.; Hussain A.; Ali B.; Jaleel F.; Imran M.; Assiri M. A.; Usman Khan M.; Ahmed S.; Abid S.; Haq S.; Saleem K.; Majeed S.; Jahrukh Tariq C. Electron Donor and Acceptor Influence on the Nonlinear Optical Response of Diacetylene-Functionalized Organic Materials (DFOMs): Density Functional Theory Calculations. Molecules 2019, 24, 2096. 10.3390/molecules24112096. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Kato K. High-efficiency high-power UV generation at 2128 Å in urea. IEEE J. Quantum Electron. 1980, 16, 810–811. 10.1109/jqe.1980.1070575. [DOI] [Google Scholar]
  73. Donaldson W. R.; Tang C. L. Urea optical parametric oscillator. Appl. Phys. Lett. 1984, 44, 25–27. 10.1063/1.94590. [DOI] [Google Scholar]
  74. Halbout J.-M.; Blit S.; Donaldson W.; Chung Tang C. Efficient phase-matched second-harmonic generation and sum-frequency mixing in urea. IEEE J. Quantum Electron. 1979, 15, 1176–1180. 10.1109/jqe.1979.1069900. [DOI] [Google Scholar]
  75. Bäuerle D.; Betzler K.; Hesse H.; Kapphan S.; Loose P. Phase-matched second harmonic generation in urea. Phys. Status Solidi A 1977, 42, K119–K121. 10.1002/pssa.2210420254. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao0c01273_si_001.pdf (2.2MB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES