Next Article in Journal
Modular Nanotransporters Delivering Biologically Active Molecules to the Surface of Mitochondria
Previous Article in Journal
The Design and Optimization of Ceramide NP-Loaded Liposomes to Restore the Skin Barrier
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of the Molecular Architectures of 2-(4-Chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic Acid Using Their Vibrational Spectra, Quantum Chemical Calculations and Molecular Docking with MMP-2 Receptor

by
Mauricio Alcolea Palafox
1,*,
Nataliya P. Belskaya
2 and
Irena P. Kostova
3,*
1
Departamento de Química Física, Facultad de Ciencias Químicas, Universidad Complutense, 28040 Madrid, Spain
2
Department of Technology for Organic Synthesis, Ural Federal University, 19 Mira Str., 620012 Yekaterinburg, Russia
3
Department of Chemistry, Faculty of Pharmacy, Medical University of Sofia, 2 Dunav Str., 1000 Sofia, Bulgaria
*
Authors to whom correspondence should be addressed.
Pharmaceutics 2023, 15(12), 2686; https://doi.org/10.3390/pharmaceutics15122686
Submission received: 16 October 2023 / Revised: 15 November 2023 / Accepted: 22 November 2023 / Published: 27 November 2023

Abstract

:
1,2,3-triazole skeleton is a valuable building block for the discovery of new promising anticancer agents. In the present work, the molecular structure of the synthesized anticancer drug 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic acid (1b) and its anionic form (2b) was characterized by means of the B3LYP, M06-2X and MP2 quantum chemical methods, optimizing their monomer, cyclic dimer and stacking forms using the Gaussian16 program package. The molecular structure was found to be slightly out of plane. The good agreement between the IR and Raman bands experimentally observed in the solid state with those calculated theoretically confirms the synthesized structures. All of the bands were accurately assigned according to functional calculations (DFT) in the monomer and dimer forms, together with the polynomic scaling equation procedure (PSE). Therefore, the effect of the substituents on the triazole ring and the effect of the chlorine atom on the molecular structure and on the vibrational spectra were evaluated through comparison with its non-substituted form. Through molecular docking calculations, it was evaluated as to how molecule 1b interacts with few amino acids of the MMP-2 metalloproteinase receptor, using Sybyl-X 2.0 software. Thus, the relevance of triazole scaffolds in established hydrogen bond-type interactions was demonstrated.

1. Introduction

Despite advances in chemotherapy, drug resistance remains a major clinical concern, creating an urgent need to explore new, more effective and safer anticancer drugs. In this task, the 1,2,3-triazole group is one of the most important classes of nitrogen-rich heterocyclic scaffolds. This heterocyclic class of compounds are considered as amide bioisosteres that can form diverse non-covalent interactions, such as Van der Waals forces and hydrogen bonds with various proteins, enzymes, and receptors with high resistance to enzymatic degradation [1,2]. This fact allows the potential use of these compounds in medicinal chemistry [3], and therefore, they are widely synthesized and tested [3,4,5,6,7]. They have received considerable attention in drug discovery for the development of anticancer agents, especially for lung [5] and breast [7] cancers, and as potential antifungal agents [8]. In addition, the 1,2,3-triazole ring possesses low multidrug resistance, low toxicity, high bioavailability, and stability in both acidic and basic conditions. The nitrogen atom in the 1,2,3-triazole ring is responsible for the enzyme–inhibitor interaction [9]. Recent reviews [3,4,5] have shown an overview of these advances.
In a previous manuscript [10], we described the synthesis, NMR and UV-Vis analysis of four new 1,2,3-triazole derivatives, two of which have now been studied in detail, but from the theoretical and experimental IR and Raman spectroscopy points of view. The selected molecules under study were 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic acid (molecule 1b) and its anionic form (molecule 2b) (Figure 1). These molecules have a strong-accepting group, the carboxylic (or carboxylate) group, in addition to the chlorine atom that can lead to special specific properties. They also have phenyl and pyrrolidine groups that confer liposolubility to the structure, which facilitates passage across the cell membrane.
Carboxylic acids are a rare type of 1,2,3-triazole derivatives. However, they have very attractive functionality, and are widespread in biomolecules. Many azoles, including pyrazoles, isoxazoles, imidazoles, and thiazoles, and some flavone analogs containing this group are widely used in medicine and biology [11,12,13,14,15,16]. Previously, we showed the extremely weak acidity of 2H-1,2,3-triazole-4-carboxylic acid (pKa = 7.65–8.08 [17]) and proposed that they do not induce the acidity of biological fluids [9], toxicity and the risk of unfavorable effects—in particular, idiosyncratic drug-induced liver injury [18,19] will be minimal in further clinical trials.
Vibrational FTIR spectra of related triazole molecules have been reported [6,20], although not in such a detailed form as presented here, nor using accurate scaling procedures. For this purpose, several dimeric forms were optimized at three DFT levels, achieving good agreement with the experimental spectra in the solid state and obtaining the following objectives: (i) to characterize and confirm the molecular structure of the synthesized compounds; (ii) to establish the effect of the different substituents on the vibrational spectra, particularly on the triazole ring, and how they affect the molecular properties; and (iii) through molecular docking calculations, to identify at the molecular level the binding sites of our compounds to active MMP-2 metalloproteinase receptors on cancer cells, and therefore, to look for new substituents with more interactions.
Our further investigations with these compounds are devoted to testing their cytotoxic effects on different tumor cell lines in accordance with literature data showing their potential, as well as to use them as ligands for the synthesis of new lanthanide complexes according to our previous studies on the activity of lanthanide coordination compounds with similar biologically active ligands.

2. Materials and Methods

2.1. Experimental Details

Molecules 1b and 2b were obtained by means of the alkaline hydrolysis of 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carbonitrile according to previously developed procedures [17,21]. 1,2,3-triazole-containing molecules form a white powder at room temperature. The IR spectra were recorded in powder form at the spectral range of 400–4000 cm–1 on a Brüker IFS-66 FTIR spectrometer equipped with a Globar source, a Ge/KBr beam splitter and a TGS detector. For spectrum acquisition, 50 interferograms were collected. The Raman spectrum was recorded in the 50–4000 cm−1 range on a Brüker (Billerica, MA, USA) IFS 66 optical bench with a FRA106 Raman module connected to a microcomputer. The sample was mounted in the sample illuminator using an optical mount without sample pre-treatment of any type. A Nd:YAG laser at 1064 nm was used as an excitation source. The laser power was set at 250 mW and the spectrum was recorded over 1000 scans at room temperature.

2.2. Theoretical Calculations

Computations were mainly carried out using density functional theory (DFT) [22]. DFT calculations on different biomolecules provide results which are quantitatively in good agreement with those obtained at the MP2 level for the same compounds [23,24], thus predicting vibrational wavenumbers more effectively than HF and MP2 methods [25]. In the present study, several levels of theory were used for geometry optimizations and computation of the NBO atomic charges: B3LYP/6-31G(d,p), M06-2X/6-31G(d,p), MP2/6-31G(d,p), B3LYP/6-311++G(3df,pd) and M06-2X/Lanl2dz [26]. All of these methods were implemented in the GAUSSIAN-16 program package [27]. The results obtained were viewed through the GaussView 6.0 graphical interface [28].
In the calculation of the harmonic IR and Raman vibrational wavenumbers, the B3LYP method was mainly used, which is one of the most cost-effective DFT methods [25,29], and for this reason it is the most widely used [30,31,32,33,34]. This method has been successfully used in many studies of biomolecules [35,36,37] and in the drug design field [38,39]. The M06-2X method was also used to optimize the dimer structures in the stacking form of 1b and 2b, because it appears as one of the best options among the meta-generalized gradient functionals for analyzing dispersion-bound systems [40]. However, the accuracy of this method in the calculation of vibrational wavenumbers is slightly lower than B3LYP, although in a previous study [29], we deduced linear scaling equations and polynomic scaling equations that can noticeably improve the calculated wavenumbers.
All optimized structures have shown only positive harmonic vibrations (local energy minima). The MP2 method was mainly used to ensure the stability of all optimized structures. Relative energies were obtained by adding zero-point vibrational energies (ZPEs) to the total energy. The Raman scattering activities (Si) calculated using the Gaussian-16 program were converted into relative Raman intensities (Ii) using the well-known relationship derived from the basic theory of Raman scattering [41].
The 1H-NMR values of 1b in the isolated state were determined at the mPW1PW91/6-311+G(2d,p)//B3LYP/6-311+G(2d,p) level using the Gauge Independent Atomic Orbital (GIAO) method [42]. This level appears to provide precise NMR chemical shifts in organic molecules [43,44], and for this reason it was chosen in the present work. Moreover, this level was implemented in the GaussView 6.0 program, automatically providing the δcalc chemical shift values (in ppm). These δcalc were scaled [45] for the isolated state using the following scaling equation:
δscaled 1H = −0.027 + 1.0933 δcalculated 1H
The SYBYL-X 2.0 software [46] package installed on Windows 10 workstations was used to prepare the protein and ligands [47]. The structures of all molecules were built using the SKETCH model in Sybyl-X 2.0. Geometry optimization was performed using Steepest Descent with Gasteiger–Marsili charges and a Tripos force field. Then, optimization using the conjugate gradient was performed, followed by optimization using BFGS [48].

2.2.1. Calculation of Interaction Energies

The interaction energies were calculated in the dimer form of the molecules under study. These computations were carried out only at the M06-2X/6-31G(d,p) level, and the values obtained were corrected for basis set superposition error, according to the procedure described by Boys and Bernardi [49], which was as follows for the present study.
The total counterpoise (CP) corrected interaction energy of the A and A′ molecules in the dimer AA′ is defined as
Δ E A A C P = A int ( A A ) + E def ( A A )
where
E int   ( A A ) = E A A A A ( A A ) E A A A ( A A ) E A A A ( A A )
In this equation, E A A A A (AA′) corresponds to the calculated electronic energy at the optimized geometry of the dimer (AA′), and E A A A ( A A ) and E A A A (AA′) represent the electronic energy of each molecule, A and A′, of the dimer (AA′).
The deformation energy is described as:
E def ( A A ) = E A d e f ( A A ) + E A d e f ( A A )
where E A d e f ( A A ) and E A d e f ( A A ) of each monomer in the dimer can be determined by the following equations:
E A d e f ( A A ) = E A A ( A A ) E A A ( A )
and
  E A d e f ( A A ) = E A A ( A A ) E A A ( A ) = E A A ( A A ) E A A ( A )
where the parentheses indicate whether the computations in the dimer form were performed for the optimized molecular geometry of each monomer form (A and A′) or that of the dimer (AA′), because the optimized form of each monomer is the same, E A A ( A ) = E A A ( A ) . The superscripts denote whether calculations were performed using the monomer or dimer basis set, and the subscripts indicate the molecular system studied.

2.2.2. Scaling the Wavenumbers

The calculated wavenumbers were scaled to adequately reproduce all experimental wavenumber patterns with enough accuracy, which will markedly improve the obtained spectra [29]. The linear scaling equation procedure was explored, using one (LSE) or two equations (TLSE), in this last case for high and low wavenumbers [29]. These procedures represent a compromise between accuracy and simplicity, and therefore they were the main procedures employed to assign the experimental bands. A new procedure called the polynomic scaling equation procedure (PSE) was also used due to its slight improvement in the calculated wavenumbers than LSE and TLSE. These procedures are based [29] on equations calculated in simpler building molecules, such as the benzene molecule used in the present study. The calculated wavenumbers using the theoretical method were represented by νcal, while the scaled values were represented by νscal. Thus, for the LSE procedure, the equation used [29] was as follows:
νscal = 22.1 + 0.9543 · νcal    at B3LYP/6-31G(d,p) level
The equations corresponding to the TLSE procedure were:
νscal = 29.7 + 0.9509 · νcal    at B3LYP/6-31G(d,p) level for the 1000–3700 cm−1 range
νscal = −16.0 + 1.0009 · νcal    at B3LYP/6-31G(d,p) level for the 0–1000 cm−1 range
The scaling equations used by the PSE procedure were:
νscal = −4.2 + 0.9909 · νcal − 0.00000929· (νcal)2    at B3LYP/6-31G(d,p) level
νscal = 6.5 + 0.9694 · νcal − 0.00000612· (νcal)2    at M06-2X/6-31G(d,p) level
The equation corresponding to the M06-2X/6-31G(d,p) level was only used for the dimer forms of 2b because of the lower accuracy [32] of this M06-2X method than B3LYP in the scaled wavenumbers.

3. Results and Discussion

3.1. Molecular Geometries in the Monomer Form

Several selected optimized geometrical parameters in the monomer form are collected in Table 1, namely bond lengths, bond angles and dihedral angles calculated at different DFT and MP2 levels. The label of the atoms shown in Figure 1 is in agreement with that reported by Safronov et al. [10]. The final optimized structure of the neutral form (1b) is shown in Figure 1a, while the anionic form (2b) is included in Figure 1b. The values of several bond lengths of interest are inserted into these figures. In addition, the total energy (E) values at all calculated levels, including the ZPE (zero-point vibrational energy) correction and the Gibbs energy (G), are shown in each plot.
Rotation of the carboxylic group –COOH leads to two possible conformations in 1b. The most stable one appears with a torsional angle ∠C8-C9-C11=O12 = 16.0° by MP2, and it is shown in Figure 1. Another conformation is about 10 kJ/mol less stable, and therefore, it is included in Figure S1 (Supplementary Material) and will not be analyzed in the present work. When comparing the optimized geometric parameters of these molecules, a large change is observed between the neutral form 1b and its anion 2b. That is, the loss of the carboxylic hydrogen in –COOH has a great effect on the molecular structure.
In these molecules, the phenyl ring has a full planar structure at all calculated levels, as can be seen from the side views of these figures. The triazole ring is also almost full planar, with torsional angle values of less than 1° in the neutral form and less than 1.5° in the anionic form. However, the pyrrolidine ring appears out of plane according to sp3 hybridization of the cyclopyrrol ring carbon atoms. This lack of planarity is more noticeable in the neutral form than in the anionic form. This can be explained by a weaker O12···H18 intramolecular H-bond in the neutral form than in the anionic form, which facilitates its lower planarity.
The three substituents on the triazole ring appear slightly rotated and out of this ring plane: (i) The phenyl ring is rotated with respect to the triazole ring, with C5-C4-N4-N10 torsional angle values being remarkably higher in 1b (−10.3° by MP2) than in 2b (−2.3°). (ii) The pyrrolidine ring plane also appears to be rotated and out of the triazole ring plane with values around 7° by MP2. (iii) The carboxylic group appears remarkably rotated related to the triazole ring, especially in 2b with MP2 values of the C8-C9-C=O12 torsional angle of 29.8°. This value is almost twice that in the neutral form 1b. This can be explained by the greater flexibility of carboxylic oxygens in the anionic form with longer C=O12 bond lengths: 1.267 Å in 2b, which is greater than the corresponding value of 1.225 Å in 1b. This greater flexibility leads to a rotation of O12 to be closer to H18 and therefore to stronger O12···H18 H-bonds in the anionic 2b form than in 1b. The coplanar form with the triazole ring corresponds to a saddle point.
These features were also confirmed by means of calculations carried out in a previous manuscript on these molecules, in which the predicted 1H, 13C-NMR spectra and fluorescent values were in good agreement with the experimental ones [10], and with those found in related triazoles [17,20,21,50]. This non-planarity of the structure is expected to be reduced in the solid state due to intermolecular H-bonds and, therefore, to crystal packing forces that will tend to compress the structure.
The chlorine atom has a very small NBO (natural bond orbital) negative charge (Table 2), which causes a small negative charge to appear on the C1 bonded atom. This feature leads to a small redistribution of the phenyl ring charges, in such a way that the charge on C4 becomes ca. 0.04me more positive in 1b and 2b with the chlorine atom than in its non-Cl-substituted form. Although this change seems small with the chlorine atom, it is enough to result in a slight rotation of the phenyl ring plane related to the triazole ring plane, −10.3° by MP2 in 1b, and only −2.3° in its anionic form 2b. In addition, the chlorine atom increases the carboxylic group rotation as compared to its non-Cl-substituted form, as well as the non-planarity of the pyrrolidine ring (N14-C15-C16-C17 torsional angle).
The loss of the H13 proton in the 2b anionic form leads to a noticeable effect on the molecular structure. The main differences appear in the carboxylic group values, with a greater negative charge on the oxygen atoms in anionic form, and with longer C=O and C9-C11 bond lengths, leading to a greater flexibility of this group. This lengthening of the C9-C11 bond in 2b also leads to an increment in the double bond character of N7=C8 and C9=N10 of the triazole ring, and consequently the C4-N4, N-N and C8-N14 bonds are lengthened. This feature modifies the triazole ring bond angles by ca. 2°, and especially its torsional angles with an increment in the non-planarity of 2b (Table 1).
In these molecules, the highest negative charge corresponds to the hydroxyl oxygen O13. This atom, along with the other oxygen atoms, seems to be the most reactive. The nitrogen atoms N7 and N14 also have high negative charges and are involved in the biological functions of these molecules. The highest positive charge appears on the C8 and C11 carbon atoms because they are bonded to large negative atoms.

3.2. Molecular Geometries in the Dimer Forms

In the solid state, molecule 1b is expected to be symmetrically H-bonded in a cyclic dimer form through the –COOH group, as in related molecules [17] with carboxylic groups [21]. Thus, the dimeric structure of 1b was optimized, as plotted in Figure 2. These predicted optimized dimer forms for the crystal unit cell were confirmed by comparing the carboxylic vibrations of their IR and Raman spectra with those obtained experimentally in the solid-state sample. Both dimer molecules are almost planar and appear H-bonded with the same bond length, 1.646 Å at the B3LYP/6-31G(d,p) level, but slightly shorter by M06-2X and larger with the 6-311++G(3df,pd) basis set. This H-bond value is slightly longer than that reported in the benzoic acid (BA) dimer at the B3LYP/6-31G(d,p) level, 1.616 Å [51]. This is because of the slight decrease in the negative charge on =O12 in 1b as compared to BA. A longer C=O bond (1.240 Å vs. 1.237 Å in BA) and consequently a shorter C9-C11 bond length (1.468 Å vs. 1.487 Å in BA) were calculated. In the dimer formation of 1b, a lengthening of the acceptor C=O bond (1.225 Å in monomer vs. 1.240 Å in dimer) and a shortening of the donor C-OH bond (1.356 Å in monomer vs. 1.319 Å in dimer) were observed, as expected. This shortening in the C-OH bond length of the carboxylic group is proportional to the lengthening of the intermolecular H-bond [52].
2b anion cannot be found as a dimer form in the solid-state crystal through an O-H···O bond as in 1b. However, a comparison of the theoretical scaled vibrational spectra of the monomer with the corresponding experimental ones reveals discrepancies in the stretching vibrations of the –COO and NNN modes. Therefore, to reduce these discrepancies, several stacking forms were optimized with only two molecules in the anionic form. In this model, π–π interactions appear, as well as interactions between the –COO and NNN moieties. In this way, we try to obtain a better calculated IR and Raman spectra and, therefore, an accurate assignment of all bands. Figure 3 shows the three stable forms optimized with the M06-2X method [53]. The most stable dimer corresponds to form I (Figure 3a) because it presents the strongest interaction, and the vibrational spectra were studied only for this form. Form I is confirmed in the optimized structure at the MP2/6-31G(d,p) level included in Figure S2.
Form I appears to be stabilized by several weak H-bonds/interactions through the oxygen atoms and the out-of-plane methylene hydrogens of the pyrrolidine ring. An increment in the twist of the –COO group is found to facilitate these H-bonds/interactions. A high planarity appears between both molecules of this dimer form, and because of this, this structure is expected to be in the crystal unit cell of 2b. Form II (Figure 3b) is slightly less stable than form I, and is also stabilized by weak C-H···O H-bonds/interactions between the –COO group and the pyrrolidine ring hydrogens. In addition, a weak C-H···N H-bond/interaction is observed. However, in this dimer form, the planarity between both molecules is remarkably reduced, and therefore, the crystal packing forces are expected to modify this form. Finally, form III with a nearly orthogonal arrangement of both molecules (Figure 3c) is unreliable for the solid state. Other dimer forms with several H-bonds/interactions through the nitrogen atoms were attempted to be optimized, but they were not stable.
The deformation energy Edef in the 1b dimer, 18.6 kJ/mol, is slightly lower than that in the most stable form I of 2b, 21.8 kJ/mol. This can be explained by a larger number of atoms in the stacking interaction than in the 1b dimer. This deformation energy has similar values in forms I and II of 2b, but it is noticeably lower in form III (14.0 kJ/mol), which is in accordance with the weaker H-bonds in this dimer. The CP-corrected interaction energy ΔECP is −82.8 kJ/mol in 1b, in agreement with that reported in the cyclic dimer form of benzoic acid, −81.7 kJ/mol [51].

3.3. Molecular Properties

Several thermodynamic parameters, rotational constants and dipole moments were also calculated in the global minimum of monomer and dimer forms and included in Table 3. In general, the values computed using B3LYP were close to those computed using M06-2X, with small differences. Rotational constant values were remarkably reduced in the dimer form, around five times lower than those in the monomer form. By contrast, the Cv and entropy (S) values were approximately two times higher in the dimer form than in monomer, as is expected.
Although the calculated values between 1b and 2b were similar, the main difference is in the dipole moment, whose value in the anionic form 2b was remarkably higher than in the neutral form 1b. This feature is in accordance with that found experimentally, with a high water solubility of the anion (form 2b). By contrast, the value in the neutral form 1b is too low and it has low water solubility. In 1b, the dipole moment value is slightly lower in the dimer form than in its monomer form, but in the anion 2b, it is remarkably reduced in its dimer form, in accordance with the –COO group arrangement.

3.4. Vibrational Analysis

The calculated wavenumbers in the global minimum of molecules 1b and 2b at the B3LYP/6-31G(d,p) level are collected in Table S1 (Supplementary Material), while a summary of the wavenumbers with high IR or Raman intensity, or those characteristics of the molecular structure, is included in Table 4 and Table 5. The scaled wavenumbers provided via two procedures, LSE (linear scaling equation) and TLSE (two linear scaling equation) are listed in the second and third columns, respectively. The relative IR and Raman intensities (%), the experimental values in the IR and Raman spectra, and the main characterization of the vibrations with their % contribution of the different modes to a computed value (Potential Electron Distribution, PEDs) are also included in these tables. Contributions lower than 10% are not included. Relative intensities are obtained by normalizing each calculated value to the intensity of the strongest one.
Scaling of the IR and Raman spectra was mainly carried out using the TLSE and PSE procedures. Since the PSE procedure leads to the best results, with errors in general being lower than 3%, it was selected as the main scaling procedure to be utilized. The scaled wavenumbers are slightly worse using the LSE and TLSE procedures than with PSE, with the LSE procedure being the worst. Thus, all scaled spectra shown in the present manuscript were obtained by means of this PSE procedure and used for discussion.
A comparison of the whole FTIR experimental spectra with those corresponding to the theoretical scaled spectra obtained using the PSE procedure in the monomer form was carried out and is plotted in Figure S3, while the same comparison with the Raman values is shown in Figure S4. For a better analysis of the different experimental and scaled vibrational wavenumbers of these figures, the spectra were divided into three regions, namely 3700–2700 cm−1, 1800–1000 cm−1, and 1000–0 cm−1 (or 1000–600 cm−1 by IR). The IR spectra of the two first regions are shown in Figure 4 and Figure 5, while for simplicity, the Raman spectrum is only presented in the 1800–1000 cm−1 range (Figure 6) and the IR spectrum in the 1000–600 cm−1 region in Figure 7. The other regions are shown in Figures S5 and S6. The assignment of the most intense and characteristic IR wavenumbers is included in these figures.
In a general comparison of the IR spectra, the following was observed: (i) A large difference between the spectra of 1b and 2b (Figure S2), which was in agreement with a significant change in the calculated geometric parameters between the neutral and anionic forms. (ii) A noticeable correlation between the scaled wavenumbers in the monomer form and the experimental ones, with only few significant differences, which were reduced with the scaled spectra of the dimer or stacking forms.
It is noted that most modes appear in the expected ranges. Due to this feature and the very small difference between the experimental and scaled values of most fundamentals, the assignments could in general be considered correct. This assignment of the vibrational bands was carried out through a detailed comparison of the experimental with the scaled spectra. This assignment was analyzed under the following sections: (i) the COOH and COO group modes, (ii) the triazole ring modes, (iii) the phenyl ring modes, and (iv) the C-X modes. The discussion was carried out focusing mainly on sections (i) and (iv) because they involved the most reactive groups.

3.4.1. The Carboxylic COOH Group Modes in 1b Molecule

The characteristic displacement vectors of these modes are similar for the monomer and dimer forms, although for each dimer vibration, two wavenumbers appear, one corresponding to the in-phase mode (Raman active) and another one corresponding to the out-of-phase mode (IR active). Table S4 collects the calculated (scaled) wavenumbers, along with the experimental ones and the main characterization with the %PED, but only a summary is included in Table 5. The theoretical values correspond mainly to the monomer form of 1b and 2b, and to a few values in the dimer form but with noticeable differences compared to the monomer form. The experimental IR spectra are characterized by a broad and complex spectral structure around 2850 cm−1, which is characteristic of the carboxylic acid association displayed by hydrogen bonds. The normal modes of the COOH and COO groups were analyzed under the following sections:
The C=O modes: The C=O stretching in the monomer form was calculated with very strong intensity in accordance with the experimental observations, and scaled (by means of PSE) at 1749 cm−1. However, this value noticeably differs (ca. 60–80 cm−1) from that observed experimentally at lower values. This feature is in accordance with a dimer formation in the solid state through the carboxylic –COOH group. Thus, the scaled values in the dimer form of 1b at 1699 cm−1 (IR) and at 1658 cm−1 (Raman) were effectively assigned to the experimental bands at 1690.5 cm−1 (IR) and at 1644.2 cm−1 (Raman), respectively. In the solid state of BA, this mode has been reported [51] at 1693 cm−1 (IR) and 1635 cm−1 (Raman), in agreement with our results.
The C=O in-plane bending is characterized by a strong contribution of the δ(C-OH) mode, and therefore it is denoted as δ(COOH) in Table 5. It was predicted (scaled) in the monomer form at 699 cm−1. However, in the dimer form, it was scaled at 745 cm−1 (IR) with medium IR intensity and at 736 cm−1 (Raman) with a weak value, in excellent accordance with the experimental IR band at 747.4 cm−1 with medium intensity and the Raman band at 739.7 cm−1, also with medium intensity.
The C=O out-of-plane bending was scaled in the monomer at 704 cm−1, very close to that determined in the dimer form at 699 and 697 cm−1. The very low IR and Raman intensities predicted for this mode make its determination difficult in the experimental spectra. Only the very weak IR band at 704.9 cm−1 was assigned to this mode. A large contribution of this mode was also observed in the scaled wavenumber at 772 cm−1.
The C-O13 modes: The C-O stretching is noticeably coupled with ν(C=O) and δ(O-H) modes, as well as with δ(C-H) modes, and it was mainly characterized as νas(COO) in Table 5. In the monomer form, it was scaled with medium IR intensity at 1112 cm−1, in excellent accordance with the experimental bands at 1112.9 cm−1 (IR) and 1117.7 cm−1 (Raman). In the dimer form, it was predicted with very weak intensity at 1255 and 1254 cm−1, and thus it was not detected in the experimental spectra. The IR band in BA at 1347 cm−1 was assigned to this mode [51]. This large difference in the experimental wavenumbers between 1b and BA is due to the longer C-O13 bond in 1b than in BA, in agreement with the shorter O13-H bond in 1b. The νs(COO) mode was scaled in the monomer form at 1091 cm−1 and it was closely related to the experimental IR band at 1088.8 cm−1 and to the Raman line at 1090.7 cm−1. This mode appears to be strongly coupled with the ν(NNN) stretching of the triazole ring, making its identification difficult. In the dimer form, this mode was predicted with very weak intensity, and therefore it was not included in Table 5.
The O13-H modes: A broad and weak IR band at 3516.0 cm−1 is observed in the experimental IR spectrum of 1b, which can only be assigned to ν(O-H) in free COOH groups, based on the scaled wavenumber in the monomer form at 3591 cm−1 with medium IR and Raman intensity, and to the experimental IR band reported at 3553.3 cm−1 in p-methoxy-benzoic acid [54] and in other related molecules [55]. These features are also in accordance with those reported in BA [51], where the monomer form was calculated using B3LYP/6-31G(d,p) at 3763 cm−1 (scaled by PSE at 3593 cm−1), in excellent accordance with our values in 1b. Also, in the experimental IR spectrum of the BA molecule [56,57], a band at 3567 cm−1 has been found with medium intensity, as we observed in the experimental spectrum of 1b, but at 3516.0 cm−1. This lower wavenumber in 1b than in BA indicates an intermolecular bond O-H···O stronger in 1b than in BA. In addition, a very weak and broad experimental IR band is detected at 3420.6 cm−1 which can only be due to the stretching of O-H groups weakly intermolecular H-bonded in contact with other molecules. It means that free and H-bonded COOH groups appear in the solid state.
In the dimer form, the calculated ν(O-H) wavenumber in 1b is observed to be red shifted (scaled) at 3035 cm−1 (IR) and at 2943 cm−1 (Raman), and it is predicted with the strongest intensity in both IR and Raman spectra. This large shift indicates strong H-bonds in the solid-state crystal through the hydroxyl groups. This highest intensity in the dimer form is in accordance with that calculated in the BA dimer [51], and that reported in carboxylic acids [58], where it is usually identified according to a broad stretching band near 3000 cm−1. In our experimental IR spectrum of 1b, a broad band is observed at 3054.1 cm−1, close to our calculations in the dimer form, but with weak–medium intensity. Bands with strong or very strong intensity were not observed around 3000 cm−1. This contradiction, together with the weak experimental bands observed at 3516.0 (IR) and 3599.7 (Raman) and the red-shifted ν(C=O) wavenumber in the dimer form, indicates that in the solid state, not all molecules appear in cyclic dimer forms through the COOH group.
The in-plane bending δ(O-H) appears to be strongly coupled with C=O and C-O modes, as well as with other modes, and it is characterized in Table 5 as δ(COOH). In the monomer form, it appears to be scaled at 1238 cm−1, in accordance with the experimental IR band observed at 1241.1 cm−1 and the Raman line at 1243.0 cm−1. In BA, it was calculated at 1218 cm−1 (scaled at 1189 cm−1), slightly lower than our calculations in 1b, and experimentally reported at 1169 cm−1 (IR) in BA [56,57]. Thus, the higher wavenumber of this mode in 1b than in BA is in agreement with the stronger H-bond in 1b than in BA.
The out-of-plane bending γ(O-H) mode is clearly characterized as an almost pure mode, and it was scaled at 580 cm−1, in accordance with the Raman band at 579.6 cm−1. In the dimer form, it was scaled at 998 and 942 cm−1 with very weak intensity, and therefore, only the very weak Raman line at 932.5 cm−1 could be assigned to this mode. However, in BA, it was reported at 628 cm−1 (in free O-H) and at 960 cm−1 (in O-H bonded) [51,56,57], in accordance with our results, as well as with the band observed at 546 cm−1 in p-methoxybenzoic acid [51].

3.4.2. The Carboxylate COO Group Modes in 2b

The νas(COO) stretching mode was predicted with very strong IR intensity in the monomer form of 2b at 1713 cm−1, far away from the very strong IR band at 1606.6 cm−1. This red shift toward lower wavenumbers in the experimental spectra can be explained by the intermolecular interactions of these molecules through the COO group that lengthened the CO bond. This feature is in accordance with the strong absorption near 1600–1560 cm−1 for asymmetric stretching vibrations reported for the carboxylate group (COO-) in the IR spectra of solid-state samples [58].
Because planar structures cannot be formed in 2b, stacking forms were optimized at the M06-2X/6-31G(d,p) level, which are stabilized by interactions/H-bonds of this COO group (Figure 3). These interactions in the stacking form shift the scaled wavenumbers of this mode, but they are also far away from the experimental values. This feature indicates that the packing crystal forces in the solid state are stronger than those calculated in our simplified optimized model. A shortening in the stacking crystal distance between planes will increase the interactions of this COO group and lead to a lengthening of their CO bond lengths, and therefore the wavenumbers will be closer to the experimental ones.
The symmetric νs(COO) stretching mode appears strongly coupled with other modes and distributed over many calculated vibrations. The highest contribution in the monomer form was determined at 1307 cm−1 with medium-strong IR intensity, in accordance with the IR band observed with medium intensity at 1301.9 cm−1. The values in the dimer form slightly differ from those of the monomer, with scaled values distributed at 1375, 1370, 1362 cm−1 and particularly at 1353 cm−1 with very strong IR intensity.
However, only the experimental IR band with medium intensity at 1369.4 cm−1 and the very weak shoulder at 1348 cm−1 appear close to the scaled values and could be assigned to this νs(COO) mode. Our values also seem to be in strong disagreement with the IR absorption near 1420–1400 cm−1 reported for this symmetric stretching in solid-state samples [58], because in this region, we could only detect a very weak band at 1417.6 cm−1.

3.4.3. The Triazole Ring Modes

The NNN modes: The νas(NNN) stretching appears to be strongly coupled with the ν(C4-N) mode as well as with other ring modes. In 1b, it was scaled at 1386 cm−1 with very weak IR intensity, and therefore it was not observed in the experimental spectrum. A similar wavenumber was calculated in its dimer form because of the small effect of the cyclic dimer on the triazole ring.
In the monomer form of 2b, this νas(NNN) mode is scaled at 1320 cm−1, but due to its very weak predicted IR intensity, it could not be related to an experimental IR band. This mode is red shifted in the stacking form because of the small π–π interaction of the triazole ring with the COO group. Therefore, it appears spread out in the scaled values at 1328, 1106 and 1100 cm−1, with very weak IR intensity. Due to this low predicted IR intensity, it could not be related to experimental bands to confirm this stacking form. The νs(NNN) mode appears scaled at 1360 cm−1 with weak IR intensity and could be related to the IR band at 1340.5 cm−1. However, in the stacking form, this mode shifts and appears at 1425 cm−1 and at 1389 cm−1, the latter being in excellent accordance with the weak IR band at 1391.6 cm−1, which was assigned to this νs(NNN) mode.
The C8-N14 modes: The stretching mode was predicted with the highest IR intensity and medium Raman activity in both 1b and 2b molecules in good agreement with the very strong band observed in the IR spectra and with the medium intensity line in the Raman spectra. It was scaled at 1566 cm−1 in 1b and closely related to the IR band at 1557.4 cm−1 and the Raman line at 1563.2 cm−1. This ν(C8-N14) mode was determined in 2b at similar values although the experimental IR band appears with medium intensity.

3.4.4. The Phenyl Ring Modes

The assignments for several of the phenyl ring modes are obvious and require no further discussion. Because of the negligible influence of the dimer and stacking forms of our simplified models in the calculated wavenumbers of the phenyl ring, their values were omitted from the discussion. The Varsanyi notation [59] for a 1,4-disubstituted benzene was followed for the ring mode assignments. Aromatic C-H stretching vibrations are generally observed in the 3120–3000 cm−1 region, and they are predicted in the molecules under study with almost null IR intensity and very weak Raman intensity, in accordance with those observed experimentally.
A 1H NMR study [10] showed chemical shifts δcalc values, which were correlated with the C-H-stretching wavenumbers. Therefore, the scaled antisymmetric C-H stretching values in the isolated state of 1b obtained using the PSE scaling procedure were closely related to the scaled δcalc 1H NMR values in the isolated state (Figure 8). This good accordance contributes to confirming our assignment. Perhaps, the small discrepancy observed in the C15H2 group of the pyrrolidine ring is due to the fact that its stretching mode appears to be strongly coupled with the C16H2 and C17H2 groups of the ring.
The experimental Raman antisymmetric stretching wavenumbers were also related to the experimental δcalc 1H NMR values (Figure 8b). Because there are not available experimental values in the gas phase, for this relation we use the 1H NMR values reported in methanol [10], although a similar relationship can be obtained with the values in DMSO and 1,4-dioxane. With the exception of the C15H2 group, in general a good relationship with the experimental values was obtained, which also confirms our assignment. For simplicity, the values of 2b were not included in these figures.

3.4.5. The C-X Modes

For a p-disubstituted benzene, the ν(C-X) stretching vibrations corresponding to the substituent are the normal modes 13 and 7a that correspond to ν(C4N) and ν(C-Cℓ) in the molecules under study. C4-N stretching was predicted to be strongly coupled with ν(NNN) stretching and with other modes. Using the displacement vector of the phenyl ring atoms, mode 13 was mainly characterized as ν(C4-N) stretching. In 1b, it appears with medium–strong intensity at 1371 cm−1, in good accordance with the IR band at 1378.1 cm−1 and to the Raman line at 1379.9 cm−1.
The C-Cℓ stretching in 1b is coupled with phenyl ring modes, and using the displacement vectors of the ring atoms, it was assigned to mode 7a. It was scaled at 1082 cm−1 and related to the strong IR band at 1088.8 cm−1 and to the Raman line at 1090.7 cm−1. Due to the strong coupling, this ν(C-Cℓ) mode appears with a noticeably higher wavenumber than in chloro-derivatives [31].

3.5. Molecular Docking Study

1,2,3-triazole derivatives have been reported to have many different pharmacological properties [4]. Among all possibilities, in the present manuscript, we only explored the binding of 1b to the MMP-2 matrix metalloproteinase, which is considered [60] to be one of the main models for angiogenesis and tumor development. Therefore, we can obtain a first idea of whether molecule 1b could act as an effective anticancer drug. For this purpose, the crystallographic structure of MMP-2 was downloaded from Protein Data Bank, PDB ID:1CK7 [60]. The interactions observed between the most stable conformation of our molecule 1b and different amino acids are shown in Figure 9 from two points of view. The final amino acid positions around 1b were in agreement with what was previously reported for similar molecules [6]. The four amino acid residues that present binding sites with 1b were as follows:
  • With aspartic acid (ASP185): one of the carboxylic oxygens of ASP185 is weakly H-bonded (at 2.347 Å) to one of the phenyl ring hydrogens of 1b, as well as to the amino hydrogen of LYS44 (at 2.128 Å). Another carboxylic oxygen of this ASP185 is H-bonded (2.007 Å) to the amino hydrogen atom of the peptide bond of glycine (GLY189), which stabilizes the system.
  • With lysine (LYS44): one of the ternary amino hydrogens (NH3+) of LYS44 is weakly H-bonded (1.979 Å) to nitrogen atom N10 of the triazole ring of 1b, as well as to the carboxylic oxygen of ASP185 (at 2.128 Å), while another of these amino hydrogens is H-bonded (2.012 Å) to the oxygen atom of the peptide bond of valine (VAL41). It is also noted that the amino hydrogen atom of the peptide bond of VAL41 appears to be specially oriented to the phenyl ring center (2.324 Å), which is rich in π electrons, stabilizing the system.
  • With leucine (LEU49): the carboxylic hydrogen H13 of our molecule 1b is H-bonded (1.897 Å) to the oxygen atom of the peptide bond of LEU49, while the carbonyl oxygen O12 of this group is H-bonded to the amino hydrogen atom of valine (VAL51). Both H-bonds fix our molecule 1b to this amino acid chain, stabilizing the system. For this reason, we include several additional amino acids, from LEU49 to TYR53, in Figure 9b. These H-bonds slightly opened the (O=C=O) angle to 123.7° vs. 121.7° in the isolated state, and notably rotated the carboxylic group with a value of the torsional angle (C8-C9-C=O12) of 25.6° vs. 10.8° in the isolated state. The weak intramolecular H-bond O12···H18 calculated in the isolated state (Figure 1a) contributes to this feature. It also contributes to planarity and the disappearance of the H-bonds with amino acid residues. The large flexibility of our molecule 1b facilitates these interactions.
In addition to these binding sites with 1b plotted in Figure 9, there are H-bonds between the amino acid chain residues and away from the position of 1b, which were not included in this figure. Additionally, there are π−π and π−alkyl interactions that stabilize the system. Therefore, the following residues are also included in Figure 9:
-
Tryptophan (TRP176): the phenyl ring of our molecule 1b appears somewhat close to the TRP176 ring, which indicates a weak π−π interaction. In addition, valine (VAL41) is located on the other side of the phenyl ring of 1b, and its hydrophobic chain appears orientated to this phenyl ring with a short distance between them. Therefore, a π−alkyl interaction is expected, which is in accordance with that reported in other 1,2,3-triazole derivatives [6]. These interactions are confirmed by the noticeable rotation of the aryl ring related to the triazole ring, with a torsional angle (C5-C4-N4-N10) of −24.8° vs. −1.7° in the isolated state. The large flexibility of the C4-N4 bond facilitates these interactions, which appear to be responsible for stabilizing TRP176 and VAL41 close to our molecule 1b.
-
The alkyl moiety of leucine (LEU190) residue appears somewhat close to the pyrrolidine ring of 1b, while glycine (GLY189) and leucine (LEU191) do not present interactions with 1b, but they show H-bonds with ASP185 that stabilize the system. The significant change in the endocyclic torsional angles of the pyrrolidine ring of 1b, with a (N14-C15-C16-C17) value of −30.2° (vs. −22.5° in the isolated state), may be due to this LEU190 residue. This pyrrolidine ring also appears to be noticeably rotated related to the triazole ring. Therefore, the (C9-C8-N14-C18) torsional angle has a value of −7.1° vs. 17.2° in the isolated state. The breaking of the weak intramolecular H-bond O12···H18 calculated in the isolated state may also contribute to this feature.

4. Summary and Conclusions

A detailed study of two 1,2,3-triazoles-containing molecules was carried out from the structural, spectroscopic and possible chemotherapeutic activity points of view. The most important findings of this study were as follows:
(1)
A conformational study of the molecules under study was performed with the help of MP2/6-31G(d,p) and several DFT levels. Via rotation on the C9-C11 bond, two conformers were found. The energy difference between both is small, at less than 10 kJ/mol.
(2)
A cyclic dimer formed through the carboxylic –COOH group was optimized in 1b, while several stacking forms were optimized in 2b using M06-2X and MP2 methods.
(3)
To improve the calculated wavenumbers, three DFT levels and two main scaling procedures were used. The PSE procedure leads to the best results, with errors less than 3%. The scaled wavenumbers are slightly worse using the LSE procedure than the TLSE procedure.
(4)
The FT-IR and FT-Raman spectra in the solid state of the molecules under study were recorded. For the first time, an accurate assignment of all observed IR and Raman vibrational bands was carried out. The scaled wavenumbers in the monomer, dimer and stacking forms were in good accordance with the experimental bands, which confirms our simplified optimized structures for the crystal unit cell of the solid state.
(5)
A comparison of the scaled and experimental wavenumbers indicates that in the solid state, free and H-bonded COOH groups appear in 1b, while stacking forms with COO anion are predicted in 2b.
(6)
New relationships between the theoretical scaled and experimental νas(C-H) stretching wavenumbers and the scaled and experimental 1H NMR chemical shifts were established in 1b.
(7)
Molecular docking calculations reveal four H-bonded sites of 1b with ASP185, LYS44, LEU49 and VAL51 amino acids of the MMP-2 matrix metalloproteinase. The carboxylic group seems to play an important role through its two H-bonds to LEU49 and VAL51 residues, which fix our molecule to the amino acid chain stabilizing the system.
These findings show that 2-aryl-1,2,3-triazole acids 1b and 2b may be good candidates as anticancer drugs. In this regard, additional studies are being carried out in our laboratory with other derivatives.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/pharmaceutics15122686/s1, Figure S1: Optimized structure of 1b at the B3LYP/6-31G(d,p) level in another conformation; Figure S2: Two views of the optimized dimer form I of 2b at the MP2/6-31G(d,p) level; Figure S3: Comparison of the theoretical scaled IR spectra at the B3LYP/6-31G(d,p) level with the experimental ones of 1b and 2b molecules in the 3750–500 cm−1 range; Figure S4: Comparison of the theoretical scaled Raman spectra at the B3LYP/6-31G(d,p) level with the experimental ones in the 3750–100 cm−1 range of 1b and 2b molecules. Figure S5: Comparison of the theoretical scaled Raman spectra at the B3LYP/6-31G(d,p) level with the experimental ones of 1b and 2b molecules in the 3700–2700 cm−1 range. Figure S6: Comparison of the theoretical scaled Raman spectra at the B3LYP/6-31G(d,p) level with the experimental ones of 1b and 2b molecules in the 1000–0 and 1000–100 cm−1 ranges, respectively. Table S1: Calculated harmonic wavenumbers (νcal, cm−1), scaled (νscal, cm−1) wavenumbers obtained and the linear scaling equation procedure (LSE) and the polynomic scaling equation procedure (PSE), experimental (Exp) wavenumbers, relative infrared intensity (A) in % and relative Raman intensity (S) in % obtained at the B3LYP/6-31G(d,p) level in 1b and 2b molecules. In the main characterization of the different bands, the number of the ring mode corresponds to Wilson’s notation [59]; Table S2: Scaled 1H NMR Chemical Shifts in 1b with different solvents.

Author Contributions

Conceptualization, M.A.P. and I.P.K.; methodology, M.A.P.; software, M.A.P.; investigation, M.A.P., N.P.B. and I.P.K.; writing—review and editing, M.A.P., N.P.B. and I.P.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the European Union-Next Generation EU, through the National Recovery and Resilience Plan of the Republic of Bulgaria (BG-RRP-2.004-0004-C01).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article and Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bonardi, E.; Chistodoulou, M.S.; Fumagalli, G.; Perdicchia, D.; Rastelli, G.; Passarella, D. The 1,2,3-triazole ring as a bioisostere in medicinal chemistry. Drug Discov. Today 2017, 22, 1572–1581. [Google Scholar] [CrossRef] [PubMed]
  2. Bozorov, K.; Zhao, J.; Aisa, H.A. 1,2,3-Triazole-containing hybrids as leads in medicinal chemistry: A recent overview. Bioorg. Med. Chem. 2019, 27, 3511–3531. [Google Scholar] [CrossRef] [PubMed]
  3. Slavova, K.I.; Todorov, L.T.; Belskaya, N.P.; Alcolea Palafox, M.; Kostova, I.P. Developments in the Application of 1,2,3-Triazoles in Cancer Treatment. Recent Pat. Anti-Cancer Drug Discov. 2020, 15, 92–112. [Google Scholar] [CrossRef] [PubMed]
  4. Alam, M.M. 1,2,3-Triazole hybrids as anticancer agents: A review. Arch. Pharm. 2022, 355, 2100158. [Google Scholar] [CrossRef] [PubMed]
  5. Liang, T.; Sun, X.; Li, W.; Hou, G.; Gao, F. 1,2,3-Triazole-Containing Compounds as Anti–Lung Cancer Agents: Current Developments, Mechanisms of Action, and Structure–Activity Relationship. Front. Pharmacol. 2021, 12, 661173. [Google Scholar] [CrossRef] [PubMed]
  6. Hrimla, M.; Oubella, A.; Laamari, Y.; Lahoucine, B.; Ghaleb, A.; Itto, M.Y.A.; Auhmani, A.; Morjani, H.; Julve, M.; Stiriba, S.-E. Click Synthesis, Anticancer Activity, and Molecular Docking Investigation of some Functional 1,2,3-triazole Derivatives. Biointerface Res. Appl. Chem. 2021, 12, 7633–7667. [Google Scholar] [CrossRef]
  7. Wu, X.; Wang, J.; Xia, S.; Cheng, S.; Shi, Y. 1,2,3-Triazole Derivatives with Anti-breast Cancer Potential. Curr. Top. Med. Chem. 2022, 22, 1406–1425. [Google Scholar] [CrossRef]
  8. Poonia, N.; Kumar, V.; Kumar, A.; Yadav, M.; Lal, K. Recent Progress in 1H-1,2,3-triazoles as Potential Antifungal Agents. Curr. Top. Med. Chem. 2021, 21, 2109–2133. [Google Scholar] [CrossRef]
  9. Khan, S.A.; Akhtar, M.J.; Gogoi, U.; Meenakshi, D.U.; Das, A. An Overview of 1,2,3-triazole-Containing Hybrids and Their Potential Anticholinesterase Activities. Pharmaceuticals 2023, 16, 179. [Google Scholar] [CrossRef]
  10. Safronov, N.E.; Kostova, I.P.; Palafox, M.A.; Belskaya, N.P. Combined NMR Spectroscopy and Quantum-Chemical Calculations in Fluorescent 1,2,3-Triazole-4-Carboxylic Acids Fine Structures Analysis. Int. J. Mol. Sci. 2023, 24, 8947–8966. [Google Scholar] [CrossRef]
  11. Li, J.; Wu, F.; Liu, X.; Zou, Y.; Chen, H.; Li, Z.; Zhang, L. Synthesis and bioevaluation of 1-phenyl-pyrazole-4-carboxylic acid derivatives as potent xanthine oxidoreductase inhibitors. Eur. J. Med. Chem. 2017, 140, 20–30. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, S.; Yan, J.; Wang, J.; Chen, J.; Zhang, T.; Zhao, Y.; Xue, M. Synthesis of some 5-phenylisoxazole-3-carboxylic acid derivatives as potent xanthine oxidase inhibitors. Eur. J. Med. Chem. 2010, 45, 2663–2670. [Google Scholar] [CrossRef] [PubMed]
  13. Chen, S.; Zhang, T.; Wang, J.; Wang, F.; Niu, H.; Wu, C.; Wang, S. Synthesis and evaluation of 1-hydroxy/methoxy-4-methyl-2-phenyl-1H-imidazole-5- carboxylic acid derivatives as non-purine xanthine oxidase inhibitors. Eur. J. Med. Chem. 2015, 103, 343–353. [Google Scholar] [CrossRef]
  14. Song, J.U.; Choi, S.P.; Kim, T.H.; Jung, C.K.; Lee, J.Y.; Jung, S.H.; Kim, G.T. Design and synthesis of novel 2-(indol-5-yl)thiazole derivatives as xanthine oxidase inhibitors. Bioorg. Med. Chem. Lett. 2015, 25, 1254–1258. [Google Scholar] [CrossRef] [PubMed]
  15. Song, J.U.; Jang, J.W.; Kim, T.H.; Park, H.; Park, W.S.; Jung, S.H.; Kim, G.T. Structure based design and biological evaluation of novel 2-(indol-2-yl) thiazole derivatives as xanthine oxidase inhibitors. Bioorg. Med. Chem. Lett. 2016, 26, 950–954. [Google Scholar] [CrossRef] [PubMed]
  16. Mehmood, A.; Ishaq, M.; Zhao, L.; Safdar, B.; Rehman, A.U.; Munir, M.; Raza, A.; Nadeem, M.; Iqbal, W.; Wang, C. Natural compounds with xanthine oxidase inhibitory activity: A review. Chem. Biol. Drug Des. 2019, 93, 387–418. [Google Scholar] [CrossRef]
  17. Safronov, N.E.; Fomin, T.O.; Minin, A.S.; Todorov, L.; Kostova, I.; Benassi, E.; Belskaya, N.P. 5-Amino-2-aryl-1,2,3-triazol-4-carboxylic acids: Synthesis, photophysical properties, and application prospects. Dye. Pigment. 2020, 178, 108343. [Google Scholar] [CrossRef]
  18. Uetrecht, J.; Kobayashi, T.; Iwaki, M.; Nogami, A.; Yoneda, M. Epidemiology and Management of Drug-induced Liver Injury: Importance of the Updated RUCAM. J. Clin. Transl. Hepatol. 2023, 11, 1239–1245. [Google Scholar] [CrossRef]
  19. Uetrecht, J. Are drugs containing a carboxylic acid functional group associated with a significant risk of idiosyncratic drug reactions? J. Mod. Med. Chem. 2020, 8, 56–64. [Google Scholar] [CrossRef]
  20. Safronov, N.E.; Minin, A.S.; Slepukhin, P.A.; Kostova, I.P.; Benassi, E.; Belskaya, N.P. 5-Amino-2-aryl-2H-1,2,3-triazole-4-carboxamides: Unique AIEE-gens and Selective Hg2+ Fluorosensors. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2023, 292, 122419. [Google Scholar] [CrossRef]
  21. Gavlik, K.D.; Lesogorova, S.G.; Sukhorukova, E.S.; Subbotina, J.O.; Slepukhin, P.A.; Benassi, E.; Belskaya, N.P. Synthesis of 2-Aryl-1,2,3-triazoles by Oxidative Cyclization of 2-(Arylazo)ethene-1,1-diamines: A One-Pot Approach. Eur. J. Org. Chem. 2016, 15, 2700–2710. [Google Scholar] [CrossRef]
  22. Seminario, J.M.; Politzer, P. Modern Density Functional Theory: A Tool for Chemistry; Elsevier: Amsterdam, The Netherlands, Volume 2, 1995. [Google Scholar]
  23. Szafran, M.; Katrusiak, A.; Koput, J.; Dega-Szafran, Z. X-ray, MP2 and DFT studies of the structure, vibrational and NMR spectra of homarine. J. Mol. Struct. 2007, 846, 1–12. [Google Scholar] [CrossRef]
  24. Rasheed, T.; Ahmad, S. Computational studies of vibrational spectra and molecular properties of 6-methyluracil using HF, DFT and MP2 methods. Indian J. Phys. 2011, 85, 239–260. [Google Scholar] [CrossRef]
  25. Palafox, M.A. Scaling factors for the prediction of vibrational spectra. I. Benzene molecule. Int. J. Quantum Chem. 2000, 77, 661–684. [Google Scholar] [CrossRef]
  26. Carpenter, J.E.; Weinhold, F. Analysis of the geometry of the hydroxymethyl radical by the “different hybrids for different spins” natural bond orbital procedure. J. Mol. Struct. 1988, 169, 41–62. [Google Scholar] [CrossRef]
  27. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2019. [Google Scholar]
  28. GaussView 6.0; Gaussian Inc.: Wallingford, CT, USA, 2017.
  29. Palafox, M.A. DFT computations on vibrational spectra: Scaling procedures to improve the wavenumbers. In Density Functional Theory: Advances in Applications (Book); Ramasami, P., Ed.; Walter de Gruyter, Inc.: Boston, MA, USA, 2019; Chapter 10; pp. 147–191. [Google Scholar]
  30. Fereyduni, E.; Vessally, E.; Yaaghubi, E.; Sundaraganesan, N. One-pot synthesis, FT-IR, NMR and density functional method (B3LYP) studies on 2-(cyclohexylamino)-2-oxo-1-(pyridin-2-yl)ethyl benzoate. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 81, 64–71. [Google Scholar] [CrossRef] [PubMed]
  31. Mahadevan, D.; Periandy, S.; Ramalingam, S. FT-IR and FT-Raman, vibrational assignments, molecular geometry, ab initio (HF) and DFT (B3LYP) calculations for 1,3-dichlorobenzene. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 79, 962–969. [Google Scholar] [CrossRef]
  32. Arivazhagan, M.; Rexalin, D.A. FT-IR, FT-Raman, NMR studies and ab initio-HF, DFT-B3LYP vibrational analysis of 4-chloro-2-fluoroaniline. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2012, 96, 668–676. [Google Scholar] [CrossRef]
  33. Szafran, M.; Ostrowska, K.; Katrusiak, A.; Dega-Szafran, Z. Spectral and structural studies of dimethylphenyl betaine hydrate. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 128, 844–851. [Google Scholar] [CrossRef]
  34. Arjunan, V.; Devi, L.; Subbalakshmi, R.; Rani, T.; Mohan, S. Synthesis, vibrational, NMR, quantum chemical and structure-activity relation studies of 2-hydroxy-4- methoxyacetophenone. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 130, 164–177. [Google Scholar] [CrossRef]
  35. Ponomareva, A.G.; Yurenko, Y.P.; Zhurakivsky, R.O.; van Mourik, T.; Hovorun, D.M. Complete conformational space of the potential HIV-1 reverse transcriptase inhibitors d4U and d4C. A quantum chemical study. Phys. Chem. Chem. Phys. 2012, 14, 6787–6795. [Google Scholar] [CrossRef] [PubMed]
  36. Shishkin, O.V.; Gorg, L.; Zhikol, O.A.; Leszczynski, J. Conformational Analysis of Canonical 2-Deoxyribonucleotides. 2. Purine Nucleotides. J. Biomol. Struct. Dyn. 2004, 22, 227–243. [Google Scholar] [CrossRef] [PubMed]
  37. Brovarets’, O.O.; Hovorun, D.M. Can tautomerization of the A·T Watson–Crick base pair via double proton transfer provoke point mutations during DNA replication? A comprehensive QM and QTAIM analysis. J. Biomol. Struct. Dyn. 2014, 32, 127–154. [Google Scholar] [CrossRef] [PubMed]
  38. Koyambo-Konzapa, S.-J.; Kongbonga, G.Y.M.; Premkumar, R.; Vamhindi, R.; Nsangou, M.; Benial, A.M.F. Spectroscopic, quantum chemical, molecular docking and molecular dynamics investigations of hydroxylic indole-3-pyruvic acid: A potent candidate for nonlinear optical applications and Alzheimer’s drug. J. Biomol. Struct. Dyn. 2022, 40, 10651–10664. [Google Scholar] [CrossRef] [PubMed]
  39. Nisha, S.V.D.; Joe, I.H. Quantum chemical computation and spectroscopic investigation on antiviral drug Acyclovir:-In-silico and in-vitro analysis. J. Mol. Struct. 2021, 1233, 130033. [Google Scholar] [CrossRef]
  40. Zhao, Y.; Truhlar, D.G. Applications and validations of the Minnesota density functionals. Chem. Phys. Lett. 2011, 502, 1–13. [Google Scholar] [CrossRef]
  41. Srivastav, G.; Yadav, B.; Kumar Yadav, R.; Yadav, R.A. DFT studies of molecular structures conformers and vibrational characteristics of sulfanilamide. Comput. Theor. Chem. 2019, 1167, 112588. [Google Scholar] [CrossRef]
  42. Ditchfield, R. Self-consistent perturbation theory of diamagnetism I. A gauge-invariant LCAO method for NMR chemical shifts. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  43. Lodewyk, M.W.; Tantillo, D.J. Prediction of the Structure of Nobilisitine A Using Computed NMR Chemical Shifts. J. Nat. Prod. 2011, 74, 1339–1343. [Google Scholar] [CrossRef]
  44. de Albuquerque, A.C.F.; Ribeiro, D.J.; de Amorim, M.B. Structural determination of complex natural products by quantum mechanical calculations of 13C NMR chemical shifts: Development of a parameterized protocol for terpenes. J. Mol. Model. 2016, 22, 183. [Google Scholar] [CrossRef]
  45. Pierens, K. 1H and 13C NMR Scaling Factors for the Calculation of Chemical Shifts in Commonly Used Solvents Using Density Functional Theory. J. Comput. Chem. 2014, 35, 1388–1394. [Google Scholar] [CrossRef] [PubMed]
  46. Tripos Inc. St. Louis, MO, USA. Available online: https://en.freedownloadmanager.org/users-choice/Sybyl-x_2.0.html (accessed on 21 November 2023).
  47. BIOVIA Studio Modeling Environment. Dassault Systemes BIOVIA; BIOVIA Studio Modeling Environment: San Diego, CA, USA, 2017. [Google Scholar]
  48. Hu, R.; Barboult, F.; Delamar, M.; Zhang, R. Receptor- and Ligand-Based 3D-QSAR Study for a Series of Non-Nucleoside HIV-1 Reverse Transcriptase Inhibitors. Bioorg. Med. Chem. 2009, 17, 2400–2409. [Google Scholar] [CrossRef]
  49. Boys, S.F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  50. Gavlik, K.D.; Sukhorukova, E.S.; Shafran, Y.M.; Slepukhin, P.A.; Benassi, E.; Belskaya, N.P. 2-Aryl-5-amino-1,2,3-triazoles: New effective blue-emitting fluorophores. Dye. Pigment. 2017, 136, 229–242. [Google Scholar] [CrossRef]
  51. Alcolea Palafox, M.; Núñez, J.L.; Gil, M. Theoretical Quantum Chemical Study of Benzoic Acid: Geometrical Parameters and Vibrational Wavenumbers. Int. J. Quantum Chem. 2002, 89, 1–24. [Google Scholar] [CrossRef]
  52. Jeffrey, G.A.; Saenger, W. Hydrogen Bonding in Biological Structures; Springer: Berlin/Heidelberg, Germany, 1991; Chapter 5. [Google Scholar]
  53. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  54. Alcolea Palafox, M.; Gil, M.; Núñez, J.L. Spectroscopy of p-methoxybenzoic acid: An AM1 and ab initio study. Appl. Spectrosc. 1994, 48, 27–36. [Google Scholar] [CrossRef]
  55. Xavier, T.S.; Joe, I.H.; Palafox, M.A.; Kumar, S.; Rastogi, V.K. Vibrational spectral investigations and density functional theory study of 4-Formylbenzoic acid. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2013, 114, 502–508. [Google Scholar] [CrossRef]
  56. Reva, I.D.; Stepanian, S.G. An infrared study on matrix-isolated benzoic acid. J. Mol. Struct. 1995, 349, 337–340. [Google Scholar] [CrossRef]
  57. Stepanian, S.G.; Reva, I.D.; Radchenko, E.D.; Sheina, G.G. Infrared spectra of benzoic acid monomers and dimers in argon matrix. Vib. Spectrosc. 1996, 11, 123–133. [Google Scholar] [CrossRef]
  58. Socrates, G. Infrared and Raman Characteristic Group Frequencies, Tables and Charts, 3rd ed.; Wiley: New York, NY, USA, 2004. [Google Scholar]
  59. Varsányi, G. Assignment for Vibrational Spectra of Seven Hundred Benzene Derivatives; Hilder, A., Ed.; Wiley: London, UK, 1974; Volume 1–2. [Google Scholar]
  60. Jezierska, A.; Motyl, T. Matrix metalloproteinase-2 involvement in breast cancer progression: A mini-review. Med. Sci. Monit. 2009, 15, RA32–RA40. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Labeling of the atoms and plot of the structure with front and lateral views forms of (a) 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic acid (in short 1b), where several bond lengths and intramolecular hydrogen bond (H-bond) values of interest are calculated at the MP2/6-31G(d,p) level and included in the figure, as well as the total energy of the system (E) including zero-point correction and the Gibbs free energy (G); (b) 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylate anion (in short 2b).
Figure 1. Labeling of the atoms and plot of the structure with front and lateral views forms of (a) 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic acid (in short 1b), where several bond lengths and intramolecular hydrogen bond (H-bond) values of interest are calculated at the MP2/6-31G(d,p) level and included in the figure, as well as the total energy of the system (E) including zero-point correction and the Gibbs free energy (G); (b) 2-(4-chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylate anion (in short 2b).
Pharmaceutics 15 02686 g001
Figure 2. Two views of the optimized dimer form of 1b. Several bond lengths and H-bond values of interest calculated at the B3LYP/6-31G(d,p) level (in black colour), the B3LYP/6-311++G(3df,pd) level (in violet colour) and the M06-2X/6-31G(d,p) level (in green colour) are also included. 1 AU = 2625.5 kJ/mol.
Figure 2. Two views of the optimized dimer form of 1b. Several bond lengths and H-bond values of interest calculated at the B3LYP/6-31G(d,p) level (in black colour), the B3LYP/6-311++G(3df,pd) level (in violet colour) and the M06-2X/6-31G(d,p) level (in green colour) are also included. 1 AU = 2625.5 kJ/mol.
Pharmaceutics 15 02686 g002
Figure 3. The three optimized dimer forms of 2b. Several bond lengths and H-bond values of interest calculated at the M06-2X/6-31G(d,p) level are included in the figure. ΔE and ΔG energy values are related to the most stable form I.
Figure 3. The three optimized dimer forms of 2b. Several bond lengths and H-bond values of interest calculated at the M06-2X/6-31G(d,p) level are included in the figure. ΔE and ΔG energy values are related to the most stable form I.
Pharmaceutics 15 02686 g003
Figure 4. Comparison of the scaled IR spectra of 1b and 2b molecules obtained using the PSE procedure in the 3700–2700 cm−1 range with the experimental ones in the 3700–2200 cm−1 range. The most intense and characteristic vibrations are assigned.
Figure 4. Comparison of the scaled IR spectra of 1b and 2b molecules obtained using the PSE procedure in the 3700–2700 cm−1 range with the experimental ones in the 3700–2200 cm−1 range. The most intense and characteristic vibrations are assigned.
Pharmaceutics 15 02686 g004
Figure 5. Comparison of the scaled IR spectra of 1b and 2b molecules obtained using the PSE procedure with the experimental ones in the 1800–1000 cm−1 range.
Figure 5. Comparison of the scaled IR spectra of 1b and 2b molecules obtained using the PSE procedure with the experimental ones in the 1800–1000 cm−1 range.
Pharmaceutics 15 02686 g005
Figure 6. Comparison of the scaled Raman spectra of 1b and 2b molecules obtained using the PSE procedure with the experimental ones in the 1800–1000 cm−1 range.
Figure 6. Comparison of the scaled Raman spectra of 1b and 2b molecules obtained using the PSE procedure with the experimental ones in the 1800–1000 cm−1 range.
Pharmaceutics 15 02686 g006
Figure 7. Comparison of the scaled IR spectra of 1b and 2b molecules obtained using the PSE procedure in the 1000–400 cm−1 range with the experimental ones in the 1000–600 cm−1 range. The most intense and characteristic vibrations are assigned.
Figure 7. Comparison of the scaled IR spectra of 1b and 2b molecules obtained using the PSE procedure in the 1000–400 cm−1 range with the experimental ones in the 1000–600 cm−1 range. The most intense and characteristic vibrations are assigned.
Pharmaceutics 15 02686 g007
Figure 8. Relationships in 1b between (a) the theoretical calculated 1H chemical shifts in the isolated state at the mPW1PW91/6-311+G(2d,p)//B3LYP/6-311+G(2d,p) level and further scaled versus the theoretical scaled stretching ν(C-H) wavenumbers at the B3LYP/6-31G(d,p) level and (b) the scaled and experimental 1H chemical shifts and the related stretching ν(C-H) wavenumbers.
Figure 8. Relationships in 1b between (a) the theoretical calculated 1H chemical shifts in the isolated state at the mPW1PW91/6-311+G(2d,p)//B3LYP/6-311+G(2d,p) level and further scaled versus the theoretical scaled stretching ν(C-H) wavenumbers at the B3LYP/6-31G(d,p) level and (b) the scaled and experimental 1H chemical shifts and the related stretching ν(C-H) wavenumbers.
Pharmaceutics 15 02686 g008
Figure 9. Optimized structure at the B3LYP/6-31G(d,p) level of the system with several amino acids of the MMP-2 receptor forming H-bonds with the 1b molecule from two viewpoints: (a) side and (b) front.
Figure 9. Optimized structure at the B3LYP/6-31G(d,p) level of the system with several amino acids of the MMP-2 receptor forming H-bonds with the 1b molecule from two viewpoints: (a) side and (b) front.
Pharmaceutics 15 02686 g009
Table 1. Several selected optimized geometrical parameters calculated in the monomer form at different levels. Bond lengths (r) are in Å, bond angles and dihedral angles (∠) are in degrees.
Table 1. Several selected optimized geometrical parameters calculated in the monomer form at different levels. Bond lengths (r) are in Å, bond angles and dihedral angles (∠) are in degrees.
ParametersB3LYPM06-2X cMP2 a
1b a1b b2b a2b b2b1b2b
r(N4-N7)1.3501.3451.3611.3551.3821.3401.350
r(N4-N10)1.3131.3061.3491.3451.3651.3341.352
r(C8-C9)1.4431.4381.4531.4451.4611.4231.430
r(C9-N10)1.3441.3381.3251.3161.3371.3571.348
r(C9-C11)1.4681.4651.5491.5371.5351.4671.541
r(CL···O12)10.28110.22410.1609.99710.25310.22710.179
∠(C4-N4-N7)121.7121.9122.4122.7122.8121.5121.9
∠(C4-N4-N10)122.4122.6122.7122.9122.9121.7122.0
∠N-N-N)115.9115.5115.0114.4114.2116.8116.1
∠(N10-C9-C11)119.5119.7121.2122.7121.2119.2120.7
∠(C9-C8-N14)133.2132.9130.8130.0131.4132.7130.9
∠(C9-C11=O12)126.0125.6113.4113.0113.9125.6113.3
∠(C9-C11-O13)112.3112.6115.9116.5116.6111.8115.4
∠(O=C=O)121.7121.8130.7130.4129.5122.6131.3
∠(C5-C4-N4-N10)−1.7−2.5−0.8−0.9−1.3−10.3−2.3
∠(N4-N10-C9-C11)175.8174.3175.0176.3175.7173.7175.9
∠(N10-N4-N7-C8)−0.4−0.5−1.1−0.9−0.4−0.8−1.4
∠(N10-C9-C=O12)−164.9−161.4−139.2−121.5−140.9−156.3−145.6
∠(C8-C9-C=O12)10.812.535.254.534.116.029.8
∠(C11-C9-C8-N14)4.86.55.43.86.98.03.1
∠(C9-C8-N14-C18)17.221.414.810.74.924.932.0
∠(C8-N14-C15-C16)−163.5−162.4−171.6−168.3−170.3−152.0−162.4
∠(N14-C15-C16-C17)−22.5−20.7−18.3−21.2−25.1−23.6−9.3
With the basis set: a 6-31G(d,p). b 6-311++G(3df,pd). c Lanl2dz.
Table 2. Natural atomic charges (in me) calculated in the monomer form at the MP2/6-31G(d,p) level.
Table 2. Natural atomic charges (in me) calculated in the monomer form at the MP2/6-31G(d,p) level.
Atom1b2bAtom1b2b
CL−0.012−0.054N10−0.183−0.240
C1−0.052−0.080C110.9790.955
C40.1600.210=O12−0.720−0.887
N4−0.005−0.073O13−0.775−0.839
N7−0.381−0.402N14−0.560−0.557
C80.4620.419C18−0.2110.219
C9−0.0910.031H180.2630.297
Table 3. Molecular properties calculated at the (a) B3LYP/6-31G(d,p) level, (b) M06-2X/6-31G(d,p) level, and (c) B3LYP/6311++G(3df,pd) level.
Table 3. Molecular properties calculated at the (a) B3LYP/6-31G(d,p) level, (b) M06-2X/6-31G(d,p) level, and (c) B3LYP/6311++G(3df,pd) level.
Molecular Properties1b2b
(a)(b)(c)(a)(b)(c)
monomerRotational constants:
(GHz)
A0.6140.6230.6190.6490.667 a0.642
B0.1380.1390.1390.1360.135 a0.134
C0.1130.1150.1140.1140.114 a0.112
Cv (cal/mol·K)64.7363.8864.8563.8864.01 a62.84
S (cal/mol·K)138.90136.12138.77138.41139.6 a136.6
Dipole moment (Debye) 2.4582.2372.41411.36611.64 a12.99
dimerRotational constants:
(GHz)
A0.1170.1220.114-0.206-
B0.0230.0230.023-0.030-
C0.0200.0200.020-0.029-
Cv (cal/mol·K)133.1131.2133.2-131.7-
S (cal/mol·K)241.5234.0240.1-230.2-
Dipole moment (Debye) 2.3902.4662.719-4.070-
a At the M06-2X/Lanl2dz level.
Table 4. Calculated harmonic wavenumbers (νcal, cm−1), relative infrared intensities (A, %) and relative Raman scattering activities (S, %) obtained at the B3LYP/6-31G(d,p) level in 1b and 2b molecules.
Table 4. Calculated harmonic wavenumbers (νcal, cm−1), relative infrared intensities (A, %) and relative Raman scattering activities (S, %) obtained at the B3LYP/6-31G(d,p) level in 1b and 2b molecules.
νcal
1b
TLSEPSEASExperimental 1bCharacterization of 1b
νscalνscalIRRaman
3123299930007112973.1 m2990.3 vwνas(C-H) in C16H2, C17H2 pyrrolidine
1650159916050100 1594.0 vs8a, ν(C=C) (100)
160915601566100201557.4 vs1563.2 mν(C8-N14) (70) + νs(N7CC) (18)
15411495150130321500.3 sh1497.6 s19a, ν(CC,CH) in aryl (85)
1512146714732561481.2 vs δ(C-H) (40) + ν(C-C11) (35)
150414601465221462.0 m1451.3 vwδs(C-H) out-of-phase in pyrrolidine (98)
142213821386129 1412.7 vwνas(NNN)(45) + ν(CN)(25) + δ(CH)(20)
14061367137118431378.1 m1379.9 vsν(C4N)(35) + ν(NNN)(30) + δ(CH)(25)
130212681270031269.1 m1268.1 wνas(NN)(65) + 3, δ(CH) in aryl (25)
121511851186201197.7 m1196.7 vwγas(C-H) out-of-phase in pyrrolidine (88)
119811691170071163.0 m1167.8 m9a, δ(C-H) in aryl (80)
118611571158321148.5 m1133.1 wγas(C-H) in pyrrolidine (81)
1108108310821371088.8 s1090.7 mν(C-CL) (60) + 18a, δ(CC) in aryl (40)
102910081006241012.6 m1011.6 w12, δ(CC, CH) in aryl (90)
9889739661510968.2 m967.3 sνs(NNN)(43) + 12,δ(CC)(25) + νs(CN14C) (20)
88987486901--870.8 mγas(C-H) in pyrrolidine (98)
84983483080825.5 s836.1 vw17b, γ(C-H) in aryl (100)
νcal
2b
TLSEPSEASExperimental 2bCharacterization of 2b
νscalνscalIRRaman
16521601160724100 1594.9 vs8a, ν(C=C) (89)
159915501556100261543.9 m1550.6 mν(C8-N14) (73) + νs(N7CC) (15)
1545149915057651497.6 s1498.6 vsν(CN4) (37) + 19a,ν(CC)(35) + δ(CH)pyrrolidine (15)
15101466147110111471.6 m1471.6 mνs(CN)in triazol (42) + δs(C-H)in pyrrolidine (35)
150114571462021462.9 m1466.7 wδs(C-H) out-of-phase in pyrrolidine (99)
1436139514000101391.6 w1395.4 s19a,ν(CC,CH) in aryl (40) + ν(C4N)(32)
1400136113652951369.4 m1373.2 vsγs(C-H) in-phase in pyrrolidine (93)
1395135613601011340.5 m1345.2 wνs(NNN) (45) + γs(CC,CH) in pyrrolidine (35)
13541317132032 1318.2 vwνas(NNN) (36) + γ(CH) in pyrrolidine (30)
132012851288101277.8 m1280.6 w3, δ(C-H) in aryl (97)
125812261228381217.0 w1232.4 wν(NN,CN) (68) + γas(CC,CH) in pyrrolidine (19)
12301199120141 1200.6 vwνas(NN,CN) (55) + 3, δ(CH) in aryl (22)
121211821183901182.3 vw1192.9 vwνas(NN,CN) (48) + ν(CN,CC) in pyrrolidine (35)
10991075107420 1079.1 vwνs(NNN) (48) + 18a, δ(C-H) in aryl (42)
1097107310722141089.7 m1089.7 sν(C-CL) (32) + 1, δ(CC) (30) + νas(NNN) (28)
1018998995141008.7 w1013.5 m12, δ(CC, CH) in aryl (96)
99397897112 970.1 vsδ(CC,CH) in pyrrolidine (73)
943928922261970.1 vs923.9 vwνs(NNN) (60) + ν(C8C)(25) + ν(CC) (15)
85283783310826.5 vs830.3 vwγas(C-H) in pyrrolidine (97)
84182682370814.9 vs821.7 w17b, γ(C-H) in aryl (99)
56454955251 555.5 wν(CCL) (38) + δ(NNN)(32) + 6b, δ(CC) (15)
44943343930 446.6 vwδ(CCL) (32) + δ(CC)(25) + δ(NNN) (22)
Observed frequencies characterized by notation: vs = very strong, s = strong, m = medium, w = weak band, vw = very weak, sh = shoulder. νas: anti-symmetric stretching, νs: symmetric stretching, δ: in-plane bending, γ: out-of-plane bending.
Table 5. Calculated harmonic wavenumbers (νcal, cm−1), relative infrared intensities (A, %), relative Raman intensities (S, %) together with the scaled values (νscal, cm−1) in the COOH (of 1b) and COO (of 2b) groups. The values in italic type correspond to the dimer form.
Table 5. Calculated harmonic wavenumbers (νcal, cm−1), relative infrared intensities (A, %), relative Raman intensities (S, %) together with the scaled values (νscal, cm−1) in the COOH (of 1b) and COO (of 2b) groups. The values in italic type correspond to the dimer form.
GroupModeνcalASTLSEPSEIRRamanCharacterization
νscalνscal
COOHν(O-H) 37611517360635913516.0 w3599.7 vvwν(O-H) (100)
31611000303530353054.1 w-m Dimer: ν(O-H) out-of-phase
δ(O-H)1269102123612381241.1 m1243.0 vwδ(OH)(40) + ν(CN)(30) + γas(CH)pyrrolidine (20)
γ(O-H)59391578580 579.6 mγ(O-H)(82)
96400949942 932.5 vwDimer: γ(O-H) out-of-phase
ν(C=O)1747180169116991690.5 vs Dimer: ν(C=O) out-of-phase
170504916511658 1644.2 wDimer: ν(C=O) in-phase
ν(C-O)1139371111311121112.9 vw1117.7 vwνas(COO)(38) + δ(C-H)(25) + δ(OH) (17)
δ(C-O)1117151109210911088.8 s1090.7 mν(NNN)(38)+νs(COO)(30)+18a,δ(CH) (25)
80901794791 799.5 vwDimer: δas(COO) in-phase
δ(C=O)76210747745747.4 m Dimer: δ(COOH) out-of-phase + ν(CC)
75202737736 739.7 mDimer: δ(COOH) in-phase
γ(C=O)71400699699704.9 vw Dimer: γ(C=O) out-of-phase
71200697697 705.9 wDimer: γ(C=O) in-phase
COOνas1762753170517131606.6 vs1613.3 vwνas(COO) (99)
νs1340468130413071301.9 m1306.7 wνs(COO)(63) + γas(CH) in pyrrolidine (18)
δas797172782780782.1 vvw767.7 vwδas(COO) (57) + γ(CC) (32)
γs81120796793 789.8 vwγs(COO) (54) + γs(CC9C) (35)
Notation used for experimental bands: vs = very strong, s = strong, m = medium, w = weak band, vw = very weak, vvw = very very weak.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Palafox, M.A.; Belskaya, N.P.; Kostova, I.P. Study of the Molecular Architectures of 2-(4-Chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic Acid Using Their Vibrational Spectra, Quantum Chemical Calculations and Molecular Docking with MMP-2 Receptor. Pharmaceutics 2023, 15, 2686. https://doi.org/10.3390/pharmaceutics15122686

AMA Style

Palafox MA, Belskaya NP, Kostova IP. Study of the Molecular Architectures of 2-(4-Chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic Acid Using Their Vibrational Spectra, Quantum Chemical Calculations and Molecular Docking with MMP-2 Receptor. Pharmaceutics. 2023; 15(12):2686. https://doi.org/10.3390/pharmaceutics15122686

Chicago/Turabian Style

Palafox, Mauricio Alcolea, Nataliya P. Belskaya, and Irena P. Kostova. 2023. "Study of the Molecular Architectures of 2-(4-Chlorophenyl)-5-(pyrrolidin-1-yl)-2H-1,2,3-triazole-4-carboxylic Acid Using Their Vibrational Spectra, Quantum Chemical Calculations and Molecular Docking with MMP-2 Receptor" Pharmaceutics 15, no. 12: 2686. https://doi.org/10.3390/pharmaceutics15122686

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

Article Metrics

Back to TopTop