Next Article in Journal
Comparative Analysis of Hydrosol Volatile Components of Citrus × Aurantium ‘Daidai’ and Citrus × Aurantium L. Dried Buds with Different Extraction Processes Using Headspace-Solid-Phase Microextraction with Gas Chromatography–Mass Spectrometry
Previous Article in Journal
Unveiling Allosteric Regulation and Binding Mechanism of BRD9 through Molecular Dynamics Simulations and Markov Modeling
Previous Article in Special Issue
Novel TiO2-Supported Gold Nanoflowers for Efficient Photocatalytic NOx Abatement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photochemical Generation and Characterization of C-Aminophenyl-Nitrilimines: Insights on Their Bond-Shift Isomers by Matrix-Isolation IR Spectroscopy and Density Functional Theory Calculations

1
University of Coimbra, CQC-IMS, Faculty of Pharmacy, 3004-295 Coimbra, Portugal
2
University of Coimbra, CQC-IMS, Department of Chemistry, 3004-535 Coimbra, Portugal
3
Istanbul Kultur University, Faculty Sciences and Letters, Department of Physics, Bakirkoy, 34158 Istanbul, Turkey
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(15), 3497; https://doi.org/10.3390/molecules29153497
Submission received: 11 July 2024 / Revised: 20 July 2024 / Accepted: 22 July 2024 / Published: 25 July 2024
(This article belongs to the Special Issue Feature Papers in Photochemistry and Photocatalysis)

Abstract

:
The intriguing ability of C-phenyl-nitrilimine to co-exist as allenic and propargylic bond-shift isomers motivated us to investigate how substituents in the phenyl ring influence this behavior. Building on our previous work on the meta- and para-OH substitution, here we extended this investigation to explore the effect of the NH2 substitution. For this purpose, C-(4-aminophenyl)- and C-(3-aminophenyl)-nitrilimines were photogenerated in an argon matrix at 15 K by narrowband UV-light irradiation (λ = 230 nm) of 5-(4-aminophenyl)- and 5-(3-aminophenyl)-tetrazole, respectively. The produced nitrilimines were further photoisomerized to carbodiimides via 1H-diazirines by irradiations at longer wavelengths (λ = 380 or 330 nm). Combining IR spectroscopic measurements and DFT calculations, it was found that the para-NH2-substituted nitrilimine exists as a single isomeric structure with a predominant allenic character. In contrast, the meta-NH2-substituted nitrilimine co-exists as two bond-shift isomers characterized by propargylic and allenic structures. To gain further understanding of the effects of phenyl substitution on the bond-shift isomerism of the nitrilimine fragment, we compared geometric and charge distribution data derived from theoretical calculations performed for C-phenyl-nitrilimine with those performed for the derivatives resulting from NH2 (electron-donating group) and NO2 (electron-withdrawing group) phenyl substitutions.

1. Introduction

Nitrilimines (R′–CNN–R″), first reported by Huisgen and co-workers [1], constitute highly versatile intermediates in organic synthesis, particularly in regio- and stereoselective 1,3-cycloaddition reactions, yielding five-membered heterocycles [2,3,4,5,6,7]. Their utility spans across diverse disciplines, from general synthetic chemistry to bioorthogonal chemistry and materials science [6,8,9,10]. These 1,3-dipolar species are typically generated in situ by photolysis or pyrolysis of tetrazoles [11,12,13]. Due to the highly reactive nature of nitrilimines, matrix-isolation infrared spectroscopy, in conjugation with ab initio calculations, is one of the most effective methodologies for their characterization [3,14,15,16,17,18,19,20].
In general, nitrilimines are monoisomeric species and, depending on the R′ and R″ substituents, exhibit predominantly propargylic or allenic structures [3,16]. However, some nitrilimines with amino (R′ = NH2, R″ = H, or NH2), hydroxy (R′ = H or OH, R″ = OH), and fluoro (R′ = F, R″ = H, or F) substitutions have been shown to possess a significant carbenic character [19,21,22] (Scheme 1). Experimental identification of nitrilimines with distinct structural characteristics can be accomplished by analyzing the infrared frequency of the antisymmetric CNN stretching vibration [3,16,19,21]. Frequencies in the 2300–2150 cm−1 range indicate nitrilimines predominantly propargylic, possessing a CN triple bond, while frequencies below the 2150–2000 cm−1 range denote allenic-type nitrilimines with a cumulenic CNN moiety. Nitrilimines characterized by weak absorption peaks below 2000 cm−1 have a strong carbenic character.
A particularly puzzling behavior has been observed for C-phenyl-nitrilimine (R′ = Ph, R″ = H). This molecule was theoretically predicted to exhibit two minima on the potential energy surface: one characterized by an allenic-like structure and the other by a propargylic-like structure [16] (Scheme 2), representing genuine bond-shift isomers [23,24]. The experimental demonstration of this theoretically predicted behavior was achieved in our laboratory through the spectroscopic identification of these two isomeric forms under matrix-isolation conditions [18]. This finding prompted us to investigate how substituents on the phenyl ring affect the bond-shift isomerism of the nitrilimine fragment. Our subsequent study on C-hydroxyphenyl-nitrilimines showed that the OH phenyl substitution leads to different outcomes depending on its position [17]. When the OH group is in the meta position, the nitrilimine has structural features similar to the parent compound, displaying two bond-shift isomers with propargylic and allenic characters. In contrast, when the OH group is in the para position, only a single isomer with a strong allenic character was observed.
Inspired by these results, the present study investigates both experimentally and theoretically the effects of the NH₂ substitution at the meta and para positions on the bond-shift isomeric properties of C-phenyl-nitrilimine. These investigations involved the use of tunable narrowband UV light to generate C-aminophenyl-nitrilimines, from matrix-isolated 5-aminophenyl-tetrazoles, and the exploration of their photochemistry. Infrared spectroscopy, supported by DFT theoretical calculations, was employed to identify and characterize the photoproduced species. In addition, to gain deeper insights into the effects of the phenyl substitution on the bond-shift isomerism of the nitrilimine fragment, we expanded the theoretical studies to compare the influence of both electron-donating (NH2) and electron-withdrawing (NO2) groups in ortho, meta, and para positions relative to the unsubstituted C-phenyl-nitrilimine. This comprehensive approach enabled us to investigate both positional effects and the inherent characteristics of substituents.

2. Results and Discussion

2.1. Structure and Matrix-Isolated IR Spectra of 5-Aminophenyl-Tetrazoles

In this study, 5-(4-aminophenyl)-tetrazole will be denoted as 1p, and 5-(3-aminophenyl)-tetrazole as 1m, in accordance with the nomenclature previously used by us for the analogous hydroxy-substituted 5-phenyl-tetrazoles (letters p and m signify the para and meta positions of the substituent group in the phenyl ring relative to the tetrazole ring, respectively) [17]. Both molecules can exist in the 1H or 2H tautomeric forms (see Table 1), abbreviated here as 1p′/1m′ or 1p″/1m″, respectively, which is also consistent with the terminology followed in our previous work [17]. However, it has been established that direct access to nitrilimines is only viable through the 2H tautomer [16,17,18,19,20,25]. Therefore, to theoretically evaluate the potential of compounds 1p and 1m as candidates for the photoproduction of the corresponding C-(4-aminophenyl)-nitrilimine 2p and C-(3-aminophenyl)-nitrilimine 2m, the former were energetically characterized at the B3LYP/6-311++G(d,p) level of theory and the obtained results are summarized in Table 1. Based on the Boltzmann populations estimated from the Gibbs energies at 380 K (temperature of the sublimated compounds before the matrix deposition, see Section 3), the 1p″/1p′ and 1m″/1m′ tautomeric ratios were determined to be 92/8 and 96/4, respectively, thus demonstrating the suitability of 1p and 1m as precursors of nitrilimines. The higher stability of the 1p″ or 1m″ isomers (2H tautomeric forms) compared to the 1p′ or 1m′ isomers (1H tautomeric forms) can be attributed to the increased electron delocalization within the tetrazole ring in the formers [26], which is reflected by the longer formal N=N and C=N double bonds and shorter formal N-N single bond (see caption of Table 1).
Molecule 1p″ does not display conformers. However, molecule 1m″ exhibits two stable conformations distinguished by syn (s) and anti (a) orientations of the tetrazole NH group relative to the phenyl NH2 group (see Table 1 for details). According to the calculations, 1m″-a is the most stable conformer. The Gibbs energy of 1m″-s relative to 1m″-a is less than 1 kJ mol−1, resulting in a computed 1m″-a/1m″-s ratio of 57/39 at 380 K (Table 1). Consequently, a mixture of 1m″-a and 1m″-s conformers should dominate the vapor phase prior to the Ar matrix deposition. Besides the gas-phase populations, evaluation of the energy barrier for the conversion of the higher (1m″-s) into the lower energy (1m″-a) conformer is also relevant for predicting the composition of the deposited Ar matrix [27,28,29,30]. The energy barrier for the conformational relaxation, obtained from the B3LYP potential energy profile (PES) for the internal rotation around the inter-rings C−C bond, was calculated to be 20 kJ mol−1 (Figure S1). This value is sufficiently large to allow efficient trapping of both conformers in an Ar matrix at 15 K [27,28,29,30].
Figure 1a and Figure 2a depict the experimental IR spectra of 1p and 1m, respectively, isolated in an Ar matrix at 15 K, accompanied by the theoretical IR spectra of 1p″ and 1m″, shown in Figure 1b and Figure 2b. For 1m″, the theoretical spectrum was derived by co-adding the B3LYP spectra calculated for the two predicted conformers (Figure 2c) with the intensities multiplied by their estimated gas-phase Boltzmann populations. Overall, there is a good correspondence between the experimental spectra of matrix-isolated 1p and 1m and the theoretical spectra computed for 1p″ and 1m″, respectively. On the contrary, the experimental spectra are not well reproduced by the spectra computed for 1p′ or 1m′ (Figures S2 and S3). This is consistent with the theoretical predictions and demonstrates that there is an overwhelming predominance of the 2H tautomeric forms in the as-deposited Ar matrices.
Tables S1 and S2 present the proposed IR band assignments for 1p and 1m, respectively, which were based on the comparison of the experimental spectra with the computed vibrational data. The higher frequency region of the two experimental mid-IR spectra is characterized by a very strong band at 3469−3467 cm−1, which is characteristic of the tetrazole ν(NH) mode [26], together with two weaker bands at 3513 and 3421 cm−1 for 1p, or at 3507 and 3417 cm−1 for 1m, corresponding to the νas(NH2) and νs(NH2) modes, respectively. In the fingerprint region, the spectrum of 1p is dominated by the bands at 1625 [δ(NH2)], 1498/1483 [δ(CNN); δ(NCN); δ(NH); ν(CN)tz], 1449 [ν(CC)ph-tz; δ(NCN)], and 1294 [ν(CCNH2)] cm−1, while the spectrum of 1m is dominated by the bands at 1621 [δ(NH2)], 1498 [δ(CNN); ν(CN)tz; δ(NH)], 763 [γ(C)tz], and 714 [γ(N)tz] cm−1. Even though the spectra calculated for the two 1m″ conformers are very similar (as shown in Figure 2c), a detailed spectral analysis allows the identification of five experimental bands specific to each of them. Spectral features of 1m″-a are observed at 1498, 1445, 1437, 1315, and 1253 cm−1 (highlighted by blue circles in Figure 2a), while spectral signatures of 1m″-s are found at 1476, 1469, 1428, 1277, and 1272 cm−1 (indicated by red squares in Figure 2a).

2.2. Photogeneration and Spectroscopic Characterization of C-(4-Aminophenyl)-Nitrilimine

Phototransformations in 1p isolated in solid Ar at 15 K were triggered by performing in situ irradiations with narrowband light at λ = 230 nm. This wavelength was chosen because of its demonstrated effectiveness in inducing photoreactions in analogous matrix-isolated molecules [17,18,19] and due to its proximity to the wavelength of maximum absorption in the UV spectrum of 5-phenyl-tetrazole dissolved in ethanol (λmax = 229.4 nm) [31]. The spectral changes originated by the UV irradiations are depicted in the difference spectrum presented in Figure 3a for the 2300–1100 cm−1 range and in Figure S4a for the 3600–3200 cm−1 range. The negative bands indicate the consumption of 1p, while the positive ones denote the appearance of photoproducts. Focusing on the bands emerging in the 2300–1700 cm−1 region, the doublet at 2098/2059 cm−1 appears shortly after 10 s of irradiation but ceases to grow for longer irradiation times. In contrast, the other two bands, observed at 2150 cm−1 and 1781 cm−1, continue to grow (see Figure S5 for details). These findings suggest that the first band should correspond to a primary photoproduct, whereas the remaining two should be associated with secondary photoproducts.
Considering the typical photochemical behavior reported for 2H-tetrazoles [11,32,33], it is highly plausible that the primary photoproduct would correspond to C-(4-aminophenyl)-nitrilimine 2p, which is produced by extrusion of molecular nitrogen from 5-(4-aminophenyl)-tetrazole 1p. To confirm this hypothesis, we have conducted subsequent irradiations at a longer wavelength (λ = 380 nm) with the aim of promoting the consumption of putative 2p without affecting 1p. The identification of 2p was definitively confirmed by the excellent agreement of the negative part of the difference IR spectrum obtained by subtracting the spectrum recorded before the irradiation at λ = 380 nm from that recorded after the irradiation at this wavelength, with the B3LYP/6-311++G(d,p) theoretical spectrum of this species, as depicted in Figure 3b,c (and additionally in Figure S4). A detailed assignment of the IR spectrum of 2p is provided in Table 2. Its most characteristic spectral feature is the doublet band at 2098/2059 cm−1, which is theoretically predicted at 2078 cm−1 and is ascribed to the antisymmetric stretching vibration of the CNN fragment [νas(CNN)]. The placement of this band within the range of 2100–2000 cm−1 strongly suggests that the photogenerated nitrilimine adopts an allenic-type structure [3,16]. Further elaboration on this topic will be provided later.
Concomitantly with the decrease in the bands assigned to 2p, other bands that emerged upon the excitations at λ = 230 nm, including those at 2150 cm−1 and 1781 cm−1, are intensified during the irradiations at λ = 380 nm (see Figure 3b). This observation demonstrates that the corresponding photoproducts are generated with the intermediacy of 2p. Building upon existing knowledge of the photochemistry of nitrilimines, these photoproducts are most likely 4-aminophenyl-1H-diazirine 3p and 4-aminophenyl-carbodiimide 4p [3,17,18,19,34,35]. Both species could be easily identified through the outcomes of additional irradiations carried out at λ = 330 nm (after the consumption of 2p). The comparison of the difference spectrum resulting from these irradiations (Figure 4a) with the calculated difference spectrum 4p3p (Figure 4b) demonstrates clearly that the negative bands are assigned to 3p while the positive ones are assigned to 4p, i.e., 3p photorearranges to 4p [16,20]. The main absorption of 3p is that identified at 1781 cm−1 (predicted at 1808 cm−1), which is particularly characteristic of 1H-azirines [17,18,20,36] and is assigned to the three-membered ring ν(C=N) vibration. Regarding 4p, its most prominent band is identified at 2150 cm−1 (predicted at 2115 cm−1), which is particularly characteristic of carbodiimides and is assigned to the νas(NCN) vibration [16,17,18]. In summary, our results reveal that once 2p is generated from 1p by N2 elimination, it photoisomerizes to yield 4p via 3p (see Scheme 3). This reasoning agrees with the mechanism of reactivity that has been established in previous photochemical studies carried out for similar systems [17,18].

2.3. Photogeneration and Spectroscopic Characterization of C-(3-Aminophenyl)-Nitrilimine

Exposing matrix-isolated 1m to narrowband UV light tuned at λ = 230 nm resulted in the appearance of new bands at 2067 (triplet), 1787, and 2142/2106 cm−1 within the 2300–1700 cm−1 region (see Figure 5a and Figure S5). These spectral positions closely match those observed during the photolysis of 1p, indicating the formation of analogous photoproducts: C-(3-aminophenyl)-nitrilimine 2m, 3-aminophenyl-1H-diazirine 3m, and 4-aminophenyl-carbodiimine 4m. The only significative difference from the photochemical behavior of 1p is the emergence of an additional band at 2239 cm−1, which displays an identical growth pattern to the band at ~2067 cm−1 (Figure S5), suggesting that it is also assigned to a primary photoproduct. This IR band falls within a spectral region where propargylic nitrilimines typically absorb [3,16,17,18]. Hence, it is most likely that both allenic and propargylic forms of 2m, designated as 2mA and 2mP, respectively, are generated from 1m, similar to the photochemistry observed for the 5-phenyl-tetrazole [18] and 5-(3-hydroxyphenyl)-tetrazole [17]. This interpretation is also supported by the computed PES as a function of the CNN bond angle, which reveals the existence of two separate energy minima for 2m but only a single minimum for 2p (Figure 6).
Experimental evidence confirming the co-existence of the two bond-shift isomers was obtained through subsequent irradiations at λ = 330 nm. In the first 20 min of irradiation, several bands that had appeared after the initial irradiation of the sample at λ = 230 nm decreased in intensity (Figure 5b). After an additional 10 min of irradiation, some of these bands were completely depleted, while others continued to diminish (Figure 5c). This allowed the distinction of two species based on their different kinetics of consumption. From the comparison of the negative parts of the two obtained experimental difference IR spectra (Figure 5b,c) with the calculated IR spectra of 2mA and 2mP (Figure 5d), the following can be concluded: (i) the bands centered at 2238 and 1221 cm−1, corresponding to the species that experienced a more rapid consumption, match the νas(CNN) and δ(NH) modes predicted for 2mP at 2236 and 1220 cm−1, respectively; (ii) the bands at 2067 and 1246/1236 cm−1, ascribed to the species that experienced a slower consumption, are well reproduced by the νas(CNN) and δ(NH) modes theoretically predicted for 2mA at 2091 and 1240 cm−1. These results clearly confirm the in situ generation of both propargylic- and allenic-type isomers of 2m upon UV irradiation of matrix-isolated 1m″ (Scheme 4). An assignment of the IR spectra of 2mA and 2mP is given in Table 3. The faster kinetics of consumption of 2mP (k1)compared to 2mA (k2) under UV irradiation at λ = 330 nm is consistent with the significantly higher absorptivity of the former relative to the latter, as deduced from the comparison of the TD-DFT UV spectra for the two isomers (Figure S6). The set of IR bands that increase upon the irradiation at λ = 330 nm, due to the consumption of 2mP and 2mA, are attributed to 3m and 4m, with the most characteristic spectral features being identified at 1787 cm−1 for 3m [ν(C=N)] and at 2067 cm−1 for 4mas(NCN)].

2.4. Stability of the Bond-Shift Isomers and Their Attempted Interconversion

As shown in Figure 6, the energy barrier separating 2mP and 2mA is very small (less than 1 kJ mol−1). Additionally, the energy difference between the two isomers is also very small (Table S3). According to these data, it can be hypothesized that the co-existence of these two isomers in the low-temperature Ar matrix is due to the fact that they are in equilibrium. However, our experimental observations do not support this hypothesis. In fact, after subjecting the sample to 20 min of UV irradiation at λ = 330 nm, resulting in the selective depletion of 2mP, we monitored the sample for several hours (13 h) in dark conditions (a metal plate was placed between the source of the spectrometer and the sample) to assess whether this isomer could be regenerated at the expense of 2mA. However, no such regeneration was observed, definitively showing that the two isomers exist independently in the cryogenic matrix.
Having established that 2mA and 2mP are not in equilibrium in the Ar matrix, we attempted their interconversion by vibrational excitation [36,37,38] with narrowband NIR light produced by a laser-OPO source or with broadband NIR/mid-IR light emitted from the IR spectrometer source. To determine the specific frequencies of vibrational modes available for performing selective NIR vibrational excitation, two spectra were recorded in the 7500–4000 cm−1 region: one immediately after the deposition of 1m in the Ar matrix and another after a series of UV irradiations at λ = 230 nm, followed by irradiations at λ = 330 nm to deplete 2mP while retaining 2mA in the matrix. The resulting NIR difference spectrum was compared with B3LYP/6-311++G(d,p) anharmonic vibrational calculations for the precursor and photoproducts (see Figure 7 and Table S6).
This spectral comparison allowed us to unequivocally identify the band corresponding to the first overtone of the ν(NH) stretching vibration of 2mA at 6307 cm−1. We also tried to identify the 2νas(NH2) and 2νs(NH2) overtones, as well as the νas(NH2) + νs(NH2) combination band corresponding to this species. However, as shown in Figure 7, these bands overlap with those of the precursor and other photoproducts, making it difficult to pinpoint the exact frequencies suitable for laser irradiations. Irradiating the matrix for 20 min with the laser tuned to 6307 cm−1 did not produce any spectral changes indicating the 2mA2mP conversion. The exposure of the sample to the broadband NIR/mid-IR radiation emitted by the spectrometer source for approximately 50 h also did not result in any detectable interconversion. Although unexpected, considering the theoretical data, these results are in line with previous findings for C-phenyl-nitrilimine [18] and C-(3-hydroxyphenyl)-nitrilimine [17]. For both molecules, two bond-shift isomers (allenic and propargylic) were also found to co-exist in an Ar matrix after photolysis of the corresponding tetrazole precursors, and in both cases, their interconversion was not achieved either by annealing experiments (for C-phenyl-nitrilimine) or by selective NIR excitation of the OH stretching overtones (for C-(3-hydroxyphenyl)-nitrilimine). Two factors may help to explain this observed behavior in the bond-shift isomers of the three phenyl-nitrilimines. Firstly, intrinsic limitations of theoretical methods that might hinder accurate calculation of the PES for the interconversion between the two bond-shift isomers. Secondly, the matrix-environment may lead to an increase in the energy barrier separating the two isomers, making their interconversion more difficult, as it has been reported in some other cryogenic matrix-isolation experiments involving systems with conformational isomers separated by very small energy barriers [24,39,40,41,42,43].

2.5. Effect of the Position and Nature of the Phenyl Substitution on the Structure of C-Phenyl-Nitrilimines

According to the experimental results, the structure of photogenerated C-aminophenyl-nitrilimines varies with the positioning of the NH2 group relative to the CNN fragment. For the para-NH2 substitution (2p), only one isomeric form is identified. Conversely, for the meta-NH2 substitution (2m), two distinct isomeric forms are observed: 2mA and 2mP. This is consistent with the PES calculated for the two nitrilimines at different levels, as depicted in Figure 6. Specifically, a single minimum is predicted for 2p, while two minima are predicted for 2m. Based on the wavenumbers corresponding to the νas(CNN) band—2098/2057 cm−1 for 2p, 2067 cm−1 for 2mA, and 2239 cm−1 for 2mP—it can be inferred that 2p and 2mA are predominantly allenic, while 2mP is predominantly propargylic. This is corroborated by the geometric data extracted from the B3LYP/cc-pVTZ fully optimized geometries, as detailed in Table 4. Both 2p and 2mA have a non-planar structure (CRCNNHH dihedral = 98–99º) and a bent CCN framework (CRCNNC angle = 142–143º). Conversely, 2mP has a more planar structure (CRCNNHH dihedral = 161º), together with a much straighter CCN framework (CRCNNC angle = 178º). Additionally, the CN bond is shorter and the NN bond is longer in 2mP compared to 2mA and 2p. The experimental and theoretical results obtained for 2p and 2m align with the results of previous studies on analogous OH-substituted phenyl nitrilimines [17]. From these findings, a general conclusion can be drawn: electron-donor substituents located at the para position relative to the CNN fragment promote an allenic-type structure, whereas meta substitution stabilizes both allenic and propargylic structures.
To deepen our understanding of how phenyl substitution affects the structure of the CNN fragment, we extended our theoretical study to include a derivative with an electron-withdrawing substituent (NO2) and also considered the ortho position for both NH2 and NO2 substitutions. The PES computed for these newly explored molecules are depicted in Figure S7, and the geometric data extracted from the B3LYP/cc-pVTZ fully optimized geometries of the respective minima are provided in Table 4. The results for the meta-NO2 substitution closely resemble those obtained for the meta-NH2 substitution, where both the allenic and propargylic structures are potentially stable minima, though the corresponding PES is even flatter. Moreover, the corresponding geometric parameters exhibit similar trends comparing the meta-NO2 with the meta-NH2 substitution. Regarding the ortho and para substitutions, there is a notable difference between the two substituents. For NH2, both substitutions stabilize allenic-type isomers exhibiting identical geometric parameters, whereas for NO2, they stabilize propargylic-type isomers that also share similar geometric features (Figure S7 and Table 4).
To better comprehend the influence of the phenyl substitution on the structure of the nitrilimine fragment, we have calculated the charge distributions in the NH2- and NO2-substituted C-phenyl-nitrilimines, as well as in the unsubstituted molecule (used as reference), and the obtained results are included in Table 4. The Hirshfeld population analysis [44] was used for this purpose because of its proven accuracy in estimating atomic charges of organic molecules [45] and its ability to provide a comprehensive description of both reactivity and regioselectivity in electrophilic aromatic substitutions [46,47,48]. For C-phenyl-nitrilimine, the interplay of resonance and inductive effects originated by the phenyl ring leads to the stabilization of both allenic and propargylic forms [18]. The allenic form is characterized by a negative charge at the nitrilimine carbon atom (CN) and a positive charge at the adjacent nitrogen atom (NC), whereas in the propargylic form, NC carries a positive charge while the terminal nitrogen (NH) bears a negative charge (see Scheme 1). The meta substitution reveals that regardless of the substituent’s nature, the charge distribution within the nitrile imine fragment is not much affected with respect to that of the analogous unsubstituted isomers. This explains why both allenic and propargylic forms remain minima on the respective PESs. Rather different results are obtained for the ortho and para substitutions, depending on the substituent nature. For the NH2 derivatives, the ring carbon atom attached to the CNN group (CR) becomes significantly more negatively charged compared to the unsubstituted molecule. This is consistent with the electron-donating nature of this substituent, which increases the electron density at the ring carbons in the ortho and para positions. According to the results presented in Table 4, the increased electron charge in CR is also partly delocalized to CN, which also becomes negatively charged. Such an increase in the negative charge at CN favors the allenic-type structure and consequently disfavors the propargylic one. Regarding the NO2 derivatives, CR becomes significantly more positively charged as compared to the unsubstituted molecule. This is also consistent with the electron-withdrawing nature of this group, which decreases the electron charge of the ring carbon in ortho or para positions with respect to the substituent. This electron withdrawal is also extended to the CN atom, which results in the destabilization of the allenic form and consequent stabilization of the propargylic one.

3. Methods

3.1. Experimental Methods

Commercial samples of 5-(4-aminophenyl)-tetrazole (1p, Alfa Aesar, 97% purity) and 5-(3-aminophenyl)-tetrazole (1m, Alfa Aesar, 97% purity) were sublimated at ~380 K inside a miniature glass oven connected to the vacuum chamber of a closed-cycle helium cryostat (APD Cryogenics, with a DE-202A expander), through a stainless steel needle valve. Before cooling the cryostat, a freeze–pump–thaw procedure was performed to purify the sample from volatile impurities. The matrices were prepared by co-deposition of vapors resulting from the sublimation of the compounds with an excess of argon (Air Liquid, N60, purity 99.999%) onto a CsI optical substrate, cooled down to 15 K. The temperature of the CsI window was measured directly at the sample holder using a silicon diode sensor connected to a digital controller (Scientific Instruments, West Palm Beach, FL, USA, model 9650-1). Spectra in the mid-IR region (4000–400 cm−1) were recorded with a resolution of 0.5 cm−1 employing a Thermo Nicolet 6700 FTIR spectrometer, equipped with a cadmium telluride (MCT-B) detector and a potassium bromide (KBr) beam splitter. Spectra in the NIR region (7500–4000 cm−1) were also recorded with a resolution of 1 cm−1, using the same spectrometer and an Indium Gallium Arsenide (InGaAs) detector with a CaF2 beam splitter. The spectrometer underwent continuous purging with a stream of dry, CO2-filtered air to eliminate atmospheric water and CO2 vapors. The prepared matrixes were then irradiated through the outer quartz window of the cryostat by applying narrowband UV radiation (fwhm = 0.2 cm−1) provided by the frequency-doubled signal beam of a Quanta-Ray MOPO-SL optical parametric oscillator (OPO), pumped with a pulsed Nd:YAG laser pro 230 from Spectra-Physics (repetition rate = 10 Hz, pulse energy ~1–3 mJ, duration = 10 ns). The UV irradiations were performed at λ = 230, 308, 330, and 380 nm, with different exposure times.

3.2. Computational Methods

All calculations were carried out using the Gaussian 16 (Revision B.01) program package [49]. The geometries of the experimentally studied molecules were fully optimized at the B3LYP/6-311++G(d,p) [50,51,52] level of theory, followed by harmonic vibrational calculations at the same level. The B3LYP/6-311++G(d,p) vibrational data were then used to assist the spectral interpretation. To account for the incomplete treatment of electron correlation, basis set limitations, and neglected anharmonic effects, the harmonic vibrational frequencies were multiplied by 0.980 for the region below 2000 cm−1 and 0.950 for the region above 2000 cm−1 [17]. Theoretical IR spectra were obtained from the vibrational data by convoluting the peaks with Lorentzian functions (fwhm = 2 cm−1) centered at the scaled wavenumbers and peak heights matching the calculated IR intensities. An approximate description of the vibrational modes was based on the results provided by the “vibAnalysis” software (version 1.2.2) [53], supported by ChemCraft animation of the vibrations [54]. Anharmonic vibrational calculations were also carried out at B3LYP/6-311++G(d,p) level using the fully automated second-order vibrational perturbative approach developed by Barone and co-workers [55,56]. For some of the studied molecules, theoretical absorption UV-spectra was simulated from time-dependent density functional theory calculations (TD-DFT) [57,58] performed at the B3LYP/6-311++G(d,p) level. The Cartesian coordinates extracted from full geometry optimizations carried out for all experimentally studied molecules are provided as Supplementary Materials.
Additional calculations were performed for C-phenyl-nitrilimine, as well as for the NH2- and NO2-substituted C-phenyl-nitrilimines in ortho, meta, and para positions. This included relaxed potential energy scans as a function of the nitrilimine CNN angle at the B3LYP/6-311++G(d,p), PBE0 [59]/6-311++G(d,p), and PW6B95 [60]/6-311++G(d,p) levels (only for the substituted C-phenyl-nitrilimines) and full geometry optimizations of the identified minima (plus vibrational calculations) at the B3LYP/cc-pVTZ level. This basis set was chosen for consistency with calculations previously performed on analogous molecules [17]. The outcomes of the optimizations at this level served to obtain geometric and charge distribution data for nitrilimines.

4. Conclusions

The effect of substitution in the phenyl ring on the structure of C-phenyl-nitrilimines has been explored in this study by considering the para- and meta-NH2-substituted compounds (C-(4-aminophenyl)- and C-(3-aminophenyl)-nitrilimines, respectively), which were experimentally studied by matrix-isolation spectroscopy and quantum chemical calculations. The comprehensive results obtained for these compounds, which bear a π electron-donating substituent, were complemented by computational studies performed on the ortho-NH2 derivative, as well as on the ortho-, meta-, and para-NO2-substituted analogous C-phenyl-nitrilimines, which bear a π electron-withdrawing substituent. Our previous data for the parent compound and the OH-substituted C-phenyl-nitrilimines [16,17,18] were also considered for the elucidation of the effects of the nature of the substituents on the structure of the C-phenyl-nitrilimines, in particular the existence of different bond-shift isomers.
The amino-substituted C-phenyl-nitrilimines were photogenerated in situ, in an argon matrix at 15 K, by excitation at λ = 230 nm of their tetrazole precursors (5-(4-aminophenyl)- and 5-(3-aminophenyl)-tetrazole, respectively). The UV-induced formation of the nitrilimines was confirmed by additional photochemistry experiments where these species were subsequently isomerized to carbodiimides via 1H-diazirines by irradiations at longer wavelengths (λ = 380–308 nm).
The combined results of the experiments and computations allowed us to conclude that, while the para-NH2 substitution stabilizes a single isomeric structure of the predominant allenic type, the meta-NH2 substitution leads to the co-existence of both propargylic- and allenic-type bond-shift isomers, a trend that we have also observed before for the analogous OH-substituted C-phenyl-nitrilimines [17].
The calculations performed on the whole set of molecules mentioned above, in particular the obtained equilibrium geometries, potential energy surfaces landscapes, and Hirshfeld charges, permitted attaining a comprehensive understanding of the effects of phenyl substitution on the bond-shift isomerism in C-phenyl nitrilimines. Parent C-phenyl-nitrilimine exists in both allenic and propargylic forms [18]. Regardless of the substituent’s nature, meta substitution (either electron donating or withdrawing) does not affect substantially the electronic features within the nitrilimine fragment, so meta substitution preserves the co-existence of the allenic and propargylic forms observed for the parent compound. On the other hand, ortho and para substitutions considerably change the electronic structure of the nitrilimine moiety. For electron donating substitutions (as in the case of the OH and NH2 derivatives), an increase in the negative charge at CN takes place, which favors the allenic-type isomer that becomes the single stable isomeric form, while in the case of electron-withdrawing substitutions (as for the NO2 derivatives), the CN atom becomes more positive, destabilizing the allenic form and leading to a consequent stabilization of the propargylic one, which then became the unique stable bond-shift isomeric form.
As a final remark, this study underscores the significant influence of substituent position and nature on the electronic structure and isomerism of C-phenyl-nitrilimines. The findings provide valuable insights for future research and potential applications involving substituted nitrilimines, particularly in designing compounds with specific electronic and structural properties.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29153497/s1, Figure S1: B3LYP potential energy scan calculated for the 2H tautomeric form of 5-(3-hydroxyphenyl)-tetrazole 1m″; Figures S2–S3: selected regions of experimental (Ar, 15 K) and theoretical IR spectra of 5-(4-aminoyphenyl)-tetrazole 1p and 5-(3-aminoyphenyl)-tetrazole 1m. Figure S4: experimental difference IR spectra in the 3600–3200 cm−1 region showing spectral changes after the irradiations of matrix-isolated 1p at λ = 230 nm followed by irradiations at λ = 380 nm and their comparison with the theoretical IR spectrum of 2p; Figure S5: fragments of the experimental IR spectra recorded after different times of exposition of matrix-isolated 1p and 1m (Ar, 15 K) to UV light at λ = 230 nm; Figure S6: TD-DFT UV spectra computed for the allenic and propargylic bond-shift isomers of 2m; Figure S7: relaxed potential energy scans as function of the CCN angle calculated for different ring-substituted C-phenyl-nitrilimines; Tables S1–S2: bands observed in the IR spectra of 1p and 1m (Ar, 15 K) and their comparison with the vibrational data extracted from B3LYP calculations carried out for tautomers 1p″ and 1m″, including an approximate assignment of the vibration modes; Table S3: structures and B3LYP relative energies calculated for the isomers of 2p and 2m; Table S4: structures and B3LYP relative energies calculated for the conformers of 4-aminophenyl-1H-diazirine 3p, 4-aminophenyl-carbodiimide 4p, 3-aminophenyl-1H-diazirine 3m and 3-aminophenyl-carbodiimide 4m; Table S5: structures and B3LYP relative energies calculated for the conformers of the ortho-NH2, ortho-NO2, and meta-NO2-substituted C-phenyl-nitrilimines; Table S6, Bands observed in the near-IR spectra recorded after isolation of 1m in an argon matrix at 15 K and after a series of UV irradiations at 230 and 330 nm, and their approximate assignment. Computational data: Cartesian coordinates and energies an extracted from the full geometry optimizations performed at the B3LYP/6-311++G(d,p) level of theory.

Author Contributions

A.J.L.J., conceptualization, investigation, computations, writing—original draft preparation, and writing—review and editing; C.M.N., laboratory work, conceptualization, and writing—review and editing; G.A.F., laboratory work; K.K., formal analysis; R.F., writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been conducted under the project PTDC/QUI-QFI/1880/2020, funded by the Portuguese Science Foundation (FCT). The Coimbra Chemistry Centre (CQC-IMS) is financially supported by FCT, projects UI0313B/00313/2020 and UI0313B/00313/2020 (national funds; DOI: 10.54499/UIDB/00313/2020 and 10.54499/UIDP/00313/2020) and LA/P/0056/2020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and the Supplementary Materials.

Acknowledgments

The authors acknowledge the Laboratory for Advanced Computing at University of Coimbra (UC-LCA) for providing computing resources that have contributed to the research results reported within this paper and Coimbra Laser Lab (CLL) for experimental facilities. Igor Reva is acknowledged for the participation in preliminary experiments and for helpful discussions.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Huisgen, R.; Seidel, M.; Sauer, J.; McFarland, J.; Wallbillich, G. Communications: The Formation of Nitrile Imines in the Thermal Breakdown of 2,5-Disubstituted Tetrazoles. J. Org. Chem. 1959, 24, 892–893. [Google Scholar] [CrossRef]
  2. Padwa, A. 1,3-Dipolar Cycloaddition Chemistry; Wiley: New York, NY, USA, 1984. [Google Scholar]
  3. Bertrand, G.; Wentrup, C. Nitrile Imines: From Matrix Characterization to Stable Compounds. Angew. Chem. Int. Ed. Engl. 1994, 33, 527–545. [Google Scholar] [CrossRef]
  4. Huisgen, R. 1,3-Dipolar Cycloadditions. Past and Future. Angew. Chem. Int. Ed. Engl. 1963, 2, 565–598. [Google Scholar] [CrossRef]
  5. Huisgen, R. Kinetics and Mechanism of 1,3-Dipolar Cycloadditions. Angew. Chem. Int. Ed. Engl. 1963, 2, 633–645. [Google Scholar] [CrossRef]
  6. Jamieson, C.; Livingstone, K. The Nitrile Imine 1,3-Dipole: Properties, Reactivity and Applications; Springer: Cham, Swirzerland, 2020. [Google Scholar]
  7. Molteni, G.; Ponti, A. The Nitrilimine–Alkene Cycloaddition Regioselectivity Rationalized by Density Functional Theory Reactivity Indices. Molecules 2017, 22, 202. [Google Scholar] [CrossRef] [PubMed]
  8. Stille, J.K.; Harris, F.W. Polymers from 1,3-dipole addition reactions: The nitrilimine dipole from acid hydrazide chlorides. J. Polym. Sci. Part A Polym. Chem. 1968, 6, 2317–2330. [Google Scholar] [CrossRef]
  9. Shawali, A.S. Reactions of heterocyclic compounds with nitrilimines and their precursors. Chem. Rev. 1993, 93, 2731–2777. [Google Scholar] [CrossRef]
  10. An, P.; Lewandowski, T.M.; Erbay, T.G.; Liu, P.; Lin, Q. Sterically Shielded, Stabilized Nitrile Imine for Rapid Bioorthogonal Protein Labeling in Live Cells. J. Am. Chem. Soc. 2018, 140, 4860–4868. [Google Scholar] [CrossRef]
  11. Frija, L.M.T.; Cristiano, M.L.S.; Gómez-Zavaglia, A.; Reva, I.; Fausto, R. Genesis of rare molecules using light-induced reactions of matrix-isolated tetrazoles. J. Photochem. Photobiol., C 2014, 18, 71–90. [Google Scholar] [CrossRef]
  12. Lucero, P.L.; Peláez, W.J.; Riedl, Z.; Hajós, G.; Moyano, E.L.; Yranzo, G.I. Flash vacuum pyrolysis of azolylacroleins and azolylbutadienes. Tetrahedron 2012, 68, 1299–1305. [Google Scholar] [CrossRef]
  13. Pagacz-Kostrzewa, M.; Mucha, M.; Weselski, M.; Wierzejewska, M. Conformational properties and photochemistry of new allyl tetrazoles: Matrix isolation FTIR and computational approach. J. Photochem. Photobiol. A 2013, 251, 118–127. [Google Scholar] [CrossRef]
  14. Wentrup, C. Flash Vacuum Pyrolysis of Azides, Triazoles, and Tetrazoles. Chem. Rev. 2017, 117, 4562–4623. [Google Scholar] [CrossRef] [PubMed]
  15. Baskir, E.G.; Platonov, D.N.; Tomilov, Y.V.; Nefedov, O.M. Infrared-spectroscopic study of amino-substituted nitrilimines and their photochemical transformations in an argon matrix. Mendeleev Commun. 2014, 24, 197–200. [Google Scholar] [CrossRef]
  16. Bégué, D.; Qiao, G.G.; Wentrup, C. Nitrile Imines: Matrix Isolation, IR Spectra, Structures, and Rearrangement to Carbodiimides. J. Am. Chem. Soc. 2012, 134, 5339–5350. [Google Scholar] [CrossRef] [PubMed]
  17. Ferreira, G.A.; Nunes, C.M.; Lopes Jesus, A.J.; Fausto, R. The meta and para OH Substitution Effect on C-Phenyl-Nitrilimine Bond-Shift Isomers. Eur. J. Org. Chem. 2023, 26, e202300310. [Google Scholar] [CrossRef]
  18. Nunes, C.M.; Reva, I.; Fausto, R.; Bégué, D.; Wentrup, C. Bond-shift isomers: The co-existence of allenic and propargylic phenylnitrile imines. Chem. Commun. 2015, 51, 14712–14715. [Google Scholar] [CrossRef] [PubMed]
  19. Nunes, C.M.; Reva, I.; Rosado, M.T.S.; Fausto, R. The Quest for Carbenic Nitrile Imines: Experimental and Computational Characterization of C-Amino Nitrile Imine. Eur. J. Org. Chem. 2015, 2015, 7484–7493. [Google Scholar] [CrossRef]
  20. Nunes, C.M.; Araujo-Andrade, C.; Fausto, R.; Reva, I. Generation and Characterization of a 4π-Electron Three-Membered Ring 1H-Diazirine: An Elusive Intermediate in Nitrile Imine–Carbodiimide Isomerization. J. Org. Chem. 2014, 79, 3641–3646. [Google Scholar] [CrossRef]
  21. Bégué, D.; Wentrup, C. Carbenic Nitrile Imines: Properties and Reactivity. J. Org. Chem. 2014, 79, 1418–1426. [Google Scholar] [CrossRef]
  22. Muchall, H.M. Computational Insight into the Carbenic Character of Nitrilimines from a Reactivity Perspective. J. Phys. Chem. A 2011, 115, 13694–13705. [Google Scholar] [CrossRef]
  23. Herges, R. Organizing Principle of Complex Reactions and Theory of Coarctate Transition States. Angew. Chem. Int. Ed. Engl. 1994, 33, 255–276. [Google Scholar] [CrossRef]
  24. Maltsev, A.; Bally, T.; Tsao, M.-L.; Platz, M.S.; Kuhn, A.; Vosswinkel, M.; Wentrup, C. The Rearrangements of Naphthylnitrenes:  UV/Vis and IR Spectra of Azirines, Cyclic Ketenimines, and Cyclic Nitrile Ylides. J. Am. Chem. Soc. 2004, 126, 237–249. [Google Scholar] [CrossRef] [PubMed]
  25. Zhao, X.; Liu, Q.; Feng, R.; Zeng, X.; Wentrup, C. Photolysis and Pyrolysis of Phenyltetrazoles: Formation of Phenylcarbodiimide, N-Phenylnitrile Imine, Phenylnitrene, Indazole, and Fulvenallene. Eur. J. Org. Chem. 2019, 2019, 6945–6950. [Google Scholar] [CrossRef]
  26. Bugalho, S.C.S.; Maçôas, E.M.S.; Cristiano, M.L.S.; Fausto, R. Low temperature matrix-isolation and solid state vibrational spectra of tetrazole. Phys. Chem. Chem. Phys. 2001, 3, 3541–3547. [Google Scholar] [CrossRef]
  27. Lopes Jesus, A.J.; Rosado, M.T.S.; Reva, I.; Fausto, R.; Eusébio, M.E.; Redinha, J.S. Conformational Study of Monomeric 2,3-Butanediols by Matrix-Isolation Infrared Spectroscopy and DFT Calculations. J. Phys. Chem. A 2006, 110, 4169–4179. [Google Scholar] [CrossRef] [PubMed]
  28. Rosado, M.T.S.; Lopes Jesus, A.J.; Reva, I.D.; Fausto, R.; Redinha, J.S. Conformational Cooling Dynamics in Matrix-Isolated 1,3-Butanediol. J. Phys. Chem. A 2009, 113, 7499–7507. [Google Scholar] [CrossRef]
  29. Barnes, A.J. Matrix isolation vibrational spectroscopy as a tool for studying conformational isomerism. J. Mol. Struct. 1984, 113, 161–174. [Google Scholar] [CrossRef]
  30. Jesus, A.J.L.; Rosado, M.T.S.; Reva, I.; Fausto, R.; Eusébio, M.E.S.; Redinha, J.S. Structure of Isolated 1,4-Butanediol: Combination of MP2 Calculations, NBO Analysis, and Matrix-Isolation Infrared Spectroscopy. J. Phys. Chem. A 2008, 112, 4669–4678. [Google Scholar] [CrossRef] [PubMed]
  31. Elpern, B.; Nachod, F.C. Absorption Spectra and Structure of Some Tetrazoles. J. Am. Chem. Soc. 1950, 72, 3379–3382. [Google Scholar] [CrossRef]
  32. Maier, G.; Eckwert, J.; Bothur, A.; Reisenauer, H.P.; Schmidt, C. Photochemical Fragmentation of Unsubstituted Tetrazole, 1,2,3-Triazole, and 1,2,4-Triazole: First Matrix-Spectroscopic Identification of Nitrilimine HCNNH. Liebigs Annalen 1996, 1996, 1041–1053. [Google Scholar] [CrossRef]
  33. Bégué, D.; Dargelos, A.; Wentrup, C. Aryl Nitrile Imines and Diazo Compounds. Formation of Indazole, Pyridine N-Imine, and 2-Pyridyldiazomethane from Tetrazoles. J. Org. Chem. 2020, 85, 7952–7958. [Google Scholar] [CrossRef] [PubMed]
  34. Bégué, D.; Santos-Silva, H.; Dargelos, A.; Wentrup, C. Imidoylnitrenes R′C(═NR)–N, Nitrile Imines, 1H-Diazirines, and Carbodiimides: Interconversions and Rearrangements, Structures, and Energies at DFT and CASPT2 Levels of Theory. J. Phys. Chem. A 2017, 121, 8227–8235. [Google Scholar] [CrossRef] [PubMed]
  35. Ismael, A.; Fausto, R.; Cristiano, M.L.S. Photochemistry of 1- and 2-Methyl-5-aminotetrazoles: Structural Effects on Reaction Pathways. J. Org. Chem. 2016, 81, 11656–11663. [Google Scholar] [CrossRef] [PubMed]
  36. Lopes Jesus, A.J.; Nunes, C.M.; Fausto, R.; Reva, I. Conformational control over an aldehyde fragment by selective vibrational excitation of interchangeable remote antennas. Chem. Commun. 2018, 54, 4778–4781. [Google Scholar] [CrossRef] [PubMed]
  37. Nunes, C.M.; Pereira, N.A.M.; Viegas, L.P.; Pinho e Melo, T.M.V.D.; Fausto, R. Inducing molecular reactions by selective vibrational excitation of a remote antenna with near-infrared light. Chem. Commun. 2021, 57, 9570–9573. [Google Scholar] [CrossRef] [PubMed]
  38. Lopes Jesus, A.J.; Nunes, C.M.; Reva, I.; Pinto, S.M.V.; Fausto, R. Effects of Entangled IR Radiation and Tunneling on the Conformational Interconversion of 2-Cyanophenol. J. Phys. Chem. A 2019, 123, 4396–4405. [Google Scholar] [CrossRef]
  39. Yoshida, K.-i.; Iiba, E.; Nozaki, Y.; Hirai, K.; Takahashi, Y.; Tomioka, H.; Lin, C.-T.; Gaspar, P.P. Di(9-anthryl)carbene Revisited. Product Analysis and Spectroscopic Studies. Bull. Chem. Soc. Jpn. 2004, 77, 1509–1522. [Google Scholar] [CrossRef]
  40. DePinto, J.T.; McMahon, R.J. Structure and rearrangements of 1,3-diphenylpropynylidene. J. Am. Chem. Soc. 1993, 115, 12573–12574. [Google Scholar] [CrossRef]
  41. Carra, C.; Nussbaum, R.; Bally, T. Experimental and Theoretical Study of 2,6-Difluorophenylnitrene, Its Radical Cation, and Their Rearrangement Products in Argon Matrices. ChemPhysChem 2006, 7, 1268–1275. [Google Scholar] [CrossRef]
  42. Reva, I.D.; Stepanian, S.G.; Adamowicz, L.; Fausto, R. Missing conformers. Comparative study of conformational cooling in cyanoacetic acid and methyl cyanoacetate isolated in low temperature inert gas matrixes. Chem. Phys. Lett. 2003, 374, 631–638. [Google Scholar] [CrossRef]
  43. Sponsler, M.B.; Jain, R.; Coms, F.D.; Dougherty, D.A. Matrix-isolation decay kinetics of triplet cyclobutanediyls. Observation of both Arrhenius behavior and heavy-atom tunneling in carbon-carbon bond-forming reactions. J. Am. Chem. Soc. 1989, 111, 2240–2252. [Google Scholar] [CrossRef]
  44. Hirshfeld, F.L. Bonded-atom fragments for describing molecular charge densities. Theor. Chim. Acc. 1977, 44, 129. [Google Scholar] [CrossRef]
  45. Wiberg, K.B.; Rablen, P.R. Atomic Charges. J. Org. Chem. 2018, 83, 15463–15469. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, S. Where does the electron go? The nature of ortho/para and meta group directing in electrophilic aromatic substitution. J. Chem. Phys. 2014, 141, 194109. [Google Scholar] [CrossRef] [PubMed]
  47. Liu, S. Quantifying Reactivity for Electrophilic Aromatic Substitution Reactions with Hirshfeld Charge. J. Phys. Chem. A 2015, 119, 3107–3111. [Google Scholar] [CrossRef] [PubMed]
  48. Nikolova, V.; Cheshmedzhieva, D.; Ilieva, S.; Galabov, B. Atomic Charges in Describing Properties of Aromatic Molecules. J. Org. Chem. 2019, 84, 1908–1915. [Google Scholar] [CrossRef]
  49. Frisch, M.E.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. GAUSSIAN 16; Revision B.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  50. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  51. Lee, C.T.; Yang, W.T.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron-density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  52. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  53. Teixeira, F.; Cordeiro, M.N.D.S. Improving Vibrational Mode Interpretation Using Bayesian Regression. J. Chem. Theory Comput. 2019, 15, 456–470. [Google Scholar] [CrossRef]
  54. Zhurko, G.A.C. Chemcraft, Version 1.8. 2016. Available online: http://www.chemcraftprog.com (accessed on June 2024).
  55. Barone, V. Anharmonic vibrational properties by a fully automated second-order perturbative approach. J. Chem. Phys. 2004, 122, 14108. [Google Scholar] [CrossRef] [PubMed]
  56. Bloino, J.; Barone, V. A second-order perturbation theory route to vibrational averages and transition properties of molecules: General formulation and application to infrared and vibrational circular dichroism spectroscopies. J. Chem. Phys. 2012, 136, 124108. [Google Scholar] [CrossRef]
  57. Bauernschmitt, R.; Ahlrichs, R. Treatment of electronic excitations within the adiabatic approximation of time dependent density functional theory. Chem. Phys. Lett. 1996, 256, 454–464. [Google Scholar] [CrossRef]
  58. Stratmann, R.E.; Scuseria, G.E.; Frisch, M.J. An efficient implementation of time-dependent density-functional theory for the calculation of excitation energies of large molecules. J. Chem. Phys. 1998, 109, 8218–8224. [Google Scholar] [CrossRef]
  59. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  60. Zhao, Y.; Truhlar, D.G. Design of Density Functionals That Are Broadly Accurate for Thermochemistry, Thermochemical Kinetics, and Nonbonded Interactions. J. Phys. Chem. A 2005, 109, 5656–5667. [Google Scholar] [CrossRef]
Scheme 1. Idealized geometric structures of nitrilimines.
Scheme 1. Idealized geometric structures of nitrilimines.
Molecules 29 03497 sch001
Scheme 2. Bond-shift isomers of C-phenyl-nitrilimine.
Scheme 2. Bond-shift isomers of C-phenyl-nitrilimine.
Molecules 29 03497 sch002
Figure 1. (a) Experimental IR spectrum of 5-(4-aminophenyl)-tetrazole 1p in an Ar matrix at 15 K; (b) B3LYP/6-311++G(d,p) theoretical IR spectrum of 1p″.
Figure 1. (a) Experimental IR spectrum of 5-(4-aminophenyl)-tetrazole 1p in an Ar matrix at 15 K; (b) B3LYP/6-311++G(d,p) theoretical IR spectrum of 1p″.
Molecules 29 03497 g001
Figure 2. (a) Experimental IR spectrum of 5-(3-aminoyphenyl)-tetrazole 1m in an Ar matrix at 15 K; (b) theoretical IR spectrum of 1m″ obtained as a sum of the B3LYP/6-311++G(d,p) spectra calculated for the 1m″-a and 1m″-s conformers with their intensities scaled by the respective gas-phase Boltzmann populations; (c) wavenumbers and IR intensities (both scaled) extracted from the B3LYP/6-311++G(d,p) harmonic vibrational calculations carried out for the two conformers.
Figure 2. (a) Experimental IR spectrum of 5-(3-aminoyphenyl)-tetrazole 1m in an Ar matrix at 15 K; (b) theoretical IR spectrum of 1m″ obtained as a sum of the B3LYP/6-311++G(d,p) spectra calculated for the 1m″-a and 1m″-s conformers with their intensities scaled by the respective gas-phase Boltzmann populations; (c) wavenumbers and IR intensities (both scaled) extracted from the B3LYP/6-311++G(d,p) harmonic vibrational calculations carried out for the two conformers.
Molecules 29 03497 g002
Figure 3. Experimental difference IR spectra (2300−1100 cm−1) showing changes (a) after irradiation of matrix-isolated 1p (Ar, 15 K) at λ = 230 nm for ~1 min and (b) after subsequent ~26 min of irradiation at λ = 380 nm (the negative bands labeled by red circles are due to the consumed photoproduct assigned to 2p); (c) theoretical IR spectrum of the most stable conformer of C-(4-aminophenyl)-nitrilimine 2p (2p-a) calculated at the B3LYP/6-311++G(d,p) level (see Table S3 for details).
Figure 3. Experimental difference IR spectra (2300−1100 cm−1) showing changes (a) after irradiation of matrix-isolated 1p (Ar, 15 K) at λ = 230 nm for ~1 min and (b) after subsequent ~26 min of irradiation at λ = 380 nm (the negative bands labeled by red circles are due to the consumed photoproduct assigned to 2p); (c) theoretical IR spectrum of the most stable conformer of C-(4-aminophenyl)-nitrilimine 2p (2p-a) calculated at the B3LYP/6-311++G(d,p) level (see Table S3 for details).
Molecules 29 03497 g003
Figure 4. (a) Experimental difference IR spectrum showing changes after 15 min of irradiation at λ = 330 nm subsequently to the consumption of 2p at 380 nm (Figure 3). The negative bands are due to the consumed 4-aminophenyl-1H-diazirine 3p while the positive ones are assigned to the photogenerated 4-aminophenyl-carbodiimide 4p. (b) B3LYP/6-311++G(d,p) theoretical difference spectrum obtained as the spectrum of the most stable conformer of 4p minus the spectrum of the most stable conformer of 3p (see Table S4 for details).
Figure 4. (a) Experimental difference IR spectrum showing changes after 15 min of irradiation at λ = 330 nm subsequently to the consumption of 2p at 380 nm (Figure 3). The negative bands are due to the consumed 4-aminophenyl-1H-diazirine 3p while the positive ones are assigned to the photogenerated 4-aminophenyl-carbodiimide 4p. (b) B3LYP/6-311++G(d,p) theoretical difference spectrum obtained as the spectrum of the most stable conformer of 4p minus the spectrum of the most stable conformer of 3p (see Table S4 for details).
Molecules 29 03497 g004
Scheme 3. Summary of the UV-induced photochemistry of the 2H-tautomeric form of 5-(4-aminophenyl)-tetrazole 1p″ (Ar, 15 K): generation of C-(4-aminophenyl)-nitrilimine 2p and its subsequent photoisomerization to 4-aminophenyl-1H-diazirine 3p and 4-aminophenyl-carbodiimide 4p. Black, red and blue arrows represent the phototransformations at 230, 380 and 330 nm, respectively.
Scheme 3. Summary of the UV-induced photochemistry of the 2H-tautomeric form of 5-(4-aminophenyl)-tetrazole 1p″ (Ar, 15 K): generation of C-(4-aminophenyl)-nitrilimine 2p and its subsequent photoisomerization to 4-aminophenyl-1H-diazirine 3p and 4-aminophenyl-carbodiimide 4p. Black, red and blue arrows represent the phototransformations at 230, 380 and 330 nm, respectively.
Molecules 29 03497 sch003
Figure 5. Experimental difference IR spectra (2300−1100 cm−1) showing changes (a) after irradiation of matrix-isolated 1m (Ar, 15 K) at λ = 230 nm and (b,c) after subsequent irradiations at λ = 330 nm (the negative bands labeled by green circles and violet squares are due to the consumed photoproducts assigned to 2mA and 2mP, respectively); (d) theoretical IR spectra of most stable conformers of 2mA (green) and 2mP (violet) calculated at the B3LYP/6-311++G(d,p) level (see Table S3 for details).
Figure 5. Experimental difference IR spectra (2300−1100 cm−1) showing changes (a) after irradiation of matrix-isolated 1m (Ar, 15 K) at λ = 230 nm and (b,c) after subsequent irradiations at λ = 330 nm (the negative bands labeled by green circles and violet squares are due to the consumed photoproducts assigned to 2mA and 2mP, respectively); (d) theoretical IR spectra of most stable conformers of 2mA (green) and 2mP (violet) calculated at the B3LYP/6-311++G(d,p) level (see Table S3 for details).
Molecules 29 03497 g005
Figure 6. Relaxed potential energy scan as function of the CCN angle for (a) C-(4-aminophenyl)-nitrilimine 2p and (b) C-(3-aminophenyl)-nitrilimine 2m. The scans were computed using the B3LYP (blue), PW6B95 (green), and PBE0 (orange) DFT functionals with the 6-311++G(d,p) basis set. Along the scans, the value of the CCN angle was incrementally fixed in steps of 2.5 degrees, and all other geometrical parameters were fully optimized.
Figure 6. Relaxed potential energy scan as function of the CCN angle for (a) C-(4-aminophenyl)-nitrilimine 2p and (b) C-(3-aminophenyl)-nitrilimine 2m. The scans were computed using the B3LYP (blue), PW6B95 (green), and PBE0 (orange) DFT functionals with the 6-311++G(d,p) basis set. Along the scans, the value of the CCN angle was incrementally fixed in steps of 2.5 degrees, and all other geometrical parameters were fully optimized.
Molecules 29 03497 g006
Scheme 4. Summary of the UV-induced photochemistry of the 2H tautomeric form of 5-(3-aminophenyl)-tetrazole 1m″ (Ar, 15 K): generation of the propargylic (2mP) and allenic (2mA) bond-shift isomers of C-(3-aminophenyl)-nitrilimine 2m and their subsequent photoisomerization to 3-aminophenyl-1H-diazirine 3m and 3-aminophenyl-carbodiimide 4m with distinct kinetics. Black and blue arrows represent the phototransformations at 230 and 330 nm, respectively.
Scheme 4. Summary of the UV-induced photochemistry of the 2H tautomeric form of 5-(3-aminophenyl)-tetrazole 1m″ (Ar, 15 K): generation of the propargylic (2mP) and allenic (2mA) bond-shift isomers of C-(3-aminophenyl)-nitrilimine 2m and their subsequent photoisomerization to 3-aminophenyl-1H-diazirine 3m and 3-aminophenyl-carbodiimide 4m with distinct kinetics. Black and blue arrows represent the phototransformations at 230 and 330 nm, respectively.
Molecules 29 03497 sch004
Figure 7. (a) Selected region of a difference near-IR spectrum (7200–6050 cm−1) showing changes after 20 min UV irradiation of 1m″ at λ = 230 nm, followed by subsequent 20 min of irradiation at λ = 330 nm. Positive bands are assigned to 2mA, 3m, and 4m species (2mP was not identified as it was consumed during the first 20 min of irradiation at λ = 330 nm, Figure 5), while the negative ones are assigned to 1m″-a and 1m″-s. (b) B3LYP/6-311++G(d,p) anharmonic wavenumbers and IR intensities (both unscaled) calculated for the first overtones of the NH stretching vibration [2ν(NH)], first overtones of the antisymmetric [2νas(NH2)] and symmetric [2νs(NH2)] stretching vibrations, and νas(NH2) + νs(NH2) combination modes (see Table S6 for details).
Figure 7. (a) Selected region of a difference near-IR spectrum (7200–6050 cm−1) showing changes after 20 min UV irradiation of 1m″ at λ = 230 nm, followed by subsequent 20 min of irradiation at λ = 330 nm. Positive bands are assigned to 2mA, 3m, and 4m species (2mP was not identified as it was consumed during the first 20 min of irradiation at λ = 330 nm, Figure 5), while the negative ones are assigned to 1m″-a and 1m″-s. (b) B3LYP/6-311++G(d,p) anharmonic wavenumbers and IR intensities (both unscaled) calculated for the first overtones of the NH stretching vibration [2ν(NH)], first overtones of the antisymmetric [2νas(NH2)] and symmetric [2νs(NH2)] stretching vibrations, and νas(NH2) + νs(NH2) combination modes (see Table S6 for details).
Molecules 29 03497 g007
Table 1. Structures, relative Gibbs energies at 380 K (ΔG, in kJ mol−1) computed at the B3LYP/6-311++G(d,p) level of theory and the respective Boltzmann populations at 380 K (Pop., in %) for the tautomers and conformers of 5-(4-aminophenyl)-tetrazole 1p and 5-(3-aminophenyl)-tetrazole 1m a.
Table 1. Structures, relative Gibbs energies at 380 K (ΔG, in kJ mol−1) computed at the B3LYP/6-311++G(d,p) level of theory and the respective Boltzmann populations at 380 K (Pop., in %) for the tautomers and conformers of 5-(4-aminophenyl)-tetrazole 1p and 5-(3-aminophenyl)-tetrazole 1m a.
5-(4-aminophenyl)-tetrazole 1p
IsomerMolecules 29 03497 i001Molecules 29 03497 i002
1p″1p′
ΔG (380 K)0.006.13
Pop. 92.27.8
5-(3-aminophenyl)-tetrazole 1m
IsomerMolecules 29 03497 i003Molecules 29 03497 i004Molecules 29 03497 i005Molecules 29 03497 i006
1m″-a1m″-s1m′-a1m′-s
ΔG (380 K)0.000.976.9110.79
Pop.57.138.63.50.7
a In the isomers designations, 1p′/1m′ or 1p″/1m″ correspond to the 1H or 2H tautomeric forms, respectively. In the case of 1m, the letter after the hyphen denotes the relative orientation of the NH and NH2 groups (a = anti; s = syn). Mean bond lengths in the tetrazole ring extracted from the full geometry optimizations carried out for 1p″, 1m″-a, and 1m″-s (2H tautomeric forms): C=N, 1.333 ± 0.001 Å; N=N, 1.303 ± 0.001 Å; N-N, 1.326 ± 0.001 Å. Mean bond lengths in the tetrazole ring extracted from the full geometry optimizations carried out for the 1p′, 1m′-a and 1m′-s (1H tautomeric forms): C=N, 1.324 ± 0.001 Å; N=N, 1.288 ± 0.001 Å; N-N, 1.354 ± 0.002 Å.
Table 2. Experimental (Ar, 15 K) and B3LYP/6-311++G(d,p) computed infrared absorptions for C-(4-aminophenyl)-nitrilimine 2p and approximate vibrational assignments.
Table 2. Experimental (Ar, 15 K) and B3LYP/6-311++G(d,p) computed infrared absorptions for C-(4-aminophenyl)-nitrilimine 2p and approximate vibrational assignments.
Ar Matrix aComputed bVibrational Assignment c
ν̃Ath
35233502.624.3νas(NH2)
3432/34283404.557.1νs(NH2)
n.o.3329.81.2ν(NH)
2098/20592077.9333.3νas(CNN)
16211631.0241.1δ(NH2)
16031610.188.5ν(CC)ph
15761564.49.8ν(CC)ph
15221515.034.7ν(CNA) − ν(CCN); ν(CC)ph; δ(CH)
13671380.4117.5ν(NN) − ν(CCN)
13001292.583.5ν(CNA); ν(CC)ph
12901287.512.4ν(CC)ph
12591258.9423.2δ(NH)
11791178.584.9δ(CH)
11241127.014.8δ(CH)
10981098.528.9ν(NN) + ν(CCN)
n.o.1046.97.4ρ(NH2)
831829.815.5ν(CC)ph
831823.445.0γ(CH)ph; γ(C)ph
a Experimental wavenumbers are given in cm−1. Only the bands in the 3600–3200 cm−1 and 2300–800 cm−1 regions are included. Abbreviations: n.o. = not observed. b Data extracted from the vibrational calculations carried out for the B3LYP/6-311++G(d,p) lowest energy conformer of 2p (2p-a); see Table S3 for details. Calculated harmonic wavenumbers were multiplied by 0.950 (above 2000 cm−1) or 0.980 (below 2000 cm−1). Vibrations predicted with very low intensities and lacking counterparts in the experimental spectrum were omitted. c Vibrational assignment was based on the results provided by the “vibAnalysis” software, supported by ChemCraft animation of the vibrations. Abbreviations: ν, stretching; δ, in-plane bending; γ, out-of-plane bending; ρ, rocking; ph, phenyl ring; s, symmetric; as, antisymmetric; CN refers to the carbon atom of the CNN moiety and NA to the nitrogen atom of the amino group. Signs “+” and “−” designate combination of vibrations occurring in the same phase and in the opposite phase, respectively.
Table 3. Experimental (Ar, 15 K) and B3LYP/6-311++G(d,p) computed infrared absorptions for C-(3-aminophenyl)-nitrilimine 2m and approximate vibrational assignments.
Table 3. Experimental (Ar, 15 K) and B3LYP/6-311++G(d,p) computed infrared absorptions for C-(3-aminophenyl)-nitrilimine 2m and approximate vibrational assignments.
Ar Matrix aComputed bVibrational Assignment c
ν̃Ath
34873489.2/3490.819.5/19.4νas(NH2)
n.o.3395.3/3396.629.2/29.8νs(NH2)
n.o. 3240.7/3277.15.1/31.5ν(NH)
22382236.1967.0νas(CNN), 2mP
20672091.1481.6νas(CNN), 2mA
16181630.5/1629.9171.3/157.1δ(NH2)
15991605.4/1602.088.6/148.9ν(CC)ph
15861579.4/1576.936.0/50.5ν(CC)ph
14901491.4/1492.42.2/18.9ν(CC)ph; δ(CH)
n.o.1454.1/1447.738.8/27.5ν(CC)ph
13661375.4/1382.677.5/63.8ν(NN) − ν(CCN)
n.o.1321.9/1323.910.8/4.5δ(CH); ν(CC)ph
13061302.0/1297.210.4/9.2ν(CC)ph; ν(CNA)
12791270.4/1273.6172.4/85.5ν(CNA); δ(NH); δ(CH)
1246/12371239.9234.3δ(NH), 2mA
12211220.4229.7δ(NH), 2mP
a Experimental wavenumbers are given in cm−1. Only the bands in the 3600–3200 cm−1 and 2300–1200 cm−1 regions are included. Abbreviations: n.o. = not observed. b Data extracted from the vibrational calculations carried out for the B3LYP/6-311++G(d,p) lowest energy conformers of 2mA (2mA-a) and 2mP (2mP-a); see Table S3 for details. Values before and after the slash refer to 2mA and 2mP, respectively. Calculated harmonic wavenumbers were multiplied by 0.950 (above 2000 cm−1) or 0.980 (below 2000 cm−1). Vibrations predicted with very low intensities and lacking counterparts in the experimental spectrum were omitted. Vibrational assignment was based on the results provided by the “vibAnalysis” software, supported by ChemCraft animation of the vibrations. c Abbreviations: ν, stretching; δ, in-plane bending; ph, phenyl ring; s, symmetric; as, antisymmetric; CN refers to the carbon atom of the CNN moiety and NA to the nitrogen atom of the amino group. Sign “−” designates a combination of vibrations occurring in the opposite phase.
Table 4. Geometric parameters and Hirshfeld atomic charges computed at the B3LYP/cc-pVTZ level for C-phenyl-nitrilimine and its ortho-, meta-, and para-NH2 and NO2 phenyl-substituted derivatives a.
Table 4. Geometric parameters and Hirshfeld atomic charges computed at the B3LYP/cc-pVTZ level for C-phenyl-nitrilimine and its ortho-, meta-, and para-NH2 and NO2 phenyl-substituted derivatives a.
Allenic-Type Structures
C-Phenyl-Nitrilimineortho-NH2meta-NH2(2mA)meta-NO2para-NH2(2p)
Molecules 29 03497 i007Molecules 29 03497 i008Molecules 29 03497 i009Molecules 29 03497 i010Molecules 29 03497 i011
Geometric parameters
CRCN (Å)1.441.441.441.431.43
CNNC (Å)1.181.191.181.181.19
NCNH (Å)1.251.251.251.241.25
CRCNNC (º)142.13142.33142.45144.93143.04
CNNCNH (º)169.82169.40169.87170.40169.39
CRCNNHH (º)−99.47−98.22−99.46−104.16−97.61
Hirshfeld charges
CR0.002−0.0250.0010.013−0.019
CN−0.019−0.023−0.018−0.004−0.025
NC0.0490.0440.0480.0580.039
NH−0.211−0.212−0.215−0.198−0.224
Propargylic-Type Structures
C-Phenyl-Nitrilimineortho-NO2meta-NH2(2mP)meta-NO2para-NO2
Molecules 29 03497 i012Molecules 29 03497 i013Molecules 29 03497 i014Molecules 29 03497 i015Molecules 29 03497 i016
Geometric parameters
CRCN (Å)1.411.401.411.411.41
CNNC (Å)1.171.171.171.171.17
NCNH (Å)1.261.251.261.251.25
CRCNNC (º)179.85170.92178.41179.80179.92
CNNCNH (º)171.76174.49171.76171.70171.69
CRCNNHH (º)180.00180.00−161.23180.00180.00
Hirshfeld charges
CR0.0070.0160.0060.0150.020
CN0.0430.0460.0440.0440.050
NC0.0460.0600.0440.0550.057
NH−0.239−0.202−0.244−0.219−0.209
a The data refer to the CNN fragment and were extracted from the optimized geometries of the lowest energy conformers of each molecule (see Tables S3 and S5). Atom labeling is given in the allenic-type structure of C-phenyl-nitrilimine.
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

Lopes Jesus, A.J.; Nunes, C.M.; Ferreira, G.A.; Keyvan, K.; Fausto, R. Photochemical Generation and Characterization of C-Aminophenyl-Nitrilimines: Insights on Their Bond-Shift Isomers by Matrix-Isolation IR Spectroscopy and Density Functional Theory Calculations. Molecules 2024, 29, 3497. https://doi.org/10.3390/molecules29153497

AMA Style

Lopes Jesus AJ, Nunes CM, Ferreira GA, Keyvan K, Fausto R. Photochemical Generation and Characterization of C-Aminophenyl-Nitrilimines: Insights on Their Bond-Shift Isomers by Matrix-Isolation IR Spectroscopy and Density Functional Theory Calculations. Molecules. 2024; 29(15):3497. https://doi.org/10.3390/molecules29153497

Chicago/Turabian Style

Lopes Jesus, A. J., Cláudio M. Nunes, Gil A. Ferreira, Kiarash Keyvan, and R. Fausto. 2024. "Photochemical Generation and Characterization of C-Aminophenyl-Nitrilimines: Insights on Their Bond-Shift Isomers by Matrix-Isolation IR Spectroscopy and Density Functional Theory Calculations" Molecules 29, no. 15: 3497. https://doi.org/10.3390/molecules29153497

Article Metrics

Back to TopTop