Yeray A. Rodríguez-Núñez*a,
Jesús Sánchez-Márquez
b,
Jorge Quintero-Saumeth
f,
Cristian J. Guerra
c,
Efraín Polo-Cuadrado
d,
David Villaman
e,
Cristopher A. Fica-Cornejo
a and
Arnold R. Romero Bohórquez
*f
aUniversidad Andrés Bello, Facultad de Ciencias Exactas, Departamento de Ciencias Químicas, Laboratorio de Síntesis y Reactividad de Compuestos Orgánicos, Santiago 8370146, Chile. E-mail: yeray.rodriguez@unab.cl; c.ficacornejo@uandresbello.edu
bDepartamento de Química-Física, Facultad de Ciencias, Campus Universitario Río San Pedro, Universidad de Cádiz, Cádiz, Spain
cDepartamento de Ciencias Químicas, Facultad de ciencias Exactas, Universidad Andrés Bello, Republica 275, Santiago, Chile. E-mail: c.guerramadera@uandresbello.edu
dDepartamento de Química Orgánica, Facultad de Ciencias Químicas, Universidad de Concepción, Concepción, Chile. E-mail: epolo@udec.cl
eLaboratorio de Química Inorgánica y Organometálica, Departamento de Química Analítica e Inorgánica, Facultad de Ciencias Químicas, Universidad de Concepción, Edmundo Larenas 129, Casilla 160-C, Concepción 4070386, Chile. E-mail: dvillaman@udec.cl
fGrupo de Investigación en Compuestos Orgánicos de Interés Medicinal CODEIM, Parque Tecnológico Guatiguará, Universidad Industrial de Santander, Piedecuesta 681011, Colombia. E-mail: arafrom@uis.edu.co
First published on 15th April 2025
In this study, the Povarov cationic reaction mechanism was explored using five different Lewis acids as catalysts for the synthesis of N-propargyl-6-methoxy-4-(2′-oxopyrrolidin-1′-yl)-1,2,3,4-tetrahydroquinoline, where the best reaction yield was obtained using InCl3. The desired product was not obtained in the absence of a catalyst. A comprehensive theoretical analysis at the density functional theory (DFT) level was conducted to study the role of the catalyst and establish a detailed reaction mechanism. Electron localization function (ELF) analyses were performed to elucidate the key bonding events during the reaction stages, highlighting the differences in bond formation among the different catalysts. Our results showed that the presence of an acid catalyst is required for obtaining the intermediary iminium ion. In this sense, the InCl3 catalyst provides the lowest energy barrier for catalytic interactions, increasing the electrophilic character and, therefore the reactivity of formaldehyde, promoting the formation of iminium ions and subsequently triggering the obtaining of the tetrahydroquinoline compound. In fact, from theoretical analysis, our findings provide evidence of the formation of the tetrahydroquinoline compound through a set of energetically favorable step reactions, ruling out a concerted process. The step involved in this part of the mechanism includes the formation of a Mannich-type adduct, obtained by the nucleophilic addition reaction between the iminium cation and an activated alkene, and a subsequent cyclization via an intramolecular Friedel–Crafts reaction. This defines the cationic Povarov reaction as a domino reaction and invites us to discard the wrong use of the name Aza Diels–Alder or imino Diels–Alder for this type of reaction.
Considering the pharmacological applications of tetrahydroquinolines and the growing interest in accessing new and promising derivatives based on these compounds, several synthetic strategies based on classical reactions have been developed. These advances have enabled rapid and efficient access to the tetrahydroquinoline structure, even with high structural functionalization.7 Some examples of these reactions involve the catalytic reduction of the quinoline core at high hydrogen pressures8 and intramolecular reactions leading to ring closure, such as aza-Michael–Michael addition,9 Friedel–Crafts reaction,10 nucleophilic substitutions (SN), and sigmatropic rearrangements.11 Nevertheless, owing to its high efficiency and facile performance, the most powerful synthetic tool for the construction of the tetrahydroquinoline skeleton is the acid-catalyzed three-component Povarov reaction. This chemical reaction was first characterized as occurring between primary arylamines and aldehydes (or ketones) to obtain an intermediate arylimine that reacts with electron-rich olefins in the presence of an acid catalyst via a formal [4 + 2] cycloaddition reaction.12
The Povarov reaction mechanism has been the subject of extensive scientific research. Indeed, for decades, the authors have defined Povarov synthesis as an inverse electron-demand aza-Diels–Alder reaction, with the subsequent transfer of protons that leads to the formation of the tetrahydroquinoline core.13 In this context, the first theoretical study of this mechanism using DFT involves the formation of tetrahydro-1,5-naphthyridine derivatives under acid catalysis, suggesting that the cycloaddition reaction occurs as an asynchronous concerted process.14 Subsequently, Domingo et al. reported two mechanistic studies on the Povarov reaction based on DFT calculations. Initially, it was suggested that the formation of the tetrahydroquinoline ring could occur through a process that involves the following reaction stages: (i) a Lewis acid-catalyzed aza-Diels–Alder (A-DA) reaction between an N-arylimine and nucleophilic ethylene, yielding a formal [4 + 2] cycloadduct and (ii) a stepwise 1,3-hydrogen shift at this intermediate. Nonetheless, the mechanistic implications of the stepwise stage are attributed solely to the acid catalyst being performed by the Lewis acid.15 While using a Brønsted acid as a catalyst, the authors pointed out that the Povarov reaction mechanism corresponds to a domino process that is initialized by the formation of a cationic intermediate, which undergoes a quick intramolecular Friedel–Crafts reaction16,17
In contrast, other experimental studies based on the stereochemistry of THQ obtained from N-arylimines under acidic conditions indicated that the mechanism of the Povarov reaction is a stepwise process. From this perspective, the Povarov reaction involves an ionic intermediate before imine activation and nucleophilic attack from an electron-rich alkene.18 Other reports have suggested that the THQ core obtained from these precursors occurs through a domino reaction beginning with the formation of a Mannich adduct, followed by an intramolecular Friedel–Crafts-type reaction.19
In addition, N-substituted THQs were synthesized by reacting secondary arylamines with formaldehyde to obtain a highly reactive intermediate iminium ion. Thus, this reaction reacts with nucleophilic olefins in a one-pot process. Initially, this variant was called the cationic imino Diels–Alder reaction.20 Similarly, the cationic Povarov reaction resembles the oxa-Povarov reaction for chromone synthesis reported by Taylor and Batey,21 which involves the reaction between a highly reactive aryl-2-oxazine cationic intermediate generated in situ using a Lewis catalyst (SnCl4) and styrene as the electron-rich olefin. Interestingly, Domingo et al. studied the oxa-Povarov reaction mechanism using DFT methods and concluded that it occurs along a stepwise mechanism; initially, there is a nucleophilic attack of styrene on the oxonium cationic system (Prins reaction) and a subsequent intramolecular Friedel–Crafts reaction.16
The use of N-arylanilines as precursors to obtain the desired THQ in high yields via N-aryl-N-alkylmethyleneiminium ions has been reported.22 In previous studies, we reported a simple one-pot methodology for the synthesis of new derivatives of 4-aryl-3-methyl-1,2,3,4-THQ using aqueous HCl as a catalyst between in situ-generated cationic 2-azadienes and arylpropenes (isoeugenol and trans-anethole).23 Hence, this work provides the basis for the most recent theoretical study on the mechanism of the cationic Povarov reaction using DFT methods. Wherein, it is assumed that the formation of the products proceeds by a stepwise mechanism. However, the role of acid catalysts in cationic intermediate formation and the subsequent stepwise mechanism has not been well established. Similarly, Lewis acids (InCl3, BF3, and BiCl3) have been used in the cationic Povarov reaction between the N-allyl/propargyl-N-arylymethyleneiminium cation and N-vinylpyrrolidone to efficiently generate N-allyl/propargyl THQs.22
In this work, the three components of the cationic Povarov reaction were carried out using five different Lewis acids as catalysts for the synthesis of N-propargyl-6-methoxy-4-(2′-oxopyrrolidin-1′-yl)-1,2,3,4-tetrahydroquinoline 4. When Lewis acids were employed, the corresponding tetrahydroquinoline core was achieved in moderate and high yields. Herein, we performed a comprehensive theoretical analysis at the DFT level to study the role of the catalyst and establish the detailed mechanism of this reaction. Finally, we conducted a real-space analysis of the electron localization function (ELF) to unravel the key chemical bonding events of both the intermediate iminium ion formation and cyclization stages.
Brown solid. m.p. 154–156 °C; IR (ATR): 3264, 2956, 2934, 2874, 2808, 1672, 1498, 1060, 803 cm−1; 1H NMR (400 MHz, CDCl3): 1.92–2.04 (2H, m), 2.05–2.16 (2H, m), 2.14 (1H, t, J = 2.4 Hz), 2.46–2.50 (2H, m), 3.09–3.3 (2H, m), 3.18–3.35 (2H, m), 3.71 (3H, s, 6-OCH3), 3.91 (1H, dd, J = 18.1, 2.4 Hz), 4.04 (1H, dd, J = 18.1, 2.4 Hz), 5.40 (1H, dd, J = 8.8, 6.4 Hz, 4-H), 6.51 (1H, d, J = 2.8 Hz), 6.72 (1H, d, J = 8.9 Hz), 6.77 (1H, ddd, J = 8.9, 2.8, 0.4 Hz); 13C NMR (100 MHz, CDCl3): 18.4, 26.9, 31.5, 41.4, 43.7, 47.8, 48.0, 55.7, 72.3, 79.2, 113.6, 114.0, 114.6, 122.9, 139.7, 152.5, 175.5; MS (ESI-IT), m/z: 307.1[M + Na]+, 591.1[2M + Na]+; Anal. Calcd for C17H20N2O2: C, 71.81; H, 7.09; N, 11.25%. Found: C, 71.72; H, 7.01; N, 11.06%. The NMR and ESI-MS data matched previously reported data.22
ΔG0 = EDFT0 + GDFTcorr + ΔGDFTSMD + ΔG1atm→1M |
To evaluate the effect of the different Lewis acids as catalysts, the reaction was performed in the absence and presence of the different Lewis catalysts, including InCl3, BiCl3, AlCl3, TiCl4 and BF3·OEt2 (20 mol% stoichiometric ratio). As expected, no conversion was observed under the evaluated reaction conditions when no catalyst was used, and the precursor substrates remained unchanged in the reaction medium. The use of any of the Lewis catalysts evaluated led to reactions taking place in high and moderate yields. In the case of InCl3, the highest yield was observed (92%), thus corroborating once again the remarkable catalytic effect of indium(III) salts in the conventional or cationic Povarov reaction (Scheme 1).
![]() | ||
Scheme 1 Synthesis of N-propargyl 1,2,3,4-tetrahydroquinoline 4 via a one-pot three-component cationic Povarov reaction. ORTEP plot of THQ 4. Thermal ellipsoids were drawn at a probability of 30%. |
Crystal data | |||
Empirical formula | C17H20N2O2 | μ mm−1 | 0.081 |
Formula weight | 284.35 | F(000) | 608.0 |
Temperature/K | 296.15 | Crystal size/mm3 | 0.226 × 0.182 × 0.052 |
Crystal system | Monoclinic | Radiation | MoKα (λ = 0.71073) |
Space group | Cc | 2Θ range for data collection/° | 4.68 to 54.986 |
a [Å] | 9.4298(3) | Index ranges | −12 ≤ h ≤ 11, −22 ≤ k ≤ 22, −12 ≤ l ≤ 12 |
b [Å] | 17.4070(7) | Reflections collected | 17![]() |
c [Å] | 9.4334(4) | Independent reflections | 3314 [Rint = 0.0455, Rsigma = 0.0329] |
α [°] | 90 | Data/restraints/parameters | 3314/2/191 |
β [°] | 96.2680(10) | Goodness-of-fit on F2 | 1.055 |
γ [°] | 90 | Final R indexes [I ≥ 2σ(I)] | R1 = 0.0440, wR2 = 0.1138 |
Volume [Å3] | 1539.18(10) | Final R indexes [all data] | R1 = 0.0546, wR2 = 0.1212 |
Z | 4 | Largest diff. peak/hole/e Å−3 | 0.13/−0.20 |
ρcalc g cm−3 | 1.227 | CCDC number | 2![]() ![]() |
It is remarkable to note that in the reactivity study, all catalysts experimentally tested were considered. However, for the reaction mechanism analysis, only three catalysts were selected, in the following order: InCl3, which demonstrated the highest yield; BiCl3, which displayed an intermediate yield; and BF3, which yielded the least.
However, for the BiCl3 catalyst, the formation of RC1 involved a small barrier of 5.0 kcal mol−1. Interestingly, based on our experimental studies, the smallest yields were obtained with the BiCl3 catalyst. It also shows that the catalyst with the lowest Free Energy barrier is InCl3, which is approximately 23.2 kcal mol−1, which agrees with the experimental results because BF3 and BiCl3 exhibit higher energy barriers of 26.9 and 24.1 kcal mol−1, respectively.
An energy barrier is also obtained without the catalyst. In this case, by optimizing the geometry of the RC1 structure, which is a reaction intermediate just before the transition state, a geometry is obtained in which the aldehyde is rejected by the reactant (the most stable C–N distance is approximately 2.82 Å). Considering the configuration of this intermediate, it seems improbable that a transition state can be reached. This justifies the experimental result obtained without the catalyst, where the proposed reaction did not give any yield. In the Povarov reaction, a cationic 2-azadiene system was generated in situ by reacting N-substituted aniline with the corresponding aldehyde. Cationic species are sufficiently electrophilic and can undergo nucleophilic attack from electron-rich olefins to yield the desired tetrahydroquinoline. The acid catalyst is involved in aldehyde activation that generates an attack of the N-substituted amine, generating the iminium ion; consequently, the reactivity of the aldehyde-catalyst complex could play a crucial role in the global cationic Povarov occurrence. Five Lewis acids (InCl3, BiCl3, AlCl3, TiCl4 and BF3) were experimentally evaluated for their ability to facilitate the formation of compound 4. Among these, InCl3 exhibited the highest reaction performance (Scheme 1). Theoretical analysis of the catalyst was performed by calculating the global and local reactivity indices using a conceptual DFT framework (CDFT). Table 2 shows several of the global reactivity indices mostly used in CDFT, the chemical potential (μ) values show that the effect of the catalyst makes the formaldehyde more electronegative, especially for BF3.
CH2O | CH2O⋯BF3 | CH2O⋯BiCl3 | CH2O⋯InCl3 | CH2O⋯AlCl3 | CH2O⋯TiCl4 | |
---|---|---|---|---|---|---|
μ | −0.186 | −0.248 | −0.211 | −0.221 | −0.225 | −0.244 |
η | 0.429 | 0.437 | 0.303 | 0.339 | 0.326 | 0.299 |
ΔNmax | 0.434 | 0.567 | 0.697 | 0.651 | 0.691 | 0.817 |
ω | 1.102 | 1.910 | 2.000 | 1.957 | 2.117 | 2.713 |
Considering the global reactivity indexes, upon O–cat coordination, formaldehyde's chemical potential (μ) becomes more negative, indicating greater electron-accepting ability, with the largest shifts observed for BF3 and TiCl4 (μ ≈ −0.248 and −0.244) and the smallest for BiCl3 (μ = −0.211). Chemical hardness (η) decreases in all cases, with BiCl3 yielding the lowest hardness (η = 0.303), enhancing reactivity. TiCl4 facilitates the highest electron transfer (ΔNmax = 0.817), while BiCl3 shows the lowest (ΔNmax = 0.697). In terms of global electrophilicity (ω), TiCl4 stands out (ω = 2.713), followed by BF3 (1.910) and BiCl3 (2.000). Overall, TiCl4 and BF3 are more effective at activating and stabilizing formaldehyde, whereas BiCl3 exhibits weaker catalytic effects.
In the context of local reactivity, the condensed Parr (Pk−, Pk+) functions and the local electrophilicity index have been used (ωk). Table 3 shows the condensed spin densities for the radical anion and radical cation (atom-condensed Parr functions) for formaldehyde and its catalyzed form.
CH2O | CH2O⋯BF3 | CH2O⋯BiCl3 | CH2O⋯InCl3 | CH2O⋯AlCl3 | CH2O⋯TiCl4 | |||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
Pk− | Pk+ | Pk− | Pk+ | Pk− | Pk+ | Pk− | Pk+ | Pk− | Pk+ | Pk− | Pk+ | |
C | 0.439 | 0.062 | 0.555 | 0.150 | 0.350 | 0.158 | 0.593 | 0.135 | 0.494 | 0.066 | 0.489 | 0.062 |
O | 0.288 | 0.751 | 0.258 | 0.614 | 0.603 | 0.622 | 0.291 | 0.647 | 0.278 | 0.760 | 0.304 | 0.759 |
H | 0.135 | 0.093 | 0.092 | 0.109 | 0.023 | 0.112 | 0.058 | 0.110 | 0.091 | 0.088 | 0.104 | 0.089 |
H | 0.135 | 0.093 | 0.094 | 0.126 | 0.024 | 0.109 | 0.058 | 0.108 | 0.138 | 0.086 | 0.103 | 0.090 |
In addition, the results show a clear correlation between the global electrophilicity index (ωk) of the catalysts and their respective reaction yields. In general, catalysts with higher ωk values tend to produce higher yields, suggesting that the catalyst's ability to accept electron density plays a key role in its efficiency. InCl3, with the highest ωk value (1.161), achieves the highest yield (92%), indicating strong stabilization of the transition state and enhanced activation of formaldehyde. BF3, with an intermediate ωk value (1.061), also shows a relatively high yield (44%), while TiCl4 and AlCl3, with lower ωk values (0.778 and 0.604, respectively), exhibit moderate yields (60% and 58%). BiCl3, despite having a lower ωk value than BF3 and InCl3 (0.700), results in a lower yield (35%), suggesting that additional factors beyond ωk may influence its catalytic performance. Overall, these findings reinforce the idea that the electronic activation of formaldehyde by the catalyst, as evidenced by variations in its local electronic properties, is directly related to reaction efficiency, with InCl3 being the most effective and BiCl3 the least favorable catalyst. Note that when observing the yield values and electrophilicity index, this leads us to believe that the effect of the catalyst increases the electrophilic character of the formaldehyde carbon, favoring the reaction with N-propargylaniline 1 and the formation of product 4 (Scheme 1) (Table 4).
Interestingly, there were differences in the formation of O–H bonds among the three catalysts. For example, for InCl3 and BiCl3, the number of electrons in O–H is close to 1.0e, whereas for BF3, this value is 1.8e. However, the valence shell of the oxygen atom in the case of BiCl3 presents two lone pairs (N(O1) and N(O2)) instead of one (N(O1)), as observed for both InCl3 and BF3.
In connection with the thermodynamic analysis discussed earlier, the accumulation of pair density on the O–H bond and the valence shell of oxygen could increase the barrier height during the formation of the iminium cation. This effect was more pronounced in the BiCl3-catalyzed system. Likewise, the structural analysis of the transition states (TS) associated with the formation of the iminium cation reveals significant differences in bond distances depending on the nature of the catalyst, as shown in Table 5. The interaction between the catalyst and the oxygen atom of formaldehyde (Cat–O) is notably stronger for BF3 (1.58 Å) compared to BiCl3 (2.16 Å) and InCl3 (2.08 Å), suggesting a greater Lewis acidity and a stronger polarization of the carbonyl group. However, despite this stronger interaction, the reaction yield follows the trend: InCl3 (92%) ≫ BF3 (44%) > BiCl3 (35%), indicating that factors beyond the strength of the catalyst–substrate interaction play a critical role in determining catalytic efficiency.
BF3 | BiCl3 | InCl3 | |
---|---|---|---|
Cat–O | 1.58 | 2.16 | 2.08 |
C–O | 1.42 | 1.44 | 1.42 |
N–C | 1.51 | 1.50 | 1.51 |
O–H | 1.27 | 1.24 | 1.26 |
N–H | 1.29 | 1.32 | 1.30 |
The C–O bond distances remain relatively consistent across all three catalysts (1.42–1.44 Å), indicating a similar degree of carbonyl activation. Likewise, the N–C bond, characteristic of iminium formation, is nearly invariant (1.50–1.51 Å), suggesting that the transition state structure is maintained regardless of the catalyst. However, subtle variations in O–H and N–H distances suggest differences in proton transfer reaction, which may impact reaction rates and product stabilization. The remarkable yield observed for InCl3 (92%), despite its weaker Cat–O interaction, suggests that additional catalytic factors—such as electronic effects, solvation, or secondary interactions—play a key role in facilitating iminium ion formation. In contrast, although BiCl3 exhibits the shortest Cat–O bond, its moderate yield (44%) suggests that an overly strong interaction may hinder turnover or lead to catalyst deactivation.
In the second stage, two possible cyclization pathways were proposed: concerted and stepwise process (Fig. 4). Following these reaction routes, the potential product formed by the stepwise pathway yields only product 4 (P-Cs), while the concerted pathway (P-Cc) yields THQ 4 and their possible regioisomer THQ 5. However, our experimental results indicate the exclusive formation of compound 4, suggesting that P-Cs are preferred. This finding is further supported by the energy profile and no covalent interaction analysis described later, which reinforces the feasibility of a stepwise mechanism.
![]() | ||
Fig. 4 Overall reaction to obtain (3-yl, 4-yl) THQs from their corresponding IMC and N-vinyl-2 pyrrolidinone (3). |
Fig. 5 depicts the energy barriers (kcal mol−1) of concerted and stepwise cyclizations. In general, the concerted via involves only a transition state (TS-Cc) with a barrier of 28 kcal mol−1, in contrast, stepwise cyclization involves two transition states: TS1-Cs and TS2-Cs. The first stage involves the formation of intermediate Int-Cs (Mannich adduct), where the C–C bond forms between the olefinic carbon of IMC and another olefinic carbon belonging to 3, which involves a barrier of ∼6 kcal mol−1.
![]() | ||
Fig. 5 Free energy kcal mol−1 for the concerted (panel a) and stepwise (panel b) cyclization stages of the IMC + 3 reaction. |
In addition, P-Cs formation (the final product of cyclization) entails TS2-Cs starting from Int-Cs with a low barrier of ∼1 kcal mol−1. In this context, a comparison of the energy profiles in Fig. 5 indicates that stepwise cyclization is favored over concerted cyclization, which occurs through intramolecular Friedel–Crafts alkylation reaction (Fig. 6).
Table 6 reports the geometrical data concerning the transition states involved in the conversion of N-vinyl-2-pyrrolidinone and the iminium cation into N-propargyl-6-methoxy-4-(2′-oxopyrrolidin-1′-yl)-1,2,3,4-tetrahydroquinoline. The bond distances in the concerted transition state (TS-Cc) indicate a synchronous bond reorganization. The C1–C2 (1.41 Å) and C3–C4 (1.43 Å) bond lengths are close to typical single/double bond values, while the C2–C3 bond remains elongated at 2.44 Å, suggesting significant electronic reorganization.
C1–C2 | C2–C3 | C3–C4 | C4–N | N–C5 | C1–C5 | |
---|---|---|---|---|---|---|
TS-Cc | 1.41 | 2.44 | 1.43 | 1.38 | 1.37 | 1.87 |
TS1-Cs | 1.38 | 3.76 | 3.21 | 1.43 | 1.33 | 2.24 |
TS2-Cs | 1.51 | 2.11 | 1.44 | 1.36 | 1.44 | 1.56 |
The N–C5 (1.37 Å) and C1–C5 (1.87 Å) distances indicate the formation of the tetrahydroquinoline ring in a single step. On the contrary, considering the stepwise pathway in TS1-Cs, the formation of key bonds is less advanced, as indicated by the significantly elongated C2–C3 (3.76 Å) and C3–C4 (3.21 Å) distances. The C1–C5 bond (2.24 Å) is also longer than in TS-Cc, suggesting an initial stage of bond rearrangement. The N–C5 distance (1.33 Å) is slightly shorter than in TS-Cc, possibly due to stronger stabilization of the iminium moiety at this stage. In addition, TS2-Cs represents the second step of the reaction, where bond formation is more advanced. The C2–C3 distance contracts significantly (2.11 Å), indicating partial bond formation, while the C1–C5 bond (1.56 Å) shortens compared to TS1-Cs but remains longer than in TS-Cc, suggesting that the final structural reorganization is still incomplete. The N–C5 bond elongates to 1.44 Å, reflecting the final electronic adjustments needed to achieve the product structure.
Chemical bonding analysis of the cyclization stage reveals bonding patterns that differentiate the concerted route from the stepwise one. In the concerted route, the formation of the first C–C bond was observed with an electron number of N(C5–C1) = 1.2e. This electron population, which is less than 2.0e, corresponds to a highly delocalized single bond formed in TS-Cc. Additionally, the double bond C1C2 in 3 was significantly weakened with N(C1–C2) = 2.6e.
Considering the stepwise route, in TS1-Cs, the first C–C bond has not yet formed, as there are non-bonding electron density centers on the C2 and C5 carbons. That is, the electron density was not shared between the carbons of IMC (C5) and 3 (C2), with electron populations of N(C5) = 0.47e and N(C2) = 0.26e. Subsequently, after IntS formation in TS2-Cs, the C2–C5 bond emerges with N(C2–C5) = 1.8e. However, the remaining C–C bond for P-Cs formation has not yet emerged in TS2-Cs, as the pairing density is divided between the C1 and C3 carbons N(C1) = 0.20e and N(C3) = 0.46e. In summary, the high energy barrier of the concerted process could be attributed to the presence of bonding density, which characterizes the formation of the first C–C bond between IMC and 3. In contrast, for the stepwise process, C–C bond formation is preceded by the appearance of non-bonding density centers, associated with relatively low energy barriers.
In addition, we performed noncovalent interaction calculations using a reduced-gradient density scalar field. The results for the TS-Cc, TS1-Cs, and TS2-Cs structures are shown in Fig. 7. Notably, in the TS-Cc structure, there are repulsive C–C and C–O interactions. Specifically, C–C repulsion arises between the six-membered ring of N-propargyl-4-methoxyanilinium and the N-vinyl-2-pyrrolidinone moiety. Similarly, a C–O repulsive interaction occurred between the carbonyl group of N-vinyl-2-pyrrolidinone and the CH2–CH fragment present in both reactants. Interestingly, these interactions exhibit different characteristics in stepwise transition states (TS1-Cs and TS2-Cs), where van der Waals interactions replace the repulsive force observed in the concerted mechanism. This is particularly relevant because the high barrier associated with the concerted mechanism correlates with repulsive interactions during cyclization. Conversely, the lower barriers of the stepwise mechanism are linked to van der Waals interactions, which mitigate the repulsive interactions characteristic of the cyclization process.
Our findings reveal that the interaction between the catalyst and reactants significantly enhances the reactivity of the aldehyde, facilitating the formation of the iminium ion intermediate and subsequent cyclization to yield the desired tetrahydroquinoline product. The analysis of the global and local reactivity indices further supports the conclusion that the catalyst increases the electrophilic character of the formaldehyde carbon, promoting the reaction. Additionally, the chemical bonding analysis of the iminium cation formation suggests that the high accumulation of pairing density over the O–H bond increases the reaction barrier.
For the cyclization stage, our results indicate that concerted cyclization involves C–C bond formation in the transition state, implying a high energy barrier. In contrast, the stepwise pathway shows that both C–C bonds emerge from non-bonding centers originating in the two transition states, each with low-energy barriers. This energetically favorable route, which leads to obtaining the final product THQ, involves the formation of a Mannich-adduct and finally a cyclization reaction via Friedel–Crafts intramolecular alkylation.
Based on these results, we propose that the term ‘Diels–Alder’ should no longer be used to describe these types of reactions. This study provides a detailed understanding of the mechanism of the cationic Povarov reaction and the influence of Lewis acid catalysts, which can be leveraged to optimize and develop new synthetic strategies for tetrahydroquinoline derivatives with potential pharmacological applications.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2415596. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5ra01375e |
This journal is © The Royal Society of Chemistry 2025 |