Next Article in Journal
The Impact of Redox, Hydrolysis and Dehydration Chemistry on the Structural and Magnetic Properties of Magnetoferritin Prepared in Variable Thermal Conditions
Next Article in Special Issue
Straightforward Access to Enantioenriched cis-3-Fluoro-dihydroquinolin-4-ols Derivatives via Ru(II)-Catalyzed-Asymmetric Transfer Hydrogenation/Dynamic Kinetic Resolution
Previous Article in Journal
Analysis of Fragmentation Pathways of Peptide Modified with Quaternary Ammonium and Phosphonium Group as Ionization Enhancers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rational Design of Simple Organocatalysts for the HSiCl3 Enantioselective Reduction of (E)-N-(1-Phenylethylidene)aniline

by
María Maciá
1,
Raúl Porcar
1,2,
Vicente Martí-Centelles
1,3,
Eduardo García-Verdugo
1,*,
Maria Isabel Burguete
1 and
Santiago V. Luis
1,*
1
Department of Inorganic and Organic Chemistry, Jaume I University, Av. Vicent Sos Baynat s/n, 12071 Castellón, Spain
2
Department of Organic and Bio-Organic Chemistry, Faculty of Science, UNED—Universidad Nacional de Educación a Distancia, Avenida de Esparta s/n, 28232 Las Rozas-Madrid, Spain
3
Instituto Interuniversitario de Investigación de Reconocimiento Molecular y Desarrollo Tecnológico (IDM), Universitat Politècnica de València, Universitat de València, 46022 Valencia, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(22), 6963; https://doi.org/10.3390/molecules26226963
Submission received: 29 October 2021 / Revised: 16 November 2021 / Accepted: 16 November 2021 / Published: 18 November 2021

Abstract

:
Prolinamides are well-known organocatalysts for the HSiCl3 reduction of imines; however, custom design of catalysts is based on trial-and-error experiments. In this work, we have used a combination of computational calculations and experimental work, including kinetic analyses, to properly understand this process and to design optimized catalysts for the benchmark (E)-N-(1-phenylethylidene)aniline. The best results have been obtained with the amide derived from 4-methoxyaniline and the N-pivaloyl protected proline, for which the catalyzed process is almost 600 times faster than the uncatalyzed one. Mechanistic studies reveal that the formation of the component supramolecular complex catalyst-HSiCl3-substrate, involving hydrogen bonding breaking and costly conformational changes in the prolinamide, is an important step in the overall process.

1. Introduction

Chiral amines are an important class of organic compounds widely used as building blocks in synthetic organic chemistry [1,2]. A large number of chiral active pharmaceutical ingredients, about one-third of the total API market, and agrochemical substances are amines or contain functional groups derived from amines. These include (S)-Rivastigmine (Alzheimer’s and Parkinson’s diseases), NPS R-568 (hyperparathyroidism treatment), and (R)-Fendiline (calcium channel blocker) [3]. In this context, the asymmetric transfer hydrogenation of imines represents a simple and attractive approach to target the synthesis of enantiopure amines, and a variety of catalytic methodologies have been developed for this purpose [4,5,6,7,8,9,10,11,12]. The use of trichlorosilane and related reagents for this purpose has attracted much attention in recent years, having shown their capacity for the transfer hydrogenation of a variety of groups, such imines [4,5,6,7,8,9,10,11,12], enones [13,14], nitro groups [15,16], and N-heteroarenes [17,18], in the synthesis of substituted hydrazines [19], and in the reduction of CO2 in the presence of amines [20,21,22,23]. Thus, the organocatalytic reduction of ketimines with trichlorosilane provides an efficient methodology for the asymmetric preparation of amines [24,25,26]. Different organocatalysts have been reported and evaluated against a series of benchmark ketimines [27,28,29,30,31,32,33] and related substrates, in particular enamines [34,35,36,37,38,39], to compare their reactivity and selectivity. Some of them have provided very good levels of asymmetric induction, but their development was often based on empirical approaches, where structural modifications were sequentially tested until the required level of chiral induction was achieved. To carry out a more rational catalyst design, better insight on the mechanisms for this process is needed to understand the effects of structural modifications. However, only a few computational or experimental systematic mechanistic studies are available, and, in most instances, tentative models are elaborated in order to understand the observed enantioselectivity at the molecular level [40,41,42,43]. In the seminal work of Schreiner and co-workers, a detailed evaluation of the possible mechanism of this reaction using density function theory calculations (B3PW91/cc-pVDZ) was reported [40]. These studies were performed using simplified models for both ketimine and catalyst. The calculated transition state for the uncatalyzed reaction involved a four-membered cyclic structure of high energy (ΔH = 40.0 kcal/mol), whereas the enthalpy of activation calculated for the reaction catalyzed by a HSiCl3-ligand complex was clearly lower (19.2 kcal/mol). The function of the ligand appeared to be to coordinate with trichlorosilane and serve as a proton donor for the imine. Based on these calculations, the authors proposed a reaction mechanism and explained the experimentally observed stereoselectivity. The catalytic reduction of ketimines with trichlorosilane is described as a formal transfer reaction of H+/H to the C = N double bond. Recently, Jones and co-workers have performed a kinetic study on an imidazole-based catalytic system able to display a dual activation mechanism, leading to high catalytic activity at unprecedented low levels of catalyst loading [41]. This study, however, did not provide any computational studies on these systems to reveal at a molecular level the activation mode proposed. A different mechanistic proposal, involving the participation of two molecules of trichlorosilane, has been recently reported by Dong et al. in the case of axially chiral biscarboline-based alcohols [43].
Natural amino acids are a simple, abundant, and cheap source of chirality for the preparation of organocatalysts [44]. Indeed, most of the catalytic systems described in the literature for the reduction of ketimines with HSiCl3 include in their structure an amino-acid-derived fragment as the source of chirality. According to the most general mechanism considered, the Lewis basic sites found in the organocatalyst interact with HSiCl3 to form a hexa-coordinated species able to transfer the H+/H couple to the C=N double bond of the ketimine. Taking this into account, any amino-acid-based catalyst for the enantioselective reduction of ketimines with HSiCl3 should contain a functional group at the nitrogen providing a second coordination site for the silicon atom of HSiCl3. Amine-protecting groups, such as carboxybenzyl (Cbz) or related fragments, contain carbonyl groups that can act as this second coordination site, and a variety of them are commercially available.
Here, we report our efforts to design l-proline-based organocatalysts for the reduction of ketimine 1 with HSiCl3 as a benchmark reaction (Scheme 1). We have used kinetic and computational studies to shed light on the mechanism of these systems, allowing a rational design of more efficient catalysts in terms of both activity and enantioselectivity.

2. Results and Discussion

2.1. Catalyst Screening

An initial screening of Cbz-protected amino acids as organocatalysts for the benchmark reduction reaction of ketimine 1 with HSiCl3 was performed. All Cbz-protected amino acids (Ala, Phe, His, Trp, Val, and Pro) provided good yields (>85%, entries 2–7, Table S1, Supplementary Materials), but it is important to note that under those standard conditions ([1] = 0.512 M, 1.5 equiv. HSiCl3, 40 mol% organocatalyst, CH2Cl2, 0 °C, 16 h), the uncatalyzed reaction providing the racemic product is important (entry 1, Table S1) [45,46]. Thus, the catalyst role was mainly analyzed through the observed enantioselectivity. The low degree of induction achieved with Cbz-l-Val was surprising, since a variety of catalysts have been reported based on this amino acid [4,5,6,7,47,48,49]. Only appreciable asymmetric induction (62% ee) was observed for Cbz-l-Pro, probably due to its cyclic structure that can reduce the conformational freedom at the transition state, favoring an efficient chirality transfer [50]. Therefore, further proline-containing organocatalysts were used in the same reaction. While no asymmetric induction was observed for l-Pro (entry 8, Table S1), the catalytic activity was maintained for functionalization at the l-Pro nitrogen atom using a carbamate group (Boc-l-Pro, entry 9, Table S1) or different amides (N-acetyl-l-Pro or Piv-l-Pro, entries 10-11, Table S1), highlighting the key role of the carbonyl group attached to the l-Pro amine group. The enantioselectivity was similar for carbamates Cbz (62% ee) and Boc (63% ee) and slightly better for amides (N-acetyl-l-Pro (68% ee) and Piv-l-Pro (77% ee). This observation was in line with previous studies suggesting pivaloyl (Piv) as an excellent N-substituent in related organocatalysts [42].
To further improve the efficiency of these l-Pro systems, a variety of N-carbamate-amides and bisamides were designed (Figure 1). The l-Pro amine group was N-protected by either Cbz or Piv groups, and the carboxylic group was functionalized using different types of aliphatic and aromatic residues. Synthesis of organocatalysts 316 was achieved from the corresponding N-protected l-proline derivatives [51,52,53,54].
Organocatalysts 39 (first generation catalysts) were tested for the reduction of the benchmark reduction of ketimine 1 with HSiCl3 under the same standard reaction conditions (Table 1). Catalysts 3, 4, 6, 7, and 8 afforded amine 2 in good yields (>83%). In general, the presence of the amide group in the structure of the catalyst improved the asymmetric induction with regard to Boc-l-Pro (from 62% to 81–85% ee for Cbz-l-Pro derivatives). Organocatalysts 5 and 9 show the importance of the proline amide proton in the catalytic cycle, absent in catalyst 5 and strongly sterically hindered in catalyst 9. For both organocatalysts, racemic product was obtained (entries 3 and 7, Table 1) in good agreement with the mechanistic proposal by Jones [41].
In light of the results in Table 1, the effect of catalyst loading (40 to 5 mol%) was evaluated for catalysts 3, 4, 6, and 8 (Figure 2). While catalysts 4 and 6 showed a significant reduction in enantiocontrol when reducing the amount of catalyst, catalysts 3 and 8 maintained their activity at low catalyst loadings. Catalyst 3 retained the asymmetric induction (83–85% ee) until a 10 mol% catalyst loading, with an appreciable decrease (ca. 70% ee) for a further decrease in loading to 5 mol%. Catalyst 8 was the most active, displaying only a minor variation of enantioselectivity (85–90% ee for 5 to 40 mol% loadings). A good correlation between yield and asymmetric induction was observed, especially for the most active catalysts, having an appreciably quicker rate than the uncatalyzed background reaction conditions (Table S2, Supplementary Materials) [45,46]. Catalyst 8 was also the most active, with only a minor decrease in yield for the lowest catalyst loading (Δyield5–40% ≈ 5% vs Δyield5–40% ≈ 36% for compound 4). Thus, there is a clear relationship between the molecular structure of the catalyst and the chiral induction, which is more relevant for low molar loadings. The comparison of Cbz-l-proline derivatives 3, 4, and 6, just differing on the amide residue, shows that the aniline derivative provided the best results [54]. On the other hand, data for catalysts 6 and 8, both being butylamine derivatives, suggested that N-Piv protection can improve the efficiency significantly. For the most active catalysts, 10 mol% catalyst loading seems to be an excellent option [55].

2.2. Kinetic Studies

For a better understanding of such differences, a kinetic study was carried out by in situ 1H-NMR monitoring of the model reduction using catalysts 3, 6, and 8 in CDCl3 (0 °C) and a 10 mol% loading. The signals characteristic for the amine formed, and the starting imine and the acetophenone obtained by hydrolysis of the imine were monitored (Figure S1, Supplementary Materials). The resulting kinetic profiles are presented in Figure 3. Changing the solvent from CH2Cl2 to CHCl3 in this kind of reaction often provides lower ee values and slower reaction rates [54,56,57,58], but slower kinetics can enable more accurate NMR kinetic analyses and reduce the contribution of the uncatalyzed reaction.
The rate constants were calculated considering a simultaneous contribution of the uncatalyzed and catalyzed reactions with first-order rate equations in both substrate and reagent, in agreement with most mechanistic studies (Scheme 2 and Equation (1)) [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42]. Although for axially chiral biscarboline-based alcohols, the participation of two molecules of HSiCl3 has been included in the reaction mechanism, to consider the quantitative reaction of this reagent with alcohols to produce hydrogen, this could be discarded in our case, as this process was not observed [43]. Accordingly, the uncatalyzed reaction has a rate v = k1[A][B] and can be obtained for the reaction in the absence of any catalyst. The catalyzed reaction should be defined as v = kcat[A][B] but taking into account the analysis of the effects of catalyst concentration, kcat can be defined as k2[D], while k represents the global experimental constant of the process (Equation (1)). The integration of Equation (1) provides the kinetic model (Equation (2)). Although in one instance, nonlinear effects (NLE) have indicated the participation of two catalyst molecules in the mechanism [59], NLE data have confirmed in most instances the involvement of one single catalyst molecule in the mechanism, in particular for catalytic systems related to the ones in this study and involving a dual activation mechanism [24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43].
v c = d C dt = k 1 A B + k 2 A B D   =   k 1 + k 2 D A [ B ] = k A B
C theoretical = A 0 B 0 ( 1   -   e B 0   -   A 0 kt ) [ A ] 0 -   B 0   ×   e B 0   -   A 0 kt
Fitting the experimental data in the presence of catalyst to Equation (2) allowed us to obtain the experimental kinetic constant (k). From this kinetic constant, it was possible to obtain the value of k2 from the concentration of catalyst and the uncatalyzed kinetic constant (k1). Table 2 summarizes the values of the rate constants obtained in this way. The fittings results confirmed the validity of the proposed kinetic model (see, for instance, Figure S2, Supplementary Materials).
The k2/k1 coefficient facilitates an easy comparison of the different catalysts. Thus, the most active one was catalyst 3, displaying a kinetic constant for the catalyzed reaction almost 100 times higher than the one for the uncatalyzed process (entries 1 and 2, Table 2). Data also show that the substitution of the Cbz group for the Piv group increases the reaction rate of the catalyzed reduction (entries 3 and 4, Table 2).
It must be borne in mind that a higher activity of the catalyst can significantly enhance the asymmetric induction, as the uncatalyzed reaction, affording a racemic mixture, will reduce its contribution. Thus, Equation (1) shows that it is required that k2[D] >> k1 for fast and highly enantioselective processes, and the concentration of the catalyst ([D]) can play a key role. Accordingly, two additional sets of experiments were performed. Initially, different concentrations of catalyst 6 (from 0.50 to 0.01 M) were analyzed, but keeping constant the substrate–catalyst ratio (40 mol%) by simultaneously reducing the concentration of imine 1 (Table S3, Supplementary Materials). For catalyst concentrations above 0.2 M, yield and enantioselectivity were high and almost identical (yield > 90%, 81% ee, entries 1–2, Table S3). A decrease of the concentration to 0.1 M led to some activity reduction (65% yield, entry 3, Table S3), although enantioselectivity was maintained (84% ee). However, a further tenfold reduction of catalyst concentration produced a drastic reduction of both activity and asymmetric induction (9% yield, 40% ee, entry 4, Table S3). In an additional study, the concentration of catalyst 6 was kept constant at 0.205 M, but its molar percentage was varied from 5 to 40 mol% (Table S4, Supplementary Materials). The reduction in catalyst loading was accompanied by a decrease from 94 to 70% in yield and from 81 to 65% in enantiomeric excess. Therefore, the activity and enantioselectivity of catalyst 6 is influenced by both its concentration and the catalyst–substrate ratio (Figure S3, Supplementary Materials). Interestingly, when the catalyst molar concentration was kept high enough, an appreciable efficiency was observed even for low molar loadings. As suggested by data in Figure 2, this dependence is very sensitive to catalyst structure. Thus, catalyst 8 maintained the enantioselectivity for the reduction of 1 even at low concentrations and for small catalyst loadings (5 mol%) (Cat 8 = 0.026 M, 85% ee; Cat 8 = 0.205 M, 82% ee, Table S5, Supplementary Materials). These differences can be explained by taking into account Equation (1). For catalyst 6, k2 is relatively close to k1, and when its concentration decreases, the product k2[A][B][D] becomes rather small, with the uncatalyzed reaction increasing its contribution and leading to an important reduction in enantioselectivity. In contrast, for catalyst 8, k2 is almost 90 times larger than k1. Thus, a much larger reduction in the concentration of 8 is needed to observe a significant increase in the relative importance of the uncatalyzed reaction. Formation of intermolecular aggregates with the increase of the concentration of catalyst could also contribute to this phenomenon, but 1H-NMR experiments with catalysts 6 and 8 at different concentrations and temperatures discarded a significant involvement of them (Figure S4, Supplementary Materials).

2.3. Structural Catalyst Optimization

The former catalyst screening and computational calculations run in parallel (see below), provided two key elements for catalyst optimization: the use of N-protected prolinamides derived from aromatic amines and the use of the pivaloyl protecting group instead of Cbz. Besides, computational studies suggested that aromatic substitution could have important effect. Thus, Cbz-protected compounds 1014 (second-generation catalysts) were synthetized, and their behavior against the molar percentage, at a constant imine concentration (0.512 M in CH2Cl2), was also evaluated (Table S6, Supplementary Materials). Catalyst 10 showed a good activity but no enantioselectivity at 40 mol% (91% yield, 4% ee, entry 1, Table S6, Supplementary Materials), while catalyst 11 presented moderate-low activity and selectivity under the same reaction conditions (78% yield, 50% ee, entry 2, Table S6), highlighting again the importance of the amide NH fragment and its environment, with groups providing significant steric hindrance in its surroundings, contributing to a decrease in the level of enantioselectivity afforded. Interestingly, catalysts 12 and 13 maintained their efficiency up to 10 mol% loading (83% ee in both cases, Figure 4 and Table S6), but with an ACE (asymmetric catalyst efficiency [60]: a formula that takes into account the amount of catalyst employed and the relative size of the catalyst to the product, as well as the ee and yield of the product, in order to measure the catalyst efficiency) slightly higher at this loading for 13 (4.39 vs. 3.35). These results were comparable with those for catalyst 3 (83% ee for 10% catalyst loading, ACE = 4.44, entry 2, Table S2). Finally, catalyst 14 only afforded an enantiomeric excess above 80% when a 40 mol% loading was used.
Kinetic constants (in CDCl3) for catalysts 12–14 are presented in Table 3 (see also Figure S5, Supplementary Materials). For catalysts displaying electron donor groups on the aromatic ring (12, R = tert-butyl; 13, R = -OCH3), k2/k1 values were approximately 58 and 106. However, for the catalyst with an electron acceptor group on the aromatic ring (14, R = -NO2), both constants were relatively similar, and therefore the enantioselectivity was low (10% ee). For this last catalyst (14), it must be noted, however, that a complex reaction crude was observed, which can be associated with the potential reduction of the NO2 group, not only consuming part of the HSiCl3 but also potentially producing a different catalytic system [15,16].
A second catalyst optimization step was performed by changing the N-protecting group from Cbz to Piv (third-generation catalysts). Thus, catalysts 15 and 16 were evaluated, and the results (Table 4) were compared with those for related 12 and 13. Both catalysts allowed the obtainment of the chiral amine with good enantioselectivity using only 1 mol% of catalyst (entries 1 and 4, Table 4), increasing the ACE values to 47.87 for 15 and to 49.05 for 16. The effect of the substitution of Cbz by Piv was also analyzed by monitoring the kinetic profile by 1H-NMR (in CDCl3) for the most efficient catalyst 16 (Figure 5 and Supplementary Materials, pages S27 and S29), revealing an increase in the observed kinetic constant (k) by almost one order of magnitude when compared with the related catalyst 13 (0.21395 and 0.04326 for 16 and 13, respectively, k2 = 3.1594 for 16). In fact, the k2/k1 ratio for 16 was 582, confirming this compound as the best catalyst in this study, leading to the highest enantioselectivity (86% ee) with the lowest amount of catalyst (1 mol%, ACE = 49.05).

2.4. Continuous Flow Reduction

Thus, mechanistic studies and experimental screening allowed the selection of prolinamide 16 as a simple and cheap catalyst with enhanced activity and enantioselectivity for the process considered. In batch, with just a 1 mol% catalyst loading of 16 and three hours of reaction, the chiral amine 2 was obtained in 98% yield and 82% ee. This facilitated the performance of the imine enantioselective reduction under continuous flow conditions. For this purpose, two flows, one containing HSiCl3 and the second the catalyst 16 and imine 1, were combined in a mixing valve and were pumped through a tubular reactor. Both flows were adjusted to reach a 10 mol% catalyst loading at the exit of the mixing valve, and samples were taken at different times after the reactor. Once the steady state was reached, a constant yield of 90% with an enantiomeric excess of 87% was obtained for more than nine hours (Figure S6, Supplementary Materials).

2.5. Computational Studies with Simplified Model Compounds

In order to better understand the mechanism for this process and rationalize catalyst optimization results, computational studies were carried out in parallel, following the sequential approach described before for experimental studies. The results will be discussed here in an integrated way for simplicity. Besides some tentative mechanistic models, there are only two reports providing detailed theoretical calculations based on DFT studies, although the mechanism considered in the more recent one cannot be applied in our case [40,43]. The original DFT-based mechanism, using simplified models of the imine and catalyst proposed, involved the formation of a catalyst–HSiCl3 complex capable of acting in a formal H+/H transfer to the C=N bond [40]. Based on this antecedent, and trying to identify the role of the different structural factors, the faster PM6 semiempirical method was chosen, in order to be able to work with structures closer to the real ones, and calculations were performed with the Gaussian 09 program [61]. To validate this method, the uncatalyzed reaction pathway was initially studied with the simplified imine model used by Schreiner. In this process (Figure 6), the reactants interaction complex (RIC) formed by the imine and HSiCl3 evolves to a transition state (TS) of four centers, transferring the hydrogen of trichlorosilane to the imine carbon atom and giving rise to the products interaction complex (PIC). The calculated energy barrier was 40 kcal/mol, as in the DFT calculations [40], and, therefore, this method of calculation, facilitating the analysis of the effects of structural changes in the catalyst, was considered valid.
The conformational space of a model for catalyst 6, where butyl and benzyl groups were replaced with methyl groups, was then studied using the same imine model presented in Figure 6. In agreement with its 1H-NMR spectrum displaying two amide NH signals of similar intensity (Figure S4A, Supplementary Materials), the two possible intramolecular amide NH hydrogen bonds in Cbz-protected compounds were considered (Figures S7 and S8A, Supplementary Materials). Both of them seem capable of interacting with HSiCl3 to form an hexacoordinate complex (Figure S8B, Supplementary Materials) in which the silicon atom is bound to the two carbonyl oxygens of the catalyst, requiring the breaking of the previous intramolecular hydrogen bonds. This complex can present different structures depending on the distribution around the silicon atom (Figure 7A). Since the structures with the hydrogen in equatorial position showed lower energy than the apical position ones, the transition state was evaluated exclusively with structures C and D. Only with structure C (Figure 7B) was it possible to obtain an optimized geometry of the TS (Figure 7C).
Figure 8 shows the energy profile for this simplified model of catalyst 6 and the calculated energies for the optimized geometries of reagents and the corresponding RIC, TS, and PIC. The energy barrier from RIC to TS decreases for the catalyzed process from 40 to 10.59 kcal/mol. This barrier is lower than the one calculated by Schreiner (19.1 kcal/mol) [40], highlighting the relevance of the substitution pattern at the two carbonyls coordinated to silicon.
Thus, the overall mechanism to be considered for N-carbamate-proline catalysts (Figure 9) involves the initial interaction of HSiCl3 with the two possible conformations of the N-carbamate-amide to form weak complexes that evolve towards a common RIC in the presence of imine. The resulting TS presents a Namide···H···Nimine interaction that results in the new amine N-H bond, while, simultaneously, the Si···H···Cimine interaction leads to the formation of the C-H bond at the new stereogenic carbon atom of the amine. Additional supramolecular interactions, which have been suggested to be relevant for this and related processes, in particular those involving the aromatic rings, can also contribute to stabilizing the TS and decreasing the energy barrier [24,25,26,42,62,63,64]. The subsequent PIC affords the final product (chiral amine–SiCl3) by a formal proton transfer of the amine to the catalyst–SiCl3 complex, which is recovered for a new catalytic cycle. This mechanism is in line with the dual activation model described for imidazole-derived compounds and other organocatalysts [41,65].

2.6. Mechanistic Studies with Real Compounds

In a second step, the energy profiles of a series of N-carbamate-amides differing in the amide substituents (compounds 3, 5, 6, 12, 13, and 14) and for the Piv derivative 16 were calculated (Table S7, Supplementary Materials). The real structures of the catalysts and imine 1 were used to calculate the energy of the interaction complexes (for both enantiomers) and the energy barrier between RIC and TS. Calculations revealed that in the optimized structure of the catalyst–HSiCl3 chelate complex, the seven-member ring can acquire two boat conformations, depending on the position of the amide NH-R group (Figure S9, Supplementary Materials, up- and down-conformation), and both conformations were considered for the calculations [66,67]. The reaction profile obtained for the pyrrolidine-derived catalyst 5 confirmed the crucial role of the Namide···H···Nimine interaction. Lacking the amide hydrogen, the value of ΔERIC-TS was very high (48.87 kcal/mol, entry 3, Table S7, Supplementary Materials) for the R product, and it was not possible to obtain the energies of the interaction complexes for the S product. For the other catalysts, the energy barrier ΔERIC-TS was always lower than for the uncatalyzed reaction. However, based on the different ΔERIC-TS obtained, calculations predicted that the reaction rates for those compounds should follow the trend v14 > v3 > v12 > v13 > v16 > v6 > vuncat, while the observed experimental order was v16 > v13 > v3 > v12 > v6 > v14 > vuncat. Although the order for catalyst 14 can be rationalized in terms of the side reactions observed, it seems clear that important discrepancies are present in both trends. According to Equation (3), a linear relationship between ln k and the energy barrier should be expected, but when the ΔERIC-TS energy barrier was compared with the experimental rate constant obtained for the catalyzed process (k2), since computational studies only considered the catalyzed reaction, this relationship was not observed (Figure S10, Supplementary Materials).
k   =   k B · T h × e Δ G RT
P S P R = k S k R = e δ Δ E RIC TS / RT δ Δ E RIC TS = Δ E RIC TS S Δ E RIC TS R e e   % = e Δ E RIC TS S Δ E RIC TS R / RT 1 e Δ E RIC TS S Δ E RIC TS R / RT + 1 × 100
The barrier for the formation of the S enantiomer was always lower than for the R (Figure S11, Supplementary Materials), in excellent agreement with the preferential formation of the S-amine in the corresponding experiments. In principle, the predicted enantiomeric excesses could be estimated by comparing the energy barriers differences for the formation of both enantiomers (Equation (4)) [43]. Thus, eecalcd. values were obtained using ΔERIC-TS for each enantiomer, as the average of the values for the up and down conformations (eecalcd.) of the catalyst–HSiCl3 complex. However, some significant discrepancies were observed between eecalcd. and eeexp. values (Table S8, Supplementary Materials). Thus, the predicted enantiomeric excess was lower than the experimental one for 3 when considering a 10 mol% catalyst loading (32% eecalcd. < 76% eeexp., entry 1, Table S8), while it was more comparable for 6 (70% eecalcd. ≈ 63% eeexp., entry 2, Table S8).
All the discrepancies observed in the former calculations could be explained because these calculations were performed considering only the energy barrier between RIC and TS (ΔERIC-TS), and RIC formation was not taken into account. According to the calculations, the observed rate should be dependent on the value of ΔERIC-TS and on (RIC). For a situation in which the RIC is quantitatively formed from the substrates and the catalyst, (RIC) would directly reflect their concentrations, and a direct comparison with the experimental rates is possible. However, if the RIC involves the formation of a weak complex, then its concentration, for a given concentration of substrates and catalyst, can significantly differ according to the structure of the catalyst. Overall, this suggests, along with the observed structure-sensitive dependence of the enantioselectivity from catalyst loading and concentration, that the formation of the catalyst–HSiCl3 complex (i.e., structures E and E’ in Figure 9), requiring the breaking of intramolecular hydrogen bonds and modifying of the relative disposition of the carbonyl groups to adopt a cis conformation, can be an important step to define the relative contribution of the catalyzed/uncatalyzed reactions and accordingly the observed enantioselectivity. When a solution of catalyst 6 and trichlorosilane in CDCl3, at the same concentrations used for imine reduction, was studied by 1H- NMR, the resulting spectra showed one single signal for H-Si (the same for 29Si-NMR) and just some changes in the chemical shifts of the signals for the two intramolecularly hydrogen-bonded conformers present in 6 (Figure S12, Supplementary Materials). This ruled out the quantitative formation of the 6–HSiCl3 complex contributing to the RIC and indicated the presence of weak 6–HSiCl3 complexes in fast equilibria (i.e., structures E and E’ in Figure 9) as the major component. Spectra taken just after addition of imine 1 again showed a similar behavior, displaying some additional shifts for the signals. The substitution of the Z group by pivaloyl (catalyst 8) afforded similar observations, although in this case, one single intramolecularly hydrogen-bonded structure is present (Figure S13, Supplementary Materials).
To take this issue into account, the energy of the initial conformations for the different catalysts was also calculated (Table S9, Supplementary Materials) and used to obtain an initial coordinate (ABC) that would be defined by the sum of reagents energies (EABC) to be included in the corresponding energy profile (Figure 10) [43]. Thus, the contribution for the formation of the RIC (requiring hydrogen bond breaking and conformational rearrangements) can be associated to ΔEABC-RIC = ERIC − EABC, and a parameter (ΔETOT) can be defined as ΔETOT = ΔEABC-RIC + ΔERIC-TS. In this approach, theoretical ee was determined using this parameter (ΔETOT, including both the energy barrier to the TS and a factor to consider the (RIC)) and Equation (5).
P S P R = k S k R = e δ Δ E TOT / RT where   δ Δ E TOT = Δ E TOT S Δ E TOT R e   ( % ) = e δ Δ E TOT / RT 1 e δ Δ E TOT / RT + 1   ×   100
The values of ΔETOT obtained using this approximation, as well as the energy difference for the R or S enantiomer (δΔETOT), allowed us to calculate a new set of expected ee values for each conformer (up and down) and the corresponding average value (Table S10, Supplementary Materials). The deviation between calculated and experimental values of ee was less than 6% for most of the catalysts. Only 6 displayed a deviation of ca. 17%. Therefore, a better fit was obtained in this case between theoretical and experimental enantioselectivity after discarding catalyst 14 according to the former discussion (Figure S14, Supplementary Materials). Figure 11 shows the linear relationship between the experimental rate constants and ΔETOT for the up conformation (see also down conformation in Figure S15, Supplementary Materials). Thus, the theoretical model developed in this work is able to reproduce in a reasonable way the results observed experimentally.

3. Materials and Methods

3.1. Experimental Section

All reagents and solvents were obtained from commercial sources (Aldrich, Fluka, Scharlab or Iris-Biotech) and used without further purification unless otherwise noted. Air and/or moisture-sensitive reactions were carried out under an inert atmosphere of nitrogen, using glass material previously dried in the oven and dry solvents supplied by a Pure Solv model-solvent-dispensing system from Innovative Technology. Moisture-sensitive reagents were handled using syringes under inert atmosphere. Reactions whose procedure required low temperature for extended time were carried out with the aid of a Neslab model CC-100 II Cryocool. Purification of synthesized products was generally performed by crystallization. In cases where it was not possible, column chromatography was performed using a silica gel 60 stationary phase with a particle size of 0.06–0.2 mm. For each occasion, the eluent used has been indicated, as well as the proportions of the solvents in volume–volume. After obtaining synthesized products, they were dried in a Binder vacuum oven at a temperature of 45 °C and then stored in a desiccator or refrigerator according to the needs of the product. Enantiomeric excess of the imine reduction was determined by high-performance liquid chromatography with a Merck HITACHI LaChrom D-7000 chromatograph and a Chiralcel OD-H chiral filler column (4.6 mm φ × 250 mm L). HPLC conditions: n-hexane/MTBE (98/2), 1 mL/min, 30 °C, 254 nm (UV/Vis detection), tR (S-amine): 26 min, tR (R-amine): 27.5 min. 1H-NMR and 13C-NMR spectra were recorded on a Varian model INOVA 500 spectrometer (1H-NMR at 500 MHz and 13C-NMR at 125 MHz) in the indicated deuterated solvent. In the kinetic studies monitored by 1H-NMR, threaded and septum tubes were used to perform the reaction in a closed system with an inert atmosphere. For the product characterization, the spectra were recorded at 30 °C. However, for the mechanism study of imine reduction, spectra were acquired at temperatures below 30 °C.

3.2. General Procedure for the Synthesis of Imine 1

In an oven-dried two-neck flask, a mixture of activated 4 Å molecular sieve (35 g), acetophenone (10 mL, 85.73 mmol), and aniline (10.16 mL, 111.44 mmol) in dry CH2Cl2 (40 mL) was gently stirred at room temperature for 24 h under nitrogen atmosphere. The reaction mixture was filtered through paper filter. The filtrate was concentrated in vacuo, and the residue was distillated by distillation under reduced pressure. The product was obtained in the distillated fraction when the temperature was 175–180 °C, and it was solidified after cooling, giving a yellow solid [68].

3.3. General Procedure for the Synthesis of Piv-l-Pro

A solution of l-proline (10 g, 86.86 mmol) in 2 M NaOH (50 mL) was cooled to 0 °C in an ice-water bath and stirred magnetically. Pivaloyl chloride (10.7 mL, 86.86 mmol) and 2 M NaOH (40 mL) were added in several alternating portions over the course of 1 h. The reaction mixture was stirred at 0 °C during this time, and the pH was checked periodically in order to confirm that the solution remained strongly alkaline. After addition was finished, the solution was stirred at room temperature for 1 h and extracted with CH2Cl2. Then, the aqueous phase was acidified to pH 1–2 with 6 M HCl and extracted with CH2Cl2. The organic phase was dried over anhydrous MgSO4, filtered, and concentrated in vacuo to obtain the product as a white solid. The product was used without further purification [69].

3.4. General Procedure for the l-Proline-Organocatalysts Synthesis (316)

Ethyl chloroformate (1 equiv.) was added to a solution of protected-l-proline (1 equiv.) and triethylamine (1 equiv.) in dry THF at 0 °C under nitrogen atmosphere. The reaction mixture was stirred at 0 °C for 30 min. Then, a solution of amine (1 equiv.) in dry THF was added, and the reaction mixture was stirred at 0 °C for 3 h, and afterwards, at room temperature overnight. The mixture was filtered, and the volatiles were removed under reduced pressure. The residual reagents were removed with acid-base extraction, and the product was purified by crystallization or column chromatography [53].

3.5. General Procedure for the Asymmetric Reduction with HSiCl3

To a stirred solution of imine 1 (0.1 g, 0.512 mmol) and the corresponding catalyst (1–40 mol%, depending of the study) in dry CH2Cl2 (1 mL) at 0 °C and under nitrogen atmosphere, trichlorosilane (77 μL, 0.768 mmol) was added. The reaction mixture was stirred at 0 °C for 16 h. After, saturated NaHCO3 was added, and the product was extracted with CH2Cl2. The organic phase was washed with brine, dried over MgSO4, and concentrated in vacuo. Yield and enantiomeric excess were determined using the crude product. In all reactions of this work, (S)-enantiomer was the one obtained in excess [52].

3.6. General Procedure for Kinetic Studies

A solution of 0.66 M imine 1 and 0.066 M corresponding catalyst in CDCl3 (dried over anhydrous MgSO4) was prepared under nitrogen atmosphere in an NMR tube with septum. The solution was cooled at 0 °C, and trichlorosilane (1.5 equiv.) was added. 1H-NMR spectra were recorded at 30 °C for different times (5 min, 0.5 h, 1 h, 2 h, 3 h, 4 h, 6 h, 8 h, and 24 h). During the reaction time, the NMR tube was cooled at 0 °C under nitrogen atmosphere. When the kinetic study was finished, saturated NaHCO3 was added, and the product was extracted with CH2Cl2. The organic phase was washed with brine, dried over MgSO4, and concentrated in vacuo. Yield and enantiomeric excess were determined using the crude product.

3.7. General Procedure for the Asymmetric Reduction with HSiCl3 in Flow Process

Two lines were connected to a PFA tubular flow reactor (length reactor: 3 m, internal diameter: 0.02 inch, residence time: 2.9 h). In the first line, a solution of 0.70 M imine 1 and 0.07 M catalyst 16 in dry CH2Cl2 under nitrogen atmosphere was pumped (with a syringe pump) into the reactor with a flow rate of 5 μL/min. In the second line, a solution of 2.63 M HSiCl3 in dry CH2Cl2 under nitrogen atmosphere was pumped with a flow rate of 2 μL/min. After the system was stabilized at 0 °C in a cryocool, various samples of outgoing solution were collected over saturated NaHCO3, using an automatic fraction collector. The sample collection was every 71 min, and the process was carried out for more than 11 h. When the fraction collection was finished, every sample was extracted with CH2Cl2, washed with brine, dried over MgSO4, and concentrated in vacuo. Yield and enantiomeric excess were determined using the crude product.

4. Conclusions

A series of experimental and computational studies have been combined to shed light on the mechanism of the HSiCl3 reduction of imine 1 catalyzed by N-protected prolinamides to facilitate catalyst optimization. It is important to note that in this process, the uncatalyzed process is always present, and achieving a process in which the catalyzed rate is much higher than the uncatalyzed one is essential to obtain an efficient and enantioselective catalytic system. Besides, the step for RIC formation is also an important factor, as it requires extensive hydrogen bond breaking and costly conformational changes, leading to the formation of weak complexes most likely at low concentrations that are sensitive to the structure of the catalyst. Regarding the structure of the prolinamides, two main structural elements can be used for the optimization of the catalytic activity: the synthesis of prolinamines from aromatic amines containing donor groups in the aromatic ring and the use of Piv instead of Cbz N-protection, most likely because only one intramolecularly hydrogen bonded conformer is possible in this case. From the systems studied, catalyst 16 caused the catalyzed reaction to be 581 times faster than the uncatalyzed reaction and allowed us to carry out the model reaction with 86% ee using a 1 mol% catalyst, obtaining good results even for low reaction times. This catalyst allowed the enantioselective reduction to be carried out under continuous flow conditions.

Supplementary Materials

Additional information of this research is included in Figures S1–S15 and Tables S1–S10, which are available online. Kinetic fitting data, computational calculations, synthetic procedures, and characterization are also included.

Author Contributions

Conceptualization, E.G.-V. and S.V.L.; methodology, R.P., E.G.-V. and S.V.L.; validation, M.M. and V.M.-C.; formal analysis, V.M.-C.; investigation, M.M. and R.P.; resources, M.I.B. and S.V.L.; data curation, M.M. and V.M.-C.; writing—original draft preparation, M.M. and V.M.-C.; writing—review and editing, E.G.-V. and S.V.L.; visualization, M.M. and V.M.-C.; supervision, E.G.-V. and S.V.L.; project administration, M.I.B.; funding acquisition, E.G.-V., M.I.B. and S.V.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Generalitat Valenciana, grant number PROMETEO/2016/071 and AICO/2021/139, Pla de Promoció de la Investigació de la Universitat Jaume I, grant number UJI-B2019-40 and Ministerio de Ciencia, Innovación y Universidades, grant number RTI2018-098233-B-C22. V.M.-C. acknowledges the financial support from Generalitat Valenciana (CIDEGENT/2020/031).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and Supplementary Materials.

Acknowledgments

The authors are very grateful to the ‘Servei Central d’Instrumentació Científica (SCIC)’ of the Universitat Jaume I.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Nugent, T.C. Chiral Amine Synthesis Methods, Developments and Applications; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar] [CrossRef]
  2. Nugent, T.C.; El-Shazly, M. Chiral Amine Synthesis—Recent Developments and Trends for Enamide Reduction, Reductive Amination, and Imine Reduction. Adv. Synth. Catal. 2010, 352, 753–819. [Google Scholar] [CrossRef]
  3. Ghislieri, D.; Turner, N.J. Biocatalytic Approaches to the Synthesis of Enantiomerically Pure Chiral Amines. Top. Catal. 2014, 57, 284–300. [Google Scholar] [CrossRef]
  4. Phillips, A.M.F.; Pombeiro, A.J.L. Recent advances in organocatalytic enantioselective transfer hydrogenation. Org. Biomol. Chem. 2017, 15, 2307–2340. [Google Scholar] [CrossRef]
  5. Herrera, R.P. Organocatalytic Transfer Hydrogenation and Hydrosilylation Reactions. Top. Curr. Chem. 2016, 374, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Rossi, S.; Benaglia, M.; Massolo, E.; Raimondi, L. Organocatalytic strategies for enantioselective metal-free reductions. Catal. Sci. Technol. 2014, 4, 2708–2723. [Google Scholar] [CrossRef] [Green Version]
  7. Foubelo, F.; Yus, M. Catalytic Asymmetric Transfer Hydrogenation of Imines: Recent Advances. Chem. Rec. 2015, 15, 907–924. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Wakchaure, V.N.; Kaib, P.S.J.; Leutzsch, M.; List, B. Disulfonimide-Catalyzed Asymmetric Reduction of N-Alkyl Imines. Angew. Chem. Int. Ed. 2015, 54, 11852–11856. [Google Scholar] [CrossRef]
  9. Genoni, A.; Benaglia, M.; Mattiolo, E.; Rossi, S.; Raimondi, L.; Barrulas, P.C.; Burke, A.J. Synthesis of an advanced precursor of Rivastigmine: Cinchona-derived quaternary ammonium salts as organocatalysts for stereoselective imine reductions. Tetrahedron Lett. 2015, 56, 5752–5756. [Google Scholar] [CrossRef]
  10. Genoni, A.; Benaglia, M.; Massolo, E.; Rossi, S. Stereoselective metal-free catalytic synthesis of chiral trifluoromethyl aryl and alkyl amines. Chem. Commun. 2013, 49, 8365–8367. [Google Scholar] [CrossRef] [PubMed]
  11. Yilmaz, N.; Sekeroglu, B. Student Performance Classification Using Artificial Intelligence Techniques. In Advances in Intelligent Systems and Computing; Springer: Cham, Switzerland, 2020; Volume 1095. [Google Scholar] [CrossRef]
  12. Brenna, D.; Pirola, M.; Raimondi, L.; Burke, A.J.; Benaglia, M. A stereoselective, catalytic strategy for the in-flow synthesis of advanced precursors of rasagiline and tamsulosin. Bioorg. Med. Chem. 2017, 25, 6242–6247. [Google Scholar] [CrossRef]
  13. Han, Z.S.; Zhang, L.; Xu, Y.; Sieber, J.D.; Marsini, M.A.; Li, Z.; Reeves, J.T.; Fandrick, K.R.; Patel, N.D.; Desrosiers, J.-N.; et al. Efficient Asymmetric Synthesis of Structurally Diverse P-Stereogenic Phosphinamides for Catalyst Design. Angew. Chem. Int. Ed. 2015, 54, 5474–5477. [Google Scholar] [CrossRef] [PubMed]
  14. Sugiura, M.; Ashikari, Y.; Takahashi, Y.; Yamaguchi, K.; Kotani, S.; Nakajima, M. Lewis Base-Catalyzed Enantioselective Conjugate Reduction of β,β-Disubstituted α,β-Unsaturated Ketones with Trichlorosilane: E/Z-Isomerization, Regioselectivity, and Synthetic Applications. J. Org. Chem. 2019, 84, 11458–11473. [Google Scholar] [CrossRef] [PubMed]
  15. Orlandi, M.; Tosi, F.; Bonsignore, M.; Benaglia, M. Metal-Free Reduction of Aromatic and Aliphatic Nitro Compounds to Amines: A HSiCl3-Mediated Reaction of Wide General Applicability. Org. Lett. 2015, 17, 3941–3943. [Google Scholar] [CrossRef]
  16. Orlandi, M.; Benaglia, M.; Tosi, F.; Annunziata, R.; Cozzi, F. HSiCl3-Mediated Reduction of Nitro-Derivatives to Amines: Is Tertiary Amine-Stabilized SiCl2 the Actual Reducing Species? J. Org. Chem. 2016, 81, 3037–3041. [Google Scholar] [CrossRef] [PubMed]
  17. Chen, L.; Wang, C.; Zhou, L.; Sun, J. Chiral 2,3-Disubstituted Indolines from Indoles and Aldehydes by Organocatalyzed Tandem Synthesis Involving Reduction by Trichlorosilane. Adv. Synth. Catal. 2014, 356, 2224–2230. [Google Scholar] [CrossRef]
  18. Fu, Y.; Sun, J. HMPA-Catalyzed Transfer Hydrogenation of 3-Carbonyl Pyridines and Other N-Heteroarenes with Trichlorosilane. Molecules 2019, 24, 401. [Google Scholar] [CrossRef] [Green Version]
  19. Wang, T.; Di, X.; Wang, C.; Zhou, L.; Sun, J. Reductive Hydrazination with Trichlorosilane: A Method for the Preparation of 1,1-Disubstituted Hydrazines. Org. Lett. 2016, 18, 1900–1903. [Google Scholar] [CrossRef]
  20. Frogneux, X.; Blondiaux, E.; Thuéry, P.; Cantat, T. Bridging Amines with CO2: Organocatalyzed Reduction of CO2 to Aminals. ACS Catal. 2015, 5, 3983–3987. [Google Scholar] [CrossRef]
  21. Hao, L.; Zhao, Y.; Yu, B.; Yang, Z.; Zhang, H.; Han, B.; Gao, X.; Liu, Z. Imidazolium-Based Ionic Liquids Catalyzed Formylation of Amines Using Carbon Dioxide and Phenylsilane at Room Temperature. ACS Catal. 2015, 5, 4989–4993. [Google Scholar] [CrossRef]
  22. Nicholls, R.L.; McManus, J.A.; Rayner, C.M.; Morales-Serna, J.A.; White, A.J.P.; Nguyen, B.N. Guanidine-Catalyzed Reductive Amination of Carbon Dioxide with Silanes: Switching between Pathways and Suppressing Catalyst Deactivation. ACS Catal. 2018, 8, 3678–3687. [Google Scholar] [CrossRef]
  23. Hulla, M.; Laurenczy, G.; Dyson, P.J. Mechanistic Study of the N-Formylation of Amines with Carbon Dioxide and Hydrosilanes. ACS Catal. 2018, 8, 10619–10630. [Google Scholar] [CrossRef] [Green Version]
  24. Guizzetti, S.; Benaglia, M. Trichlorosilane-Mediated Stereoselective Reduction of C=N Bonds. Eur. J. Org. Chem. 2010, 29, 5529–5541. [Google Scholar] [CrossRef]
  25. Jones, S.; Warner, C.J.A. Trichlorosilane mediated asymmetric reductions of the C=N bond. Org. Biomol. Chem. 2012, 10, 2189–2200. [Google Scholar] [CrossRef] [PubMed]
  26. Qinglong, S.; Zhangpei, C.; Jianshe, H. Recent Advances in Organocatalyzed Asymmetric Hydrosilylations. Curr. Org. Chem. 2018, 22, 557–580. [Google Scholar] [CrossRef]
  27. Malkov, A.V.; Stewart-Liddon, A.J.P.; McGeoch, G.D.; Ramírez-López, P.; Kočovský, P. Catalyst development for organocatalytic hydrosilylation of aromatic ketones and ketimines. Org. Biomol. Chem. 2012, 10, 4864–4877. [Google Scholar] [CrossRef] [Green Version]
  28. Pan, W.; Deng, Y.; He, J.-B.; Bai, B.; Zhu, H.-J. Highly efficient asymmetric-axle-supported N-O amides in enantioselective hydrosilylation of ketimines with trichlorosilane. Tetrahedron 2013, 69, 7253–7257. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, Z.; Wang, C.; Zhoub, L.; Sun, J. L-Pipecolinic acid derived Lewis base organocatalyst for asymmetric reduction of N-aryl imines by trichlorosilane: Effects of the side amide group on catalytic performances. Org. Biomol. Chem. 2013, 11, 787–797. [Google Scholar] [CrossRef]
  30. Pei, Y.-N.; Deng, Y.; Li, J.-L.; Liu, L.; Zhu, H.-J. New chiral biscarboline N,N’-dioxide derivatives as catalyst in enantioselective reduction of ketoimines with trichlorosilane. Tetrahedron Lett. 2014, 55, 2948–2952. [Google Scholar] [CrossRef]
  31. Brenna, D.; Porta, R.; Massolo, E.; Raimondi, L.; Benaglia, M. A New Class of Low-Loading Catalysts for Highly Enantioselective, Metal-Free Imine Reduction of Wide General Applicability. ChemCatChem 2017, 9, 941–945. [Google Scholar] [CrossRef]
  32. Skrypai, V.; Varjosaari, S.E.; Azam, F.; Gilbert, T.M.; Adler, M.J. Chiral Brønsted Acid-Catalyzed Metal-Free Asymmetric Direct Reductive Amination Using 1-Hydrosilatrane. J. Org. Chem. 2019, 84, 5021–5026. [Google Scholar] [CrossRef]
  33. Chen, W.; Tan, C.-H.; Wang, H.; Ye, X. The Development of Organocatalytic Asymmetric Reduction of Carbonyls and Imines Using Silicon Hydrides. Eur. J. Org. Chem. 2021, 21, 3091–3112. [Google Scholar] [CrossRef]
  34. Jones, S.; Li, X. Synthesis of chiral β-amino acid derivatives by asymmetric hydrosilylation with an imidazole derived organocatalyst. Tetrahedron 2012, 68, 5522–5532. [Google Scholar] [CrossRef]
  35. Jones, S.; Zhao, P. Evaluating dynamic kinetic resolution strategies in the asymmetric hydrosilylation of cyclic ketimines. Tetrahedron: Asymmetry 2014, 25, 238–244. [Google Scholar] [CrossRef]
  36. Ye, J.; Wang, C.; Chen, L.; Wu, X.; Zhou, L.; Sun, J. Chiral Lewis Base-Catalyzed, Enantioselective Reduction of Unprotected β-Enamino Esters with Trichlorosilane. Adv. Synth. Catal. 2016, 358, 1042–1047. [Google Scholar] [CrossRef]
  37. Dai, X.; Weng, G.; Yu, S.; Chen, H.; Zhang, J.; Cheng, S.; Xu, X.; Yuan, W.; Wang, Z.; Zhang, X. One-pot diastereo- and enantioselective hydrosilylation–transacylation of α-acyloxy β-enamino esters. Org. Chem. Front. 2018, 5, 2787–2793. [Google Scholar] [CrossRef]
  38. Sugiura, M.; Kashiwagi, T.; Ito, M.; Kotani, S.; Nakajima, M. Stereoselective Synthesis of Nitrogen-Containing Compounds from Enamines. J. Org. Chem. 2017, 82, 10968–10979. [Google Scholar] [CrossRef] [PubMed]
  39. Dai, X.; Liu, M.; Huang, M.; Zhang, J.; Xu, X.; Yuan, W.; Zhang, X. Synthesis of Chiral α-Mercapto-β-acylamido Esters via One-pot Asymmetric Hydrosilylation-transacylation of α-Acylthio-β-Enamino Esters. Asian J. Org. Chem. 2019, 8, 456–461. [Google Scholar] [CrossRef]
  40. Zhang, Z.; Rooshenas, P.; Hausmann, H.; Schreiner, P.R. Asymmetric Transfer Hydrogenation of Ketimines with Trichlorosilane: Structural Studies. Synthesis 2009, 2009, 1531–1544. [Google Scholar] [CrossRef] [Green Version]
  41. Li, X.; Reeder, A.T.; Torri, F.; Adams, H.; Jones, S. Mechanistic investigations of the asymmetric hydrosilylation of ketimines with trichlorosilane reveals a dual activation model and an organocatalyst with enhanced efficiency. Org. Biomol. Chem. 2017, 15, 2422–2435. [Google Scholar] [CrossRef] [Green Version]
  42. Bonsignore, M.; Benaglia, M.; Raimondi, L.; Orlandi, M.; Celentano, G. Enantioselective reduction of ketoimines promoted by easily available (S)-proline derivatives. Beilstein J. Org. Chem. 2013, 9, 633–640. [Google Scholar] [CrossRef] [Green Version]
  43. Dong, M.; Wang, J.; Wu, S.; Zhao, Y.; Ma, Y.; Xing, Y.; Cao, F.; Li, L.; Li, Z.; Zhu, H. Catalytic Mechanism Study on the 1,2- and 1,4-Transfer Hydrogenation of Ketimines and β-Enamino Esters Catalyzed by Axially Chiral Biscarboline-Based Alcohols. Adv. Synth. Catal. 2019, 361, 4602–4610. [Google Scholar] [CrossRef]
  44. Shaikh, I.R. Organocatalysis: Key Trends in Green Synthetic Chemistry, Challenges, Scope Towards Heterogenization, and Importance from Research and Industrial Point of View. J. Catal. 2014, 2014, 1–35. [Google Scholar] [CrossRef] [Green Version]
  45. Iwasaki, F.; Onomura, O.; Mishima, K.; Kanematsu, T.; Maki, T.; Matsumura, Y. First chemo- and stereoselective reduction of imines using trichlorosilane activated with N-formylpyrrolidine derivatives. Tetrahedron Lett. 2001, 42, 2525–2527. [Google Scholar] [CrossRef]
  46. Ever, Y.; Dimililer, K.; Sekeroglu, B. Comparison of Machine Learning Techniques for Prediction Problems. In Advances in Intelligent Systems and Computing; Springer: Cham, Switzerland, 2019; Volume 927. [Google Scholar] [CrossRef]
  47. Ge, X.; Qian, C.; Chen, X. Synthesis of novel carbohydrate-based valine-derived formamide organocatalysts by CuAAC click chemistry and their application in asymmetric reduction of imines with trichlorosilane. Tetrahedron Asymmetry 2014, 25, 1450–1455. [Google Scholar] [CrossRef]
  48. Ge, X.; Qian, C.; Yea, X.; Chen, X. Asymmetric reduction of imines with trichlorosilane catalyzed by valine-derived formamide immobilized onto magnetic nano-Fe3O4. RSC Adv. 2015, 5, 65402–65407. [Google Scholar] [CrossRef]
  49. Wagner, C.; Kotthaus, A.F.; Kirsch, S.F. The asymmetric reduction of imidazolinones with trichlorosilane. Chem. Commun. 2017, 53, 4513–4516. [Google Scholar] [CrossRef] [PubMed]
  50. Pedrosa, R.; Andres, J.M. Prolinamides as Asymmetric Organocatalysts. In Sustainable Catalysis: Without Metals or Other Endangered Elements; part 1; North, M., Ed.; RSC: London, UK, 2016; pp. 120–138. [Google Scholar]
  51. Becerril, J.; Bolte, M.; Burguete, M.I.; Galindo, F.; Garcia-España, E.; Luis, S.V.; Miravet, J.F. Efficient Macrocyclization of U-Turn Preorganized Peptidomimetics:  The Role of Intramolecular H-Bond and Solvophobic Effects. J. Am. Chem. Soc. 2003, 125, 6677–6686. [Google Scholar] [CrossRef] [PubMed]
  52. Burguete, M.I.; Collado, M.; Escorihuela, J.; Luis, S.V. Efficient Chirality Switching in the Addition of Diethyzinc to Aldehydes in the Presence of Simple Chiral β-Amino Amides. Angew. Chem. Int. Ed. 2007, 46, 9002–9005. [Google Scholar] [CrossRef]
  53. Obregón-Zúñiga, A.; Juaristi, E. (2S,4R)-Hyp-(S)-Phe-OMe dipeptide supported on imidazolium tagged molecules as recoverable organocatalysts for asymmetric aldol reactions using water as reaction medium. Tetrahedron 2017, 73, 5373–5380. [Google Scholar] [CrossRef]
  54. Kanemitsu, T.; Umehara, A.; Haneji, R.; Nagata, K.; Itoh, T. A simple proline-based organocatalyst for the enantioselective reduction of imines using trichlorosilane as a reductant. Tetrahedron 2012, 68, 3893–3898. [Google Scholar] [CrossRef]
  55. Barrulas, P.C.; Genoni, A.; Benaglia, M.; Burke, A.J. Cinchona-Derived Picolinamides: Effective Organocatalysts for Stereoselective Imine Hydrosilylation. Eur. J. Org. Chem. 2014, 2014, 7339–7342. [Google Scholar] [CrossRef] [Green Version]
  56. Hu, F.-Z.; Chen, H.; Xu, X.-Y.; Yuan, W.-C.; Zhang, X.-M. Dynamic Kinetic Resolution in Enantioselective Reductive Amination of a-Branched Aldehydes by Lewis Base Organocatalyzed Hydrosilylation. ChemistrySelect 2017, 2, 4076–4078. [Google Scholar] [CrossRef]
  57. Xing, Y.; Wu, S.; Dong, M.; Wang, J.; Liu, L.; Zhu, H. Synthesis and application of axially chiral biscarbolines with functional N-O and sulfone for 1,2-transfer hydrogenations of ketimines. Tetrahedron 2019, 75, 130495. [Google Scholar] [CrossRef]
  58. Hu, X.-Y.; Zhang, M.-M.; Shu, C.; Zhang, Y.-H.; Liao, L.-H.; Yuan, W.-C.; Zhang, X.-M. Enantioselective Lewis-Base-Catalyzed Asymmetric Hydrosilylation of Substituted Benzophenone N-Aryl Imines: Efficient Synthesis of Chiral (Diarylmethyl)amines. Adv. Synth. Catal. 2014, 356, 3539–3544. [Google Scholar] [CrossRef]
  59. Warner, C.J.A.; Reeder, A.T.; Jones, S. P-Chiral phosphine oxide catalysed reduction of prochiral ketimines using trichlorosilane. Tetrahedron: Asymmetry 2016, 27, 136–141. [Google Scholar] [CrossRef] [Green Version]
  60. El-Fayyoumy, S.; Todd, M.H.; Richards, C.J. Can we measure catalyst efficiency in asymmetric chemical reactions? A theoretical approach. Beilstein J. Org. Chem. 2009, 5, 67. [Google Scholar] [CrossRef]
  61. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09, Revision, B.01; Gaussian, Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  62. Cortez, P.; Silva, A. Role of noncovalent interactions in the enantioselective reduction of aromatic ketimines with trichlorosilane. In Proceedings of the 5th FUture BUsiness TEChnology Conference (FUBUTEC), Porto, Portugal, 9–11 April 2008; pp. 5–12, ISBN 978-9077381-39-7. [Google Scholar]
  63. Onomura, O.; Kouchi, Y.; Iwasaki, F.; Matsumura, Y. New organic activators for the enantioselective reduction of aromatic imines with trichlorosilane. Tetrahedron Lett. 2006, 47, 3751–3754. [Google Scholar] [CrossRef] [Green Version]
  64. Reep, C.; Morgante, P.; Peverati, R.; Takenaka, N. Axial-Chiral Biisoquinoline N,N′-Dioxides Bearing Polar Aromatic C-H Bonds as Catalysts in Sakurai-Hosomi-Denmark Allylation. Org. Lett. 2018, 20, 5757–5761. [Google Scholar] [CrossRef] [PubMed]
  65. Warner, C.J.A.; Berry, S.S.; Jones, S. Evaluation of bifunctional chiral phosphine oxide catalysts for the asymmetric hydrosilylation of ketimines. Tetrahedron 2019, 75, 130733. [Google Scholar] [CrossRef]
  66. Entrena, A.; Campos, J.M.; Gallo, M.A.; Espinosa, A. Rules for Predicting the Conformational Behavior of Saturated Seven-Membered Heterocycles. ARKIVOC 2005, 6, 88–108. [Google Scholar] [CrossRef] [Green Version]
  67. Gawley, R.E. Do the Terms “% ee” and “% de” Make Sense as Expressions of Stereoisomer Composition or Stereoselectivity? J. Org. Chem. 2006, 71, 2411–2416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Gautier, F.-M.; Jones, S.; Martin, S.J. Asymmetric reduction of ketimines with trichlorosilane employing an imidazole derived organocatalyst. Org. Biomol. Chem. 2009, 7, 229–231. [Google Scholar] [CrossRef] [PubMed]
  69. Pirkle, W.H.; Murray, P.G.; Rausch, D.J.; McKenna, S.T. Intermolecular 1H–1H Two-Dimensional Nuclear Overhauser Enhancements in the Characterization of a Rationally Designed Chiral Recognition System. J. Org. Chem. 1996, 61, 4769–4774. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Benchmark asymmetric reduction of ketimine 1 used to develop the organocatalysts in this work.
Scheme 1. Benchmark asymmetric reduction of ketimine 1 used to develop the organocatalysts in this work.
Molecules 26 06963 sch001
Figure 1. Catalysts used for the reduction of the ketimine 1 in this study.
Figure 1. Catalysts used for the reduction of the ketimine 1 in this study.
Molecules 26 06963 g001
Figure 2. Influence of catalyst loading on enantioselectivity for the reduction of ketimine 1 (0.512 M) with HSiCl3 (1.5 equiv.) in CH2Cl2 at 0 °C for 16 h.
Figure 2. Influence of catalyst loading on enantioselectivity for the reduction of ketimine 1 (0.512 M) with HSiCl3 (1.5 equiv.) in CH2Cl2 at 0 °C for 16 h.
Molecules 26 06963 g002
Figure 3. Variation of the yield as a function of time in the asymmetric reduction. See Supplementary Materials (pages S22–S25); a Concentration 1 = 0.66 M, 1.5 equiv. HSiCl3, CDCl3, 0 °C; b Concentration 1 = 0.66 M, [cat.] = 0.066 M, 1.5 equiv. HSiCl3, CDCl3, 0 °C.
Figure 3. Variation of the yield as a function of time in the asymmetric reduction. See Supplementary Materials (pages S22–S25); a Concentration 1 = 0.66 M, 1.5 equiv. HSiCl3, CDCl3, 0 °C; b Concentration 1 = 0.66 M, [cat.] = 0.066 M, 1.5 equiv. HSiCl3, CDCl3, 0 °C.
Molecules 26 06963 g003
Scheme 2. Kinetic model for uncatalyzed and catalyzed reactions.
Scheme 2. Kinetic model for uncatalyzed and catalyzed reactions.
Molecules 26 06963 sch002
Figure 4. Variation of enantioselectivity and ACE as a function of the molar percentage of catalysts 12, 13, and 14. Reaction conditions: Concentration 1 = 0.512 M, 1.5 equiv. HSiCl3, CH2Cl2, 0 °C, 16 h.
Figure 4. Variation of enantioselectivity and ACE as a function of the molar percentage of catalysts 12, 13, and 14. Reaction conditions: Concentration 1 = 0.512 M, 1.5 equiv. HSiCl3, CH2Cl2, 0 °C, 16 h.
Molecules 26 06963 g004
Figure 5. Comparison of kinetic curves for catalysts 13 and 16 with different N-protecting groups. See Supplementary Materials (pages S27 and S29). a Concentration 1 = 0.66 M, 1.5 equiv. HSiCl3, 0 °C; b Concentration 1 = 0.66 M, (catalyst) = 0.066 M, 1.5 equiv. HSiCl3, 0 °C.
Figure 5. Comparison of kinetic curves for catalysts 13 and 16 with different N-protecting groups. See Supplementary Materials (pages S27 and S29). a Concentration 1 = 0.66 M, 1.5 equiv. HSiCl3, 0 °C; b Concentration 1 = 0.66 M, (catalyst) = 0.066 M, 1.5 equiv. HSiCl3, 0 °C.
Molecules 26 06963 g005
Figure 6. Uncatalyzed reduction of an imine.
Figure 6. Uncatalyzed reduction of an imine.
Molecules 26 06963 g006
Figure 7. (A) Calculated structures and energies for the catalyst –HSiCl3 complex. (B) Reaction scheme from structures C and D. (C) Optimized geometry for the TS calculated from structure C.
Figure 7. (A) Calculated structures and energies for the catalyst –HSiCl3 complex. (B) Reaction scheme from structures C and D. (C) Optimized geometry for the TS calculated from structure C.
Molecules 26 06963 g007
Figure 8. Calculated energy profile for the simplified model reaction.
Figure 8. Calculated energy profile for the simplified model reaction.
Molecules 26 06963 g008
Figure 9. Reaction mechanism considered for the hydrogenation of ketimine 1 with the corresponding N-carbamate-amide in the presence of HSiCl3.
Figure 9. Reaction mechanism considered for the hydrogenation of ketimine 1 with the corresponding N-carbamate-amide in the presence of HSiCl3.
Molecules 26 06963 g009
Figure 10. Energy profile calculated for the enantioselective reduction of imine 1 using catalyst 16, obtaining the S-amine 2 as the product.
Figure 10. Energy profile calculated for the enantioselective reduction of imine 1 using catalyst 16, obtaining the S-amine 2 as the product.
Molecules 26 06963 g010
Figure 11. Representation of ln k2 (experimental) versus ΔETOT (theoretical), and linear fitting (not including catalyst 14) for R and S enantiomers in the up conformation.
Figure 11. Representation of ln k2 (experimental) versus ΔETOT (theoretical), and linear fitting (not including catalyst 14) for R and S enantiomers in the up conformation.
Molecules 26 06963 g011
Table 1. Enantioselective reduction of ketimine 1 catalyzed by 40 mol% of organocatalyts 39 a.
Table 1. Enantioselective reduction of ketimine 1 catalyzed by 40 mol% of organocatalyts 39 a.
EntryCatalystConv. b (%)Yield b (%)eec (%)
1 d--230
23919185
34988385
4599950
56999481
67989782
78999988
8993910
a Reaction conditions: Concentration 1 = 0.512 M, 1.5 equiv. HSiCl3, CH2Cl2, 0 °C, 16 h. b Conversions and yields determined by 1H-NMR on the crude reaction mixture. c Enantiomeric excess determined by chiral HPLC (S configuration for the major enantiomer). d See reference [46]. (Reaction conditions: Concentration 1 = 0.2 M, 3 equiv. HSiCl3, CH2Cl2, room temperature, 20 h).
Table 2. Experimental rate constants obtained for catalysts 3, 6, and 8 using Equations (1) and (2).
Table 2. Experimental rate constants obtained for catalysts 3, 6, and 8 using Equations (1) and (2).
EntryCatalystk
(M−1 h−1)
k2
(M−2 h−1)
k2/k1
(M−1)
1- a0.00543 c --
23b0.040000.523896.5
36b0.011240.088016.2
48b0.036470.470386.6
a Reduction of ketimine 1 at Concentration 1 = 0.66 M, 1.5 equiv. HSiCl3, CDCl3, 0 °C, 24 h. b Reduction of ketimine 1 at concentration 1 = 0.66 M, [catalyst] = 0.066 M, 1.5 equiv. HSiCl3, CDCl3, 0 °C, 24 h. c The observed kinetic constant corresponds to k1 in this case.
Table 3. Experimental rate constants obtained for catalysts 1214 using Equations (1) and (2) a.
Table 3. Experimental rate constants obtained for catalysts 1214 using Equations (1) and (2) a.
EntryCat.k
(M−1 h−1)
k2
(M−2 h−1)
k2/k1
(M−1)
eeexpb (%)
1120.026170.314257.877
2130.043260.5732105.579
3140.005910.00731.310
a Reduction of ketimine 1 in CDCl3 at concentration 1 = 0.66 M, catalyst concentration = 0.066 M, 1.5 equiv. HSiCl3, 0 °C, 24 h. b Enantiomeric excess determined by chiral HPLC (S configuration for the major enantiomer).
Table 4. Enantioselective reduction of ketimine 1 with HSiCl3 catalyzed by 15 and 16 a.
Table 4. Enantioselective reduction of ketimine 1 with HSiCl3 catalyzed by 15 and 16 a.
EntryCat.mol% Cat. (%)[Cat] (M)Conv. b (%)Yield b (%)eec (%)ACE
11510.00599998147.87
21550.0269999829.69
315400.2059999821.21
41610.00596888649.05
51650.0261001008711.28
616400.2059999861.38
a Reaction conditions: 1.5 equiv. HSiCl3, CH2Cl2, 0 °C, 16 h. b Conversions and yields determined by 1H-NMR on the crude reaction mixture. c Enantiomeric excess determined by chiral HPLC (S configuration for the major enantiomer).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Maciá, M.; Porcar, R.; Martí-Centelles, V.; García-Verdugo, E.; Burguete, M.I.; Luis, S.V. Rational Design of Simple Organocatalysts for the HSiCl3 Enantioselective Reduction of (E)-N-(1-Phenylethylidene)aniline. Molecules 2021, 26, 6963. https://doi.org/10.3390/molecules26226963

AMA Style

Maciá M, Porcar R, Martí-Centelles V, García-Verdugo E, Burguete MI, Luis SV. Rational Design of Simple Organocatalysts for the HSiCl3 Enantioselective Reduction of (E)-N-(1-Phenylethylidene)aniline. Molecules. 2021; 26(22):6963. https://doi.org/10.3390/molecules26226963

Chicago/Turabian Style

Maciá, María, Raúl Porcar, Vicente Martí-Centelles, Eduardo García-Verdugo, Maria Isabel Burguete, and Santiago V. Luis. 2021. "Rational Design of Simple Organocatalysts for the HSiCl3 Enantioselective Reduction of (E)-N-(1-Phenylethylidene)aniline" Molecules 26, no. 22: 6963. https://doi.org/10.3390/molecules26226963

Article Metrics

Back to TopTop