ACS Publications. Most Trusted. Most Cited. Most Read
Origin of Stereoselectivity in FLP-Catalyzed Asymmetric Hydrogenation of Imines
My Activity
  • Open Access
Research Article

Origin of Stereoselectivity in FLP-Catalyzed Asymmetric Hydrogenation of Imines
Click to copy article linkArticle link copied!

Open PDFSupporting Information (1)

ACS Catalysis

Cite this: ACS Catal. 2020, 10, 23, 14290–14301
Click to copy citationCitation copied!
https://doi.org/10.1021/acscatal.0c04263
Published November 23, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Abstract

Click to copy section linkSection link copied!

Development of metal-free strategies for stereoselective hydrogenation of unsaturated substrates is of particular interest in asymmetric synthesis. The emerging chemistry of frustrated Lewis pairs offers a promising approach along this line as demonstrated by recent achievements. However, the stereocontrol elements in these reactions are not clearly recognized thus far. Herein, we analyze the origin of stereoinduction in direct hydrogenation of imines catalyzed by a set of chiral boranes. We use the tools of computational chemistry to describe the elementary steps of the catalytic cycle, and we pay special attention to the stereoselectivity-determining hydride transfer process. The enantioselectivities predicted by the applied computational approach are in very good agreement with previous experimental observations. We find that the stereoselectivity is governed by a thermodynamically less favored conformer of the borohydride intermediate and not by the experimentally observed form. The most favored hydride transfer transition states are stabilized by specific aryl–aryl and alkyl–aryl noncovalent interactions, which play an important role in stereoinduction. This computational insight is exploited in proposing additional borane variants to improve the enantioselectivity, which could be demonstrated experimentally

Copyright © 2020 American Chemical Society

Introduction

Click to copy section linkSection link copied!

Chiral amines, particularly those bearing α-stereogenic carbon atoms, are widely used intermediates in the production of pharmaceuticals, agrochemicals, and fine chemicals. Therefore, the development of effective methods for enantioselective synthesis of these compounds is of high interest. (1) Among the available synthetic strategies, the asymmetric hydrogenation of prochiral imines and enamines by the direct use of H2 as the hydrogen source represents a promising atom-economic approach. (2) A variety of transition metal (TM) complexes comprising chiral organic ligands have been developed, and some of them have proved to be efficient catalysts for direct enantioselective imine hydrogenations. Iridium complexes are known to be particularly powerful hydrogenation catalysts; however, achieving very high enantioselectivities for a wide range of substrates remains challenging. (3)
Metal-free routes to the synthesis of chiral amines via catalytic hydrogenation have also been developed in the past decade. (4) Several successful organocatalytic transfer hydrogenation reactions catalyzed by chiral phosphoric acids as well as hydrosilylations mediated by chiral Lewis bases have been reported, but these transformations require a stoichiometric amount of other hydrogen sources, such as Hantzsch ester or trichlorosilane. (5) The discovery that sterically hindered Lewis acid–base pairs are able to cleave molecular hydrogen reversibly under mild reaction conditions (6) opened a metal-free strategy for direct catalytic hydrogenation of unsaturated molecules. (7) These so-called frustrated Lewis pairs (FLPs) (8) were shown to catalyze various unsaturated organic substrates (9) even under water-tolerant conditions. (10)
The FLP concept has been successfully adopted to design asymmetric catalytic hydrogenation processes as well. (11) Pioneering contributions from Klankermayer et al. (12) provided the first examples of chiral induction by FLPs. Boranes 14 (Chart 1) were prepared via hydroboration of chiral olefins with Piers’ borane (C6F5)2BH, (13) and they were employed for the hydrogenation of imines. The (+)-α-pinene derived borane 1 gave only a low enantiomeric excess (ee) in the product amine (13% ee); (12a) however, the asymmetric induction was notably improved (up to 83% ee) using boranes 24, which were derived from (1R)-(+)-camphor. (12b)

Chart 1

Chart 1. Selection of Chiral FLP Components Employed in Enantioselective Catalytic Hydrogenationsa

aAr denotes aromatic substituents; ArF = C6F5 or p-C6F4H.

Repo and co-workers developed intramolecular FLPs with chiral amine moieties (e.g., 5); however, they resulted in only moderate enantioselectivities in catalytic hydrogenation of imines and quinolines (up to 37% ee). (14) The same group later introduced chirality via the binaphthyl framework into ansa-aminoboranes, and intramolecular FLP 6 was shown to be an efficient hydrogenation catalyst, particularly for asymmetric hydrogenation of enamines (ee values up to 99%). (15) In 2013, Du proposed a simple and powerful strategy to generate enantiomerically pure bis-boranes 7 in situ by direct hydroboration of binaphthyl-based terminal dienes. (16) Some representatives of this family of chiral boranes, especially those with very bulky Ar substituents on the binaphthyl framework, were demonstrated to be highly effective catalysts in enantioselective hydrogenation of imines, (16) silyl enol ethers, (17) and various heteroarenes (18) as well. Wang and co-workers have recently developed a series of C2-symmetric bicyclic bis-boranes derived from cis-fused bicyclic dienes by hydroboration. (19) Stereoisomer 8 of this series was found to be exceptionally efficient for imine hydrogenation providing ee’s above 90% for the first time. In subsequent work, spiro-bicyclic bis-boranes 9 were also introduced, (20) exhibiting excellent catalytic activities and stereoselectivities for the hydrogenation of quinolines and 2-vinylpyridines. Attempts to control the stereochemistry of FLP-type hydrogenation by chiral Lewis bases resulted in only modest enantioselectivities, (9b) until a very recent study reported by the Du group, (21) which established this approach as a promising direction for the design of new chiral FLPs. Quite remarkably, chiral oxazolines such as 10, combined with achiral boranes were shown to induce high degree of enantioselectivity in asymmetric hydrogenation of ketones and enones. In addition to direct catalytic hydrogenation, FLP-type asymmetric hydrosilylation has also been successfully applied to reduce various unsaturated substrates with high enantioselectivities. (22)
Despite these impressive advances, our knowledge regarding the stereocontrol elements in FLP-catalyzed asymmetric hydrogenation processes is quite scanty. So far, only a few computational studies have been reported that supplemented the synthetic developments and addressed the issue of stereoselectivity. (15,21,23) Density functional theory (DFT) calculations carried out for the highly selective enamine hydrogenation with catalyst 6 indicated that the energy difference obtained for the hydride transfer transition states leading to the two enantiomeric amine products stems from a combination of repulsive steric and attractive noncovalent interactions. (15) The stereoselectivity determining transition states for asymmetric hydrogenation of ketones catalyzed with the B(p-C6F4H)3/10 pair were also identified computationally. (21) The predicted energy differences were consistent with experimental observations, but the origin of stereoinduction was not examined.
To gain more insight into the nature of molecular interactions responsible for the stereoselectivity of FLP-catalyzed hydrogenation, in our present work we investigated a reaction that is frequently used as a test case in the evaluation of developed catalysts, namely, the direct hydrogenation of imine (E)-N-(1-phenylethylidene)-aniline (im, see Scheme 1). As shown previously, the enantioselectivity of this reaction varies in a fairly broad range (from poor to good) when structurally analogous boranes 13 are applied as catalysts; thus, this series of reactions allows us to inspect the effect of a single chiral borane substituent on the stereochemical outcome of the hydrogenation process.

Scheme 1

Scheme 1. Reactions Examined Computationally
In this work, we focus primarily on the stereoselectivity-determining hydride transfer step of these reactions; however, for borane 2, we provide a detailed analysis of the H2 activation process as well. We closely inspect the conformational space of the iminium borodyride species, which is a key intermediate prior to product formation. Accurate prediction of stereoselectivities from quantum chemical calculations is rather challenging. (24) In our present analysis, we attempted to pay respect to the critical issues (e.g., conformational complexity, the choice of the electronic structure method, estimation of entropic contributions, and solvent effects), and computed the enantioselectivity of the examined reactions accordingly (for computational details, see the Supporting Information (SI)). The good agreement obtained between the predicted and observed stereoselectivities allowed us to probe the origin of stereoinduction in these reactions. Based on the new computational insights, structural modifications in borane 2 were proposed, and two of these new chiral boranes were synthesized and tested as catalysts in asymmetric hydrogenation of imines.

Results

Click to copy section linkSection link copied!

Alternative Catalytic Cycles

As described by Klankermayer et al., (12b) the camphor-derived boranes 2 and 3 could not be isolated in pure stereoisomeric forms; however, kinetically controlled product formation in the reaction of the 2/3 mixture with phosphine PtBu3 (P) and H2 enabled isolation of diastereomerically pure ion pair compounds PH+/2H and PH+/3H. These phosphonium/hydroborate FLP salts were then used as catalysts in the hydrogenation of imines. In principle, the catalysis in these reactions can take place via two distinct cycles, as illustrated in Scheme 2.

Scheme 2

Scheme 2. Alternative Catalytic Cycles in Imine Reductiona

aNotations: P, im and am are defined in the text, B denotes chiral boranes 2 and 3. HT refers to hydride transfer from BH to the prochiral carbon of imH+.

In cycle 1, the heterolytic H2 splitting is induced by the P/B pair resulting in the PH+/BH ion pair. Proton transfer from PH+ to im gives the imH+/BH intermediate, which then undergoes hydride transfer (HT) to yield the chiral amine product am. Alternatively, the imine can serve as a base component of an FLP, so the imH+/BH intermediate is formed directly via H2 activation (cycle 2). The HT process represents the stereoselectivity-determining step in both cycles. Our computational analysis suggests that cycle 2 is a more feasible pathway. DFT calculations carried out for the reaction with borane 2 predict a significantly higher barrier for H2 activation with the P/2 pair as compared to that with im/2 (19.9 versus 15.8 kcal/mol; see Figure 1). (25)

Figure 1

Figure 1. Transition states of H2 activation by the P/2 and im/2 FLPs. Relative stabilities (in kcal/mol, with respect to the base + B + H2 reactant states) are given in parentheses. H atoms of the FLPs are omitted for clarity.

Even though PtBu3 is certainly more basic than imine im, (26) steric hindrance between the chiral borane substituent and the bulky phosphine destabilizes the transition state of H2 splitting (TSHHP/2) leading to an increased barrier. Furthermore, the protonation of im on the P/2 pathway is computed to have even higher barrier (31.6 kcal/mol) rendering cycle 1 far less favorable. (27) Our calculations thus indicate that although the catalytic process in the hydrogenation of im is initiated by the PH+/2H ion pair, the catalysis follows the im/2 pathway. The high barrier obtained for the initiation step is consistent with the elevated temperature (T = 65 °C) used in the experimental setup. (12b)

Iminium Borohydride Intermediates for Borane 2

The conformational space of the imH+/BH intermediate formed upon H2 activation is rather complex because both ionic components can adopt different structures and their relative positions can vary as well. Conformations of borohydrides BH can be classified into three groups that differ in the orientation of the B–H unit with respect to the chiral backbone (Chart 2).

Chart 2

Chart 2. Main Borohydride Conformers as Exemplified by 2H
The C6F5 borane substituent may also display altered orientations within these classes, resulting in several structural forms. For borohydride 2H, DFT calculations predict a c1-type conformation to be the most stable form, wherein one of the C6F5 rings is in stacking arrangement with the phenyl substituent of the chiral unit. (28) This borohydride structure was characterized experimentally via X-ray diffraction analysis of compound PH+/2H, and it was assumed to play an important role in the catalytic process. (12b) In principle, both E and Z isomers of iminium imH+ could be produced in the H2 activation step, so we have considered this possibility in our computational analysis as well.
For the imH+/2H ion pair intermediate, we carried out an extensive conformational analysis and identified 14 different isomeric forms all lying within a 5 kcal/mol free energy range. (29) A few representative structures are displayed in Figure 2.

Figure 2

Figure 2. Selected structures of imH+/2H ion pair intermediates. In the labeling, H···H and H···C refer to structures with B–H units pointing to iminium N–H bond or to prochiral C atom; c1 and c2 denote two different borohydride conformers. Relative stabilities (in kcal/mol, with respect to the im + 2 + H2) are given in parentheses. Selected bond distances are in angstroms.

The most stable form of imH+/2H involves a c1 borohydride conformer with the B–H bond interacting closely with the iminium N–H group (H···H/c1 in Figure 2). This imH+/2H isomer is formed directly after transition state TSHHim/2, and it is predicted to be 2.0 kcal/mol above the reactant state (im + 2 + H2). The dihydrogen bond is apparent from the very short intermolecular H···H distance (1.53 Å), and it provides considerable stabilization for the ion-pair intermediate. (30) Isomer H···H/c2 features a dihydrogen bonding interaction as well; however, the borohydride anion adopts a c2-type conformation. This form lies at 2.8 kcal/mol in free energy. Isomeric forms with B–H bonds pointing toward the electrophilic carbon atom of the substrate (H···C/c1 and H···C/c2 in Figure 2) are notably less stable as compared to their H···H/ci analogues; however, these isomers are structurally well prepared for the HT step of the catalytic cycle. Our computational analysis indicates that the H···H and H···C type imH+/2H isomers can easily interconvert with each other. For instance, the transformation of isomer H···H/c2 to H···C/c2 shown in Figure 2 can take place via a barrier of only 5.3 kcal/mol. Computations also suggest that facile transformation between the c1 and c2 borohydride conformers is feasible; the free energy barrier estimated for the c1c2 conversion is only 8.7 kcal/mol. (31) These results imply that the array of conformational isomers identified computationally for the structurally flexible ion-pair intermediate imH+/2H is likely in fast equilibrium even at room temperature.

Hydride Transfer Transition States for Borane 2

The conformational space of HT transition states leading to the (R)-am and (S)-am enantiomeric products were explored comprehensively, and we identified a set of energetically low-lying conformers on both pathways. The relative stabilities of the most stable structures are summarized in a free energy diagram shown in Figure 3, wherein selected transition-state structures are also depicted.

Figure 3

Figure 3. Hydride transfer transition states identified computationally for hydrogenation of im with borane 2. Each line on the free energy diagram represents a specific isomeric form with the computed relative stability. TS-2-Ri and TS-2-Si denote transition states leading to (R)-am and (S)-am products (index i defines the stability order). Full and dotted lines refer to transition state isomers involving c2 and c1 borohydride conformers. Selected structures are depicted and marked with arrows; their relative stabilities are given in parentheses (in kcal/mol, with respect to the most stable form). The iminium component is highlighted in blue for clarity. Green and red dotted arrows indicate attractive and repulsive intermolecular contacts. Computed and experimental (in brackets) ee data are shown below the diagram.

The most favored HT transition state corresponds to the formation of product (R)-am (see TS-2-R1 in Figure 3), and it is predicted to be 14.8 kcal/mol with respect to the reactant state (im + 2 + H2). In this structure, the borohydride that donates H to the iminium has a c2 conformation, which is somewhat surprising because the c2 form of 2H is predicted to be 4.7 kcal/mol less stable as compared to the c1 form. (28) Several intermolecular contacts, such as π–π stacking between the iminium phenyl and the borohydride C6F5 substituents, or CH···π interaction between the iminium CH3 and the borohydride phenyl groups, are noticeable in TS-2-R1. As shown previously for TM-catalyzed hydrogenation reactions, these types of noncovalent interactions can strongly influence the stereochemical outcome; (32,33) therefore, they may provide notable stabilization to transition state TS-2-R1 as well. Additional HT transition states involving the c2 form of 2H could be identified on the (R) reaction pathway (TS-2-R2 and TS-2-R3), but they are computed to be 2–3 kcal/mol less stable than TS-2-R1. The extent of noncovalent interactions in higher lying transition states is reduced, which corroborates the stabilizing nature of these intermolecular contacts.
Interestingly, the most favored transition state comprising a c1-type 2H conformation (TS-2-R4) is predicted to be even higher in free energy (at 4.3 kcal/mol). This structure is destabilized by steric repulsion between the chiral borane substituent and the methyl group of iminium imH+, which becomes too close as demonstrated by rather short H···H distances measured in the optimized structure (∼2.1 Å). (34)
On the HT reaction pathway that furnishes the minor (S)-am enantiomeric product, iminium imH+ approaches 2H with the other face. The most stable transition state TS-2-S1 also involves a c2-type borohydride; however, the intermolecular contacts are different from those in TS-2-R1. Although the CH···π interaction between the iminium CH3 and the borohydride phenyl groups is still present in TS-2-S1, no close π–π stacking interactions are perceived. The phenyl substituent of the borohydride interacts with the iminium phenyl group, but this aryl–aryl interaction is far from the ideal stacking or T-shaped arrangements. Overall, transition state TS-2-S1 is found to be 1.5 kcal/mol less stable than TS-2-R1. Several other close-lying transition states were located on the (S) pathway, of which the next in the stability order (TS-2-S2) incorporates a c1-type borohydride. Steric hindrance is also a characteristic feature of this transition state structure (similarly to TS-2-R4), but due to more favorable π–π stacking interactions, TS-2-S2 is more stable than TS-2-R4.
The enantioselectivity of catalytic imine hydrogenation is under kinetic control. (35) The rapid equilibration of various isomers of the imH+/2H intermediate enables the application of the Curtin–Hammett principle, so the enantiomeric excess (ee) can be estimated from the Boltzmann-weighted relative Gibbs free energies of the identified HT transition states. Using this procedure, the ee of the kinetically favored (R)-am product in the present reaction is computed to be 80.7%, which is in good agreement with the experimental observation (79%). (12b,36)

Reaction with Borane 3

Assuming that the concepts discussed in the previous sections and our conclusions regarding the hydrogenation mechanism can be extended to the analogous reactions with other boranes, we carried out a systematic computational study for the HT process with borane 3 as well. Borane 3 is the diastereomeric pair of 2 having the same bridged bicyclic (R)-camphor derived framework with the altered stereochemical arrangement of the B(C6F5)2 and Ph units (see Chart 1). The free energy diagram illustrating the relative stabilities of HT transition states computed for im hydrogenation with borane 3 is shown in Figure 4 along with the most stable structures identified on the (R) and (S) pathways.

Figure 4

Figure 4. Hydride transfer transition states identified computationally for hydrogenation of im with borane 3. For further relevant information, see the caption of Figure 3.

The most favored transition state TS-3-S1 can be regarded as a pseudoenantiomeric form of TS-2-R1 displaying a c2-type borohydride conformation and the same type of stabilizing intermolecular contacts (π–π stacking and CH3–π interactions). In this reaction, however, HT transition states involving the c1 borohydride conformation tend to be energetically more favored as compared to those in the reaction with borane 2, and in fact, the lowest lying transition state on the minor (R) pathway (TS-3-R1) encompasses a c1-type borohydride. This latter transition state is computed to be only 0.4 kcal/mol apart from TS-3-S1. Our structural analysis points to a somewhat reduced steric hindrance between the methyl group of iminium imH+ and the chiral bicyclic 3H substituent in transition states with c1 borohydrides, which explains the tendency of having these transition state structures more populated in the reaction with borane 3. (37) As a result of reduced steric repulsion, two transition states (TS-3-R1 and TS-3-S2) shift closer to the most favored TS-3-S1 structure in free energy, which of course influences the enantioselectivity as well. Computations predict ee = 33.7% for this reaction, which is consistent with the measured enantioselectivity (48%). (12b)

Reaction with Borane 1

HT transition state isomers for the reaction with borane 1 were also explored, and we identified a set of structures, some of which were found to have very similar relative stabilities (see Figure 5). (38)

Figure 5

Figure 5. Hydride transfer transition states identified computationally for hydrogenation of im with borane 1. For further relevant information, see the caption of Figure 3. The classification of transition states according to the borohydride conformations is not relevant in this case. The lowest lying energy level in the (S) ensemble (at 0.4 kcal/mol) represents two different structures, TS-1-S1 and TS-1-S2, of which only the former is depicted.

Due to the lack of any bulky substituent on the α-pinene framework in borane 1 (such as Ph in 2 and 3), the chiral environment around the reactive B–H bond of borohydride 1H is less defined. For instance, no multiple combinations of stabilizing π–π stacking and CH3–π contacts are developed between 1H and imH+; therefore, no structure with particularly enhanced stabilities exists. For the same reason, the most favored transition states on the competing (R) and (S) pathways become less separated in free energy. Indeed, calculations predict five different transition state structures within a 0.5 kcal/mol range, and the most stable diastereomeric forms (TS-1-R1 and TS-1-S1) display very similar intermolecular contacts (a single π–π stacking and a CH···π type interaction; see Figure 5). Consequently, a very low enantioselectivity is predicted (ee = 1.5%), which is again in line with experimental observations (13%). (12a)

On the Origin of Stereoselectivity

Several important findings that emerged from our computational analysis are worth highlighting when discussing the origin of enantioselectivity in the examined reactions. First, it appears that for camphor-derived boranes 2 and 3, the stereoselectivity of imine hydrogenation is dictated by thermodynamically less favored borohydride conformers of c2-type and not by the most stable c1 form that is observed experimentally. The c2 conformer of 2H is clearly more reactive in hydride transfer to imH+ and with its phenyl substituent arranged next to the B–H bond provides a chiral binding site for the approaching protonated substrate. The chiral environment is defined by the three aromatic rings of the borohydride resulting in facial selectivity for the hydride transfer (Figure 6).

Figure 6

Figure 6. Noncovalent interactions (NCI) in hydride transfer transition states TS-2-R1 and TS-2-S1. The borohydride is represented by a gray isodensity surface (ρ = 0.01 au); the iminium is shown in blue. The applied cutoff for reduced density gradient is s = 0.3 au. π–π stacking and CH3–π interactions are highlighted by green dotted arrows.

The stereoinduction in the present reactions is found to be influenced by multiple stabilizing noncovalent interactions between 2H and imH+, which are apparent from the NCI plots (39) generated for transition states TS-2-R1 and TS-2-S1. The green surface areas on these plots represent weak attractive noncovalent interactions corresponding to π–π stacking, CH3–π, Ph–Ph, etc. intermolecular contacts. As noted above, and also illustrated in Figure 6, the higher stability of transition state TS-2-R1 could be associated with more favorable aromatic (C6F5···C6H5) interactions in this structure as compared to that in TS-2-S1.

Proposed Modifications in Borane 2

Our computational analysis indicates that the phenyl substituent of camphor derived boranes is an important stereocontrol element in catalytic hydrogenation of im. This is further supported by calculations carried out for a model reaction catalyzed by a borane derived from 2 by omitting the Ph substituent. These calculations predict very low ee (only 5.6%) in this case. (40) We envisioned that additional substitutions implemented on the Ph group of borane 2, or replacing the Ph group by a larger aromatic ring, could alter the enantioselectivity of hydrogenation. The boranes considered for additional computational analysis are shown in Figure 7. (41)

Figure 7

Figure 7. Modified boranes and predicted ee data.

DFT calculations carried out for reactions catalyzed by boranes derived from 2 by adding either electron-withdrawing (F and CF3) or electron-donating (CH3 and tBu) groups at the meta positions of the Ph ring predicted significantly enhanced enantioselectivities (ee’s above 98%). We find that in these reactions, HT transition state conformers analogous to TS-2-S1 are notably destabilized with respect to the corresponding R1 structures. This is due to steric effects induced by the meta substituents on the catalyst Ph group. As discussed above, transition state TS-2-S1 is displaced only by 1.5 kcal/mol from the most stable TS-2-R1 structure (Figure 3), but this free energy difference increases to 4–5 kcal/mol with the modified boranes. The strength of the CH3–π interaction is also altered by introducing the meta substituents, (42) but these interactions are equally present in the most stable diastereomeric transition states (R1 and S1; see Figure 6), so they have no considerable influence on the enantioselectivity. For borane 2-ant, calculations predict somewhat lower ee (90.5%). In this borane variant, the c1-type borohydride conformations attain further stabilization via the extended aromatic stacking interactions, so all HT transition states of c1-type shift closer in free energy to the most stable form, reducing the enantioselectivity.

Experiments with Borane 2-F

To assess the reliability of DFT predictions, we first synthesized borane 2-F according to the procedure established for the reported camphor-derived boranes 2 and 3 (12) (Scheme 3). Reaction of enantiopure (1R)-(+)-camphor (11) with aryllithium 12, followed by dehydration of resulting tertiary alcohol 13 with thionyl chloride/pyridine gave olefin 14. Hydroboration of 14 with Piers’ borane under solvent-free conditions resulted in a mixture of diastereomeric boranes 2-F and 2-F′ in a 7:1 ratio according to multinuclear NMR. Pure 2-F could be isolated from the crude mixture by recrystallization from pentane at −20°C as a colorless powder in 52% yield, and its structure was confirmed by single-crystal X-ray diffraction analysis (Figure 8).

Scheme 3

Scheme 3. Synthesis of Chiral Borane 2-F

Figure 8

Figure 8. Crystal structure of borane 2-F. H atoms are omitted for clarity.

Catalytic hydrogenation of imine im with 2-F gave high conversions (above 90%) in 1 h at room temperature, however, no improvement could be obtained for the enantioselectivity as compared to that observed by Klankermayer et al.; the ee measured in our experiments reached 75% at most, depending on the reaction conditions (see Table S6 in the Supporting Information). Noteworthy, hydrogenation of im in the presence of external base 1,2,2,6,6-pentamethylpiperidine (PMP, 5 mol %) gave only 13% conversion. These results support the view emerging from the computational analysis that catalytic hydrogenation takes place via cycle 2 (Scheme 2) even in the presence of an external base.
The reason for the disagreement we see between the computed and measured enantioselectivities is not fully settled. The discrepancy may certainly arise from the inaccuracy of the DFT calculations, although the good match obtained for boranes 13 points to other, yet unknown, factors. In an attempt to provide a plausible explanation, we decided to assess the stability of 2-F upon the hydrogenation process. In a thick-wall gas-tight NMR tube, a stoichiometric mixture of 2-F and im in deuterated toluene was pressurized with 10 bar of dihydrogen. After 1 h, the sample was analyzed by 1H NMR, which revealed appearance of a new set of signals related to camphor scaffold in addition to those of 2-F (44% conversion). Although we could not reliably identify the structure of the newly formed species, a detailed NMR data analysis (19F, 11B, HH COSY and HH NOESY, see the SI) suggested that it was isomeric to both 2-F and 2-F′. The same species, albeit to a lesser extent (with 18% conversion), was formed when am was probed instead of im under the same reaction condition. On the basis of these observations, it is reasonable to assume that the new borane species produced in the catalytic process may catalyze a parallel low-enantioselective hydrogenation of im, which deteriorates the overall enantioselectivity of this reaction.

Experiments with Borane 2-tBu

Next, we decided to synthesize and to test borane 2-tBu, which involves bulkier 3,5-substituents on the phenyl group. Attempt to produce 2-tBu following the established procedure as for 2-F was unsuccessful due to predominant enolization of 11 upon addition of either 3,5-di-tert-butylphenyllithium or -magnesium bromide. Therefore, we designed an alternative synthetic route outlined in Scheme 4. Camphor 11 was transformed into bromobornene 17 via the Shapiro reaction. The subsequent Suzuki-Miyaura cross-coupling of 17 with 3,5-di-tert-butylphenyl boronic acid 18 gave alkene 19 in 79% yield. The hydroboration of 19 using Piers’ borane under solvent-free conditions gave exclusively diastereomer 2-tBu in nearly quantitative yield. Recrystallization of 2-tBu from n-pentane at −20°C provided colorless crystals (83% yield) suitable for X-ray crystallographic analysis (Figure 9).

Scheme 4

Scheme 4. Synthesis of Chiral Borane 2-tBu

Figure 9

Figure 9. Crystal structure of borane 2-tBu. H atoms are omitted for clarity.

Hydrogenation of imine im using 5 mol % borane 2-tBu at room temperature and 50 bar H2 in toluene gave only moderate conversion (52%) after 1 h, however, the enantioselectivity was high (91% ee; see Table S7 in the Supporting Information). Prolongation of reaction time to 24 h led to quantitative hydrogenation of im with similarly high enantiomeric excess (92%), demonstrating the stability of the catalyst under the applied conditions. Solvent screening showed only a small effect on the stereoselectivity, giving slightly lower ees in etherial solvents (see Table S7 in the Supporting Information). We note that this level of selectivity in FLP-type imine hydrogenation could only be accomplished by recently developed bicyclic borane 8 at significantly lower temperature (−40 °C).
Broadening of substrate scope was not an objective of our present work, but catalyst 2-tBu was tested for the hydrogenation of a few additional imines as well (Table 1 and also Table S8 in the Supporting Information). Various N-aryl-substituted-imines 20a20d and imine 20e bearing nonaromatic cyclohexyl substituent at nitrogen could be reduced with high stereoselectivities (91-94% ee). On the other hand, hydrogenation of the benzyl-substituted imine 20f resulted in a dramatically lower enantioselectivity (33%). Quantitative hydrogenation of 2-phenyl quinoline 20g to 2-phenyl-1,2,3,4-tetrahydroquinoline 21g could also be achieved, albeit in 48 h with doubling of catalyst loading, and with a significantly reduced ee (21%). Hydrogenation of selected substrates im and 20c at lower temperature (−15°C) afforded slightly enhanced enantioseletivities: 95% and 96% ee, respectively.
Table 1. Asymmetric Hydrogenations of Imines with 2-tBua
a

Substrate (0.25 mmol), PhMe (0.5ml), conversion by 1H NMR spectroscopy, ee by HPLC (Chiralcel OD-H or OJ-H column).

b

For detailed optimization, see Table S7 in the Supporting Information.

c

Reaction time 48 h.

d

10 mol % of 2-tBu.

These results highlight and confirm the importance of specific aryl–aryl and aryl–alkyl catalyst–substrate interactions in stereoinduction. For the particular borane catalyst 2-tBu and imines with topology analogous to im (20a20e) the overall effect of attractive noncovalent and destabilizing steric interactions is highly beneficial; however, this balance is far from optimal for other substrates (20f and 20g).

Summary and Conclusions

Click to copy section linkSection link copied!

Utilization of chiral FLP catalysts in direct asymmetric hydrogenation of unsaturated compounds is a potential metal-free strategy in stereoselective synthesis. Remarkable developments have been achieved along this line over the past decade; however, the current level of comprehension concerning the stereoselectivity governing factors in these catalytic processes is not sufficient thus far to facilitate new catalyst design. In our present work, we performed a detailed computational analysis for imine hydrogenation reactions reported previously by the Klankermayer group, namely those catalyzed by chiral boranes derived from (+)-α-pinene (borane 1 in Chart 1) and (1R)-(+)-camphor (boranes 2 and 3). Although only a single imine substrate (im) was considered in our computational study, and the selection of borane catalysts is limited as well, yet several interesting findings and conclusions have emerged from our analysis, which considerably improve our understanding.
Computations carried out for imine hydrogenation with borane 2 (the most selective reaction in the series) revealed that catalysis operates preferably via H2 activation with the im/2 pair even if the borane is introduced in a phosphonium–borohydride form, as in the original experiments. This computational insight could be supported experimentally in our work. The analysis focusing on the hydride transfer step pointed to fast equilibration of various isomers of the imH+/2H intermediate. We showed that the stereoselectivity is dictated by a thermodynamically unfavored borohydride isomer, and not by the most stable, experimentally observed form. This latter 2H isomer is less reactive for steric reasons. In the active form of 2H, the phenyl substituent of the camphor framework, along with the two C6F5 aromatic rings, define a chiral pocket for the approaching protonated substrate (imH+). The most stable hydride transfer transition states are found to be stabilized by specific noncovalent interactions, such as π–π stacking, CH3–π, and Ph–Ph intermolecular contacts. The enantioselectivity predicted by DFT calculations is in close agreement with experimental observations, and we find similarly good match between computed and measured ee data for the reactions with boranes 1 and 3.
Based on these promising results, we anticipated that alteration of the Ph substituent in borane 2 could be beneficial for stereoinduction, and indeed, calculations for boranes with 3,5-disubstituted Ph groups predicted significantly improved enantioselectivities. Two of these new borane candidates were selected for experimental studies. Boranes 2-F and 2-tBu were synthesized and probed as catalysts in imine hydrogenation. No improvement in the measured ee could be demonstrated with borane 2-F, but the experiments inferred that catalytic hydrogenation could be performed at room temperature in the absence of an external base. Borane 2-tBu, however, was shown to be a robust and efficient FLP catalyst in imine hydrogenation, providing ee’s above 90%, which could only be achieved so far at significantly lower temperature.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscatal.0c04263.

  • Details regarding the computational analysis, total energies and Cartesian coordinates for the considered stationary points, and experimental details (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Andrea Hamza - Institute of Organic Chemistry, Research Centre for Natural Sciences, Magyar tudósok körútja 2, H-1117 Budapest, HungaryOrcidhttp://orcid.org/0000-0002-9614-8522
    • Kristina Sorochkina - Department of Chemistry, University of Helsinki, A. I. Virtasen aukio 1, 00014 Helsinki, Finland
    • Bianka Kótai - Institute of Organic Chemistry, Research Centre for Natural Sciences, Magyar tudósok körútja 2, H-1117 Budapest, Hungary
    • Konstantin Chernichenko - Department of Chemistry, University of Helsinki, A. I. Virtasen aukio 1, 00014 Helsinki, FinlandOrcidhttp://orcid.org/0000-0002-6425-5075
    • Dénes Berta - Institute of Organic Chemistry, Research Centre for Natural Sciences, Magyar tudósok körútja 2, H-1117 Budapest, HungaryOrcidhttp://orcid.org/0000-0002-8299-3784
    • Michael Bolte - Institute of Inorganic Chemistry, Goethe-University, Max-von-Laue-Strasse 7, D-60438 Frankfurt am Main, GermanyOrcidhttp://orcid.org/0000-0001-5296-6251
    • Martin Nieger - Department of Chemistry, University of Helsinki, A. I. Virtasen aukio 1, 00014 Helsinki, FinlandOrcidhttp://orcid.org/0000-0003-1677-0109
  • Author Contributions

    A.H. and K.S. contributed equally.

  • Notes
    The authors declare no competing financial interest.

    CCDC 2007899 (2-F), and 2007900 (2-tBu) contain the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/.

Acknowledgments

Click to copy section linkSection link copied!

Financial support for this work was provided by the Hungarian Scientific Research Fund Grant No. K-115660) and from the Academy of Finland (Grant No. 316207). A.H. acknowledges the János Bolyai Scholarship from Hungarian Academy of Sciences. Computer facilities provided by NIIF HPC Hungary (project 85708 kataproc) are also acknowledged. We are grateful to Dr. Tibor András Rokob for insightful discussions and to Dr. Péter Nagy for assisting with benchmark studies.

References

Click to copy section linkSection link copied!

This article references 42 other publications.

  1. 1
    (a) Nugent, T. C. Chiral Amine Synthesis: Methods, Developments and Applications; Wiley-VCH: Weinheim, 2010.
    (b) 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, 753819,  DOI: 10.1002/adsc.200900719
  2. 2

    For selected reviews, see:

    (a) Spindler, F.; Blaser, H.-U. Enantioselective Hydrogenation of C=N Functions and Enamines. In Handbook of Homogenous Hydrogenation; de Vries, J. G., Elsevier, C. J., Eds.; Wiley-VCH: Weinheim, 2007; Vol. 3, pp 11931214.
    (b) Xie, J. H.; Zhu, S. F.; Zhou, Q. L. Transition Metal-Catalyzed Enantioselective Hydrogenation of Enamines and Imines. Chem. Rev. 2011, 111, 17131760,  DOI: 10.1021/cr100218m
    (c) Yu, Z.; Jin, W.; Jiang, Q. Brønsted Acid Activation Strategy in Transition-Metal Catalyzed Asymmetric Hydrogenation of N-Unprotected Imines, Enamines, and N-Heteroaromatic Compounds. Angew. Chem., Int. Ed. 2012, 51, 60606072,  DOI: 10.1002/anie.201200963
    (d) Xie, J.-H.; Zhu, S.-F.; Zhou, Q.-L. Recent Advances in Transition Metal-Catalyzed Enantioselective Hydrogenation of Unprotected Enamines. Chem. Soc. Rev. 2012, 41, 4126,  DOI: 10.1039/c2cs35007f
    (e) Stereoselective Formation of Amines; Li, W., Zhang, X., Eds.; Springer-Verlag: New York, 2014; Vol. 343.
    (f) Liu, Y.; Yue, X.; Luo, C.; Zhang, L.; Lei, M. Mechanisms of Ketone/Imine Hydrogenation Catalyzed by Transition-Metal Complexes. Energy Environ. Mater. 2019, 2, 292312,  DOI: 10.1002/eem2.12050
  3. 3

    For related reviews, see:

    (a) Hopmann, K. H.; Bayer, A. Enantioselective Imine Hydrogenation with Iridium-Catalysts: Reactions, Mechanisms and Stereocontrol. Coord. Chem. Rev. 2014, 268, 5982,  DOI: 10.1016/j.ccr.2014.01.023
    (b) Mwansa, J. M.; Page, M. I. Catalysis, Kinetics and Mechanisms of Organo-Iridium Enantioselective Hydrogenation-Reduction. Catal. Sci. Technol. 2020, 10, 590612,  DOI: 10.1039/C9CY02147G
    (c) Cui, C.-X.; Chen, H.; Li, S.-J.; Zhang, T.; Qu, L.-B.; Lan, Y. Mechanism of Ir-Catalyzed Hydrogenation: A Theoretical View. Coord. Chem. Rev. 2020, 412, 213251,  DOI: 10.1016/j.ccr.2020.213251
  4. 4

    For a review on metal-free hydrogenation strategies, see:

    Rossi, S.; Benaglia, M.; Massolo, E.; Raimondi, L. Organocatalytic Strategies for Enantioselective Metal-Free Reductions. Catal. Sci. Technol. 2014, 4, 27082723,  DOI: 10.1039/C4CY00033A
  5. 5

    For a review on organocatalytic transfer hydrogenation and hydrosilation reactions, see:

    Herrera, R. P. Organocatalytic Transfer Hydrogenation and Hydrosilylation Reactions. Top. Curr. Chem. 2016, 374, 29,  DOI: 10.1007/s41061-016-0032-4
  6. 6
    Welch, G. C.; Juan, R. R. S.; Masuda, J. D.; Stephan, D. W. Reversible, Metal-Free Hydrogen Activation. Science 2006, 314, 11241126,  DOI: 10.1126/science.1134230
  7. 7

    For the first influential works, see:

    (a) Chase, P. A.; Welch, G. C.; Jurca, T.; Stephan, D. W. Metal-Free Catalytic Hydrogenation. Angew. Chem., Int. Ed. 2007, 46, 80508053,  DOI: 10.1002/anie.200702908
    (b) Chase, P. A.; Jurca, T.; Stephan, D. W. Lewis Acid-Catalyzed Hydrogenation: B(C6F5)3-Mediated Reduction of Imines and Nitriles with H2. Chem. Commun. 2008, 14, 1701,  DOI: 10.1039/b718598g
    (c) Wang, H.; Fröhlich, R.; Kehr, G.; Erker, G. Heterolytic Dihydrogen Activation with the 1,8-Bis(Diphenylphosphino)Naphthalene/B(C6F5)3 Pair and Its Application for Metal-Free Catalytic Hydrogenation of Silyl Enol Ethers. Chem. Commun. 2008, 45, 5966,  DOI: 10.1039/b813286k
    (d) Spies, P.; Schwendemann, S.; Lange, S.; Kehr, G.; Fröhlich, R.; Erker, G. Metal-Free Catalytic Hydrogenation of Enamines, Imines, and Conjugated Phosphinoalkenylboranes. Angew. Chem., Int. Ed. 2008, 47, 75437546,  DOI: 10.1002/anie.200801432
    (e) Sumerin, V.; Schulz, F.; Atsumi, M.; Wang, C.; Nieger, M.; Leskelä, M.; Repo, T.; Pyykkö, P.; Rieger, B. Molecular Tweezers for Hydrogen: Synthesis, Characterization, and Reactivity. J. Am. Chem. Soc. 2008, 130, 1411714119,  DOI: 10.1021/ja806627s
  8. 8

    For reviews on FLP chemistry, see:

    (a) Topics in Current Chemistry; Erker, G., Stephan, D. W., Eds.; Springer-Verlag, 2013; Vols. 332 and 334.
    (b) Stephan, D. W.; Erker, G. Frustrated Lewis Pair Chemistry : Development and Perspectives. Angew. Chem., Int. Ed. 2015, 54, 64006441,  DOI: 10.1002/anie.201409800
    (c) Stephan, D. W. Frustrated Lewis Pairs. J. Am. Chem. Soc. 2015, 137, 1001810032,  DOI: 10.1021/jacs.5b06794
    (d) Stephan, D. W. Frustrated Lewis Pairs: From Concept to Catalysis. Acc. Chem. Res. 2015, 48, 306316,  DOI: 10.1021/ar500375j
    (e) Stephan, D. W. The Broadening Reach of Frustrated Lewis Pair Chemistry. Science 2016, 354, aaf7229aaf7229,  DOI: 10.1126/science.aaf7229
    (f) Jupp, A. R.; Stephan, D. W. New Directions for Frustrated Lewis Pair Chemistry. Trends in Chemistry 2019, 1, 3548,  DOI: 10.1016/j.trechm.2019.01.006
  9. 9

    For reviews on FLP-type catalytic hydrogenations, see:

    (a) Stephan, D. W.; Erker, G. Frustrated Lewis Pairs: Metal-free Hydrogen Activation and More. Angew. Chem., Int. Ed. 2010, 49, 4676,  DOI: 10.1002/anie.200903708
    (b) Stephan, D. W.; Greenberg, S.; Graham, T. W.; Chase, P.; Hastie, J. J.; Geier, S. J.; Farrell, J. M.; Brown, C. C.; Heiden, Z. M.; Welch, G. C.; Ullrich, M. Metal-Free Catalytic Hydrogenation of Polar Substrates by Frustrated Lewis Pairs. Inorg. Chem. 2011, 50, 1233812348,  DOI: 10.1021/ic200663v
    (c) Stephan, D. W.; Erker, G. Frustrated Lewis Pair Mediated Hydrogenations. Topics in Current Chemistry; Springer: Berlin, 2013; pp 85110.
    (d) Sumerin, V.; Chernichenko, K.; Schulz, F.; Leskelä, M.; Rieger, B.; Repo, T. Amine-Borane Mediated Metal-Free Hydrogen Activation and Catalytic Hydrogenation. Topics in Current Chemistry; Springer: Berlin, 2012; pp 111155.
    (e) Paradies, J. Metal-Free Hydrogenation of Unsaturated Hydrocarbons Employing Molecular Hydrogen. Angew. Chem., Int. Ed. 2014, 53, 35523557,  DOI: 10.1002/anie.201309253
    (f) Hounjet, L. J.; Stephan, D. W. Hydrogenation by Frustrated Lewis Pairs: Main Group Alternatives to Transition Metal Catalysts?. Org. Process Res. Dev. 2014, 18, 385391,  DOI: 10.1021/op400315m
    (g) Scott, D. J.; Fuchter, M. J.; Ashley, A. E. Designing Effective ‘Frustrated Lewis Pair’ Hydrogenation Catalysts. Chem. Soc. Rev. 2017, 46, 56895700,  DOI: 10.1039/C7CS00154A
    (h) Paradies, J. From Structure to Novel Reactivity in Frustrated Lewis Pairs. Coord. Chem. Rev. 2019, 380, 170183,  DOI: 10.1016/j.ccr.2018.09.014
    (i) Lam, J.; Szkop, K. M.; Mosaferi, E.; Stephan, D. W. FLP Catalysis: Main Group Hydrogenations of Organic Unsaturated Substrates. Chem. Soc. Rev. 2019, 48, 35923612,  DOI: 10.1039/C8CS00277K
    (j) Paradies, J. Mechanisms in Frustrated Lewis Pair-Catalyzed Reactions. Eur. J. Org. Chem. 2019, 2019, 283294,  DOI: 10.1002/ejoc.201800944
  10. 10

    For development of water-tolerant FLP catalysts, see:

    (a) Mahdi, T.; Stephan, D. W. Enabling Catalytic Ketone Hydrogenation by Frustrated Lewis Pairs. J. Am. Chem. Soc. 2014, 136, 1580915812,  DOI: 10.1021/ja508829x
    (b) Scott, D. J.; Fuchter, M. J.; Ashley, A. E. Nonmetal Catalyzed Hydrogenation of Carbonyl Compounds. J. Am. Chem. Soc. 2014, 136, 1581315816,  DOI: 10.1021/ja5088979
    (c) Gyömöre, Á.; Bakos, M.; Földes, T.; Pápai, I.; Domján, A.; Soós, T. Moisture-Tolerant Frustrated Lewis Pair Catalyst for Hydrogenation of Aldehydes and Ketones. ACS Catal. 2015, 5, 53665372,  DOI: 10.1021/acscatal.5b01299
    (d) Scott, D. J.; Simmons, T. R.; Lawrence, E. J.; Wildgoose, G. G.; Fuchter, M. J.; Ashley, A. E. Facile Protocol for Water-Tolerant “Frustrated Lewis Pair”-Catalyzed Hydrogenation. ACS Catal. 2015, 5, 55405544,  DOI: 10.1021/acscatal.5b01417
    (e) Fasano, V.; Radcliffe, J. E.; Ingleson, M. J. B(C6F5)3-Catalyzed Reductive Amination Using Hydrosilanes. ACS Catal. 2016, 6, 17931798,  DOI: 10.1021/acscatal.5b02896
    (f) Scott, D. J.; Phillips, N. A.; Sapsford, J. S.; Deacy, A. C.; Fuchter, M. J.; Ashley, A. E. Versatile Catalytic Hydrogenation Using A Simple Tin(IV) Lewis Acid. Angew. Chem., Int. Ed. 2016, 55, 1473814742,  DOI: 10.1002/anie.201606639
    (g) Dorkó, É.; Szabó, M.; Kótai, B.; Pápai, I.; Domján, A.; Soós, T. Expanding the Boundaries of Water-Tolerant Frustrated Lewis Pair Hydrogenation: Enhanced Back Strain in the Lewis Acid Enables the Reductive Amination of Carbonyls. Angew. Chem., Int. Ed. 2017, 56, 95129516,  DOI: 10.1002/anie.201703591
    (h) Sapsford, J. S.; Scott, D. J.; Allcock, N. J.; Fuchter, M. J.; Tighe, C. J.; Ashley, A. E. Direct Reductive Amination of Carbonyl Compounds Catalyzed by a Moisture Tolerant Tin(IV) Lewis Acid. Adv. Synth. Catal. 2018, 360, 10661071,  DOI: 10.1002/adsc.201701418
    (i) Hoshimoto, Y.; Kinoshita, T.; Hazra, S.; Ohashi, M.; Ogoshi, S. Main-Group-Catalyzed Reductive Alkylation of Multiply Substituted Amines with Aldehydes Using H2. J. Am. Chem. Soc. 2018, 140, 72927300,  DOI: 10.1021/jacs.8b03626
    (j) Fasano, V.; LaFortune, J. H. W.; Bayne, J. M.; Ingleson, M. J.; Stephan, D. W. Air- and Water-Stable Lewis Acids: Synthesis and Reactivity of P-Trifluoromethyl Electrophilic Phosphonium Cations. Chem. Commun. 2018, 54, 662665,  DOI: 10.1039/C7CC09128A
    (k) Fasano, V.; Ingleson, M. J. Recent Advances in Water-Tolerance in Frustrated Lewis Pair Chemistry. Synthesis 2018, 50, 17831795,  DOI: 10.1055/s-0037-1609843
  11. 11

    For review articles, see:

    (a) Chen, D.; Klankermayer, J. Frustrated Lewis Pairs: From Dihydrogen Activation to Asymmetric Catalysis. Top. Curr. Chem. 2013, 334, 126,  DOI: 10.1007/128_2012_402
    (b) Feng, X.; Du, H. Metal-Free Asymmetric Hydrogenation and Hydrosilylation Catalyzed by Frustrated Lewis Pairs. Tetrahedron Lett. 2014, 55, 69596964,  DOI: 10.1016/j.tetlet.2014.10.138
    (c) Shi, L.; Zhou, Y.-G. Enantioselective Metal-Free Hydrogenation Catalyzed by Chiral Frustrated Lewis Pairs. ChemCatChem 2015, 7, 5456,  DOI: 10.1002/cctc.201402838
    (d) Paradies, J. Chiral Borane-Based Lewis Acids for Metal Free Hydrogenations. Topics in Organometallic Chemistry; Springer International Publishing, 2018; pp 193216.
    (e) Meng, W.; Feng, X.; Du, H. Frustrated Lewis Pairs Catalyzed Asymmetric Metal-Free Hydrogenations and Hydrosilylations. Acc. Chem. Res. 2018, 51, 191201,  DOI: 10.1021/acs.accounts.7b00530
    (f) Meng, W.; Feng, X.; Du, H. Asymmetric Catalysis with Chiral Frustrated Lewis Pairs. Chin. J. Chem. 2020, 38, 625634,  DOI: 10.1002/cjoc.202000011
  12. 12
    (a) Chen, D.; Klankermayer, J. Metal-Free Catalytic Hydrogenation of Imines with Tris(Perfluorophenyl)Borane. Chem. Commun. 2008, 21302131,  DOI: 10.1039/b801806e
    (b) Chen, D.; Wang, Y.; Klankermayer, J. Enantioselective Hydrogenation with Chiral Frustrated Lewis Pairs. Angew. Chem., Int. Ed. 2010, 49, 94759478,  DOI: 10.1002/anie.201004525
    (c) Ghattas, G.; Chen, D.; Pan, F.; Klankermayer, J. Asymmetric Hydrogenation of Imines with a Recyclable Chiral Frustrated Lewis Pair Catalyst. Dalton Trans. 2012, 41, 90269028,  DOI: 10.1039/c2dt30536d
  13. 13
    (a) Parks, D. J.; von H. Spence, R. E.; Piers, W. E. Bis(Pentafluorophenyl)Borane: Synthesis, Properties, and Hydroboration Chemistry of a Highly Electrophilic Borane Reagent. Angew. Chem., Int. Ed. Engl. 1995, 34, 809811,  DOI: 10.1002/anie.199508091
    (b) Parks, D. J.; Piers, W. E.; Yap, G. P. a. Synthesis, Properties, and Hydroboration Activity of the Highly Electrophilic Borane Bis(Pentafluorophenyl)Borane. Organometallics 1998, 17, 54925503,  DOI: 10.1021/om980673e
    (c) Patrick, E. A.; Piers, W. E. Twenty-Five Years of Bis-Pentafluorophenyl Borane: A Versatile Reagent for Catalyst and Materials Synthesis. Chem. Commun. 2020, 56, 841853,  DOI: 10.1039/C9CC08338C
  14. 14
    Sumerin, V.; Chernichenko, K.; Nieger, M.; Leskelä, M.; Rieger, B.; Repo, T. Highly Active Metal-Free Catalysts for Hydrogenation of Unsaturated Nitrogen-Containing Compounds. Adv. Synth. Catal. 2011, 353, 20932110,  DOI: 10.1002/adsc.201100206
  15. 15
    Lindqvist, M.; Borre, K.; Axenov, K.; Kótai, B.; Nieger, M.; Leskela, M.; Pápai, I.; Repo, T. Chiral Molecular Tweezers: Synthesis and Reactivity in Asymmetric Hydrogenation. J. Am. Chem. Soc. 2015, 137, 40384041,  DOI: 10.1021/ja512658m
  16. 16
    Liu, Y.; Du, H. Chiral Dienes as “Ligands” for Borane-Catalyzed Metal-Free Asymmetric Hydrogenation of Imines. J. Am. Chem. Soc. 2013, 135, 68106813,  DOI: 10.1021/ja4025808
  17. 17
    Wei, S.; Du, H. A Highly Enantioselective Hydrogenation of Silyl Enol Ethers Catalyzed by Chiral Frustrated Lewis Pairs. J. Am. Chem. Soc. 2014, 136, 1226112264,  DOI: 10.1021/ja507536n
  18. 18
    (a) Zhang, Z.; Du, H. Cis-Selective and Highly Enantioselective Hydrogenation of 2,3,4-Trisubstituted Quinolines. Org. Lett. 2015, 17, 28162819,  DOI: 10.1021/acs.orglett.5b01240
    (b) Zhang, Z.; Du, H. Enantioselective Metal-Free Hydrogenations of Disubstituted Quinolines. Org. Lett. 2015, 17, 62666269,  DOI: 10.1021/acs.orglett.5b03307
    (c) Zhang, Z.; Du, H. A Highly cis-Selective and Enantioselective Metal-Free Hydrogenation of 2,3-Disubstituted Quinoxalines. Angew. Chem., Int. Ed. 2015, 54, 623626,  DOI: 10.1002/anie.201409471
    (d) Wei, S.; Feng, X.; Du, H. A Metal-Free Hydrogenation of 3-Substituted 2H-1,4-Benzoxazines. Org. Biomol. Chem. 2016, 14, 80268029,  DOI: 10.1039/C6OB01556E
  19. 19
    Tu, X.; Zeng, N.; Li, R.; Zhao, Y.; Xie, D.; Peng, Q.; Wang, X. C2-Symmetric Bicyclic Bisborane Catalysts: Kinetic or Thermodynamic Products of a Reversible Hydroboration of Dienes. Angew. Chem., Int. Ed. 2018, 57, 1509615100,  DOI: 10.1002/anie.201808289
  20. 20
    (a) Li, X.; Tian, J.; Liu, N.; Tu, X.; Zeng, N.; Wang, X. Spiro-Bicyclic Bisborane Catalysts for Metal-Free Chemoselective and Enantioselective Hydrogenation of Quinolines. Angew. Chem., Int. Ed. 2019, 58, 46644668,  DOI: 10.1002/anie.201900907
    (b) Tian, J.; Yang, Z.; Liang, X.; Liu, N.; Hu, C.; Tu, X.; Li, X.; Wang, X. Borane-Catalyzed Chemoselective and Enantioselective Reduction of 2-Vinyl-Substituted Pyridines. Angew. Chem., Int. Ed. 2020, 59, 1845218456,  DOI: 10.1002/anie.202007352
  21. 21
    Gao, B.; Feng, X.; Meng, W.; Du, H. Asymmetric Hydrogenation of Ketones and Enones with Chiral Lewis Base Derived Frustrated Lewis Pairs. Angew. Chem., Int. Ed. 2020, 59, 44984504,  DOI: 10.1002/anie.201914568
  22. 22

    For selected contributions, see:

    (a) Hermeke, J.; Mewald, M.; Oestreich, M. Experimental Analysis of the Catalytic Cycle of the Borane-Promoted Imine Reduction with Hydrosilanes: Spectroscopic Detection of Unexpected Intermediates and a Refined Mechanism. J. Am. Chem. Soc. 2013, 135, 1753717546,  DOI: 10.1021/ja409344w
    (b) Süsse, L.; Hermeke, J.; Oestreich, M. The Asymmetric Piers Hydrosilylation. J. Am. Chem. Soc. 2016, 138, 69406943,  DOI: 10.1021/jacs.6b03443
    (c) Mercea, D. M.; Howlett, M. G.; Piascik, A. D.; Scott, D.J.; Steven, A.; Ashley, A. E.; Fuchter, M. J. Enantioselective reduction of N-alkyl ketimines with frustrated Lewis pair catalysis using chiral borenium ions. Chem. Commun. 2019, 55, 70777080,  DOI: 10.1039/C9CC02900A
    (d) Liu, X.; Wang, Q.; Han, C.; Feng, X.; Du, H. Chiral Frustrated Lewis Pairs Catalyzed Highly Enantioselective Hydrosilylations of Ketones. Chin. J. Chem. 2019, 37, 663666,  DOI: 10.1002/cjoc.201900121
    (e) Lundrigan, T.; Welsh, E. N.; Hynes, T.; Tien, C.; Adams, M. R.; Roy, K. R.; Robertson, K. N.; Speed, A. W. H. Enantioselective Imine Reduction Catalyzed by Phosphenium Ions. J. Am. Chem. Soc. 2019, 141, 1408314088,  DOI: 10.1021/jacs.9b07293
  23. 23

    For computational mechanistic studies on FLP-type catalytic hydrogenations (not addressing the issue of stereoselectivity), see:

    (a) Rokob, T. A.; Hamza, A.; Stirling, A.; Pápai, I. On the Mechanism of B(C6F5)3-Catalyzed Direct Hydrogenation of Imines: Inherent and Thermally Induced Frustration. J. Am. Chem. Soc. 2009, 131, 20292036,  DOI: 10.1021/ja809125r
    (b) Nyhlén, J.; Privalov, T. On the Possibility of Catalytic Reduction of Carbonyl Moieties with Tris(Pentafluorophenyl)Borane and H2: A Computational Study. Dalton Trans. 2009, 29, 57805786,  DOI: 10.1039/b905137f
    (c) Privalov, T. The Role of Amine-B(C6F5)3 Adducts in the Catalytic Reduction of Imines with H2: A Computational Study. Eur. J. Inorg. Chem. 2009, 2009, 22292237,  DOI: 10.1002/ejic.200900194
    (d) Li, H.; Zhao, L.; Lu, G.; Huang, F.; Wang, Z.-X. Catalytic Metal-Free Ketone Hydrogenation: A Computational Experiment. Dalton Trans. 2010, 39, 55195526,  DOI: 10.1039/c001399d
    (e) Zhao, L.; Li, H.; Lu, G.; Huang, F.; Zhang, C.; Wang, Z.-X. Metal-Free Catalysts for Hydrogenation of Both Small and Large Imines: A Computational Experiment. Dalton Trans. 2011, 40, 19291937,  DOI: 10.1039/c0dt01297a
    (f) Zhao, L.; Lu, G.; Huang, F.; Wang, Z.-X. A Computational Experiment to Study Hydrogenations of Various Unsaturated Compounds Catalyzed by a Rationally Designed Metal-Free Catalyst. Dalton Trans. 2012, 41, 46744684,  DOI: 10.1039/c2dt12152b
    (g) Chernichenko, K.; Madarász, Á.; Pápai, I.; Nieger, M.; Leskelä, M.; Repo, T. A Frustrated-Lewis-Pair Approach to Catalytic Reduction of Alkynes to Cis-Alkenes. Nat. Chem. 2013, 5, 718723,  DOI: 10.1038/nchem.1693
    (h) Wang, Z.-X.; Zhao, L.; Lu, G.; Li, H. X.; Huang, F. Computational Design of Metal-Free Molecules for Activation of Small Molecules, Hydrogenation, and Hydroamination. Top. Curr. Chem. 2012, 332, 231266,  DOI: 10.1007/128_2012_385
    (i) Zhao, J.; Wang, G.; Li, S. Mechanistic Insights into the Full Hydrogenation of 2,6-Substituted Pyridine Catalyzed by the Lewis Acid C6F5(CH2)2B(C6F5)2. Dalton Trans. 2015, 44, 92009208,  DOI: 10.1039/C5DT00978B
    (j) Das, S.; Pati, S. K. On the Mechanism of Frustrated Lewis Pair Catalysed Hydrogenation of Carbonyl Compounds. Chem. - Eur. J. 2017, 23, 10781085,  DOI: 10.1002/chem.201602774
    (k) Heshmat, M.; Privalov, T. Carbonyl Activation by Borane Lewis Acid Complexation: Transition States of H2 Splitting at the Activated Carbonyl Carbon Atom in a Lewis Basic Solvent and the Proton-Transfer Dynamics of the Boroalkoxide Intermediate. Chem. - Eur. J. 2017, 23, 90989113,  DOI: 10.1002/chem.201700437
    (l) Mane, M. V.; Vanka, K. Less Frustration, More Activity-Theoretical Insights into Frustrated Lewis Pairs for Hydrogenation Catalysis. ChemCatChem 2017, 9, 30133022,  DOI: 10.1002/cctc.201700289
    (m) Heshmat, M.; Privalov, T. Computational Elucidation of a Role That Brønsted Acidification of the Lewis Acid-Bound Water Might Play in the Hydrogenation of Carbonyl Compounds with H2 in Lewis Basic Solvents. Chem. - Eur. J. 2017, 23, 1148911493,  DOI: 10.1002/chem.201700937
    (n) Heshmat, M.; Privalov, T. Theory-Based Extension of the Catalyst Scope in the Base-Catalyzed Hydrogenation of Ketones: RCOOH-Catalyzed Hydrogenation of Carbonyl Compounds with H2 Involving a Proton Shuttle. Chem. - Eur. J. 2017, 23, 1819318202,  DOI: 10.1002/chem.201702149
    (o) Das, S.; Pati, S. K. Unravelling the Mechanism of Tin-Based Frustrated Lewis Pair Catalysed Hydrogenation of Carbonyl Compounds. Catal. Sci. Technol. 2018, 8, 51785189,  DOI: 10.1039/C8CY01227J
  24. 24

    For recent reviews on concepts and challenges in computing stereoselectivities, see:

    (a) Hopmann, K. H. Quantum Chemical Studies of Asymmetric Reactions: Historical Aspects and Recent Examples. Int. J. Quantum Chem. 2015, 115, 12321249,  DOI: 10.1002/qua.24882
    (b) Peng, Q.; Duarte, F.; Paton, R. S. Computing Organic Stereoselectivity-from Concepts to Quantitative Calculations and Predictions. Chem. Soc. Rev. 2016, 45, 60936107,  DOI: 10.1039/C6CS00573J
    (c) Krenske, E. H. Challenges in Predicting Stereoselectivity. In Applied Theoretical Organic Chemistry; Tantillo, D. J., Ed.; World Scientific: New Jersey, 2018; pp 583604.
    (d) Gridnev, I. D.; Dub, P. A. Enantioselection in Asymmetric Catalysis; CRC Press: Boca Raton, 2017.
  25. 25

    For a detailed computational analysis of H2 activation pathways with the im/2 and P/2 pairs, see the SI (section 2.1).

  26. 26

    Computed proton affinities of PtBu3 and im are −277.5 and −231.7 kcal/mol, respectively. These values are obtained as solution phase Gibbs free energies of base → baseH+ reactions.

  27. 27

    For a detailed comparison of the energetics of the two catalytic cycles, see the SI (section 2.2).

  28. 28

    For conformational analysis of borohydride 2H, see the SI (section 2.3).

  29. 29

    For details of the conformational analysis carried out for the imH+/2H ion pair intermediate, see the SI (section 2.4).

  30. 30

    For related studies, see, for example:

    (a) Rokob, T. A.; Hamza, A.; Pápai, I. Rationalizing the Reactivity of Frustrated Lewis Pairs: Thermodynamics of H2 Activation and the Role of Acid–Base Properties. J. Am. Chem. Soc. 2009, 131, 1070110710,  DOI: 10.1021/ja903878z
    (b) Schulz, F.; Sumerin, V.; Heikkinen, S.; Pedersen, B.; Wang, C.; Atsumi, M.; Leskelä, M.; Repo, T.; Pyykkö, P.; Petry, W.; Rieger, B. Molecular Hydrogen Tweezers: Structure and Mechanisms by Neutron Diffraction, NMR, and Deuterium Labeling Studies in Solid and Solution. J. Am. Chem. Soc. 2011, 133, 2024520257,  DOI: 10.1021/ja206394w
    (c) Zaher, H.; Ashley, A. E.; Irwin, M.; Thompson, A. L.; Gutmann, M. J.; Krämer, T.; O’Hare, D. Structural and Theoretical Studies of Intermolecular Dihydrogen Bonding in [(C6F5)2(C6Cl5)B]–H···H–[TMP]. Chem. Commun. 2013, 49, 97559757,  DOI: 10.1039/c3cc45889j
    (d) Zhivonitko, V. V.; Sorochkina, K.; Chernichenko, K.; Kótai, B.; Földes, T.; Pápai, I.; Telkki, V.-V.; Repo, T.; Koptyug, I. Nuclear Spin Hyperpolarization with Ansa-Aminoboranes: A Metal-Free Perspective for Parahydrogen-Induced Polarization. Phys. Chem. Chem. Phys. 2016, 18, 2778427795,  DOI: 10.1039/C6CP05211H
  31. 31

    For the estimation of barriers relevant to the interconversion of imH+/2H isomers, see the SI (section 2.4).

  32. 32

    For computational studies examining the role of stabilizing noncovalent interactions in TM-catalyzed stereoselective hydrogenations, see:

    (a) Hopmann, K. H.; Bayer, A. On the Mechanism of Iridium-Catalyzed Asymmetric Hydrogenation of Imines and Alkenes: A Theoretical Study. Organometallics 2011, 30, 24832497,  DOI: 10.1021/om1009507
    (b) Václavík, J.; Kuzma, M.; Přech, J.; Kačer, P. Asymmetric Transfer Hydrogenation of Imines and Ketones Using Chiral RuIICl(η6-p-cymene)[(S,S)-N-TsDPEN] as a Catalyst: A Computational Study. Organometallics 2011, 30, 48224829,  DOI: 10.1021/om200263d
    (c) Wang, T.; Zhuo, L.-G.; Li, Z.; Chen, F.; Ding, Z.; He, Y.; Fan, Q.-H.; Xiang, J.; Yu, Z.-X.; Chan, A. S. C. Highly Enantioselective Hydrogenation of Quinolines Using Phosphine-Free Chiral Cationic Ruthenium Catalysts: Scope, Mechanism, and Origin of Enantioselectivity. J. Am. Chem. Soc. 2011, 133, 98789891,  DOI: 10.1021/ja2023042
    (d) Pablo, Ó.; Guijarro, D.; Kovács, G.; Lledós, A.; Ujaque, G.; Yus, M. A Versatile Ru Catalyst for the Asymmetric Transfer Hydrogenation of Both Aromatic and Aliphatic Sulfinylimines. Chem. - Eur. J. 2012, 18, 19691983,  DOI: 10.1002/chem.201102426
    (e) Hopmann, K. H. Iron/Brønsted Acid Catalyzed Asymmetric Hydrogenation: Mechanism and Selectivity-Determining Interactions. Chem. - Eur. J. 2015, 21, 1002010030,  DOI: 10.1002/chem.201500602
    (f) Tutkowski, B.; Kerdphon, S.; Limé, E.; Helquist, P.; Andersson, P. G.; Wiest, O.; Norrby, P.-O. Revisiting the Stereodetermining Step in Enantioselective Iridium-Catalyzed Imine Hydrogenation. ACS Catal. 2018, 8, 615623,  DOI: 10.1021/acscatal.7b02386
    (g) Salomó, E.; Gallen, A.; Sciortino, G.; Ujaque, G.; Grabulosa, A.; Lledós, A.; Riera, A.; Verdaguer, X. Direct Asymmetric Hydrogenation of N-Methyl and N-Alkyl Imines with an Ir(III)H Catalyst. J. Am. Chem. Soc. 2018, 140, 1696716970,  DOI: 10.1021/jacs.8b11547
    (h) Chen, J.; Gridnev, I. D. Size is Important: Artificial Catalyst Mimics Behaviour of Natural Enzymes. iScience 2020, 23, 100960,  DOI: 10.1016/j.isci.2020.100960
  33. 33

    For a selection of related studies on TM-catalyzed stereoselective hydrogenation of other substrates, see:

    (a) Hopmann, K. H. Cobalt–Bis(Imino)Pyridine-Catalyzed Asymmetric Hydrogenation: Electronic Structure, Mechanism, and Stereoselectivity. Organometallics 2013, 32, 63886399,  DOI: 10.1021/om400755k
    (b) Dub, P. a.; Henson, N. J.; Martin, R. L.; Gordon, J. C. Unravelling the Mechanism of the Asymmetric Hydrogenation of Acetophenone by [RuX2(Diphosphine)(1,2-Diamine)] Catalysts. J. Am. Chem. Soc. 2014, 136, 35053521,  DOI: 10.1021/ja411374j
    (c) Nakatsuka, H.; Yamamura, T.; Shuto, Y.; Tanaka, S.; Yoshimura, M.; Kitamura, M. Mechanism of Asymmetric Hydrogenation of Aromatic Ketones Catalyzed by a Combined System of Ru(π-CH2C(CH3)CH2)2(Cod) and the Chiral Sp2N/Sp3NH Hybrid Linear N4 Ligand Ph-BINAN-H-Py. J. Am. Chem. Soc. 2015, 137, 81388149,  DOI: 10.1021/jacs.5b02350
    (d) Dub, P. A.; Gordon, J. C. The Mechanism of Enantioselective Ketone Reduction with Noyori and Noyori–Ikariya Bifunctional Catalysts. Dalton Trans. 2016, 45, 67566781,  DOI: 10.1039/C6DT00476H
    (e) Nakane, S.; Yamamura, T.; Manna, S. K.; Tanaka, S.; Kitamura, M. Mechanistic Study of the Ru-Catalyzed Asymmetric Hydrogenation of Nonchelatable and Chelatable Tert-Alkyl Ketones Using the Linear Tridentate Sp3P/Sp3NH/Sp2N-Combined Ligand PN(H)N: RuNH- and RuNK-Involved Dual Catalytic Cycle. ACS Catal. 2018, 8, 1105911075,  DOI: 10.1021/acscatal.8b02671
  34. 34

    For structures of all HT transition states in the reaction with borane 2 and related structural analysis, see the SI (section 2.5).

  35. 35

    The HT step of the catalytic cycle of im hydrogenation is an irreversible process. For the reaction with borane 2, the am + 2 product state is predicted to be 16.1 kcal/mol below the most favored imH+/2H.

  36. 36

    The nearly quantitative agreement between computed and observed ee values is no doubt fortuitous and cannot be regarded as a measure of accuracy of the applied computational methodology.

  37. 37

    For structures of all HT transition states in the reaction with borane 3 and related structural analysis, see the SI (section 2.6).

  38. 38

    For structures of all HT transition states in the reaction with borane 1 and related structural analysis, see the SI (section 2.7).

  39. 39

    The RDG data were computed by using the NCIPLOT program:

    (a) Johnson, E. R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A. J.; Yang, W. Revealing Noncovalent Interactions. J. Am. Chem. Soc. 2010, 132, 64986506,  DOI: 10.1021/ja100936w
    (b) Contreras-García, J.; Johnson, E. R.; Keinan, S.; Chaudret, R.; Piquemal, J.-P.; Beratan, D. N.; Yang, W. NCIPLOT: A Program for Plotting Non-Covalent Interaction Regions. J. Chem. Theory Comput. 2011, 7, 625632,  DOI: 10.1021/ct100641a
  40. 40

    For details on the reaction with the simplified borane, see the SI (section 2.8).

  41. 41

    For details on computational analysis for hydrogenation reactions with boranes 2-F, 2-CF3, 2-CH3, 2-tBu, and 2-ant, see the SI (section 2.9).

  42. 42

    For the influence of various substituents on the strength of CH···π interaction, see:

    Karthikeyan, S.; Ramanathan, V.; Mishra, B. K. Influence of the Substituents on the CH···π Interaction: Benzene–Methane Complex. J. Phys. Chem. A 2013, 117, 66876694,  DOI: 10.1021/jp404972f

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 30 publications.

  1. Clara Faure, Salim Benmaouche, Philippe Belmont, Etienne Brachet, Diana Lamaa. N–H Insertion of Anilines on N-Tosylhydrazones Induced by Visible Light Irradiation. The Journal of Organic Chemistry 2024, 89 (16) , 11620-11630. https://doi.org/10.1021/acs.joc.4c01317
  2. Jiyang Zhao, Shaoxian Liu, Shanshan Liu, Wenwen Ding, Sijia Liu, Yao Chen, Pan Du. A Theoretical Study on the Borane-Catalyzed Reductive Amination of Aniline and Benzaldehyde with Dihydrogen: The Origins of Chemoselectivity. The Journal of Organic Chemistry 2022, 87 (2) , 1194-1207. https://doi.org/10.1021/acs.joc.1c02491
  3. Utku Yolsal, Thomas A. R. Horton, Meng Wang, Michael P. Shaver. Cyclic Ether Triggers for Polymeric Frustrated Lewis Pair Gels. Journal of the American Chemical Society 2021, 143 (33) , 12980-12984. https://doi.org/10.1021/jacs.1c06408
  4. Haiting Ye, Takuma Sato, Taishi Nakanishi, Shigetomo Ito, Shigenobu Umemiya, Masahiro Terada. DFT calculation-aided optimisation of a chiral phosphoric acid catalyst: case study of kinetic resolution of racemic secondary alcohols through lactonisation. Catalysis Science & Technology 2024, 14 (23) , 6869-6881. https://doi.org/10.1039/D4CY01029A
  5. Benjámin Kovács, Tamás Földes, Márk Szabó, Éva Dorkó, Bianka Kótai, Gergely Laczkó, Tamás Holczbauer, Attila Domján, Imre Pápai, Tibor Soós. Illuminating the multiple Lewis acidity of triaryl-boranes via atropisomeric dative adducts. Chemical Science 2024, 15 (38) , 15679-15689. https://doi.org/10.1039/D4SC00925H
  6. Péter R. Nagy. State-of-the-art local correlation methods enable affordable gold standard quantum chemistry for up to hundreds of atoms. Chemical Science 2024, 15 (36) , 14556-14584. https://doi.org/10.1039/D4SC04755A
  7. Shanti Gopal Patra. Asymmetric catalysis by chiral FLPs: A computational mini‐review. Chirality 2024, 36 (5) https://doi.org/10.1002/chir.23671
  8. B. Kótai, G. Laczkó, A. Hamza, I. Pápai. Stereocontrol via Propeller Chirality in FLP‐Catalyzed Asymmetric Hydrogenation. Chemistry – A European Journal 2024, 30 (21) https://doi.org/10.1002/chem.202400241
  9. Himangshu Mondal, Pratim Kumar Chattaraj. Unraveling Reactivity Pathways: Dihydrogen Activation and Hydrogenation of Multiple Bonds by Pyramidalized Boron‐Based Frustrated Lewis Pairs. ChemistryOpen 2024, 13 (4) https://doi.org/10.1002/open.202300179
  10. Jing Guo, Lvnan Jin, Lei Xiao, Maying Yan, Jiangkun Xiong, Zheng-Wang Qu, Stefan Grimme, Douglas W. Stephan. Readily accessible chiral frustrated Lewis pair catalysts: Asymmetric hydrosilylation of imines and quinolines. Chem Catalysis 2024, 4 (4) , 100959. https://doi.org/10.1016/j.checat.2024.100959
  11. Xiangqing Feng, Wei Meng, Haifeng Du. Asymmetric catalysis with FLPs. Chemical Society Reviews 2023, 52 (24) , 8580-8598. https://doi.org/10.1039/D3CS00701D
  12. Dániel Csókás, Bivas Mondal, Miloš Đokić, Richa Gupta, Beatrice J. Y. Lee, Rowan D. Young. Stereoselective Synthesis of Fluoroalkanes via FLP Mediated Monoselective C─F Activation of Geminal Difluoroalkanes. Advanced Science 2023, 10 (36) https://doi.org/10.1002/advs.202305768
  13. Zhenli Luo, Ji Yang, Zhen Yao, Jianbo Yang, Lijin Xu, Qian Shi. B(C 6 F 5 ) 3 ‐Catalyzed One‐Pot Tandem Diastereoselective Synthesis of cis ‐2,3‐Disubstituted 1,2,3,4‐Tetrahydroquinoxalines and cis ‐2,4‐Disubstituted 2,3,4,5‐Tetrahydro‐1 H ‐1,5‐benzodiazepines. Advanced Synthesis & Catalysis 2023, 365 (20) , 3527-3534. https://doi.org/10.1002/adsc.202300817
  14. Jin-Feng Li, Lan Luo, Zhi-Hui Bai, Bing Yin. Unveiling the correlation between the catalytic efficiency and acidity of a metal-free catalyst in a hydrogenation reaction. A theoretical case study of the hydrogenation of ethene catalyzed by a superacid arising from a superhalogen. Physical Chemistry Chemical Physics 2023, 25 (32) , 21684-21698. https://doi.org/10.1039/D3CP03147K
  15. Taishi Nakanishi, Masahiro Terada. Computational molecular refinement to enhance enantioselectivity by reinforcing hydrogen bonding interactions in major reaction pathway. Chemical Science 2023, 14 (21) , 5712-5721. https://doi.org/10.1039/D3SC01637D
  16. Rundong Zhou, Zoleykha Pirhadi Tavandashti, Jan Paradies. Frustrated Lewis Pair Catalyzed Reactions. SynOpen 2023, 07 (01) , 46-57. https://doi.org/10.1055/a-2005-5443
  17. Choon Wee Kee. Molecular Understanding and Practical In Silico Catalyst Design in Computational Organocatalysis and Phase Transfer Catalysis—Challenges and Opportunities. Molecules 2023, 28 (4) , 1715. https://doi.org/10.3390/molecules28041715
  18. Yin Zhang, Songbo Chen, Abdullah M. Al‐Enizi, Ayman Nafady, Zhiyong Tang, Shengqian Ma. Chiral Frustrated Lewis Pair@Metal‐Organic Framework as a New Platform for Heterogeneous Asymmetric Hydrogenation. Angewandte Chemie 2023, 135 (2) https://doi.org/10.1002/ange.202213399
  19. Yin Zhang, Songbo Chen, Abdullah M. Al‐Enizi, Ayman Nafady, Zhiyong Tang, Shengqian Ma. Chiral Frustrated Lewis Pair@Metal‐Organic Framework as a New Platform for Heterogeneous Asymmetric Hydrogenation. Angewandte Chemie International Edition 2023, 62 (2) https://doi.org/10.1002/anie.202213399
  20. Michael G. Guerzoni, Ayan Dasgupta, Emma Richards, Rebecca L. Melen. Enantioselective applications of frustrated Lewis pairs in organic synthesis. Chem Catalysis 2022, 2 (11) , 2865-2875. https://doi.org/10.1016/j.checat.2022.09.007
  21. Braulio Aranda, Gonzalo Valdebenito, Sebastián Parra-Melipán, Vicente López, Sergio Moya, Andrés Vega, Pedro Aguirre. Hydrogenation of imines catalyzed by ruthenium(II) complexes containing phosphorus-nitrogen ligands via hydrogen transfer reaction. Molecular Catalysis 2022, 526 , 112374. https://doi.org/10.1016/j.mcat.2022.112374
  22. Wentian Zou, Liuzhou Gao, Jia Cao, Zhenxing Li, Guoao Li, Guoqiang Wang, Shuhua Li. Mechanistic Insight into Hydroboration of Imines from Combined Computational and Experimental Studies. Chemistry – A European Journal 2022, 28 (11) https://doi.org/10.1002/chem.202104004
  23. Andrea Hamza, Daniel Moock, Christoph Schlepphorst, Jacob Schneidewind, Wolfgang Baumann, Frank Glorius. Unveiling a key catalytic pocket for the ruthenium NHC-catalysed asymmetric heteroarene hydrogenation. Chemical Science 2022, 13 (4) , 985-995. https://doi.org/10.1039/D1SC06409F
  24. Meera Mehta. Recent Development in the Solution-State Chemistry of Boranes and Diboranes. 2022, 122-195. https://doi.org/10.1016/B978-0-12-820206-7.00125-6
  25. Tanzeela Fazal, Fayaz Ali, Narayan S. Hosmane, Yinghuai Zhu. Boron compounds for catalytic applications. 2022, 169-199. https://doi.org/10.1016/bs.acat.2022.04.005
  26. Jun Li, Constantin G. Daniliuc, Gerald Kehr, Gerhard Erker. Three‐Component Reaction to 1,4,2‐Diazaborole‐Type Heteroarene Systems. Angewandte Chemie 2021, 133 (52) , 27259-27267. https://doi.org/10.1002/ange.202111946
  27. Jun Li, Constantin G. Daniliuc, Gerald Kehr, Gerhard Erker. Three‐Component Reaction to 1,4,2‐Diazaborole‐Type Heteroarene Systems. Angewandte Chemie International Edition 2021, 60 (52) , 27053-27061. https://doi.org/10.1002/anie.202111946
  28. . Borane-based Asymmetric FLP Hydrogenations. 2021, 80-94. https://doi.org/10.1039/9781839162442-00080
  29. Mojgan Heshmat. Lewis Acidity of Carbon in Activated Carbonyl Group vs . B(C 6 F 5 ) 3 for Metal‐Free Catalysis of Hydrogenation of Carbonyl Compounds. ChemPhysChem 2021, 22 (14) , 1535-1542. https://doi.org/10.1002/cphc.202100003
  30. . Frustrated Lewis Pair Catalyzed Asymmetric Hydrogenation of Imines. Synfacts 2021, 0202. https://doi.org/10.1055/s-0040-1706120

ACS Catalysis

Cite this: ACS Catal. 2020, 10, 23, 14290–14301
Click to copy citationCitation copied!
https://doi.org/10.1021/acscatal.0c04263
Published November 23, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Article Views

5237

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Chart 1

    Chart 1. Selection of Chiral FLP Components Employed in Enantioselective Catalytic Hydrogenationsa

    aAr denotes aromatic substituents; ArF = C6F5 or p-C6F4H.

    Scheme 1

    Scheme 1. Reactions Examined Computationally

    Scheme 2

    Scheme 2. Alternative Catalytic Cycles in Imine Reductiona

    aNotations: P, im and am are defined in the text, B denotes chiral boranes 2 and 3. HT refers to hydride transfer from BH to the prochiral carbon of imH+.

    Figure 1

    Figure 1. Transition states of H2 activation by the P/2 and im/2 FLPs. Relative stabilities (in kcal/mol, with respect to the base + B + H2 reactant states) are given in parentheses. H atoms of the FLPs are omitted for clarity.

    Chart 2

    Chart 2. Main Borohydride Conformers as Exemplified by 2H

    Figure 2

    Figure 2. Selected structures of imH+/2H ion pair intermediates. In the labeling, H···H and H···C refer to structures with B–H units pointing to iminium N–H bond or to prochiral C atom; c1 and c2 denote two different borohydride conformers. Relative stabilities (in kcal/mol, with respect to the im + 2 + H2) are given in parentheses. Selected bond distances are in angstroms.

    Figure 3

    Figure 3. Hydride transfer transition states identified computationally for hydrogenation of im with borane 2. Each line on the free energy diagram represents a specific isomeric form with the computed relative stability. TS-2-Ri and TS-2-Si denote transition states leading to (R)-am and (S)-am products (index i defines the stability order). Full and dotted lines refer to transition state isomers involving c2 and c1 borohydride conformers. Selected structures are depicted and marked with arrows; their relative stabilities are given in parentheses (in kcal/mol, with respect to the most stable form). The iminium component is highlighted in blue for clarity. Green and red dotted arrows indicate attractive and repulsive intermolecular contacts. Computed and experimental (in brackets) ee data are shown below the diagram.

    Figure 4

    Figure 4. Hydride transfer transition states identified computationally for hydrogenation of im with borane 3. For further relevant information, see the caption of Figure 3.

    Figure 5

    Figure 5. Hydride transfer transition states identified computationally for hydrogenation of im with borane 1. For further relevant information, see the caption of Figure 3. The classification of transition states according to the borohydride conformations is not relevant in this case. The lowest lying energy level in the (S) ensemble (at 0.4 kcal/mol) represents two different structures, TS-1-S1 and TS-1-S2, of which only the former is depicted.

    Figure 6

    Figure 6. Noncovalent interactions (NCI) in hydride transfer transition states TS-2-R1 and TS-2-S1. The borohydride is represented by a gray isodensity surface (ρ = 0.01 au); the iminium is shown in blue. The applied cutoff for reduced density gradient is s = 0.3 au. π–π stacking and CH3–π interactions are highlighted by green dotted arrows.

    Figure 7

    Figure 7. Modified boranes and predicted ee data.

    Scheme 3

    Scheme 3. Synthesis of Chiral Borane 2-F

    Figure 8

    Figure 8. Crystal structure of borane 2-F. H atoms are omitted for clarity.

    Scheme 4

    Scheme 4. Synthesis of Chiral Borane 2-tBu

    Figure 9

    Figure 9. Crystal structure of borane 2-tBu. H atoms are omitted for clarity.

  • References


    This article references 42 other publications.

    1. 1
      (a) Nugent, T. C. Chiral Amine Synthesis: Methods, Developments and Applications; Wiley-VCH: Weinheim, 2010.
      (b) 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, 753819,  DOI: 10.1002/adsc.200900719
    2. 2

      For selected reviews, see:

      (a) Spindler, F.; Blaser, H.-U. Enantioselective Hydrogenation of C=N Functions and Enamines. In Handbook of Homogenous Hydrogenation; de Vries, J. G., Elsevier, C. J., Eds.; Wiley-VCH: Weinheim, 2007; Vol. 3, pp 11931214.
      (b) Xie, J. H.; Zhu, S. F.; Zhou, Q. L. Transition Metal-Catalyzed Enantioselective Hydrogenation of Enamines and Imines. Chem. Rev. 2011, 111, 17131760,  DOI: 10.1021/cr100218m
      (c) Yu, Z.; Jin, W.; Jiang, Q. Brønsted Acid Activation Strategy in Transition-Metal Catalyzed Asymmetric Hydrogenation of N-Unprotected Imines, Enamines, and N-Heteroaromatic Compounds. Angew. Chem., Int. Ed. 2012, 51, 60606072,  DOI: 10.1002/anie.201200963
      (d) Xie, J.-H.; Zhu, S.-F.; Zhou, Q.-L. Recent Advances in Transition Metal-Catalyzed Enantioselective Hydrogenation of Unprotected Enamines. Chem. Soc. Rev. 2012, 41, 4126,  DOI: 10.1039/c2cs35007f
      (e) Stereoselective Formation of Amines; Li, W., Zhang, X., Eds.; Springer-Verlag: New York, 2014; Vol. 343.
      (f) Liu, Y.; Yue, X.; Luo, C.; Zhang, L.; Lei, M. Mechanisms of Ketone/Imine Hydrogenation Catalyzed by Transition-Metal Complexes. Energy Environ. Mater. 2019, 2, 292312,  DOI: 10.1002/eem2.12050
    3. 3

      For related reviews, see:

      (a) Hopmann, K. H.; Bayer, A. Enantioselective Imine Hydrogenation with Iridium-Catalysts: Reactions, Mechanisms and Stereocontrol. Coord. Chem. Rev. 2014, 268, 5982,  DOI: 10.1016/j.ccr.2014.01.023
      (b) Mwansa, J. M.; Page, M. I. Catalysis, Kinetics and Mechanisms of Organo-Iridium Enantioselective Hydrogenation-Reduction. Catal. Sci. Technol. 2020, 10, 590612,  DOI: 10.1039/C9CY02147G
      (c) Cui, C.-X.; Chen, H.; Li, S.-J.; Zhang, T.; Qu, L.-B.; Lan, Y. Mechanism of Ir-Catalyzed Hydrogenation: A Theoretical View. Coord. Chem. Rev. 2020, 412, 213251,  DOI: 10.1016/j.ccr.2020.213251
    4. 4

      For a review on metal-free hydrogenation strategies, see:

      Rossi, S.; Benaglia, M.; Massolo, E.; Raimondi, L. Organocatalytic Strategies for Enantioselective Metal-Free Reductions. Catal. Sci. Technol. 2014, 4, 27082723,  DOI: 10.1039/C4CY00033A
    5. 5

      For a review on organocatalytic transfer hydrogenation and hydrosilation reactions, see:

      Herrera, R. P. Organocatalytic Transfer Hydrogenation and Hydrosilylation Reactions. Top. Curr. Chem. 2016, 374, 29,  DOI: 10.1007/s41061-016-0032-4
    6. 6
      Welch, G. C.; Juan, R. R. S.; Masuda, J. D.; Stephan, D. W. Reversible, Metal-Free Hydrogen Activation. Science 2006, 314, 11241126,  DOI: 10.1126/science.1134230
    7. 7

      For the first influential works, see:

      (a) Chase, P. A.; Welch, G. C.; Jurca, T.; Stephan, D. W. Metal-Free Catalytic Hydrogenation. Angew. Chem., Int. Ed. 2007, 46, 80508053,  DOI: 10.1002/anie.200702908
      (b) Chase, P. A.; Jurca, T.; Stephan, D. W. Lewis Acid-Catalyzed Hydrogenation: B(C6F5)3-Mediated Reduction of Imines and Nitriles with H2. Chem. Commun. 2008, 14, 1701,  DOI: 10.1039/b718598g
      (c) Wang, H.; Fröhlich, R.; Kehr, G.; Erker, G. Heterolytic Dihydrogen Activation with the 1,8-Bis(Diphenylphosphino)Naphthalene/B(C6F5)3 Pair and Its Application for Metal-Free Catalytic Hydrogenation of Silyl Enol Ethers. Chem. Commun. 2008, 45, 5966,  DOI: 10.1039/b813286k
      (d) Spies, P.; Schwendemann, S.; Lange, S.; Kehr, G.; Fröhlich, R.; Erker, G. Metal-Free Catalytic Hydrogenation of Enamines, Imines, and Conjugated Phosphinoalkenylboranes. Angew. Chem., Int. Ed. 2008, 47, 75437546,  DOI: 10.1002/anie.200801432
      (e) Sumerin, V.; Schulz, F.; Atsumi, M.; Wang, C.; Nieger, M.; Leskelä, M.; Repo, T.; Pyykkö, P.; Rieger, B. Molecular Tweezers for Hydrogen: Synthesis, Characterization, and Reactivity. J. Am. Chem. Soc. 2008, 130, 1411714119,  DOI: 10.1021/ja806627s
    8. 8

      For reviews on FLP chemistry, see:

      (a) Topics in Current Chemistry; Erker, G., Stephan, D. W., Eds.; Springer-Verlag, 2013; Vols. 332 and 334.
      (b) Stephan, D. W.; Erker, G. Frustrated Lewis Pair Chemistry : Development and Perspectives. Angew. Chem., Int. Ed. 2015, 54, 64006441,  DOI: 10.1002/anie.201409800
      (c) Stephan, D. W. Frustrated Lewis Pairs. J. Am. Chem. Soc. 2015, 137, 1001810032,  DOI: 10.1021/jacs.5b06794
      (d) Stephan, D. W. Frustrated Lewis Pairs: From Concept to Catalysis. Acc. Chem. Res. 2015, 48, 306316,  DOI: 10.1021/ar500375j
      (e) Stephan, D. W. The Broadening Reach of Frustrated Lewis Pair Chemistry. Science 2016, 354, aaf7229aaf7229,  DOI: 10.1126/science.aaf7229
      (f) Jupp, A. R.; Stephan, D. W. New Directions for Frustrated Lewis Pair Chemistry. Trends in Chemistry 2019, 1, 3548,  DOI: 10.1016/j.trechm.2019.01.006
    9. 9

      For reviews on FLP-type catalytic hydrogenations, see:

      (a) Stephan, D. W.; Erker, G. Frustrated Lewis Pairs: Metal-free Hydrogen Activation and More. Angew. Chem., Int. Ed. 2010, 49, 4676,  DOI: 10.1002/anie.200903708
      (b) Stephan, D. W.; Greenberg, S.; Graham, T. W.; Chase, P.; Hastie, J. J.; Geier, S. J.; Farrell, J. M.; Brown, C. C.; Heiden, Z. M.; Welch, G. C.; Ullrich, M. Metal-Free Catalytic Hydrogenation of Polar Substrates by Frustrated Lewis Pairs. Inorg. Chem. 2011, 50, 1233812348,  DOI: 10.1021/ic200663v
      (c) Stephan, D. W.; Erker, G. Frustrated Lewis Pair Mediated Hydrogenations. Topics in Current Chemistry; Springer: Berlin, 2013; pp 85110.
      (d) Sumerin, V.; Chernichenko, K.; Schulz, F.; Leskelä, M.; Rieger, B.; Repo, T. Amine-Borane Mediated Metal-Free Hydrogen Activation and Catalytic Hydrogenation. Topics in Current Chemistry; Springer: Berlin, 2012; pp 111155.
      (e) Paradies, J. Metal-Free Hydrogenation of Unsaturated Hydrocarbons Employing Molecular Hydrogen. Angew. Chem., Int. Ed. 2014, 53, 35523557,  DOI: 10.1002/anie.201309253
      (f) Hounjet, L. J.; Stephan, D. W. Hydrogenation by Frustrated Lewis Pairs: Main Group Alternatives to Transition Metal Catalysts?. Org. Process Res. Dev. 2014, 18, 385391,  DOI: 10.1021/op400315m
      (g) Scott, D. J.; Fuchter, M. J.; Ashley, A. E. Designing Effective ‘Frustrated Lewis Pair’ Hydrogenation Catalysts. Chem. Soc. Rev. 2017, 46, 56895700,  DOI: 10.1039/C7CS00154A
      (h) Paradies, J. From Structure to Novel Reactivity in Frustrated Lewis Pairs. Coord. Chem. Rev. 2019, 380, 170183,  DOI: 10.1016/j.ccr.2018.09.014
      (i) Lam, J.; Szkop, K. M.; Mosaferi, E.; Stephan, D. W. FLP Catalysis: Main Group Hydrogenations of Organic Unsaturated Substrates. Chem. Soc. Rev. 2019, 48, 35923612,  DOI: 10.1039/C8CS00277K
      (j) Paradies, J. Mechanisms in Frustrated Lewis Pair-Catalyzed Reactions. Eur. J. Org. Chem. 2019, 2019, 283294,  DOI: 10.1002/ejoc.201800944
    10. 10

      For development of water-tolerant FLP catalysts, see:

      (a) Mahdi, T.; Stephan, D. W. Enabling Catalytic Ketone Hydrogenation by Frustrated Lewis Pairs. J. Am. Chem. Soc. 2014, 136, 1580915812,  DOI: 10.1021/ja508829x
      (b) Scott, D. J.; Fuchter, M. J.; Ashley, A. E. Nonmetal Catalyzed Hydrogenation of Carbonyl Compounds. J. Am. Chem. Soc. 2014, 136, 1581315816,  DOI: 10.1021/ja5088979
      (c) Gyömöre, Á.; Bakos, M.; Földes, T.; Pápai, I.; Domján, A.; Soós, T. Moisture-Tolerant Frustrated Lewis Pair Catalyst for Hydrogenation of Aldehydes and Ketones. ACS Catal. 2015, 5, 53665372,  DOI: 10.1021/acscatal.5b01299
      (d) Scott, D. J.; Simmons, T. R.; Lawrence, E. J.; Wildgoose, G. G.; Fuchter, M. J.; Ashley, A. E. Facile Protocol for Water-Tolerant “Frustrated Lewis Pair”-Catalyzed Hydrogenation. ACS Catal. 2015, 5, 55405544,  DOI: 10.1021/acscatal.5b01417
      (e) Fasano, V.; Radcliffe, J. E.; Ingleson, M. J. B(C6F5)3-Catalyzed Reductive Amination Using Hydrosilanes. ACS Catal. 2016, 6, 17931798,  DOI: 10.1021/acscatal.5b02896
      (f) Scott, D. J.; Phillips, N. A.; Sapsford, J. S.; Deacy, A. C.; Fuchter, M. J.; Ashley, A. E. Versatile Catalytic Hydrogenation Using A Simple Tin(IV) Lewis Acid. Angew. Chem., Int. Ed. 2016, 55, 1473814742,  DOI: 10.1002/anie.201606639
      (g) Dorkó, É.; Szabó, M.; Kótai, B.; Pápai, I.; Domján, A.; Soós, T. Expanding the Boundaries of Water-Tolerant Frustrated Lewis Pair Hydrogenation: Enhanced Back Strain in the Lewis Acid Enables the Reductive Amination of Carbonyls. Angew. Chem., Int. Ed. 2017, 56, 95129516,  DOI: 10.1002/anie.201703591
      (h) Sapsford, J. S.; Scott, D. J.; Allcock, N. J.; Fuchter, M. J.; Tighe, C. J.; Ashley, A. E. Direct Reductive Amination of Carbonyl Compounds Catalyzed by a Moisture Tolerant Tin(IV) Lewis Acid. Adv. Synth. Catal. 2018, 360, 10661071,  DOI: 10.1002/adsc.201701418
      (i) Hoshimoto, Y.; Kinoshita, T.; Hazra, S.; Ohashi, M.; Ogoshi, S. Main-Group-Catalyzed Reductive Alkylation of Multiply Substituted Amines with Aldehydes Using H2. J. Am. Chem. Soc. 2018, 140, 72927300,  DOI: 10.1021/jacs.8b03626
      (j) Fasano, V.; LaFortune, J. H. W.; Bayne, J. M.; Ingleson, M. J.; Stephan, D. W. Air- and Water-Stable Lewis Acids: Synthesis and Reactivity of P-Trifluoromethyl Electrophilic Phosphonium Cations. Chem. Commun. 2018, 54, 662665,  DOI: 10.1039/C7CC09128A
      (k) Fasano, V.; Ingleson, M. J. Recent Advances in Water-Tolerance in Frustrated Lewis Pair Chemistry. Synthesis 2018, 50, 17831795,  DOI: 10.1055/s-0037-1609843
    11. 11

      For review articles, see:

      (a) Chen, D.; Klankermayer, J. Frustrated Lewis Pairs: From Dihydrogen Activation to Asymmetric Catalysis. Top. Curr. Chem. 2013, 334, 126,  DOI: 10.1007/128_2012_402
      (b) Feng, X.; Du, H. Metal-Free Asymmetric Hydrogenation and Hydrosilylation Catalyzed by Frustrated Lewis Pairs. Tetrahedron Lett. 2014, 55, 69596964,  DOI: 10.1016/j.tetlet.2014.10.138
      (c) Shi, L.; Zhou, Y.-G. Enantioselective Metal-Free Hydrogenation Catalyzed by Chiral Frustrated Lewis Pairs. ChemCatChem 2015, 7, 5456,  DOI: 10.1002/cctc.201402838
      (d) Paradies, J. Chiral Borane-Based Lewis Acids for Metal Free Hydrogenations. Topics in Organometallic Chemistry; Springer International Publishing, 2018; pp 193216.
      (e) Meng, W.; Feng, X.; Du, H. Frustrated Lewis Pairs Catalyzed Asymmetric Metal-Free Hydrogenations and Hydrosilylations. Acc. Chem. Res. 2018, 51, 191201,  DOI: 10.1021/acs.accounts.7b00530
      (f) Meng, W.; Feng, X.; Du, H. Asymmetric Catalysis with Chiral Frustrated Lewis Pairs. Chin. J. Chem. 2020, 38, 625634,  DOI: 10.1002/cjoc.202000011
    12. 12
      (a) Chen, D.; Klankermayer, J. Metal-Free Catalytic Hydrogenation of Imines with Tris(Perfluorophenyl)Borane. Chem. Commun. 2008, 21302131,  DOI: 10.1039/b801806e
      (b) Chen, D.; Wang, Y.; Klankermayer, J. Enantioselective Hydrogenation with Chiral Frustrated Lewis Pairs. Angew. Chem., Int. Ed. 2010, 49, 94759478,  DOI: 10.1002/anie.201004525
      (c) Ghattas, G.; Chen, D.; Pan, F.; Klankermayer, J. Asymmetric Hydrogenation of Imines with a Recyclable Chiral Frustrated Lewis Pair Catalyst. Dalton Trans. 2012, 41, 90269028,  DOI: 10.1039/c2dt30536d
    13. 13
      (a) Parks, D. J.; von H. Spence, R. E.; Piers, W. E. Bis(Pentafluorophenyl)Borane: Synthesis, Properties, and Hydroboration Chemistry of a Highly Electrophilic Borane Reagent. Angew. Chem., Int. Ed. Engl. 1995, 34, 809811,  DOI: 10.1002/anie.199508091
      (b) Parks, D. J.; Piers, W. E.; Yap, G. P. a. Synthesis, Properties, and Hydroboration Activity of the Highly Electrophilic Borane Bis(Pentafluorophenyl)Borane. Organometallics 1998, 17, 54925503,  DOI: 10.1021/om980673e
      (c) Patrick, E. A.; Piers, W. E. Twenty-Five Years of Bis-Pentafluorophenyl Borane: A Versatile Reagent for Catalyst and Materials Synthesis. Chem. Commun. 2020, 56, 841853,  DOI: 10.1039/C9CC08338C
    14. 14
      Sumerin, V.; Chernichenko, K.; Nieger, M.; Leskelä, M.; Rieger, B.; Repo, T. Highly Active Metal-Free Catalysts for Hydrogenation of Unsaturated Nitrogen-Containing Compounds. Adv. Synth. Catal. 2011, 353, 20932110,  DOI: 10.1002/adsc.201100206
    15. 15
      Lindqvist, M.; Borre, K.; Axenov, K.; Kótai, B.; Nieger, M.; Leskela, M.; Pápai, I.; Repo, T. Chiral Molecular Tweezers: Synthesis and Reactivity in Asymmetric Hydrogenation. J. Am. Chem. Soc. 2015, 137, 40384041,  DOI: 10.1021/ja512658m
    16. 16
      Liu, Y.; Du, H. Chiral Dienes as “Ligands” for Borane-Catalyzed Metal-Free Asymmetric Hydrogenation of Imines. J. Am. Chem. Soc. 2013, 135, 68106813,  DOI: 10.1021/ja4025808
    17. 17
      Wei, S.; Du, H. A Highly Enantioselective Hydrogenation of Silyl Enol Ethers Catalyzed by Chiral Frustrated Lewis Pairs. J. Am. Chem. Soc. 2014, 136, 1226112264,  DOI: 10.1021/ja507536n
    18. 18
      (a) Zhang, Z.; Du, H. Cis-Selective and Highly Enantioselective Hydrogenation of 2,3,4-Trisubstituted Quinolines. Org. Lett. 2015, 17, 28162819,  DOI: 10.1021/acs.orglett.5b01240
      (b) Zhang, Z.; Du, H. Enantioselective Metal-Free Hydrogenations of Disubstituted Quinolines. Org. Lett. 2015, 17, 62666269,  DOI: 10.1021/acs.orglett.5b03307
      (c) Zhang, Z.; Du, H. A Highly cis-Selective and Enantioselective Metal-Free Hydrogenation of 2,3-Disubstituted Quinoxalines. Angew. Chem., Int. Ed. 2015, 54, 623626,  DOI: 10.1002/anie.201409471
      (d) Wei, S.; Feng, X.; Du, H. A Metal-Free Hydrogenation of 3-Substituted 2H-1,4-Benzoxazines. Org. Biomol. Chem. 2016, 14, 80268029,  DOI: 10.1039/C6OB01556E
    19. 19
      Tu, X.; Zeng, N.; Li, R.; Zhao, Y.; Xie, D.; Peng, Q.; Wang, X. C2-Symmetric Bicyclic Bisborane Catalysts: Kinetic or Thermodynamic Products of a Reversible Hydroboration of Dienes. Angew. Chem., Int. Ed. 2018, 57, 1509615100,  DOI: 10.1002/anie.201808289
    20. 20
      (a) Li, X.; Tian, J.; Liu, N.; Tu, X.; Zeng, N.; Wang, X. Spiro-Bicyclic Bisborane Catalysts for Metal-Free Chemoselective and Enantioselective Hydrogenation of Quinolines. Angew. Chem., Int. Ed. 2019, 58, 46644668,  DOI: 10.1002/anie.201900907
      (b) Tian, J.; Yang, Z.; Liang, X.; Liu, N.; Hu, C.; Tu, X.; Li, X.; Wang, X. Borane-Catalyzed Chemoselective and Enantioselective Reduction of 2-Vinyl-Substituted Pyridines. Angew. Chem., Int. Ed. 2020, 59, 1845218456,  DOI: 10.1002/anie.202007352
    21. 21
      Gao, B.; Feng, X.; Meng, W.; Du, H. Asymmetric Hydrogenation of Ketones and Enones with Chiral Lewis Base Derived Frustrated Lewis Pairs. Angew. Chem., Int. Ed. 2020, 59, 44984504,  DOI: 10.1002/anie.201914568
    22. 22

      For selected contributions, see:

      (a) Hermeke, J.; Mewald, M.; Oestreich, M. Experimental Analysis of the Catalytic Cycle of the Borane-Promoted Imine Reduction with Hydrosilanes: Spectroscopic Detection of Unexpected Intermediates and a Refined Mechanism. J. Am. Chem. Soc. 2013, 135, 1753717546,  DOI: 10.1021/ja409344w
      (b) Süsse, L.; Hermeke, J.; Oestreich, M. The Asymmetric Piers Hydrosilylation. J. Am. Chem. Soc. 2016, 138, 69406943,  DOI: 10.1021/jacs.6b03443
      (c) Mercea, D. M.; Howlett, M. G.; Piascik, A. D.; Scott, D.J.; Steven, A.; Ashley, A. E.; Fuchter, M. J. Enantioselective reduction of N-alkyl ketimines with frustrated Lewis pair catalysis using chiral borenium ions. Chem. Commun. 2019, 55, 70777080,  DOI: 10.1039/C9CC02900A
      (d) Liu, X.; Wang, Q.; Han, C.; Feng, X.; Du, H. Chiral Frustrated Lewis Pairs Catalyzed Highly Enantioselective Hydrosilylations of Ketones. Chin. J. Chem. 2019, 37, 663666,  DOI: 10.1002/cjoc.201900121
      (e) Lundrigan, T.; Welsh, E. N.; Hynes, T.; Tien, C.; Adams, M. R.; Roy, K. R.; Robertson, K. N.; Speed, A. W. H. Enantioselective Imine Reduction Catalyzed by Phosphenium Ions. J. Am. Chem. Soc. 2019, 141, 1408314088,  DOI: 10.1021/jacs.9b07293
    23. 23

      For computational mechanistic studies on FLP-type catalytic hydrogenations (not addressing the issue of stereoselectivity), see:

      (a) Rokob, T. A.; Hamza, A.; Stirling, A.; Pápai, I. On the Mechanism of B(C6F5)3-Catalyzed Direct Hydrogenation of Imines: Inherent and Thermally Induced Frustration. J. Am. Chem. Soc. 2009, 131, 20292036,  DOI: 10.1021/ja809125r
      (b) Nyhlén, J.; Privalov, T. On the Possibility of Catalytic Reduction of Carbonyl Moieties with Tris(Pentafluorophenyl)Borane and H2: A Computational Study. Dalton Trans. 2009, 29, 57805786,  DOI: 10.1039/b905137f
      (c) Privalov, T. The Role of Amine-B(C6F5)3 Adducts in the Catalytic Reduction of Imines with H2: A Computational Study. Eur. J. Inorg. Chem. 2009, 2009, 22292237,  DOI: 10.1002/ejic.200900194
      (d) Li, H.; Zhao, L.; Lu, G.; Huang, F.; Wang, Z.-X. Catalytic Metal-Free Ketone Hydrogenation: A Computational Experiment. Dalton Trans. 2010, 39, 55195526,  DOI: 10.1039/c001399d
      (e) Zhao, L.; Li, H.; Lu, G.; Huang, F.; Zhang, C.; Wang, Z.-X. Metal-Free Catalysts for Hydrogenation of Both Small and Large Imines: A Computational Experiment. Dalton Trans. 2011, 40, 19291937,  DOI: 10.1039/c0dt01297a
      (f) Zhao, L.; Lu, G.; Huang, F.; Wang, Z.-X. A Computational Experiment to Study Hydrogenations of Various Unsaturated Compounds Catalyzed by a Rationally Designed Metal-Free Catalyst. Dalton Trans. 2012, 41, 46744684,  DOI: 10.1039/c2dt12152b
      (g) Chernichenko, K.; Madarász, Á.; Pápai, I.; Nieger, M.; Leskelä, M.; Repo, T. A Frustrated-Lewis-Pair Approach to Catalytic Reduction of Alkynes to Cis-Alkenes. Nat. Chem. 2013, 5, 718723,  DOI: 10.1038/nchem.1693
      (h) Wang, Z.-X.; Zhao, L.; Lu, G.; Li, H. X.; Huang, F. Computational Design of Metal-Free Molecules for Activation of Small Molecules, Hydrogenation, and Hydroamination. Top. Curr. Chem. 2012, 332, 231266,  DOI: 10.1007/128_2012_385
      (i) Zhao, J.; Wang, G.; Li, S. Mechanistic Insights into the Full Hydrogenation of 2,6-Substituted Pyridine Catalyzed by the Lewis Acid C6F5(CH2)2B(C6F5)2. Dalton Trans. 2015, 44, 92009208,  DOI: 10.1039/C5DT00978B
      (j) Das, S.; Pati, S. K. On the Mechanism of Frustrated Lewis Pair Catalysed Hydrogenation of Carbonyl Compounds. Chem. - Eur. J. 2017, 23, 10781085,  DOI: 10.1002/chem.201602774
      (k) Heshmat, M.; Privalov, T. Carbonyl Activation by Borane Lewis Acid Complexation: Transition States of H2 Splitting at the Activated Carbonyl Carbon Atom in a Lewis Basic Solvent and the Proton-Transfer Dynamics of the Boroalkoxide Intermediate. Chem. - Eur. J. 2017, 23, 90989113,  DOI: 10.1002/chem.201700437
      (l) Mane, M. V.; Vanka, K. Less Frustration, More Activity-Theoretical Insights into Frustrated Lewis Pairs for Hydrogenation Catalysis. ChemCatChem 2017, 9, 30133022,  DOI: 10.1002/cctc.201700289
      (m) Heshmat, M.; Privalov, T. Computational Elucidation of a Role That Brønsted Acidification of the Lewis Acid-Bound Water Might Play in the Hydrogenation of Carbonyl Compounds with H2 in Lewis Basic Solvents. Chem. - Eur. J. 2017, 23, 1148911493,  DOI: 10.1002/chem.201700937
      (n) Heshmat, M.; Privalov, T. Theory-Based Extension of the Catalyst Scope in the Base-Catalyzed Hydrogenation of Ketones: RCOOH-Catalyzed Hydrogenation of Carbonyl Compounds with H2 Involving a Proton Shuttle. Chem. - Eur. J. 2017, 23, 1819318202,  DOI: 10.1002/chem.201702149
      (o) Das, S.; Pati, S. K. Unravelling the Mechanism of Tin-Based Frustrated Lewis Pair Catalysed Hydrogenation of Carbonyl Compounds. Catal. Sci. Technol. 2018, 8, 51785189,  DOI: 10.1039/C8CY01227J
    24. 24

      For recent reviews on concepts and challenges in computing stereoselectivities, see:

      (a) Hopmann, K. H. Quantum Chemical Studies of Asymmetric Reactions: Historical Aspects and Recent Examples. Int. J. Quantum Chem. 2015, 115, 12321249,  DOI: 10.1002/qua.24882
      (b) Peng, Q.; Duarte, F.; Paton, R. S. Computing Organic Stereoselectivity-from Concepts to Quantitative Calculations and Predictions. Chem. Soc. Rev. 2016, 45, 60936107,  DOI: 10.1039/C6CS00573J
      (c) Krenske, E. H. Challenges in Predicting Stereoselectivity. In Applied Theoretical Organic Chemistry; Tantillo, D. J., Ed.; World Scientific: New Jersey, 2018; pp 583604.
      (d) Gridnev, I. D.; Dub, P. A. Enantioselection in Asymmetric Catalysis; CRC Press: Boca Raton, 2017.
    25. 25

      For a detailed computational analysis of H2 activation pathways with the im/2 and P/2 pairs, see the SI (section 2.1).

    26. 26

      Computed proton affinities of PtBu3 and im are −277.5 and −231.7 kcal/mol, respectively. These values are obtained as solution phase Gibbs free energies of base → baseH+ reactions.

    27. 27

      For a detailed comparison of the energetics of the two catalytic cycles, see the SI (section 2.2).

    28. 28

      For conformational analysis of borohydride 2H, see the SI (section 2.3).

    29. 29

      For details of the conformational analysis carried out for the imH+/2H ion pair intermediate, see the SI (section 2.4).

    30. 30

      For related studies, see, for example:

      (a) Rokob, T. A.; Hamza, A.; Pápai, I. Rationalizing the Reactivity of Frustrated Lewis Pairs: Thermodynamics of H2 Activation and the Role of Acid–Base Properties. J. Am. Chem. Soc. 2009, 131, 1070110710,  DOI: 10.1021/ja903878z
      (b) Schulz, F.; Sumerin, V.; Heikkinen, S.; Pedersen, B.; Wang, C.; Atsumi, M.; Leskelä, M.; Repo, T.; Pyykkö, P.; Petry, W.; Rieger, B. Molecular Hydrogen Tweezers: Structure and Mechanisms by Neutron Diffraction, NMR, and Deuterium Labeling Studies in Solid and Solution. J. Am. Chem. Soc. 2011, 133, 2024520257,  DOI: 10.1021/ja206394w
      (c) Zaher, H.; Ashley, A. E.; Irwin, M.; Thompson, A. L.; Gutmann, M. J.; Krämer, T.; O’Hare, D. Structural and Theoretical Studies of Intermolecular Dihydrogen Bonding in [(C6F5)2(C6Cl5)B]–H···H–[TMP]. Chem. Commun. 2013, 49, 97559757,  DOI: 10.1039/c3cc45889j
      (d) Zhivonitko, V. V.; Sorochkina, K.; Chernichenko, K.; Kótai, B.; Földes, T.; Pápai, I.; Telkki, V.-V.; Repo, T.; Koptyug, I. Nuclear Spin Hyperpolarization with Ansa-Aminoboranes: A Metal-Free Perspective for Parahydrogen-Induced Polarization. Phys. Chem. Chem. Phys. 2016, 18, 2778427795,  DOI: 10.1039/C6CP05211H
    31. 31

      For the estimation of barriers relevant to the interconversion of imH+/2H isomers, see the SI (section 2.4).

    32. 32

      For computational studies examining the role of stabilizing noncovalent interactions in TM-catalyzed stereoselective hydrogenations, see:

      (a) Hopmann, K. H.; Bayer, A. On the Mechanism of Iridium-Catalyzed Asymmetric Hydrogenation of Imines and Alkenes: A Theoretical Study. Organometallics 2011, 30, 24832497,  DOI: 10.1021/om1009507
      (b) Václavík, J.; Kuzma, M.; Přech, J.; Kačer, P. Asymmetric Transfer Hydrogenation of Imines and Ketones Using Chiral RuIICl(η6-p-cymene)[(S,S)-N-TsDPEN] as a Catalyst: A Computational Study. Organometallics 2011, 30, 48224829,  DOI: 10.1021/om200263d
      (c) Wang, T.; Zhuo, L.-G.; Li, Z.; Chen, F.; Ding, Z.; He, Y.; Fan, Q.-H.; Xiang, J.; Yu, Z.-X.; Chan, A. S. C. Highly Enantioselective Hydrogenation of Quinolines Using Phosphine-Free Chiral Cationic Ruthenium Catalysts: Scope, Mechanism, and Origin of Enantioselectivity. J. Am. Chem. Soc. 2011, 133, 98789891,  DOI: 10.1021/ja2023042
      (d) Pablo, Ó.; Guijarro, D.; Kovács, G.; Lledós, A.; Ujaque, G.; Yus, M. A Versatile Ru Catalyst for the Asymmetric Transfer Hydrogenation of Both Aromatic and Aliphatic Sulfinylimines. Chem. - Eur. J. 2012, 18, 19691983,  DOI: 10.1002/chem.201102426
      (e) Hopmann, K. H. Iron/Brønsted Acid Catalyzed Asymmetric Hydrogenation: Mechanism and Selectivity-Determining Interactions. Chem. - Eur. J. 2015, 21, 1002010030,  DOI: 10.1002/chem.201500602
      (f) Tutkowski, B.; Kerdphon, S.; Limé, E.; Helquist, P.; Andersson, P. G.; Wiest, O.; Norrby, P.-O. Revisiting the Stereodetermining Step in Enantioselective Iridium-Catalyzed Imine Hydrogenation. ACS Catal. 2018, 8, 615623,  DOI: 10.1021/acscatal.7b02386
      (g) Salomó, E.; Gallen, A.; Sciortino, G.; Ujaque, G.; Grabulosa, A.; Lledós, A.; Riera, A.; Verdaguer, X. Direct Asymmetric Hydrogenation of N-Methyl and N-Alkyl Imines with an Ir(III)H Catalyst. J. Am. Chem. Soc. 2018, 140, 1696716970,  DOI: 10.1021/jacs.8b11547
      (h) Chen, J.; Gridnev, I. D. Size is Important: Artificial Catalyst Mimics Behaviour of Natural Enzymes. iScience 2020, 23, 100960,  DOI: 10.1016/j.isci.2020.100960
    33. 33

      For a selection of related studies on TM-catalyzed stereoselective hydrogenation of other substrates, see:

      (a) Hopmann, K. H. Cobalt–Bis(Imino)Pyridine-Catalyzed Asymmetric Hydrogenation: Electronic Structure, Mechanism, and Stereoselectivity. Organometallics 2013, 32, 63886399,  DOI: 10.1021/om400755k
      (b) Dub, P. a.; Henson, N. J.; Martin, R. L.; Gordon, J. C. Unravelling the Mechanism of the Asymmetric Hydrogenation of Acetophenone by [RuX2(Diphosphine)(1,2-Diamine)] Catalysts. J. Am. Chem. Soc. 2014, 136, 35053521,  DOI: 10.1021/ja411374j
      (c) Nakatsuka, H.; Yamamura, T.; Shuto, Y.; Tanaka, S.; Yoshimura, M.; Kitamura, M. Mechanism of Asymmetric Hydrogenation of Aromatic Ketones Catalyzed by a Combined System of Ru(π-CH2C(CH3)CH2)2(Cod) and the Chiral Sp2N/Sp3NH Hybrid Linear N4 Ligand Ph-BINAN-H-Py. J. Am. Chem. Soc. 2015, 137, 81388149,  DOI: 10.1021/jacs.5b02350
      (d) Dub, P. A.; Gordon, J. C. The Mechanism of Enantioselective Ketone Reduction with Noyori and Noyori–Ikariya Bifunctional Catalysts. Dalton Trans. 2016, 45, 67566781,  DOI: 10.1039/C6DT00476H
      (e) Nakane, S.; Yamamura, T.; Manna, S. K.; Tanaka, S.; Kitamura, M. Mechanistic Study of the Ru-Catalyzed Asymmetric Hydrogenation of Nonchelatable and Chelatable Tert-Alkyl Ketones Using the Linear Tridentate Sp3P/Sp3NH/Sp2N-Combined Ligand PN(H)N: RuNH- and RuNK-Involved Dual Catalytic Cycle. ACS Catal. 2018, 8, 1105911075,  DOI: 10.1021/acscatal.8b02671
    34. 34

      For structures of all HT transition states in the reaction with borane 2 and related structural analysis, see the SI (section 2.5).

    35. 35

      The HT step of the catalytic cycle of im hydrogenation is an irreversible process. For the reaction with borane 2, the am + 2 product state is predicted to be 16.1 kcal/mol below the most favored imH+/2H.

    36. 36

      The nearly quantitative agreement between computed and observed ee values is no doubt fortuitous and cannot be regarded as a measure of accuracy of the applied computational methodology.

    37. 37

      For structures of all HT transition states in the reaction with borane 3 and related structural analysis, see the SI (section 2.6).

    38. 38

      For structures of all HT transition states in the reaction with borane 1 and related structural analysis, see the SI (section 2.7).

    39. 39

      The RDG data were computed by using the NCIPLOT program:

      (a) Johnson, E. R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A. J.; Yang, W. Revealing Noncovalent Interactions. J. Am. Chem. Soc. 2010, 132, 64986506,  DOI: 10.1021/ja100936w
      (b) Contreras-García, J.; Johnson, E. R.; Keinan, S.; Chaudret, R.; Piquemal, J.-P.; Beratan, D. N.; Yang, W. NCIPLOT: A Program for Plotting Non-Covalent Interaction Regions. J. Chem. Theory Comput. 2011, 7, 625632,  DOI: 10.1021/ct100641a
    40. 40

      For details on the reaction with the simplified borane, see the SI (section 2.8).

    41. 41

      For details on computational analysis for hydrogenation reactions with boranes 2-F, 2-CF3, 2-CH3, 2-tBu, and 2-ant, see the SI (section 2.9).

    42. 42

      For the influence of various substituents on the strength of CH···π interaction, see:

      Karthikeyan, S.; Ramanathan, V.; Mishra, B. K. Influence of the Substituents on the CH···π Interaction: Benzene–Methane Complex. J. Phys. Chem. A 2013, 117, 66876694,  DOI: 10.1021/jp404972f
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acscatal.0c04263.

    • Details regarding the computational analysis, total energies and Cartesian coordinates for the considered stationary points, and experimental details (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.