Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Sep 9.
Published in final edited form as: J Am Chem Soc. 2015 Aug 26;137(35):11491–11497. doi: 10.1021/jacs.5b07201

Catalytic Kinetic Resolution of Disubstituted Piperidines by Enantioselective Acylation: Synthetic Utility and Mechanistic Insights

Benedikt Wanner , Imants Kreituss , Osvaldo Gutierrez , Marisa C Kozlowski ‡,*, Jeffrey W Bode †,*
PMCID: PMC4601563  NIHMSID: NIHMS724694  PMID: 26308097

Abstract

The catalytic kinetic resolution of cyclic amines with achiral N-heterocyclic carbenes (NHC) and chiral hydroxamic acids has emerged as a promising method to obtain enantioenriched amines with high selectivity factors. In this report, we describe the catalytic kinetic resolution of disubstituted piperdines with practical selectivity factors (s up to 52) in which we uncovered an unexpected and pronounced conformational effect resulting in disparate reactivity and selectivity between the cis and trans-substituted piperidine isomers. Detailed experimental and computational (DFT) studies of the kinetic resolution of various disubstituted piperidines revealed a strong preference for the acylation of conformers in which the α-substituent occupies the axial position. This work provides further experimental and computational support for the concerted 7-member transition state model for acyl transfer reagents and expands the scope and functional group tolerance of the secondary amine kinetic resolution.

INTRODUCTION

Disubstituted piperidines are increasingly attractive platforms for drug development, as the precise placement of substituents about the basic, low-molecular weight, three-dimensional scaffold is ideally suited for structure-activity relationship studies.1 Although the synthesis of such saturated N-heterocycles remains challenging,2 recent advances including diastereoselective metalation/cross-coupling,3 conjugate additions of dihydropyridinones,4 hydrogenation of disubstituted pyridines5 and cyclization methods6 are bringing these once exotic building blocks into the mainstream. Unfortunately, most of the available methods for the preparation of multisubstituted N-heterocycles deliver racemic products and prospects for enantioselective approaches to many substitution patterns are limited.7

For such challenging cases, kinetic resolution can be an efficient and powerful alternative to enantioselective synthesis of chiral N-heterocycles.8 Although kinetic resolution has the disadvantage of a maximum yield of 50% of a desired stereoisomer, for early stage applications it can be a reliable, effective, and easily implemented tool for the production of enantiopure material. Provided that selectivity factors (s) greater than 20 can be obtained, kinetic resolution can be used to separate enantiomers with acceptable yield and outstanding enantiopurity.9

In 2011, our group reported a new approach to the catalytic, kinetic resolution of saturated N-heterocycles consisting of two catalysts working synergistically.10 Our key advance was the use of a chiral hydroxamic acid, employed in either catalytic or stoichiometric amounts, as highly enantio– and chemoselective acylating agent, which operates by a unique mechanism involving concerted acyl and proton transfer via a 7-member transition state.11,12 To date, our studies on the kinetic resolution of chiral N-heterocycles using chiral hydroxamic acids have been limited to mono-, α-substituted N-heterocyclic substrates. In contrast to the resolution of α-monosubstituted N-heterocycles, for which only a single enantiomeric pair exists, the resolution of disubstituted piperidines requires an enantioselective catalyst to discrimitate among several different ring conformations (Scheme 1).

Scheme 1.

Scheme 1

Four possible positional isomers of disubstituted piperidines that serve as targets for our study.

In this report, we examine the reactivity and selectivity of disubstituted piperidines in the kinetic resolution by means of enantioselective acylation with a chiral hydroxamic acid. Kinetic resolution studies and density functional theory (DFT) calculations exposed remarkable conformational preferences in both reactivity and selectivity, revealing a requirement for axial-α substitution on the reactive form of the N-heterocycle for effective resolution. This preference extends even to substitution patterns where the substituents must occupy bis-axial positions. This work provides, 1) access to valuable enantioenriched disubstituted piperidines, 2) general guidelines for selecting suitable N-heterocycles as substrates for the kinetic resolution via chiral hydroxamic acids, and 3) affirmation of the transition state model involving a high energy conformation of the substrate that we had previously developed on the basis of DFT calculations.

RESULTS AND DISCUSSION

As depicted in Scheme 1, there are four types of disubstituted piperidines that could serve as target substrates for our study. Piperidines without α-substituents, such as 3,5-disubstituted piperidines, undergo acylation in an unselective manner. Further, 2,6-disubstituted substrates were found to be unreactive and were excluded from further studies.

Our work therefore targeted the six isomers arising from 2,3-, 2,4-, and 2,5-disubstitution of piperidine, with cis and trans stereoisomers in each case. Further disubstituted compounds with an exocyclic 4-methylene group were also prepared and evaluated. The results of the catalytic kinetic resolution are summarized in Table 1. Surprisingly, we observed substantial differences in the reactivity and selectivity between the cis- and trans-isomers of the otherwise identical positional isomers. For the 2,3- (Table 1, entries 1 and 2, entries 7 and 8) and 2,5- (Table 1, entries 5 and 6) disubstituted piperidines the cis-isomer underwent faster reactions with higher selectivity, while the trans-isomer gave rise to superior outcomes for the 2,4- (Table 1, entries 3 and 4) disubstituted compounds. The decahydroquinoline isomers were particularly interesting, as the trans isomer is conformationally locked. These results already hinted at strict stereochemical requirements for effective enantioselective acylation.

Table 1.

Catalytic Kinetic resolution of disubstituted piperidines.

graphic file with name nihms724694u1.jpg
Entry Substrate sa Conv.b (%) reaction timec (h) er amined yielde (%) er amided yielde (%)
1 graphic file with name nihms724694t1.jpg 11 28 72 64:36 (43) 89:11 (25)
2 graphic file with name nihms724694t2.jpg 2 3 72 51:49 (58) 67:33 (5)
3 graphic file with name nihms724694t3.jpg 3 4 48 51:49 (43) 78:22 (4)
4 graphic file with name nihms724694t4.jpg 21 59 20 99:1 (28) 83:17 (26)
5 graphic file with name nihms724694t5.jpg 19 17 20 59:41 (35) 94:6 (20)
6 graphic file with name nihms724694t6.jpg 8 5 48 52:48 (54) 89:11 (9)
7 graphic file with name nihms724694t7.jpg 18 63 48 99:1 (20) 79:21 (40)
8f graphic file with name nihms724694t8.jpg 9 11 65 45:55 (61) 10:90 (9)
9 graphic file with name nihms724694t9.jpg 19 43 20 81:19 (40) 91:9 (42)
10 graphic file with name nihms724694t10.jpg 14 41 20 77:23 (43) 89:11 (40)
11 graphic file with name nihms724694t11.jpg 21 55 20 96:4 (42) 87:13 (43)
a

Calculated selectivity.15

b

Calculated conversion.15

c

Reaction time until the α-hydroxyenone is fully consumed.

d

Determined by SFC or HPLC on a chiral support.

e

Isolated as the Cbz-derivative. Yield after column chromatography.

f

Opposite sense of induction observed. RMesClO4 = (2,4,6-trimethylphenyl)-2,5,6,7-tetrahydro-pyrrolo[2,1-c] [1,2,4]triazol- 4-ylium perchlorate.

In light of these initially confusing results, we undertook an extensive study to assess the role of the positional isomers and their stereoisomers. Our goal was to reliably predict whether a given substrate would undergo a selective kinetic resolution. We elected to evaluate at least three substrate pairs (cis/trans) for each of the three positional isomers, giving 20 substrates in total. Considerations for compound selection included ease of synthesis, the presence of groups for further functionalization of the products, and ready access to diastereomerically pure forms of the targets (See Supporting Information for experimental procedures and characterization data).

To rule out selectivity arising from the NHC-catalyst in the kinetic resolution, we used our previously reported stoichiometric variant in the resolution of these N-heterocycles.10 Slightly higher reaction rates and selectivities are observed with this reagent and a recyclable, resin supported variant makes it an attractive option for practical amine resolution.13 We first examined the kinetic resolution of the compounds in the 2,3-disubstitution series (Table 2). The data presented in Table 1 implied that substrates with a cis configuration undergo faster reaction with much higher selectivities than those observed for the trans isomers. With cis-2-phenylpiperidin-3-ol (Table 2, entry 1) an s-factor of 24 and a conversion of 33% could be obtained after 48 hours, whereas the trans diastereomer (Table 2, entry 2) reached only 14% conversion after 72 hours with essentially no selectivity. This trend could also be observed for the other two sets of substrates (Table 2, entries 3 and 4, entries 5 and 6). Decahydroquinolines showed a similar behavior in terms of reaction rate, although acceptable selectivity was obtain in both cases. In the resolution of cis-decahydroquinoline (Table 2, entry 7), 65% conversion was obtained after 15 hours whereas the conformationally locked trans-decahydroquinoline (Table 2, entry 8) reacted much more slowly. Notably, the opposite sense of induction was observed with this substrate, which contradicted a simple stereochemical model bases only on the stereochemistry of the acyl transfer reagent (vide infra). The kinetic resolution reactions were stopped once the stoichiometric reagent was fully consumed.14

Table 2.

Kinetic resolution of 2,3-disubstituted piperidines with stoichiometric reagent 1.

graphic file with name nihms724694u2.jpg
Entry Substrate sa Conv.b (%) reaction timec (h) er amined yielde (%) er amided yielde (%)
1 graphic file with name nihms724694t12.jpg 24 33 48 72:28 (31) 94:6 (30)
2 graphic file with name nihms724694t13.jpg 1 14 72 51:49 (32) 56:44 (5)
3 graphic file with name nihms724694t14.jpg 19 31 72 69:31 (34) 93:7 (28)
4 graphic file with name nihms724694t15.jpg 2 46 96 56:44 (39) 57:43 (7)
5 graphic file with name nihms724694t16.jpg 23 50 72 90:10 (39) 91:9 (50)
6 graphic file with name nihms724694t17.jpg 4 26 96 59:41 (40) 75:25 (21)
7f graphic file with name nihms724694t18.jpg 20 65 15 99:1 (17) 73:27 (50)
8g graphic file with name nihms724694t19.jpg 20 36 48 24:76 (25) 7:93 (30)
a

Calculated selectivity.

b

Calculated conversion.

c

Reaction time until stoichiometric reagent is fully consumed.

d

Determined by SFC or HPLC on a chiral support.

e

Isolated as the Cbz-derivative. Yield after column chromatography.

f

0.6 equiv of stochiometric reagent 1 were used.

g

Opposite sense of induction observed.

For the 2,4-disubstituted piperidines (Table 3), the trend was reversed in comparison to that observed for the 2,3-disubstituted examples in Table 2. For this series, the cis-diastereomers exhibit poor selectivities (s = 3–7) and low reactivity with reaction times between 72 and 120 hours (Table 3, entries 1, 3, and 5). In contrast, the trans-diastereomers reached full conversion in 20 hours or less with high selectivities (s = 10–29) (Table 3, entries 2, 4, and 6).

Table 3.

Kinetic resolution of 2,4-disubstituted piperidines with stoichiometric reagent 1.

graphic file with name nihms724694u3.jpg
Entry Substrate sa Conv.b (%) reaction timec (h) er amined yielde (%) er amided yielde (%)
1 graphic file with name nihms724694t20.jpg 3 22 120 57:43 (39) 78:22 (4)
2f graphic file with name nihms724694t21.jpg 10 65 20 97:3 (15) 75:25 (43)
3 graphic file with name nihms724694t22.jpg 7 32 72 66:34 (56) 84:16 (29)
4 graphic file with name nihms724694t23.jpg 29 52 20 94:6 (46) 91:9 (45)
5 graphic file with name nihms724694t24.jpg 6 38 96 68:32 (46) 80:20 (34)
6f graphic file with name nihms724694t25.jpg 15 62 20 98:2 (19) 80:20 (40)
a

Calculated selectivity.

b

Calculated conversion.

c

Reaction time until stoichiometric reagent is fully consumed.

d

Determined by SFC or HPLC on a chiral support.

e

Isolated as the Cbz-derivative. Yield after column chromatography.

f

0.6 equiv of stochiometric reagent 1 were used.

In the case of the 2,5-disubstituted piperidines (Table 4) we observed a similar trend as with the 2,3-disubstituted substrates. Namely, higher s-factors were observed for the cis-diastereomers as well as higher reaction rates compared to the corresponding trans-diastereomers. The biggest contrast with regards to s-factors can be noticed in entry 5; for cis-6-propylpiperidin-3-ol an s-factor of 52 was obtained while the trans-isomer (Table 4, entry 6) afforded an s-factor of only 4. Such large differences in selectivity were not evident in entries 1–4; however, the data supports the conjecture that cis isomers in this series are superior to the trans isomers. These substrates also demonstrate the chemoselectivity of the kinetic resolution for N-heterocycles bearing a variety of functional groups suitable for further functionalization.

Table 4.

Kinetic resolution of 2,5-disubstituted piperidines with stoichiometric reagent 1.

graphic file with name nihms724694u4.jpg
Entry Substrate sa Conv.b (%) reaction timec (h) er amined yielde (%) er amided yielde (%)
1 graphic file with name nihms724694t26.jpg 22 42 24 81:19 (44) 92:8 (33)
2 graphic file with name nihms724694t27.jpg 20 40 96 78:22 (47) 92:8 (33)
3 graphic file with name nihms724694t28.jpg 13 54 72 90:10 (22) 84:16 (28)
4 graphic file with name nihms724694t29.jpg 9 29 72 65:35 (45) 87:13 (19)
5 graphic file with name nihms724694t30.jpg 52 54 20 99:1 (25) 92:2 (38)
6 graphic file with name nihms724694t31.jpg 4 53 48 74:26 (29) 71:29(51)
a

Calculated selectivity.

b

Calculated conversion.

c

Reaction time until stoichiometric reagent is fully consumed.

d

Determined by SFC or HPLC on a chiral support.

e

Isolated as the Cbz-derivative. Yield after column chromatography.

The divergent reactivity of the diastereomeric pairs in the different series can be rationalized on the basis of our recent DFT calculations of the transition states of α-methyl piperidine reacting with the chiral acylating agent.12 These studies concluded that acylation occurred via a concerted 7-membered transition state with concomitant proton transfer guided by the hydroxamic acid and with the α-substituent placed in the axial position. In contrast, other work from our group on the resolution of morpholines with stoichiometric chiral acyl donors and achiral hydroxamic acids was most consistent with a stereochemical model featuring an equatorial substituent of a morpholine.16 However, no computational studies were performed on this system and significant conformational differences between morpholine and piperidines can be expected. Both models suggest that the stereoselective step involves a seven-membered transition state with the lone pair of the nitrogen in the equatorial position. This mechanism of acyl transfer was found to be general for a variety of widely used acyl transfer reagents including N-hydroxysuccinimide, and HOAt.

In order to reconcile these two models with the results obtained for the resolution of disubstituted piperidines, we undertook detailed DFT studies on the relevant transition states using the cis and trans geometric isomers of 2,3-dimethylpiperidine as a model system and the chiral hydroxamic acid.17 All calculations were carried out using the same computational methods previously employed in related systems using Gaussian 09.18 Structures were optimized with B3LYP/6-31G(d).19,20 Single-point energy calculations in the condensed phase (CH2Cl2; ε = 8.93) were undertaken using the IEPCM solvation model with M06-2X/6-311+G(d,p).21 Scheme 2 shows the lowest energy transition states for acyl transfer to each enantiomer of cis and trans 2,3-dimethyl piperidine. As anticipated from our previous work, the concerted 7-membered transition states feature an intramolecular proton shuttle between the secondary amine and the hydroxamate carbonyl. The calculations are in agreement with experimental rates and selectivity trends listed in Table 2 (entries 1–6) in which the cis 2,3-disubstituted piperidines showed significantly faster rates and superior s-factors (s = up to 24) in comparison to the trans 2,3-disubstituted piperidines. For example, the lowest transition state free energy for the II-cis (25.8 kcal/mol gas phase) is approximately 3 kcal/mol lower in energy than the lowest for II-trans (29.1 kcal/mol gas phase), which is consistent with the relative rates of these two amine diastereomers.

Scheme 2. Relative acyl transfer reaction barriers.

Scheme 2

Relative acyl transfer reaction barriers (free energies are in kcal/mol; gas phase energetics calculated using B3LYP/6-31G(d,p)//(in parenthesis IEPCM-CH2Cl2-M06-2X/6-311+G(d,p)) for the competing diastereomeric transtion states between cis and trans 2,3- dimethyl substituted piperidines and the acyl transfer agent.

This finding is in line with our postulate that the lower energy transition states for acyl transfer feature an axial α-substituent, which requires the trans-2,3-dimethyl piperidine to adopt an unfavorable di-axial conformation. In contrast, the chair conformation of the cis-2,3-dimethylpiperidine with an α-axial substituent is the preferred ground state conformation. Further, our calculations correctly predict the fastest reacting enantiomer via II-trans-TS and II-cis-TS, respectively. In agreement with experiment, the quantum mechanical calculations indicate a larger gap between the cis-1,2-disubsituted systems (here modeled as 2,3-dimethyl piperidine II-cis and ent-II-cis) than the trans-1,2-disubsituted systems, which reflects the significant difference in selectivity between cis and trans diastereomers.22 Notably, the lowest energy diastereomeric transition states place the α-substituent away from both the aromatic ring of the acyl transfer agent and the acyl chain. Exhaustive conformational analysis, including other chair conformations, revealed that all other transition states are much higher in energy (see Figure 1 and Supporting Information). For kinetic resolution, the energy difference between the diastereomeric acylation transition states determines the selectivity. As shown in Figure 1, the significant energy difference (3.7 kcal/mol, s = 19–24) between II-cis-TS and ent-II-cis-TS arises from the additional unfavorable gauche interactions between α-substituent in the equatorial position and the carbonyl group. In addition, ent-II-cis-TS has an further interaction between the axial β-methyl group and the hydroxamate carbonyl.

Figure 1.

Figure 1

Proposed reaction pathways for possible conformers/enantiomers in the kinetic resolution. Relative acyl transfer reaction barriers (free energies are in kcal/mol; gas phase energetics calculated using B3LYP/6-31G(d,p)//(in parenthesis IEPCM-CH2Cl2-M06-2X/6-311+G(d,p)) for the competing diastereomeric transtion states between cis and trans 2,3- dimethyl substituted piperidines and the acyl transfer agent.

In contrast, the energy difference between II-trans-TS and ent-II-trans-TS is small at 0.5 kcal/mol, which is consistent with the poor s-factors (s = 1–4) observed for the trans substituted systems. It is remarkable, however, that the preferred conformation for the lowest energy transition state is the trans-2,3-bis-axial, which demonstrates again the strong preference for α-axial substituents in the acylation step and offers an excellent example of the Curtin-Hammett principle (Figure 1, bottom). The dimethyl groups are in the axial position in II-trans-TS and in the equatorial position in ent-II-trans-TS; in this case, the additional 1,3-diaxial interactions in II-trans-TS raise its energy leading to a smaller energy gap between the two enantiomers and smaller selectivity values.

The disparate results seen for the decahydroquinolines serve to further confirm this model. The trans-decahydroquinoline leads to the opposite sense of induction relative to the cis-decahydroquinoline even though the same configuration of the chiral acylating reagent is employed. Since the trans-decahydroquinoline cannot react in the conformation where the α-substituent is axial, the only transition states that are acessible have the α-substituent in equatorial position (Figure 1, lower right). When only these two transition states are considered, selectivitey is expected for the opposite enantiomer due to the large calculated energy gap. Further, the reaction rate of trans-decahydroquinoline is expected to be slower, as observed experimentally, since the overall, energy of this pathway is higher.

The same results can be applied to 2,5-disubstituted piperidines. Just as with the 2,3-disubstituted compounds, the cis-diastereomer consists of two conformers that differ only slightly in energy whereas the trans-diastereomer exists primarily in the diequatorial conformer rather than the higher energy diaxial conformation. An analogous situation occurs for the 2,4-disubstituted compounds; however, it is the trans isomer that always has one axial substituent in the lowest energy conformations, and is more selective. The cis-2,4-disubstituted isomers, in contrast, exist almost exclusively in the more favorable di-equatorial conformation and gives rise to lower selectivity.

This analysis suggests that disubstituted piperdines with a single chiral center and no large conformational preferences between the axial and equatorial α-substituents should be good substrates for the resolution. This hypothesis is confirmed by the excellent results obtained with substrates bearing an exocyclic 4-methylene and a single α-substituent (Table 5, entries 1–3). In addition 2,4,4-trisubstituted piperidines containing a cyclic ketal were also resolved with modest selectivity (Table 5, entries 4 and 5). All five substrates evaluated show fast reaction rates and useful selectivity, indicating no adverse affect on the resolution of substrates containing additional achiral substitution. The alkene or protected carbonyl in the products provide a means for post-resolution derivatization.

Table 5.

Kinetic resolution of 2,4,4-trisubstituted piperidines with stoichiometric reagent 1.

graphic file with name nihms724694u5.jpg
Entry Substrate sa Conv.b (%) reaction timec (h) er amined yielde (%) er amided yielde (%)
1 graphic file with name nihms724694t32.jpg 27 51 20 92:8 (45) 91:9 (43)
2 graphic file with name nihms724694t33.jpg 18 49 20 88:12 (46) 89:11 (43)
3f graphic file with name nihms724694t34.jpg 18 56 20 95: 5 (39) 86:14 (54)
4 graphic file with name nihms724694t35.jpg 8 48 20 79:21 (25) 81:19 (44)
5f graphic file with name nihms724694t36.jpg 9 64 20 94:6 (23) 75:25 (21)
a

Calculated selectivity.

b

Calculated conversion.

c

Reaction time until stoichiometric reagent is fully consumed.

d

Determined by SFC or HPLC on a chiral support.

e

Isolated as the Cbz-derivative. Yield after column chromatography.

f

0.6 equiv of stochiometric reagent 1 were used.

In summary we have examined the kinetic resolution of disubstituted piperidines using a chiral hydroxamic acid catalyst and its corresponding stoichiometric reagent. For most cases, synthetically useful relative rates and outstanding functional group tolerance were observed. The exceptions – stereoisomeric substrates that gave poor reactivity and selectivity – revealed a pronounced conformational preference in which the lowest energy transition states require that the α-substituent populates the axial position. This conjecture was fully supported by DFT calculations on the transition states for all of the relevant stereoisomers and ring conformations. These finding both rationalized the initially unexpected differences in reactivity and selectivity between cis and trans stereoisomers and strengthen our stereochemical model.

Supplementary Material

spectra
xray

Acknowledgments

This work was supported by the European Research Council (ERC Starting Grant No. 306793 − CASAA) and National Institutes of Health (GM-112684 to M.C.K.) and the National Science Foundation (CHE1464778 to M.C.K.). Computational support was provided by XSEDE on SDSC Gordon (TG-CHE120052). Computational support was provided by the XSEDE on SDSC Gordon (TG-CHE120052). We thank LOC MS Service (ETH Zürich) for analyses and Dr. Nils Trapp and Dr. Bernd Schweizer for the acquisition of X-ray structures. We are also grateful to Dr. Kimberly Geoghegan, Sheng-Ying Hsieh and Dr. Hidetoshi Noda for helpful discussions.

Footnotes

Supporting Information

Experimental procedures and spectroscopic data for all new compounds. Transition state structures of all minimized TSs. This material is available free of charge via the Internet at http://pubs.acs.org.

Notes

The authors declare no competing financial interest.

References

  • 1.Lovering F, Bikker J, Humblet C. J Med Chem. 2009;52:6752–6756. doi: 10.1021/jm901241e. [DOI] [PubMed] [Google Scholar]
  • 2.For reviews on the a functionalisation of unsubstituted heterocycles see:; (a) Mitchell EA, Peschiulli A, Lefevre N, Meerpoal L, Maes BUW. Chem Eur J. 2012;18:10092–10142. doi: 10.1002/chem.201201539. [DOI] [PubMed] [Google Scholar]; (b) Campos KR. Chem Soc Rev. 2007;36:1069–1084. doi: 10.1039/b607547a. [DOI] [PubMed] [Google Scholar]; (c) Vo CT, Bode JW. J Org Chem. 2014;79:2809–2815. doi: 10.1021/jo5001252. [DOI] [PubMed] [Google Scholar]
  • 3.(a) Millet A, Larini P, Clot E, Baudoin O. Chem Sci. 2013;4:2241–2247. [Google Scholar]; (b) Cordier CJ, Lundgren RJ, Fu GC. J Am Chem Soc. 2013;135:10946–10949. doi: 10.1021/ja4054114. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) McNally A, Prier CK, MacMillan DWC. Science. 2011;334:1114–1117. doi: 10.1126/science.1213920. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.(a) Comins DL, Abdullah AH. J Org Chem. 1982;47:4315–4319. [Google Scholar]; (b) Chau ST, Lutz JP, Wu K, Doyle AG. Angew Chem Int Ed. 2013;52:9153–9156. doi: 10.1002/anie.201303994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.(a) Cai W, Colony JL, Frost H, Hudspeth JP, Kendall PM, Krishnan AM, Maskowski T, Mazur DJ, Phillips J, Brown Ripin DH, Gut Ruggeri S, Stearns JF, White TD. Org Process Res Dev. 2005;9:51–56. [Google Scholar]; (b) Ye Z-S, Chen M-W, Chen Q-A, Shi L, Duan Y, Zhou Y-G. Angew Chem Int Ed. 2012;51:10181–10184. doi: 10.1002/anie.201205187. [DOI] [PubMed] [Google Scholar]; (c) Wang D-S, Chen Q-A, Lu S-M, Zhou Y-G. Chem Rev. 2012;112:2557–2590. doi: 10.1021/cr200328h. [DOI] [PubMed] [Google Scholar]; (d) Iimuro A, Yamaji K, Kandula S, Nagano T, Kita Y, Mashima K. Angew Chem Int Ed. 2013;52:2046–2050. doi: 10.1002/anie.201207748. [DOI] [PubMed] [Google Scholar]; (e) Romanov-Michailidis F, Sedillo KF, Neely JM, Rovis T. J Am Chem Soc. 2015;137:8892–8895. doi: 10.1021/jacs.5b04946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.(a) Pablo O, Guijarro D, Yus M. J Org Chem. 2013;78:9181–9189. doi: 10.1021/jo4014386. [DOI] [PubMed] [Google Scholar]; (b) Launay GG, Slawin AMZ, O’Hagan D. Beilstein J Org Chem. 2010;6:41. doi: 10.3762/bjoc.6.41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.(a) Chen S, Mercado BQ, Bergman RG, Ellman JA. J Org Chem. 2015;80:6660–6668. doi: 10.1021/acs.joc.5b00816. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Duttwyler S, Chen S, Lu C, Mercado BQ, Bergman RG, Ellman JA. Angew Chem Int Ed. 2014;53:3877–3880. doi: 10.1002/anie.201310517. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Ischay M, Takase MK, Bergma RG, Ellman JA. J Am Chem Soc. 2013;135:2478–2481. doi: 10.1021/ja312311k. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Duttwyler S, Chen S, Takase MK, Wiberg KB, Bergman RG, Ellman JA. Science. 2013;339:678–682. doi: 10.1126/science.1230704. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.(a) Mittal N, Lippert KM, Kanta De C, Klauber EG, Emge TJ, Schreiner PR, Seidel D. J Am Chem Soc. 2015;137:5748–5758. doi: 10.1021/jacs.5b00190. [DOI] [PubMed] [Google Scholar]; (b) Kanta De C, Klauber EG, Seidel D. J Am Chem Soc. 2009;131:17060–17061. doi: 10.1021/ja9079435. [DOI] [PubMed] [Google Scholar]; (c) Klauber EG, Kanta De C, Shah TK, Seidel D. J Am Chem Soc. 2010;132:13624–13626. doi: 10.1021/ja105337h. [DOI] [PubMed] [Google Scholar]; (d) Yang X, Bumbu VD, Liu P, Li X, Jiang H, Uffman EW, Guo L, Zhang W, Jiang X, Houk KN, Birman VB. J Am Chem Soc. 2012;134:17605–17612. doi: 10.1021/ja306766n. [DOI] [PubMed] [Google Scholar]; (e) Fu GC. Acc Chem Res. 2000;33:412–420. doi: 10.1021/ar990077w. [DOI] [PubMed] [Google Scholar]; (f) Fu GC. Acc Chem Res. 2004;37:542–547. doi: 10.1021/ar030051b. [DOI] [PubMed] [Google Scholar]; (g) Fowler BS, Mikochick PJ, Miller SJ. J Am Chem Soc. 2010;132:2870–2871. doi: 10.1021/ja9107897. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Birman VB, Uffman EW, Jiang H, Li X, Kilbane CJ. J Am Chem Soc. 2004;126:12226–12227. doi: 10.1021/ja0491477. [DOI] [PubMed] [Google Scholar]; (i) Li X, Liu P, Houk KN, Birman VB. J Am Chem Soc. 2008;130:13836–13837. doi: 10.1021/ja805275s. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Brown MK, Blewett MM, Colombe JR, Corey EJ. J Am Chem Soc. 2010;132:11165–11170. doi: 10.1021/ja103103d. [DOI] [PMC free article] [PubMed] [Google Scholar]; (k) Larionov E, Mahesh M, Spivey AC, Wei Y, Zipse H. J Am Chem Soc. 2012;134:9390–9399. doi: 10.1021/ja302420g. [DOI] [PubMed] [Google Scholar]
  • 9.(a) Keith JM, Larrow JF, Jacobsen EN. Adv Synth Catal. 2001;343:5–26. [Google Scholar]; (b) Breuer M, Ditrich K, Habicher T, Hauer B, Keßeler M, Stürmer R, Zelinski T. Angew Chem. 2004;116:806–843. doi: 10.1002/anie.200300599. Angew. Chem. Int. Ed2004, 43, 788–824. [DOI] [PubMed] [Google Scholar]
  • 10.Binanzer M, Hsieh SY, Bode JW. J Am Chem Soc. 2011;133:19698–19701. doi: 10.1021/ja209472h. [DOI] [PubMed] [Google Scholar]
  • 11.Hsieh S-Y, Binanzer M, Kreituss I, Bode JW. Chem Commun. 2012;48:8892–8894. doi: 10.1039/c2cc34907h. [DOI] [PubMed] [Google Scholar]
  • 12.Allen SE, Hsieh S-Y, Gutierrez O, Bode JW, Kozlowski MC. J Am Chem Soc. 2014;136:11783–11791. doi: 10.1021/ja505784w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Kreituss I, Murakami Y, Binanzer M, Bode JW. Angew Chem Int Ed. 2012;51:10660–10663. doi: 10.1002/anie.201204991. [DOI] [PubMed] [Google Scholar]
  • 14.For the purposes of method development and study we performed the kinetic resolutions under conditions that allow for accurate calculations of the relative rates. At longer reaction times, the hydroxyenone and stochiometric reagent can undergo side reactions. This protocol explains the low conversion even with full consumption of the reagent with slow-reacting substrates. For preparative work, the resolutions are allowed to proceed to higher conversion, which affords virtually enantiopure recovered amines. As a guideline, s = 20 can deliver >99% ee at 62% conversion. See; Vedejs E, Jure M. Angew Chem. 2005;117:4040–4069. doi: 10.1002/anie.200460842. Angew. Chem. Int. Ed.200544, 3974–4001 for a discussion. [DOI] [PubMed] [Google Scholar]
  • 15.Calculated according to:; Kagan HB, Fiaud JC. Topic in Stereochemistry. 1988;18:249–330. [Google Scholar]
  • 16.Hsieh S-Y, Wanner B, Wheeler P, Beauchemin AM, Rovis T, Bode JW. Chem Eur J. 2014;20:7228–7231. doi: 10.1002/chem.201402818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.All calculcations were carried out using Gaussian 09, Revision D.01
  • 18.Frisch MJ. Gaussian 09, Revision D.01. Gaussian, Inc; Wallingford CT: 2013. Gaussian, Inc.; Wallingford, CT, 2013. [Google Scholar]
  • 19.(a) Becke AD. J Chem Phys. 1993;98:5648–5652. [Google Scholar]; (b) Lee C, Yang W, Parr RG. Phys Rev B. 1988;37:785–789. doi: 10.1103/physrevb.37.785. [DOI] [PubMed] [Google Scholar]; (c) Vosko SH, Wilk L, Nusair M. Can J Phys. 1980;58:1200. [Google Scholar]
  • 20.(a) Ditchfield R, Hehre WJ, Pople JA. J Chem Phys. 1971;54:724–728. [Google Scholar]; (b) Frisch MJ, Pople JA, Binkley JS. J Chem Phys. 1984;80:3265–3269. [Google Scholar]
  • 21.(a) Cancès E, Mennucci B, Tomasi J. J Chem Phys. 1997;107:3032–3041. [Google Scholar]; (b) Zhao Y, Truhlar DG. Theor Chem Acc. 2008;120:215–241. [Google Scholar]
  • 22.Additional calculations using cis and trans-2-phenyl-3-hydroxypiperidine showed a similar trend as the model, cis and trans 2,3-dimethylpiperidine. For instance the enthalpic energy difference (in solvent) between cis-2-phenylpiperidin-3-ol was computed to be 1.7 kcal/mol vs only 0.4 kcal/mol for the corresponding trans diastereomer. This calculation is in excellent agreement with experiment for kinetic resolution of the cis and trans-2- phenylpiperidin-3-ol systems (Table 2, entries 1 and 2) which exhibited a s-values of 24 and 1, respectively.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

spectra
xray

RESOURCES