Abstract
Asymmetric Michael additions are powerful tools to meet the growing need for stereochemically complex products. While 1,3-dicarbonyls are common nucleophiles, the successful use of configurationally unstable β-keto esters in diastereoselective variants remains understudied. In this Letter, crystalline β-keto esters were leveraged in a two-phase, one-pot merger of an asymmetric Michael addition with a crystallization-induced diastereomer transformation. Tuning the crystallinity of β-keto ester adducts enabled stereoconvergence of the products, which were isolated by filtration.
Graphical Abstract

Stereochemical complexity is a key feature in many products from agriscience to pharmaceuticals, where a higher number of stereogenic centers has been linked to improvements in environmental safety and outcomes in clinical drug trials.1–5 To meet the growing demand for stereochemically complex compounds, chemists regularly turn to asymmetric catalysis. Of the wealth of methods available, one of the most powerful is the enantioselective Michael addition,6–13 an atom-efficient carbon-carbon (C-C) bond forming reaction that produces complex products from simple, inexpensive starting materials.
1,3-Dicarbonyls, whose low pKa values enable facile deprotonation under mild conditions by a variety of chiral catalysts, are prototypical pronucleophiles in such processes. Symmetrical 1,3-dicarbonyls have been used to synthesize enantioenriched compounds with β-stereogenic centers14 (orange circle, Figure 1). However, Michael additions with unsymmetrical 1,3-dicarbonyls typically yield mixtures of diastereomers15–17 since the acidic α-proton results in facile epimerization of the second stereogenic center (gold circle, Figure 1), eroding any kinetic selectivity provided by the chiral catalyst. A small number of diastereoenriched outliers exist,15,18 but general methods remain rare. A few approaches achieve high diastereoselectivity but must prevent epimerization with cryogenic temperatures,19,20 careful catalyst tuning,21 or incorporation of an α-substituent on the 1,3-dicarbonyl nucleophile.22–24 This prevailing strategy has largely ignored unsubstituted β-keto ester nucleophiles, whose increased acidity poses challenges for approaches relying solely on kinetic selectivity (pKa = 13–15 in DMSO vs pKa 18 for related β-keto amide nucleophiles).21,25 To complement current strategies that seek to avoid epimerization, we sought to productively leverage the acidity of the products to access the Michael adducts as single stereoisomers through crystallization-based stereocontrol.
Figure 1.

Previous strategies to access diastereoenriched Michael adducts require restricted epimerization.
Crystallization-induced diastereomer transformations (CIDTs) have been successfully merged with asymmetric catalysis in a two-phase approach that achieves excellent stereocontrol and yields complex products.26–29 In the first phase, a chiral catalyst enables skeletal construction while lowering the activation energy of one enantiodetermining transition state, creating a kinetic preference for one configuration at the new stereogenic center (i.e., β-carbon, Figure 2a). After C-C bond formation, a CIDT enables the thermodynamic driving force of crystallization to funnel an equilibrating mixture of diastereomers into a single crystalline stereoisomer.30–32 The two phases occur in series in one pot and generate products that can be isolated by the direct filtration of the reaction mixture. By avoiding solvent-intensive chromatography, a CIDT can decrease the production cost of stereochemically complex products on industrial scale.33,34
Figure 2.

Merging an asymmetric Michael addition with a CIDT and tuning product crystallinity promotes diastereoselectivity.
(Diamine)2Ni(II)-catalyzed Michael additions of β-keto esters to nitrostyrenes afforded crystalline products with excellent enantioselectivity but no diastereoselectivity.17 We wondered if the synthetically flexible β-keto ester35,36 could be carefully chosen to tune the Michael adducts’ crystallinity and promote a CIDT37 to achieve diastereoselectivity (Figure 2b). In this Letter, we report the asymmetric Michael addition of crystalline β-keto esters to nitroolefins merged with a CIDT to access stereochemically complex products by filtration in up to 88% yield with excellent enantio- and diastereoselectivity.
We began by examining the enantioselective Michael addition of a crystalline β-keto ester 1a to nitrostyrene, including an aromatic 4,4′-biphenyl ester group to promote crystallinity in product 3a (biphenyl ester 3a mp: 105–107 °C vs ethyl ester mp: 71–74 °C, Figure 2b). We were excited to see nickel diamine catalyst I17,38,39 facilitated both the enantioselective C-C bond formation and a CIDT at the resulting α-carbon stereogenic center. After slight modifications to previously successful CIDT conditions,27,28 which included ethereal solvent and a concentrated reaction mixture ([nitrostyrene]0 = 1.0 M), we obtained crystalline product 3a in good yield with excellent enantio- and diastereoselectivity upon filtration (see Supporting Information for full optimization).
Next, the method was expanded to include more diverse nitroolefins, requiring only slight modifications to the solvent system across the substrate scope (Scheme 1). Unsubstituted and para- substituted nitrostyrenes performed well in this manifold, affording crystalline products in fair to excellent yields with excellent enantio- and diastereoselectivity. Nitro-styrenes with electron-donating (3f, 3g, 3j) and electron-withdrawing (3d, 3e) substituents were well-tolerated, as was a meta- substituted nitrostyrene (3h). Ortho-substituted, aliphatic, and heteroaromatic nitrostyrenes proved difficult (see Supporting Information for more details).
Scheme 1. Scope of the Enantio- and Diastereoselective Michael Additiona.

aReaction conditions: Nucleophile 1 (0.300 mmol, 1.5 equiv), nitroolefin 2 (0.200 mmol, 1.0 equiv), catalyst I (5 mol %), rt, 24 h. Solvent (Et2O, MTBE, toluene, etc.) and concentration were adjusted by substrate. Yields refer to isolated yields. Diastereoselectivity was determined by 1H NMR spectroscopic analysis of the solid product following filtration. Enantiomeric ratios were determined by chiral HPLC analysis of the solid product. For compounds without X-ray crystal structures, the relative stereo-chemistry of the major diastereomer shown must be considered tentative because of the possibility for thermodynamic crystallization properties to vary depending on the substrate.27 X-ray structure of adduct 3j is shown at 50% thermal ellipsoids.
Both aromatic and branched aliphatic ketone groups were also tolerated (3k and 3l, respectively), as were various aromatic ester groups (3a, 3b, 3c). The crystallinity of the product could be tuned by altering the keto ester. Using fluorenol-derived keto ester 1b in place of keto ester 1a increased the yield and diastereoselectivity of the Michael addition to a lipophilic nitrostyrene (cf. adducts 3g vs 3j). This difference in crystallinity was reflected in their respective melting points (biphenyl methyl ester 3g, mp 109–112 °C; fluoren-9-yl ester 3j, mp 133–135 °C). An X-ray diffraction study of keto ester 3j revealed the absolute stereochemistry to be (2R,3S), and together with the X-ray structure of hydroxy ester 4 derived from keto ester 3a (vide inf ra), the configurations of the other Michael adducts were assigned by analogy. An improved yield and excellent selectivity were realized on larger scale: the synthesis of adduct 3a on gram scale afforded the desired product in 88% yield, >20:1 dr, and >99:1 er (see Supporting Information for more details).
To investigate if the reaction was indeed operating by a CIDT, we analyzed by 1H NMR spectroscopy the reaction filtrate from the synthesis of adduct 3a. The filtrate was revealed to contain a mixture of adduct 3a and its epimer as a 1:1 mixture of diastereomers in solution, indicating any kinetic selectivity during C-C bond formation was removed by epimerization and that the enrichment was not simply a consequence of selective nondynamic crystallization. A second 1H NMR spectroscopic analysis revealed that a single stereoisomer (>20:1 dr) of 3a was converted to a 1.4:1 mixture of 3a:epi-3a in 1 h by catalyst I (7 mol %) in the presence of an equivalent of keto ester 1a, highlighting the catalyst’s role in epimerizing the Michael adducts to establish the equilibrium required for a CIDT (Scheme 2). Finally, under homogeneous conditions, adduct 3a was synthesized with an er of 95:5 relative to the 99:1 er obtained under CIDT conditions, indicating that the latter provides a concurrent upgrade to the enantioselectivity during crystallization (see Supporting Information for more details).
Scheme 2. Ni(II) Catalyst I Rapidly Epimerizes the Michael Adduct To Facilitate a CIDT.

Having developed a highly diastereo- and enantioselective reaction platform, we sought to exploit the polyfunctional nature of the products in a variety of secondary transformations (Scheme 3). The ketone was reduced with a noncoordinating borohydride to yield β-hydroxy ester 4 in good yield and excellent dr.40 An X-ray diffraction study of alcohol 4 was conducted to assign the absolute stereochemistry as (1R,2R,3S). The nitro group of alcohol 4 was further reduced to the amine,41 which spontaneously cyclized to form β-hydroxy lactam 5. The latter could be epimerized to a second diastereomer, epi-5, under basic conditions. Hydrogenolysis of benzyl ester 4 allowed for simple functional group interconversion, revealing β-hydroxy acid 6.
Scheme 3. Reactions of Michael Adducts.

a(a) Bu4NBH4 (3 equiv), DCM:MeOH (1:1, 0.05 M), −50 °C, 9 h. (b) Fe (10 equiv), HCl (10 equiv), EtOH (0.05 M), reflux, 1 h. (c) NaOH (33 equiv), MeOH (0.03 M), 40 °C, 2.5 h. (d) Pd/C (10 wt %), EtOH:EtOAc:MeOH (1:1.4:3, 0.03 M), H2 (2 atm), 20 min. (e) Number in parentheses represents dr following flash column chromatography. X-ray structure of alcohol 4 is shown at 50% thermal ellipsoids.
The precise molecular drivers for selectivity remain to be determined; however, a closer examination of the X-ray structures of Michael adduct 3j and alcohol 4 revealed possible clues involving the aromatic ester groups and their intra-molecular interactions that possibly enhance the crystal packing42 of the major diastereomer. The iso-butyl group of Michael adduct 3j exhibited a contrasteric conformation, possibly due to attractive London dispersion interactions with the fluorenol ester group. The biphenyl ester of alcohol 4 was found to have a C-H π interaction between the biphenyl ester and the phenyl ring (Figure 3). How important (or incidental) such interactions are and the extent to which they can be integrated into designed CIDTs are topics of ongoing investigation.
Figure 3.

Intramolecular interactions for keto ester 3j and biphenyl ester 4 in the solid state.
In summary, we have developed a method that merges a CIDT with the asymmetric Michael addition of unsubstituted β-keto esters to nitroolefins, affording complex products as single (or major) diastereomers that can be isolated by direct filtration of the reaction mixture. Three crystalline β-keto esters promoted the product crystallinity required for a CIDT and could be tuned to improve the reaction’s yield and diastereoselectivity. 1H NMR spectroscopy studies indicate the dual role of nickel diamine catalyst I in forming the key C-C bond and epimerizing the resulting α-carbon stereogenic center to promote a CIDT. The stereochemically complex Michael adducts were then diversified to access useful products, including a γ-lactam and β-keto acid. Our laboratory continues to investigate the potential of this merged manifold in other asymmetric Michael additions.
Supplementary Material
ACKNOWLEDGMENTS
The project described was supported by award no. R35 GM118055 from the National Institute of General Medical Sciences. The X-ray crystallography analysis was performed by E. T. Crawford of these laboratories (UNC Chapel Hill) on instrumentation acquired under the NSF MRI program under grant no. CHE-2117287. We thank the UNC Dept. of Chemistry Mass Spectrometry Core Laboratory for their assistance with mass spectrometry analysis, and the UNC Dept. of Chemistry NMR Core Laboratory for assistance with NMR analysis.
Footnotes
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.orglett.3c02799.
Experimental details, materials and methods, characterization data, NMR spectra, chromatograms for chiral separations, and information on X-ray diffraction experiments (PDF)
Accession Codes
CCDC 2252712 and 2267674 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.
The authors declare no competing financial interest.
Contributor Information
Emily R. Sherman, Department of Chemistry, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599-3290, United States
William R. Cassels, Department of Chemistry, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599-3290, United States
Jeffrey S. Johnson, Department of Chemistry, University of North Carolina at Chapel Hill, Chapel Hill, North Carolina 27599-3290, United States
Data Availability Statement
The data underlying this study are available in the published article and its Supporting Information.
REFERENCES
- (1).Lovering F; Bikker J; Humblet C Escape from Flatland: Increasing Saturation as an Approach to Improving Clinical Success. J. Med. Chem 2009, 52 (21), 6752–6756. [DOI] [PubMed] [Google Scholar]
- (2).Lovering F Escape from Flatland 2: Complexity and Promiscuity. MedChemComm 2013, 4 (3), 515–519. [Google Scholar]
- (3).Jeschke P Current Status of Chirality in Agrochemicals: Chirality in Agrochemicals. Pest Manag. Sci 2018, 74 (11), 2389–2404. [DOI] [PubMed] [Google Scholar]
- (4).Méndez-Lucio O; Medina-Franco JL The Many Roles of Molecular Complexity in Drug Discovery. Drug Discovery Today 2017, 22 (1), 120–126. [DOI] [PubMed] [Google Scholar]
- (5).Beaver MG; Caille S; Farrell RP; Rötheli AR; Smith AG; Tedrow JS; Thiel OR Building Complexity and Achieving Selectivity through Catalysis - Case Studies from the Pharmaceutical Pipeline. Synlett 2020, 32 (5), 457–471. [Google Scholar]
- (6).Das T; Mohapatra S; Mishra NP; Nayak S; Raiguru BP Recent Advances in Organocatalytic Asymmetric Michael Addition Reactions to α, β-Unsaturated Nitroolefins. ChemistrySelect 2021, 6 (15), 3745–3781. [Google Scholar]
- (7).Zheng K; Liu X; Feng X Recent Advances in Metal-Catalyzed Asymmetric 1,4-Conjugate Addition (ACA) of Nonorganometallic Nucleophiles. Chem. Rev 2018, 118 (16), 7586–7656. [DOI] [PubMed] [Google Scholar]
- (8).Alonso DA; Baeza A; Chinchilla R; Gómez C; Guillena G; Pastor IM; Ramón DJ Recent Advances in Asymmetric Organocatalyzed Conjugate Additions to Nitroalkenes. Molecules 2017, 22 (6), 895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Zhang Y; Wang W Recent Advances in Organocatalytic Asymmetric Michael Reactions. Catal. Sci. Technol 2012, 2 (1), 42–53. [Google Scholar]
- (10).Pellissier H Recent Developments in Enantioselective Nickel(II)-Catalyzed Conjugate Additions. Adv. Synth. Catal 2015, 357 (13), 2745–2780. [Google Scholar]
- (11).Mondal A; Bhowmick S; Ghosh A; Chanda T; Bhowmick KC Advances on Asymmetric Organocatalytic 1,4-Conjugate Addition Reactions in Aqueous and Semi-Aqueous Media. Tetrahedron Asymmetry 2017, 28 (7), 849–875. [Google Scholar]
- (12).Hui C; Pu F; Xu J Metal-Catalyzed Asymmetric Michael Addition in Natural Product Synthesis. Chem. - Eur. J 2017, 23 (17), 4023–4036. [DOI] [PubMed] [Google Scholar]
- (13).Reyes E; Uria U; Vicario JL; Carrillo L The Catalytic, Enantioselective Michael Reaction. In Organic Reactions; John Wiley & Sons, Ltd: 2016; pp 1–898. DOI: 10.1002/0471264180.or090.01. [DOI] [Google Scholar]
- (14).Pasuparthy SD; Maiti B Enantioselective Organocatalytic Michael Addition Reactions Catalyzed by Proline/Prolinol/Supported Proline Based Organocatalysts: An Overview. ChemistrySelect 2022, 7. DOI: 10.1002/slct.202104261. [DOI] [Google Scholar]
- (15).Reznikov AN; Sibiryakova AE; Baimuratov MR; Golovin EV; Rybakov VB; Klimochkin YN Synthesis of Non-Racemic 4-Nitro-2-Sulfonylbutan-1-Ones via Ni(II)-Catalyzed Asymmetric Michael Reaction of β-Ketosulfones. Beilstein J. Org. Chem 2019, 15 (1), 1289–1297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).Ma Z; Liu Y; Huo L; Gao X; Tao J Doubly Stereocontrolled Asymmetric Michael Addition of Acetylacetone to Nitroolefins Promoted by an Isosteviol-Derived Bifunctional Thiourea. Tetrahedron Asymmetry 2012, 23 (6), 443–448. [Google Scholar]
- (17).Evans DA; Mito S; Seidel D Scope and Mechanism of Enantioselective Michael Additions of 1,3-Dicarbonyl Compounds to Nitroalkenes Catalyzed by Nickel(II)-Diamine Complexes. J. Am. Chem. Soc 2007, 129 (37), 11583–11592. [DOI] [PubMed] [Google Scholar]
- (18).Xie Z-B; Wang N; Wu M-Y; He T; Le Z-G; Yu X-Q Catalyst-Free and Solvent-Free Michael Addition of 1,3-Dicarbonyl Compounds to Nitroalkenes by a Grinding Method. Beilstein J. Org. Chem 2012, 8 (1), 534–538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Jiang X; Zhang Y; Liu X; Zhang G; Lai L; Wu L; Zhang J; Wang R Enantio- and Diastereoselective Asymmetric Addition of 1,3-Dicarbonyl Compounds to Nitroalkenes in a Doubly Stereo-controlled Manner Catalyzed by Bifunctional Rosin-Derived Amine Thiourea Catalysts. J. Org. Chem 2009, 74 (15), 5562–5567. [DOI] [PubMed] [Google Scholar]
- (20).Goldys AM; Nuñez MG; Dixon DJ Creation through Immobilization: A New Family of High Performance Heterogeneous Bifunctional Iminophosphorane (BIMP) Superbase Organocatalysts. Org. Lett 2014, 16 (24), 6294–6297. [DOI] [PubMed] [Google Scholar]
- (21).Du H; Rodriguez J; Bugaut X; Constantieux T Organo-catalytic Enantio- and Diastereoselective Conjugate Addition to Nitroolefins: When β-Ketoamides Surpass β-Ketoesters. Chem. - Eur. J 2014, 20 (27), 8458–8466. [DOI] [PubMed] [Google Scholar]
- (22).Chen G; Liang G; Wang Y; Deng P; Zhou H A Homodinuclear Cobalt Complex for the Catalytic Asymmetric Michael Reaction of β-Ketoesters to Nitroolefins. Org. Biomol. Chem 2018, 16 (20), 3841–3850. [DOI] [PubMed] [Google Scholar]
- (23).Okino T; Hoashi Y; Takemoto Y Enantioselective Michael Reaction of Malonates to Nitroolefins Catalyzed by Bifunctional Organocatalysts. J. Am. Chem. Soc 2003, 125 (42), 12672–12673. [DOI] [PubMed] [Google Scholar]
- (24).Ju Y-D; Xu L-W; Li L; Lai G-Q; Qiu H-Y; Jiang J-X; Lu Y Noyori’s Ts-DPEN Ligand: An Efficient Bifunctional Primary Amine-Based Organocatalyst in Enantio- and Diastereoselective Michael Addition of 1,3-Dicarbonyl Indane Compounds to Nitroolefins. Tetrahedron Lett 2008, 49 (48), 6773–6777. [Google Scholar]
- (25).Okino T; Hoashi Y; Furukawa T; Xu X; Takemoto Y Enantio- and Diastereoselective Michael Reaction of 1,3-Dicarbonyl Compounds to Nitroolefins Catalyzed by a Bifunctional Thiourea. J. Am. Chem. Soc 2005, 127 (1), 119–125. [DOI] [PubMed] [Google Scholar]
- (26).De Jesús Cruz P; Johnson JS Crystallization-Enabled Henry Reactions: Stereoconvergent Construction of Fully Substituted [N]-Asymmetric Centers. J. Am. Chem. Soc 2022, 144, 15803–15811. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (27).de Jesus Cruz P; Cassels WR; Chen C-H; Johnson JS Doubly Stereoconvergent Crystallization Enabled by Asymmetric Catalysis. Science 2022, 376 (6598), 1224–1230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Cassels WR; Crawford ET; Johnson JS Enantio- and Diastereoselective Mannich Reactions of ß-Dicarbonyls by Second Stage Diastereoconvergent Crystallization. ACS Catal 2023, 13 (10), 6518–6524. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Cassels WR; Johnson JS Development of Organic Reactions That Productively Leverage Physical Properties. Sci. Adv 2023, 9 (27), eadg6765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Brands KMJ; Davies AJ Crystallization-Induced Diastereomer Transformations. Chem. Rev. 2006, 106 (7), 2711–2733. [DOI] [PubMed] [Google Scholar]
- (31).Kolarovič A; Jakubec P State of the Art in Crystallization-Induced Diastereomer Transformations. Adv. Synth. Catal 2021, 363 (17), 4110–4158. [Google Scholar]
- (32).Anderson NG Developing Processes for Crystallization-Induced Asymmetric Transformation. Org. Process Res. Dev 2005, 9 (6), 800–813. [Google Scholar]
- (33).Anderson NG Assessing the Benefits of Direct Isolation Processes. Org. Process Res. Dev 2004, 8 (2), 260–265. [Google Scholar]
- (34).Constable DJC; Jimenez-Gonzalez C; Henderson RK Perspective on Solvent Use in the Pharmaceutical Industry. Org. Process Res. Dev 2007, 11 (1), 133–137. [Google Scholar]
- (35).Benetti S; Romagnoli R; De Risi C; Spalluto G; Zanirato V Mastering.Beta.-Keto Esters. Chem. Rev 1995, 95 (4), 1065–1114. [Google Scholar]
- (36).Govender T; Arvidsson PI; Maguire GEM; Kruger HG; Naicker T Enantioselective Organocatalyzed Transformations of β-Ketoesters. Chem. Rev 2016, 116 (16), 9375–9437. [DOI] [PubMed] [Google Scholar]
- (37).Decicco CP; Buckle RN Chirality as a Probe in.Beta.-Keto Ester Tautomerism. J. Org. Chem 1992, 57 (3), 1005–1008. [Google Scholar]
- (38).Evans DA; Seidel D Ni(II)-Bis[(R,R)-N,N’-Dibenzylcy-clohexane-1,2-Diamine]Br2 Catalyzed Enantioselective Michael Additions of 1,3-Dicarbonyl Compounds to Conjugated Nitroalkenes. J. Am. Chem. Soc 2005, 127 (28), 9958–9959. [DOI] [PubMed] [Google Scholar]
- (39).Sohtome Y; Komagawa S; Nakamura A; Hashizume D; Lectard S; Akakabe M; Hamashima Y; Uchiyama M; Sodeoka M Experimental and Computational Investigation of Facial Selectivity Switching in Nickel-Diamine-Acetate-Catalyzed Michael Reactions. J. Org. Chem 2023, 88, 7764. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (40).Raber DJ; Guida WC Tetrabutylammonium Borohydride. Borohydride Reductions in Dichloromethane. J. Org. Chem 1976, 41 (4), 690–696. [Google Scholar]
- (41).Matsuda Y; Kitajima M; Takayama H First Total Synthesis of Trimeric Indole Alkaloid, Psychotrimine. Org. Lett 2008, 10 (1), 125–128. [DOI] [PubMed] [Google Scholar]
- (42).Corpinot MK; Bučar D-K A Practical Guide to the Design of Molecular Crystals. Cryst. Growth Des 2019, 19 (2), 1426–1453. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data underlying this study are available in the published article and its Supporting Information.
