Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Jun 30.
Published in final edited form as: J Am Chem Soc. 2021 Jun 21;143(25):9343–9349. doi: 10.1021/jacs.1c05212

Enantioselective Iridium-Catalyzed Allylation of Nitroalkanes: Entry to β-Stereogenic α-Quaternary Primary Amines

Woo-Ok Jung , Binh Khanh Mai , Brian J Spinello , Zachary J Dubey , Seung Wook Kim , Craig E Stivala §, Jason R Zbieg §, Peng Liu , Michael J Krische
PMCID: PMC8284932  NIHMSID: NIHMS1720704  PMID: 34152145

Abstract

The first systematic study of simple nitronate nucleophiles in iridium-catalyzed allylic alkylation are described. Using a tol-BINAP-modified π-allyliridium C,O-benzoate catalyst, α,α-disubstituted nitronates substitute racemic branched alkyl-substituted allylic acetates, thus providing entry to β-stereogenic α-quaternary primary amines. DFT calculations reveal early transition states that render the reaction less sensitive to steric effects and distinct trans-effects of diastereomeric chiral-at-iridium π-allyl complexes that facilitate formation of congested tertiary-quaternary C-C bonds.

Graphical Abstract

graphic file with name nihms-1720704-f0001.jpg


Chiral amines are prevalent among FDA approved drugs,1 as are protocols for their synthesis.2 The vast majority of catalytic enantioselective methods for the construction of chiral amines deliver N-substituted carbon stereocenters.2 Catalytic enantioselective methods applicable to the formation of acyclic chiral β-stereogenic amines are far less common and encompass asymmetric hydrogenation of enamines,3 allylic amines/amides4 or nitrolefins,5 1,4-reductions and 1,4-additions to nitroolefins,6,7 and dynamic kinetic resolutions via aldehyde reductive amination.8 These methods do not deliver chiral β-stereogenic α-quaternary primary amines. Catalytic enantioselective Tsuji–Trost nitronate allylic alkylation-reduction potentially provides access to acyclic chiral β-stereogenic amines, but such methods are underdeveloped.912 Palladium-catalyzed allylic alkylations typically display linear regioselectivity, and consequently focus on reactions that proceed via symmetric π-allylpalladium intermediates9 or prochiral nucleophiles (α-nitroesters).12 Beyond vinyl epoxides,10a only a single system for catalytic asymmetric allylic alkylation of unactivated nitronates with mono-substituted π-allyl precursors that displays branched regioselectivity has been described.10b,c Only two examples of corresponding iridium-catalyzed nitronate allylic alkylations are reported.11 Both the palladium- and iridium-based catalyst systems rely on linear aryl-substituted π-allyl precursors and display incomplete diastereo- and regioselectivity (Figure 1).

Figure 1.

Figure 1.

Unactivated nitronates in catalytic enantioselective allylic alkylations of mono-substituted π-allyl precursors.

In connection with studies of π-allyliridium C,O-benzoate-catalyzed nucleophilic allylations of carbonyl compounds,13 we recently found these same complexes are competent catalysts for electrophilic allylation, as demonstrated in regio- and enantioselective aminations of racemic branched alkyl-substituted allylic acetates to form acyclic chiral α-stereogenic amines.14 Aspiring to access corresponding chiral β-stereogenic amines, a study of nitronate nucleophiles was undertaken. Here, we demonstrate that α,α-disubstituted nitroalkanes participate in highly enantioselective allylic alkylation with racemic branched alkyl-substituted allylic acetates. Despite forming congested contiguous quaternary-tertiary C-C bonds, complete branched regioselectivities are observed. DFT calculations reveal early transition states that render the reaction less sensitive to steric effects and distinct trans-effects of diastereomeric chiral-at-iridium π-allyl complexes that facilitate formation of congested tertiary-quaternary C-C bonds.

In initial experiments, allylic acetate 1a (100 mol%) was exposed to 2-nitropropane 2a (1125 mol%) as solvent (1.0 M) in the presence of Cs2CO3 (500 mol%) and the π-allyliridium C,O-benzoate complex modified by (S)-tol-BINAP, Ir-I, at 100 °C. The homoallylic nitroalkane 3a was formed in 51% yield and 95% ee (Table 1, entry 1). Lower loading of 2-nitropropane 2a (450 mol%) with DMF (0.5 M) as solvent improved yield and enantioselectivity (Table 1, entry 2), but further reduction in the loading of 2a led to a small decrease in yield (Table 1, entry 3). A slight decrease in temperature (80 °C) improved the yield of 3a without diminishing enantioselectivity (Table 1, entry 4). Lower loading of Cs2CO3 decreased the yield of 3a (Table 1, entry 5) and water (100 mol%) dramatically decreased conversion of 1a (Table 1, entry 6). Other π-allyliridium C,O-benzoates were evaluated but did not improve the yield of 3a (Table 1, entries 7–10).

Table 1.

Selected optimization experiments in the enantioselective iridium-catalyzed allylic alkylation of racemic branched alkyl-substituted allylic acetate 1awith nitronate 2a.a

graphic file with name nihms-1720704-t0002.jpg
Entry 2a (mol%) catalyst Cs2C03 (mol%) T°C Yield (%) ee (%)
1b 1125 lr-I 500 mol% 100 51 95
2 450 lr-I 500 mol% 100 67 99
3 240 Ir-I 500 mol% 100 35 99
graphic file with name nihms-1720704-t0003.jpg 4 450 lr-I 500 mol% 80 71 99
5 450 lr-I 300 mol% 80 56 99
6C 450 lr-I 500 mol% 80 <10 -
7 450 lr-II 500 mol% 80 51 99
8 450 lr-III 500 mol% 80 53 99
9 450 Ir-IV 500 mol% 80 43 99
10 450 Ir-V 500 mol% 80 50 99
graphic file with name nihms-1720704-t0004.jpg
a

Yields are of material isolated by silica gel chromatography. Diastereoselectivities were determined by 1H NMR of crude reaction mixtures. Enantioselectivities were determined by HPLC analysis.

b

2a = solvent (1.0 M).

c

H2O (100 mol%). See Supporting Information for experimental details.

To evaluate reaction scope, optimal conditions for the formation of 3a (Table 1, entry 4) were applied to the allylic alkylation of α,α-disubstituted nitroalkanes 2a-2j with racemic branched alkyl substituted allylic acetates 1a-1n (Scheme 1). As illustrated by the conversion of 2-nitropropane 2a to adducts 3a-3h, diverse (hetero)aromatic groups (3a-3e), (thio)ethers (3f, 3g), cyclopropyl groups (3h) are tolerated. As demonstrated by the formation of adducts 3i-3x, cyclic nitroalkanes 2b-2h are also competent partners for allylic alkylation. This includes nitrocyclobutane (2b), nitrocyclopentane (2c), N-Cbz- and N-Boc-4-nitropiperidines (2d and 2g, respectively), as well as 1,1-disubstituted 4-nitrocyclohexanes (2e, 2f), nitrocycloheptane (2h) and the spirocyclic nitroalkanes (2i, 2j). In all cases, good to excellent yields of adducts 3a-3x were obtained with high levels of enantiomeric enrichment (87–99% ee) and complete branched regioselectivity. The nonsymmetric nitroalkane (1-nitroethyl)benzene provided high yields of allylic alkylation product, but low diastereomeric ratios were observed. Low conversions were observed using allylic acetates bearing α- or β-branched alkyl moieties. As revealed by the conversion of 3f, 3k, 3l, 3s and 3v to amines 4f, 4k, 4l, 4s and 4v, respectively, zinc-mediated reduction of the allylation products provides entry to chiral β-stereogenic α-quaternary primary amines (Scheme 2).

Scheme 1.

Scheme 1.

Iridium-catalyzed allylation nitroalkanes 2a-2j using allylic acetates 1a-1n to form homoallylic nitroalkanes 3a-3x.a

aYields are of material isolated by silica gel chromatography. Enantioselectivities were determined by chiral stationary phase HPLC analysis. bNMP was used as the reaction solvent. c(S)-Ir-tol-BINAP, X= OMe. dIr-catalyst (10 mol%). See Supporting Information for experimental details.

Scheme 2.

Scheme 2.

Zinc-mediated reduction of 3f, 3k, 3l, 3s and 3v to form β-stereogenic α-quaternary primary amines 4f, 4k, 4l, 4s and 4v.

aYields are of material isolated after filtration through celite. See Supporting Information for experimental details

Density functional theory (DFT) calculations were carried out to investigate the origin of regio- and enantioselectivity of the Ir-catalyzed allylic alkylation.15,16 As numerous diastereomeric π-allyliridium C,O-benzoate complexes may exist in equilibration prior to the allylation, the relative stabilities, reactivities and regioselectivities of all 16 possible stereoisomers of the π-crotyliridium(III) complex were computed (Figure 2). These include diastereomers A, B, C, and D, where the C,O-benzoate resides in different coordination sites of the octahedral Ir complex,17 with four different coordination modes of the π-crotyl ligand for each diastereomer. Here, the suffix exo or endo describes whether the allyl C2−H points away or towards the C,O-benzoate ligand, and distal or proximal describes if the C3−methyl points away or towards the C,O-benzoate (see Figure S2 for the relative stabilities of all 16 isomers). Computed activation barriers for the addition of CMe2NO2 anion to each π-crotyliridium(III) isomer leading to branched and linear products are shown in Figure 2B. The 16 π-crotyliridium(III) isomers exhibited very different reactivity and regioselectivity. The most reactive isomer, D_exo_proximal, strongly favors the branched product (ΔΔG(B-L) = −8.5 kcal/mol), while some less reactive isomers, such as A_endo_proximal and B_endo_proximal, prefer the linear product. These results suggest stereogenicity at iridium not only affects reactivity but also regioselectivity. To assess the origin of these effects, computational analysis on the electronic and structural properties of the diastereomeric π-crotyliridium(III) complexes were performed.

Figure 2.

Figure 2.

Computed activation free energies of the addition of CMe2NO2 to different π-crotyliridium(III) isomers. All activation barriers (ΔG) are with respect to A_endo_distal.

The computed transition state structures of the outer sphere addition of the nitronate nucleophile all involve long forming C–C distances (usually >2.4 Å in branch-selective TSs and >2.2 Å in linear-selective TSs. See Figure 3 and Figure S3 in the SI). These early transition states suggest that the addition is not sensitive to steric effects, and thus, the regioselectivity is mainly controlled by the electronic properties of the π-allyl complexes. The computed ground state properties of the π-crotyliridium(III) isomers and the distortion/interaction model analysis of the allylation transition states revealed enormously different electronic properties and their influences on the regioselectivity. In branch-selective isomers, such as D_exo_proximal (Figure 3A), the Ir−C3 distance is much longer than Ir−C1. This geometry leads to weaker d→π* backbonding at C3 and more positive charge on C3 compared to C1 (Figure 3A),18 making the more substituted C3 terminus more electrophilic.19 In addition, the transition state of nitronate addition to C3 (TS-1) is promoted by the smaller distortion energy of the π-crotyliridium complex (ΔEdist-Ir) as the ground state Ir−C3 bond of D_exo_proximal is predistorted to a longer distance (2.50 Å) that is closer to that in TS-1 (2.87 Å). In contrast, electronic effects strongly disfavor branch-selective addition with A_endo_proximal (Figure 2B), because the more substituted terminus C3 is less electrophilic than C1, as evidenced by the negative charge on C3. In addition, the Ir−C3 bond becomes shorter than Ir−C1, leading to greater distortion energy (ΔEdist-Ir) to elongate the Ir−C3 bond in the branch-selective transition state (TS-3).

Figure 3.

Figure 3.

Origin of regioselectivity. Bisphosphine ligand and some H-atoms are omitted for clarity. All Gibbs free energies are in kcal/mol with respect to A_endo_distal. ΔEdist-Ir: distortion energy of the π-crotyliridium complex to reach its geometry in the TS for nucleophilic addition.

The same analyses on all 16 isomers of the π-crotyliridium complex revealed good correlations between the computed regioselectivity for each isomer and the C3/C1 difference in NPA charge and Ir−C3 bond distance, i.e. diastereomers with more positive charge on C3 and a longer Ir−C3 bond give higher branch-selectivity (Figure 4).20 These results are consistent with our findings that regioselectivity is controlled by electronic effects, including the allyl electrophilicity and the distortion of the π-allyl complex. The electronic properties of the diastereomeric π-crotyl complexes also are affected by the trans effect21 of the bisphosphine and the chelating C,O-benzoate ligands on the stereogenic Ir center. In diastereomers A and B, due to the stronger trans effect of the aryl vs phosphine groups, addition occurs at sites trans to the aryl group, leading to branched products in “distal” isomers and linear products in “proximal” isomers. In disastereomers C and D, “proximal” isomers, in which C3 is trans to a phosphorus atom, give higher branched regioselectivity. This is consistent with the stronger trans effects of phosphine vs carboxylate ligands. These branch-selective “proximal” isomers of C and D are also among the most reactive in all 16 stereoisomers (Figure 2B), as the π-acceptor ability of the phosphine ligand promotes addition at sites trans to phosphorus.22

Figure 4.

Figure 4.

Electronic effects on allylation regioselecitivy of different stereoisomers of the π-crotyliridium complex.

Stereogenicity at iridium is also a critical determinant of enantioselectivity. Due to trans effects (vide supra), the “proximal” isomers of C and D are the most reactive among all 16 stereoisomers. The four most favorable pathways involve D_exo_proximal (TS-1) and C_endo_proximal (TS-7), which give (S)-product, and D_endo_proximal (TS-6) and C_exo_proximal (TS-5), which give (R)-product. Optimized structures of these branch-selective transition states are shown in Figure 5. The two transition states leading to the (R)-product, TS-5 and TS-6, are destabilized because the allyl C2−H moiety in TS-5 and the C3−methyl in TS-6 clash with a P-tolyl group of (S)-tol-BINAP. Thus, TS-5 and TS-6 are 1.6 and 1.8 kcal/mol less stable than the lowest-energy transition state, TS-1 that gives the (S)-product. In TS-1, the steric repulsions between tol-BINAP and the allyl group are absent. Our calculations are consistent with the experimentally observed enantioselectivity for the (S)-product.

Figure 5.

Figure 5.

Optimized structures for additions of CMe2NO2 anion giving branched product. Gibbs free energies are with respect to A_endo_distal.

In summary, we report iridium-catalyzed allylic alkylations of nitronate nucleophiles. This method, which employs an air and water stable π-allyliridium C,O-benzoate catalyst modified by tol-BINAP, enables highly regio- and enantioselective substitution of racemic branched alkyl-substituted allylic acetates by α,α-disubstituted nitronates and, hence, entry to β-stereogenic α-quaternary primary amines. As revealed by DFT calculations, early transition states that render the reaction less sensitive to steric effects and distinct trans-effects of diastereomeric chiral-at-iridium π-allyl complexes facilitate formation of congested tertiary-quaternary C-C bonds. Related asymmetric allylic alkylations of nonstabilized carbanions are currently underway.23

Supplementary Material

Supporting Info

Acknowledgments.

The Robert A. Welch Foundation (F-0038), the NIH-NIGMS (RO1-GM069445, 1 S10 OD021508-01, R35 GM128779). We thank Mr. Weijia Shen and Ms. Sakiho Sumikura for skillful technical assistance. Genentech is acknowledged for summer predoctoral internship support (W.-O. J., B. J. S. and S. W. K.). DFT calculations were carried out at the Center for Research Computing at the University of Pittsburgh, the Extreme Science and Engineering Discovery Environment (XSEDE), and the TACC Frontera Supercomputer supported by the National Science Foundation grant number ACI-1548562. We thank Dr. Michael Ruf of Bruker AXS for acquisition of crystallographic data.

Footnotes

The authors declare no competing financial interest.

Supporting Information Available: Experimental procedures, spectroscopic and chromatographic data for all new compounds (1H NMR, 13C NMR, IR, HRMS), including HPLC traces for racemic and enantiomerically enriched compounds, and computational details. Single crystal X-ray diffraction data for compound 4l. This material is available free of charge via the internet at http://pubs.acs.org.

REFERENCES

  • (1).It was estimated in 2014 that chiral amines constitute roughly 40% of new FDA approved small molecule drugs:; Jarvis LM Chem. Eng. News 2016, 94, 12–17. [Google Scholar]
  • (2).For selected reviews on the synthesis of chiral amines, see:; (a) Enders D; Reinhold U Asymmetric Synthesis of Amines by Nucleophilic 1,2-Addition of Organometallic Reagents to the CN-Double Bond. Tetrahedron: Asymmetry 1997, 8, 1895–1946. [Google Scholar]; (b) Bloch R Additions of Organometallic Reagents to C=N Bonds: Reactivity and Selectivity. Chem. Rev 1998, 98, 1407–1438. [DOI] [PubMed] [Google Scholar]; (c) Robak MT; Herbage MA; Ellman JA Synthesis and Applications of tert-Butanesulfinamide. Chem. Rev 2010, 110, 3600–3740. [DOI] [PubMed] [Google Scholar]; (d) Chiral Amine Synthesis: Methods, Developments and Applications; Nugent TC, Eds.; Wiley-VCH, 2010. [Google Scholar]; (e) Nugent TC; El-Shazly M Chiral Amine Synthesis − Recent Developments and Trends for Enamide Reduction, Reductive Amination, and Imine Reduction. Adv. Synth. Catal 2010, 352, 753–819. [Google Scholar]; (f) Stereoselective Formation of Amines; Li W, Zhang X, Eds.; Topics in Current Chemistry,; Vol. 343; Springer, 2014. [DOI] [PubMed] [Google Scholar]; (g) Mailyan AK; Eickhoff JA; Minakova AS; Gu Z; Lu P; Zakarian A Cutting-Edge and Time-Honored Strategies for Stereoselective Construction of C–N Bonds in Total Synthesis. Chem. Rev 2016, 116, 4441–4557. [DOI] [PubMed] [Google Scholar]; (h) Abdine RAA; Hedouin G; Colobert F; Wencel-Delord J Metal-Catalyzed Asymmetric Hydrogenation of C═N Bonds. ACS Catal. 2021, 11, 215–247. [Google Scholar]
  • (3).For chiral β-stereogenic amines via catalytic enantioselective hydrogenation of enamines derivatives (including dehydro-β-amino acid derivatives), see:; (a) Elaridi J; Thaqi A; Prosser A; Jackson WR Robinson, A. J. An Enantioselective Synthesis of β2-Amino Acid Derivatives. Tetrahedron: Asymmetry 2005, 16, 1309–1319. [Google Scholar]; (b) Remarchuk T; Babu S; Stults J; Zanotti-Gerosa A; Roseblade S; Yang S; Huang P; Sha C; Wang Y An Efficient Catalytic Asymmetric Synthesis of a β2-Amino Acid on Multikilogram Scale. Org. Process Res. Dev 2014, 18, 135–141. [Google Scholar]; (c) Saito N; Abdullah I; Hayashi K; Hamada K; Koyama M; Sato Y Enantioselective Synthesis of β-Amino Acid Derivatives via Nickel-Promoted Regioselective Carboxylation of Ynamides and Rhodium-Catalyzed Asymmetric Hydrogenation. Org. Biomol. Chem 2016, 14, 10080–10089. [DOI] [PubMed] [Google Scholar]; (d) Han C; Savage S; Al-Sayah M; Yajima H; Remarchuk T; Reents R; Wirz B; Iding H; Bachmann S; Fantasia SM; Scalone M; Hell A; Hidber P; Gosselin F Asymmetric Synthesis of Akt Kinase Inhibitor Ipatasertib. Org. Lett 2017, 19, 4806–4809. [DOI] [PubMed] [Google Scholar]; (e) Zhang J; Liu C; Wang X; Chen J; Zhang Z; Zhang W Rhodium-Catalyzed Asymmetric Hydrogenation of β-Branched Enamides for the Synthesis of β-Stereogenic Amines. Chem. Commun 2018, 54, 6024–6027. [DOI] [PubMed] [Google Scholar]
  • (4).For chiral β-stereogenic amines via catalytic enantioselective hydrogenation of allylic amines or amides to form acyclic chiral β-stereogenic amines, see:; (a) Fahrang R; Sinou D Asymmetric Reduction of Allylamines and their Derivatives with Rhodium Complexes. Bull. Soc. Chim. Belg 1989, 98, 387–393. [Google Scholar]; (b) Wang C-J; Sun X; Zhang X Enantioselective Hydrogenation of Allylphthalimides: An Efficient Method for the Synthesis of β-Methyl Chiral Amines. Angew. Chem., Int. Ed 2005, 44, 4933–4935. [DOI] [PubMed] [Google Scholar]; (c) Steinhuebel DP; Krska SW; Alorati A; Baxter JM; Belyk K; Bishop B; Palucki M; Sun Y; Davies IW Asymmetric Hydrogenation of Protected Allylic Amines. Org. Lett 2010, 12, 4201–4203. [DOI] [PubMed] [Google Scholar]; (d) Ma S; Grinberg N; Haddad N; Rodriguez S; Busacca CA; Fandrick K; Lee H; Song JJ; Yee N; Krishnamurthy D; Senanayake CH; Wang J; Trenck J; Mendonsa S; Claise PR; Gilman RJ; Evers TH Org. Process Res. Dev 2013, 17, 806–810. [Google Scholar]; (e) Cabre A; Verdaguer X; Riera A Enantioselective Synthesis of β-Methyl Amines via Iridium-Catalyzed Asymmetric Hydrogenation of N-Sulfonyl Allyl Amines. Adv. Synth. Catal 2019, 361, 4196–4200. [Google Scholar]
  • (5).For chiral β-stereogenic amines via catalytic enantioselective hydrogenation of nitrolefins, see:; (a) Li S; Huang K; Cao B; Zhang J; Wu W; Zhang X Highly Enantioselective Hydrogenation of β,β-Disubstituted Nitroalkanes. Angew. Chem., Int. Ed 2012, 51, 8573–8576. [DOI] [PubMed] [Google Scholar]; (b) Li S; Huang K; Zhang J; Wu W; Zhang X Rh-Catalyzed Highly Enantioselective Hydrogenation of Nitroalkenes under Basic Conditions. Chem. Eur. J 2013, 19, 10840–10844. [DOI] [PubMed] [Google Scholar]; (c) Zhao Q; Li S; Huang K; Wang R; Zhang X A Novel Chiral Bisphosphine-Thiourea Ligand for Asymmetric Hydrogenation of β,β-Disubstituted Nitroalkenes. Org. Lett 2013, 15, 4014–4017. [DOI] [PubMed] [Google Scholar]; (d) Liu M, Kong D, Li M, Zi G and Hou G, Iridium-Catalyzed Enantioselective Hydrogenation of β,β-Disubstituted Nitroalkenes. Adv. Synth. Catal 2015, 357, 3875–3879. [Google Scholar]; (e) Li S; Xiao T; Li D; Zhang X First Iridium-Catalyzed Highly Enantioselective Hydrogenation of β-Nitroacrylates. Org. Lett 2015, 17, 3782–3785. [DOI] [PubMed] [Google Scholar]; (f) Yu Y-B; Cheng L; Li Y-P; Fu Y; Zhu S-F; Zhou Q-L Enantioselective Iridium-Catalyzed Hydrogenation of β,β-Disubstituted Nitroalkenes. Chem. Commun 2016, 52, 4812–4815. [DOI] [PubMed] [Google Scholar]
  • (6).For chiral β-stereogenic amines via conjugate reduction of nitroolefins (selected examples), see:; (a) Czekelius C; Carreira EM Catalytic Enantioselective Conjugate Reduction of β,β-Disubstituted Nitroalkenes. Angew. Chem., Int. Ed 2003, 42, 4793–4795. [DOI] [PubMed] [Google Scholar]; (b) Soltani O; Ariger MA; Carreira EM Transfer Hydrogenation in Water: Enantioselective, Catalytic Reduction of (E)-β,β-Disubstituted Nitroalkenes. Org. Lett 2009, 11, 4196–4198. [DOI] [PubMed] [Google Scholar]; (c) Chen L-A; Xu W; Huang B; Ma J; Wang L; Xi J; Harms K; Gong L; Meggers E Asymmetric Catalysis with an Inert Chiral-at-Metal Iridium Complex. J. Am. Chem. Soc 2013, 135, 10598–10601. [DOI] [PubMed] [Google Scholar]; (d) Martinelli E; Vicini AC; Mancinelli M; Mazzanti A; Zani P; Bernardi L; Fochi M Catalytic Highly Enantioselective Transfer Hydrogenation of β-Trifluoromethyl Nitroalkenes. An Easy and General Entry to Optically Active β-Trifluoromethyl Amines. Chem. Commun 2015, 51, 658–660. [DOI] [PubMed] [Google Scholar]; (e) Bertolotti M; Brenna E; Crotti M; Gatti FG; Monti D; Parmeggiani F; Santangelo S Substrate Scope Evaluation of the Enantioselective Reduction of β‐Alkyl‐β‐arylnitroalkenes by Old Yellow Enzymes 1–3 for Organic Synthesis Applications. ChemCatChem 2016, 8, 577–583. [Google Scholar]; (f) Xu W; Arieno M; Löw H; Huang K; Xie X; Cruchter T; Ma Q; Xi J; Huang B; Wiest O; Gong L; Meggers E Metal-Templated Design: Enantioselective Hydrogen-Bond-Driven Catalysis Requiring Only Parts-per-Million Catalyst Loading. J. Am. Chem. Soc 2016, 138, 8774–8780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).For chiral β-stereogenic amines via conjugate addition to nitroolefins (selected examples), see:; (a) Mampreianand DM; Hoveyda AH Efficient Cu-Catalyzed Asymmetric Conjugate Additions of Alkylzinc Reagents to Aromatic and Aliphatic Acyclic Nitroalkenes. Org. Lett 2004, 6, 2829–2832. [DOI] [PubMed] [Google Scholar]; (b) Herrera RP; Sgarzani V; Bernardi L; Ricci A Catalytic Enantioselective Friedel–Crafts Alkylation of Indoles with Nitroalkenes by Using a Simple Thiourea Organocatalyst. Angew. Chem., Int. Ed 2005, 44, 6576–6579. [DOI] [PubMed] [Google Scholar]; (c) Itoh J; Fuchibe K; Akiyama T Catalytic Enantioselective Friedel–Crafts Alkylation of Indoles with Nitroalkenes by Using a Simple Thiourea Organocatalyst. Angew. Chem., Int. Ed 2008, 47, 4016–4018. [DOI] [PubMed] [Google Scholar]; (d) Sheng Y-F; Li G-Q; Kang Q; Zhangand A-J; You S-L Asymmetric Friedel-Crafts Reaction of 4,7‐Dihydroindoles with Nitroolefins by Chiral Brønsted Acids under Low Catalyst Loading. Chem. Eur. J 2009, 15, 3351–3354. [DOI] [PubMed] [Google Scholar]; (e) Nishimura T; Sawano T; Tokuji S; Hayashi T Rhodium-Catalyzed Asymmetric Conjugate Alkynylation of Nitroalkenes. Chem. Commun 2010, 46, 6837–6839. [DOI] [PubMed] [Google Scholar]; (f) Arai T; Tsuchida A; Miyazaki T; Awata A Catalytic Asymmetric Synthesis of Chiral 2-Vinylindole Scaffolds by Friedel–Crafts Reaction. Org. Lett 2017, 19, 758–761. [DOI] [PubMed] [Google Scholar]
  • (8).For chiral β-stereogenic amines via dynamic kinetic asymmetric reductive amination, see:; (a) Hoffmann S; Nicoletti M List, B. Catalytic Asymmetric Reductive Amination of Aldehydes via Dynamic Kinetic Resolution. J. Am. Chem. Soc 2006, 128, 13074–13075. [DOI] [PubMed] [Google Scholar]; (b) Fuchs CS; Hollauf M; Meissner M; Simon RC; Besset T; Reek JNH; Riethorst W; Zepeck F; Kroutil W Dynamic Kinetic Resolution of 2‐Phenylpropanal Derivatives to Yield β‐Chiral Primary Amines via Bioamination. Adv. Synth. Catal 2014, 356, 2257–2265. [Google Scholar]; (c) Villa-Marcos B; Xiao J Metal and Organo‐Catalysed Asymmetric Hydroaminomethylation of Styrenes. Chin. J. Catal 2015, 36, 106–112. [Google Scholar]; (d) Meng J; Li X-H; Han Z-Y Enantioselective Hydroaminomethylation of Olefins Enabled by Rh/Brønsted Acid Relay Catalysis. Org. Lett 2017, 19, 1076–1079. [DOI] [PubMed] [Google Scholar]; (e) Matzel P; Wenske S; Merdivan S; Günther S; Höhne M Synthesis of β-Chiral Amines by Dynamic Kinetic Resolution of α-Branched Aldehydes Applying Imine Reductases. ChemCatChem 2020, 11, 4281–4285. [Google Scholar]
  • (9).For use of unactivated nitronate nucleophiles in enantioselective palladium-catalyzed allylic alkylations that proceed by way of symmetric π-allylpalladium intermediates or display linear regioselectivity, see:; (a) Rieck H; Helmchen G Palladium Complex Catalyzed Asymmetric Allylic Substitutions with Nitromethane: Enantioselectivities Exceeding 99.9 % ee. Angew. Chem. Int. Ed 1995, 34, 2687–2689. [Google Scholar]; (b) Trost BM; Surivet J-P Diastereo- and Enantioselective Allylation of Substituted Nitroalkanes. J. Am. Chem. Soc 2000, 122, 6291–6292. [Google Scholar]; (c) Trost BM; Surivet J-P Asymmetric Allylic Alkylation of Nitroalkanes. Angew. Chem. Int. Ed 2000, 39, 3122–3124. [DOI] [PubMed] [Google Scholar]; (d) Uozumi Y; Suzuka T π-Allylic C1-Substitution in Water with Nitromethane Using Amphiphilic Resin-Supported Palladium Complexes. J. Org. Chem 2006, 71, 8644–864. [DOI] [PubMed] [Google Scholar]; (e) Maki K; Kanai M; Shibasaki M Pd-Catalyzed Allylic Alkylation of Secondary Nitroalkanes. Tetrahedron 2007, 63, 4250–4257. [Google Scholar]
  • (10).Isolated examples of the use of nitronate nucleophiles in enantioselective palladium-catalyzed allylic alkylation exist, but incomplete regioselectivities are typically observed:; (a) Trost BM; Jiang C Atom Economic Asymmetric Creation of Quaternary Carbon: Regio- and Enantioselective Reactions of a Vinylepoxide with a Carbon Nucleophile. J. Am. Chem. Soc 2001, 123, 12907–12908. [DOI] [PubMed] [Google Scholar]; (b) Yang X-F; Ding C-H; Li X-H; Huang J-Q; Hou X-L; Dai L-X; Wang P-J Regio- and Enantioselective Palladium-Catalyzed Allylic Alkylation of Nitromethane with Monosubstituted Allyl Substrates: Synthesis of (R)-Rolipram and (R)-Baclofen. J. Org. Chem 2012, 77, 8980–8985. [DOI] [PubMed] [Google Scholar]; (c) Yang X-F; Yu W-H; Ding C-H; Wan S-H; Hou X-L; Dai L-X; Wang P-J Palladium-Catalyzed Regio-, Diastereo-, and Enantioselective Allylation of Nitroalkanes with Monosubstituted Allylic Substrates. J. Org. Chem 2013, 78, 6503–6509. [DOI] [PubMed] [Google Scholar]; (d) Pd-Catalyzed Allylic Alkylation of Dienyl Carbonates with Nitromethane with High C-5 Regioselectivity.Yang X-F; Li X-H; Ding C-H; Xu C-F; Dai L-X; Hou X-L Chem. Commun 2014, 50, 484–486. [DOI] [PubMed] [Google Scholar]
  • (11).Only one prior study of enantioselective iridium-catalyzed allylic alkylation of nitronate nucleophiles is reported, which mainly focused on reactions of α-nitroesters:; Dahnz A; Helmchen G Iridium-Catalyzed Enantioselective Allylic Substitutions with Aliphatic Nitro Compounds as Prenucleophiles. Synlett 2006, 697–700. [Google Scholar]
  • (12).For examples of enantioselective metal-catalyzed allylic alkylation of activated nitronates (α-nitroesters) the reader is referred to the following studies and references cited therein:; (a) Ohmatsu K; Ito M; Kunieda T; Ooi T Ion-Paired Chiral Ligands for Asymmetric Palladium Catalysis. Nat. Chem 2012, 4, 473–477. [DOI] [PubMed] [Google Scholar]; (b) Trost BM; Schultz JE; Bai Y Development of Chemo- and Enantioselective Palladium-Catalyzed Decarboxylative Asymmetric Allylic Alkylation of α-Nitroesters. Angew. Chem. Int. Ed 2019, 58, 11820–11825. [DOI] [PubMed] [Google Scholar]; (c) Davison RT; Parker PD; Hou X; Chung CP; Augustine SA; Dong VM Enantioselective Coupling of Nitroesters and Alkynes. Angew. Chem. Int. Ed 2021, 60, 4599–4603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).For a recent reviews on alcohol-mediated carbonyl allylation, see:; Kim SW; Zhang W; Krische MJ Catalytic Enantioselective Carbonyl Allylation and Propargylation via Alcohol-Mediated Hydrogen Transfer: Merging the Chemistry of Grignard and Sabatier. Acc. Chem. Res 2017, 50, 2371–2380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).(a) Meza AT; Wurm T; Smith L; Kim SW; Zbieg JR; Stivala CE; Krische MJ Amphiphilic π-Allyliridium C,O-Benzoates Enable Regio- and Enantioselective Amination of Branched Allylic Acetates Bearing Linear Alkyl Groups. J. Am. Chem. Soc 2018, 140, 1275–1279. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Kim SW; Schwartz LA; Zbieg JR; Stivala CE; Krische MJ Regio- and Enantioselective Iridium-Catalyzed Amination of Racemic Branched Alkyl-Substituted Allylic Acetates with Primary and Secondary Aromatic and Heteroaromatic Amines. J. Am. Chem. Soc 2019, 141, 671–676. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Kim SW; Schempp TT; Zbieg JR; Stivala CE; Krische MJ Regio- and Enantioselective Iridium-Catalyzed N-Allylation of Indoles and Related Azoles with Racemic Branched Alkyl-Substituted Allylic Acetates. Angew. Chem. Int. Ed 2019, 58, 7762–7766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15). DFT calculations were performed at the M06/6–311+G(d,p)-SDD(Ir)/SMD(NMP)//B3LYP-D3/6–31G(d)-SDD(Ir)/SMD(NMP) level of theory. See the Supporting Information for computational details.
  • (16).For recent computational studies of Ir-catalyzed allylation reactions, see:; (a) Liu W-B; Zheng C; Zhuo C-X; Dai L-X; You S-L Iridium-Catalyzed Allylic Alkylation Reaction with N-Aryl Phosphoramidite Ligands: Scope and Mechanistic Studies. J. Am. Chem. Soc 2012, 134, 4812–4821. [DOI] [PubMed] [Google Scholar]; (b) Bhaskararao B; Sunoj RB Origin of Stereodivergence in Cooperative Asymmetric Catalysis with Simultaneous Involvement of Two Chiral Catalysts. J. Am. Chem. Soc 2015, 137, 15712–15722. [DOI] [PubMed] [Google Scholar]; (c) Madrahimov ST; Li Q; Sharma A; Hartwig JF Origins of Regioselectivity in Iridium Catalyzed Allylic Substitution. J. Am. Chem. Soc 2015, 137, 14968–14981. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Butcher TW; Hartwig JF Enantioselective Synthesis of Tertiary Allylic Fluorides by Iridium-Catalyzed Allylic Fluoroalkylation. Angew. Chem. Int. Ed 2018, 57, 13125–13129. [DOI] [PubMed] [Google Scholar]; (e) Sorlin AM; Mixdorf JC; Rotella ME; Martin RT; Gutierrez O; Nguyen HM The Role of Trichloroacetimidate To Enable Iridium-Catalyzed Regio- and Enantioselective Allylic Fluorination: A Combined Experimental and Computational Study. J. Am. Chem. Soc 2019, 141, 14843–14852. [DOI] [PubMed] [Google Scholar]
  • (17).Kim SW; Meyer CC; Mai BK; Liu P; Krische MJ Inversion of Enantioselectivity in Allene Gas vs Allyl Acetate Reductive Aldehyde Allylation Mediated by 2-Propanol Guided by Metal-Centered Stereogenicity: An Experimental and Computational Study. ACS Catalysis 2019, 9, 9158–9163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18). Here, the computed charges on the hydrogen atoms were summed into the carbon atoms. The NPA charges of the carbon atoms alone reveal the same electrophilicity trends. See SI for details.
  • (19).Ward TR Regioselectivity of Nucleophilic Attack on [Pd(allyl)(phosphine)(imine)] Complexes: A Theoretical Study. Organometallics 1996, 15, 2836–2838. [Google Scholar]
  • (20).When the C3 of crotyl is trans to the carbon of the C,O-benzoate ligand, the η1 coordination of crotyl is more favorable than the η3 coordination. These complexes with long Ir-C3 distances are highlighted using red circles in Figure 3B.
  • (21).Coe BJ; Glenwright SJ Trans-effects in Octahedral Transition Metal Complexes. Coord. Chem. Rev 2000, 203, 5–80. [Google Scholar]
  • (22).Mata Y; Pàmies O; Diéguez M Pyranoside Phosphite-Oxazoline Ligand Library: Highly Efficient Modular P,N Ligands for Palladium-Catalyzed Allylic Substitution Reactions. A Study of the Key Palladium Allyl Intermediates. Adv. Synth. Catal 2009, 351, 3217–3234. [Google Scholar]
  • (23).The π-allyliridium C,O-benzoate catalyst developed in our laboratory was recently used to catalyze asymmetric allylic alkylations of malonates:; Zhang T-Y; Deng Y; Wei K; Yang Y-R Enantioselective Iridium-Catalyzed Allylic Alkylation of Racemic Branched Alkyl-Substituted Allylic Acetates with Malonates. Org. Lett 2021, 23, 1086–1089. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Supporting Info

RESOURCES