LETTER doi:10.1038/nature17438 Asymmetric catalytic formation of quaternary carbons by iminium ion trapping of radicals John J. Murphy1*, David Bastida1*, Suva Paria1, Maurizio Fagnoni2 & Paolo Melchiorre1,3 An important goal of modern organic chemistry is to develop new catalytic strategies for enantioselective carbon–carbon bond formation that can be used to generate quaternary stereogenic centres. Whereas considerable advances have been achieved by exploiting polar reactivity1, radical transformations have been far less successful2. This is despite the fact that open-shell intermediates are intrinsically primed for connecting structurally congested carbons, as their reactivity is only marginally affected by steric factors3. Here we show how the combination of photoredox4 and asymmetric organic catalysis5 enables enantioselective radical conjugate additions to β,β-disubstituted cyclic enones to obtain quaternary carbon stereocentres with high fidelity. Critical to our success was the design of a chiral organic catalyst, containing a redox-active carbazole moiety, that drives the formation of iminium ions and the stereoselective trapping of photochemically generated carbon-centred radicals by means of an electron-relay mechanism. We demonstrate the generality of this organocatalytic radicaltrapping strategy with two sets of open-shell intermediates, formed through unrelated light-triggered pathways from readily available substrates and photoredox catalysts—this method represents the application of iminium ion activation6 (a successful catalytic strategy for enantioselective polar chemistry) within the realm of radical reactivity. Organic chemists generally rely on polar reactivity to address the challenge of forming quaternary carbon stereocentres in a catalytic enantioselective fashion1. Of the stereoselective methods available, metal-catalysed conjugate addition of organometallic nucleophilic species to trisubstituted unsaturated carbonyl substrates has recently emerged as a powerful technology7–11 (Fig. 1a). These additions are reliable processes, but they generally require controlled reaction conditions and preformed organometallic reagents7–10. In contrast, there has been limited success in developing analogous transformations with nucleo­ hilic carbon-centred radicals. Although a few p examples of metal-catalysed enantioselective radical conjugate additions (RCAs) have been reported12–15, none of these approaches provide for the formation of sterically demanding quaternary carbons. The work we report here was prompted by the desire to address this gap in catalytic enantio­ elective methodology. s Our initial motivation stems from the notion that, because of the long incipient carbon–carbon bond in the early transition state16, additions of radicals to electron-deficient olefins are rather insensitive to steric hindrance3. This makes radical reactivity particularly suited to connecting structurally complex carbon fragments while forging quaternary carbons17. We also recognized that the emerging field of photoredox catalysis4 had recently provided an effective way of generating radicals from bench-stable precursors and under mild conditions. As a result, novel transformations have been invented that capitalize upon non-traditional open-shell mechanisms18. We sought to combine this effective radical generation strategy, which does not require pre-functionalized reagents, with a suitable chiral catalyst that could drive the stereoselective trapping of photogenerated radicals while forging quaternary stereocentres. If successful, this combination would provide direct access to chiral molecules that could not be synthesized using polar conjugate additions. We used the iminium ion activation strategy6 to attack the problem of identifying a suitable chiral catalyst. This chemistry exploits the electrophilic nature of the iminium ion A (Fig. 1b), generated upon condensation of chiral amine catalysts and α,β-unsaturated ketones, to facilitate enantioselective conjugate additions of nucleophiles19. This catalytic platform has found many applications in the polar domain5,6. However, to date, iminium ions A have not been used to trap nucleophilic radicals. This is most surprising, given the high tendency of openshell species to react with electron-deficient olefins3. We reasoned that this dearth of applications could stem from the nature of the radical intermediate B, generated upon carbon–carbon bond formation (Fig. 1c). Generally, olefinic radical traps are electrically neutral and afford long-lived, neutral radical intermediates. In contrast, radical addition to the cationic iminium ion A generates a short-lived, highly reactive α-iminyl radical cation B, which, in line with the classical behaviour of radical ions20, has a high tendency to undergo radical elimination (β-scission)21 to re-form the more stable iminium ion A. The instability of B is the main obstacle to productive RCA to iminium ions (Fig. 1d). We considered the possibility of reducing the radical cation B in situ to generate the corresponding enamine C, which can be hydrolysed in a facile manner to release both the catalyst and the conjugate addition product. From the outset, we identified three design elements as key to realizing this goal. First, the high reactivity of B requires a very rapid single electron transfer (SET) reduction. We hypothesized that using a chiral amine catalyst with a redox active, electron-rich moiety (‘e− pool’ unit in Fig. 1d) attached would secure a fast, proximity-driven intramolecular reduction of B. This idea finds support in the mechanism of electron transfer within biological systems, where even endergonic redox processes can be achieved via electron tunnelling if the redox centres are in close proximity22. Second, we needed to identify a rapid process to interrupt a possible equilibrium between B and the nascent enamine C established by an intramolecular back electron transfer (BET). Since secondary enamines are known to exist mainly as tautomeric electron poor imines D23, the use of a chiral primary amine catalyst potentially offered an efficient mechanism to preclude the BET by triggering a tautomeric equilibrium which converts C into D. Last, the oxidized centre (‘e− hole’ unit in Fig. 1d), arising from the intramolecular SET, had to be long-lived enough to undergo SET reduction from the photoredox catalysts, restoring the redox-active moiety while facilitating productive catalysis. Achieving a high level of stereocontrol further complicated matters. To test the feasibility of this electron-relay strategy24, we explored the reaction between β-methyl cyclohexenone 1a and benzodioxole 2a (Table 1). We used the photocatalyst tetrabutylammonium 1 Institute of Chemical Research of Catalonia (ICIQ), Barcelona Institute of Science and Technology, Avda. Països Catalans 16, 43007 Tarragona, Spain. 2Photogreen Laboratory, Department of Chemistry, University of Pavia, viale Taramelli 12, 27100 Pavia, Italy. 3Catalan Institution for Research and Advanced Studies (ICREA), Passeig Lluís Companys 23, 08010 Barcelona, Spain. *These authors contributed equally to this work. 2 1 8 | NAT U R E | VO L 5 3 2 | 1 4 A P R I L 2 0 1 6 © 2016 Macmillan Publishers Limited. All rights reserved LETTER RESEARCH a Polar reactivity domain O O Metal-based chiral catalyst R R1 R1-M R Preformed organometallic reagents R1-M Cryogenic conditions Metal-based processes Table 1 | Exploratory studies of the feasibility of the electron-relay strategy O + H Quaternary carbon R2 H Simple substrates O Iminium ion activation 1a •R2 O Chiral amine catalyst A R β-scission B H N SET B R R2 BET e– hole N R R2 R R2 e– hole SET PCred PC H2O Amine catalyst Tautomerization C N R1 Unstable radical cation D R R2 + quaternarized product Figure 1 | Conjugate addition technology for forging quaternary stereocentres. a, Established metal-catalysed enantioselective conjugate additions of organometallic reagents (R1-M) via classical polar pathways. b, Design plan for dual photoredox and iminium ion catalysis of radical conjugate additions (RCAs); the grey circle represents the chiral organic catalyst scaffold. c, Challenges associated with implementing iminium ion-catalysed conjugate addition of radicals (·R2). d, Our electron-relay strategy to rapidly remove the short-lived α-iminyl radical cation (B) by intramolecular reduction, and the role of tautomerization to prevent back electron transfer (BET). SET, single electron transfer. PC, photocatalyst; PCred, reduced form of the photocatalyst. The blue ellipse represents an electron rich, reducing moiety, while the red ellipse represents a stable oxidizing species. decatungstate25 (TBADT, 5 mol%) because, upon light excitation, it can easily generate a nucleophilic carbon-centred radical by homolytically cleaving the methylene C–H bond in 2a26 via a hydrogen atom transfer (HAT) mechanism. The experiments were conducted at 35 °C in acetonitrile (CH3CN) and under irradiation by a single ultraviolet (UV)-light emitting diode (UV LED, λmax = 365 nm). We observed a negligible racemic background process in the absence of any amine catalyst, which is necessary for realizing a stereoselective process (entry 1). We then focused on identifying a redox-active moiety that, when installed within the chiral primary amine catalyst, could instigate a fast intramolecular reduction of the transient radical cation B and thus trigger the entire RCA. We identified carbazole as a suitable scaffold because of (i) its excellent electron-donating capabilities, which would provide the e− pool unit, and (ii) the high stability of the long-lived carbazole radical cation27, which makes it a possible e− hole moiety. These properties form the basis of the wide application of carbazoles in hole-transport materials for light-emitting diodes and photovoltaic cells28. The chiral cyclohexylamine scaffold 4b adorned with the carbazole provided the product 3a with appreciable yield and stereoselectivity (33% yield, 82% enantiomeric excess (e.e.), entry 3). In consonance with the proposed electron-relay mechanism, the reaction could not be catalysed by cyclohexylamine 4a, which mimics the catalyst 4b’s scaffold while lacking the redox-active moiety (entry 2). An equimolar combination of 4a and exogenous N-cyclohexyl-3, 6-di-tert-butyl-carbazole (20 mol%) also proved unsuitable for O Me O 3a N R1 4 H + N •R 2 d e– pool 4b–d R = 4e R = Radical addition R H + N R NH2 Quaternarized product R H + N 4a R = H R R2 A R c 2a O H + N O Benzoic acid (40 mol%) TBABF4 (1 equiv.) CH 3CN, 35 oC O Me b Elusive radical reactivity Photoredox activation Catalyst (20 mol%) TBADT (5 mol%) UV LED (365 nm) O Entry 1 2† 3 4† 5 6 7 8 4b R 1 = H 4c R 1 = tBu 4d R 1 = CF3 Catalyst Time (h) Epox/Epred (4) versus Ag/AgCl (V) 3a yield (%) e.e. (%) None 4a 4b 4a* 4c 4d 4e 4e 48 48 48 48 48 48 48 84 NA NA +1.15/−1.32 NA +1.05/−1.28 +1.51/−1.86 +1.10/−1.38 +1.10/−1.38 Trace 4 33 5 35 52 46 75 ND 0 82 0 84 77 93 93 Redox potential values (Epox and Epred) describe the electrochemical properties of the redox active carbazole moiety in 4. NA, not applicable; ND, not determined. *In combination with 20 mol% of exogenous N-cyclohexyl-3,6-di-tert-butyl-carbazole. †Using 40 mol% of trifluoroacetic acid (TFA) instead of benzoic acid. catalysis, suggesting the importance of a proximity-driven intramolecular SET process. We then modified the redox properties of the carbazole scaffold by introducing substituents at the 3- and 6-positions. It is known that this substitution pattern can further stabilize the carbazole radical cation27. Indeed, we could isolate a bench-stable carbazoliumyl radical cation salt from N-cyclohexyl-3, 6-di-tert-butyl-carbazole upon treatment with SbCl 5 (see Supplementary Information). Concurrently, the increased steric hindrance carried the additional benefit of conferring a higher stereocontrol. These considerations explain the high yield and enantio­ selectivity achieved when using the encumbered primary amine catalyst 4e (75% yield and 93% e.e., entry 8). Finally, no product formation was detected in the absence of TBADT, catalyst 4e, or UV light, demonstrating the need for all these components. We then undertook studies to better investigate the role of the active intermediates in the electron-relay mechanism (Fig. 2a). We could synthesize stable tetrafluoroborate salts of the chiral iminium ion A-1, generated upon condensation of catalyst 4c and substrate 1a, which were characterized by X-ray single-crystal analysis (Fig. 2b). The unusual stability of the iminium ion A-1 and the well-defined (Z)-configuration of the C=N double bond originate from a stabilizing intramolecular charge transfer π–π interaction between the electron-rich carbazole and the electron-deficient iminium ion. As a result, the measured interatomic separation in the solid state between the carbazole nitrogen and the sp2 α-carbon of the iminium ion (3.10 Å) is significantly less than the van der Waals distance. This highly organized topology of A-1, which NMR spectroscopic analy­ ses confirmed to be also dominant in solution, plays a critical dual role. On the one hand, it governs the stereocontrol of the RCA, since the bulky carbazole unit is positioned in such a way as to effectively shield the diastereotopic Si face of the iminium ion, leaving the Re face exposed for enantioselective bond formation (Re and Si are stereochemical descriptors for heterotopic faces). Importantly, the sense of asymmetric induction observed in the model reaction is consistent 1 4 A P R I L 2 0 1 6 | VO L 5 3 2 | NAT U R E | 2 1 9 © 2016 Macmillan Publishers Limited. All rights reserved RESEARCH LETTER a Iminium ion cycle O R-H H Me 1a X– N N + PCred R Photoredox cycle h PC PC* System 1: TBADT - HAT mechanism h O TBADT* H oxidant O 3 Me A-1 N R Me NH2 H X– TBADT 4– N N+ 4 – H+ N E-1 SET D-1 Å H O H +N N + Ar – H+ reductant O O Me R1 + Acid (40 mol%) TBABF4 (1 equiv.) CH3CN, 35 °C, 72 h 1b Ar Ar Ir PCred Amine (20 mol%) TBADT (5 mol%) UV LED (365 nm) R1-H (2a) H N N II E-1 c –BF 4 7 C-1 Me SET R D-1 Me b 09 IrIII Tautomerization PC Ar *IrIII oxidant R Me PC red 3. h e– hole N Z-configured C=N bond System 2: Iridium catalyst - SET mechanism HN X– + N R Me O reductant E-1 X– + N N TBADT 5– –H PCred SET D-1 Electron relay mechanism H2O 2a O B-1 R Me O HAT e– pool R1 A-1 Re face exposed R1 = CCDC 1437991 exo-5 3b O Amine 4a + TFA 6% yield 3% yield O Me Amine 4e + PhCO2H Traces 40% yield 83% e.e. Figure 2 | Proposed mechanism and mechanistic investigations. a, Synergistic activities of the iminium ion and the photoredox catalytic cycles to realize the enantioselective RCA to enone 1a. Upon radical addition to the iminium ion A-1, the electron-relay mechanism rapidly reduces the unstable radical cation B-1 producing a carbazoliumyl radical cation C-1, which is prevented from undergoing BET by tautomerization of the secondary enamine to the corresponding imine D-1. Regeneration of the photocatalyst (PC) is achieved by reduction of the carbazoliumyl radical cation in D-1, while the aminocatalyst 4 is liberated upon hydrolysis of imine E-1. b, X-ray crystal structure of the carbazole-based iminium ion A-1: the distance between the carbazole nitrogen and the sp2 α-carbon of the cyclohexene moiety is highlighted. c, Cyclization experiments indicating that the α-iminyl radical intermediate B-1 has been bypassed when using the carbazole-based catalyst 4e. with this stereochemical model (Fig. 2b). On the other hand, the three-dimensional assembly of A-1 suggests that the α-iminyl radical cation B-1, arising from the radical trapping, is generated in close proximity to the electron-rich carbazole, allowing for a proximitydriven22 intramolecular reduction. Once the carbazoliumyl radical cation (e− hole) is generated, the fast tautomerization of the secondary (S,S)-4e (20 mol%) TBADT (5 mol%) Single UV LED (365 nm) O + () n n = 0–3 O R2 1 2 Me O 3a 75% yield 93% e.e. 3c 69% yield 97% e.e. Et O Me O Me O O O Me 3g 70% yield 94% e.e. 3h 63% yield 91% e.e. Me O O OTBS 3e 57% yield 85% e.e. 3d 56% yield 96% e.e. O Me O () n O R1 O Me 3i 49% yield 91% e.e. N NH2 Me O Me O 3f 99% yield 88% e.e. (S,S)-4e Organic catalyst O Br Me O Br Figure 3 | Substrate scope for the enantioselective trapping of benzodioxole-derived radicals via the dual photoredox organocatalytic strategy. Survey of the cyclic enones 1 and substituted benzodioxoles 2 that can participate in the organocatalytic asymmetric radical conjugate addition (RCA) to forge quaternary stereocentres (as in 3). Yields and enantiomeric excesses of the isolated products are indicated below each O 3 Me O Me O O Me R2 O O O O Me O Benzoic acid (40 mol%) TBABF4 (1 equiv.) CH3CN, 35 oC, 72–96 h O R1 O O O Photoredox system 1 TBADT - HAT mechanism O O Me CCDC 1437992 Br 3j 67% yield 1.5:1 d.r., 98% e.e. O Me O Cl O Me 3k 61% yield 1:1 d.r., 98% e.e. entry (3a to 3k). Details of the TBADT-mediated photoredox cycle to produce carbon-centred radicals from 2 via a HAT mechanism are reported in Fig. 2a. The carbazole-based organic catalyst 4e is drawn in the boxed inset. TBS, tert-butyldimethylsilyl; TBABF4, tetrabutylammonium tetrafluoroborate; d.r., diastereomeric ratio. 2 2 0 | NAT U R E | VO L 5 3 2 | 1 4 A P R I L 2 0 1 6 © 2016 Macmillan Publishers Limited. All rights reserved LETTER RESEARCH a O () n b R N Me + n = 0–3 (R,R)-4e (20 mol%) Ir III catalyst 8 (1 mol%) White LEDs strip R2 1 Benzoic acid (40 mol%) Toluene, 15 °C, 48 h 6 1 Organic catalyst (R,R)-4e O CF 3 2 R1 R N () n t-Bu N N 7 N O O O Me Me N O Et Me N 7a 78% yield 88% e.e. N 7c Me 52% yield 64% e.e. 7b 85% yield 84% e.e. O O Me Me 7f 83% yield 82% e.e. 7g 81% yield 80% e.e. O Me N Ph Me 7k 85% yield 1.6:1 d.r., 94% e.e.major 72% e.e.minor c O + Me 1a Me Br O Me N (R,R)-4e (20 mol%) Benzophenone 9 (10 mol%) UV LEDs strip 6a Benzoic acid (40 mol%) Toluene, 15 °C, 72 h Me F Photoredox system 2 Iridium catalyst 8 - SET mechanism 7e 62% yield 90% e.e. Me N 7i 52% yield 80% e.e. CCDC 1437993 F Ir[dF(CF 3)ppy ] 2(dtbbpy)PF6 (8) N 7h 69% yield 87% e.e. Br F CF 3 O N Me 7d 92% yield 88% e.e. O Me Me N Me Me N – PF 6 Ir N t-Bu O Me Me N Me F Br O Me N 7j 71% yield 88% e.e. O Me Me N 7a 76% yield 88% e.e. Me N Me 9 N Me Me Figure 4 | Substrate scope for the enantioselective trapping of α-amino radicals via the dual photoredox organocatalytic strategy. a, The photochemical organocatalytic radical conjugate addition (RCA) developed to forge quaternary stereocentres. b, Survey of the cyclic enones 1 and tertiary amines 6 that can participate in the reaction. Yields and enantiomeric excesses (e.e.) of the isolated products 7a–k are indicated below each entry. Details of the iridium-mediated photoredox cycle to produce carbon-centred radicals via a SET mechanism are reported in Fig. 2a. The structure of the iridium photoredox catalyst 8 is given in the boxed inset. c, Fully organocatalytic enantioselective RCA using the benzophenone photocatalyst 9. enamine C-1 to afford the imines D-1 precludes the BET. At this point, the long-lived radical cation in D-1 (Epred > −1.28 V versus Ag/Ag+ in CH3CN, see Table 1) can be reduced by the photocatalyst (half-wave potential E1/2 [TBADT4−/TBADT5−−H] = −0.96 V versus Ag/Ag+ in CH3CN)29, closing the photoredox cycle (Fig. 2a, right panel). The iminium ion cycle terminates with the imine E-1 hydrolysis to regenerate the catalyst 4 while liberating the product 3. To gain further evidence supporting the electron-relay mechanism, we used the enone 1b, bearing a β-homoallyl substituent, to trap the radical photogenerated from 2a (Fig. 2c). The reaction catalysed by cyclohexylamine 4a provides preferentially the cyclized exo-adduct 5 (5:3b in a 2:1 ratio), demonstrating the propensity of the α-iminyl radicals, emerging from the radical addition, to undergo cyclization with unactivated olefins. In sharp contrast, the process catalysed by the amine 4e almost exclusively affords the conjugate addition product 3b (40% yield, 83% e.e., 3b:5 in a >10:1 ratio). This result is consonant with a fast redox process, governed by the carbazole-based catalyst, which rapidly reduces the α-iminyl radical cation B-1 preventing cyclization. Adopting the optimized conditions described in Table 1, entry 8, we then demonstrated the generality of the RCA by evaluating a variety of cyclic enones 1 and benzodioxoles 2 (Fig. 3). The presence of a methyl substituent at the methylene position of 2 provides the corresponding product 3c, bearing two adjacent tetrasubstituted carbons, with nearly perfect enantioselectivity. Experiments to probe the scope of the enone component revealed that a wide range of carbocycles and β-olefin substituents are well tolerated. For example, high levels of stereocontrol are achieved with different β-alkyl groups (products 3d, e) and with a diverse range of ring sizes, including cyclopentenyl, cycloheptenyl, and cyclooctenyl architecture (adducts 3f–h). One limitation is that the presence of an aromatic β-substituent completely inhibits the reaction. As for the benzodioxole substrates 2, different substituents can be installed at the aromatic ring without compromising the efficiency of the reaction (adducts 3i–k). Crystals of compound 3i were suitable for X-ray analysis, which secured the absolute configuration of the products. We then wondered if the synthetic utility of the amine carbazole catalyst 4e could be expanded to trap other carbon-centred radicals, formed through an unrelated light-triggered mechanism, while forging quaternary stereocentres. Specifically, we used the commercially available photocatalyst Ir[dF(CF3)ppy]2(dtbbpy)PF6 (8) which, upon absorption of visible light, can generate α-amino radicals directly from tertiary amines 6 via single electron oxidation18 (SET mechanism in Fig. 2a). The conjugate addition adducts 7 were provided with high stereoselectivity by using the combination of catalysts (R,R)-4e and 8 while conducting the reactions with enones 1 at 15 °C in toluene and under irradiation by a white LED (λemiss > 400 nm) (Fig. 4a). We next explored the scope of both substrates in this dual photoredox organocatalytic strategy. As highlighted in Fig. 4b, cyclic enones of different ring sizes (7c–e) and bearing alkyl β-substituents (products 7a, b) are suitable substrates, while both mixed N-alkyl-N-aryl (adducts 7a, f, g) and N,N-diaryl tertiary amines (7h–j) efficiently participated in the RCA. Substituents of different electronic nature were easily accommodated at the aryl para (7f, g, i) or ortho position (7j), while a cyclic amine afforded compound 7k with high enantiomeric purity, albeit with a 3:2 diastereomeric ratio. For this enantioselective trap of α-amino radicals, we determined a quantum yield of 0.4 (λ = 400 nm), while Stern–Volmer fluorescence quenching experiments demonstrated that the excited state of the photocatalyst 8 is quenched by the amine 6. Both experiments are consistent with the electron-relay mechanism depicted in Fig. 2a. Notably, the RCA can be performed without any metal when replacing the photocatalyst 8 with the benzophenone 9, which can generate the radical acting as an organic photosensitizer30 (Fig. 4c). We have developed the first (to our knowledge) catalytic strategy that allows quaternary carbon stereocentres to be obtained with high fidelity using an enantioselective RCA manifold. The approach 1 4 A P R I L 2 0 1 6 | VO L 5 3 2 | NAT U R E | 2 2 1 © 2016 Macmillan Publishers Limited. All rights reserved RESEARCH LETTER requires mild conditions and unfunctionalized substrates, effectively complementing established polar conjugate addition technologies based on preformed organometallic reagents. Received 23 November 2015; accepted 17 February 2016. 1. Quasdorf, K. W. & Overman, L. E. Catalytic enantioselective synthesis of quaternary carbon stereocentres. Nature 516, 181–191 (2014). 2. Murakata, M., Jono, T., Mizuno, Y. & Hoshino, O. Construction of chiral quaternary carbon centers by catalytic enantioselective radical-mediated allylation of α-iodolactones using allyltributyltin in the presence of a chiral Lewis acid. J. Am. Chem. Soc. 119, 11713–11714 (1997). 3. Fischer, H. & Radom, L. Factors controlling the addition of carbon-centered radicals to alkenes — an experimental and theoretical perspective. Angew. Chem. Int. Ed. 40, 1340–1371 (2001). 4. Schultz, D. M. & Yoon, T. P. Solar synthesis: prospects in visible light photocatalysis. Science 343, 1239176 (2014). 5. MacMillan, D. W. C. The advent and development of organocatalysis. Nature 455, 304–308 (2008). 6. Lelais, G. & MacMillan, D. W. C. Modern strategies in organic catalysis: the advent and development of iminium activation. Aldrichim. Acta 39, 79–87 (2006). 7. Hawner, C. & Alexakis, A. Metal-catalyzed asymmetric conjugate addition reaction: formation of quaternary stereocenters. Chem. Commun. 46, 7295–7306 (2010). 8. Hird, A. W. & Hoveyda, A. H. Catalytic enantioselective alkylations of tetrasubstituted olefins. Synthesis of all-carbon quaternary stereogenic centers through Cu-catalyzed asymmetric conjugate additions of alkylzinc reagents to enones. J. Am. Chem. Soc. 127, 14988–14989 (2005). 9. Shintani, R., Tsutsumi, Y., Nagaosa, M., Nishimura, T. & Hayashi, T. Sodium tetraarylborates as effective nucleophiles in rhodium/diene-catalyzed 1,4-addition to β,β-disubstituted α,β-unsaturated ketones: catalytic asymmetric construction of quaternary carbon stereocenters. J. Am. Chem. Soc. 131, 13588–13589 (2009). 10. Liu, Y., Han, S.-J., Liu, W.-B. & Stoltz, B. M. Catalytic enantioselective construction of quaternary stereocenters: assembly of key building blocks for the synthesis of biologically active molecules. Acc. Chem. Res. 48, 740–751 (2015). 11. Sidera, M., Roth, P. M. C., Maksymowicz, R. M. & Fletcher, S. P. Formation of quaternary centers by copper-catalyzed asymmetric conjugate addition of alkylzirconium reagents. Angew. Chem. Int. Ed. 52, 7995–7999 (2013). 12. Srikanth, G. S. C. & Castle, S. L. Advances in radical conjugate additions. Tetrahedron 61, 10377–10441 (2005). 13. Sibi, M. P., Ji, J., Wu, J. H., Gürtler, S. & Porter, N. A. Chiral Lewis acid catalysis in radical reactions: enantioselective conjugate radical additions. J. Am. Chem. Soc. 118, 9200–9201 (1996). 14. Gansäuer, A., Lauterbach, T., Bluhm, H. & Noltemeyer, M. A catalytic enantioselective electron transfer reaction: titanocene-catalyzed enantioselective formation of radicals from meso-epoxides. Angew. Chem. Int. Ed. 38, 2909–2910 (1999). 15. Ruiz Espelt, L., McPherson, I. S., Wiesnsch, E. M. & Yoon, T. P. Enantioselective conjugate additions of α-amino radicals via cooperative photoredox and Lewis acid catalysis. J. Am. Chem. Soc. 137, 2452–2455 (2015). 16. Damm, W. et al. Diastereofacial selectivity in reactions of substituted cyclohexyl radicals. An experimental and theoretical study. J. Am. Chem. Soc. 114, 4067–4079 (1992). 17. Schnermann, M. J. & Overman, L. E. A concise synthesis of (−)-Aplyviolene facilitated by a strategic tertiary radical conjugate addition. Angew. Chem. Int. Ed. 51, 9576–9580 (2012). 18. Prier, C. K., Rankic, D. A. & MacMillan, D. W. C. Visible light photoredox catalysis with transition metal complexes: applications in organic synthesis. Chem. Rev. 113, 5322–5363 (2013). 19. Gu, X. et al. A general, scalable, organocatalytic nitro-Michael addition to enones: enantioselective access to all-carbon quaternary stereocenters. Org. Lett. 17, 1505–1508 (2015). 20. Poniatowski, A. J. & Floreancig, P. E. in Carbon-Centered Free Radicals and Radical Cations: Structure, Reactivity, and Dynamics Ch. 3 (ed. Forbes, M. D. E.) 43–49 (Wiley & Sons, 2010). 21. Jakobsen, H. J., Lawesson, S. O., Marshall, J. T. B., Schroll, G. & Williams, D. H. Mass spectrometry. XII. Mass spectra of enamines. J. Chem. Soc. B 940–946 (1966). 22. Page, C. C., Moser, C. C., Chen, X. & Dutton, P. L. Natural engineering principles of electron tunnelling in biological oxidation-reduction. Nature 402, 47–52 (1999). 23. Häfelinger, G. & Mack, H.-G. in The Chemistry of Enamines Ch. 1 (ed. Rappaport, Z.) 1–87 (Wiley & Sons, 1994). 24. Okada, Y., Nishimoto, A., Akaba, R. & Chiba, K. Electron-transfer-induced intermolecular [2+2] cycloaddition reactions based on the aromatic “redox tag” strategy. J. Org. Chem. 76, 3470–3476 (2011). 25. Tzirakis, M. D., Lykakis, I. N. & Orfanopoulos, M. Decatungstate as an efficient photocatalyst in organic chemistry. Chem. Soc. Rev. 38, 2609–2621 (2009). 26. Ravelli, D., Albini, A. & Fagnoni, M. Smooth photocatalytic preparation of 2-substituted 1,3-benzodioxoles. Chem. Eur. J. 17, 572–579 (2011). 27. Prudhomme, D. R., Wang, Z. & Rizzo, C. J. An improved photosensitizer for the photoinduced electron-transfer deoxygenation of benzoates and m-(trifluoromethyl)benzoates. J. Org. Chem. 62, 8257–8260 (1997). 28. Blouin, N. & Leclerc, M. Poly(2,7-carbazole)s: structure-property relationships. Acc. Chem. Res. 41, 1110–1119 (2008). 29. Yamase, T., Takabayashi, N. & Kaji, M. Solution photochemistry of tetrakis(tetrabutylammonium) decatungstate(VI) and catalytic hydrogen evolution from alcohols. J. Chem. Soc. Dalton Trans. 793–799 (1984). 30. Bertrand, S., Hoffmann, N. & Pete, J.-P. Highly efficient and stereoselective radical addition of tertiary amines to electron-deficient alkenes — application to the enantioselective synthesis of Necine bases. Eur. J. Org. Chem. 2227–2238 (2000). Supplementary Information is available in the online version of the paper. Acknowledgements Financial support was provided by the ICIQ Foundation, MINECO (project CTQ2013-45938-P and Severo Ochoa Excellence Accreditation 2014-2018, SEV-2013-0319), AGAUR (2014 SGR 1059), and the European Research Council (ERC 278541, ORGA-NAUT). J.J.M. and S.P. thank the Marie Curie COFUND (291787-ICIQ-IPMP) and the CELLEX Foundation, respectively, for postdoctoral fellowships. We thank M. Minozzi, M. Nappi and E. Raluy for preliminary investigations, D. Merli and D. Dondi for assistance with EPR experiments, and D. Ravelli for discussions. Author Contributions J.J.M., D.B. and S.P. performed the experiments and analysed the data. J.J.M., D.B., S.P., and P.M. designed the experiments. M.F. and P.M. conceived the project. P.M. directed the project, and P.M. and J.J.M. wrote the manuscript with contributions from all the authors. Author Information Crystallographic data for the iminium ion A-1 and for compounds 3i and 7h have been deposited with the Cambridge Crystallographic Data Centre, accession numbers CCDC 1437991, 1437992 and 1437993, respectively. Reprints and permissions information is available at www.nature.com/reprints. The authors declare no competing financial interests. Readers are welcome to comment on the online version of the paper. Correspondence and requests for materials should be addressed to P.M. (pmelchiorre@iciq.es). 2 2 2 | NAT U R E | VO L 5 3 2 | 1 4 A P R I L 2 0 1 6 © 2016 Macmillan Publishers Limited. All rights reserved