Abstract
The substitution of an alkyl electrophile by a nucleophile is a foundational reaction in organic chemistry that enables the efficient and convergent synthesis of organic molecules. Whereas substantial progress has been reported in recent years in exploiting transition-metal catalysis to dramatically expand the scope of nucleophilic substitution reactions using carbon nucleophiles1–,2,3,4, there has been limited progress in corresponding reactions with nitrogen nucleophiles5–6,7,8. Furthermore, for many substitution reactions, the bond construction itself is not the only challenge, as there is a need to control stereochemistry at the same time. Here, we describe a method for the enantioconvergent substitution of unactivated racemic alkyl electrophiles by a ubiquitous nitrogen-containing functional group, an amide, through the use of a photoinduced catalyst system based on copper, an earth-abundant metal. This process for asymmetric N-alkylation relies upon three distinct ligands (a bisphosphine, a phenoxide, and a chiral diamine) that assemble, in situ, a copper/bisphosphine/phenoxide complex that serves as a photocatalyst and a chiral copper/diamine complex that catalyzes enantioselective C–N bond formation. This study thus expands enantioselective N-substitution by alkyl electrophiles beyond activated electrophiles (those bearing at least one sp- or sp2-hybridized substituent on the carbon undergoing substitution)8–,9,10,11,12,13 to include unactivated electrophiles.
Amides, including chiral compounds of type A (Fig. 1a), are a ubiquitous functional group in bioactive molecules14–,15,16, and, correspondingly, amide bond formation through the coupling of a carboxylic acid (or derivative) with an amine is perhaps the most widely used reaction in medicinal chemistry17,18. The most common method for the synthesis of enantioenriched amides such as A is via the acylation of an enantioenriched amine (Fig. 1a: N-acylation). Although the N-alkylation of a primary amide by an enantioenriched electrophile would represent a complementary approach to chiral secondary amides of type A (Fig. 1a: N-alkylation), to the best of our knowledge such an SN2 reaction has not been described, presumably due to the relatively poor reactivity of unactivated secondary alkyl electrophiles.
Fig. 1. Chiral secondary amides.
a, Strategies for their synthesis. b, This study. Ar, Ar1 = aromatic substituents, R = carbon substituent, X = leaving/anionic group, t-Bu = tert-butyl, e.e. = enantiomeric excess, equiv. = equivalent, Et = ethyl, Me = methyl, Ph = phenyl, i-Pr = isopropyl.
Despite substantial limitations such as this, constructing C–N bonds via the SN2 reaction of an alkyl electrophile with a nitrogen nucleophile is a widely exploited transformation in organic synthesis, reportedly the most frequently used method for achieving N-substitution19 and the fifth most-used reaction overall in the pharmaceutical industry17. Unfortunately, expanding the scope of C–N coupling reactions of alkyl electrophiles with nitrogen nucleophiles via classical substitution pathways (i.e., SN2 or SN1) is challenging, stymied by E2 elimination (SN2) and by deactivation of the nucleophile by protonation or formation of a Lewis acid-base complex (SN1); furthermore, neither of these classical pathways is readily amenable to enantioconvergent C–N bond formation beginning with racemic electrophiles (generally, inverted stereochemistry for SN2 reactions and racemic products for SN1 reactions).
Recognizing the potential impact of the development of powerful new strategies for N-alkylation, we and others have pursued the discovery of catalyzed processes, establishing that transition metals such as copper can enable bond constructions that had not previously been achieved, via new pathways5,6,10,12,20–,21,22,23. For example, in the case of photoinduced, copper-catalyzed reactions, the alkyl electrophile is transformed into an alkyl radical prior to C–N bond formation6.
With regard to our goal of developing a new, complementary approach to the synthesis of chiral secondary amides of type A (via N-alkylation, rather than N-acylation; Fig. 1a), a radical-based pathway opens the door to exploiting catalysis not only to increase the scope of suitable reaction partners, but also to control the stereochemistry of the product using a racemic electrophile. Thus, rather than needing to start with an enantiopure electrophile (Fig. 1a: failed alternative approach), a chiral catalyst would activate a readily available racemic electrophile and simultaneously achieve C–N bond formation and stereochemical control (Fig. 1b). It is important to note that, to obtain good enantioselectivity in the synthesis of chiral amide A, the chiral catalyst must effectively distinguish the two alkyl substituents of the unactivated secondary electrophile (alkyl vs alkyl1 in Fig. 1); in asymmetric synthesis, differentiating between alkyl groups (e.g., ethyl vs n-propyl) is generally substantially more difficult than differentiating between an alkyl and an unsaturated group (e.g., ethyl vs phenyl)24.
In this report, we describe the achievement of our proposed strategy for the enantioconvergent synthesis of chiral secondary amides via couplings of primary amides with unactivated racemic alkyl electrophiles through the use of a copper-based catalyst system that is activated by blue-LED irradiation (Fig. 1b). Key to this new method is the use of three ligands, a bisphosphine and a phenoxide that generate a copper photocatalyst that is highly reducing in its excited state and thereby capable of activating the unactivated electrophile, as well as a diamine that affords a chiral copper catalyst that simultaneously achieves C–N bond formation and control of stereochemistry.
As noted above, to achieve enantioconvergent N-alkylation in the case of an unactivated alkyl electrophile, the chiral catalyst must differentiate between two alkyl groups, which is often a difficult task in asymmetric catalysis. To obtain good enantioselectivity in challenging situations such as this, a common strategy is to exploit a Lewis-basic functional group that can provide a two-point interaction between the substrate and the chiral catalyst in the stereochemistry-determining step of the process, thereby enabling the differentiation of two similar substituents25. Accordingly, in our initial studies, we examined photoinduced, copper-catalyzed enantioconvergent N-alkylations with electrophiles that include a phosphonyl group, since γ-aminophosphonic acids have been shown to be bioactive26,27. Upon exploring a variety of reaction conditions, we determined that the desired catalytic asymmetric synthesis of secondary amides can be achieved in good yield and enantiomeric excess (e.e.) upon irradiating a copper-based catalyst system with blue-LED lamps (Fig. 2, 1; 95% yield and 93% e.e.). By comparison, under the same conditions, racemic 3-bromoheptane undergoes substitution to furnish the secondary amide in 59% yield and 8% e.e.
Fig. 2. An array of amides serve as nucleophilic coupling partners.
Couplings were generally conducted using 0.5 mmol of the amide. All data represent the average of two experiments. The percent yield represents purified product. Boc = tert-butoxycarbonyl, i-Bu = isobutyl, n-Bu = n-butyl, d.r. = diastereomer ratio.
The stereochemistry of the product (93% e.e. (S)) is the same regardless of whether (rac)-P, (R)–P, or (S)–P is used (see the Supplementary Information), establishing that the stereochemistry of diamine N1*, not that of bisphosphine P, determines the stereochemistry of the amide. Control experiments established that, in the absence of copper, light, or bisphosphine P, essentially no C–N coupling is observed (<1% yield; Fig. 2, 1). Furthermore, in the absence of diamine N1*, only a small amount of the coupling product is obtained (7% yield), pointing to cooperativity between diamine N1* and bisphosphine P in achieving C–N bond formation.
Under our standard conditions, the corresponding alkyl iodide undergoes C–N coupling with similar yield and e.e. (86% yield, 92% e.e.) as the bromide (Fig. 2, 1), whereas the corresponding chloride is essentially unreactive (<5% conversion). From a practical point of view, it is worthwhile to note that the reaction is not highly air- or moisture-sensitive, proceeding well in the presence of 0.1 mL of air or of one equivalent of water (>85% yield, 92% e.e.). On a gram scale, C–N coupling occurs with slightly diminished yield and similar enantioselectivity (85% yield and 92% e.e.). The combination of high yield and high e.e., while using only a small excess of the racemic electrophile (1.2 equiv.), establishes that a stereoconvergent coupling, not a simple kinetic resolution, of the electrophile is operative.
A diverse array of amides serve as useful nucleophiles in this photoinduced, copper-catalyzed enantioconvergent N-alkylation by unactivated electrophiles (Fig. 2). For example, a variety of aromatic amides, including para-, meta-, and ortho-substituted compounds, are suitable nucleophiles, providing secondary amides in good yield and enantioselectivity (1–8). Furthermore, heterocyclic amides (bearing a furan, a thiophene, or an indole; 9–11) and aliphatic amides, including sterically demanding nucleophiles (12–15), are effective coupling partners. Moreover, this asymmetric N-alkylation can be applied to more complex amides (16–19); for couplings of chiral primary amides, the stereochemistry of the chiral catalyst, rather than the stereochemistry of the chiral nucleophile, primarily determines the stereochemistry of the product (17–19). In contrast to the classical approach to the synthesis of such chiral secondary amides (N-acylation in Fig. 1a; exception: kinetic resolution/acylation of amines18), this method enables a new, convergent strategy wherein stereochemical control and C–N bond formation are accomplished in a single step.
The scope with respect to the phosphonyl-substituted unactivated secondary electrophile is also fairly broad. For example, the R substituent can vary in steric demand from Me to isobutyl (Fig. 3, 20–22). Furthermore, the method is compatible with an array of functional groups, including trifluoromethyl, an acetal, an alkene, an alkyne, and a ketone (23–28); moreover, through additive studies we have determined that the presence of an unactivated alkyl chloride, aniline, aryl triflate, benzofuran, benzothiophene, epoxide, N-methylindole, nitrile, tertiary amide, or thioether does not significantly impact the yield or the e.e. of the coupling to generate 1 in Fig. 2 (>80% yield and >90% e.e. of product, along with >95% recovery of the additive; see the Supplementary Information). Finally, the phosphonyl group can bear a range of substituents other than OEt (29–31).
Fig. 3. An array of alkyl bromides serve as electrophilic coupling partners.
Couplings were generally conducted using 0.5 mmol of the amide, and N1* was used as the diamine, unless otherwise noted. All data represent the average of two experiments. The percent yield represents purified product. a: N2* as the diamine, no Cs2CO3, 2-Me-THF as solvent. b: N2* as the diamine. c: electrophile (1.5 equiv.), Cu(CH3CN)4PF6 (15 mol%), P (5 mol%), N1* (20 mol%), K3PO4•H2O (1.5 equiv.; in place of Cs2CO3), 10 °C. Ar1 = p-(F3C)C6H4, DG = directing group, Ac = acetyl, Bn = benzyl, TIPS = triisopropylsilyl.
Furthermore, the same approach can be applied to photoinduced, copper-catalyzed enantioconvergent N-alkylations of amides by other classes of racemic unactivated secondary alkyl electrophiles. Thus, in addition to a phosphonyl group (Fig. 2 and 20–31 in Fig. 3), a diverse array of other Lewis-basic functional groups, including an amide, ester, ketone, sulfone, sulfonamide, and phosphine oxide, enable the often-challenging differentiation between two alkyl substituents to provide good enantioselectivity in the nucleophilic substitution reaction (32–53 in Fig. 3 and 54–64 in Extended Data), leading to derivatives of γ-amino acids, 1,2-diamines, and the like28–,29,30. Whereas ligand-directed asymmetric catalysis is now well-established, the ability to utilize such a wide array of directing groups for a particular reaction is noteworthy31.
We have pursued mechanistic studies to elucidate the role that the various ligands play in this photoinduced, copper-catalyzed enantioconvergent N-alkylation process. Our working hypothesis is that, in contrast to all previous examples of enantioconvergent N-alkylations (all of which utilize activated electrophiles8–,9,10,11,12,13), more than one catalytic cycle is involved in this new process.21 Specifically, we believe that one copper complex serves as a photocatalyst, cleaving the relatively strong (unactivated) C–Br bond, and a second copper complex effects enantioselective C–N bond formation (Fig. 4a).
Fig. 4. Mechanism.
a, Outline of a possible catalytic cycle. b, Redox potentials (vs Fc+/Fc). c, Stern–Volmer quenching of [PCuI(OPh)]* by 2: irradiation at 350 nm, emission at 496 nm. d, EPR studies: X-band spectra (9.4 GHz, 77 K). e, Investigation of an organic radical (R•) as an intermediate: TEMPO trapping and radical-clock experiments. f, Reactions of enantioenriched electrophile. Ar = p-(F3C)C6H4, TEMPO = 2,2,6,6-tetramethyl-1-piperidinyloxy.
With respect to the left-hand catalytic cycle (Fig. 4a), our observation that, for the N-alkylation illustrated in Fig. 2 (1), the electrophile is consumed in the absence of diamine N1* (>99% conversion of the electrophile; 7% yield of the product), but not in the absence of bisphosphine P (<2% conversion of the electrophile), is consistent with a P-bound copper complex serving as the photocatalyst. With respect to the right-hand catalytic cycle (Fig. 4a), our observation that the stereochemistry of diamine N1*, not that of bisphosphine P, determines the stereochemistry of the product is consistent with an N1*-bound copper complex being responsible for C–N bond construction.
Regarding the left-hand catalytic cycle, we determined that two representative unactivated electrophiles have reduction potentials of −2.3 to −2.4 V (Fig. 4b: 65 and 66; all redox potentials are referenced to Fc+/Fc), placing them beyond the reach of conventional ruthenium/iridium photocatalysts32 but within the reach of CuI/bisphosphine complexes33. Two PCuIX complexes that might plausibly serve as photocatalysts under our conditions are PCuI(OPh) and PCuI(amidate). We synthesized each complex via the treatment of CuCl with bisphosphine P and X− (X = OPh, amidate; amidate = p-(F3C)C6H4CONH) and then crystallographically characterized them (see the Supplementary Information).
Our observations point to PCuI(OPh) complex as the more likely photocatalyst under our reaction conditions. Thus, during a coupling, PCuI(OPh) accounts for ~35% of the copper-containing species that are present, whereas PCuI(amidate) is not detected (see the Supplementary Information). Furthermore, the estimated excited-state reduction potential of PCuI(OPh) (−2.8 V), but not of PCuI(amidate) (−2.3 V), is lower than that of the electrophiles (−2.3 and −2.4 V). The excited state of PCuI(OPh) has a relatively long lifetime (4.6 μs), and a Stern–Volmer study established that luminescence of the excited state of PCuI(OPh) (emission at 496 nm) is readily quenched by electrophile 66 with a Stern−Volmer constant of 189 M−1 and a quenching rate constant of 4.1 × 107 M−1s−1 (Fig. 4c); in contrast, the excited state of PCuI(amidate) is not effectively quenched by this electrophile (see the Supplementary Information). Alternatively, the reaction of the excited state of PCuI(OPh) with the electrophile could proceed by an inner-sphere electron-transfer pathway (halogen-atom transfer).34
The two catalytic cycles of the proposed coupling pathway intersect at the reaction of D and G to generate B and E (Fig. 4a). Two of the plausible mechanisms by which this process might occur are via electron transfer or via ligand exchange. To corroborate our intuition about the likely feasibility of electron transfer between complexes D and G, we carried out DFT calculations (BP86, def2-TZVP, SMD(THF) level of theory), which predict redox potentials (E0 for CuI/CuII) of 1.3 V and 1.2 V for PCuII(OPh)2 and PCuII(OPh)Br, respectively, and −0.6 V for both (N2*)CuI(OPh) and (N2*)CuI(amidate) (Fig. 4b). Taken together, these computed redox potentials support the viability of electron transfer from (N2*)CuIX (G) to PCuIIX2 (D).
Alternatively, D and G could generate B and E via ligand exchange. To explore this possibility, we treated PCuI(OPh) with Magic Blue (tris(4-bromophenyl)ammoniumyl hexachloroantimonate, a one-electron oxidant; 1.0 equiv.) in the presence of N2* (1.0 equiv.) and NaOPh (10 equiv.) in 2-Me-THF. An EPR spectrum taken after 2 minutes at room temperature (experiment 1 in Fig. 4d) revealed a spectrum consistent with (N2*)CuII(OPh)2 (see below). This observation is consistent with rapid ligand exchange (P → N2*) upon oxidation of CuI to CuII, with the latter preferring the harder diamine ligand to the softer bisphosphine ligand. Thus, either or both of these pathways for the formation of B and E from D and G may be operative.
We carried out a series of EPR spectroscopic studies in order to gain insight into the CuII species that are present under our coupling conditions. Treatment of CuCl2 with N2* (1.0 equiv.) and NaOPh (2.0 equiv.) in 2-Me-THF afforded a blue solution with the EPR spectrum illustrated in Fig. 4d (experiment 2), which we assign to (N2*)CuII(OPh)2. Furthermore, reaction of CuBr2 with N2* (1.0 equiv.) and NaOPh (1.0 equiv.) furnished a blue solution with the EPR spectrum shown in Fig. 4d (experiment 3), which we attribute to (N2*)CuII(OPh)Br. An EPR spectrum of a coupling reaction after four hours at −5 °C appeared to be an ~1:1 mixture of the species shown in experiments 2 and 3 (Fig. 4d, experiment 4; for a simulation, see the Supplementary Information); spin quantification of the EPR-active copper species indicated that ~40% of the copper in the reaction mixture is present as these two CuII complexes. ESI–MS analysis of a reaction mixture confirmed the presence of (N2*)CuII(OPh)2 during the coupling process (see the Supplementary Information).
With regard to the proposed right-hand catalytic cycle (Fig. 4a), nucleophilic substitution of one or both of these CuII/N2* complexes (E) by the amidate anion leads to (N2*)CuIIX(amidate) (F), which reacts with organic radical R• in an out-of-cage process, via coordination of the directing group to CuII followed by C–N bond formation, to furnish the desired coupling product and (N2*)CuIX (G). Consistent with the hypothesis that C–N bond formation proceeds via an out-of-cage reaction of R•, we observe no C–N coupling (<1%) in the presence of TEMPO, a radical trap; instead, a TEMPO adduct of R• can be detected (20% yield; Fig. 4e; the sodium salt of TEMPO does not react with alkyl bromide 2 under these conditions (see the Supplementary Information)).
A radical-cyclization study (Fig. 4e) provides further support for the intermediacy of an organic radical and for the out-of-cage pathway for C–N bond formation that is depicted in Fig. 4a. Thus, under our standard reaction conditions, alkyl bromide 67 reacts with a primary amide to predominantly generate cyclized product 68 with 3:1 diastereoselectivity (<5% of the acyclic coupling product). The diastereoselectivity matches that observed in the n-Bu3SnH-mediated reductive cyclization of the same electrophile, consistent with a common intermediate (a secondary organic radical) in the two processes. The rate constant for the cyclization of related radicals is ~1×105 s−1 at 25 °C35, which is much slower than typical diffusion rates (generally >108 s−1)36. These observations are thus consistent with the formation of an organic radical and its out-of-cage coupling to produce a C–N bond, as depicted in the mechanism outlined in Fig. 4a.
Our studies of couplings of enantiomerically enriched electrophiles have also been informative. Under our standard reaction conditions, the two enantiomers of electrophile 69 react at comparable rates in the presence of (R,R)–N1*, and the e.e. of the product is independent of the original stereochemistry of the electrophile (Fig. 4f, Entries 1 and 2). Slow racemization of the electrophile occurs during the photoinduced C–N coupling (Entries 1–3), but racemization ceases when irradiation ceases (Entry 3 vs Entry 4), suggesting that racemization is proceeding through a photoinduced pathway, e.g., reversible copper-mediated C–Br bond cleavage. Similar racemization occurs in the absence of N1*, consistent with CuII complex D, not E or F, serving as the Br donor (Entry 5).
In conclusion, by exploiting photoinduced copper catalysis, the first enantioconvergent substitution reactions of unactivated alkyl electrophiles by nitrogen nucleophiles (specifically, primary amides) have been accomplished, thereby enabling a new, convergent strategy for the asymmetric synthesis of secondary amides wherein C–N bond formation and control of stereochemistry are achieved in a single step. The method exploits three classes of ligands, which assemble the necessary catalysts in situ: a bidentate phosphine and a phenoxide bind to CuI to produce a photocatalyst with a sufficient excited-state lifetime and reduction potential to activate the comparatively strong C–Br bond of the unactivated electrophile, and a bidentate chiral diamine binds to CuII to effect enantioselective C–N bond formation. Efforts to exploit copper, an earth-abundant metal, to address other unsolved challenges in asymmetric catalysis are underway.
Extended Data
Extended Data Fig. 1 |. Continued scope of alkyl bromides serving as electrophilic coupling partners.
Couplings were generally conducted using 0.5 mmol of the amide, and N1* was used as the diamine, unless otherwise noted. All data represent the average of two experiments. The percent yield represents purified product. a: electrophile (1.5 equiv.), Cu(CH3CN)4PF6 (15 mol%), P (5 mol%), N1* (20 mol%), K3PO4•H2O (1.5 equiv.; in place of Cs2CO3), 10 °C. Ar1 = p-(F3C)C6H4, DG = directing group, Bn = benzyl.
Supplementary Material
Acknowledgements
This manuscript is dedicated to the memory of Gregory P. Harlow. Support has been provided by the National Institutes of Health (National Institute of General Medical Sciences, R01-GM109194), the Beckman Institute (support of the Laser Resource Center, as well as the Center for Catalysis and Chemical Synthesis, the EPR facility, and the X-ray crystallography facility), the Gordon and Betty Moore Foundation (support for the Center for Catalysis and Chemical Synthesis), the Dow Next-Generation Educator Fund (grant to Caltech), and Boehringer–Ingelheim Pharmaceuticals. We thank C. Citek, T.M. Donnell, J. Dørfler, P. Garrido Barros, L.M. Henling, P.H. Oyala, F. Schneck, M. Shahgoli, D. VanderVelde, and J.R. Winkler, for assistance and discussions.
Footnotes
Competing interests The authors declare no competing interests.
Data availability
The data that support the findings of this study are available within the paper, its Supplementary Information (experimental procedures and characterization data) and from the Cambridge Crystallographic Data Centre (https://www.ccdc.cam.ac.uk/structures; crystallographic data are available free of charge under CCDC reference numbers CCDC 2055329-2055338.
References
- 1.Fu GC Transition-metal catalysis of nucleophilic substitution reactions: a radical alternative to SN1 and SN2 processes. ACS Cent. Sci. 3, 692–700 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Choi J & Fu GC Transition metal-catalyzed alkyl–alkyl bond formation: another dimension in cross-coupling chemistry. Science 356, eaaf7230 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Kaga A & Chiba S Engaging radicals in transition metal-catalyzed cross-coupling with alkyl electrophiles: recent advances. ACS Catal. 7, 4697–4706 (2017). [Google Scholar]
- 4.Iwasaki T & Kambe N Ni-catalyzed C–C couplings using alkyl electrophiles. Top. Curr. Chem 374, 66 (2016). [DOI] [PubMed] [Google Scholar]
- 5.For early examples, see: Bissember AC, Lundgren RJ, Creutz SE, Peters JC & Fu GC Transition-metal-catalyzed alkylations of amines with alkyl halides: photoinduced, copper-catalyzed couplings of carbazoles. Angew. Chem. Int. Ed 52, 5129–5133 (2013).
- 6.Do H-Q, Bachman S, Bissember AC, Peters JC & Fu GC Photoinduced, copper-catalyzed alkylation of amides with unactivated secondary alkyl halides at room temperature. J. Am. Chem. Soc. 136, 2162–2167 (2014). [DOI] [PubMed] [Google Scholar]
- 7.Peacock DM, Roos CB & Hartwig JF Palladium-catalyzed cross coupling of secondary and tertiary alkyl bromides with a nitrogen nucleophile. ACS Cent. Sci 2, 647–652 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Grangea RL, Clizbe EA & Evans PA Recent developments in asymmetric allylic amination reactions. Synthesis 48, 2911–2968 (2016). [Google Scholar]
- 9.Zhang H et al. Construction of the N1-C3 linkage stereogenic centers by catalytic asymmetric amination reaction of 3-bromooxindoles with indolines. Org. Lett 16, 2394–2397 (2014). [DOI] [PubMed] [Google Scholar]
- 10.Kainz QM et al. Asymmetric copper-catalyzed C–N cross-couplings induced by visible light. Science 351, 681–684 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Zhang X et al. An enantioconvergent halogenophilic nucleophilic substitution (SN2X) reaction. Science 363, 400–404 (2019). [DOI] [PubMed] [Google Scholar]
- 12.Bartoszewicz A, Matier CD & Fu GC Enantioconvergent alkylations of amines by alkyl electrophiles: copper-catalyzed nucleophilic substitutions of racemic α-halolactams by indoles. J. Am. Chem. Soc. 141, 14864–14869 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Wang Y, Wang S, Shan W & Shao Z Direct asymmetric N-propargylation of indoles and carbazoles catalyzed by lithium SPINOL phosphate. Nat. Commun 11, 226 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Greenberg A, Breneman CM & Liebman JF (eds) The Amide Linkage: Structural Significance in Chemistry, Biochemistry, and Materials Science (Wiley: New York, 2000). [Google Scholar]
- 15.Ponra S, Boudet B, Phansavath P & Ratovelomanana-Vidal V Recent developments in transition-metal-catalyzed asymmetric hydrogenation of enamides. Synthesis 53, 193–214 (2021). [Google Scholar]
- 16.Wang J-W et al. Catalytic asymmetric reductive alkylation of enamines to chiral aliphatic amines. ChemRxiv (2020), 10.26434/chemrxiv.13102307.v1. [DOI] [Google Scholar]
- 17.Brown DG & Böstrom J Analysis of past and present synthetic methodologies on medicinal chemistry: where have all the new reactions gone? J. Med. Chem. 59, 4443–4458 (2016). [DOI] [PubMed] [Google Scholar]
- 18.Taylor JE & Bull SD N-Acylation reactions of amines. Comprehensive Organic Synthesis 2nd edn, Vol. 6 (eds Knochel P & Molander GA) 427–478 (Elsevier, Amsterdam, 2014). [Google Scholar]
- 19.Carey JS, Laffan D, Thomson C & Williams MT Analysis of the reactions used for the preparation of drug candidate molecules. Org. Biomol. Chem 2006, 4, 2337–2347.Table 11 [DOI] [PubMed] [Google Scholar]
- 20.Matier CD, Schwaben J, Peters JC & Fu GC Copper-catalyzed alkylation of aliphatic amines induced by visible light. J. Am. Chem. Soc 139, 17707–17710 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Ahn JM, Peters JC & Fu GC Design of a photoredox catalyst that enables the direct synthesis of carbamate-protected primary amines via photoinduced, copper-catalyzed N-alkylation reactions of unactivated secondary halides. J. Am. Chem. Soc 139, 18101–18106 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Hossain A, Bhattacharyya A & Reiser O Copper’s rapid ascent in visible-light photoredox catalysis. Science 364, 450–461 (2019). [DOI] [PubMed] [Google Scholar]
- 23.Cheng L-J & Mankad NP C–C and C–X coupling reactions of unactivated alkyl electrophiles using copper catalysis. Chem. Soc. Rev 49, 8036–8064 (2020). [DOI] [PubMed] [Google Scholar]
- 24.Carreira EM & Yamamoto H (eds) Comprehensive Chirality (Academic, Amsterdam, 2012). [Google Scholar]
- 25.Reyes RL et al. Asymmetric remote C–H borylation of aliphatic amides and esters with a modular iridium catalyst. Science 369, 970–974 (2020). [DOI] [PubMed] [Google Scholar]
- 26.Ordonez M, Labastida-Galvan V & Lagunas-Rivera S Stereoselective synthesis of GABOB, carnitine and statine phosphonates analogues. Tetrahedron: Asymmetry 21, 129–147 (2010). [Google Scholar]
- 27.Jeffrey J et al. The discovery of 3-(N-alkyl)aminopropylphosphonic acids as potent S1P receptor agonists. Bioorg. Med. Chem. Lett 14, 3495–3499 (2004). [DOI] [PubMed] [Google Scholar]
- 28.Ordonez M, Cativiela C & Romero-Estudillo I An update on the stereoselective synthesis of γ-amino acids. Tetrahedron: Asymmetry 27, 999–1055 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Souto JA & Muniz K Asymmetric diamination of alkenes. Asymmetric Synthesis, Vol. 2 (eds Christmann M & Brase S) 371–375 (Wiley–VCH, Weinheim, Germany, 2012). [Google Scholar]
- 30.Antonopoulou G et al. Synthesis of 2-oxoamides based on sulfonamide analogs of γ -amino acids and their activity on phospholipase A2. J. Pept. Sci. 14, 1111–1120 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Comprehensive Asymmetric Catalysis, Vol. I–III (eds Jacobsen EN, Pfaltz A & Yamamoto H) (Springer, Berlin, Germany, 1999). [Google Scholar]
- 32.Teegardin K, Day JI, Chan J & Weaver J Advances in photocatalysis: a microreview of visible light mediated ruthenium and iridium catalyzed organic transformations. Org. Process Res. Dev. 20, 1156–1163 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Creutz SE, Lotito KJ, Fu GC & Peters JC Photoinduced Ullmann C–N coupling: demonstrating the viability of a radical pathway. Science 338, 647–651 (2012). [DOI] [PubMed] [Google Scholar]
- 34.Ribelli TG, Lorandi F, Fantin M, Matyjaszewski K Atom Transfer Radical Polymerization: Billion Times More Active Catalysts and New Initiation Systems. Macromol. Rapid Comm. 40, 1–44 (2019). [DOI] [PubMed] [Google Scholar]
- 35.Lusztyuk J, Maillard B, Deycard S, Lindsay DA & Ingold KU Kinetics for the reaction of a secondary alkyl radical with tri-n-butylgermanium hydride and calibration of a secondary alkyl radical clock reaction. J. Org. Chem 52, 3509–3514 (1987). [Google Scholar]
- 36.Anslyn EV & Dougherty DA in Modern Physical Organic Chemistry, 156 (University Science Books, Sausalito, CA, 2006). [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 that support the findings of this study are available within the paper, its Supplementary Information (experimental procedures and characterization data) and from the Cambridge Crystallographic Data Centre (https://www.ccdc.cam.ac.uk/structures; crystallographic data are available free of charge under CCDC reference numbers CCDC 2055329-2055338.