Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Aug 2.
Published in final edited form as: Org Lett. 2019 Jul 17;21(15):6089–6095. doi: 10.1021/acs.orglett.9b02234

Sulfamyl Radicals Direct Photoredox-Mediated Giese Reactions at Unactivated C(3)–H Bonds

Anastasia L G Kanegusuku 1, Thomas Castanheiro 1, Suraj K Ayer 1, Jennifer L Roizen 1,*
PMCID: PMC7359209  NIHMSID: NIHMS1586955  PMID: 31313933

Abstract

Alcohol-anchored sulfamate esters guide the alkylation of tertiary and secondary aliphatic C(3)–H bonds. The transformation proceeds directly from N–H bonds with a catalytic oxidant, a contrast to prior methods which have required pre-oxidation of the reactive nitrogen center, or employed stoichiometric amounts of strong oxidants to obtain the sulfamyl radical. These sulfamyl radicals template otherwise rare 1,6-hydrogen-atom transfer (HAT) processes via seven-membered ring transition states to enable C(3)–H functionalization during Giese reactions.

Keywords: sulfamate ester, directed, C–H functionalization, masked alcohol, Giese reaction

Graphical Abstract

graphic file with name nihms-1586955-f0006.jpg

INTRODUCTION

Nitrogen-centered radicals are an important and versatile class of chemical intermediates.1 Yet, nitrogen-centered radicals remain underutilized, as most methods for their generation rely on harsh conditions to oxidize the nitrogen center. Recently, photocatalytic strategies have been developed as mild processes to form nitrogen-centered radicals,26 however, only amides, carbamates, and sulfonamides have served as precursors to neutral nitrogen-centered radicals.46 These groups template position-selective C–H functionalization technologies, transforming C(4)–H bonds through 1,5-HAT processes (Scheme 1A).7 In-contrast, sulfamate esters guide functionalization to C(3)–H centers through otherwise rare 1,6-HAT processes (Scheme 1B), providing complementary positional selectivity to established processes.816 As a complement to known guided methods, sulfamate esters are attractive directing groups because they derive from alcohols, which are ubiquitous in biologically active small molecules. To date, protocols templated by sulfamate ester substrates require pre-oxidation of the reactive nitrogen center, or the use of strong stoichiometric oxidants to access nitrogen-centered sulfamyl radicals.14, 15 As such, a strategy to facilitate sulfamyl radical formation directly from N–H bonds under mild conditions would enhance the substrate tolerance of these directing motifs, and could also enable previously unrealized synthetic disconnections.17, 18 Herein disclosed is the first catalytic process to access sulfamyl free radicals directly from N–H bonds.19 These sulfamyl radicals have been engaged in Giese reactions, in the only examples of C(3)–H alkylation reactions guided by alcohol surrogates (Scheme 1C, 12).

Scheme 1.

Scheme 1.

Sulfamate esters offer complementary site selectivity to developed templates for C–H alkylation reactions

RESULTS AND DISCUSSION

At the outset of these investigations, we sought conditions that would facilitate oxidation of sulfamate ester 1a to sulfamyl radical 3a. We envisioned that this could occur through an initial deprotonation to provide sulfamate ester anion 4a. Anion 4a could undergo single electron oxidation to generate sulfamyl radical 3a, with concurrent reduction of the excited iridium catalyst 6a (Scheme 2). Consistent with this proposal, in acetonitrile, sodiated 5-methylhexyl N-tert-butyl sulfamate ester anion 4a has a half-peak potential (Ep/2) of +0.753 V versus saturated calomel electrode (SCE), indicating that oxidation of the anion by a strongly oxidizing and broadly used iridium(III) photocatalyst [Ir(dF(CF3)ppy)2(dtbbpy)]PF6 (6a, E1/2[IrIII*/IrII]= +1.21 V versus SCE, dF(CF3)ppy = 2-(2,4-difluorophenyl)-3-trifluoromethylpyridine, dtbbpy = 4,4’-di-tert-butyl-2,2’- bipyridine) is thermodynamically feasible.20 Stern-Volmer analysis of the related sulfamate ester anion 4b quenches 6a. Based on this data, reductive quenching of [IrIII*] 6a to [IrII] 7a by an in situ generated sulfamate ester anion, such as 4a, can furnish sulfamyl radical 3a. Alternatively, oxidation could also occur through a concerted proton-coupled electron transfer (PCET) event. Presently, we cannot exclude the possibility that electron transfer could proceed through a concerted PCET event.

Scheme 2.

Scheme 2.

Proposed mechanism for sulfamate ester directed Giese reaction

By analogy to previously reported sulfamyl radical-mediated C–H functionalization reactions,14, 15 3a is poised to guide an intramolecular 1,6-HAT process to furnish carbon-centered radical 8a. Nucleophilic radical 8a engages electron-deficient olefin 9a in a Giese reaction. At this point, reduction of an ester stabilized radical (E0[R•/R–] ≈ –0.63 V versus SCE),21 such as 10a, by strong reductant 7a (E1/2[IrIII/IrII]= –1.37 V versus SCE)20 to furnish carbon anion 11a should be thermodynamically favorable. Finally, anion protonation would provide C(3)-alkylated product 2a.

To develop this guided Giese reaction, we employed tert-butyl acrylate (9a) as a radical trapping agent with 5-methylhexyl N-tert-butylsulfamate ester (1a, Table 1). With this system, we could probe the ability of this process to overcome the innate position-selectivity that dictates the site of unguided C–H functionalization reactions. In undirected processes, functionalization would be expected to proceed preferentially at the weakest and the most electron-rich C–H bond in the molecule, which would be the distal tertiary C–H center (bond dissociation energy (BDE) ≈ 96 kcal/mol).22 The targeted C(3)–H bond is expected to be stronger (BDE ≈ 98 kcal/mol), and thus less likely to engage in an unguided functionalization. Nevertheless, the directed Giese reaction proceeds in the presence of light, photocatalyst 5a, and K2CO3 to afford desired 2a in 77% yield (entry 1). Using a range of strongly oxidizing excited state photocatalysts 5a5e, those that become stronger reductants prove more efficient in this transformation (entries 1, 2, 5). Consistent with mechanistic hypotheses, control experiments confirm that the photocatalyst, light, base, and inert atmosphere are required for product formation (See supplementary information).

Table 1.

Photocatalysts that are stronger reductants upon initial quenching are more efficient

graphic file with name nihms-1586955-t0007.jpg

entrya photo-catalyst E1/2(PC*/PC) [V]b E1/2(PC/PC) [V]b 1a [%]c yield 2a [%]d
1 5a +1.21 −1.37 17 77
2 5b +0.97 −1.23 46 44
3 5c +1.68 −0.69 98 nde
4 5d +1.65 −0.79 96 nde
5 5e +1.37 −1.21 12 34

graphic file with name nihms-1586955-t0008.jpg
a

General reaction conditions: 1.0 equiv sulfamate ester, 2.2 equiv tert-butyl acrylate, two 34 W blue lamps, photocatalyst (1.0 mol %), 1.0 equiv K2CO3, MeCN (0.2 M), 900 mot/min stirring, 21–28 °C, 48 h.

b

V versus SCE. 5, 20, 23

c

Recovered 1.

d

Isolated yields.

e

Not detected in 1H NMR of the crude reaction mixture.

Under the optimized reaction conditions, the methylene center in sulfamate ester 1a engages selectively in a single alkylation event to furnish Giese product 2a, which incorporates a tertiary C(5)–H bond. This tertiary C–H bond is weaker than the secondary C(3)–H bond in substrate 1a, and might be expected to be more susceptible to further alkylation to furnish fully-substituted 12. To explain this selectivity, we hypothesize that the newly generated tertiary C(3)–H bond may prove too sterically hindered to engage in further functionalization. Consistent with this hypothesis, we found that less sterically encumbered 1b–1d undergo initial monoalkylation with tert-butyl acrylate, followed by a second Giese reaction to generate fully-substituted 12b–d (Scheme 3). Furthermore, compounds with distal branching patterns (i.e. 1e–h) react to furnish exclusive monoalkylated products 2e–h.

Scheme 3. Secondary centers engage in productive Giese reactionsa,b.

Scheme 3.

a General reaction conditions: 1.0 equiv sulfamate ester, 2.2 equiv electrophilic olefin, two 34 W blue lamps, photocatalyst (1.0 mol %), 1.0 equiv K2CO3, MeCN (0.2 M), 900 mot/min stirring, 21–28 °C, 48h. b Isolated yields. c 11 days, additional 2.2 equiv acrylate added after 142 h. d 144 h, 4.4 equiv ethyl acrylate, and additional 4.4 equiv methyl acrylate added after 48 and 96 h.

In more complex substrates, C(3)-methylene centers react with substrate-induced site- and diastereoselectivity (Scheme 3). To our delight, the reaction of isosteviol derivative 1i gives product 2i as a 2:1 mixture of diastereomers in 58% isolated yield. Furthermore, while many electron-rich aryl groups are oxidatively labile, these conditions are sufficiently mild to affect functionalization of dehydroabietyl-derived 1j. Indeed, 1j provides alkylated products 2j as a 1.6:1 mixture of diastereomers in a remarkable 81% isolated yield. These are the first documented cases of sulfamate ester-guided process where the radical-containing intermediate unambiguously induces a diastereoselective reaction.24

In spite of sensitivity to steric hindrance, sulfamate esters guide Giese reactions at tertiary centers with predictable selectivity (Scheme 4). While the electron densities of the C(5)–H and C(8)–H bonds of menthol-derived 13a are predicted to be similar,25 this Giese reaction functionalizes solely the C(8)–H bond. Analogous site-selectivity is displayed in sulfamate ester-guided chlorination14 and bromination15 reactions, as well as in iron- and manganese-mediated intramolecular amination reactions.10d, 10g Presumably, the C(8)–H bond is geometrically disposed to interact with the directing groups that guide these reactions. By comparison, unguided intermolecular oxidation25 and amination26 processes engage the more sterically accessible C(5)–H bond selectively.

Scheme 4. Reactions can form fully-substituted centers and tolerate N-substituent variationsa,b.

Scheme 4.

a General reaction conditions: 1.0 equiv sulfamate ester (0.2 mmol), 2.2 equiv electrophilic olefin, two 34 W blue lamps, photocatalyst (1.0 mol %), 1.0 equiv K2CO3, MeCN (0.2 M), 900 mot/min stirring, 21–28 °C, 48h. b Isolated yields. c 16 h. d 24 h. e Reaction performed with 13f (1 mmol). f Yield as average of two trials. g Product not detected by 1H NMR of the crude reaction mixture in a reaction with tert-butyl acrylate.

Furthermore, this directed reaction surmounts electronically-induced preferences for functionalization. In undirected processes, inductively electron-withdrawing groups can deactivate proximal C–H bonds to oxidation. When 3,7-dimethyloctanol is masked with an electron withdrawing group, it undergoes preferential oxidation at C(7) in fluorination,27 oxygenation,28 amination,29 azidation,30 and trifluoromethylthiolation31 reactions. By contrast, this sulfamate ester guided reaction installs new carbon–carbon bonds predictably and selectively at C(3).

Additional fully-substituted C(3) centers can be generated from unactivated acyclic or cyclic tertiary C(sp3)–H bonds, or ethereal centers (i.e. 13cg). Remarkably, while ethereal 13e features more sterically accessible and electronically activated C–H bonds, the reaction occurs selectively at the C(3)–H site.

Notably, optically active 13h is converted to racemic 14h, such that the generated radical must have a sufficient lifetime to epimerize. Nevertheless, remaining substrate is recovered without detectable epimerization, which suggests that the C–H abstraction is not reversible on a time scale that would allow for epimerization.

While tertiary and secondary C–H bonds engage in productive Giese reactions, primary centers, and secondary benzylic centers do not. Additionally, while racemic substrate 13g reacts productively in this transformation, analogue 13k is ineffective, providing further evidence that the reaction is sensitive to steric encumbrance. To date, the efficient transformation is highly dependent on nitrogen substitution. Unfortunately, N-iso-propyl, N-(1,1,1-trifluoro-2-methylpropyl), and N-2-(2-phenylpropyl) sulfamate esters (13ln) guide the Giese reaction less effectively than N-tert-butyl sulfamate esters.

A wide range of electron-deficient olefins engage in this directed Giese reaction (Table 2). Alkyl, aryl, allyl, and alkynyl acrylates react in good to excellent yields (entries 1–4). This is remarkable, as many aryl, allyl, and alkynyl groups would undergo deleterious processes under more oxidizing conditions. Furthermore, as alkynes are well-developed as handles to facilitate structure-activity-relationship or mode-of-action studies, incorporation of an alkyne-bearing trapping agent suggests that this Giese reaction could be applied to complex small molecules to generate chemical probes.32

Table 2.

Reaction traps various electron deficient olefinsa

graphic file with name nihms-1586955-t0009.jpg

entry olefin product R2 yield [%]b yield [%]c
1 graphic file with name nihms-1586955-t0010.jpg graphic file with name nihms-1586955-t0011.jpg 15a Me 22 74
2 15b Ph 07 72
3 15c graphic file with name nihms-1586955-t0012.jpg 38 58
4 15d graphic file with name nihms-1586955-t0013.jpg 09 66
R2 R3
5 graphic file with name nihms-1586955-t0014.jpg graphic file with name nihms-1586955-t0015.jpg 16a Et Me ndd 58
6 16b Me Ph e 80
7f 16c Me NBoc2 ndd 89
8f graphic file with name nihms-1586955-t0016.jpg graphic file with name nihms-1586955-t0017.jpg 16d 21 78
9 graphic file with name nihms-1586955-t0018.jpg graphic file with name nihms-1586955-t0019.jpg 16e 11 87
R2
10 graphic file with name nihms-1586955-t0020.jpg graphic file with name nihms-1586955-t0021.jpg 15e Ph 47 50
11 15f Me 24 72
12 15g H 58 43
13 15h NMe2 04 66
14 15i NHMe 04 55
15 graphic file with name nihms-1586955-t0022.jpg graphic file with name nihms-1586955-t0023.jpg 15j 52 44
16 graphic file with name nihms-1586955-t0024.jpg graphic file with name nihms-1586955-t0025.jpg 16f ndd ≥98
>20:1
dr
a

General reaction conditions: 1.0 equiv sulfamate ester, 2.2 equiv olefin, two 34 W blue lamps, photocatalyst (1.0 mol %), 1.0 equiv K2CO3, M eCN (0.2 M), 900 mot/min stirring, 21–28 °C.

b

Recovered 13b or 13f

c

Isolated yields.

d

Not detected in crude 1H NMR of the crude reaction mixture.

e

Not isolated.

f

1.2 equiv olefin.

Despite the documented challenges with substrate steric encumbrance, bulky electrophilic olefin engage effectively, including 2-substituted acrylates, a dehydroalanine derivative, and dimethyl fumarate, which contains an internal olefin (entries 5–8). Moreover electrophilic olefins also react effectively, including α,β-unsaturated ketones that are cyclic, acyclic, or aryl-substituted, and acrolein (entries 9–12). Finally, nitrogen-containing olefins are competent in the reaction (entries 13–15).

Radical conjugate addition reactions benefit from a rich library of enantioenriched electrophilic olefins, designed to induce asymmetry in the course of radical addition reactions, even with acyclic nucleophilic free radicals.33 In principle, this rich array of enantioenriched electrophilic olefins could be used to intercept this radical translocation protocol and induce asymmetry in the course of these Giese reactions. Gratifyingly, sulfamate ester 13f reacts with a known enantioenriched methyleneoxazolidinone34, 35 to furnish amino acid derivative 16f in quantitative yield with complete diastereocontrol (>20:1 d.r., entry 16). This approach could provide access to a broad array of enantioenriched non-natural alkylated amino acids.

Ultimately, to maximize the utility of this process, it is necessary to cleave or displace the sulfamate ester template for this alkylation process. To make sulfamate esters more liable to displacement, typically they are N-acylated.10a To our disappointment, the N-tert-butyl sulfamate ester products are not efficiently carbamoylated under standard conditions. Fortunately, N-vinylation of N-tert-butyl sulfamate esters can be achieved with diethyl acetylene dicarboxylate (Scheme 5). This renders vinylated 17 susceptible to displacement to furnish azide,15 iodide,15 acetate,10a thioacetate,10a or free alcohol10a analogues.

Scheme 5.

Scheme 5.

Strategies for sulfamate ester displacement

CONCLUSION

This sulfamate ester-guided photoredox-mediated reaction provides a powerful and general platform for directed C–H functionalization. This is the first research to establish that sulfamate ester anions can be photochemically oxidized. Moreover, the resultant nitrogen-centered radicals guide 1,6-HAT to furnish tertiary or secondary carbon-centered radicals, which can be trapped to generate racemic and diastereoselective products. Thereby, this process affords complementary position-selectivity to that achieved using known photoredox-mediated carbon–carbon bond-forming reactions. Finally, a new tactic for sulfamate ester vinylation and displacement introduces further diversity to the array of accessed small molecules.

Supplementary Material

Supporting Information

ACKNOWLEDGMENT

Funding was provided by the National Institutes of Health (R35GM128741–01). Characterization data were obtained on instrumentation secured with funding from the NSF (CHE-0923097, ESI-MS, George Dubay, the Duke Dept. of Chemistry Instrument Center), or the NSF, the NIH, HHMI, the North Carolina Biotechnology Center and Duke University (Duke Magnetic Resonance Spectroscopy Center). A Bruker-Nonius X8 Kappa APEXII (CCD) instrument using Mo Kα radiation with an Oxford Cryostream 700 cold stream was used for crystallographic analyses, and measurements were made at the Molecular Education, Technology, and Research Innovation Center (METRIC) at NC State University (Roger Sommer). We thank Dr. Peter Silinski for performing high-resolution mass spectrometry, Dr. Ben Bobay for NMR support, and Dr. Todd Woerner for cyclic voltammetry support (Duke University). We gratefully acknowledge Prof. Nathan Jui (Emory University) for generously donating a sample of dehydroalanine-derived electrophilic olefin.

Footnotes

Notes

The authors declare no competing financial interests.

The Supporting Information is available free of charge on the ACS publications website at DOI:

All experimental procedures and characterization for new compounds (PDF)

Crystallographic data for dehydroabietyl-derived alkylated sulfamate ester 2j (CIF)

REFERENCES

  • (1).Ruppel JV; Kamble RM; Zhang XP Cobalt-Catalyzed Intramolecular C−H Amination with Arylsulfonyl Azides. Org. Lett 2007, 9, 4889–4892. [DOI] [PubMed] [Google Scholar]; (b) Lu H; Lang K; Jiang H; Wojtas L; Zhang XP Intramolecular 1,5-C(sp3)–H radical amination via Co(II)-based metalloradical catalysis for five-membered cyclic sulfamides. Chem. Sci 2016, 7, 6934–6939. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Xiong T; Zhang Q New amination strategies based on nitrogen-centered radical chemistry. Chem. Soc. Rev 2016, 45, 3069–3087. [DOI] [PubMed] [Google Scholar]
  • (2).Kärkäs MD Photochemical Generation of Nitrogen-Centered Amidyl, Hydrazonyl, and Imidyl Radicals: Methodology Developments and Catalytic Applications. ACS Catal. 2017, 7, 4999–5022. [Google Scholar]
  • (3).Ischay MA; Anzovino ME; Du J; Yoon TP Efficient Visible Light Photocatalysis of [2+2] Enone Cycloadditions. J. Am. Chem. Soc 2008, 130, 12886–12887.Nicewicz DA; MacMillan DWC Merging Photoredox Catalysis with Organocatalysis: The Direct Asymmetric Alkylation of Aldehydes. Science 2008, 322, 77–80.Narayanam JM; Tucker JW; Stephenson CR Electron-Transfer Photoredox Catalysis: Development of a Tin-Free Reductive Dehalogenation Reaction. J. Am. Chem. Soc 2009, 131, 8756–8757.
  • (4).Shu W; Nevado C Visible-Light-Mediated Remote Aliphatic C-H Functionalizations through a 1,5-Hydrogen Transfer Cascade. Angew. Chem. Int. Ed 2017, 56, 1881–1884.Dauncey EM; Morcillo SP; Douglas JJ; Sheikh NS; Leonori D Photoinduced Remote Functionalizations via Iminyl Radical-Promoted C–C and C–H Bond Cleavage Cascades. Angew. Chem. Int. Ed 2018, 57, 744–748.Jiang H; Studer A α-Aminoxy-Acid-Auxiliary-Enabled Intermolecular Radical γ-C(sp3)–H Functionalization of Ketones. Angew. Chem. Int. Ed 2018, 57, 1708–1712.
  • (5).Choi GJ; Zhu Q; Miller DC; Gu CJ; Knowles RR Catalytic alkylation of remote C–H bonds enabled by proton-coupled electron transfer. Nature 2016, 539, 268–271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (6).(a) Chu JCK; Rovis T Amide-directed photoredox-catalysed C–C bond formation at unactivated sp3 C–H bonds. Nature 2016, 539, 272–275. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Chen D; Chu JCK; Rovis T Directed γ-C(sp3)–H Alkylation of Carboxylic Acid Derivatives through Visible Light Photoredox Catalysis. J. Am. Chem. Soc 2017, 139, 14897–14900. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (7).Zhu Y; Huang K; Pan J; Qiu X Luo X; Qin Q; Wei J; Wen X; Zhang L; Jiao N Silver-catalyzed remote Csp3–H functionalization of aliphatic alcohols. Nat. Commun 2018, 9, 2625.Wi X; Zhang H; Tang N; Wu Z; Wang D; Ji M; Xu Y; Wang M; Zhu C Metal-free alcohol-directed regioselective heterarylation of remote unactivated C(sp3)–H bonds. Nat. Commun 2018, 9, 3343.Hu A; Guo J-J; Pan H, Tang H; Gao Z; Zuo Z δ-Selective Functionalization of Alkanol, Enables by Visible-Light-Induced Ligand-to-Metal Charge Transfer. J. Am. Chem. Soc 2018, 140, 1612–1616.Li G-X; Hu X; He G; Chen G Photoredox-mediated remote C(sp3)–H heteroarylation of free alcohols. Chem. Sci 2019, 10, 688–693.
  • (8).Mo F; Tabor JR; Dong G Alcohols or Masked Alcohols as Directing Groups for C–H Bond Functionalization. Chem. Lett 2014, 43, 264–271. [Google Scholar]
  • (9).Simmons EM; Hartwig JF Catalytic functionalization of unactivated primary C–H bonds directed by an alcohol. Nature 2012, 483, 70–73. For carbamate-guided oxidation that is selective for oxidation of benzylic or tertiary beta-C(sp3)–H or gamma-C(sp3)–H centers, see:Chen K; Richter JM; Baran PS 1,3-Diol Synthesis via Controlled, Radical-Mediated C−H Functionalization. J. Am. Chem. Soc 2008, 130, 7247–7249. For a silyl ether-guided radical relay Heck reaction, seeChuentragool P; Yadagiri D; Morita T; Sarkar S; Parasram M; Wang Y; Gevorgyan V Aliphatic Radical Relay Heck Reaction at Unactived C(sp3)–H Sites of Alcohols. Angew. Chem. Int. Ed 2019, 58, 1794–1798.Xia G; Weng J; Liu L; Verma P; Li Z; Yu J-Q Reversing conventional site-selectivity in C(sp3)–H bond activation. Nature, 2019, 11, 571–577.
  • (10).Espino CG; Wehn PM; Chow J; Du Bois J Synthesis of 1,3-Difunctionalized Amine Derivatives through Selective C−H Bond Oxidation. J. Am. Chem. Soc 2001, 123, 6935–6936.Duran F; Leman FL; Ghini A; Burton G; Dauban P; Dodd RH Intramolecular PhI=O Mediated Copper-Catalyzed Aziridination of Unsaturated Sulfamates: A New Direct Access to Polysubstituted Amines from Simple Homoallylic Alcohols. Org. Lett 2002, 4, 2481–2483.Milczek E; Boudet N; Blakey S Enantioselective C-H amination using cationic ruthenium(II)-pybox catalysts. Angew. Chem. Int. Ed 2008, 47, 6825–6828.Liu Y; Guan X; Lai-Ming Wong E; Liu P; Huang J-S; Che C-M Nonheme Iron-Mediated Amination of C(sp3)–H Bonds. Quinquepyridine-Supported Iron-Imide/Nitrene Intermediates by Experimental Studies and DFT Calculations. J. Am. Chem. Soc 2013, 135, 7194–7204.Alderson JM; Phelps AM; Scamp RJ; Dolan NS; Schomaker JM Ligand-Controlled, Tunable Silver-Catalyzed C–H Amination. J. Am. Chem. Soc 2014, 136, 16720–16723.Subbarayan V; Jin L-M; Cui X; Zhang XP Room Temperature Activation of Aryloxysulfonyl Azides by [Co(II)(TPP)] for Selective Radical Aziridination of Alkenes via Metalloradical Catalysis. Tetrahedron Lett. 2015, 56, 3431–3434.Paradine SM; Griffin JR; Zhao J; Petronico AL; Miller SM; White MC A Manganese Catalyst for Highly Reactive Yet Chemoselective Intramolecular C(sp3)–H Amination. Nature Chem. 2015, 7, 987–994.
  • (11).Blackburn JM; Short MA; Castanheiro T; Ayer SK; Muellers TD; Roizen JL Synthesis of N-Substituted Sulfamate Esters from Sulfamic Acid Salts by Activation with Triphenylphosphine Ditriflate Org. Lett 2017, 19, 6012–6015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (12).Nechab M; Mondal S; Bertrand MP 1,n‐Hydrogen‐Atom Transfer (HAT) Reactions in Which n≠5: An Updated Inventory. Chem. Eur. J 2014, 20, 16034–16059. [DOI] [PubMed] [Google Scholar]
  • (13).Zalatan DN; Du Bois J Oxidative Cyclization of Sulfamate Esters Using NaOCl - A Metal-Mediated Hoffman-Löffler-Freytag Reaction. Synlett 2009, 143–146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (14).Short MA; Blackburn JM; Roizen JL Sulfamate Esters Guide Selective Radical-Mediated Chlorination of Aliphatic C-H Bonds. Angew. Chem. Int. Ed 2018, 57, 296–299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (15).Sathyamoorthi S; Banerjee S; Du Bois J; Burns N; Zare RN Site-selective bromination of sp3 C–H bonds. Chem. Sci 2018, 9, 100–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).Liu T; Mei T-S; Yu J-Q gamma, delta, epsilon-C(sp3)–H Functionalization through Directed Radical C–H-Abstraction. J. Am. Chem. Soc 2015, 137, 5871–5874.Wappes EA; Nakafuka KM; Nagib DA Directed beta-C–H Amination of Alcohols via Radical Relay Chaperones. J. Am. Chem. Soc 2017, 139, 10204–10207.Wappes EA; Vanitcha A; Nagib DA beta-C–H di-halogenation via iterative hydrogen atom transfer. Chem. Sci 2018, 9, 4500–4504.Del Castillo E; Martínez MD; Bosnidou AE; Duhamel T; O’Broin CQ; Zhang H; Escudero-Adán EC; Martínez-Belmonte M; Muñiz K Multiple Halogenation of Aliphatic C–H Bonds within the Hofmann-Löffler Manifold. Chem. Eur. J 2018, 24, 17225–17229.
  • (17).For select reviews, see:Bergman RG C–H activation. Nature 2007, 446, 391–393.Davies HML; Manning JR Catalytic C–H functionalization by metal carbenoid and nitrenoid insertion. Nature 2008, 451, 417–424.Lyons TW; Sanford MS Palladium-Catalyzed Ligand-Directed C–H Functionalization Reactions. Chem. Rev 2010, 110, 1147–1169.White MC Adding Aliphatic C−H Bond Oxidations to Synthesis. Science 2012, 335, 807–809.He J; Wasa M; Chan KSL; Shao Q; Yu J-Q Palladium-Catalyzed Transformations of Alkyl C-H Bonds. Chem. Rev 2017, 117, 8754–8786.
  • (18).For select resources describing site-selectivity in C–H functionalization, see:Chen K; Eschenmoser A; Baran PS Strain-Release in C–H Bond Activation? Angew. Chem. Int. Ed 2009, 48, 9705–9708.Brückl T; Baxter RD; Ishihara Y; Baran PS Innate and Guided C–H Functionalization Logic. Acc. Chem. Res 2012, 45, 826–839.Cernak T; Dykstra KD; Tyagarajan S; Vachal P; Krska SW The medicinal chemist’s toolbox for late stage functionalization of drug-like molecules. Chem. Soc. Rev 2016, 45, 546–576.
  • (19).A preprint of this manuscript was published on May 31, 2019:Kanegusuku ALG; Castanheiro T; Ayer SK; Roizen JL Sulfamyl Radicals Direct Photoredox-Mediated Giese Reactions at Unactivated C(3)–H Bonds. ChemRxiv. 2019, Preprint. On June 28, 2019 we became aware of a similar manuscriptMa Z-Y; Guo L-N; You Y; Yang F; Hu M; Duan X-H Visible Light Driven Alkylation of C(sp3) –H Bonds Enabled by 1,6-Hydrogen Atom Transfer/Radical Relay Addition. Org. Lett 2019, 10.1021/acs.orglett.9b01804.
  • (20).(a) Lowry MS; Goldsmith JI; Slinker JD; Rohl R; Pascal RA; Malliaras GG; Bernhard S Single-Layer Electroluminescent Devices and Photoinduced Hydrogen Production from an Ionic Iridium(III) Complex. Chem. Mater 2005, 17, 5712–5719. [Google Scholar]; (b) Cline ED; Bernhard S The Transformation and Storage of Solar Energy: Progress Towards Visible-Light Induced Water Splitting. Chimia 2009, 709–713. [Google Scholar]
  • (21).Bartolamei N; Isse AA; Gennaro A Estimation of standard reduction potentials of alkyl radicals involved in atom transfer radical polymerization. Electrochim. Acta 2010, 55, 8312–8318. [Google Scholar]
  • (22).Rumble John R., Ed., CRC Handbook of Chemistry and Physics 98th Edition (Internet Version 2018) CRC Press/Taylor & Francis, Boca Raton, FL. [Google Scholar]
  • (23).(a) Haans D; Freys J; Bernardinelli G; Wegner OS Cyclometalated Iridium (III) Complexes as Photosensitizers for Long-Range Electron Transfer: Occurrence of a Coulomb Barrier. Eur. J. Inorg. Chem 2009, 4850–4859. [Google Scholar]; (b) Uoyama H; Goushi K; Shizu K; Nomura H; Adachi C Highly efficient organic light-emitting diodes from delayed fluorescence. Nature 2012, 492, 234–238. [DOI] [PubMed] [Google Scholar]
  • (24).Ayer SK, Roizen JL, J. Org. Chem, 2019, 84, 3508–3523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (25).Chen MS; White MC A Predictably Selective Aliphatic C–H Oxidation Reaction for Complex Molecule Synthesis. Science 2007, 318, 783–787. [DOI] [PubMed] [Google Scholar]
  • (26).Roizen JL; Zalatan DN; Du Bois J Selective intermolecular amination of C–H bonds at tertiary carbon centers. Angew. Chem. Int. Ed 2013, 52, 11343–11346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (27).(a) Gal C; Ben-Shoshan G; Rozen S Selective fluorination on tertiary carbon-hydrogen single bonds in the aliphatic series. Tetrahedron Lett. 1980, 21, 5067–5070. [Google Scholar]; (b) Rozen S; Gal C Selective substitution of aliphatic remote tertiary hydrogens by fluorine. J. Org. Chem 1987, 52, 4928–4933. [Google Scholar]
  • (28).For a pioneering example, see:Brodsky BH; Du Bois J Oxaziridine-Mediated Catalytic Hydroxylation of Unactivated 3° C−H Bonds Using Hydrogen Peroxide. J. Am. Chem. Soc 2005, 127, 15391–15393.
  • (29).For a pioneering example, see:Collet F; Lescot C; Liang C; Dauban P Studies in catalytic C–H amination involving nitrene C–H insertion. Dalton Trans. 2010, 39, 10401–10413.
  • (30).Sharma A; Hartwig JF Metal-catalysed azidation of tertiary C–H bonds suitable for late-stage functionalization. Nature 2015, 517, 600–604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (31).Mukherjee S; Maji B; Tlahuext-Aca A; Glorius F Visible-Light-Promoted Activation of Unactivated C(sp3)–H Bonds and Their Selective Trifluoromethylthiolation. J. Am. Chem. Soc 2016, 138, 16200–16203. [DOI] [PubMed] [Google Scholar]
  • (32).Peddibhotla S; Dang Y; Liu JO; and Romo D Simultaneous Arming and Structure/Activity Studies of Natural Products Employing O–H Insertions: An Expedient and Versatile Strategy for Natural Products-Based Chemical Genetics. J. Am. Chem. Soc 2007, 129, 12222–12231.Li J; Cisar JS; Zhou C-Y; Vera B; Williams H; Rodríguez AD; Cravatt BF; Romo D Simultaneous structure-activity studies and arming of natural products by C–H amination reveal cellular targets of eupalmerin acetate. Nature Chem. 2013, 5, 510–517. For a review of application of click-chemistry to chemical biology and drug development, see:Thirumurugan P, Matosiuk D, and Jozwiak K Click Chemistry for Drug Development and Diverse Chemical–Biology Applications. Chem. Rev 2013, 113, 4905–4979.
  • (33).Porter NA; Giese B; Curran DP Acyclic Stereochemical Control in Free-Radical Reactions. Acc. Chem. Res 1991, 24, 296–304.Srikanth GSC; Castle SL Advances in radical conjugate additions. Tetrahedron 2005, 61, 10377–10441.
  • (34).Axon JR; Beckwith ALJ Diastereoselective Radical Addition to Methyleneoxazolidinones: an Enantioselective Route to α-Amino Acids. J. Chem. Soc., Chem. Commun 1995, 549–550. [Google Scholar]
  • (35).Hiskey RG; Jung JM Azomethine Chemistry. II. Formation of Peptides from Oxazolidine-5-ones. J. Am. Chem. Soc 1963, 85, 578–582.Karady S; Amato JS; Weinstock LM Enantioretentive Alkylation of Acyclic Amino Acids. Tetrahedron Lett. 1984, 25, 4337–4340.Seebach D; Fadel AN,O-Acetals from Pivaldehyde and Amino Acids for the alpha-Alkylation with Self-Reproduction of the Center of Chirality. Enolates of 3-Benzoyl-2-(tert-butyl)-1,3-oxazolidin-5-ones. Helv. Chim. Acta, 1985, 68, 1243–1250.Beckwith ALJ; Chai CLL Diastereoselective Radical Addition to Derivatives of Dehydroalanine Dehydrolactiv Acid. J. Chem. Soc., Chem. Commun 1990, 1087–1088.

Associated Data

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

Supplementary Materials

Supporting Information

RESOURCES