1,3-Diamine Formation from an Interrupted Hofmann–Löffler Reaction

Jul 29, 2019 - An iodine-catalyzed Ritter-type amination of nonactivated C–H bonds is presented enabling the formation of 1,3-α-tertiary diamines...
0 downloads 0 Views 1MB Size
Letter Cite This: ACS Catal. 2019, 9, 7741−7745

pubs.acs.org/acscatalysis

1,3-Diamine Formation from an Interrupted Hofmann−Löffler Reaction: Iodine Catalyst Turnover through Ritter-Type Amination Thomas Duhamel,†,‡ Mario D. Martínez,† Ioanna K. Sideri,† and Kilian Muñiz*,†,§ †

Downloaded via GUILFORD COLG on July 29, 2019 at 22:43:20 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Institute of Chemical Research of Catalonia (ICIQ), The Barcelona Institute of Science and Technology, 16 Avgda. Països Catalans, 43007 Tarragona, Spain ‡ Universidad de Oviedo, Julian Clavería, s/n, 33006 Oviedo, Spain § ICREA, Pg. Lluís Companys 23, 08010 Barcelona, Spain S Supporting Information *

ABSTRACT: An iodine-catalyzed Ritter-type amination of nonactivated C−H bonds is presented enabling the formation of 1,3-αtertiary diamines. A sulfamidyl radical serves as the promoter in a guided tertiary C−H iodination through an exclusive 1,6-HAT process. The subsequent Ritter reaction furnishes the C−N bond and establishes an unprecedented concept for catalyst turnover in iodine redox catalysis. The general robustness of the methodology, including broad functional group tolerance, was demonstrated for 24 different 1,3-diamine derivatives, which were synthesized in yields of 42%−99%. KEYWORDS: amination, 1,3-diamines, Hofmann-Löffler reaction, iodine catalysis, Ritter reaction

A

mong the plethora of amino groups, the 1,3-diamine motif is an important structural unit, which is present in several molecules of strategic biological interest, including natural products such as carfentanil, nankakurine A, and the manzacidin family.1 In addition, it has been employed as building blocks in synthetic organic chemistry2 and in ligand engineering for asymmetric transition-metal catalysis.3 In contrast to the related family of 1,2-diamines,4 the corresponding 1,3-diamine scaffold has received significantly less attention in methodology development, and its synthesis has largely remained a domain of transition-metal catalysis.5 We now gave consideration to the concept of the venerable Hofmann−Löffler reaction,6 which mechanistically enables position-selective C−H amination7 through a sulfamidylradical mediated 1,n-hydrogen atom transfer (1,n-HAT). Previously, we have developed C(sp3)−H amination protocols under iodine8 and bromine9 catalysis. While for the class of 1,3-diamines, the direct strategy of C−H functionalization within amidyl radicals would require a challenging 1,4-HAT process and an external nucleophile, we explored alternative amidyl radical precursors based on tracelessly removable functional entities and turned to sulfamides (sulfuric diamides; see Figure 1, conceptual outline). Following the behavior of related sulfamate esters,10 these compounds should enable guided 1,6-HAT processes. We decided to start to investigate the reactivity of sulfamides under the conditions of iodine oxidation catalysis11 for model compound 1a, which contains an activated benzylic carbon position for the rapid cyclization and C−N bond installment (Scheme 1). After optimization of the iodine(III) oxidant,12 the corresponding 1,3-diamine 2a was obtained with an © XXXX American Chemical Society

Figure 1. Structure and synthetic strategy.

Scheme 1. 1,3-Diamine Formation through Hofmann− Löffler Reaction: Initial Substrate Exploration

excellent yield of 90% using molecular iodine as the catalyst source and the iodine(III) reagent PhI(oFBA)2 (oFBA = 2Received: April 17, 2019 Revised: July 22, 2019

7741

DOI: 10.1021/acscatal.9b01566 ACS Catal. 2019, 9, 7741−7745

Letter

ACS Catalysis fluorobenzoate) as the terminal oxidant. Heterocycle 2a was formed as a single diastereoisomer and the expected stereochemistry of two equatorial phenyl substituents was confirmed by X-ray analysis. However, the observed cyclization to 1,3diamines turned out to be limited to benzylic positions. The subsequent attempt to move from the activated benzylic carbon position in 1a to a nonactivated secondary carbon position in 1b did not provide the desired cyclization product. Instead, the corresponding alkyl iodide 2b was isolated in 40% yield for commercial (diacetoxyiodo)benzene (PIDA) as the oxidant.12 Conversion to the 1,3-diamine 2c was only possible with the strong oxidant PhICl2, thus preventing an overall catalytic transformation. To arrive at a synthetically useful catalytic amination reaction at nonactivated C(sp3)−H positions, we decided to tackle tertiary alkyl positions. Amination reactions of this type13 are highly sought-after, since they provide access to the important class of tert-alkyl amines, which constitute important subgroups in naturally occurring alkaloids.14 Among all the reported synthetic approaches, the Ritter-type amination15 at nonactivated C− H positions stands out as a potent strategy, in which harsh conditions16 or transition-metal catalysis17 have so far dominated the stage. Recently, a new concept was designed to perform multiple selective halogenation reactions within the Hofmann−Löffler manifold.10,18 For halogenations at tertiary C−H bonds, the related sulfamate ester group was of unique effectiveness, because of its low tendency to undergo amination itself.10a,b,d This encouraged us to investigate a guided Ritter-type amination under iodine catalysis to access the formation of 1,3-α-tertiary diamines. Indeed, the underlying strategy (Table 1) was to direct the iodination within the Hofmann−Löffler manifold (1,6-HAT) at a nonactivated tertiary carbon position. Following the manifest stability of 2b toward intramolecular amination, we envisaged that a plausible intermolecular

amination could derive from a Ritter-type reaction at the more reactive tertiary carbon position. We used 3a as a model substrate to optimize the reaction conditions. While trying the same system of 20 mol % of molecular iodine and 2 equiv of PhI(mClBA)2 (mClBA = 3chlorobenzoate) as an oxidant,8a we already achieved the good yield of 81% for 4a (Table 1, entry 1). The use of PhI(oFBA)2 slightly improved the yield to 90% (Table 1, entry 2). Moving to PIDA, a yield of 73% was obtained, together with unidentified decomposition products (Table 1, entry 3). Reducing the amount of molecular iodine and PIDA to 10 mol % and 1.2 equiv, respectively, enhanced the selectivity and generated an excellent yield of 95% (Table 1, entry 4). Decreasing the amount of catalyst further, affected the effectiveness of the catalysis (Table 1, entries 5 and 6). Despite an excellent yield of 95% using N-iodosuccinimide NIS,19 more ionic iodine sources did not promote the reaction (Table 1, entries 7 and 8). Finally, exploration of irradiation conditions either under blue or purple LEDs (Table 1, entries 9 and 10) determined optimum conditions for the latter with an isolated yield of 99% for 4a. With optimized conditions in hand, the general applicability of the new methodology for 1,3-α-tertiary diamine formation was explored (Scheme 2). At the outset, it was explored whether a branched substitution pattern was required for the reaction. Gratifyingly, the primary alkyl substitution in 3b led to a comparable yield for 4b (90%), indicating that a Thorpe− Ingold effect is not required. Performing the reaction with 2propionitrile as a solvent in the presence of 20 mol % molecular iodine and 1.5 equiv of PIDA, the corresponding 1,3-diamine 4c was obtained in 83% yield. Comparison with an N-tert-butyl substituent at the sulfamidyl radical provided similar outcome for 4d. To implement the pharmaceutically important class of benzylamines into the 1,3-diamination products, a large number of substrates with different aryl substituents were found compatible, demonstrating the reaction tolerance to either electron-rich or poor arenes (4e−4k, 62%−99%). Next, we investigated the reactivity of a benzyl-protected derivative 3l. The reaction is efficient, providing 4l with a yield of 70%. A successful experiment at 1 mmol scale was also performed, indicating the robustness of the catalysis (4l, 62%). The same holds true for amination at chiral tertiary centers, which form the corresponding chiral tert-alkyl amines 4m and 4n in good yields as equimolar mixtures of diastereomers. For the substrates 3o−3q with different alkyl chains at the internal nitrogen atom containing potentially competing methyl and methylene positions, complete selectivity for the tertiary C−H bond was encountered and 4o−4q were isolated in 60%−82% yield. The C−H amination proceeds equally well at cyclic positions as demonstrated for 4r and 4s. No benzylic functionalization was detected for the former example, and generally, no free radical-derived byproducts have ever been observed under the present protocol. To further demonstrate the exclusive dominance of the guided C−H functionalization, compound 3t was investigated, which exclusively formed 4t, without any reaction at the remote tertiary C−H group. The reaction tolerates functionalized side chains, as demonstrated for the examples of nitrile 4u and acetate 4v. Finally, amination of enantiopure chiral pool derivatives was successfully explored for the proline derivative 4w (70% yield). For the more complex structure of the 5α-androsterone derivative 3x, stereoselective formation of the 5β-aminated product 4x was

Table 1. Ritter-Type Amination through the Hofmann− Löffler Manifold: Reaction Optimization

entry

iodine source (x mol %)

oxidant (y equiv)

yield [%]

1 2 3 4 5 6 7 8 9 10

I2 (20 mol %) I2 (20 mol %) I2 (20 mol %) I2 (10 mol %) I2 (5 mol %) I2 (2.5 mol %) NIS (20 mol %) Bu4NI (20 mol %) I2 (10 mol %) I2 (10 mol %)

PhI(mClBA)2 (2 equiv) PhI(oFBA)2 (2 equiv) PIDA (2 equiv) PIDA (1.2 equiv) PIDA (1.2 equiv) PIDA (1.2 equiv) PIDA (1.2 equiv) PIDA (1.2 equiv) PIDA (1.2 equiv) PIDA (1.2 equiv)

81 90 73 95 48 30 95 0 80a 99b

a Reaction performed under blue LED irradiation. performed under purple LED irradiation.

b

Reaction 7742

DOI: 10.1021/acscatal.9b01566 ACS Catal. 2019, 9, 7741−7745

Letter

ACS Catalysis Scheme 2. Guided Ritter-Type C(sp3)−H Amination: Scope

strated for 4l. It can be deprotected in an orthogonal manner to the two amides 5a and 5b, respectively, and consecutively to monobenzylated 1,3-diamine 5c, maintaining an amino differentiation. Regarding the mechanistic basis for the present transformation, acylhypoiodite (IOAc) is generated at the outset from comproportionation between molecular iodine and PIDA (Figure 2).20 This electrophilic iodine(I) catalyst state

Figure 2. Mechanism of the guided Ritter-type amination within an interrupted iodine-catalyzed Hofmann−Löffler reaction.

iodinates the NH bond of the substrate 3a to form the required intermediate N−I bond A. Unlike other amide sources,8a sulfamides provide effective N-iodination with IOAc. In the presence of light, homolysis to the N-centered radical species B occurs.21 A subsequent 1,6-HAT process to C occurs, followed by an iodination through a radical chain mechanism22 with a quantum yield of 120, to selectively provide the tert-alkyl iodide D. This is the first example of an iodine-catalyzed Hofmann−Löffler process comprising a 1,6HAT. As in the related cases involving a 1,5-HAT,8,9 the experimental value kH/kD = 5.5 from a competition reaction points to the H-abstraction as the most probable slow step. At this stage, a Ritter reaction can proceed to the product and to the regeneration of the iodine catalyst through oxidation. As documented previously,8a,c oxidation to an alkyl iodine(III) prior to the Ritter reaction should provide a more efficient iodine catalysis cycle. The successful application of alkyl iodine(III) intermediates in nonguided Ritter-type aminations has previously been explored by Minakata.23 In this work, an alkyl iodine(III) was required for the Ritter-type amination to occur. Under our conditions, PIDA can act as terminal oxidant to generate the tert-alkyl iodine(III) intermediate E. Despite the application of the iodine high oxidation state as a driving force for nucleophilic substitution, 8a,c no cyclization by the sulfamide is observed interrupting the Hofmann−Löffler pathway.24 Instead, acetonitrile solvent undergoes rapid and irreversible Ritter-type amination at the tertiary carbon position, releasing the original acylhypoiodite(I) catalyst. Water traces in acetonitrile provide the corresponding acetamide 4a. Based on the observed acid

a

Reaction performed with 20 mol % of I2 and 1.5 equiv of PIDA. Isopropionitrile was used as a solvent. bReaction performed over 40 h. c Reaction performed with 15 mol % of I2 and 1.3 equiv of PIDA. d Reaction performed with 1 mmol of 3l. eA 1:1-mixture of diastereoisomers was obtained. fYield based on recovering starting material. gReaction performed over 24 h. hReaction performed with 2 equiv of AcOH as additive. iStarting from the 5α-androsterone isomer. jNaOH, ethylene glycol, 200 °C. k1,3-Propanediamine, 150 °C.

observed (yield of 65%) and confirmed by single-crystal X-ray diffraction. This stereochemical outcome can be rationalized from shielding of the α-face at the stage of acetonitrile attack. Deprotection of the sulfamide directing group was demon7743

DOI: 10.1021/acscatal.9b01566 ACS Catal. 2019, 9, 7741−7745

Letter

ACS Catalysis

nucleophile engages in the final carbon−heteroatom bond formation.

effect, we cannot exclude that the corresponding N-protonated species may be participating throughout the catalytic cycle.25 Regarding the overall transformation, so far, there has been no precedence for a metal-free catalytic Ritter-type amination of nonactivated tertiary C−H bonds. Our work demonstrates that such a process is feasible. It also introduces the intermolecular Ritter reaction as a useful concept for ensuring catalyst turnover for the given example of sulfamides, which do not generate a nucleophilic nitrogen group for intramolecular C−N bond formation. As a result, the present protocol is the first to use the sulfamide group as a source for sulfamidyl radical formation under catalytic conditions. Previous work on related sulfamate esters in Hofmann-Löffler chemistry has remained stoichiometric with regard to halogen10a,b,d or pseudo-halogen26 reagents. Importantly, when turning to the corresponding sulfamate ester 6, the iodine-catalyzed amination did not proceed (Scheme 3). In order to generate an iodine-catalyzed entry



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acscatal.9b01566. Details on experimental procedures for the catalytic reactions, spectroscopic data for the products (PDF) Crystallographic data for 2a (CIF) Crystallographic data for 4m (CIF) Crystallographic data for 4w (CIF) Crystallographic data for 2c (CIF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Scheme 3. Guided Oxygenation Reactions

Kilian Muñiz: 0000-0002-8109-1762 Author Contributions

Experiments were performed by T.D., M.D.M., and I.S., and the manuscript was written by T.D. and K.M. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors thank Dr. E. C. Escudero-Adán and Dr. M. Martı ́nez for the X-ray structure determinations, Dr. G. Goti for support with the quantum yield determination, and the Spanish Ministry for Economy and Competitiveness and FEDER (Grant No. CTQ2014-56474R to K.M.), and the region of Catalonia for financial support.



REFERENCES

(1) (a) Misailidi, N.; Papoutsis, I.; Nikolaou, P.; Dona, A.; Spiliopoulou, C.; Athanaselis, S. Fentanyls continue to replace heroin in the drug arena: the cases of ocfentanil and carfentanil. Forensic Toxicol. 2018, 36, 12−32. (b) Kang, S. H.; Kang, S. Y.; Lee, H. S.; Buglass, A. J. Total Synthesis of Natural tert-Alkylamino Hydroxy Carboxylic Acids. Chem. Rev. 2005, 105, 4537−4558. (c) Hirasawa, Y.; Morita, H.; Kobayashi, J. Nankakurine A, a Novel C16N2-Type Alkaloid from Lycopodium hamiltonii. Org. Lett. 2004, 6, 3389−3391. (d) Busscher, G. F.; Rutjes, F. P. J. T.; van Delft, F. L. 2Deoxystreptamine: Central Scaffold of Aminoglycoside Antibiotics. Chem. Rev. 2005, 105, 775−791. (2) (a) Halli, J.; Bolte, M.; Bats, J.; Manolikakes, G. Modular TwoStep Approach for the Stereodivergent Synthesis of 1,3-Diamines with Three Continuous Stereocenters. Org. Lett. 2017, 19, 674−677. (b) Liew, S. K.; He, Z.; St. Denis, J. D.; Yudin, A. K. Stereocontrolled Synthesis of 1,2- and 1,3-Diamine Building Blocks from Aziridine Aldehyde Dimers. J. Org. Chem. 2013, 78, 11637−11645. (3) Facchetti, G.; Gandolfi, R.; Fusè, M.; Zerla, D.; Cesarotti, E.; Pellizzoni, M.; Rimoldi, I. Simple 1,3-diamines and their application as ligands in ruthenium(II) catalysts for asymmetric transfer hydrogenation of aryl ketones. New J. Chem. 2015, 39, 3792−3800. (4) (a) Lucet, D.; Le Gall, T.; Mioskowski, C. The Chemistry of Vicinal Diamines. Angew. Chem., Int. Ed. 1998, 37, 2580−2627. (b) Bennani, Y. L.; Hanessian, S. trans-1,2-Diaminocyclohexane Derivatives as Chiral Reagents, Scaffolds, and Ligands for Catalysis: Applications in Asymmetric Synthesis and Molecular Recognition. Chem. Rev. 1997, 97, 3161−3196. (c) Muñiz, K.; Martínez, C. The Development of Intramolecular Vicinal Diamination of Alkenes: From Palladium to Bromine Catalysis. J. Org. Chem. 2013, 78, 2168−2174.

into 1,3-aminoalcohol derivatives, we briefly explored an oxygenative termination of the C−H functionalization. Indeed, upon working in acetic acid as a cosolvent, substrate 3a is cleanly converted to the corresponding product 7 containing an acetoxylated tertiary carbon center.27 A related intramolecular oxygenation was obtained for the two substrates 8a and 8b with a lateral ester and ether substituent, respectively. In their cases, the corresponding iodine(III) intermediate D of the catalytic cycle is nucleophilically intercepted by the oxygenated functional groups to provide THF derivative 9a or lactone 9b, respectively. The propensity of the oxygenated groups to undergo C−O bond formation contrasts with the apparent absent nucleophilicity of the current sulfamide unit.22 These results demonstrate that reactions other than amination may be obtained from the present iodine catalysis. In summary, we have identified the sulfamide as a nontransferable directing group in C−H functionalization. Applying this concept, we have developed a new entry into the iodine-catalyzed C(sp3)−H amination reaction using the Hofmann−Löffler approach to address a guided 1,6-HAT, followed by a Ritter amination, which interrupts the usual cyclization and enables an elusive intermolecular nucleophilic regeneration of the iodine catalyst. These results demonstrate the kinetic feasibility of 1,6-HAT processes to participate in catalytic Hofmann−Löffler reactions and constitute the first example of such transformations, in which an external 7744

DOI: 10.1021/acscatal.9b01566 ACS Catal. 2019, 9, 7741−7745

Letter

ACS Catalysis

(15) For a review for the Ritter reaction, see: Jiang, D.; He, T.; Ma, L.; Wang, Z. Recent developments in Ritter reaction. RSC Adv. 2014, 4, 64936−64946. (16) (a) Olah, G. A.; Balaram Gupta, B. G. Synthetic methods and reactions. 86. Novel synthesis of N-(1-adamantyl)amides from adamantane. J. Org. Chem. 1980, 45, 3532−3533. (b) Kalkhambkar, R. G.; Waters, S. N.; Laali, K. K. Highly efficient synthesis of amides via Ritter chemistry with ionic liquids. Tetrahedron Lett. 2011, 52, 867−871. (17) Michaudel, Q.; Thevenet, D.; Baran, P. S. Intermolecular Ritter-Type C−H Amination of Unactivated sp3 Carbons. J. Am. Chem. Soc. 2012, 134, 2547−2550. (18) Wappes, E.; Vanitcha, A.; Nagib, D. β C−H di-halogenation via iterative hydrogen atom transfer. Chem. Sci. 2018, 9, 4500−4504. (19) O’Broin, C. Q.; Fernández, P.; Martínez, C.; Muñiz, K. NIodosuccinimide-Promoted Hofmann−Löffler Reactions of Sulfonimides under Visible Light. Org. Lett. 2016, 18, 436−439. (20) (a) De Armas, P.; Carrau, R.; Concepción, J. L.; Francisco, C. G.; Hernández, R.; Suárez, E. Synthesis of 1,4-epimine compounds. Iodosobenzene diacetate, an efficient reagent for neutral nitrogen radical generation. Tetrahedron Lett. 1985, 26, 2493−2496. (b) Carrau, R.; Hernández, R.; Suárez, E.; Betancor, C. Intramolecular functionalization of N-cyanamide radicals: synthesis of 1,4-and 1,5-Ncyanoepimino compounds. J. Chem. Soc., Perkin Trans. 1 1987, 1, 937−943. (c) De Armas, P.; Francisco, C. G.; Hernández, R.; Salazar, J. A.; Suárez, E. Steroidal N-nitroamines. Part 4. Intramolecular functionalization of N-nitroamine radicals: synthesis of 1,4-nitroimine compounds. J. Chem. Soc., Perkin Trans. 1 1988, 1, 3255−3265. (d) Fan, R.; Pu, D.; Wen, F.; Wu, J. δ and α sp3 C−H Bond Oxidation of Sulfonamides with PhI(OAc)2/I2 under Metal-Free Conditions. J. Org. Chem. 2007, 72, 8994−8997. (e) Achar, T. K.; Maiti, S.; Mal, P. PIDA-I2 mediated direct vicinal difunctionalization of olefins: iodoazidation, iodoetherification and iodoacyloxylation. Org. Biomol. Chem. 2016, 14, 4654−4663. (21) (a) Kärkäs, M. D. Photochemical Generation of NitrogenCentered Amidyl, Hydrazonyl, and Imidyl Radicals: Methodology Developments and Catalytic Applications. ACS Catal. 2017, 7, 4999− 5022. (b) Sakić, D.; Zipse, H. Radical Stability as a Guideline in C−H Amination Reactions. Adv. Synth. Catal. 2016, 358, 3983−3991. (22) Cismesia, M. A.; Yoon, T. P. Characterizing chain processes in visible light photoredox catalysis. Chem. Sci. 2015, 6, 5426−5434. (23) (a) Kiyokawa, K.; Takemoto, K.; Minakata, S. Ritter-type amination of C−H bonds at tertiary carbon centers using iodic acid as an oxidant. Chem. Commun. 2016, 52, 13082−13085. (b) Kiyokawa, K.; Watanabe, T.; Fra, L.; Kojima, T.; Minakata, S. Hypervalent Iodine(III)-Mediated Decarboxylative Ritter-Type Amination Leading to the Production of α-Tertiary Amine Derivatives. J. Org. Chem. 2017, 82, 11711−11720. (24) At present, we cannot rule out an intermediary participation of the sulfamide in a reversible addition to the tert-alkyl position, which may involve the sulfonyl groups. For a related nonproductive Oalkylation, see: (a) Funes-Ardoiz, I.; Sameera, W. M. C.; Romero, R. M.; Martínez, C.; Souto, J. A.; Sampedro, D.; Muñiz, K.; Maseras, F. DFT Rationalization of the Diverse Outcomes of the Iodine(III)Mediated Oxidative Amination of Alkenes. Chem. - Eur. J. 2016, 22, 7545−7553. (b) Muñiz, K.; Barreiro, L.; Romero, R. M.; Martínez, C. Catalytic Asymmetric Diamination of Styrenes. J. Am. Chem. Soc. 2017, 139, 4354−4357. (25) A radical cation B would follow the original Hofmann−Löffler mechanism and would also explain the low nucleophilicity of the nitrogen group toward nucleophilic substitution. (26) Ayer, S. K.; Roizen, J. L. Sulfamate Esters Guide C(3)-Selective Xanthylation of Alkanes. J. Org. Chem. 2019, 84, 3508−3523. (27) Kiyokawa, K.; Okumatsu, D.; Minakata, S. Hypervalent iodine(III)-mediated decarboxylative acetoxylation at tertiary and benzylic carbon centers. Beilstein J. Org. Chem. 2018, 14, 1046−1050.

(d) De Jong, S.; Nosal, D. G.; Wardrop, D. J. Methods for direct alkene diamination, new & old. Tetrahedron 2012, 68, 4067−4105. (5) (a) Kurokawa, T.; Kim, M.; Du Bois, J. Synthesis of 1,3Diamines Through Rhodium-Catalyzed C-H Insertion. Angew. Chem., Int. Ed. 2009, 48, 2777−2779. (b) Wehn, P. M.; Du Bois, J. Enantioselective Synthesis of the Bromopyrrole Alkaloids Manzacidin A and C by Stereospecific C−H Bond Oxidation. J. Am. Chem. Soc. 2002, 124, 12950−12951. (c) Li, C.; Lang, K.; Lu, H.; Hu, Y.; Cui, X.; Wojtas, L.; Zhang, P. X. Catalytic Radical Process for Enantioselective Amination of C(sp3)−H Bonds. Angew. Chem., Int. Ed. 2018, 57, 16837−16841. For a review on 1,3-diamines, see: (d) Ji, X.; Huang, H. Synthetic methods for 1,3-diamines. Org. Biomol. Chem. 2016, 14, 10557−10566. (e) Paradine, S. M.; Griffin, J. R.; Zhao, J.; Petronico, A. L.; Miller, S. M.; White, M. C. A manganese catalyst for highly reactive yet chemoselective intramolecular C(sp3)H amination. Nat. Chem. 2015, 7, 987−994. (6) (a) Wolff, M. E. Cyclization of N-Halogenated Amines (The Hofmann-Löffler Reaction). Chem. Rev. 1963, 63, 55−64. (b) Neale, R. S. Nitrogen Radicals as Synthesis Intermediates. N-Halamide Rearrangements and Additions to Unsaturated Hydrocarbons. Synthesis 1971, 1971, 1−15. (c) Stella, L. Homolytic Cyclizations of N-Chloroalkenylamines. Angew. Chem., Int. Ed. Engl. 1983, 22, 337− 350. (7) For reviews on C−H aminations, see: Díaz-Requejo, M. M.; Pérez, P. J. Chem. Rev. 2008, 108, 3379−3394. (b) Collet, F.; Dodd, R. H.; Dauban, P. Chem. Commun. 2009, 5061−5074. (c) Collet, F.; Lescot, C.; Dauban, P. Chem. Soc. Rev. 2011, 40, 1926−1936. (8) (a) Martínez, C.; Muñiz, K. An Iodine-Catalyzed HofmannLöffler Reaction. Angew. Chem., Int. Ed. 2015, 54, 8287−8291. (b) Becker, P.; Duhamel, T.; Stein, C. J.; Reiher, M.; Muñiz, K. Cooperative Light-Activated Iodine and Photoredox Catalysis for the Amination of Csp3-H Bonds. Angew. Chem., Int. Ed. 2017, 56, 8004− 8008. (c) Duhamel, T.; Stein, C. J.; Martínez, C.; Reiher, M.; Muñiz, K. Engineering Molecular Iodine Catalysis for Alkyl−Nitrogen Bond Formation. ACS Catal. 2018, 8, 3918−3925. (9) Becker, P.; Duhamel, T.; Martínez, C.; Muñiz, K. Designing Homogeneous Bromine Redox Catalysis for Selective Aliphatic C−H Bond Functionalization. Angew. Chem., Int. Ed. 2018, 57, 5166−5170. (10) (a) Short, M. A.; Blackburn, J. M.; Roizen, J. L. Sulfamate Esters Guide Selective Radical-Mediated Chlorination of Aliphatic C−H Bonds. Angew. Chem., Int. Ed. 2018, 57, 296−299. (b) Sathyamoorthi, S.; Banerjee, S.; Du Bois, J.; Burns, N. Z.; Zare, R. N. Site-selective bromination of sp3-C-H bonds. Chem. Sci. 2018, 9, 100−104. (c) Zalatan, D. N.; Du Bois, J. Oxidative Cyclization of Sulfamate Esters Using NaOCl - A Metal-Mediated Hofmann-LöfflerFreytag Reaction. Synlett 2009, 2009, 143−146. (d) Del Castillo, E.; Martínez, M. D.; Bosnidou, A. E.; Duhamel, T.; O’Broin, C. Q.; Zhang, H.; Escudero-Adán, E. C.; Martínez-Belmonte, M.; Muñiz, K. Multiple Halogenation of Aliphatic C−H Bonds within the Hofmann−Löffler Manifold. Chem. - Eur. J. 2018, 24, 17225−17229. (11) (a) Finkbeiner, P.; Nachtsheim, B. Iodine in Modern Oxidation Catalysis. Synthesis 2013, 45, 979−999. (b) Yusubov, M. S.; Zhdankin, V. V. Iodine catalysis: A green alternative to transition metals in organic chemistry and technology. Resource-Efficient Technologies 2015, 1, 49−67. (c) Uyanik, M.; Ishihara, K. Catalysis with In Situ-Generated (Hypo)iodite Ions for Oxidative Coupling Reactions. ChemCatChem 2012, 4, 177−185. (12) See the Supporting Information for more details. (13) Trowbridge, A.; Reich, D.; Gaunt, M. Multicomponent synthesis of tertiary alkylamines by photocatalytic olefin-hydroaminoalkylation. Nature 2018, 561, 522−527. (14) (a) Hager, A.; Vrielink, N.; Hager, D.; Lefranc, J.; Trauner, D. Synthetic approaches towards alkaloids bearing α-tertiary amines. Nat. Prod. Rep. 2016, 33, 491−522. (b) Clayden, J.; Donnard, M.; Lefranc, J.; Tetlow, D. J. Quaternary centres bearing nitrogen (α-tertiary amines) as products of molecular rearrangements. Chem. Commun. 2011, 47, 4624−4632. 7745

DOI: 10.1021/acscatal.9b01566 ACS Catal. 2019, 9, 7741−7745