“This document is the Accepted Manuscript version of a Published Work that appeared in final form in ACS Catalysis, copyright © American Chemical Society after peer review and technical editing by the publisher. To access the final edited and published work see https://pubs.acs.org/doi/abs/10.1021/acscatal.8b00286“. Engineering Molecular Iodine Catalysis for AlkylNitrogen Bond Formation Thomas Duhamel,†,§ Christopher J. Stein, ‡ Claudio Martínez,† Markus Reiher‡* and Kilian Muñiz†,∫* † Institute of Chemical Research of Catalonia (ICIQ), Av. Països Catalans 16, 43007 Tarragona, Spain. § Facultad de Química, Universidad de Oviedo, C/Julián Claveria, 33006 Oviedo, Spain. ‡ Laboratorium für Physikalische Chemie, ETH Zürich, Vladimir-Prelog-Weg 2, 8093 Zürich, Switzerland. ∫ ICREA, Pg. Lluís Companys 23, 08010 Barcelona, Spain. KEYWORDS: C-H Amination, Catalysis, Hofmann-Löffler reaction, Iodine, Oxidation. ABSTRACT An advanced protocol for the intramoleclar C-H amination of alkyl groups via amidyl radicals (Hofmann-Löffler reaction) under homogeneous iodine catalysis is reported. This protocol employs common mCPBA as terminal oxidant. It proceeds under mild conditions, with complete chemoselectivity, is compatible with radical intermediates and allows for the selective intramolecular amination reaction of secondary and tertiary hydrocarbon bonds and is not restricted to benzylic C-H amination. The involvement of an iodine(III) catalyst state in the C-N bond formation derives from selective oxidation at the stage of the corresponding alkyl iodide with mCPBA. Its formation is corroborated by quantum-chemical calculations. This new catalysis thus proceeds within a defined iodine(I/III) catalysis manifold. Key Words Amination, C-H Functionalisation, Hofmann-Löffler Reaction, Iodine, Oxidation, Radicals 1 Introduction Homogeneous oxidation catalysis based on molecular iodine1 has emerged as a viable answer to the quest for alternative concepts complimenting conventional transition metal based transformations. Within this context, seminal iodine-catalyzed a-oxygenation reactions of ketones were developed by Ishihara.2 We have recently presented selective Csp3-H amination reactions that constitute the first Hofmann-Löffler reactions employing only a catalytic amount of iodine (Scheme 1).3 Despite the common feature of deriving from iodine catalysis, the two protocols display significantly different behavior depending on the respective re-oxidant. The initial protocol made use of a stoichiometric amount of a non-commercial iodine(III) reagent as terminal oxidant (eq. 1). This variant optimizes earlier stoichiometric protocols4,5 and demonstrates a remarkably robust main group catalysis based on a defined iodine(I/III) manifold.6 The broad scope of this catalytic transformation includes aliphatic primary, secondary and tertiary C-H bonds alike. The hypervalent iodine(III) reagent as terminal oxidant (i) provides an alkyliodine(III) intermediate for the observed general aliphatic C-N bond formation and (ii) leads to the generation of electrophilic iodine species I-O2CR, as demonstrated independently.3a,7,8 Despite the clean reaction conditions, the requirement for a preformed iodine(III) reagent may be considered less optimum from economic and practical perspectives. In search for a more economic re-oxidant, a terminal oxidation based on an organic photoredox catalyst was developed (eq. 2).3b This reaction proceeds under significantly more benign conditions, however, due to its iodine(-I/I) manifold, it is limited to the C-H amination at benzylic positions. Iodine(I/III) catalysis: requirement for a preformed iodine(III) oxidant3a I2 (2.5 - 5 mol%) H R´´ (ArCO2)2IPh (1 equiv) NHSO2R R´ R´ R´´ N DCE, RT, 12h, SO2R unfunctionalized [visible light] C-H amination products remote C-H bond (1) Iodine(-I/I) catalysis: limitation to benzylic C-H positions3b H H NHSO2R Ar unfunctionalized remote benzylic C-H bond I2 (5 mol%), TPT (1 mol%) HFIP/DCE, RT, 18h, [blue LED] Ar H N SO2R (2) C-H amination products 2 Scheme 1. Iodine-catalyzed Hofmann-Löffler reactions using hypervalent iodine as terminal oxidant (1) or photoredox catalysis for terminal oxidation (2). Ar = 3-Cl-C6H4. Despite its undisputable progress,1 iodine redox catalysis still requires the development of defined general and benign re-oxidation conditions for a broader application in organic synthesis. Some interesting contributions have arisen in the recent past, which include the combination of molecular iodine with cleaner oxidants such as peracids or peroxides.1,2 We here report on a modified protocol of increased economic nature for catalytic Hofmann-Löffler reactions of general substrate scope. It relies on a suitable engineering of the iodine catalyst, which is generated from molecular iodine as a single iodine component in combination with a commercial terminal oxidant. 2. Results and Discussion 2.1. Reaction Development In previous investigation, we had disclosed that several attempts to implement common hypoiodite catalysis2 for catalytic Hofmann-Löffler reactions were unsuccessful. In order to render the reaction more economic and replace the required preformed hypervalent iodine as terminal oxidant, we were intrigued to test whether common organic peroxides or peracids could promote the desired Hofmann-Löffler reaction (Table 1).9 While tert-butyl hydroperoxide (TBHP) was completely ineffective (entry 1), peracetic acid showed some low conversion (entries 2,3). In contrast, mCPBA in acetonitrile/acetic acid directly provided a 52% of yield (entry 4). Using a limiting amount of this oxidant in acetonitrile, 44% product was observed (entry 5). Screening of alternative solvents in the presence of a three-fold excess of mCPBA revealed that acetonitrile represented the best solvent (entries 6-11). Finally, addition of tertbutanol as co-solvent provided quantitative conversion and 98% isolated yield of 2a (entry 12), which was maintained at 2.2 equivalents of terminal oxidant (entry 13). Under these conditions, reduced iodine loadings of even 2.5 mol% provide lower, albeit synthetically useful yields (entries 13-15). 3 Table 1. Iodine-catalyzed Hofmann-Löffler reactions: Optimization of reaction conditions for mCPBA as terminal oxidant. NHTs Ph conditions Ph N Ts 12 h, 25 ºC 1a 2a No Conditions Conversiona 1 I2 (15%), TBHP (3.0 eq), CH3CN/tBuOH (1/1) s.m 2 I2 (15%), CH3CO3H (3 eq), DCE 12% 3 I2 (15%), CH3CO3H (3 eq), CH3CN s.m 4 I2 (15%), mCPBA (3 eq), CH3CN/AcOH (1/1) 52% 5 I2 (15%), mCPBA (1 eq), CH3CN 44% 6 I2 (15%), mCPBA (3 eq), PhCN 38% 7 I2 (15%), mCPBA (3 eq), AcOH 32% 8 I2 (15%), mCPBA (3 eq), TFE 37% 9 I2 (15%), mCPBA (3 eq), MeOH 26% 10 I2 (15%), mCPBA (3 eq), EtOAc 56% 11 I2 (15%), mCPBA (3 eq), CH3CN 60% 12 I2 (15%), mCPBA (3 eq), CH3CN/tBuOH (1/1) 98% b 13 I2 (15%), mCPBA (2.2 eq), CH3CN/tBuOH (1/1) 98% b 14 I2 (5%), mCPBA (2.2 eq), CH3CN/tBuOH (1/1) 70% b 15 I2 (2.5%), mCPBA (2.2 eq), CH3CN/tBuOH (1/1) 56% b a Estimated conversion from integration of the 1H NMR spectrum of the crude reaction product. b Isolated yield after purification. 4 2.2. Scope of the iodine/mCPBA catalyzed Hofmann-Löffler reaction Under the optimized conditions, a number of pyrrolidines were prepared from the corresponding linear sulfonamides 1a-aa (Scheme 2). In agreement with earlier protocols,3a cyclization onto benzylic positions is straightforward and includes variation at the nitrogen substitution (2b), along the aliphatic chain (2c,k), at the arene (2d-2i, including an acetylenic group) and at the benzylic position (2l,m). The accurate screening of re-oxidants has identified mCPBA to provide suitable conditions for the important substrate class of C-H amination even at non-benzylic aliphatic positions (2n-r), including the new compounds 2o and 2p/2p’. Here, the synthesis of the a-tertiary amine derivative 2r stands out, as it represents an uncommon synthetic access to this important class of compounds in natural alkaloids.10 The reaction proceeds at alpha-position to heteroatoms (2s) and with excellent diastereoselectivity for cyclic stereocontrol (2t-u), while acyclic stereocontrol is not possible due to the involved radical pathway (2v/v’). Transannular CH amination11 is also possible (2w). 5 R'' I2 (15 mol%) mCPBA (2.2 equiv.) R' RO2SHN R'' R' N SO2R visible light, CH3CN/tBuOH, 25 ºC, 12 h H 1a-z 2a-z Ph N SO2R Ph 2a (R = Tol): 95% 2b (R = 4-NO2-Ph): 63%a N Ts N Ts N Ts 2l: 36%a 2k: 54% X 2d (X = F): 83% 2e (X = Cl): 82% 2f (X = Br): 73% 2g (X = Me): 96% 2h (X = OMe): 99% 2i (X = CCPh): 58% 2c: 99% Ph Ph N Ts N Ts Ph Ph 2m: 84%a Ph N Ts Ph N Ts N Ts 2n: 58% 2o: 54% N Ts N Ts 2r: Ph 2p/2p´: 1:1 d.r., 69% N Ts N Ts 2q: 65% Ph + 61%a,b OMe TsHN NHTs 2s: 60% Ph Ph N Ts 2t: 100:0 d.r., 96% O2 S N Ts 2u: 100:0 d.r., 85% O2S N N Ph N Ts 2v/2v´: 1:1 d.r., 94% + 2x (R = Me): 63% 2y(R = Et): 85% NHTs Ph + R Ph 2w: 70% N Ts N Ts Ph Ph N Ts 2z/2z': 1:1 d.r., 67% I2 (15 mol%) mCPBA (3 equiv) NTs CH3CN/tBuOH (1/1) 3a 4a (40%)c 6 Scheme 2. Iodine-catalyzed Hofmann-Löffler reaction with mCPBA: Scope. a Reaction in 1,2-dichloroethane. b with 3 equivalents of mCPBA. c >55% recovered starting material. New substrates were investigated for the amination of 1x and 1y, which cyclized under the present conditions, while they are entirely unreactive under the previous ones.3 Compound 1z demonstrates that in the case of competitive secondary hydrocarbon bonds, the common 1,5HAT12,13 is preferred for the weaker benzylic C-H bond, thus providing 2z/2z’ as the exclusive amination products. The reaction was further expanded to an interesting example of sixmembered ring formation. For the case of compound 3a, clean formation of the corresponding tetrahydroisoquinoline 4a was observed (40%, together with amost fully recovered remaining starting material). This example represents a rare case of a 1,6-HAT reaction,14-16 which in the present case is the result of an impossible 1,5-HAT. However, the potentially competing aelimination to an aldimine6f was not observed at all for this case. Noteworthily, previous iodine catalyses do not promote this particular reaction. In order to further demonstrate the difference between the individual protocols, we explored the performance of the benzyl methylether derivative 3b (Scheme 3). While the iodine/iodine(III) system3a provides the expected C-H amination product 4c, the present iodine/mCPBA variant promotes overoxidation to the lactam 4b. Finally, the inherent reaction differences between the protocols of the reoxidants iodine(III) and mCPBA may open new synthetic possibilities. For the case of a substrate with a tertiary alcohol, clean Hofmann-Löffler cyclization is obtained for the substrates 5a,b in the presence of the iodine/mCPBA system, yielding the pyrrolidine products 6a,b in good isolated yields. In contrast, the iodine/iodine(III) system leads to concomitant opening of the tertiary alcohol to form w-iodo ketones. Such a scission reaction had previously been reported by Barluenga and González for the use of a cationic iodonium reagent.17 Since the active iodine species is also an electrophilic iodine(I) in the present case, ketone formation represents the expected outcome. However, the incorporation of the iodine into the carbon framework of the product prevents the progress of the catalytic Hofmann-Löffler cyclization, and, consequently, yields could not exceed the range of 10%. When conducting the reaction in the presence of equimolar reagent amounts, synthetically relevant yields of 7a,b could be obtained. 7 NTs O 4b: 99% NHTs I2 (15 mol%) mCPBA (2.2 equiv) CH3CN/tBuOH (1/1) I2 (5 mol%) PhI(O2CAr)2 (1 equiv) OMe NTs (CH2Cl)2 OMe 3b 4c: 89% I2 (15 mol%) mCPBA (3 equiv) CH3CN/tBuOH (1/1) N Ts HO TsHN HO 5a (n = 1) 5b (n = 3) 6a: 60% 6b: 50% n n I2 (1 equiv) PhI(O2CAr)2 (2 equiv) (CH2Cl)2 N Ts I n 7a (n = 10): 63% 7b (n = 13): 79% O Scheme 3. Iodine-catalyzed Hofmann-Löffler reactions with mCPBA: Influence of the terminal oxidant. Ar = 3-Cl-C6H4. 2.3 Mechanistic Discussion Several control experiments were conducted to gain insight into the mechanism of the present transformation (Scheme 4). At the outset, it is noteworthy that mCPBA as the optimum oxidant generates the 3-chlorobenzoate anion, which had been identified as a component in the best hypervalent iodine reagent in our earlier reaction.3a This could suggest that these two iodine catalyses actually proceed through the formation of identical iodine(I) states such as I-O2CAr.7 However, control experiments with the isolated compound 8 showed a completely different picture. First, a reaction employing 8 as catalyst source led to no reaction either with mCPBA alone or in the presence of additional free 3-chlorobenzoic acid. Next, it was observed that tetrabutylammonium iodide could also not serve as catalyst source upon oxidation with mCPBA. 8 NHTs conditions Ph 1a visible light CH3CN/tBuOH, 12h, 25 ºC Ph N Ts 2a conditions [(ArCO2)2I]NBu4 8 (30 mol%), mCPBA (1.5 eq) No reaction [(ArCO2)2I]NBu4 8 (30 mol%), mCPBA (1.5 eq) ArCO2H (2 eq) No reaction INBu4 (30 mol%), mCPBA (1.5 eq) No reaction IOtBu (1.2 equiv), c-C6H12 51% IOtBu (30 mol%), c-C6H12, mCPBA (1.5 equiv) 83% I2 (15 mol%), mCPBA (2.2 eq), dark laboratory No reaction Scheme 4. Control Experiments. Ar = 3-Cl-C6H4. These observations make an iodine(I) catalyst state IO2CAr less probable. In view of the positive influence of the tBuOH co-solvent, the participation of tert-butyl hypoiodite IOtBu becomes an intriguing alternative. We investigated the participation of this reagent using its in situ formation following Wirth’s protocol.18 Indeed, a stoichiometric amount of this oxidant promotes a 51% isolated yield for the standard transformation of 1a into 2a. When IOtBu is employed as the catalyst source in the presence of mCPBA as terminal oxidant under standard conditions, the reaction outcome closely resembles the yield from the corresponding reaction with in situ catalyst formation. 9 ArCO2H Ts NH O I Ph N Ts 2a ArCO3H Ph Ts NH I E Ph Ph D I-OH Ts N HOtBu H2O I Ts NH C Ts N I I-OtBu Ph NHTs HOtBu A B Ph Initiation: initiation:18a I2 + HOtBu Ph 1a hυ, - I kinetic isotope effect: 1.4 [mCPBA] vs 4.0 [PhI(O2CAr)2]3a quantum yield: 3.0 [mCPBA] vs 49 [PhI(O2CAr)2]3a Figure 1. Catalytic cycle for the iodine-catalyzed Hofmann-Löffler reaction with mCPBA as terminal oxidant. Values for control experiments are given with the corresponding oxidant in brackets. Ar = 3-Cl-C6H4. With these observations in hand, it is plausible to conclude that the reaction proceeds through an iodine catalysis that involves IOtBu as a competent catalyst state (Figure 1). This compound is formed at the outset from molecular iodine and tBuOH.18 Reaction with the substrate 1a provides the crucial intermediate A. In the absence of tBuOH, intermediate A is directly formed from IOH.3b Obviously, this latter pathway is slower than the one through tBuOI, and this context explains the observed acceleration effect of tBuOH. Intermediate A then undergoes photolytic NI bond cleavage as investigated previously.3 This homolytic light-induced amidyl radical formation19 is in agreement with the corresponding control experiment in the absence of light, which does not show product formation (Scheme 4). It obviously overrides the potentially competing unselective radical C-H iodination by the tBuOI reagent itself.18a The resulting intermediate B promotes a 1,5-HAT to C,3,19,20 which engages in a radical chain reaction with A 10 to finally generated the aliphatic iodide D. Alternatively, radical recombination with iodine would provide the same outcome in a direct transformation. In view of the demonstrated successful reaction scope including aliphatic, non-benzylic C-H amination, it can be concluded that the reaction should involve an alkyl-iodine(III) intermediate E prior to nucleophilic C-N bond formation to product 2a. The enhanced leaving group capacity21 of alkyliodine(III) E over the comparable iodine(I) state D enables amination reactions outside the activated benzylic or neighboring heteroatom scaffold. This postulation of an alkyliodine(III) catalyst state E is in agreement with literature precedence on the formation of iodosyl alkanes from alkyl iodides upon oxidation with mCPBA.22,23 A somewhat slower leaving group capacity than for related alkyl[(dicarboxy)iodine](III) intermediates3 is generally expected for iodosyl derivative E. This assumption should affect the relative rate of C-N bond formation, and the subsequent nucleophilic substitution should be slightly slower than in the case of comparable di(acyloxy)iodosylalkanes,3 which display a comparably higher nucleophuge character.24 This observation is corroborated by Hammett studies9 and by a pronounced observation of an electronic influence of the nitrogen substituent on the reaction. For a competition experiment between 1a and 1b, the less nucleophilic nosyl group displays a decreased reactivity and the two products are formed in a 1.7:1 ratio in favor of 2a (Scheme 5). NHTs NHNs + I2 (15 mol%), mCPBA (1 equiv) CH3CN/tBuOH (1/1), 12h, 25 ºC 1a 1b N Ts + N Ns 89%, 2a:2b = 1.7:1 Scheme 5. Competition experiment between 1a and 1b. Conversion based on the terminal oxidant mCPBA as limiting agent. In any case, the leaving group ability of the alkyliodine(III) intermediate is required to guarantee a sufficient rate in the iodine-based catalytic cycle, in particular for those examples, that promote amination of non-activated carbon atoms. In comparison to the previous catalytic HofmannLöffler reaction with iodine(III) PhI(O2CAr)2 as the terminal oxidant, the relative rates of the individual steps of both catalytic cycles are different. First, the 1,5-HAT step now occurs at comparably faster relative rate as suggested from the corresponding KIE of 1.49 measured for the 11 overall transformation from 1a to 2a. In addition, the observed quantum yield of only 39 suggests that significant interruption of the radical chain reaction occurs. This may be the result of partial radical recombination with an iodine radical at the intermediary state B, and due to the comparably slow regeneration of the crucial intermediate A, even at a comparably higher iodine catalyst loading. 38.29 35.42 ‡ 0.00 1.37 E / kcal mol 1 24.23 -15.57 -20.25 ‡ -34.09 -63.26 Figure 2. Electronic energy profile of the proposed catalytic cycle from Figure 1. All energies are calculated either with the PBE functional and with (light green) and without (dark green) D3 dispersion correction, or with the PBE0 hybrid functional again with (gray) and without (black) dispersion correction. Transition state energies, indicated by the double dagger, are connected to the energies of the minimum structures by solid lines, whereas the dashed lines connect minimum energies for which no transition state was calculated or already discussed previously.3b All structures were optimized with PBE/def2-TZVPP and the green numbers indicate the exact values for these relative energies. 12 In order to understand in more detail the mechanism and in particular the two crucial steps of 1,5-HAT and alkyliodine oxidation in particular, we carried out density functional theory calculations for the catalytic cycle from Figure 1. We optimized the structures with the PBE functional25 and the def2-TZVPP basis set26 in Turbomole27 and solvent effects were modeled with the COSMO continuum model28 employing a dielectric constant of e = 24.95, which is the average of the dielectric constants of MeCN and tBuOH. Transition state structures were optimized with the QST3 transition state search29 as implemented in Gaussian 09.30 The effect of exact exchange and dispersion interaction on the energies was investigated with the PBE0 hybrid functional31 and the empirical D3 dispersion correction,32 respectively. As in our previous work,3b we adopted a closed conformation for all structures but note that the relative energy barriers to the open structure are on the order of 2 kcal mol-1. Figure 2 displays the electronic energy profile of the proposed catalytic cycle. As expected, the photochemical initiation step, corresponding to the homolytic N-I bond cleavage,3b leads to radical structure B that is destabilized by about 35 kcal mol-1 compared to structure A. The intramolecular radical transfer to form the more stable benzylic radical C is almost barrierless. Structure D is then significantly stabilized by a recombination with an iodine radical in the next step. The crucial iodine oxidation with mCPBA has a barrier (TS D-E) of 4.68 (11.63) kcal mol-1 as calculated with the PBE (PBE0) functional. The corresponding transition state structure is displayed in Figure 3. It is important to note, however, that both transition states have several vibrations with very low wave numbers (four (twelve) vibrations with less than 50 cm-1 for TS B-C (TS D-E)). Therefore, the concept of a transition state is not strictly valid but we can confirm that both transformations correspond to very flat regions of the potential energy hyper-surface. Despite the prominence of mCPBA in iodine(I/III) catalysis,33 a more detailed picture on the course of the oxidation of organic iodine(I) compounds had not been generated previously. The detected low barrier now provides the rationale for the success of this particular oxidant. The direct reaction from D to 2a is endothermic, with energies ranging from 1.32 kcal mol-1 (PBE0) to 5.10 kcal mol-1 (PBE-D3), underlining the necessity of the iodine oxidation with mCPBA. It is interesting that for the alternative peroxide TBHP as oxidant (Table 1, entry 1) the corresponding transition state from D to E is significantly increased by almost 20 kcal/mol (see the Supporting Information), indicating that iodine(I) to iodine(III) oxidation is a kinetically less feasible process with this oxidant.9,34 These results again demonstrate that the involvement of an 13 iodine(III) intermediate is crucial in order to provide kinetically competent pyrrolidine formation. We further calculated the energy of a radical quenching reaction by a solvent molecule. For the most stable radical structure C, H-atom abstraction from an acetonitrile molecule has an enthalpic penalty of 12 kcal mol-1, whereas it is as high as 19 kcal mol-1 for H-atom abstraction from tBuOH (both values are obtained from PBE/def2-TZVPP calculations). Even if present, these reactions do not affect the overall yield because they lead to regeneration of the starting material 1a. However, they contribute an explanation to the comparably low quantum yield of the present radical chain.34 a b c TS D-E Ar O O H I Ar O O O Ar O O H I OH O I Ph Ph Ph NHTs NHTs NHTs Figure 3. Three structures obtained from displacements along the imaginary vibrational mode corresponding to the transition state structure TS D-E. The transition state is shown in the middle (b), whereas the structure on the left (a) is closer to the starting structures mCPBA and D and the 14 structure on the right (c) resembles the oxidized structure E. Hydrogen atoms were removed for clarity. The combined experimental and theoretical observations demonstrate that the development of catalytic iodine catalyzes for Hofmann-Löffler reactions require an accurate engineering of a number of different factors in order to arrive at suitably balanced rates for all involved individual steps. It includes the identification of tBuOH as a suitable co-solvent for stabilization of the electrophilic iodine catalyst in the form of tBuOI, mCPBA as a suitably rapid oxidant to generate the crucial alkyliodine(III) intermediate and acetonitrile as non-chlorinated solvent. As a consequence, the catalytic system based on mCPBA that is presented here is only the second catalytic reaction that is general for C-H amination of aliphatic positions including non-benzylic ones and it provides the advantage of employing only conventional bulk reagents and solvents. 3 Conclusion We have identified suitable conditions for the selective intramolecular amination of aliphatic positions, which include activated and non-activated aliphatic C-H bonds. The reaction proceeds with high selectivity and synthetically useful yields. The robust homogeneous catalyst is conveniently generated in situ from molecular iodine and tert-butanol and is continuously regenerated from 3-chloro perbenzoic acid as terminal oxidant, which are all commodity reagents. The reaction represents the currently most economic variant of an iodine-catalyzed Hofmann-Löffler reaction. It proceeds through an iodine(I/III) catalytic manifold, which is active in the presence of visible light irradiation. Although related to alternative catalysis with an iodine(III) reagent as terminal oxidant, the present system shows significant changes in the rate of the relative steps of the catalytic cycle and for some substrate classes can even provide alternative reactivity. The present work thus demonstrates that careful engineering of molecular iodine catalysis can provide an entry into versatile improvement of the economic conditions of homogeneous catalysis. 15 ASSOCIATED CONTENT Supporting Information. The Supporting Information is available free of charge on the ACS Publications website. The following files are available free of charge. Details on experimental procedures for the catalytic reactions, spectroscopic data for the products and xyz structure files for all molecules included in the calculations (PDF). AUTHOR INFORMATION Corresponding Author * E-mail for K.M.: kmuniz@iciq.es * E-mail for M.R.: markus.reiher@phys.chem.ethz.ch ORCID Thomas Duhamel: 0000-0003-3397-0639 Christopher J. Stein: 0000-0003-2050-4866 Markus Reiher: 0000-0002-9508-1565 Kilian Muñiz: 0000-0002-8109-1762 Author Contributions TD and CM conducted the experimental research and CJS and MR designed and conducted the theoretical investigation. The manuscript was written by KM after consultation with all authors. All authors have given approval to the final version of the manuscript. Notes The authors declare no competing financial interest. ACKNOWLEDGMENT We thank the Spanish Ministry for Economy and Competitiveness and FEDER (CTQ201456474R grant to K. M., Severo Ochoa Excellence Accreditation 2014-2018 to ICIQ, SEV-20130319), the region of Catalonia and the Schweizerischer Nationalfonds (No. 20020 169120 to M.R.) for financial support. 16 REFERENCES 1 (a) Finkbeiner, P.; Nachtsheim, B. Iodine in Modern Oxidation Catalysis. Synthesis 2013, 45, 979. (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. (c) Uyanik, M.; Ishihara, K. Catalysis with In Situ-Generated (Hypo)iodite Ions for Oxidative Coupling Reactions. ChemCatChem 2012, 4, 177. (d) Minakata, S. Utilization of N−X Bonds in The Synthesis of N-Heterocycles. Acc. Chem. Res. 2009, 42, 1172. (e) Wu, X.-F.; Gong, J.-L.; Qi, X. A powerful combination: recent achievements on using TBAI and TBHP as oxidation system. Org. Biomol. Chem. 2014, 12, 5807. (f) Li, J.; Lear, M. J.; Kawamoto, Y.; Umemiya, S.; Wong, A. R.; Kwon, E.; Sato, I.; Hayashi, Y. Oxidative Amidation of Nitroalkanes with Amine Nucleophiles using Molecular Oxygen and Iodine. Angew. Chem. Int. Ed. 2015, 54, 12986. 2 (a) Uyanik, M.; Okamoto, H.; Yasui, T.; Ishihara, K. Quaternary Ammonium (Hypo)iodite Catalysis for Enantioselective Oxidative Cycloetherification. Science 2010, 328, 1376. (b) Uyanik, M.; Suzuki, D.; Yasui, T.; Ishihara, K. In Situ Generated (Hypo)Iodite Catalysts for the Direct α-Oxyacylation of Carbonyl Compounds with Carboxylic Acids. Angew. Chem. Int. Ed. 2011, 50, 5331. (c) Uyanik, M.; Hayashi, H.; Ishihara, K. High-turnover hypoiodite catalysis for asymmetric synthesis of tocopherols. Science 2014, 345, 291. (d) Uyanik, M.; Hayashi, H.; Iwata, H.; Ishihara, K. Chiral Ammonium Hypoiodite Salt-catalyzed Enantioselective Oxidative Cycloetherification to 2-Acyl Tetrahydrofurans. Chem. Lett. 2016, 45, 353. (e) Uyanik, M.; Sasakura, N.; Kaneko, E.; Ohori, K.; Ishihara, K. Chiral Ammonium Hypoiodite-catalyzed Enantioselective Oxidative Dearomatization of 1-Naphthols Using Hydrogen Peroxide. Chem. Lett. 2015, 44, 179. (f) Uyanik, M.; Suzuki, D.; Watanabe, M.; Tanaka, H.; Furukawa, K.; Ishihara, K. High-performance Hypoiodite/Hydrogen Peroxide Catalytic System for the Oxylactonization of Aliphatic γ-Oxocarboxylic Acids. Chem. Lett. 2015, 44, 387. 3 (a) Martínez, C.; Muñiz, K. An Iodine-Catalyzed Hofmann-Löffler Reaction. Angew. Chem. Int. Ed. 2015, 54, 8287. (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. 4 (a) Wolff, M. E. Cyclization of N-Halogenated Amines (The Hofmann-Löffler Reaction). Chem. Rev. 1963, 63, 55. (b) Neale, R. S. Nitrogen Radicals as Synthesis Intermediates. NHalamide Rearrangements and Additions to Unsaturated Hydrocarbons. Synthesis 1971, 1. (c) 17 Stella, L. Homolytic Cyclizations of N-Chloroalkenylamines. Angew. Chem. Int. Ed. 1983, 22, 337. (d) Jeffrey, J. L.; Sarpong, R. Intramolecular C(sp3)–H amination. Chem. Sci. 2013, 4, 4092. 5 For recent developments: (a) Wappes, E. A.; Nakafuku, K. M.; Nagib, D. A. Directed β C–H Amination of Alcohols via Radical Relay Chaperones. J. Am. Chem. Soc. 2017, 139, 10204. (b) Long, J.; Cao, X.; Zhu, L.; Qiu, R.; Au, C. -T.; Yin, S. -F.; Iwasaki, T.; Kambe, N. Intramolecular, Site-Selective, Iodine-Mediated, Amination of Unactivated (sp3)C–H Bonds for the Synthesis of Indoline Derivatives. Org. Lett. 2017, 19, 2793. (c) Wappes, E. A.; Fosu, S. C.; Chopko, T. C.; Nagib, D. A. Triiodide-Mediated δ-Amination of Secondary C−H Bonds. Angew. Chem. Int. Ed. 2016, 55, 9974. (d) Paz, N. R.; Rodríguez-Sosa, D.; Valdés, H.; Marticorena, R.; Melián, D.; Copano, M. B.; González, C. C.; Herrera, A. J. Chemoselective Intramolecular Functionalization of Methyl Groups in Nonconstrained Molecules Promoted by NIodosulfonamides. Org. Lett. 2015, 17, 2370. (e) O’Broin, C. Q.; Fernández, P.; Martínez, C.; Muñiz, K. N-Iodosuccinimide-Promoted Hofmann–Löffler Reactions of Sulfonimides under Visible Light. Org. Lett. 2016, 18, 436. For the alternative photochemical cleavage of N-O bonds: (f) Dauncey, E. M.; Morcillo, S. P.; Douglas, J. J.; Sheikh, N. S; Leonori, D. Photoinduced Remote Functionalisations by Iminyl Radical Promoted C-C and C-H Bond Cleavage Cascades. Angew. Chem. Int. Ed. 2018, 57, 744. (g) Jiang, H.; Studer, A. α-AminoxyAcid-Auxiliary-Enabled Intermolecular Radical γ-C(sp3 )-H Functionalization of Ketones. Angew. Chem. Int. Ed. 2018, 57, 1692. 6 Classic work on iodine/iodine(III): (a) Hernández, R.; Rivera, A.; Salazar, J. A.; Suárez, E. Nitroamine radicals as intermediates in the functionalization of non-activated carbon atoms. Chem. Commun. 1980, 958. (b) Betancor, C.; Concepción, J. I.; Hernández, R.; Salazar, J. A.; Suárez, E. Intramolecular functionalization of nonactivated carbons by amidylphosphate radicals. Synthesis of 1,4-epimine compounds. J. Org. Chem. 1983, 48, 4430. (c) 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. (d) 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. I 1987, 937. (e) De Armas, P.; Francisco, C. G.; Hernández, R.; Salazar, J. A.; Suárez, E. Steroidal N-nitroamines. Part 4. 18 Intramolecular functionalization of N-nitroamine radicals: synthesis of 1,4-nitroimine compounds. J. Chem. Soc., Perkin Trans. I 1988, 3255. (f) 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. 7 Muñiz, K.; García, B.; Martínez, C.; Picinelli, A. Dioxoiodane Compounds as Versatile Sources for Iodine(I) Chemistry. Chem. Eur. J. 2017, 23, 1539. 8 (a) Kirschning, A.; Monenschein, H.; Schmeck, C. Stabiles festphasengebundenes Iodazid. Angew. Chem. 1999, 111, 2720. (b) Kirschning, A.; Hashem, A.; Monenschein, H.; Rose, L.; Schöning, K.-U. Preparation of Novel Haloazide Equivalents by Iodine(III)-Promoted Oxidation of Halide Anions. J. Org. Chem. 1999, 64, 6522. (c) Domann, S.; Sourkouni-Argirusi, G.; Merayo, N.; Schönberger, A.; Kirschning, A. Cohalogenation of Allyl and Vinylsilanes using Polymer-bound Haloate(l)-Reagents. Molecules 2001, 6, 127. 9 See Supporting Information for details. 10 Hager, A.; Vrielink, N.; Hager, D.; Lefranc, J.; Trauner, D. Synthetic approaches towards alkaloids bearing α-tertiary amines. Nat. Prod. Synth. 2016, 33, 491. 11 Reyes, E.; Uria, U.; Carrillo, L.; Vicario, J. L. Transannular reactions in asymmetric total synthesis. Tetrahedron 2014, 70, 9461. 12 For reviews, see (a) Mayer, J. M. Understanding Hydrogen Atom Transfer: From Bond Strengths to Marcus Theory. Acc. Chem. Res. 2011, 44, 36. (b) Liu, W.; Huang, X.; Cheng, M.J.; Nielsen, R. J.; Goddard III, W. A.; Groves, J. T. Oxidative Aliphatic C-H Fluorination with Fluoride Ion Catalyzed by a Manganese Porphyrin. Science 2012, 337, 1322. (c) Majetich, G.; Wheless, K. Remote intramolecular free radical functionalizations: An update. Tetrahedron 1995, 51, 7095. (d) Cekovic, Z. Reaction of carbon radicals generated by 1,5-transposition of reactive centers. J. Serb. Chem. Soc. 2005, 70 , 287. (e) Feray, L.; Kuznetzov, N.; Renaud P. in Radicals in Organic Synthesis Vol. 2 (Renaud, P.; Sibi, M. P., Eds.), Wiley-VCH, Weinheim, 2001, pp. 246 –278. (f) Robertson, J.; Pillai, J.; Lush, R. K. Radical translocation reactions in synthesis. Chem. Soc. Rev. 2001, 30, 94. (g) Sperry, J.; Liu, Y.-C.; Brimble, M. A. Synthesis of natural products containing spiroketals via intramolecular hydrogen abstraction. Org. Biomol. Chem. 2010, 8, 29. (h) Gansäuer, A.; Lauterbach, T.; Narayan, S. Strained heterocycles in radical chemistry. Angew. Chem. Int. Ed. 2003, 42, 5556. (i) Dénés, F.; Beaufils, F.; Renaud, P. Preparation of Five-Membered Rings via the Translocation-Cyclization of Vinyl Radicals. 19 Synlett 2008, 2389. (j) Lathbury, D. C.; Parsons, P. J.; Pinto, I. A route to the pyrrolizidine ring system using a novel radical cyclisation. J. Chem. Soc. Chem. Commun. 1988, 81. (k) Renaud, P.; Beaufils, F.; Dénés, F.; Feray, L.; Imboden, C.; Kuznetsov, N. Stereoselective Radical Translocations. Chimia 2008, 62, 510. (l) Chiba, S.; Chen, H. sp3 C–H oxidation by remote Hradical shift with oxygen- and nitrogen-radicals: a recent update. Org. Biomol. Chem. 2014, 12, 4051. 13 Recent examples: (a) Chen, D. F.; Chu, J. C. K.; Rovis, T. Directed γ-C(sp3)–H Alkylation of Carboxylic Acid Derivatives through Visible Light Photoredox Catalysis. J. Am. Chem. Soc. 2017, 139, 14897. (b) Chu, J. C. K.; Rovis, T. Amide-directed photoredox-catalysed C–C bond formation at unactivated sp3 C–H bonds. Nature 2016, 539, 272. (c) Choi, G. J.; Zhu, Q.; Miller, D. C.; Gu, C. J.; Knowles, R. R. Catalytic alkylation of remote C–H bonds enabled by protoncoupled electron transfer. Nature 2016, 539, 268. (d) Hu, X.-Q.; Chen, J.-R.; Xiao, W.-J. Controllable Remote C-H Bond Functionalization by Visible-Light Photocatalysis. Angew. Chem. Int. Ed. 2017, 56, 1960. 14 Nechab, M.; Mondal, S.; Bertrand, M. P. Amide-Directed Photoredox Catalyzed C-C Bond Formation at Unactivated sp3 C-H Bonds. Chem. Eur. J. 2014, 20, 16034. 15 (a) Short, M. A.; Blackburn, J. M.; Roizen, J. L. Sulfamate Esters Guide Selective RadicalMediated Chlorination of Aliphatic C-H Bonds. Angew. Chem. Int. Ed. 2018, 57, 296. (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. 16 For an alternative approach: Zhang, H.; Muñiz, K. Selective Piperidine Synthesis Exploiting Iodine-Catalyzed Csp3–H Amination under Visible Light. ACS Catal. 2017, 7, 4122. 17 (a) Barluenga, J.; González-Bobes, F.; Ananthoju, S. R.; García-Martín, M. A.; González, J. M. Oxidative Opening of Cycloalkanols: An Efficient Entry to ω-Iodocarbonyl Compounds. Angew. Chem. Int. Ed. 2001, 40, 3389. (b) Barluenga, J.; González-Bobes, F.; Murguía, M. C.; Ananthoju, S. R.; González, J. M. Bis(pyridine)iodonium Tetrafluoroborate (IPy2BF4): A Versatile Oxidizing Reagent. Chem. Eur. J. 2004, 10, 4206. (c) Suárez-Pantiga, S.; González, J. M. Electrophilic activation of unsaturated systems: Applications to selective organic synthesis. Pure Appl. Chem. 2013, 85, 721. (d) Bieti, M.; Lanzalunga, O.; Salamone, M. Involvement of alkoxyl radical intermediates in the photolysis of 1-alkylcycloalkanols in the presence of 20 bis(pyridine)iodonium tetrafluoroborate: Comparison with the (diacetoxyiodo)benzene/I2 system. J. Photochem Photobiol. A: Chem. 2006, 182, 33. 18 (a) Montoro, R.; Wirth, T. Direct Iodination of Alkanes. Org. Lett. 2003, 5, 4729. (b) Barton, D. H. R.; Beckwith, A. L. J.; Goosen, A. Photochemical transformations. Part XVI. A novel synthesis of lactones. J. Chem. Soc. 1965, 181. 19 Amidyl radicals: (a) Kärkäs, M. D. Photochemical Generation of Nitrogen-Centered Amidyl, Hydrazonyl, and Imidyl Radicals: Methodology Developments and Catalytic Applications. ACS Catal. 2017, 7, 4999. (b) Xiong, T.; Zhang, Q. New amination strategies based on nitrogen-centered radical chemistry. Chem. Soc. Rev. 2016, 45, 3069. (c) Chen, J.-R.; Hu, X.­Q.; Lu, L.-Q.; Xiao, W.-J. Visible light photoredox-controlled reactions of N-radicals and radical ions. Chem. Soc. Rev. 2016, 45, 2044. (d) Nguyen, L. Q.; Knowles, R. R. Catalytic C–N Bond-Forming Reactions Enabled by Proton-Coupled Electron Transfer Activation of Amide N–H Bonds. ACS Catal. 2016, 6, 2894. 20 For a discussion of the nature of amides and their amidoyl radicals in Hofmann-Löffler reactions, see: Šakić, D.; Zipse, H. Radical Stability as a Guideline in C–H Amination Reactions. Adv. Synth. Catal. 2016, 358, 3983. 21 Okuyama, T.; Takino, T.; Sueda, T.; Ochiai, M. Solvolysis of Cyclohexenyliodonium Salt, a New Precursor for the Vinyl Cation: Remarkable Nucleofugality of the Phenyliodonio Group and Evidence for Internal Return from an Intimate Ion-Molecule Pair. J. Am. Chem. Soc. 1995, 117, 3360. 22 (a) Guo, W.; Vallcorba, O.; Vallribera, A.; Shafir, A.; Pleixats, R.; Rius, J. Oxidative Breakdown of Iodoalkanes to Catalytically Active Iodine Species: A Case Study in the αTosyloxylation of Ketones. ChemCatChem 2014, 6, 468. (b) Davidson, R. I.; Kropp, P. J. Oxidatively assisted nucleophilic substitution/elimination of alkyl iodides in alcoholic media. A further study. J. Org. Chem. 1982, 47, 1904. (c) McCabe, P. H.; de Jenga, C. I.; Stewart, A. Oxidative cleavages of aliphatic and cyclic iodides. Tetrahedron Lett. 1981, 22, 3679. (d) Macdonald, T. L.; Narasimhan, N.; Burka, L. T. Crystal lattice control of unimolecular photorearrangements. Differences in cyclohexenone photochemistry in solution and the solid state. Solid-state results. J. Am. Chem. Soc, 1980, 102, 1160. (e) Cambie, R. C.; Chambers, D.; Lindsay, B. G.; Rutledge, P. S.; Woodgate, P. D. Oxidative displacement of hypervalent iodine from alkyl iodides. J. Chem. Soc. Perkin Trans. 1 1980, 822. (f) Reich, H. J.; Peake, S. L. 21 Hypervalent organoiodine chemistry. Syn elimination of alkyl iodoso compounds. J. Am. Chem. Soc. 1978, 100, 4888. (g) Cambie, R. C.; Noall, W. I.; Potter, G. J.; Rutledge, P. S.; Woodgate, P. D. Reactions of alkenes with electrophilic iodine in tetramethylene sulphone-chloroform. J. Chem. Soc., Perkin Trans. 1. 1977, 226. 23 The assumption of a successful pyrrolidine synthesis within a nucleophilic substitution of an alkyl iodide intermediate with tosylamide has recently been proposed: Leger, P. R.; Murphy, R. A.; Pushkarskaya, E.; Sarpong, R. Synthetic Efforts toward the Lycopodium Alkaloids Inspires a Hydrogen Iodide Mediated Method for the Hydroamination and Hydroetherification of Olefins. Chem. Eur. J. 2015, 21, 4377. 24 For a comparison on leaving group capacity of related selenium(IV) derivatives: Andreou, T.; Burés, J.; Vilarrasa, J. Reaction of Dess–Martin periodinane with 2-(alkylselenyl)pyridines. Dehydration of primary alcohols under extraordinarily mild conditions. Tetrahedron Lett. 2010, 51, 1863. 25 Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. 26 (a) Ahlrichs, R.; May, K. Contracted all-electron Gaussian basis sets for atoms Rb to Xe. Phys. Chem. Chem. Phys., 2000, 2, 943. (b) Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297. (c) Weigend, F.; Häser, M.;, Patzelt, H.; Ahlrichs, R. RI-MP2: optimized auxiliary basis sets and demonstration of efficiency. Chem. Phys. Lett. 1998, 294, 143. 27 Ahlrichs, R.; Bär, M.; Häser, M.; Horn, H.; Kölmel, C. Electronic structure calculations on workstation computers: The program system Turbomole. Chem. Phys. Lett. 1989, 162, 165. 28 Klamt, A; Schüürmann, G. COSMO: a new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradient. J. Chem. Soc. Perkin Trans. 1993, 2, 799. 29 Peng, C.; Avala, Y.; Schlegel, H. B.; Frisch, M. J. Using redundant internal coordinates to optimize equilibrium geometries and transition states. J. Comp. Chem. 1996, 17, 49. 30 Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, 22 M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr, J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09 Revision D.01. 2013; Gaussian Inc. Wallingford CT (Gaussian 09 Revision D.01 (program package)). 31 Perdew, J. P.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982. 32 (a) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (b) Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456. 33 (a) Richardson, R. D.; Wirth, T. Hypervalent iodine goes catalytic. Angew. Chem. Int. Ed. 2006, 45, 4402. (b) Ochiai, M.; Miyamoto, K. Catalytic Version of and Reuse in Hypervalent Organo-λ3- and -λ5-iodane Oxidation. Eur. J. Org. Chem. 2008, 4229. (c) Ochiai, M. Stoichiometric and catalytic oxidations with hypervalent organo-λ3-iodanes. Chem. Rec. 2007, 7, 12. (d) Dohi, T.; Kita, Y. Hypervalent iodine reagents as a new entrance to organocatalysts. Chem. Commun. 2009, 2073. (e) Uyanik, M.; Ishihara, K. Hypervalent iodine-mediated oxidation of alcohols. Chem. Commun. 2009, 2086. 34 Initial guesses for the transition state structures were obtained from a spline-based automated transition state search: Vaucher, A. C.; Reiher, M. Minimum energy paths and transition states by curve optimizations. arXiv:1802.05669. 35 (a) Yan, M.; Lo, J. C.; Edwards, J. T.; Baran, P. S. Radicals: Reactive Intermediates with Translational Potential. J. Am. Chem. Soc. 2016, 138, 12692. (b) Studer, A.; Curran, D. P. Catalysis of Radical Reactions: A Radical Chemistry Perspective. Angew. Chem. Int. Ed. 2016, 55, 58. 23 Synopsis Ar O NHSO2R' R I [mCPBA] O O H I Ph R N SO2R' NHTs Oxidation to Iodine(III) Catalyst 24