Skip to main content
Skip to main content
Chem Rev. 2020 Sep 9; 120(17): 9790–9833.
Published online 2020 Aug 6. doi: 10.1021/acs.chemrev.0c00278
PMCID: PMC8009483
PMID: 32786419

Generation of Alkyl Radicals: From the Tyranny of Tin to the Photon Democracy

Abstract

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0120.jpg

Alkyl radicals are key intermediates in organic synthesis. Their classic generation from alkyl halides has a severe drawback due to the employment of toxic tin hydrides to the point that “flight from the tyranny of tin” in radical processes was considered for a long time an unavoidable issue. This review summarizes the main alternative approaches for the generation of unstabilized alkyl radicals, using photons as traceless promoters. The recent development in photochemical and photocatalyzed processes enabled the discovery of a plethora of new alkyl radical precursors, opening the world of radical chemistry to a broader community, thus allowing a new era of photon democracy.

1. Introduction

Among all the open-shell species, carbon-centered radicals are intriguing neutral intermediates that find extensive use in organic synthesis, despite the initial distrust about their possible application.15 In particular, the generation of unstabilized alkyl radicals under mild conditions granting the controlled and selective outcome of the ensuing reactions has been a challenge for many years. The first and more obvious way to form such species is the homolytic cleavage of a labile C–X bond; alkyl halides appeared as the ideal choice in this respect. The real breakthrough in radical chemistry was the discovery of Bu3SnH to promote radical chain reactions as reported about 60 years ago in the reduction of bromocyclohexane.6 Reduction of an organotin halide by lithium aluminum hydride formed the reactive tin hydride in solution. In subsequent modifications of the protocol, both sodium borohydride7 and sodium cyanoborohydride8 acted as effective reducing agents. Alkyl radicals generated via tin chemistry were then used for C–C bond formation mainly via the addition to (electron-poor) olefins, the well-known Giese reaction,912 an evolution of the original process which made use of organomercury compounds.12,13

As illustrated in Figure Figure11a, tributyltin hydride has the double role of allowing the formation of Bu3Sn as the radical chain carrier and as a hydrogen donor to close the catalytic cycle. The unique features of this catalytic cycle are attributed to the forging of stronger Sn–X and C–H bonds at the expense of the cleavage of the more labile Sn–H and C–X ones. A more quantitative aspect of this reaction can be appreciated comparing the different bond dissociation energies (BDE) associated with the steps mentioned above (see Figure Figure11a).14,15 More recent applications showcase the crucial role of tin intermediates in controlling the outcome of different reactions. Sn–O interactions direct the regioselective addition of the radical in the radical stannylation of the triple bond in propargyloxy derivatives,16 whereas tin radicals induced the synthesis of stannylated polyarenes via double radical peri-annulations, increasing the solubility of the products.17

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0001.jpg

(A) Thermal generation of radicals from alkyl halides in the Giese reaction. (B) LD50 values for selected organotin compounds. (C) Thermal generation of radicals from alcohols via xanthates (I). (D) Thermal and photochemical generation of radicals from carboxylic acids via Barton esters (II).

The performance of Bu3SnH was so competitive9,1820 that more than 20 years ago it was claimed that it was improbable to have “flight from the tyranny of tin” in radical processes,21,22 a hard statement that subtly introduces the problem of the substantial toxicity and high biological activity of triorganotin compounds.23 The LD50 of 0.7 mmol/kg in murine species (see Figure Figure11b) combined with the long half-lives in aquatic environment represent the biggest concerns for the application of these otherwise extremely versatile species, especially in the absence of viable alternatives.24 Indeed, O-thiocarbonyl derivatives like xanthates (I, obtained from alcohols) were considered an alternative to the alkyl halides, albeit the radical generation required in most cases the use of tin hydrides (Figure Figure11c).3,22,25,26

Efforts in substituting toxic tin derivatives with other hydrides such as (TMS)3SiH27 or lauroyl peroxides and xanthates met some success, however, only in limited cases.2830 Other initiators to promote tin-free radical chain reactions were organoboranes,31 thiols,32 P–H-based reagents,32 and 1-functionalized cyclohexa-2,5-dienes,3234 but nowadays they are not commonly used in synthetic planning.

The use of metal oxidants (MnIII acetate)35 or metal reductants (TiIII catalyst36 or SmII iodide37) were sparsely used, but only in the latter case unstabilized alkyl radicals were formed from alkyl iodides.

The introduction of the Barton esters II in 1985 represented a step forward in solving the conundrum of the tyranny of tin: the conversion of the strong O–H bond of an acid into the (photo)labile O–N bond of the corresponding thiohydroxamate ester (Figure Figure11d).38,39 Barton esters have the advantage of being slightly colored, allowing the use of visible light irradiation to induce the cleavage of the O–N bond. The last point is significant, demonstrating the formation of alkyl radicals in a very mild way under tin-free conditions with no need of further additives, albeit Barton esters have currently a limited application. On this occasion, photochemistry showed an attractive potential for the development of novel synthetic strategies based on radical chemistry. However, Barton esters remained for several years an isolated niche. In most cases, the photochemical generation of radicals required harmful UV radiation and dedicated equipment.40 Since the milestone represented by the development of the chemistry of Barton esters, new photochemical ways were sought toward more efficient ways to generate radicals. The photon appears to be the ideal component for a chemical reaction, assuming the form of a traceless reagent, catalyst, or promoter that leaves no toxic residues in the final mixture.4144 The breakthrough that would allow moving forward from the “tyranny of tin” to a greener “photon democracy” can be associated with the use of solar or visible photons, freely available from the sun that shines throughout the scientific world. The renaissance of the photocatalyzed processes that we have witnessed in the last years represents a significant step toward this direction.4558

The multifaceted use of photoredox catalysis and photocatalyzed hydrogen transfer reactions expanded the range of possible radical precursors and unconventional routes for the generation of several carbon (or heteroatom based) radicals, including the challenging formation of unstabilized alkyl radicals.5965 Consequently, in this review, we aim to present a summary of the novel ways to generate alkyl radicals by photochemical means that, in the last years66 have revolutionized the way to carry out radical chemistry. This work will focus exclusively on the reactions promoting the formation of unstabilized alkyl radicals, and not the stabilized ones, e.g., α-oxy, α-amino, benzylic, or allylic.

Figure Figure22 collects the main paradigmatic approaches to the photogeneration of alkyl radicals (either photochemically or photocatalyzed). The more classical, although the less employed, path to generate alkyl radicals consists of the introduction of a photoauxiliary group which renders a bond labile to a direct photochemical cleavage (Figure Figure22A).67 The Barton esters are the archetypal moiety belonging to this class.39 A conspicuous body of literature have been focusing on the development of suitable alkyl substituents able to facilitate redox reactions making the derivatives more oxidizable or reducible. The strategy that is followed in Figure Figure22B consists in the conversion of a common functional group (e.g., OH or COOH), which in most cases is tethered to the alkyl group, into a different electroauxiliary group68 (Figure Figure22B). As a result, the interaction of the activated species with an excited photoredox catalyst (PCSET) able to induce a single electron transfer (SET) process generates the corresponding radical ions, either by an oxidative pathway or a reductive pathway. The desired alkyl radical is then formed by fragmentation of these radical ion intermediates. The oxidative pathway is efficient when the radical precursor is negatively charged (see further Figure Figure33) as in the case of alkyl carboxylates and alkyl sulfinates causing the CO2 or SO2 loss, respectively, despite the fact that the exothermicity of the process is verified only in the C–C cleavage rather than the C–S cleavage.69 On the contrary, positively charged Katritzky salts were ideal candidates for the releasing of radicals via the reductive pathway (Figure Figure33).70

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0002.jpg

Different approaches for the photogeneration of alkyl radicals (A) by photochemical means through the introduction of a photoauxiliary group (B) via fragmentation of a radical cation (oxidative pathway) or anion (reductive pathway) formed by photoredox catalysis (C) via a halogen atom transfer reaction (XAT) with a photogenerated radical (D) through the photocatalyzed cleavage of a C–H bond via direct (d-HAT) or indirect (i-HAT) hydrogen atom transfer (E) by the remote-controlled C–H activation via a photogenerated heteroatom based radical (F) by a ring-opening via a photogenerated heteroatom (nitrogen) based radical.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0003.jpg

On the left, substrates used to promote the photochemical formation of alkyl radicals divided according to the C–Y bond cleaved. The oxidation potentials (Eox, in orange) or the reduction potentials (Ered, in green) of the precursors as well as the BDE values of the bond that is broken (highlighted in gray) by direct photocleavage are reported. On the right, a selection of common photoredox catalysts with their main redox features are collected.

A viable alternative is the photogeneration (often from a photoredox process) of a reactive radical on a heteroatom like a silyl radical, which can exploit a halogen atom transfer reaction to afford an alkyl radical through the smooth Si-X bond formation (XAT, Figure Figure22C).71,72 This strategy provides an elegant way to overcome the Giese conditions in the tin mediated activation of alkyl halides. Recently, an α-amino radical was used in the same strategy, promoting the formation of alkyl radicals via C–X bond cleavage.73

A more challenging approach requires the photocatalyzed selective cleavage of a strong alkyl-H bond, via a direct hydrogen atom transfer reaction (d-HAT, Figure Figure22D) operated by an excited photocatalyst (PCHAT).72,7476 The indirect version of the latter path exploits the photogeneration of a stable heteroatom based radical (i-HAT, Figure Figure22D) that will become the competent intermediate in the abstraction of the H atom from the alkyl moiety.76 An indirect HAT (i-HAT) may also take place by intramolecular hydrogen transfer thus releasing an alkyl radical (Figure Figure22E).7681 As an alternative, the photochemical radical generation may induce a ring-opening in strained structures like cyclobutanes, to form a substituted alkyl radical (Figure Figure22F).82

Figure Figure33 showcases a collection of the main alkyl radical precursors devised for the generation under photochemical conditions of unstabilized alkyl radicals. In this figure, the radical precursors were collected depending on the C(sp3)-Y bond cleaved during the radical release. As apparent, the photochemically triggered cleavage of several C-heteroatom bonds like C–X,71,8388 C–O,8998 C–B,99102 C–S,103106 or C–N70 (Figure Figure33) affords carbon-centered radicals. The alkyl radical generation is granted by the very versatile photochemical tool. This feature includes particular cases such as C–Se (in alkyl selenides),107,108 C–Te [in (aryltelluro)formates,109,110 for a previous thermal generation of alkyl radical from diorganyl tellurides, see ref (111)], and C–Si (in tetra alkyl silanes and bis-chatecolates)112114 to be added to C–Sn (in alkyl stannanes).112,115 Interestingly, even the more resilient C–H15,76 or C–C38,83,94,116130 bonds may be cleaved for alkyl radical generation, opening up new exciting possibilities for the synthetic (photo)chemist (Figure Figure33).

For the clarity of the reader, each radical precursor is accompanied by its oxidation potential (EOX, in orange) or its reduction potential (ERED, in green) to guide the feasibility on the generation of the radical via the oxidative or reductive pathway (type B, Figure Figure22B), respectively. Since the redox potentials may vary with the nature of the alkyl group, the values reported are referred to known structures. In alternative, the BDE values of the bond that is broken by direct photocleavage (type A, Figure Figure22A) or by photocatalyzed hydrogen abstraction (type D, Figure Figure22D) are reported. Figure Figure33 (right part) likewise collects the redox properties of commonly used photoredox catalysts including metal-free photoorganocatalysts (POC) to be used in the oxidative/reductive pathways.131137

The reactions collected and commented on in this review are primarily divided according to the type of the bond formed, namely the forging of C–C or C–heteroatom bonds, along with the construction of rings of different sizes. When possible, in each section, we will further categorize the reactions depending on the mechanism of the radical generation, ascribing them to the six types (A–F) described in Figure Figure22.

2. Formation of a C(sp3)-C Bond

Photochemically generated alkyl radicals have been employed to forge C(sp3)-C(spn) bonds (n = 1–3) in an intermolecular fashion following different strategies. In most cases, a conjugate addition onto a Michael acceptor or a Minisci-like reaction occurred, albeit alkenylations, acylations, or oxyalkylations are likewise used.

2.1. Formation of a C(sp3)-C(sp3) Bond

2.1.1. Addition to C–C Double Bonds: Hydroalkylations

Many reactions belonging to this class involve the nucleophilic alkyl radical addition onto an electrophilic Michael acceptor, resulting in a formal hydroalkylation of the double bond viz. the incorporation of an alkyl group (in position β with respect to the EWG group in the starting unsaturated compound) and a hydrogen atom (in position α). This reaction is usually one of the first that many authors would test during the discovery process of a new radical precursor, as testified by the plethora of reagents that are used in this transformation. Photoredox catalysis is by far the preferred approach here, especially by using the oxidation of a negatively charged precursor (oxidative pathway in Figure Figure22B).

A typical example is the oxidation of carboxylates138,139 that releases an alkyl radical via CO2 loss from the carboxyl radical intermediate (Scheme 1). Adamantylation of both acrylonitrile (Scheme 1a)140 and dimethyl 2-ethylidenemalonate starting from adamantane carboxylic acid 1-1 (Scheme 1b)141 were carried out following this strategy. In the former case, the authors employed 1,4-dicyanonaphthalene (DCN) as the POC under UV light irradiation, while visible light and an IrIII complex in the latter case. The approach used in Scheme 1b was also useful for the three steps preparation of the medicinal agent (±)-pregabalin.141 Also, the Fukuzumi catalyst (9-mesitylene-10-methylacridinium perchlorate, [Acr+Mes]ClO4) can promote this Giese-type reaction,142 allowing the alkylation of α-aryl ethenylphosphonates for the synthesis of fosmidomycin analogues.143

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0005.jpg
Different Strategies for the Decarboxylative Adamantylation of Electron-Poor Alkenes

A variation of this procedure is the decarboxylative-decarbonylative process occurring on an α-keto acid 2-1 under sunlight-driven photoredox catalyzed reaction conditions (Scheme 2).125

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0006.jpg
Decarboxylative-Decarbonylation of an α-Keto Acid

Oxalates are another class of electron-donors having two carboxylate moieties that can be lost upon photocatalyzed oxidation. These species may be introduced in situ by reaction of the alcohol with oxalyl chloride. The process induced the cleavage of a C–O bond, and the resulting radical could be trapped by butenolide 31 to form the menthyl derivative 32 used for the enantioselective preparation of cheloviolene A (33, Scheme 3).144 An IrIII-based photocatalyst efficiently promoted the reaction also in this case, allowing the synthesis of quaternary centers89 and the total synthesis of trans-clerodane diterpenoids.145

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0007.jpg
Enantioselective Preparation of Cheloviolene A

Alkyl trifluoroborates stand out as another important class of easily oxidizable moieties.146 The photocatalyzed oxidation of these salts (e.g., 41) causes the smooth release of BF3 and the formation of the reactive alkyl radical. Such a reaction was employed to functionalize Michael acceptors under sunlight irradiation (Scheme 4a) exploiting Acr+Mes as the POC.147 Complexation of 44 by a chiral rhodium complex (Λ-RhS, Scheme 4b) delivered 45 in good yields with 97% ee.148

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0008.jpg
Visible and Solar Light Photocatalyzed Functionalization of Michael Acceptors with Alkyl Trifluoroborate Salts

This synthetic strategy can be extended to neutral boronic acids or esters, upon in situ activation by a Lewis base (LB). The so formed negatively charged species is consequently more prone to oxidation, which eventually will provide the formation of the alkyl radicals. A typical example is illustrated in Scheme 5 where the boronic acid 51 was activated by 4-dimethylaminopyridine (DMAP) and then oxidized by an IrIII complex. The resulting cyclobutyl radical was trapped by methyl vinyl ketone to access the substituted ketone 52 in a good yield.100 This reaction was later scaled up under flow conditions by using the Photosyn reactor. In such a way, the authors could synthesize gram amounts per hour of the analogues of some drugs belonging to the GABA family.149

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0009.jpg
Activation of Boronic Acids with a Lewis Base

Following the examples of the carboxylate derivatives, the electron-donating species may be generated in situ by deprotonation, as in the case of sulfonamides, employed in the desulfurative conjugate addition of alkyl radicals onto Michael acceptors (Scheme 6). Again, the process is based on a photocatalyzed oxidation pathway. The starting sulfonamide (61) was first deprotonated by a mild base (K2HPO4), and the resulting anion was easily oxidized to a N-centered radical. Loss of N-sulfinylbenzamide generates the desired radical that gave the adduct 63 upon reaction with 62 in 75% yield.103

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0010.jpg
Desulfurative Strategy for the Conjugate Addition of Alkyl Radicals onto Michael Acceptors

In some instances, the radical precursor is a neutral compound. This situation is possible only when the derivative contains a highly oxidizable or reducible moiety. 4-Alkyl-1,4-dihydropyridines (alkyl-DHPs) under PC-free conditions act as radical precursors when combined with photoexcited iminium ion catalysis (Scheme 7). Here, enal 71 formed a chiral iminium ion 74+ by reaction with amine 73. Cation 74+ upon visible light excitation oxidized the alkyl-DHP 72 that in turn released the radical 75 upon fragmentation, along with radical 74. Radical recombination followed by hydrolysis gave the desired alkylated dihydrocinnamaldehyde 76 in a satisfactory yield with a good enantiomeric excess (Scheme 7).150 A similar Giese reaction was later proposed, where the alkyl-DHP was excited and a SET reaction with Ni(bpy)32+, acting as an electron mediator, took place. The alkyl radical derived from the radical cation of alkyl-DHP readily attacked a series of Michael acceptors.151

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0011.jpg
Alkyl-DHPs as Radical Precursors in Combination with Iminium Catalysis

Looking at the other edge of the redox spectrum, easily reducible compounds were devised as radical precursors via a photocatalyzed process. As an example, the incorporation of a N-phthalimidoyl moiety in an organic compound helps its photocatalyzed reduction, ultimately leading to the release of the alkyl radical. A typical case is represented by N-(acyloxy)phthalimides.126 A stereoselective variant of this reaction was applied to the synthesis of (−)-solidagolactone (84, Scheme 8). Thus, the photocatalyzed reduction of phthalimide 81 by a RuII complex released a tertiary carbon radical. Attack to the terminal carbon of the unsaturated core present in β-vinylbutenolide 82 yields 83 with a very high diastereomeric excess. Further elaboration of compound 83 gave 84 in a single step.152 This reaction emerges as a very interesting tool to construct quaternary carbons153 and to synthesize biologically active derivatives, e.g., (−)-aplyviolene.154

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0012.jpg
Synthesis of (−)-Solidagolactone via N-(Acyloxy)phthalimides

Interesting results were also obtained using N-phthalimidoyl oxalates such as 91 in place of the N-(acyloxy)phthalimides for the generation of alkyl radicals starting from tertiary alcohols (Scheme 9).92,97 The similarity of this reaction to the one presented in Scheme 8 is striking, despite a less atom economical radical generation.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0013.jpg
Generation of Alkyl Radicals from N-Phthalimidoyl Oxalates

Reduction of an organic compound may be carried out even on organic iodides by using cyanoborohydride anion as the reducing agent. The reaction is chemoselective, since no alkyl bromides or chlorides could be activated following this way. Giese adducts were formed by irradiation with a Xe lamp of the reaction mixture in good yields as illustrated by the formation of 102 from 101 in Scheme 10.155 This is another interesting example to circumvent the use of tin hydrides in the activation of alkyl halides.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0014.jpg
Giese-Type Reaction of Iodides in the Presence of BH3CN

Alkyl chlorides can be activated using Ir(dtbby)(ppy)2PF6 in the presence of micelles. The micellar environment stabilizes the photogenerated [Ir(dtbby)•–(ppy)2] species (−1.51 V vs SCE), unable to directly reduce the alkyl chlorides (ca. −2.8 V vs SCE). A second excitation of this long-lived intermediate allows the electron transfer to the halide, which could react with different electron-poor olefins, forging a novel C–C bond. The micellar system allowed intramolecular cyclizations to form five-membered cycles.156

Reduction of the alkyl halide 111 could be avoided applying a halogen transfer reaction. In fact, due to the strong BDE of the Si-halogen bond, an alkyl radical is formed thanks to the action of a purposely generated silyl radical (from (Me3Si)3SiH, TTMSS) by a photoredox catalytic step. Radical addition onto an unsaturated amide (112) gave the 1,8-difunctionalized derivative 113, a key compound in the preparation of Vorinostat 114, a histone deacetylase (HDAC) inhibitor active against HIV and cancer (Scheme 11).157 This is the typical case where the radical is formed by the cleavage of an Alk-Br bond without the assistance of tin derivatives. It is interesting that the reaction requires a substoichiometric amount of silane to proceed. Indeed, with higher loadings the product yield decreases, possibly due to the presence of competing nonproductive pathways. A chain reaction mechanism could be envisaged; however, the quantum yield for this reaction (Φ = 0.45) does not fully clarify the mechanistic details of the transformation.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0015.jpg
Synthesis of Vorinostat via XAT Strategy

In many instances the formation of the alkyl radical arose from a direct or indirect photocatalyzed C–H homolytic cleavage. The excited state of the decatungstate anion in its tetrabutylammonium salt form (TBADT) promoted in several cases the direct chemoselective cleavage of a C–H bond.75,158Scheme 12 depicts two examples involving the hydroalkylation of acrylonitrile. Unsubstituted cycloalkanes were suitable hydrogen donors under flow conditions (yielding 121Scheme 12a).159 Moreover, the chemoselective cleavage of the methine hydrogen in isovaleronitrile allowed the preparation of dinitrile 122 in 73% yield (Scheme 12b).160

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0016.jpg
TBADT-Photocatalyzed Hydroalkylation of Acrylonitrile

Similarly, the presence of a tertiary hydrogen was the driving force of the chemoselective TBADT-photocatalyzed C–H cleavage in several derivatives, as depicted in Scheme 13. As an example, alkylpyridine 131 was selectively functionalized and gave derivative 132 as the exclusive product in the reaction with a vinyl sulfone (Scheme 13a).161 Interestingly, the labile benzylic hydrogens present in 131 remained untouched under these reaction conditions. Noteworthy, steric and polar effects cooperatively operated in the derivatization of lactone 133. As a result only the methine hydrogen of the isopropyl group was selectively abstracted and afforded 134 in very high yields by reaction with fumaronitrile (Scheme 13b), albeit the seven different types of hydrogens present in 133.162 The C–H cleavage may also take place in branched alkanes as witnessed by the derivatization of 135 to form the succinate derivative 136 (Scheme 13c).163

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0017.jpg
TBADT-Driven Functionalization of Tertiary Carbons

In rare instances, the hydroalkylation reaction may be applied to olefins different to the usual Michael acceptors. Thus, substituted vinylpyridines were functionalized by TBADT-photocatalyzed addition of cycloalkanes. Scheme 14 showed the smooth synthesis of 142 starting from 141 simply by irradiation of the reaction mixture containing a slight excess of cyclohexane in the presence of a catalytic amount of the decatungstate salt.164

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0018.jpg
Functionalization of a Vinylpyridine with a Cycloalkane

Recently, alternative PCs have been developed for the direct photocatalyzed activation of C–H bonds in cycloalkanes, namely uranyl cation165 and Eosin Y,166 both having the advantage of absorbing in the visible light region. The alkyl radical formation may be induced by a photogenerated stable radical which acts as a radical mediator. An IrIII based photoredox catalyst oxidized the chloride anion (being the counterion of the Ir complex) to the corresponding chlorine atom, which abstracted a hydrogen atom from cyclopentane, thus forming adduct 151 in 69% yield upon addition onto a maleate ester (Scheme 15).167

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0019.jpg
Indirect HAT Mediated by a Cl Radical

Another intriguing way to induce the cleavage of unactivated C(sp3)-H bonds is by a photocatalyzed intramolecular hydrogen abstraction. Usually a photoredox or a proton-coupled electron transfer (PCET)47 step induced the formation of a heteroatom centered radical that abstracts a tertiary C–H bond intramolecularly in a selective fashion, following a 1,5-HAT process mimicking the Hoffmann-Löffler-Freytag reaction (Scheme 16).168170

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0020.jpg
Intramolecular 1,5-HAT Forming Tertiary Alkyl Radicals

When the reaction was applied to compound 161, an oxidative PCET generated a neutral amidine radical that promotes the 1,5-hydrogen atom abstraction forming a tertiary radical which is able to functionalize olefin 162 in a complete regioselective fashion affording 163 (Scheme 16a).171 The reaction was also applied to medicinally relevant molecules such as the steroid-derived trifluoroacetamide 164 (Scheme 16b). Despite the fact that this compound has several labile C–H bonds including tertiary C–H bonds and C–H bonds adjacent to heteroatoms, the intramolecular hydrogen abstraction followed by conjugate addition onto 165 gave 166 as the sole product.172

The remote activation of the C–H bond in the δ-position following this approach is a general reaction as demonstrated in related systems applied to amides protected with a carbamate group173 or in simple benzamide derivatives.174 In the latter case, the reaction was carried out in the presence of a chiral Rh-based Lewis acid catalyst that allowed the asymmetric alkylation of α,β-unsaturated 2-acyl imidazoles.174

The abstracting species could be likewise a photogenerated iminyl radical as illustrated in Scheme 17. Here a carbonyl group is converted in an oxime derivative (e.g., 171) by reaction with an α-aminoxy acid. Photocatalyzed oxidation followed by fragmentation of the resulting carbonyloxy radical gave an iminyl radical prone to a 1,5-HAT to afford a tertiary radical that upon addition to acrylate 172 gave compound 173 in 77% yield.175

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0021.jpg
1,5-HAT Promoted by an Iminyl Radical

2.1.2. Heteroalkylation of C–C Double Bonds

An interesting variation of the functionalization of a double bond is the formation of a C–C bond (upon an alkylation step) followed by the formation of another C–Y bond (Y ≠ H) on the adjacent carbon. As an example, alkyl diacyl peroxides were reduced photocatalytically and the fragmentation released an alkyl radical and a carboxylate anion both incorporated in the structure of the product. Thus, 2-vinylnaphthalene 182 was converted into compound 184 in a very good yield upon reaction with lauroyl peroxide 181 upon an oxidative quenching process by consecutive C–C and C–O formation (Scheme 18).176 The reaction was made possible by the oxidation of the resulting radical adduct 183 (by RuIII, the oxidized form of the PC) that generated the cation 183+ that was easily trapped by the carboxylate anion previously released.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0022.jpg
Oxyalkylation via Alkyl Diacyl Peroxides

N-(acyloxy)phthalimide 191 as radical precursor found use in a similar multicomponent oxyalkylation of styrenes. The addition of the alkyl radical onto the vinylarene followed by the incorporation of water present in the reaction mixture afforded derivative 192 in 72% yield (Scheme 19).177 Noteworthy, the labile C–Br bond in 191 remained untouched in the process.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0023.jpg
Multicomponent Oxyalkylation of Styrenes

The use of water as the oxygen source was likewise used in the difunctionalization of aryl alkenes where the carbon-centered radical was formed by an intramolecular 1,5-HAT of a photogenerated iminyl radical.178

Performing the reaction in DMSO, allows for the use of the solvent as an oxygen donor adopting the Kornblum oxidation. The intermediate benzyl radical formed after the alkylation step reacts with the solvent and eventually forming a carbonyl group in place of a simple C–O bond. An elegant example is shown in (Scheme 20) for the synthesis of ketonitrile 204.179 A cycloketone oxime ester (201) was photocatalytically reduced, inducing a ring opening on the resulting iminyl radical. The resulting cyano-substituted alkyl radical reacted with styrene 202, and the addition with DMSO formed the intermediate 203, that, upon Me2S loss, afforded the product.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0024.jpg
Photocatalyzed Oxyalkylation of Styrenes Based on the Kornblum Oxidation

A related oxyalkylation of styrenes made again the use of the Kornblum oxidation as the last step in the synthesis of substituted acetophenones. Indeed, N-hydroxyphthalimides (e.g., 211) were employed as the radical source, and an IrIII complex was used as the PC, obtaining good yields even on a 7 mmol scale (64% of 213, Scheme 21a).180,181 Ester 211 was also adopted for the preparation of the aryl alkyl ketone 215 in 61% yield (Scheme 21b). In this case, however, the decarboxylative alkylation was applied to silyl enol ethers having the carbonyl oxygen already incorporated in the initial structure such as 214.182 The same process described in Scheme 21b can be carried out under uncatalyzed conditions under blue LED irradiation in the presence of an excess of NaI (150 mol %) and PPh3 (20 mol %). The reaction was based on the photoactivation of a complex formed by N-(acyloxy)phthalimide with NaI and PPh3 through Coulombic and cation-π interactions. In this case, the excitation caused the reduction of the phthalimide by a SET reaction within the complex.183

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0025.jpg
Oxyalkylation by Using N-(Acyloxy)phthalimide Derivatives as Radicals Source

Alkylated ketones 22–3a–d were likewise obtained by the IrIII-photocatalyzed reaction between a 2-mercaptothiazolinium salt (221, as alkyl radical precursor) and silyl enol ethers 22–2a–d (Scheme 22).106

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0026.jpg
Synthesis of Alkylated Ketones from Mercaptothiazolinium Salts

Lauryl peroxide (LPO, see Scheme 18) was adopted for the Ru-catalyzed three-component carbofluorination of styrenes as illustrated in Scheme 23a. The vinylic double bond of compound 231 derived from estrone was functionalized twice by using triethylamine trihydrofluoride Et3N·HF as the fluoride anion source to deliver the desired alkyl-fluorinated olefin 232 in 61% yield. The reduction of LPO is mediated by the presence of a copper salt in the role of a cocatalyst in a dual catalytic process.184 The carbofluorination was later applied to dehydroalanine derivative 234 by using alkyltrifluoroborates and an excess of Selectfluor as an electrophilic fluorine source (Scheme 23b). The use of a visible light POC ([Acr+Mes]ClO4) allowed for the synthesis of a wide range of unnatural α-fluoro-α-amino acids including F-Leu (235).185

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0027.jpg
Carbofluorination of (a) Styrenes and (b) Dehidroalanine Derivatives

In rare instances, two C–C bonds could be formed in the adjacent position of the double bond as in cyanoalkylations. The enantioselectivity of the reaction was controlled exploiting the capability of a copper catalyst to form complexes with chiral Box ligands. Thus, a methyl radical was obtained by IrIII-photocatalyzed reduction of phthalimide 241 that readily attacked styrene (Scheme 24). Meanwhile, the CuI salt incorporated Box 242 as the ligand, and the resulting complex reacted with the adduct radical in the presence of TMSCN. As a result, cyanoalkylated 243 was obtained in a good yield and in good ee.186

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0028.jpg
Enantioselective Cyanoalkylation of Styrenes

A particular case of cyanoalkylation was later reported in the photocatalyzed reaction between cyclopropanols and cyanohydrins having a pendant C=C bond. Oxidative ring opening of the three-membered ring followed by addition onto the double bond and cyano migration gave a series of multiply functionalized 1,8-diketones incorporating the cyano group.187

2.1.3. Allylation

Allylation reactions can be easily performed by reaction of an alkyl radical with substituted allyl sulfones (mainly with 1,2-bis(phenylsulfonyl)-2-propene 251, Scheme 25a). The alkyl radical was generated under visible light irradiation by hydrogen abstraction from cycloalkanes by an aromatic ketone, e.g., 5,7,12,14-pentacenetetrone 252. Addition of a cycloalkyl radical onto 251 followed by sulfonyl radical elimination gave access to vinyl sulfones 25–3a–b in good yields (Scheme 25a).188

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0029.jpg
Allylation through Alkyl Radicals Generated (a) via HAT and (b) from Si bis-Catecholates

Other related reactions were designed to forge C(sp3)–allyl bonds following this simple scheme. The alkyl radical was formed by photocatalytic oxidation of hypervalent bis-catecholato silicon compounds as shown in Scheme 25b. Thus, compound 254 upon oxidation released the desired substituted alkyl radical, and addition onto allyl sulfone 255 gave the corresponding allylated derivative 256 in 70% yield.113 Olefin 255 was also used in a decarboxylative allylation of alkyl N-acyloxyphthalimides under RuII photocatalysis. The great advantage of the process was the reaction time since the allylation was completed in a few minutes at room temperature.189

A particular class of phthalimides could be employed with no need of a photocatalyst to promote the reaction. N-alkoxyphthalimide (261) is able to form a donor–acceptor complex with electron donor compounds, such as the Hantzsch ester HE. Upon excitation, an electron transfer occurred within the complex generating a radical anion, which released an alkoxy radical upon N–O bond cleavage. Loss of formaldehyde formed the desired alkyl radical which reacted with 262, to obtain product 263 in 60% yield (Scheme 26).127

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0030.jpg
Hantzsch Ester Mediated Photocleavage of N-Alkoxyphthalimides

Photocatalyzed reduction of Katritzky salts 27–1a–c obtained from the corresponding amines (Scheme 27) gave access to the allylated compounds 27–3a–c. Thus, the monoelectronic reduction of pyridinium salts 27–1a–c caused the release of the corresponding pyridines along with the substituted cyclohexyl radicals than upon trapping by 272 efficiently afforded acrylates 27–3a–c.190

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0031.jpg
Photocatalyzed Allylation by Using Katritzky Salts

A remote allylation under visible light irradiation was devised starting from amide 281 making use of eosin Y (EY) as the PC (Scheme 28). The excited EY is able to reduce 281 thanks to the electron-withdrawing capability of the substituted phenoxy group on the nitrogen of the amide. The amidyl radical formed upon fragmentation of 281•– gave rise to a tertiary radical upon 1,5-HAT, allowing the remote allylation, forming 282 in 75% yield.191

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0032.jpg
Remote Allylation via Amidyl Radicals

A different approach involved the use of trifluoromethyl-substituted alkenes (e.g., 291) that upon addition of the alkyl radical gave access to valuable gem-difluoroalkenes such as 29–2a–b (Scheme 29). The oxidation of alkyltrifluoroborates was here assured by the organic photocatalyst 4CzIPN, leading to nonstabilized primary, secondary, and tertiary radicals. The defluorinative alkylation resulted from the reduction of the radical adduct, followed by an E1cB-like fluoride elimination.192

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0033.jpg
Synthesis of gem-Difluoroalkenes from Alkyl Trifluoroborates

A dual catalytic approach was designed for valuable allylation using vinyl epoxides as allylating agents (Scheme 30). The mechanism was investigated by quantum mechanical calculations [by DFT and DLPNO–CCSD(T)] and supported an initial complexation of Ni0 to 302 that quickly underwent a SN2-like ring opening, followed by the incorporation of the alkyl radical formed by DHP-derived compounds 30–1a,b into the metal complex. Allyl alcohols 30–3a,b were then formed by inner sphere C(sp2)-C(sp3) bond formation from the resulting NiIII complex.193

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0034.jpg
Dual-Catalytic Allylation of Vinyl Epoxides

2.1.4. sp3–sp3 Cross-coupling

Another intriguing possibility offered by the photochemical approach to alkyl radicals is the formation of a C–C bond by a sp3–sp3 cross-coupling reaction. The transformation could lead to novel pathways to interesting targets, as represented by the synthesis of the drug tirofiban in only four steps, starting from easily available compounds. The protocol made use of two consecutive photocatalyzed reactions applying a metallaphotoredox strategy (Scheme 31). The key step is the coupling between carboxylic acid 311 and alkyl halide 312. The halide is first complexed by a Ni0 catalyst and the resulting NiI complex trapped the alkyl radical (obtained by photocatalyzed decarboxylation) to yield a NiIII complex that in turn released the sp3–sp3 coupled product 313 after desilylation with TBAF. The desired tirofiban was then obtained by elaboration of 313 in two subsequent steps.194

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0035.jpg
Synthesis of a Precursor of Tirofiban by a Metallaphotoredox Strategy

Another example of a C(sp3)–C(sp3) cross-coupling process is the reaction between alkylsilicates and alkyl halides. As in the previous case, a dual catalytic Ir/Ni system was required.195 The alkyl radical may be likewise generated from an alkyl halide by a halogen atom transfer with a photogenerated silyl radical (from a silanol). The radical that is hence formed could be coupled with another alkyl bromide, e.g., methyl bromide, using a Ni0 catalyst to perform valuable methylation reactions.196 Aliphatic carboxylic acids were used to form alkyl-CF3 bonds via a photocatalyzed reaction making use of Togni’s reagent as the trifluoromethylating agent. The reaction was promoted under visible light irradiation employing an IrIII photocatalyst coupled with a CuI salt. This process tolerates various functionalities including olefins, alcohols, heterocycles, and even strained ring systems.197

The alkylation of a benzylic position in N-aryl tetrahydroisoquinoline 321 was reported following two different approaches (Scheme 32). The first allowed the reaction of an unactivated alkyl bromide (322) by the excitation of a Pd0 complex. Compound 321 was oxidized in the catalytic cycle, and the resulting α-amino radical coupled with the isopropyl radical to form 324 in 81% yield (Scheme 32a).198 An alternative heavy-metal-free route catalyzed by a dye-sensitized semiconductor is depicted in Scheme 32b. Excitation of an inexpensive dye (erythrosine B) caused the reduction of titanium dioxide that in turn was able to reduce phthalimide 323 that eventually yielded quinoline 324.199

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0036.jpg
Different Strategies in the Alkylation of N-Aryl Tetrahydroisoquinolines

2.1.5. Other Reactions

In particular cases, a C=N bond can be made sufficiently electrophilic to undergo alkyl radical addition as in the case of N-sulfinimines, exploited for the preparation of protected amines. A high degree of diastereoselectivity can be obtained when starting from chiral N-sulfinimines (33–2a–c, Scheme 33). Thus, the asymmetric addition of an isopropyl radical (formed from derivative 331) onto 33–2a–c allowed the isolation of sulfinamides 33–3a–c in good yields.200

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0037.jpg
Addition of an Alkyl Radical to Chiral Sulfinimines

The alkylation of related imines can be carried out by using ammonium alkyl bis(catecholato)silicates as the radical precursors under metal-free conditions adopting 4CzIPN as the POC201 or by using potassium alkyltrifluoroborates in the alkylation of N-phenylimines.202

Another particular case is the alkylative semipinacol rearrangement devised for the synthesis of 2-alkyl-substituted cycloalkanones. The reaction involved the photocatalytic reaction between TMS protected α-styrenyl substituted cyclic alcohol 342 and the unactivated bromoalkane 341 (Scheme 34a). The reaction was promoted by the dimeric gold complex [Au2(dppm)2]Cl2. This complex is able to reduce 341 (ca. −2.5 V vs SCE) despite having an oxidation potential in the excited state considerably lower for the reaction to occur (ca. −1.63 V vs SCE). This can be explained by the formation an inner sphere exciplex between the excited dimeric catalyst and 341 that promotes the otherwise thermodynamically unfeasible redox process, generating an AuI–AuII dimer and 344. The combination of the latter species formed an AuIII complex that induces a semipinacol rearrangement coupled with C(sp3)–C(sp3) reductive elimination, which furnished 343 in 84% yield (Scheme 34b).203

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0038.jpg
Gold Catalyzed Activation of Bromoalkanes

A similar reaction was later developed starting from cycloalkanol-substituted styrenes and N-acyloxyphthalimides under IrIII photocatalysis.204

2.2. Formation of a C(sp3)-C(sp2) Bond

2.2.1. Alkenylation

The reaction between an alkyl radical with a cinnamic acid followed by loss of the COOH group is one of the more common approaches to promote an alkenylation reaction. Thus, the radical formed from salt 353 attacked the benziodoxole adduct 352, synthesized from acrylic acid 351. The reaction yielded 83% of the diphenylethylene derivative 354 upon a deboronation/decarboxylation sequence (Scheme 35).205 The benziodoxole moiety gave efficient results in promoting the radical elimination step, while other noncyclic IIII reagents were ineffective.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0039.jpg
Alkenylations Mediated by Benziodoxole

Different decarboxylative alkenylations have been reported by changing the radical source and the photocatalyst (Scheme 36). The homolytic cleavage of an alkyl-I bond has been promoted by a CuI complex and the resulting cyclohexyl radical afforded styrene 363 in 68% yield upon addition onto cinnamic acid 361 (path a).206 The same product may be formed as well starting from the same substrate by using phthalimide 362 under visible light irradiation with the help of an IrIII207 (path b) or a RuII photocatalyst.208 As an additional bonus, the formation of adduct 363 was obtained with a preferred E configuration.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0040.jpg
Different Strategies Toward Decarboxylative Alkenylations

The alkenylation may mimic a Heck reaction as in the visible light-induced Pd-catalyzed reaction between a vinyl (hetero)arene and an α-heteroatom-substituted alkyl iodide or bromide (see Scheme 37). Here, the generation of the radical is induced by the reduction of the TMS-derivative 371 by the excited Pd0 species. Radical addition onto 372 followed by β-H-elimination from the adduct radical delivered allyl silane 373 in 81% yield.209 Noteworthy, the same reaction did not take place under usual thermal Pd-catalysis.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0041.jpg
Heck-Like Alkenylation of an Alkyl Iodide

Alkylation of styrenes could be carried out using an inexpensive palladium source (Pd(PPh3)4) with no need of any base or classical photocatalyst. The reaction was promoted by visible light, adopting N-hydroxyphthalimides as radical sources.210 Other visible light Pd promoted alkenylations include the reaction of vinyl arenes with carboxylic acids211 or tertiary alkyl halides212 as radical precursors. Other metal catalysts, however, were helpful for the substitution of a vinylic hydrogen atom with an alkyl group. In this respect, a dinuclear gold complex was employed for the activation of an alkyl bromide to promote a photocatalyzed Heck-like reaction.213 The synergistic combination of a POC and a cobaloxime catalyst promoted the photocatalyzed decarboxylative coupling between 381 and styrene 382 to give the alkenylated product 383 in 82% yield and with a complete E/Z selectivity as illustrated in Scheme 38.214

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0042.jpg
Cobaloxime-Mediated Decarboxylative Coupling of Carboxylic Acids with Styrenes

The addition of the alkyl radical may take place even on substituted alkenes via an ipso-substitution reaction. An example is shown in Scheme 39 where a vinyl iodide (392) is used for an alkenylation by the reaction with a radical generated from silicate 391, obtaining compound 393. The RuII photocatalyst in the dual catalytic system has the role of generating the radical, while the Ni0 catalyst activates the C(sp2)-I bond.215

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0043.jpg
Alkenylation via Alkenyl Iodides

Alkenylation of alkyl iodide 401 can also take place starting from an alkenyl sulfone (402). Also in this case, an ipso-substitution is central to the novel bond formation and the Pd0 catalyst formed the radical by a SET reaction with 401. After the addition of the radical onto 402, the sequence is completed by the elimination of a sulfonyl radical affording 53% yield of 403 (Scheme 40).216

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0044.jpg
Alkenylation of an Alkyl Iodide with Alkenyl Sulfones

Alkyl bromides were used in alkenylations by reaction with vinyl sulfones made possible by the photocatalytic generation of silicon centered radical that in turn formed the alkyl radical by a halogen atom transfer reaction.217

2.2.2. Acylation

Acylation owes its importance to the possibility to convert an alkyl radical into a ketone, a reaction that proceeds in most cases with the intermediacy of an acyl radical.218,219 A classical approach is based on the homolytic cleavage of an alkyl-I bond followed by carbonylation with CO and reaction with electrophiles of the resulting nucleophilic acyl radical. Scheme 41 illustrates the concept. Irradiation of iodide 411 with a Xe lamp in the presence of CO (45 atm) and a Pd0 complex led to an electron transfer reaction which formed an alkyl radical that, upon carbonylation and addition onto phenyl acetylene, gave ynone 412 in 63% yield.220 The reaction is supposed to proceed via a photoinduced electron transfer from the Pd0 catalyst to the iodoalkane, furnishing a PdII species and the alkyl radical. The carbon-centered radical promptly reacts with CO to generate an acyl radical. The PdI catalyst intervenes here again to couple the acyl derivative with the alkyne, preserving the triple bond in the final product. This reaction was later applied to the acylation of styrenes to give the corresponding enones.221 The electrophilic nature of the alkyne could be exploited if the moiety is placed in the same reagent bearing the iodide. In this case, the first reaction observed was an intramolecular cyclization forming an alkenyl radical which eventually reacted with CO, furnishing an α,β-unsaturated ketone.222

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0045.jpg
Photocatalyzed Synthesis of Ynones

A reductive step induced the generation of the alkyl radical through an IrIII-photocatalyzed C–N bond activation in pyridinium salt 421 (Scheme 42). Trapping of the alkyl radical with CO followed by addition onto 1,1-diphenylethylene gave access to the Heck-type product 422 with no interference by the 2,4-dioxo-3,4-dihydropyrimidin-1-yl ring.223

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0046.jpg
Synthesis of Enones via Photocatalyzed C–N Bond Activation

The alkyl radical to be carbonylated was likewise formed starting from a cycloalkane for the preparation of unsymmetrical ketones via radical addition onto Michael acceptors. The reaction proceeded via a photocatalyzed decatungstate hydrogen atom transfer reaction224 When cyclopentanones were subjected to the photocatalyzed C–H activation, a regioselective β-functionalization occurred. Thus, 1,4-diketones 43–3a–c were smoothly formed by reaction of the photogenerated acyl radical 431 onto Michael acceptors 43–2a–c (Scheme 43).225

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0047.jpg
Photocatalyzed Synthesis of Unsymmetrical Ketones

Unsymmetrical ketones have been likewise formed by carbonylation of alkyl radicals generated from organosilicates by using 4CzIPN as POC under visible-light irradiation.226

Potassium alkyltrifluoroborates were extensively used for acylation reactions having recourse to a dual photocatalytic system. The unstabilized alkyl radical was generated from trifluoroborate 441 with the help of an IrIII PC (Scheme 44). Meanwhile, the acid 442 was converted in situ into a mixed anhydride (by reaction with dimethyl dicarbonate, DMDC) that was activated by a Ni0 complex. Addition of the alkyl radical onto the resulting complex led to the acylated product 443.227 In a similar vein, Ir-photoredox/nickel catalytic cross-coupling reactions were devised by using acyl chlorides228 and N-acylpyrrolidine-2,5-diones229 as acylating reagents.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0048.jpg
Dual Catalytic Acylation of Alkyl Trifluoroborates

A Ni/Ru, dual-catalyzed amidation protocol was possible thanks to the coupling between an alkylsilicate and an isocyanate. Even in the latter case, the alkyl radical attacked the complex formed between the isocyanate and a Ni0 species and, as a result, the mild formation of substituted amides took place.230

The acylation of the radical was also exploited for the synthesis of esters. This elegant approach involves the generation of radicals from unactivated C(sp3)–H bonds (e.g., in cycloalkanes). The hydrogen abstraction on cycloalkanes was induced by a chlorine atom released from the photocleavage of the complex formed between chloroformate 451 and a Ni0 complex, allowing one to synthesize scaffolds with different ring sizes (45–2a–d in Scheme 45).231

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0049.jpg
Dual Catalytic Acylation of Cycloalkanes

2.2.3. Minisci-Like Reactions

A fundamental transformation for the construction of C(sp2)-C(sp3) bond is the Minisci reaction, where the functionalization of heteroaromatics took place by substituting a H atom with an alkyl group. The reaction was widely investigated in the last years and mainly involves the functionalization of a nitrogen-containing heterocycle.232 An interesting example is the methylation reported in Scheme 46.94 A methyl radical was formed by using a peracetate such as 461. The protonation of 461 by acetic acid facilitates a PCET reduction of the peracetate by the IrIII PC. A double fragmentation ensued, and the resulting methyl radical may attack the protonated form of biologically active heterocycles (e.g., fasudil 462) in a mild selective manner to afford 463 in 43% yield.94

Another approach made use of an alkyl boronic acid as the radical precursor. The process is initiated by the RuII-photocatalyzed reduction of acetoxybenziodoxole (BI-OAc) that liberated the key species ArCOO (Scheme 47). Upon addition onto an alkyl boronic acid, this ortho-iodobenzoyloxy radical made available the alkyl radical that in turn functionalized pyridine 471 in position 2 in 52% yield (472, Scheme 47).233

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0051.jpg
Minisci Reaction by Using Alkyl Boronic Acids

The generation of the alkyl radical from boron-containing derivatives was made easier starting from alkyltrifluoroborates. A POC (Acr+Mes) is, however, required, but in all cases, the regioselective functionalization of various nitrogen-containing heterocycles was achieved.234 A related chemical oxidant-free approach process was later developed where alkyl radicals were formed by merging electro and photoredox catalysis.235

Alkyl halides are versatile substrates for the photoinduced functionalization (e.g., butylation) of lepidine 481 (Scheme 48). An uncatalyzed redox process is a rare occurrence here, since alkyl halides reduction is more demanding. This drawback can be overcome by the adoption of a dimeric AuI complex (see Scheme 34) that upon excitation coordinates an unactivated haloalkane promoting an inner sphere PET. This interaction pushes the activation of R-Br despite its larger Ered with respect to the PC (Scheme 48a).236 A different approach promoting the homolytic cleavage of the R-I bond is shown in Scheme 48b. Decacarbonyldimanganese Mn2(CO)10 was cleaved upon visible light irradiation, and the resulting Mn-based radical was able to abstract the iodine atom from an alkyl iodide thus generating the desired butyl radical. This route was smoothly applied to the late-stage functionalization of complex nitrogen-containing substrates.237 Moreover, the activation of alkyl halides may be obtained by the photogeneration of a silyl based radical derived by TTMSS (Scheme 48c, see also Scheme 11). The robustness and the mildness of this approach was witnessed by the broad substrate scope and the compatibility of several functional groups present in the radical.238 The use of acidic conditions (required to make the nitrogen heterocycle more electrophilic) may however be avoided. Excited [Ir(ppy)2(dtbbpy)]PF6 was sufficiently reducing to convert alkyl iodides to alkyl radicals under basic conditions by combining conjugate and halogen ortho-directing effects.239

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0052.jpg
Photocatalyzed Butylation of Lepidine

In general, lepidine is the preferred substrate to test new ways for the C–H alkylation of heteroarenes. Accordingly, adamantane carboxylic acid 491 served for the visible light induced synthesis of 493 starting from lepidine 492 (Scheme 49). An IrIII PC was adopted to alkylate various nitrogen heterocycles, making use of a large excess of persulfate anion as the terminal oxidant (path a).240 The presence of a PC is not mandatory for the adamantylation reaction with (bis(trifluoroacetoxy)-iodo)benzene as starting material. This compound in the presence of a carboxylic acid gave the corresponding hypervalent iodineIII reagent that upon irradiation generates the alkyl radical. The TFA liberated in the process was crucial for the activation of the nitrogen heterocycle and adduct 493 was isolated in 95% yield (path b).241

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0053.jpg
Decarboxylative Minisci Alkylation

Very recently, an interesting approach for the generation of alkyl radicals from the C–C cleavage in alcohols was reported making use of a CFL lamp as irradiation source. The combination of 2,2-dimethylpropan-1-ol (501) with benziodoxole acetate (BI-OAc) gave adduct 503. Photocatalytic reduction of compound 503 released and alkoxy radical that upon fragmentation formed a tbutyl radical that reacted with N-heteroarene 502 to form 504 in 57% yield (Scheme 50).130

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0054.jpg
Aliphatic Alcohols as Radical Precursors in Minisci Reaction

The use of hypervalent iodineIII in promoting the decarboxylation of R-COOH was effective in the derivatization of drugs or drug-like molecules. As a result, the quinine analogue 512 was formed in a 76% yield from quinine 51–1, utilizing Acr+Mes as the POC (Scheme 51).242

Azoles can be adamantylated starting from adamantane carboxylic acid by a dual catalytic approach (Acr+Mes as the POC and [Co(dmgH)(dmgH2)Cl2] as the cocatalyst)243 or simply C2-alkylated under photoorganocatalyzed conditions.244

The photocatalyzed reduction of N-(acyloxy)phthalimide 521 induced by an IrIII* complex is an alternative approach for the functionalization of N-heterocycles such as 2-chloroquinoxaline 522 to form the cyclopentenyl derivative 523 (Scheme 52).245

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0056.jpg
Cyclopentenylation of 2-Chloroquinoxaline

The reductive pathway is feasible even when the generation of the alkyl radical was carried out starting from the redox-active pyridinium salt 531. In this case, the obtained cycloalkyl radical gave a regioselective addition onto 6-chloroimidazo[1,2-b]pyridazine 532 to yield 533 under mild conditions (Scheme 53).70

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0057.jpg
Functionalization of 6-Chloroimidazo[1,2-b]pyridazine

The alkyl radical could be formed even from simple hydrocarbons via hydrogen atom transfer reaction. A valuable example is reported in Scheme 54. The hypervalent iodine oxidant PFBI–OH is reduced by an excited RuII complex generating a carbonyloxy radical that acted as hydrogen atom abstracting agent. Functionalization of isoquinoline 542 by the resulting radical (derived from 541) afforded adduct 543 in 65% yield (>15:1 dr).246 The high selectivity observed in the functionalization of 541 was ascribed to the slow addition of the tertiary alkyl radical possibly formed onto 542.246 The direct (rather than indirect) C–H cleavage in cycloalkane was possible by using decatungstate anion as PC. Various nitrogen-containing heterocycles were then easily derivatized even under simulated solar light irradiation.247

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0058.jpg
PFBI–OH Mediated Minisci Reaction

PFBI–OH was likewise used for the remote C(sp3)–H heteroarylation of alcohols (Scheme 55). As an example, the reaction of pentanol with PFBI–OH gave adduct 551 that was reduced by the photocatalyst releasing the alkoxy radical 552. 1,5-HAT and addition onto protonated phthalazine 553 afforded adduct 554 and the functionalized heterocycle 555 from it in 72% yield after sequential oxidation and deprotonation.248

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0059.jpg
Remote C(sp3)–H Heteroarylation of Alcohols

As previously stressed, an acid is often required for an efficient Minisci-like reaction. To overcome this problem the alkylation may be carried out on the corresponding N-oxide derivatives as it is the case of pyridine N-oxides (562, Scheme 56). The radical is generated from a trifluoroborate salt (561) and the alkylation is regioselective in position 2 (forming compound 563).249 The process is efficient thanks to the photocatalytic degradation of BI-OAc that promoted a hydrogen abstraction, operated by the resulting carbonyloxy radical, on the Minisci radical cation adduct.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0060.jpg
Minisci Alkylation of Pyridine Oxides

On the other hand, the pyridine N-oxide 571 can be acylated in situ with suitable acyl chlorides to furnish the electron-poor 57–2a–c+ derivatives. Photocatalytic reduction of these intermediates leads to the generation of alkyl radicals prone to attack the pyridine nucleus itself in the ortho position resulting in a decarboxylative alkylation (57–3a–c, Scheme 57).122

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0061.jpg
Decarboxylative Alkylation of Heterocycles

2.2.4. Ipso-Substitution Reactions

The forging of an alkyl-sp2 bond (e.g., an alkyl-Ar bond) is undoubtfully one of the most crucial goals pursued by a synthetic organic chemist. Alkyl radicals generated via different mild routes can be successfully employed for the arene ipso functionalization, given the presence of a suitable group X on the (hetero)aromatic ring that directs the selective formation of a new Ar–C bond at the expense of an Ar-X bond. Dual catalysis (with the help a Ni-based complex) is one of the preferred approaches.

In a recent example, the hydrogen atom transfer ability of the excited TBADT catalyst (see also Scheme 12) is used to form an alkyl radical starting from different aliphatic moieties (see Scheme 58).250 The combined action of the tungstate anion and the nickel catalyst (Ni(dtbbpy)Br2) allowed the coupling of (hetero)aromatic bromides with unactivated alkanes, overcoming their high bond dissociation energies (ca. 90–100 kcal/mol) and low acidities. Both linear (41–56% yield) and cyclic (57–70% yield) alkanes could be functionalized with a vast range of competent partners. Interestingly, the radicals are generated preferentially on the less sterically demanding secondary carbons in alkanes, affording a remarkable selectivity.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0062.jpg
Dual TBADT-Ni Catalysis for the Synthesis of Pyridyl-Functionalized Bicycles

The scope of this method could be proved by the functionalization of natural products and drugs, such as in the preparation of the bicyclic derivative 58–3a (61% yield) and the N-Boc protected epibatidine alkaloid 58–3b (28%, Scheme 58).250 A very similar approach was later reported for the dual photocatalytic formation of an Ar–C bond starting from aryl bromides and cycloalkanes.251

Another dual-catalytic approach allowed the coupling reaction of aryl bromides (592, Scheme 59) and alkyl sulfinates (591), in the presence of Ni(COD)2 and tetramethylheptanedione (TMHD, Scheme 59a) to give 593 in 84% yield under air.104 The photogenerated radical was trapped by the Ni complex that mediated the coupling with the aryl halide 592. The method was then applied to the synthesis of 59–5, selective ATP-competitive inhibitors of the casein kinase 1δ, an enzyme related to the regulation of the circadian rhythm (Scheme 59b).104

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0063.jpg
Dual Catalytic Cross-Coupling of Aryl Bromides with Alkyl Sulfinates

A very similar strategy to access C(sp3) radicals involves the photoredox induced cleavage of alkyl oxalate 601, starting from the corresponding alcohols (see Scheme 60, see also Scheme 3).252 The rapid in situ formation of the oxalate (without purification) was followed by the metallaphotoredox sequence based on Ni catalysis, allowing to obtain the C(sp2)-C(sp3) coupling to give derivatives 60–3a–e in good yields.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0064.jpg
Coupling of Alkyl Oxalates with Aryl Bromides

The advantage of the use of potassium and ammonium bis-catecholato silicates relies in the smooth generation of unstabilized primary and secondary alkyl radicals to be engaged in dual catalysis.253,254 An example is the consecutive functionalization of bromo(iodo)arene 612 (Scheme 61, see also Scheme 25) for the preparation of 614 where the radical (from 611) is trapped by Ni0 (stabilized by a phenanthroline ligand). The synthesis of 613 can be achieved in high yields on 10 mmol scale with reduced effect on yield (75%) and selectivity (98%). The crude bromide 613 was further functionalized by a second Ni/photoredox cross-coupling of the alkylsilicate 615, affording product 614 in 66% yield.255 The procedure was extended successfully to alkyl triflates, tosylates and mesylates,256 and to brominated borazaronaphthalene cores.257 The latter approach was crucial to access previously unknown isosteres of azaborines.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0065.jpg
Consecutive Functionalization of Bromo-Iodo Arenes with bis-Catecholato Silicates

The action of a silyl radical on an alkyl bromide 621 forms an alkyl radical that, again with the help of a Ni based catalyst, reacted with aryl bromides 622 (Scheme 62, see also Scheme 11). The scope of products 623 that can be obtained is varied and includes both aromatic and heteroaromatic substrates, along with cycloalkanes of different size.71

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0066.jpg
Ir/Ni Complex Mediated Coupling Between Alkyl and Aryl Bromides

A peculiar case is when the ipso-substitution took place via a radical rearrangement such as shown in Scheme 63.258 Thus, the heteroaromatic sulfonamide 631 was subjected to the Finkelstein reaction, obtaining the corresponding iodide 632 that acted as the source of radical able to induce a Smiles rearrangement via 635. Intermediate 635 has lost its aromaticity; however, the radical has become tertiary, gaining further stabilization from the ester group nearby. Restoration of the aromaticity is followed by a presumable hydrogen atom transfer to obtain compound 636 in 95% yield.258

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0067.jpg
Photocatalyzed Smiles Rearrangement

Dual photoredox/nickel catalysis was successfully applied to couple β-trifluoroboratoketones 641 with aryl bromides 64–2a–f (Scheme 64, see also Scheme 4). Arylated compounds 64–3a–f were efficiently prepared with substituents of different electronic nature on the aryl ring.259

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0068.jpg
Photoredox/Nickel Dual Catalytic Coupling of β-Trifluoroboratoketones with Aryl Bromides

Potassium tetrafluoroborate salts have been applied to generate secondary alkyl radicals via Ir photocatalysis coupled with Ni.260 However, they were found to be likewise suitable for cross-coupling reactions devoted to the forging of quaternary carbon centers (Scheme 65) without the need of using reactive organometallic species.261 In the adamantylation of bromides 65–1a–d better yields were obtained when the aryl ring was substituted with electron-withdrawing groups.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0069.jpg
Adamantylation of Aryl Bromides

An interesting application of this synthetic strategy is the functionalization of 7-azaindole pharmacophores with cycloalkyl scaffolds to improve the drug likeness of the azaindole core structure. Different potential drug candidates (66–3a–c, Scheme 66) were prepared via dual photocatalysis in a flow setup varying the dimension and substitution of the ring.262

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0070.jpg
Functionalization of 7-Azaindole Pharmacophores in Flow

In a similar way, a DHP-functionalized cyclohexene 671 was used to generate a secondary alkyl radical. In this case, the authors promoted the oxidation of 671 by using the strongly oxidizing 4CzIPN photocatalyst. Coupling with bromopyridine 672 gave substituted pyridine 673 in moderate yields (Scheme 67, see also Scheme 7).118

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0071.jpg
Coupling of DHP-Cyclohexene with Cyanobromopyridine

DHP-derivatives (681) may be used in ipso-substitution reaction even in the absence of a photocatalyst (Scheme 68). Violet light LED illumination directly excited didehydropyridine 681, fueling electrons to the NiII species which formed the catalytic competent Ni0 along with the desired radical by the fragmentation of 681•+. Noteworthy, the alkyl aromatic 683 was then formed where the use of electron-withdrawing groups on the ring contributes to the good yields of the process.263

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0072.jpg
Photocatalyst-Free Activation of DHPs

2.3. Formation of a C(sp3)-C(sp) Bond

2.3.1. Cyanation

Alkyl radicals have been used for the synthesis of alkyl nitriles by using different cyanide sources. The cyanation may be carried out in the presence of cyanide anion as tetrabutyl ammonium salt (TBACN). The C–C bond formation here may be carried out under very mild conditions by using the inexpensive precatalyst CuI, starting from unactivated alkyl chlorides (e.g., 69–1a–c, Scheme 69). Probably, a CuI-cyanide adduct is the species that was excited and engaged an electron transfer reaction with the alkyl halide to form a CuII-cyanide adduct. This intermediate combines with the alkyl radical formed to release nitriles 69–2a–c. The CuI-halide complex formed in the reaction restores the initial photocatalyst by exchange with the cyanide anion.264

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0073.jpg
CopperI-Mediated Synthesis of Nitriles

TMSCN was instead used for the remote δ-C(sp3)-H cyanation of alcohols under Ir/Cu-photocatalyzed conditions. The reduction of an N-alkoxypyridinium salt generated an alkoxy radical that upon intramolecular 1,5-HAT formed an alkyl radical that is cyanated with the help of the copper catalyst.265

A typical cyanation procedure, however, makes use of tosyl cyanide as cyanating agent. Thus, the radical obtained by oxidation of trifluoroborate 70–1a (by excited Acr+Mes)266 or acid 70–1b (by riboflavin tetraacetate RFTA)267 was trapped by tosyl cyanide to afford nitrile 702 by a substitution reaction (Scheme 70).

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0074.jpg
Cyanation of (a) Trifluoroborates and (b) Carboxylic Acids

A related RuII-photocatalyzed cyanation employing Ts-CN starting from alkyl trifluoroborates but requiring BI-OAc as a mild oxidant has been likewise reported.268

An elegant way to forge an alkyl-CN bond required the photocatalyzed elaboration of cyanohydrines 71–1a–d. At first, the interaction of the OH group with the sulfate anion (generated by the decomposition of persulfate anion) allowed its oxidation by a proton-coupled electron transfer (PCET) process promoted by an in situ formed IrIV species. Alkoxy radicals 712a–d were then formed and promoted a regioselective cyanation of remote C(sp3)–H bonds by a 1,5-HAT followed by cyano migration to form cyanoketones 71–3a–d (Scheme 71).269

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0075.jpg
Photocatalyzed Cyano Migration in Cyanohydrines

2.3.2. Alkynylation

Direct alkynylation of photogenerated alkyl radicals could be accomplished utilizing a reagent or catalyst that activates the alkyne moiety, making it more prone to the forging of a novel C(sp3)–C(sp) bond. One of the first strategies that were employed made use of benziodoxole-functionalized alkynes to promote the reaction with the alkyl radical.270 A representative case is illustrated in Scheme 72. [Ru(bpy)3](PF6)2 promoted the alkyl radical formation from trifluoroborate salt 721 that upon addition onto the alkynyl derivatives 72–2a–d induced the alkynylation via the intermediacy of vinyl radicals 723a–d. This deboronative alkynylation strategy could be performed in neutral DCM:water (1:1) at room temperature, giving access to the alkynylation of primary, secondary, and tertiary derivatives. To further prove the mildness of the conditions used, the authors carried out the reaction in PBS at pH 7.4 in the presence of biomolecules such as amino acids, but also single-stranded DNA and proteins (e.g., bovine serum albumin), obtaining satisfactory yields ranging from 68 to 86% of selectively alkynylated product.270

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0076.jpg
Alkynylation of Alkyl Trifluoroborates

The nature of the substituents on the alkynylbenziodoxole reagent were proved to determine the outcome of the alkynylation process. The electron-rich compounds performed better in the photocatalyzed transformation, both as radical acceptor and oxidative quencher of the RuII* photocatalyst.271

A similar strategy to the one mentioned before consists in the IrIII-photoredox-catalyzed alkynylation of carboxylic acids 732 (see Scheme 73a, path a).272,273 In this case benziodoxole derivatives 731 were again used to activate the sp carbon of the alkyne to the radical attack, affording good yields of products 73–3. Following these results, they developed a reaction to synthesize ynones 734 utilizing the same reaction conditions in the presence of gaseous CO (see Scheme 73a, path b and Section 2.2.2). Gram-scale reactions and late-stage functionalization of natural terpenoids such as ursolic acid (735, Scheme 73b) were likewise reported.273

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0077.jpg
IrIII-Catalyzed Alkynylation of Carboxylic Acids

Alkynyl sulfones were extensively employed as alkynylating reagents, with a mechanism like the one described in Scheme 72. Alkynyl phenyl sulfone was used in combination with N-acyloxyphthalimide derivatives as radical precursors in a RuII-photocatalyzed reaction that gave direct access to TIPS-substituted alkynes.274N-Phthalimidoyl oxalates and tolyl alkynyl sulfones were found to be competent for the reaction (even for the preparation of internal alkynes having quaternary carbons),275,276 the latter even in combination with pyridinium salts as radical precursors.277 The consecutive photoredox decarboxylative coupling of doubly functionalized adipic acid derivatives with alkynyl phenyl sulfones induced the cascade formation of interesting cyclic derivatives with an exo double bond (Scheme 74). In this case, compound 741 underwent two efficient consecutive photoredox decarboxylative couplings leading first to alkyne 743 that it was subjected to radical cyclization to form radical 744 and styrene 745 from it.278 The authors reported the formation of five-membered rings via the consecutive formation of two C–C bonds, along with one example showing the application to the synthesis of six-membered derivatives (31% yield).278

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0078.jpg
Cascade Double Alkynylation of Functionalized Adipic Acids

In rare instances alkynyl bromides could be used as sp counterpart in the radical addition of alkyl derivatives obtained from the oxidative decomposition of various Hantzsch esters under visible light conditions promoted by 4CzIPN.279

The versatility of the photocatalytic method, however, allowed to obtain functionalized alkynes starting from terminal alkynes (Scheme 75). The first approach involves the UV light induced cleavage of the C–I bond in iodide 751 (used in large excess) in basic aqueous media. Addition of the cyclohexyl radical onto alkyne 752 followed by the incorporation of the iodine atom gave vinyl iodide 75–3. The strong basic conditions used (NaOtBu) coupled with heating (up to 50 °C) favored an elimination of HI to yield the desired alkyne 754 under metal-free conditions (Scheme 75a).280 Visible-light (450 nm) was used in the copper-catalyzed coupling of an alkyl iodide (755) and again a terminal alkyne (756, Scheme 75b). The success of the reaction was ascribed to the use of terpyridine ligand 758 that avoided the photoinduced copper-catalyzed polymerization of the starting substrates. Probably, the reaction started by the excitation of the first formed copper acetylide that upon SET with 755 promoted the synthesis of alkyne 757 in high yields.281

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0079.jpg
Alkylation of Terminal Alkynes

3. Formation of a C(sp3)-Y Bond

3.1. C–B Bond

Borylation of an alkyl derivative to access differently substituted boron containing compounds can be carried out under mild conditions, employing different photochemical approaches. Thus, the alkyl radical formed from an N-hydroxyphthalimide 761 (derived from dehydrocholic acid) may be trapped either by bis(pinacolato)diboron (B2pin2) to give the corresponding alkyl pinacol boronates 762 (Scheme 76, path a) or by tetrahydroxydiboron (B2(OH)4) followed by treatment with KHF2 to give alkyl tetrafluoroborates (Scheme 76, path b).282

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0080.jpg
Photocatalyzed Borylation of N-Hydroxyphthalimides

A variation of the previous methodology involves the irradiation of N-hydroxyphthalimide esters 772 in the presence of B2cat2 (771) with the help of N,N-dimethylacetamide (DMAc) as the solvent under uncatalyzed conditions (Scheme 77). These components formed a heteroleptic ternary complex able to be excited by blue light and ultimately leading to the corresponding benzo[1,3,2]dioxaborole 774 that upon treatment with pinacol and TEA released the desired pinacol boronic ester 773. The functionalization of a series of drugs and natural products, such as pinonic acid and fenbufen were likewise effective, underlying the broad scope and functional group tolerance of the method.283

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0081.jpg
Photocatalyst-Free Borylation

Two related approaches were later developed and involve the irradiation of the ternary complex formed by differently substituted N-alkyl pyridinium salts, B2cat2 and DMAc. The reaction gave again pinacol boronic esters in what is considered a deaminative protocol for the borylation of aliphatic primary amines since the latter compounds were used for the synthesis of the pyridinium salts.284286

Interestingly, 2-iodophenyl thionocarbonates were later adopted as radical precursor for the preparation of boronic ester via photocatalyzed reaction with B2cat2 (Scheme 78).95 The strategy is based on the photoinduced reduction of compound 781 that upon iodide anion elimination formed aryl radical 782 that underwent a 5-endo-trig cyclization causing the release of benzo[d][1,3]oxathiol-2-one 783 and alkyl radical 784. Usual borylation gave boronic ester 785 in 85% yield.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0082.jpg
Borylation of 2-Iodophenyl Thionocarbonates

3.2. C–N Bond

Diverse structural motifs based on the C–N bond such as hydrazine and hydrazide cores were accessed by the photochemical addition of alkyl radicals onto the N=N of azodicarboxylates. TBADT-photocatalyzed HAT was applied to synthesize hydrazines by the coupling of cycloalkyl radicals with diisopropyl azodicarboxylate (DIAD). A synthetically challenging three component reaction can be achieved in the presence of CO, allowing the synthesis of the corresponding hydrazides.287

The C–H amination can be smoothly achieved even starting from light hydrocarbons, such as methane (791, Scheme 79), with ditert-butylazodicarboxylate (DBAD, 792) in the presence of CeIII salts. This inexpensive photocatalyst furnished the desired product 793 in 63% yield, with a turnover number up to 2900. The authors proposed a ligand-to-metal charge transfer excitation between the cerium salt and trichloroethanol as the source of alkoxy radicals that acted as hydrogen atom transfer agents.288

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0083.jpg
Photocatalyzed Amination of Methane

Aminated alkanes can be obtained by reacting aliphatic carboxylates with DIAD making use of Acr+Mes as a photoredox catalyst.289 A cerium catalyst was adopted for the generation of several alkyl radicals starting from carboxylic acids, under basic conditions, allowing for the functionalization of a broad range of substrates, including natural products such as drugs like gemfibroxil (802) and tolmetin (801, Scheme 80).290

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0084.jpg
Cerium-Catalyzed Decarboxylative Amination

The N=N bond of differently substituted azobenzenes (81–1a–e) can be functionalized on both nitrogens with a tandem N-methylation and N-sulfonylation, by cleavage of DMSO by UV irradiation of the Fenton reagent (FeSO4/H2O2,Scheme 81).105

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0085.jpg
Tandem N-Methylation and N-Sulfonylation of Azobenzenes

Synthesis of amides can be achieved recurring to copper photocatalysis. Secondary alkyl bromide 821 could be efficiently coupled with cyclohexane carboxyamide 822 in 90% yield using CuI in catalytic amounts (Scheme 82a). The authors were able to isolate the catalytic species (a copper–amidate complex), formed by the assembly of four copper ions and four amides.291

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0086.jpg
Photocatalyzed Synthesis of (a) Amides and (b) Carbamates

The same group reported the functionalization of carbamates with secondary alkyl bromides by shifting the wavelength of irradiation in the visible region by developing a tridentate carbazolide/bisphosphine ligand 824 for the copper catalyst thus able to prepare Boc-pregnenolone 825 in 90% yield (Scheme 82b).292 A variation of this protocol was applied to the synthesis of amines, using secondary unactivated alkyl iodides and CuI/BINOL as the catalytic system.293

Several reagents can be used as an azide source to synthesize synthetically valuable C–N3 bonds. Tertiary aliphatic C–H bonds can be selectively functionalized via Zhdankin azidoiodane reagent 832. Visible light was used to excite Ru(bpy)3Cl2 that cleaves the labile I–N3 bond, triggering the cascade of radical reactions that leads to the product formation (Scheme 83). The selectivity and compatibility of this reaction with different groups is underlined by the conversion of the dipeptide 831 to 833 in 30% yield.294 A related C–H azidation was performed by using tosyl azide as an alternative azide source with the help of 4-benzoylpyridine to promote the photocatalytic C–H cleavage in various cycloalkanes.295

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0087.jpg
Azidation of Tertiary Aliphatic C–H Bonds

Another example of the functionalization of unactivated C–H bonds is depicted in Scheme 84 making use of tosyl azide 842. The reaction needs the intermediacy of an oxygen radical center on a phosphate group, previously oxidized by the action of the mesityl acridinium photocatalyst 84–3. This allows the C–H to C–N3 conversion in menthol benzoate 841 to give azide 844 in a satisfying yield, with a regioselectivity favoring the more electron-rich tertiary position.296

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0088.jpg
Photocatalyzed C–H to C–N3 Conversion

The synthesis of amines is undoubtedly more challenging to be dealt with, relying on radical chemistry. However, several strategies were developed to effectively forge this fundamental functional group. A classic reaction for the synthesis of amine is the Curtius reaction that has the drawback in handling of potentially dangerous azides. A dual copper/photoredox catalytic approach mimicked this process for the obtainment of N-protected amines from the N-hydroxyphthalimide ester of cholic acid triacetate 851 (Scheme 85, see also Scheme 8). The alkyl radical was again formed by CuI-photocatalyzed reduction of 851, but this recombine with the CuII–phthalimide complex formed to release 852 (52% yield) by a formal decarboxylation process. A great variety of functional groups are compatible with this reaction including steroidal structures.297

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0089.jpg
Decarboxylative C–N Coupling in Cholic Acid Triacetate

A very interesting approach to synthesize β-aminoalcohols from the unfunctionalized alcohol 861 relies on the introduction of a radical relay chaperone to direct the C–H functionalization of the β position of the OH group (Scheme 86). Imidate radicals can be accessed via the photodecomposition of PhI(OAc)2. A transient sp2N-centered radical is generated from 862, which allows a 1,5-hydrogen atom transfer. A source of iodine promotes the formal transfer of an iodine radical to the β-position to the imidate, followed by cyclization to obtain 863 which can be promptly hydrolyzed to 864. The nature of the substituents on acetimidate 862 may affect the overall yield.298

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0090.jpg
Radical Relay Chaperone Strategy Driven by the Photodecomposition of PhI(OAc)2

Direct cross-coupling between alkyl carboxylic acids and nitrogen nucleophiles can be achieved by dual copper/photoredox catalysis through iodonium activation. The scope of the transformation is broad and applicable to a diverse array of nitrogen nucleophiles such as heterocycles, amides, sulfonamides, and anilines to give the corresponding C–N coupling product in excellent yields on short time scales (5 min to 1 h). The high regioselectivity obtained in late stage functionalization of complex pharmaceuticals such as Skelaxin 872 (to give 873 in 90% yield from 871, Scheme 87, see also Scheme 48) gave an idea of the importance of the approach.299

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0091.jpg
Late Stage Functionalization of Skelaxin

Similar strategies were explored for the synthesis of amines via C(sp3)–N cross-coupling combining a copper catalyst and the action of a photoredox catalyst by using anilines300 or benzophenone imines301 as nitrogen source. Hydroxylamines were instead formed under photoorganocatalytic conditions by reaction of carboxylic acids and nitrosoarenes.302

3.3. C–O Bond

The C–O bond formation is without doubt a prerogative of polar chemistry. However, there are examples of photochemically driven reactions making use of an alkyl radical for the introduction of different oxygen-containing functional groups. In Scheme 88, the nonenolizable ester 881 is transformed into 882 via a photochemically promoted decarboxylation of the NPhth-ester (see also Scheme 8) in the presence of Hantzsch ester to yield a tertiary radical. The intermediate is promptly quenched by TEMPO, affording 882 in 91% yield, in a multigram scale reaction.303

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0092.jpg
Decarboxylative Oxygenation of Phthalimide Esters

A similar reaction was employed to synthesize alkyl aryl ethers, given their importance in medicinal and agricultural chemistry. A tandem photoredox and copper catalysis approach allows the decarboxylative coupling of alkyl N-hydroxyphthalimide esters (NHPI) with phenols (892Scheme 89). Various NHPI esters of different drugs and natural products easily underwent a late-stage decarboxylative etherification. As an example, the chlorambucil derivative 891 was converted into the corresponding 2-MeO phenyl ether 893 in 49% yield.304

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0093.jpg
Decarboxylative C(sp3)-O Cross-Coupling

Following a similar strategy, carboxylates are converted into alcohols via a photocatalytic decarboxylative hydroxylation mediated by the mesityl acridinium salt. In this case, molecular oxygen is used as the oxidant, to promote the formation of the desired C–O bond. Since the reactions gave mainly a mixture of ketones and hydroperoxides, reduction in situ by sodium borohydride allowed the synthesis of alcohols in good yields.305 A decarboxylative hydroxylation may be carried out with the intermediacy of Barton esters that upon irradiation in oxygen-saturated toluene followed by treatment with P(OEt)3 afforded an alcohol intermediate for the total synthesis of Crotophorbolone.306 The more challenging oxidation of unactivated alkanes to alcohols or ketones can be achieved through a photoelectrochemical approach, as testified by the C–H bond activation of cyclohexane to prepare a mixture of cyclohexanone and cyclohexanol (the so-called KA oil) with high partial oxidation selectivity (99%) and high current utilization ratio (76%). The highest current ratio was obtained illuminating the solution with 365 nm wavelength.307 Decatungstate photocatalysis was efficiently applied to oxidize activated and unactivated C–H bonds. Taking advantage of a microflow reactor setup, a late stage regioselective CH2/C=O conversion in several natural compounds, such as artemisinin 901 to form artemisitone-9 902 was readily pursued even in a 5 mmol scale (Scheme 90).308

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0094.jpg
Late Stage Regioselective Carbonylation of Artemisinin

3.4. C-Halogen Bond

Halogenation of alkanes through a radical reaction under UV irradiation is one of the core pathways to chemically activate a paraffin. Industrially, chlorine gas is used to functionalize methane. A major drawback of the classical chain reaction using either Cl2 or Br2 under direct irradiation is the formation of di or polyhalogenated products. The application of microflow technology in combination with visible light irradiation (with an absorption maximum in the near UV at ca. 350 nm) allowed the monobromination of different alkanes with molecular bromine. High selectivity for the monobrominated compound and excellent overall yields (between 60 and 99%) could be achieved for secondary and tertiary alkanes, along with primary benzylic positions.309

Chlorination with molecular chlorine, on the other hand, suffers from the low yields of the reaction, typically around 50%, from the high concentrations of HCl generated in the process and from the toxicity of the chlorine gas itself. However, when Cl2 was generated by mixing NaClO with HCl and the chlorination took place under flow conditions, efficient C–H to C–Cl conversion resulted.310,311 A photochemical alternative using NaCl as chlorine source was developed.312 In the reaction, Cl2 was formed in situ by oxidation of the chloride anion with oxone. The monochlorination of cyclohexane 911 to give 912 could be obtained in 93% isolated yield thus overcoming the limitation of the classical chlorination process with chlorine gas (Scheme 91, see also Scheme 15).

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0095.jpg
Photoinduced Monochlorination of Cyclohexane

Fluorination is essential to modern medicinal chemistry, both as a viable way to insert radiotracers or to deactivate specific degradation pathways in drugs. Photochemistry is a reliable tool to achieve the fluorination of C–H bonds, following different strategies. Excited TBADT may formed a radical intermediate (from unactivated alkanes) that abstracts the fluorine atom from the labile N–F bond of the fluorinating agent N-fluorobenzenesulfonimide (NFSI). An N-centered radical resulted which closes the radical cycle oxidizing the reduced photocatalyst. Acetate 921 was fluorinated in 40% yield following this procedure to yield 922 (Scheme 92). The reaction applied to sclareolide, however, was not selective and gave a mixture of fluorinated regioisomers (68% overall yield).313

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0096.jpg
TBADT-Catalyzed Fluorination of Alkanes

Following a similar reaction scheme, uranyl acetate was employed in combination with NFSI to promote the fluorination of secondary alkanes but poorly on the benzylic positions. Indeed, in the absence of an aromatic scaffold, the excited U=O abstract a hydrogen atom through HAT, while the presence of an aromatic ring deactivated the excited state of the catalyst via exciplex formation preventing the fluorination to occur.314 Acetophenone in its excited state promoted the hydrogen abstraction of secondary alkanes, with the advantage that a common CFL housebulb can be used to promote an efficient conversion, using Selectfluor as the fluoride source.315 In this case, the authors irradiated the tail of the n-π* absorption band of the ketone which can be found in the visible region due to the high concentration of the photocatalyst present in solution. N-Alkyl phthalimides having an alkyl chain linked to the nitrogen was fluorinated by using Selectfluor under photocatalyst-free conditions. An exciplex was supposed to be formed between the reagents and it was proposed that the C–F bond formation took place concomitantly with hydrogen atom abstraction with the nitrogen radical of the fluorinating agent.316

A considerable regioselectivity in the fluorination reaction can be achieved using carboxylates as alkyl radical precursors and again Selectfluor as a fluorinating reagent. The reaction is possibly initiated by reduction of Selectfluor 932 by means of Ir[dF(CF3)ppy]2(dtbbpy)PF6. Fluorination of different carboxylic acids can be achieved in a very high yields (between 70 and 99%), and 931 was readily converted into 933 in 90% yield (Scheme 93).317 In case of unactivated primary substrates, a prolonged irradiation (12–15 h) was mandatory to achieve a high conversion of the substrate.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0097.jpg
Regioselective Fluorination of Carboxylates

An interesting case is the fluorination of compounds having the MOM group to direct the halogenation event. In this case, the PC oxidized an imidine base (DBN) that acted as hydrogen atom abstractor of the dioxolanyl group in compound 941 (Scheme 94). The resulting α,α-dioxy radical 942 released an alkyl radical (upon formiate loss) that was fluorinated by Selectfluor. This metal-free approach again used visible light and is particularly successful when applied to tertiary alkyl ethers to give sterically hindered alkyl fluorides (e.g., 943).318

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0098.jpg
DBN-Mediated HAT in C–F Bond Formation

Interestingly, fluorination of carboxylates with Selectfluor was also reported to occur under heterogeneous photocatalytic conditions, using titania as photocatalyst to promote the oxidation of the carboxylate anion.319 Fluorination and chlorination of nitriles and ketones could be obtained starting from oximes, using Selectfluor and NCS as halogen sources, respectively. With this methodology, γ-functionalization of ketones and a complex photoinduced ring-opening/halogenation of oximes via the intermediacy of an iminyl radical was pursued. The C=N moiety of the reagent (e.g., 951) was preserved in the products (95–2a,b) in its oxidized nitrile form (Scheme 95). Several natural products could be functionalized following this methodology, such as androsterone (95–3a,b) and camphor (95–4a,b) derivatives (Scheme 95, see also Scheme 20).320

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0099.jpg
Ring-Opening Halogenation of Oximes

Alcohols were converted into their corresponding pyruvates that upon irradiation in the presence of an IrIII photocatalyst with blue LEDs released an alkyl radical prone to be chlorinated by 2,2,2-trichloroacetate as the chlorine atom source. A series of secondary and tertiary chlorides could be obtained in good to excellent yields.321

An Ir-based photocatalyst was used to promote bromination of carboxylic acid (961) with bromomalonate as brominating agent (Scheme 96).322 The acids used in this work were likewise converted into the corresponding alkyl chlorides and iodides in the presence of the corresponding N-halosuccinimides.322

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0100.jpg
Bromomalonate as Brominating Agent

A radical relay strategy was employed to synthesize gem-diiodides through successive intramolecular 1,5-HAT processes and iodine trapping. Indeed, the excitation with visible light of an N–I imidate group, formed in situ from the reaction of a trichloroacetimidate with PhI(OAc)2 as an iodine source, allowed the synthesis of a small library of gem di-I compounds in good yields. As an example, the cholic acid derivative 971 has been converted to its corresponding di-iodo derivative 972 in 71% yield (Scheme 97). Moreover, the authors could also achieve a dibromination using NaBr and TBABr and visible light, while only monochlorination is reported when NaCl, TBACl, and UV light were adopted.323

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0101.jpg
Gem di-Iodination of a Cholic Acid Derivative

3.5. C–S or C–Se Bonds

Alkyl radicals were sparsely used for the unusual introduction of a SCF2X (X = F, H) or an SAr moiety in an organic compound. The introduction of a SCF2X group has recently sparked attention due to the remarkable hydrogen donor nature of the group when X = H, making it the lipophilic surrogate for OH or NH groups.324 On the other hand, the trifluoromethylthio group increases the metabolic stability and the lipophilicity of drugs.

One strategy for the introduction of a SCF2X group is the photocatalyzed (by IrIII PC) oxidation of alkyl carboxylates via visible light irradiation in the presence of PhthN-SCF2H (982) as the sulfur donor. Indeed, 981 was converted into 983 in high yields (Scheme 98). The reaction was sustained by the stability of the imidyl radical liberated in the process, that was able to promote a chain reaction oxidizing a further carboxylate group. Indeed, the quantum yield for the reaction was found to be 1.7.325 Bis-methyltiolation was observed in different cases, possibly due to HAT triggered by an intermediate of the reaction, presumably PhthN and following transfer of SCF2X from the reactant. To avoid the formation of byproducts either mesitylene or 3-(methyl) toluate were added as sacrificial hydrogen donors.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0102.jpg
Difluorothiomethylation of Carboxylic Acids

An interesting follow-up for this methodology from the same group made use of the hydrogen atom transfer process previously reported as detrimental for the reaction yield. In fact, when using an aryl carboxylate instead of an aliphatic one, the carboxyl radical that is formed upon electron transfer with the excited Ir catalyst is now stable enough to act as a hydrogen abstractor, selectively targeting secondary or tertiary H in alkyl chains. Also, in this case, PhthN-SCF2X acted as the sulfur source. The conversion of ambroxide 991 to its trifluorothiomethyl derivative 992 proceeded with 95% yield (Scheme 99).326

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0103.jpg
Trifluoromethylthiolation of Ambroxide

A photocatalyst-free decarboxylative arylthiation took place by mixing an N-acyloxyphthalimide (e.g., 1002) in the presence of an aryl thiol (1001) under basic conditions (by Cs2CO3) upon visible light irradiation. In this case, a SET between 1001 and 1002 caused the formation of 1002•– along with thiyl radical 1003 (that easily dimerized to disulfide 1004). Trapping of the resulting cyclohexyl radical (by loss of PhthN from 1002•–) with 1004 afforded alkylaryl sulfide 1005 in 89% yield (Scheme 100).327

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0104.jpg
Arylthiation of N-Acyloxyphthalimides

The most widely used reaction for the C–S bond synthesis requires the incorporation of sulfur dioxide by using DABSO (DABCO(SO2)2) as its surrogate as depicted in Scheme 101.328 Thus, excited mesityl acridinium salts promoted the oxidation of an alkyl-BF3K salt that generated a nucleophilic radical able to react with DABSO. The sulfonyl radical intermediate formed has been employed in a three-component reaction with electron poor olefins (e.g., a vinyl piridine 1011, Scheme 101a)329 or an alkyne (phenyl acetylene, Scheme 101b),330 affording alkyl sulfone (1012) or (E)-vinyl sulfone (1013), respectively.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0105.jpg
Sulfonylation of (a) Styrenes and (b) Alkynes

Alternatively, alkyl iodides can be used to react with olefins decorated with EWGs and DABSO to generate a broad range of alkyl sulfones.331 A very similar strategy was implemented by the same authors using differently substituted Hantzsch esters as alkyl radical precursors, upon irradiation in the presence of Eosin Y.332 In the latter case, the sulfonyl radical added onto vinyl azides and, after releasing of molecular nitrogen, an imidyl radical resulted which reacted with the reduced photocatalyst, forming an anion that is easily protonated. After a tautomeric equilibrium, (Z)-2-(alkylsulfonyl)-1-arylethen-1-amines were formed, with good regioselectivity and complete control over the configuration of the double bond.332

Cyclobutanone oximes can be reduced via photocatalytic means in the presence of Ir(dtbbpy)(ppy)2PF6 to form γ-cyanoalkyl radicals after radical fragmentation. In this process, a vinyl sulfone was used having the dual role of radical acceptors and SO2 source, allowing the synthesis of β-ketosulfones or allylsulfones through a radical transfer mechanism.333

3.6. C–H Bond

Classical radical reductive dehalogenation is one of the most successful reactions based on tin chemistry.21 Photocatalysis and photochemistry propose a milder and more environmentally friendly alternative to this process, via different strategies. As an example, fac-Ir(ppy)3 was used to convert alkyl iodides 1021 into their corresponding alkyl radicals using Hantzsch ester or HCO2H as the hydrogen atom source for the HAT process that drives the reaction to the formation of 1022 (Scheme 102). The authors optimized their procedure by using tributylamine as the sacrificial electron donor to reduce the oxidized form of the catalyst and restore the catalytic cycle.334 A variation of this protocol using p-toluenethiol, DIPEA, and fac-Ir(mppy)3 was used to synthesize D-albucidin.335

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0106.jpg
Photocatalyzed Reduction of Alkyl Iodides

Other catalytic systems were proved to be competent in the reduction of halides. In particular, unactivated aryl and alkyl bromides could be reduced using [Ir(ppy)2(dtbbpy)]PF6 in combination with TTMSS as a reducing agent. The mild conditions typical of the reaction were critical to obtain both the mono and the bis reduction of a gem-dibromocyclopropane in a selective fashion.336

Alkyl iodides and bromides were reduced under metal-free conditions via irradiation of 4-carbazolyl-3-(trifluoromethyl)-benzoic acid as the photocatalyst and 1,4 cyclohexadiene as sacrificial hydrogen donor.337 The reduction of C–X bonds to C–H bonds can take place under photocatalyst-free conditions by PET reactions between the halide and an amine as sacrificial reductant. In this way, adamantane was obtained in 95% yield by photochemical reduction of 1-bromoadamantane.338 Borohydride-mediated radical photoreduction of alkyl halides (iodides, bromides, and chlorides) is another valuable tool for the formation of a C–H bond.339

The C–H bond formation could be achieved via a hydrodecarboxylation of carboxylic acids. In fact, carboxylic acid 1031 could be reduced in 97% yield to 1032 by generating the corresponding carboxyl radical through excitation of an acridinium photocatalyst with 450 nm LEDs, in the presence of 10% mol of (PhS)2 (Scheme 103). The authors achieved good yields in the decarboxylation of different carboxylic acids. Most notably they succeeded in the double reduction of doubly substituted malonic acids, although with the necessity of longer irradiation times and higher catalyst loading to compensate for the increased amount of substrate to be reduced.340

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0107.jpg
Hydrodecarboxylation of Carboxylic Acids

The challenging reduction of alcohols to the corresponding alkane can take place via functionalization of the OH group to form an O-thiocarbamate. This compound is the substrate of a photocatalyzed Barton-McCombie deoxygenation in combination with Ir(ppy)3 and DIPEA under an oxidative quenching. Accordingly, the xylofuranose derivative 1041 was cleanly reduced to 1042 in 70% yield by maintaining the benzoyl group in position 5 (Scheme 104).93 The reaction was studied mostly on secondary alcohol derivatives being another interesting alternative to the usual tin-mediated reaction.22

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0108.jpg
Reduction of a Xylofuranose Derivative

An alternative pathway to reduce the hydroxy function required a more sophisticated functionalization of the OH group making use of two consecutive photochemical reactions. Conducting the reaction in CBr4 under UVA irradiation, the hydroxy groups of a series of primary alcohols were converted into their bromides and then subjected to a one-pot photoreduction mediated by the dimeric gold complex [Au2(dppm)2]Cl2 in the presence of DIPEA.341

4. Formation of a Ring

4.1. Three/Four-Membered Rings

In this last section, selected examples will be given when a photogenerated alkyl radical is used for the construction of a ring. Scheme 105 shows one example of formation of a three-membered ring. 1,1-Disubstituted cyclopropanes 105–3a–d were obtained through the addition of an alkyl radical (from silicate 1051) onto homoallylic tosylates 105–2a–d. The trick here is a radical/polar crossover process where the reduction of the benzyl radical adducts to benzyl anions (by SET with the reduced form of the photoorganocatalyst 4-CzIPN) followed by intramolecular substitution gave the three-membered ring (Scheme 105).342 The versatility of the method was demonstrated by using alkyl trifluoroborates or 4-alkyldihydropyridines as radical precursors and a good tolerance of various functional groups.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0109.jpg
Synthesis of 1,1-Disubstituted Cyclopropanes

A related approach was adopted for the construction of cyclobutanes.102 Here, the alkyl radical was formed by easily oxidizable electron-rich alkyl arylboronate complexes and added to an iodide-tethered alkene such as methyl 5-iodo-2-methylenepentanoate. However, changing the length of the chain in the haloalkyl alkenes led to the synthesis of three-, five-, six-, and seven-membered rings.102

4.2. Five-Membered Rings

Five-membered ring is one of the privileged structures accessible via photogenerated alkyl radicals. A common approach is the cyclization onto an alkyne to form an exocyclic double bond as exemplified in Scheme 106. In most cases, an alkyl halide is reduced by an excited photocatalyst and the resulting radical cyclizes in a 5-exo dig fashion to form the desired alkene. When using a dimeric gold complex 1065 the reaction of alkyl bromide 1061 generates diester 1062 in 93% yield (Scheme 106a).343 Cyclopentanes were likewise formed starting from an unactivated alkyl iodide that underwent an intramolecular radical closure by using a strong reductant in the excited state (Ir(ppy)2(dtb-bpy)PF6). The iodine atom, however, was incorporated in the final product forming an alkenyl iodide.344 The same metal-based photocatalyst was effective to induce a visible light-promoted preparation of five-membered heterocycles (Scheme 106b). The cyclization step was applied on a Ueno–Stork reaction starting from 2-iodoethyl propargyl ethers (e.g., 106–3a,b) to construct a tetrahydrofuran ring (in 106–4a,b).345 The examples described in Scheme 106 required an amine as a sacrificial donor. However, amines can be used as efficient reducing agents by a PET reaction with excited alkynyl halides. The resulting photocyclization may then be carried out under metal-free conditions and in a flow photomicroreactor providing the preparation of five-membered rings in a 4 g scale.346

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0110.jpg
Metal-Photocatalyzed Synthesis of Cyclopentanes from Alkyl Halides

The dehalogenation/cyclization strategy was explored even under heterogeneous conditions by using platinum nanoparticles on titania (PtNP@TiO2) as the photoredox catalyst. The pyrrolidine scaffold was then obtained by reaction of N-(2-iodoethyl)-4-methyl-N-(prop-2-yn-1-yl)benzenesulfonamide under irradiation (DIPEA as sacrificial donor).347

As an alternative, a biphasic system may be adopted (Scheme 107). In fact, a polyisobutylene-tagged fac-Ir(ppy)3 complex (Ir(ppy)2(PIB-ppy)) soluble in heptane was prepared. The substrate 1071 along with the reagents were soluble in a MeCN phase. However, heating at 85 °C allowed the two phases to mix. Preparation of tetrahydrofuran derivative 1072 was then accomplished in continuous flow in a satisfying yield with an automatic recovery and reuse of the catalyst (Scheme 107).348

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0111.jpg
Photocatalyzed Synthesis of Tetrahydrofurans in Flow

Alkyl N-hydroxyphthalimide esters were used as alkylation reagents in the functionalization of alkenoic acid 1082 (Scheme 108). The alkyl radical added onto the double bond, and the resulting benzyl radical was oxidized to a benzyl cation readily trapped by water and cyclization of the resulting hydroxy acid gave alkyl-substituted lactones 108–3a–e in moderate yields.349

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0112.jpg
Photocatalyzed Lactonization of Alkenoic Acids

Alkyl N-hydroxyphthalimide esters were exploited for the photocatalyzed (by a RuII complex) alkylation of N-arylacrylamides that caused the cyclization of the adduct radical onto the phenyl ring to afford 3,3-dialkyl substituted oxindoles.350 Moreover, the same radical precursors have been used for the derivatization of alkynylphosphine oxides under metal- and oxidant-free conditions to form benzo[b]phospholes in very good yields.351

A five-membered ring may be accessed via late-stage C(sp3)-H functionalization in N-chlorosulfonamides 1091 (Scheme 109a). The IrIII-photocatalyzed reduction of 1091 induced the elimination of the chloride anion along the formation of a N-centered radical prone to abstract a hydrogen atom from a remote position to afford an alkyl radical. Oxidation of this radical to the cation followed by incorporation of the chloride anion gave the corresponding chloride 1092 that upon treatment with solid NaOH formed pyrrolidine 1093 by an intramolecular nucleophilic substitution.352 The mildness of the process allowed the application onto biologically important (−)-cis-myrtanylamine and (+)-dehydroabietylamine derivatives.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0113.jpg
Photochemically Induced Synthesis of Pyrrolidines

A related intramolecular 1,5-HAT transfer was induced in N-tosyl amides 109–4a–e. In this case, the haloamide is formed in situ by iodination by reaction of 109–4a-e with iodine (obtained by oxidation of iodide anion by PhI(OAc)2). The excess of iodine allowed for tuning the amount of iodine released in solution by forming the triiodide anion. Then, visible light irradiation of the mixture induced the cyclization to give N-tosylpyrrolidines 109–5a–e (Scheme 109b).353

Even primary nonactivated sp3-hybridized positions were functionalized again by a remote intramolecular radical 1,5-hydrogen abstraction in γ-bromoamides to produce several γ-lactones in a one-pot fashion.354 Trifluoroethyl amides were found useful as the directing group increasing the efficiency of the hydrogen abstraction process.

It is also possible to incorporate more than one heteroatom in the ring starting from benzyl amine 1102 and unactivated bromides 110–1a–e (Scheme 110). Compound 1102 incorporates CO2 (with the help of the base TBD), and the resulting carbamate underwent attack by an alkyl radical photogenerated by reaction of 110–1a–e and an excited Pd0 photocatalyst (Pd(PPh3)4). Ring closing yielded valuable 2-oxazolidinones 110–3a–e under very mild conditions and easy scalability.355

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0114.jpg
Three-Component Synthesis of 2-Oxazolidinones

4.3. Six-Membered or Larger Rings

Different approaches were devised to form a six-membered ring even cointaining heteroatoms. A cyclohexane ring was constructed by ring opening of an iminyl radical by IrIII-photocatalyzed reduction of a 3-phenyl O-acyl oxime (e.g., 1111) to give radical 1112 that upon addition onto unsaturated esters 111–3a–e and ensuing cyclization led to cyanoalkylated 1,2,3,4-tetrahydrophenanthrenes (111–4a–e, Scheme 111).356

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0115.jpg
Photocatalyzed Preparation of 1,2,3,4-Tetrahydrophenanthrenes

Six-membered rings have been likewise obtained by ring expansion in cycloalkanone derivatives. This expansion was caused by the photocatalyzed decarboxylation of α-(ω-carboxyalkyl) β-keto esters, followed by an exo-trig cyclization of the resulting radical onto the carbonyl group that ultimately led to the one-carbon expanded cycloalkanones by β-cleavage.357

Reduction of indoles having an unactivated haloalkane chain is a useful approach to construct a ring. As an example, bromo derivatives 112–1a,b were reduced by a AuI photocatalyst and radical cyclization onto the heteroaromatic ring afforded 6,7,8,9-tetrahydropyrido[1,2-a]indoles 112–2a,b in excellent yield (Scheme 112a).358 Interestingly, changing the reaction conditions and starting from N-(2-iodoethyl)indoles 112–3a,b in place of 112–1a,b in the presence of Michael acceptors 112–4a–c caused a dearomatizative tandem [4 + 2] cyclization to deliver tri- and tetracyclic benzindolizidines 112–5aa–bc with high diastereoselectivity and yield (Scheme 112b).359

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0116.jpg
Intramolecular C–C Bond Formation in Indoles

The phenanthridine core is one of the elective scaffolds to be prepared by using a cyclization step induced by photogenerated alkyl radicals. Scheme 113 illustrated a representative case where an alkyl radical added onto a vinyl azide 1132, and after nitrogen loss the resulting iminyl radicals 113–3a-c yielded phenanthridines 113–4a-c by ring closure.360 The method has several advantages including metal-free conditions (a dye as a POC) an excellent functional group tolerance and a broad substrate scope.

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0117.jpg
Photocatalyzed Synthesis of Functionalized Phenanthridines

An alternative way to prepare phenanthridines is by having recourse to photoredox gold catalysis employing bromoalkanes as alkyl radical source. In this case, radicals attack a biaryl isonitrile thus forming a sp2-hybridized radical that readily cyclizes upon the pendant arene.361 Aryl isocyanides (e.g., 114–2a–d) were largely used for the construction of heterocycles such as pyrrolo[1,2-a]quinoxalines 114–4a–d. PhenyliodineIII dicarboxylate 1141 was used for the incorporation of the cyclohexyl group both in batch and flow under IrIII-photocatalyzed conditions (Scheme 114, see also Scheme 49).362

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0118.jpg
Photoredox Preparation of Pyrrolo[1,2-a]quinoxalines

The photocatalyzed insertion of SO2 into an unactivated C(sp3)-H bond was designed to prepare 1,2-thiazine 1,1-dioxide derivatives under uncatalyzed conditions. In fact, visible light irradiation of the complex between an electron-poor O-aryl oxime and DABCO·(SO2)2 releases an iminyl radical that upon 1,5-HAT, SO2 incorporation and cyclization gave the hoped-for heterocycle in a satisfying yield.363

In rare instances a ring larger than six may be constructed. By using the approach depicted in Scheme 115, it was possible to pursue a late stage functionalization on ursolic acid (a compound having excellent pharmaceutical activity). Accordingly, the NHPI ester of ursolic acid acetate (1152) underwent a radical addition cascade by a photocatalyzed reaction with acrylamide-tethered styrene (1151) with the intermediacy of radical 1153. As a result, the benzazepine unit was incorporated in the end compound 1154 combining two privileged bioactive scaffolds.364

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0119.jpg
Late Stage Functionalization of Ursolic Acid

5. Conclusions and Outlook

This review provides a concise and up-to-date selection of modern methods to generate alkyl radicals via photochemistry and photocatalysis. The effort and interest of the chemical community in developing and applying these new methods is witnessed by the rapid increase in the number of articles devoted to this topic that appeared in the literature in the last two decades. Indeed, the rediscovery of photocatalysis and the renaissance of visible light-driven processes have contributed to elevate radical chemistry from the isolated (yet efficient) niche of the tyrannical organotin compounds to a vast plethora of methodologies that relies on more environmental benign compounds. The facile synthesis of the precursors necessary for these transformations, along with the readily available setups (a vast number of reactions can occur by simple irradiation with visible LEDs), made radical chemistry approachable, appointing the photon as the agent of this revolutionary democracy.

Photocatalysis has reached the stage of maturity; however, we are still far from the statement of Ciamician envisioning “industrial colonies without smoke [···] forests of glass tubes [···]; inside of these will take place the photochemical processes that hitherto have been the guarded secret of the plants, but that will have been mastered by human industry which will know how to make them bear even more abundant fruit than nature, for nature is not in a hurry and mankind is”.365 New practical methods and theoretical assumptions are needed to foster the revolution that has just started. A promising approach makes use of the upconversion of reductants to generate strongly reductive species, but the method was not applied so far to alkyl radicals.366 This phenomenon can be exploited, for example, if the reaction of a radical anion R•– to give P•– is less exoergonic (see the ΔG•– value in Figure Figure44A) than its neutral counterpart (ΔG, referred to RP conversion). The difference between these two free energies defines the upconversion energy (ΔGup = ΔG•– – ΔG). The high quantum yields associated with the transformation of R into P in Figure Figure44A (Φ = 44) were attributed to the presence of electrocatalytic cycles propagated by P•–, which is able to transfer an electron to the reactant, closing the catalytic cycle. This phenomenon is attributed to P•– being a better reductant than R•–, due to the diminished conjugation (Figure Figure44A).

An external file that holds a picture, illustration, etc.
Object name is cr0c00278_0004.jpg

(A) Upconversion of the reducing power of the intermediates in a photocatalytic/photoinitiated cyclization. (B) Two pathways to employ the photoelectrocatalytic strategy: either promoting a single electron transfer with photocatalysis first and a second one with electrocatalysis or vice versa.

The novel approach granted by the merging of homogeneous photocatalysis with electrocatalysis (see Figure Figure44B) is surfacing as the new challenge in this constantly evolving topic.367369

Joining the almost unlimited potential of these two interchangeable fields of research would open unprecedented scenarios in chemical synthesis, allowing one to tweak the reactivity of intermediates and excited state species at will, walking on the path carved by the institution of the photon democracy.

Acknowledgments

S.C. gratefully acknowledges MIUR (Ministry of University and Research) for the support.

Glossary

Abbreviations

Acr+Mes9-mesitylene-10-methylacridinium
alkyl-DHPs4-alkyl-1,4-dihydropyridines
BDEbond dissociation energies
BI-OAcbenziodoxole acetate
BI–OHhydroxylbenziodoxole
4CzIPN1,2,3,5-tetrakis(carbazol-9-yl)-4,6-dicyanobenzene
DABSO1,4-diazabicyclo[2.2.2]octane bis(sulfur dioxide)
DBADdi-tert-butylazodicarboxylate
DCN1,4-dicyanonaphthalene
DIADdiisopropyl azodicarboxylate
DMAcN,N-dimethylacetamide
DMAP4-dimethylaminopyridine
EYEosin Y
HAThydrogen atom transfer
d-HATdirect hydrogen atom transfer reaction
i-HATindirect hydrogen atom transfer reaction
HDAChistone deacetylase
HEHantzsch ester
Ir[dF(CF3)ppy]2(dtbbpy)2+bis(2-(2,4-difluorophenyl)-5-trifluoromethylpyridine)(ditert-butylbipyridine)iridium
fac-Ir(ppy)3fac-(tris(2,2′-phenylpyridine))iridium
[Ir(ppy)2(dtbbpy)]+bis(2-phenylpyridine) (di-tert-butylbipyridine)iridium
LBLewis base
NFSIN-fluorobenzenesulfonimide
NHPIalkyl N-hydroxyphthalimide esters (Phth-ester)
PCphotocatalyst
PCETproton-coupled electron transfer
PFBI–OHperfluorohydroxylbenziodoxole
POCphotoorganocatalysts
RFTAriboflavin tetraacetate
RuII(bpy)32+tris(2,2′-bipyridine)ruthenium
SETsingle electron transfer
TBACNtetrabutylammonium cyanide
TBADTtetrabutylammonium decatungstate
TBD1,5,7-triazabicyclo[4.4.0]dec-5-ene
TMHDtetramethylheptanedione
TMSCNtrimethylsilyl cyanide
TTMSStris(trimethylsilyl)silane
XAThalogen atom transfer reaction

Biographies

•Â 

Maurizio Fagnoni is currently an Associate Professor at the PhotoGreen Lab (Department of Chemistry, University of Pavia, Italy). His academic and professional background is in organic photochemistry and his activity has always been focused on the exploration of the photochemistry of organic molecules and the attending applications in various fields. The photochemical generation of intermediates, e.g., radicals and cations and radical ions by photochemical means, is the main topic of his research. Particular attention has been given to the significance of such mild synthetic procedures in the frame of the increasing interest for sustainable/green chemistry. He was the recipient in 2019 of the “Elsevier Lectureship Award” from the Japanese Photochemical Association. He was recently coeditor of the book Photoorganocatalysis in Organic Synthesis (World Scientific, 2019). Since 2019, he has been the President of the Didactic Council in Chemistry of the University of Pavia.

•Â 

Stefano Crespi received his Ph.D. in 2017 at the University of Pavia (Italy) under the supervision of Maurizio Fagnoni. He won a two-year fellowship as a Post-Doc in the same University focusing on the study of novel heteroaryl azo photoswitches. He joined the workgroup of Burkhard König at the University of Regensburg, where he studied new scaffolds based on heteroaryl azo dyes and novel photocatalytic transformations. In 2019, he moved to Groningen to work on molecular motors in the group of Ben Feringa as a Marie Skłodowska-Curie fellow. His research interests lie in the combination of reaction design in organic (photo)chemistry with computational models.

Author Contributions

S.C. and M.F. discussed and contributed to the final manuscript. M.F. conceived the original idea.

Notes

The authors declare no competing financial interest.

References

  • Radicals in Organic Synthesis; Renaud P., Sibi M. P., Eds.; Wiley-VCH, 2001. [Google Scholar]
  • Encyclopedia of Radicals in Chemistry, Biology and Materials; Chatgilialoglu C., Studer A., Eds.; John Wiley & Sons, Ltd: Chichester, UK, 2012. [Google Scholar]
  • Zard S. Z.Radical Reactions in Organic Synthesis; Oxford University Press, 2003. [Google Scholar]
  • Yan M.; Lo J. C.; Edwards J. T.; Baran P. S. Radicals: Reactive Intermediates with Translational Potential. J. Am. Chem. Soc. 2016, 138, 12692–12714. 10.1021/jacs.6b08856. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Studer A.; Curran D. P. Catalysis of Radical Reactions: A Radical Chemistry Perspective. Angew. Chem., Int. Ed. 2016, 55, 58–102. 10.1002/anie.201505090. [PubMed] [CrossRef] [Google Scholar]
  • Kuivila H. G.; Menapace L. W. Reduction of Alkyl Halides by Organotin Hydrides 1, 2. J. Org. Chem. 1963, 28, 2165–2167. 10.1021/jo01044a001. [CrossRef] [Google Scholar]
  • Corey E. J.; Suggs J. W. Method for Catalytic Dehalogenations via Trialkyltin Hydrides. J. Org. Chem. 1975, 40, 2554–2555. 10.1021/jo00905a039. [CrossRef] [Google Scholar]
  • Stork G.; Sher P. M. A Catalytic Tin System for Trapping of Radicals from Cyclization Reactions. Regio- and Stereocontrolled Formation of Two Adjacent Chiral Centers. J. Am. Chem. Soc. 1986, 108, 303–304. 10.1021/ja00262a024. [CrossRef] [Google Scholar]
  • Giese B. Formation of CC Bonds by Addition of Free Radicals to Alkenes. Angew. Chem., Int. Ed. Engl. 1983, 22, 753–764. 10.1002/anie.198307531. [CrossRef] [Google Scholar]
  • Giese B.; González-Gómez J. A.; Witzel T. The Scope of Radical CC-Coupling by the “Tin Method”. Angew. Chem., Int. Ed. Engl. 1984, 23, 69–70. 10.1002/anie.198400691. [CrossRef] [Google Scholar]
  • Giese B.; Dupuis J. Diastereoselective Syntheses of C-Glycopyranosides. Angew. Chem., Int. Ed. Engl. 1983, 22, 622–623. 10.1002/anie.198306221. [CrossRef] [Google Scholar]
  • Giese B. Syntheses with Radicals-C-C Bond Formation via Organotin and Organomercury Compounds [New Synthetic Methods (52)]. Angew. Chem., Int. Ed. Engl. 1985, 24, 553–565. 10.1002/anie.198505531. [CrossRef] [Google Scholar]
  • Giese B.; Meister J. Die Addition von Kohlenwasserstoffen an Olefine Eine Neue Synthetische Methode. Chem. Ber. 1977, 110, 2588–2600. 10.1002/cber.19771100717. [CrossRef] [Google Scholar]
  • Galli C.; Pau T. The Dehalogenation Reaction of Organic Halides by Tributyltin Radical: The Energy of Activation vs. the BDE of the C-X Bond. Tetrahedron 1998, 54, 2893–2904. 10.1016/S0040-4020(98)83025-1. [CrossRef] [Google Scholar]
  • Luo Y.-R.Comprehensive Handbook of Chemical Bond Energies; CRC Press: Boca Raton, 2007. [Google Scholar]
  • Hale K. J.; Manaviazar S.; Watson H. A. The O-Directed Free Radical Hydrostannation of Propargyloxy Dialkyl Acetylenes with Ph3SnH/Cat. Et3B. A Refutal of the Stannylvinyl Cation Mechanism. Chem. Rec. 2019, 19, 238–319. 10.1002/tcr.201700104. [PubMed] [CrossRef] [Google Scholar]
  • Gonzalez-Rodriguez E.; Abdo M. A.; dos Passos Gomes G.; Ayad S.; White F. D.; Tsvetkov N. P.; Hanson K.; Alabugin I. V. Twofold π-Extension of Polyarenes via Double and Triple Radical Alkyne Peri -Annulations: Radical Cascades Converging on the Same Aromatic Core. J. Am. Chem. Soc. 2020, 142, 8352–8366. 10.1021/jacs.0c01856. [PubMed] [CrossRef] [Google Scholar]
  • Curran D. P. The Design and Application of Free Radical Chain Reactions in Organic Synthesis. Part 1. Synthesis 1988, 1988, 417–439. 10.1055/s-1988-27600. [CrossRef] [Google Scholar]
  • Jasperse C. P.; Curran D. P.; Fevig T. L. Radical Reactions in Natural Product Synthesis. Chem. Rev. 1991, 91, 1237–1286. 10.1021/cr00006a006. [CrossRef] [Google Scholar]
  • Ryu I.; Sonoda N.; Curran D. P. Tandem Radical Reactions of Carbon Monoxide, Isonitriles, and Other Reagent Equivalents of the Geminal Radical Acceptor/Radical Precursor Synthon. Chem. Rev. 1996, 96, 177–194. 10.1021/cr9400626. [PubMed] [CrossRef] [Google Scholar]
  • Baguley P. A.; Walton J. C. Flight from the Tyranny of Tin: The Quest for Practical Radical Sources Free from Metal Encumbrances. Angew. Chem., Int. Ed. 1998, 37, 3072–3082. 10.1002/(SICI)1521-3773(19981204)37:22<3072::AID-ANIE3072>3.0.CO;2-9. [PubMed] [CrossRef] [Google Scholar]
  • McCombie S. W.; Motherwell W. B.; Tozer M. J.. The Barton-McCombie Reaction. In Organic Reactions; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2012; pp 161–432. [Google Scholar]
  • Davies A. G.Organotin Chemistry, Second, Completely Revised and Updated ed.; John Wiley & Sons, Ltd, 2004. [Google Scholar]
  • Sax’s Dangerous Properties of Industrial Materials; Lewis R. J., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2004. [Google Scholar]
  • Zard S. Z.Xanthates and Related Derivatives as Radical Precursors. In Encyclopedia of Radicals in Chemistry, Biology and Materials; Chatgilialoglu C., Studer A., Eds.; John Wiley & Sons, Ltd: Chichester, UK, 2012. [Google Scholar]
  • Sato R.; Okamoto R.; Ishizuka T.; Nakayama A.; Karanjit S.; Namba K. Microwave-Assisted Tertiary Carbon Radical Reaction for Construction of Quaternary Carbon Center. Chem. Lett. 2019, 48, 414–417. 10.1246/cl.190040. [CrossRef] [Google Scholar]
  • Chatgilialoglu C.; Ferreri C.; Landais Y.; Timokhin V. I. Thirty Years of (TMS)3SiH: A Milestone in Radical-Based Synthetic Chemistry. Chem. Rev. 2018, 118, 6516–6572. 10.1021/acs.chemrev.8b00109. [PubMed] [CrossRef] [Google Scholar]
  • Gagosz F.; Moutrille C.; Zard S. Z. A New Tin-Free Source of Amidyl Radicals. Org. Lett. 2002, 4, 2707–2709. 10.1021/ol026221m. [PubMed] [CrossRef] [Google Scholar]
  • Tsai L.-C.; You M.-L.; Ding M.-F.; Shu C.-M. Thermal Hazard Evaluation of Lauroyl Peroxide Mixed with Nitric Acid. Molecules 2012, 17, 8056–8067. 10.3390/molecules17078056. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Quiclet-Sire B.; Zard S. Z. Powerful Carbon-Carbon Bond Forming Reactions Based on a Novel Radical Exchange Process. Chem. - Eur. J. 2006, 12, 6002–6016. 10.1002/chem.200600510. [PubMed] [CrossRef] [Google Scholar]
  • Darmency V.; Renaud P.. Tin-Free Radical Reactions Mediated by Organoboron Compounds. In Radicals in Synthesis I; Springer-Verlag: Berlin, 2006; pp 71–106. [Google Scholar]
  • Studer A.; Amrein S. Tin Hydride Substitutes in Reductive Radical Chain Reactions. Synthesis 2002, 2002, 835–849. 10.1055/s-2002-28507. [CrossRef] [Google Scholar]
  • Walton J. C.; Studer A. Evolution of Functional Cyclohexadiene-Based Synthetic Reagents: The Importance of Becoming Aromatic. Acc. Chem. Res. 2005, 38, 794–802. 10.1021/ar050089j. [PubMed] [CrossRef] [Google Scholar]
  • Studer A.; Amrein S. Silylated Cyclohexadienes: New Alternatives to Tributyltin Hydride in Free Radical Chemistry. Angew. Chem., Int. Ed. 2000, 39, 3080–3082. 10.1002/1521-3773(20000901)39:17<3080::AID-ANIE3080>3.0.CO;2-E. [PubMed] [CrossRef] [Google Scholar]
  • Snider B. B. Manganese(III)-Based Oxidative Free-Radical Cyclizations. Chem. Rev. 1996, 96, 339–364. 10.1021/cr950026m. [PubMed] [CrossRef] [Google Scholar]
  • Streuff J. Reductive Umpolung Reactions with Low-Valent Titanium Catalysts. Chem. Rec. 2014, 14, 1100–1113. 10.1002/tcr.201402058. [PubMed] [CrossRef] [Google Scholar]
  • Molander G. A.; Harris C. R. Sequencing Reactions with Samarium(II) Iodide. Chem. Rev. 1996, 96, 307–338. 10.1021/cr950019y. [PubMed] [CrossRef] [Google Scholar]
  • Allonas X.; Dietlin C.; Fouassier J.-P.; Casiraghi A.; Visconti M.; Norcini G.; Bassi G. Barton Esters as New Radical Photoinitiators for Flat Panel Display Applications. J. Photopolym. Sci. Technol. 2008, 21, 505–509. 10.2494/photopolymer.21.505. [CrossRef] [Google Scholar]
  • Saraiva M. F.; Couri M. R. C.; Le Hyaric M.; de Almeida M. V. The Barton Ester Free-Radical Reaction: A Brief Review of Applications. Tetrahedron 2009, 65, 3563–3572. 10.1016/j.tet.2009.01.103. [CrossRef] [Google Scholar]
  • Handbook of Synthetic Photochemistry; Albini A., Fagnoni M., Eds.; Wiley, 2009. [Google Scholar]
  • Albini A.; Protti S.. Paradigms in Green Chemistry and Technology; Springer, 2016. [Google Scholar]
  • Protti S.; Manzini S.; Fagnoni M.; Albini A.. Chapter 2. The Contribution of Photochemistry to Green Chemistry. In Eco-Friendly Synthesis of Fine Chemicals; Roberto Ballini., Ed.; The Royal Society of Chemistry, 2009; pp 80–111. [Google Scholar]
  • Albini A.; Fagnoni M.. The Greenest Reagent in Organic Synthesis: Light. In Green Chemical Reactions; Tundo P., Esposito V., Eds.; Springer Netherlands: Dordrecht, 2008; pp 173–189 10.1007/978-1-4020-8457-7_8. [CrossRef] [Google Scholar]
  • Oelgemöller M.; Jung C.; Mattay J. Green Photochemistry: Production of Fine Chemicals with Sunlight. Pure Appl. Chem. 2007, 79, 1939–1947. 10.1351/pac200779111939. [CrossRef] [Google Scholar]
  • Staveness D.; Bosque I.; Stephenson C. R. J. Free Radical Chemistry Enabled by Visible Light-Induced Electron Transfer. Acc. Chem. Res. 2016, 49, 2295–2306. 10.1021/acs.accounts.6b00270. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Douglas J. J.; Sevrin M. J.; Stephenson C. R. J. Visible Light Photocatalysis: Applications and New Disconnections in the Synthesis of Pharmaceutical Agents. Org. Process Res. Dev. 2016, 20, 1134–1147. 10.1021/acs.oprd.6b00125. [CrossRef] [Google Scholar]
  • Gentry E. C.; Knowles R. R. Synthetic Applications of Proton-Coupled Electron Transfer. Acc. Chem. Res. 2016, 49, 1546–1556. 10.1021/acs.accounts.6b00272. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Xie J.; Jin H.; Hashmi A. S. K. The Recent Achievements of Redox-Neutral Radical C–C Cross-Coupling Enabled by Visible-Light. Chem. Soc. Rev. 2017, 46, 5193–5203. 10.1039/C7CS00339K. [PubMed] [CrossRef] [Google Scholar]
  • Savateev A.; Antonietti M. Heterogeneous Organocatalysis for Photoredox Chemistry. ACS Catal. 2018, 8, 9790–9808. 10.1021/acscatal.8b02595. [CrossRef] [Google Scholar]
  • Silvi M.; Melchiorre P. Enhancing the Potential of Enantioselective Organocatalysis with Light. Nature 2018, 554, 41–49. 10.1038/nature25175. [PubMed] [CrossRef] [Google Scholar]
  • Reiser O. Shining Light on Copper: Unique Opportunities for Visible-Light-Catalyzed Atom Transfer Radical Addition Reactions and Related Processes. Acc. Chem. Res. 2016, 49, 1990–1996. 10.1021/acs.accounts.6b00296. [PubMed] [CrossRef] [Google Scholar]
  • Levin M. D.; Kim S.; Toste F. D. Photoredox Catalysis Unlocks Single-Electron Elementary Steps in Transition Metal Catalyzed Cross-Coupling. ACS Cent. Sci. 2016, 2, 293–301. 10.1021/acscentsci.6b00090. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Lee K. N.; Ngai M.-Y. Recent Developments in Transition-Metal Photoredox-Catalysed Reactions of Carbonyl Derivatives. Chem. Commun. 2017, 53, 13093–13112. 10.1039/C7CC06287G. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Riente P.; Noël T. Application of Metal Oxide Semiconductors in Light-Driven Organic Transformations. Catal. Sci. Technol. 2019, 9, 5186–5232. 10.1039/C9CY01170F. [CrossRef] [Google Scholar]
  • Skubi K. L.; Blum T. R.; Yoon T. P. Dual Catalysis Strategies in Photochemical Synthesis. Chem. Rev. 2016, 116, 10035–10074. 10.1021/acs.chemrev.6b00018. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Tellis J. C.; Kelly C. B.; Primer D. N.; Jouffroy M.; Patel N. R.; Molander G. A. Single-Electron Transmetalation via Photoredox/Nickel Dual Catalysis: Unlocking a New Paradigm for sp3–sp2 Cross-Coupling. Acc. Chem. Res. 2016, 49, 1429–1439. 10.1021/acs.accounts.6b00214. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Twilton J.; Le C.; Zhang P.; Shaw M. H.; Evans R. W.; MacMillan D. W. C. The Merger of Transition Metal and Photocatalysis. Nat. Rev. Chem. 2017, 1, 0052. 10.1038/s41570-017-0052. [CrossRef] [Google Scholar]
  • Zhang L.; Meggers E. Steering Asymmetric Lewis Acid Catalysis Exclusively with Octahedral Metal-Centered Chirality. Acc. Chem. Res. 2017, 50, 320–330. 10.1021/acs.accounts.6b00586. [PubMed] [CrossRef] [Google Scholar]
  • Goddard J. P.; Ollivier C.; Fensterbank L. Photoredox Catalysis for the Generation of Carbon Centered Radicals. Acc. Chem. Res. 2016, 49, 1924–1936. 10.1021/acs.accounts.6b00288. [PubMed] [CrossRef] [Google Scholar]
  • Ravelli D.; Protti S.; Fagnoni M. Carbon–Carbon Bond Forming Reactions via Photogenerated Intermediates. Chem. Rev. 2016, 116, 9850–9913. 10.1021/acs.chemrev.5b00662. [PubMed] [CrossRef] [Google Scholar]
  • Roslin S.; Odell L. R. Visible-Light Photocatalysis as an Enabling Tool for the Functionalization of Unactivated C(sp3)-Substrates. Eur. J. Org. Chem. 2017, 2017, 1993–2007. 10.1002/ejoc.201601479. [CrossRef] [Google Scholar]
  • Matsui J. K.; Lang S. B.; Heitz D. R.; Molander G. A. Photoredox-Mediated Routes to Radicals: The Value of Catalytic Radical Generation in Synthetic Methods Development. ACS Catal. 2017, 7, 2563–2575. 10.1021/acscatal.7b00094. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Protti S.; Fagnoni M.; Albini A.. Photochemical Synthesis. In Green Techniques for Organic Synthesis and Medicinal Chemistry; John Wiley & Sons, Ltd: Chichester, UK, 2018; pp 373–406. [Google Scholar]
  • Milligan J. A.; Phelan J. P.; Badir S. O.; Molander G. A. Alkyl Carbon–Carbon Bond Formation by Nickel/Photoredox Cross-Coupling. Angew. Chem., Int. Ed. 2019, 58, 6152–6163. 10.1002/anie.201809431. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Pitre S. P.; Weires N. A.; Overman L. E. Forging C(sp3)–C(sp3) Bonds with Carbon-Centered Radicals in the Synthesis of Complex Molecules. J. Am. Chem. Soc. 2019, 141, 2800–2813. 10.1021/jacs.8b11790. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Photochemically-Generated Intermediates in Synthesis; John Wiley & Sons, Inc.: Hoboken, NJ, 2013. (ISBN: 978-0-470-91534-9) [Google Scholar]
  • Crespi S.; Protti S.; Fagnoni M. Wavelength Selective Generation of Aryl Radicals and Aryl Cations for Metal-Free Photoarylations. J. Org. Chem. 2016, 81, 9612–9619. 10.1021/acs.joc.6b01619. [PubMed] [CrossRef] [Google Scholar]
  • Yoshida J.; Shimizu A.; Hayashi R. Electrogenerated Cationic Reactive Intermediates: The Pool Method and Further Advances. Chem. Rev. 2018, 118, 4702–4730. 10.1021/acs.chemrev.7b00475. [PubMed] [CrossRef] [Google Scholar]
  • dos Passos Gomes G.; Wimmer A.; Smith J. M.; König B.; Alabugin I. V. CO2 or SO2: Should It Stay, or Should It Go?. J. Org. Chem. 2019, 84, 6232–6243. 10.1021/acs.joc.9b00503. [PubMed] [CrossRef] [Google Scholar]
  • Klauck F. J. R.; James M. J.; Glorius F. Deaminative Strategy for the Visible-Light-Mediated Generation of Alkyl Radicals. Angew. Chem., Int. Ed. 2017, 56, 12336–12339. 10.1002/anie.201706896. [PubMed] [CrossRef] [Google Scholar]
  • Zhang P.; Le C. C.; MacMillan D. W. C. Silyl Radical Activation of Alkyl Halides in Metallaphotoredox Catalysis: A Unique Pathway for Cross-Electrophile Coupling. J. Am. Chem. Soc. 2016, 138, 8084–8087. 10.1021/jacs.6b04818. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Protti S.; Fagnoni M.; Ravelli D. Photocatalytic C-H Activation by Hydrogen-Atom Transfer in Synthesis. ChemCatChem 2015, 7, 1516–1523. 10.1002/cctc.201500125. [CrossRef] [Google Scholar]
  • Constantin T.; Zanini M.; Regni A.; Sheikh N. S.; Juliá F.; Leonori D. Aminoalkyl Radicals as Halogen-Atom Transfer Agents for Activation of Alkyl and Aryl Halides. Science 2020, 367, 1021–1026. 10.1126/science.aba2419. [PubMed] [CrossRef] [Google Scholar]
  • Hoffmann N. Electron and Hydrogen Transfer in Organic Photochemical Reactions. J. Phys. Org. Chem. 2015, 28, 121–136. 10.1002/poc.3370. [CrossRef] [Google Scholar]
  • Ravelli D.; Fagnoni M.; Fukuyama T.; Nishikawa T.; Ryu I. Site-Selective C–H Functionalization by Decatungstate Anion Photocatalysis: Synergistic Control by Polar and Steric Effects Expands the Reaction Scope. ACS Catal. 2018, 8, 701–713. 10.1021/acscatal.7b03354. [CrossRef] [Google Scholar]
  • Capaldo L.; Ravelli D. Hydrogen Atom Transfer (HAT): A Versatile Strategy for Substrate Activation in Photocatalyzed Organic Synthesis. Eur. J. Org. Chem. 2017, 2017, 2056–2071. 10.1002/ejoc.201601485. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Chiba S.; Chen H. sp3 C–H Oxidation by Remote H-Radical Shift with Oxygen- and Nitrogen-Radicals: A Recent Update. Org. Biomol. Chem. 2014, 12, 4051–4060. 10.1039/C4OB00469H. [PubMed] [CrossRef] [Google Scholar]
  • Stateman L.; Nakafuku K.; Nagib D. Remote C–H Functionalization via Selective Hydrogen Atom Transfer. Synthesis 2018, 50, 1569–1586. 10.1055/s-0036-1591930. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • White M. C.; Zhao J. Aliphatic C–H Oxidations for Late-Stage Functionalization. J. Am. Chem. Soc. 2018, 140, 13988–14009. 10.1021/jacs.8b05195. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Barton D. H. R.; Beaton J. M.; Geller L. E.; Pechet M. M. A NEW PHOTOCHEMICAL REACTION. J. Am. Chem. Soc. 1960, 82, 2640–2641. 10.1021/ja01495a061. [CrossRef] [Google Scholar]
  • Barton D. H. R.; Beaton J. M.; Geller L. E.; Pechet M. M. A New Photochemical Reaction 1. J. Am. Chem. Soc. 1961, 83, 4076–4083. 10.1021/ja01480a030. [CrossRef] [Google Scholar]
  • Morcillo S. P. Radical-Promoted C–C Bond Cleavage: A Deconstructive Approach for Selective Functionalization. Angew. Chem., Int. Ed. 2019, 58, 14044–14054. 10.1002/anie.201905218. [PubMed] [CrossRef] [Google Scholar]
  • Roth H.; Romero N.; Nicewicz D. Experimental and Calculated Electrochemical Potentials of Common Organic Molecules for Applications to Single-Electron Redox Chemistry. Synlett 2016, 27, 714–723. 10.1055/s-0035-1561297. [CrossRef] [Google Scholar]
  • Cossy J.; Ranaivosata J.-L.; Bellosta V. Formation of Radicals by Irradiation of Alkyl Halides in the Presence of Triethylamine. Tetrahedron Lett. 1994, 35, 8161–8162. 10.1016/0040-4039(94)88271-1. [CrossRef] [Google Scholar]
  • Isse A. A.; Lin C. Y.; Coote M. L.; Gennaro A. Estimation of Standard Reduction Potentials of Halogen Atoms and Alkyl Halides. J. Phys. Chem. B 2011, 115, 678–684. 10.1021/jp109613t. [PubMed] [CrossRef] [Google Scholar]
  • Lambert F. L.; Ingall G. B. Voltammetry of Organic Halogen Compounds. IV. The Reduction of Organic Chlorides at the Vitreous (Glassy) Carbon Electrode. Tetrahedron Lett. 1974, 15, 3231–3234. 10.1016/S0040-4039(01)91870-2. [CrossRef] [Google Scholar]
  • Lambert F. L.; Kobayashi K. Polarography of Organic Halogen Compounds. I. Steric Hindrance and the Half-Wave Potential in Alicyclic and Aliphatic Halides 1,2. J. Am. Chem. Soc. 1960, 82, 5324–5328. 10.1021/ja01505a014. [CrossRef] [Google Scholar]
  • Vasudevan D. Direct and Indirect Electrochemical Reduction of Organic Halides in Aprotic Media. Russ. J. Electrochem. 2005, 41, 310–314. 10.1007/s11175-005-0067-2. [CrossRef] [Google Scholar]
  • Nawrat C. C.; Jamison C. R.; Slutskyy Y.; MacMillan D. W. C.; Overman L. E. Oxalates as Activating Groups for Alcohols in Visible Light Photoredox Catalysis: Formation of Quaternary Centers by Redox-Neutral Fragment Coupling. J. Am. Chem. Soc. 2015, 137, 11270–11273. 10.1021/jacs.5b07678. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Han J. Bin; Guo A.; Tang X. Y. Alkylation of Allyl/Alkenyl Sulfones by Deoxygenation of Alkoxyl Radicals. Chem. - Eur. J. 2019, 25, 2989–2994. 10.1002/chem.201806138. [PubMed] [CrossRef] [Google Scholar]
  • Capaldo L.; Ravelli D. Alkoxy Radicals Generation: Facile Photocatalytic Reduction of N -Alkoxyazinium or Azolium Salts. Chem. Commun. 2019, 55, 3029–3032. 10.1039/C9CC00035F. [PubMed] [CrossRef] [Google Scholar]
  • Lackner G. L.; Quasdorf K. W.; Overman L. E. Direct Construction of Quaternary Carbons from Tertiary Alcohols via Photoredox-Catalyzed Fragmentation of Tert-Alkyl N-Phthalimidoyl Oxalates. J. Am. Chem. Soc. 2013, 135, 15342–15345. 10.1021/ja408971t. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Chenneberg L.; Baralle A.; Daniel M.; Fensterbank L.; Goddard J. P.; Ollivier C. Visible Light Photocatalytic Reduction of O-Thiocarbamates: Development of a Tin-Free Barton-McCombie Deoxygenation Reaction. Adv. Synth. Catal. 2014, 356, 2756–2762. 10.1002/adsc.201400729. [CrossRef] [Google Scholar]
  • DiRocco D. A.; Dykstra K.; Krska S.; Vachal P.; Conway D. V.; Tudge M. Late-Stage Functionalization of Biologically Active Heterocycles through Photoredox Catalysis. Angew. Chem., Int. Ed. 2014, 53, 4802–4806. 10.1002/anie.201402023. [PubMed] [CrossRef] [Google Scholar]
  • Wu J.; Bär R. M.; Guo L.; Noble A.; Aggarwal V. K. Photoinduced Deoxygenative Borylations of Aliphatic Alcohols. Angew. Chem., Int. Ed. 2019, 58, 18830–18834. 10.1002/anie.201910051. [PubMed] [CrossRef] [Google Scholar]
  • Vara B. A.; Patel N. R.; Molander G. A. O-Benzyl Xanthate Esters under Ni/Photoredox Dual Catalysis: Selective Radical Generation and Csp3–Csp2 Cross-Coupling. ACS Catal. 2017, 7, 3955–3959. 10.1021/acscatal.7b00772. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Lackner G. L.; Quasdorf K. W.; Pratsch G.; Overman L. E. Fragment Coupling and the Construction of Quaternary Carbons Using Tertiary Radicals Generated from Tert-Alkyl N-Phthalimidoyl Oxalates by Visible-Light Photocatalysis. J. Org. Chem. 2015, 80, 6012–6024. 10.1021/acs.joc.5b00794. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Syroeshkin M. A.; Krylov I. B.; Hughes A. M.; Alabugin I. V.; Nasybullina D. V.; Sharipov M. Y.; Gultyai V. P.; Terent’ev A. O. Electrochemical Behavior of N-Oxyphthalimides: Cascades Initiating Self-Sustaining Catalytic Reductive N-O Bond Cleavage. J. Phys. Org. Chem. 2017, 30, e3744 10.1002/poc.3744. [CrossRef] [Google Scholar]
  • Yasu Y.; Koike T.; Akita M. Visible Light-Induced Selective Generation of Radicals from Organoborates by Photoredox Catalysis. Adv. Synth. Catal. 2012, 354, 3414–3420. 10.1002/adsc.201200588. [CrossRef] [Google Scholar]
  • Lima F.; Sharma U. K.; Grunenberg L.; Saha D.; Johannsen S.; Sedelmeier J.; Van der Eycken E. V.; Ley S. V. A Lewis Base Catalysis Approach for the Photoredox Activation of Boronic Acids and Esters. Angew. Chem., Int. Ed. 2017, 56, 15136–15140. 10.1002/anie.201709690. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Li G. X.; Morales-Rivera C. A.; Wang Y.; Gao F.; He G.; Liu P.; Chen G. Photoredox-Mediated Minisci C-H Alkylation of N-Heteroarenes Using Boronic Acids and Hypervalent Iodine. Chem. Sci. 2016, 7, 6407–6412. 10.1039/C6SC02653B. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Shu C.; Noble A.; Aggarwal V. K. Photoredox-Catalyzed Cyclobutane Synthesis by a Deboronative Radical Addition–Polar Cyclization Cascade. Angew. Chem., Int. Ed. 2019, 58, 3870–3874. 10.1002/anie.201813917. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Xue F.; Wang F.; Liu J.; Di J.; Liao Q.; Lu H.; Zhu M.; He L.; He H.; Zhang D.; et al. A Desulfurative Strategy for the Generation of Alkyl Radicals Enabled by Visible-Light Photoredox Catalysis. Angew. Chem., Int. Ed. 2018, 57, 6667–6671. 10.1002/anie.201802710. [PubMed] [CrossRef] [Google Scholar]
  • Knauber T.; Chandrasekaran R.; Tucker J. W.; Chen J. M.; Reese M.; Rankic D. A.; Sach N.; Helal C. Ru/Ni Dual Catalytic Desulfinative Photoredox Csp2-Csp3 Cross-Coupling of Alkyl Sulfinate Salts and Aryl Halides. Org. Lett. 2017, 19, 6566–6569. 10.1021/acs.orglett.7b03280. [PubMed] [CrossRef] [Google Scholar]
  • Xu N.; Zhang Y.; Chen W.; Li P.; Wang L. Photoinduced N-Methylation and N-Sulfonylation of Azobenzenes with DMSO Under Mild Reaction Conditions. Adv. Synth. Catal. 2018, 360, 1199–1208. 10.1002/adsc.201701548. [CrossRef] [Google Scholar]
  • Zemtsov A. A.; Ashirbaev S. S.; Levin V. V.; Kokorekin V. A.; Korlyukov A. A.; Dilman A. D. Photoredox Reaction of 2-Mercaptothiazolinium Salts with Silyl Enol Ethers. J. Org. Chem. 2019, 84, 15745–15753. 10.1021/acs.joc.9b02478. [PubMed] [CrossRef] [Google Scholar]
  • Pandey G.; Rao K. S. S. P.; Sekhar B. B. V. S. Photosensitized One-Electron Reductive Cleavage of a Carbon–Selenium Bond: A Novel Chemoselective Deselenenylation and Phenylselenenyl Group Transfer Radical Chain Reaction. J. Chem. Soc., Chem. Commun. 1993, 21, 1636–1638. 10.1039/C39930001636. [CrossRef] [Google Scholar]
  • Pandey G.; Sesha Poleswara Rao K. S.; Nageshwar Rao K. V. Photosensitized Electron Transfer Promoted Reductive Activation of Carbon-Selenium Bonds to Generate Carbon-Centered Radicals: Application for Unimolecular Group Transfer Radical Reactions. J. Org. Chem. 1996, 61, 6799–6804. 10.1021/jo960805i. [PubMed] [CrossRef] [Google Scholar]
  • Lucas M. A.; Schiesser C. H. (Aryltelluro)Formates as Precursors of Alkyl Radicals: Thermolysis and Photolysis of Primary and Secondary Alkyl (Aryltelluro)Formates. J. Org. Chem. 1996, 61, 5754–5761. 10.1021/jo960838y. [CrossRef] [Google Scholar]
  • Yamago S. Development of Organotellurium-Mediated and Organostibine-Mediated Living Radical Polymerization Reactions. J. Polym. Sci., Part A: Polym. Chem. 2006, 44, 1–12. 10.1002/pola.21154. [CrossRef] [Google Scholar]
  • Han L. B.; Ishihara K. I.; Kambe N.; Ogawa A.; Ryu I.; Sonoda N. Carbotelluration of Alkynes. J. Am. Chem. Soc. 1992, 114, 7591–7592. 10.1021/ja00045a058. [CrossRef] [Google Scholar]
  • Kyushin S.; Masuda Y.; Matsushita K.; Nakadaira Y.; Ohashi M. Novel Alkylation of Aromatic Nitriles via Photo-Induced Electron Transfer of Group 14 Metal-Carbon σ Donors. Tetrahedron Lett. 1990, 31, 6395–6398. 10.1016/S0040-4039(00)97074-6. [CrossRef] [Google Scholar]
  • Corcé V.; Chamoreau L. M.; Derat E.; Goddard J. P.; Ollivier C.; Fensterbank L. Silicates as Latent Alkyl Radical Precursors: Visible-Light Photocatalytic Oxidation of Hypervalent Bis-Catecholato Silicon Compounds. Angew. Chem., Int. Ed. 2015, 54, 11414–11418. 10.1002/anie.201504963. [PubMed] [CrossRef] [Google Scholar]
  • Yoshida J.; Nishiwaki K. Redox Selective Reactions of Organo-Silicon and -Tin Compounds. J. Chem. Soc., Dalton Trans. 1998, 16, 2589–2596. 10.1039/a803343i. [CrossRef] [Google Scholar]
  • Klingler R. J.; Kochi J. K. Electron-Transfer Kinetics from Cyclic Voltammetry. Quantitative Description of Electrochemical Reversibility. J. Phys. Chem. 1981, 85, 1731–1741. 10.1021/j150612a028. [CrossRef] [Google Scholar]
  • Togo H.; Aoki M.; Kuramochi T.; Yokoyama M. Radical Decarboxylative Alkylation onto Heteroaromatic Bases with Trivalent Iodine Compounds. J. Chem. Soc., Perkin Trans. 1 1993, 20, 2417. 10.1039/p19930002417. [CrossRef] [Google Scholar]
  • Hypervalent Iodine Chemistry; Wirth T., Ed.; Topics in Current Chemistry; Springer: Berlin, 2003; Vol. 224. [Google Scholar]
  • Gutierrez-Bonet Á.; Tellis J. C.; Matsui J. K.; Vara B. A.; Molander G. A. 1,4-Dihydropyridines as Alkyl Radical Precursors: Introducing the Aldehyde Feedstock to Nickel/Photoredox Dual Catalysis. ACS Catal. 2016, 6, 8004–8008. 10.1021/acscatal.6b02786. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Nakajima K.; Nojima S.; Sakata K.; Nishibayashi Y. Visible-Light-Mediated Aromatic Substitution Reactions of Cyanoarenes with 4-Alkyl-1,4-dihydropyridines through Double Carbon–Carbon Bond Cleavage. ChemCatChem 2016, 8, 1028–1032. 10.1002/cctc.201600037. [CrossRef] [Google Scholar]
  • Okada K.; Okamoto K.; Oda M. A New and Practical Method of Decarboxylation: Photosensitized Decarboxylation of N-Acyloxyphthalimides via Electron-Transfer Mechanism. J. Am. Chem. Soc. 1988, 110, 8736–8738. 10.1021/ja00234a047. [CrossRef] [Google Scholar]
  • Mella M.; Fasani E.; Albini A. Electron Transfer Photoinduced Cleavage of Acetals. A Mild Preparation of Alkyl Radicals. J. Org. Chem. 1992, 57, 3051–3057. 10.1021/jo00037a020. [CrossRef] [Google Scholar]
  • Sun A. C.; McClain E. J.; Beatty J. W.; Stephenson C. R. J. Visible Light-Mediated Decarboxylative Alkylation of Pharmaceutically Relevant Heterocycles. Org. Lett. 2018, 20, 3487–3490. 10.1021/acs.orglett.8b01250. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Benson S. W. Kinetics of Pyrolysis of Alkyl Hydroperoxides and Their O–O Bond Dissociation Energies. J. Chem. Phys. 1964, 40, 1007–1013. 10.1063/1.1725239. [CrossRef] [Google Scholar]
  • Liu J.; Liu Q.; Yi H.; Qin C.; Bai R.; Qi X.; Lan Y.; Lei A. Visible-Light-Mediated Decarboxylation/Oxidative Amidation of α-Keto Acids with Amines under Mild Reaction Conditions Using O2. Angew. Chem., Int. Ed. 2014, 53, 502–506. 10.1002/anie.201308614. [PubMed] [CrossRef] [Google Scholar]
  • Chen J.-Q.; Chang R.; Wei Y.-L.; Mo J.-N.; Wang Z.-Y.; Xu P.-F. Direct Decarboxylative–Decarbonylative Alkylation of α-Oxo Acids with Electrophilic Olefins via Visible-Light Photoredox Catalysis. J. Org. Chem. 2018, 83, 253–259. 10.1021/acs.joc.7b02628. [PubMed] [CrossRef] [Google Scholar]
  • Okada K.; Okamoto K.; Morita N.; Okubo K.; Oda M. Photosensitized Decarboxylative Michael Addition through N-(Acyloxy)Phthalimides via an Electron-Transfer Mechanism. J. Am. Chem. Soc. 1991, 113, 9401–9402. 10.1021/ja00024a074. [CrossRef] [Google Scholar]
  • Zhang J.; Li Y.; Xu R.; Chen Y. Donor–Acceptor Complex Enables Alkoxyl Radical Generation for Metal-Free C(sp3)–C(sp3) Cleavage and Allylation/Alkenylation. Angew. Chem., Int. Ed. 2017, 56, 12619–12623. 10.1002/anie.201707171. [PubMed] [CrossRef] [Google Scholar]
  • Beatty J. W.; Douglas J. J.; Cole K. P.; Stephenson C. R. J. A Scalable and Operationally Simple Radical Trifluoromethylation. Nat. Commun. 2015, 6, 7919. 10.1038/ncomms8919. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Yang J.-D.; Li M.; Xue X.-S. Computational I(III)-X BDEs for Benziodoxol(on)e-Based Hypervalent Iodine Reagents: Implications for Their Functional Group Transfer Abilities. Chin. J. Chem. 2019, 37, 359–363. 10.1002/cjoc.201800549. [CrossRef] [Google Scholar]
  • Hu X.; Li G.-X.; He G.; Chen G. Minisci C–H Alkylation of N-Heteroarenes with Aliphatic Alcohols via β-Scission of Alkoxy Radical Intermediates. Org. Chem. Front. 2019, 6, 3205–3209. 10.1039/C9QO00786E. [CrossRef] [Google Scholar]
  • Fagnoni M.; Protti S.; Ravelli D.. Photoorganocatalysis in Organic Synthesis; Catalytic Science Series; World Scientific (Europe), 2019; Vol. 18. [Google Scholar]
  • Romero N. A.; Nicewicz D. A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116, 10075–10166. 10.1021/acs.chemrev.6b00057. [PubMed] [CrossRef] [Google Scholar]
  • Bogdos M. K.; Pinard E.; Murphy J. A. Applications of Organocatalysed Visible-Light Photoredox Reactions for Medicinal Chemistry. Beilstein J. Org. Chem. 2018, 14, 2035–2064. 10.3762/bjoc.14.179. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Zeitler K.Metal-Free Photo(Redox) Catalysis. In Visible Light Photocatalysis in Organic Chemistry; Wiley-VCH: Weinheim, 2018; pp 159–232. [Google Scholar]
  • Zilate B.; Fischer C.; Sparr C. Design and Application of Aminoacridinium Organophotoredox Catalysts. Chem. Commun. 2020, 56, 1767–1775. 10.1039/C9CC08524F. [PubMed] [CrossRef] [Google Scholar]
  • Shang T.-Y.; Lu L.-H.; Cao Z.; Liu Y.; He W.-M.; Yu B. Recent Advances of 1,2,3,5-Tetrakis(Carbazol-9-Yl)-4,6-Dicyanobenzene (4CzIPN) in Photocatalytic Transformations. Chem. Commun. 2019, 55, 5408–5419. 10.1039/C9CC01047E. [PubMed] [CrossRef] [Google Scholar]
  • Montanaro S.; Ravelli D.; Merli D.; Fagnoni M.; Albini A. Decatungstate As Photoredox Catalyst: Benzylation of Electron-Poor Olefins. Org. Lett. 2012, 14, 4218–4221. 10.1021/ol301900p. [PubMed] [CrossRef] [Google Scholar]
  • Hunsdiecker H.; Hunsdiecker C. Über Den Abbau Der Salze Aliphatischer Säuren Durch Brom. Ber. Dtsch. Chem. Ges. B 1942, 75, 291–297. 10.1002/cber.19420750309. [CrossRef] [Google Scholar]
  • Borodine A. Ueber Bromvaleriansäure Und Brombuttersäure. Ann. der Chemie und Pharm. 1861, 119, 121–123. 10.1002/jlac.18611190113. [CrossRef] [Google Scholar]
  • Yoshimi Y.; Washida S.; Okita Y.; Nishikawa K.; Maeda K.; Hayashi S.; Morita T. Radical Addition to Acrylonitrile via Catalytic Photochemical Decarboxylation of Aliphatic Carboxylic Acids. Tetrahedron Lett. 2013, 54, 4324–4326. 10.1016/j.tetlet.2013.06.020. [CrossRef] [Google Scholar]
  • Chu L.; Ohta C.; Zuo Z.; MacMillan D. W. C. Carboxylic Acids as A Traceless Activation Group for Conjugate Additions: A Three-Step Synthesis of (±)-Pregabalin. J. Am. Chem. Soc. 2014, 136, 10886–10889. 10.1021/ja505964r. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ramirez N. P.; Gonzalez-Gomez J. C. Decarboxylative Giese-Type Reaction of Carboxylic Acids Promoted by Visible Light: A Sustainable and Photoredox-Neutral Protocol. Eur. J. Org. Chem. 2017, 2017, 2154–2163. 10.1002/ejoc.201601478. [CrossRef] [Google Scholar]
  • Guo T.; Zhang L.; Fang Y.; Jin X.; Li Y.; Li R.; Li X.; Cen W.; Liu X.; Tian Z. Visible-Light-Promoted Decarboxylative Giese Reactions of α-Aryl Ethenylphosphonates and the Application in the Synthesis of Fosmidomycin Analogue. Adv. Synth. Catal. 2018, 360, 1352–1357. 10.1002/adsc.201701285. [CrossRef] [Google Scholar]
  • Slutskyy Y.; Jamison C. R.; Zhao P.; Lee J.; Rhee Y. H.; Overman L. E. Versatile Construction of 6-Substituted Cis-2,8-Dioxabicyclo[3.3.0]Octan-3-Ones: Short Enantioselective Total Syntheses of Cheloviolenes A and B and Dendrillolide C. J. Am. Chem. Soc. 2017, 139, 7192–7195. 10.1021/jacs.7b04265. [PubMed] [CrossRef] [Google Scholar]
  • Slutskyy Y.; Jamison C. R.; Lackner G. L.; Müller D. S.; Dieskau A. P.; Untiedt N. L.; Overman L. E. Short Enantioselective Total Syntheses of Trans-Clerodane Diterpenoids: Convergent Fragment Coupling Using a Trans-Decalin Tertiary Radical Generated from a Tertiary Alcohol Precursor. J. Org. Chem. 2016, 81, 7029–7035. 10.1021/acs.joc.6b00697. [PubMed] [CrossRef] [Google Scholar]
  • Sorin G.; Martinezmallorquin R.; Contie Y.; Baralle A.; Malacria M.; Goddard J. P.; Fensterbank L. Oxidation of Alkyl Trifluoroborates: An Opportunity for Tin-Free Radical Chemistry. Angew. Chem., Int. Ed. 2010, 49, 8721–8723. 10.1002/anie.201004513. [PubMed] [CrossRef] [Google Scholar]
  • Chinzei T.; Miyazawa K.; Yasu Y.; Koike T.; Akita M. Redox-Economical Radical Generation from Organoborates and Carboxylic Acids by Organic Photoredox Catalysis. RSC Adv. 2015, 5, 21297–21300. 10.1039/C5RA01826A. [CrossRef] [Google Scholar]
  • Huo H.; Harms K.; Meggers E. Catalytic, Enantioselective Addition of Alkyl Radicals to Alkenes via Visible-Light-Activated Photoredox Catalysis with a Chiral Rhodium Complex. J. Am. Chem. Soc. 2016, 138, 6936–6939. 10.1021/jacs.6b03399. [PubMed] [CrossRef] [Google Scholar]
  • Lima F.; Grunenberg L.; Rahman H. B. A.; Labes R.; Sedelmeier J.; Ley S. V. Organic Photocatalysis for the Radical Couplings of Boronic Acid Derivatives in Batch and Flow. Chem. Commun. 2018, 54, 5606–5609. 10.1039/C8CC02169D. [PubMed] [CrossRef] [Google Scholar]
  • Verrier C.; Alandini N.; Pezzetta C.; Moliterno M.; Buzzetti L.; Hepburn H. B.; Vega-Peñaloza A.; Silvi M.; Melchiorre P. Direct Stereoselective Installation of Alkyl Fragments at the β-Carbon of Enals via Excited Iminium Ion Catalysis. ACS Catal. 2018, 8, 1062–1066. 10.1021/acscatal.7b03788. [CrossRef] [Google Scholar]
  • van Leeuwen T.; Buzzetti L.; Perego L. A.; Melchiorre P. A Redox-Active Nickel Complex That Acts as an Electron Mediator in Photochemical Giese Reactions. Angew. Chem., Int. Ed. 2019, 58, 4953–4957. 10.1002/anie.201814497. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Müller D. S.; Untiedt N. L.; Dieskau A. P.; Lackner G. L.; Overman L. E. Constructing Quaternary Stereogenic Centers Using Tertiary Organocuprates and Tertiary Radicals. Total Synthesis of Trans-Clerodane Natural Products. J. Am. Chem. Soc. 2015, 137, 660–663. 10.1021/ja512527s. [PubMed] [CrossRef] [Google Scholar]
  • Pratsch G.; Lackner G. L.; Overman L. E. Constructing Quaternary Carbons from N-(Acyloxy)Phthalimide Precursors of Tertiary Radicals Using Visible-Light Photocatalysis. J. Org. Chem. 2015, 80, 6025–6036. 10.1021/acs.joc.5b00795. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Schnermann M. J.; Overman L. E. A Concise Synthesis of (−)-Aplyviolene Facilitated by a Strategic Tertiary Radical Conjugate Addition. Angew. Chem., Int. Ed. 2012, 51, 9576–9580. 10.1002/anie.201204977. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kawamoto T.; Uehara S.; Hirao H.; Fukuyama T.; Matsubara H.; Ryu I. Borohydride-Mediated Radical Addition Reactions of Organic Iodides to Electron-Deficient Alkenes. J. Org. Chem. 2014, 79, 3999–4007. 10.1021/jo500464q. [PubMed] [CrossRef] [Google Scholar]
  • Giedyk M.; Narobe R.; Weiß S.; Touraud D.; Kunz W.; König B. Photocatalytic Activation of Alkyl Chlorides by Assembly-Promoted Single Electron Transfer in Microheterogeneous Solutions. Nat. Catal. 2020, 3, 40–47. 10.1038/s41929-019-0369-5. [CrossRef] [Google Scholar]
  • Elmarrouni A.; Ritts C. B.; Balsells J. Silyl-Mediated Photoredox-Catalyzed Giese Reaction: Addition of Non-Activated Alkyl Bromides. Chem. Sci. 2018, 9, 6639–6646. 10.1039/C8SC02253D. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ravelli D.; Protti S.; Fagnoni M. Decatungstate Anion for Photocatalyzed “Window Ledge” Reactions. Acc. Chem. Res. 2016, 49, 2232–2242. 10.1021/acs.accounts.6b00339. [PubMed] [CrossRef] [Google Scholar]
  • Bonassi F.; Ravelli D.; Protti S.; Fagnoni M. Decatungstate Photocatalyzed Acylations and Alkylations in Flow via Hydrogen Atom Transfer. Adv. Synth. Catal. 2015, 357, 3687–3695. 10.1002/adsc.201500483. [CrossRef] [Google Scholar]
  • Yamada K.; Okada M.; Fukuyama T.; Ravelli D.; Fagnoni M.; Ryu I. Photocatalyzed Site-Selective C-H to C-C Conversion of Aliphatic Nitriles. Org. Lett. 2015, 17, 1292–1295. 10.1021/acs.orglett.5b00282. [PubMed] [CrossRef] [Google Scholar]
  • Fukuyama T.; Nishikawa T.; Yamada K.; Ravelli D.; Fagnoni M.; Ryu I. Photocatalyzed Site-Selective C(sp3)-H Functionalization of Alkylpyridines at Non-Benzylic Positions. Org. Lett. 2017, 19, 6436–6439. 10.1021/acs.orglett.7b03339. [PubMed] [CrossRef] [Google Scholar]
  • Yamada K.; Fukuyama T.; Fujii S.; Ravelli D.; Fagnoni M.; Ryu I. Cooperative Polar/Steric Strategy in Achieving Site-Selective Photocatalyzed C(sp3)–H Functionalization. Chem. - Eur. J. 2017, 23, 8615–8618. 10.1002/chem.201701865. [PubMed] [CrossRef] [Google Scholar]
  • Fukuyama T.; Yamada K.; Nishikawa T.; Ravelli D.; Fagnoni M.; Ryu I. Site-Selectivity in TBADT-Photocatalyzed C(sp3)H Functionalization of Saturated Alcohols and Alkanes. Chem. Lett. 2018, 47, 207–209. 10.1246/cl.171068. [CrossRef] [Google Scholar]
  • Capaldo L.; Fagnoni M.; Ravelli D. Vinylpyridines as Building Blocks for the Photocatalyzed Synthesis of Alkylpyridines. Chem. - Eur. J. 2017, 23, 6527–6530. 10.1002/chem.201701346. [PubMed] [CrossRef] [Google Scholar]
  • Capaldo L.; Merli D.; Fagnoni M.; Ravelli D. Visible Light Uranyl Photocatalysis: Direct C-H to C-C Bond Conversion. ACS Catal. 2019, 9, 3054–3058. 10.1021/acscatal.9b00287. [CrossRef] [Google Scholar]
  • Fan X.-Z.; Rong J.-W.; Wu H.-L.; Zhou Q.; Deng H.-P.; Tan J. Da; Xue C.-W.; Wu L.-Z.; Tao H.-R.; Wu J. Eosin Y as a Direct Hydrogen-Atom Transfer Photocatalyst for the Functionalization of C–H Bonds. Angew. Chem., Int. Ed. 2018, 57, 8514–8518. 10.1002/anie.201803220. [PubMed] [CrossRef] [Google Scholar]
  • Rohe S.; Morris A. O.; McCallum T.; Barriault L. Hydrogen Atom Transfer Reactions via Photoredox Catalyzed Chlorine Atom Generation. Angew. Chem., Int. Ed. 2018, 57, 15664–15669. 10.1002/anie.201810187. [PubMed] [CrossRef] [Google Scholar]
  • Hofmann A. W. Zur Kenntniss Des Piperidins Und Pyridins. Ber. Dtsch. Chem. Ges. 1879, 12, 984–990. 10.1002/cber.187901201254. [CrossRef] [Google Scholar]
  • Löffler K.; Freytag C. Über Eine Neue Bildungsweise von N-Alkylierten Pyrrolidinen. Ber. Dtsch. Chem. Ges. 1909, 42, 3427–3431. 10.1002/cber.19090420377. [CrossRef] [Google Scholar]
  • Li J. J.Hofmann–Löffler–Freytag Reaction. In Name Reactions; Springer: Berlin, 2009; pp 292–293. [Google Scholar]
  • Choi G. J.; Zhu Q.; Miller D. C.; Gu C. J.; Knowles R. R. Catalytic Alkylation of Remote C-H Bonds Enabled by Proton-Coupled Electron Transfer. Nature 2016, 539, 268–271. 10.1038/nature19811. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Chu J. C. K.; Rovis T. Amide-Directed Photoredox-Catalysed C-C Bond Formation at Unactivated sp3 C-H Bonds. Nature 2016, 539, 272–275. 10.1038/nature19810. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • 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, 14897–14900. 10.1021/jacs.7b09306. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Yuan W.; Zhou Z.; Gong L.; Meggers E. Asymmetric Alkylation of Remote C(sp3)-H Bonds by Combining Proton-Coupled Electron Transfer with Chiral Lewis Acid Catalysis. Chem. Commun. 2017, 53, 8964–8967. 10.1039/C7CC04941B. [PubMed] [CrossRef] [Google Scholar]
  • Jiang H.; Studer A. α-Aminoxy-Acid-Auxiliary-Enabled Intermolecular Radical γ-C(sp3)–H Functionalization of Ketones. Angew. Chem., Int. Ed. 2018, 57, 1692–1696. 10.1002/anie.201712066. [PubMed] [CrossRef] [Google Scholar]
  • Ge L.; Li Y.; Jian W.; Bao H. Alkyl Esterification of Vinylarenes Enabled by Visible-Light-Induced Decarboxylation. Chem. - Eur. J. 2017, 23, 11767–11770. 10.1002/chem.201702385. [PubMed] [CrossRef] [Google Scholar]
  • Tlahuext-Aca A.; Garza-Sanchez R. A.; Glorius F. Multicomponent Oxyalkylation of Styrenes Enabled by Hydrogen-Bond-Assisted Photoinduced Electron Transfer. Angew. Chem., Int. Ed. 2017, 56, 3708–3711. 10.1002/anie.201700049. [PubMed] [CrossRef] [Google Scholar]
  • Ma Z. Y.; Guo L. N.; Gu Y. R.; Chen L.; Duan X. H. Iminyl Radical-Mediated Controlled Hydroxyalkylation of Remote C(sp3)-H Bond via Tandem 1,5-HAT and Difunctionalization of Aryl Alkenes. Adv. Synth. Catal. 2018, 360, 4341–4347. 10.1002/adsc.201801198. [CrossRef] [Google Scholar]
  • He B. Q.; Yu X. Y.; Wang P. Z.; Chen J. R.; Xiao W. J. A Photoredox Catalyzed Iminyl Radical-Triggered C-C Bond Cleavage/Addition/Kornblum Oxidation Cascade of Oxime Esters and Styrenes: Synthesis of Ketonitriles. Chem. Commun. 2018, 54, 12262–12265. 10.1039/C8CC07072E. [PubMed] [CrossRef] [Google Scholar]
  • Tlahuext-Aca A.; Garza-Sanchez R. A.; Schäfer M.; Glorius F. Visible-Light-Mediated Synthesis of Ketones by the Oxidative Alkylation of Styrenes. Org. Lett. 2018, 20, 1546–1549. 10.1021/acs.orglett.8b00272. [PubMed] [CrossRef] [Google Scholar]
  • Xia Z.-H.; Zhang C.-L.; Gao Z.-H.; Ye S. Switchable Decarboxylative Heck-Type Reaction and Oxo-Alkylation of Styrenes with N -Hydroxyphthalimide Esters under Photocatalysis. Org. Lett. 2018, 20, 3496–3499. 10.1021/acs.orglett.8b01268. [PubMed] [CrossRef] [Google Scholar]
  • Kong W.; Yu C.; An H.; Song Q. Photoredox-Catalyzed Decarboxylative Alkylation of Silyl Enol Ethers To Synthesize Functionalized Aryl Alkyl Ketones. Org. Lett. 2018, 20, 349–352. 10.1021/acs.orglett.7b03587. [PubMed] [CrossRef] [Google Scholar]
  • Fu M. C.; Shang R.; Zhao B.; Wang B.; Fu Y. Photocatalytic Decarboxylative Alkylations Mediated by Triphenylphosphine and Sodium Iodide. Science 2019, 363, 1429–1434. 10.1126/science.aav3200. [PubMed] [CrossRef] [Google Scholar]
  • Deng W.; Feng W.; Li Y.; Bao H. Merging Visible-Light Photocatalysis and Transition-Metal Catalysis in Three-Component Alkyl-Fluorination of Olefins with a Fluoride Ion. Org. Lett. 2018, 20, 4245–4249. 10.1021/acs.orglett.8b01658. [PubMed] [CrossRef] [Google Scholar]
  • Sim J.; Campbell M. W.; Molander G. A. Synthesis of α-Fluoro-α-Amino Acid Derivatives via Photoredox-Catalyzed Carbofluorination. ACS Catal. 2019, 9, 1558–1563. 10.1021/acscatal.8b04284. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Sha W.; Deng L.; Ni S.; Mei H.; Han J.; Pan Y. Merging Photoredox and Copper Catalysis: Enantioselective Radical Cyanoalkylation of Styrenes. ACS Catal. 2018, 8, 7489–7494. 10.1021/acscatal.8b01863. [CrossRef] [Google Scholar]
  • Ji M.; Wu Z.; Zhu C. Visible-Light-Induced Consecutive C-C Bond Fragmentation and Formation for the Synthesis of Elusive Unsymmetric 1,8-Dicarbonyl Compounds. Chem. Commun. 2019, 55, 2368–2371. 10.1039/C9CC00378A. [PubMed] [CrossRef] [Google Scholar]
  • Kamijo S.; Kamijo K.; Maruoka K.; Murafuji T. Aryl Ketone Catalyzed Radical Allylation of C(sp3)-H Bonds under Photoirradiation. Org. Lett. 2016, 18, 6516–6519. 10.1021/acs.orglett.6b03586. [PubMed] [CrossRef] [Google Scholar]
  • Hu C.; Chen Y. Chemoselective and Fast Decarboxylative Allylation by Photoredox Catalysis under Mild Conditions. Org. Chem. Front. 2015, 2, 1352–1355. 10.1039/C5QO00187K. [CrossRef] [Google Scholar]
  • Zhang M. M.; Liu F. Visible-Light-Mediated Allylation of Alkyl Radicals with Allylic Sulfones: Via a Deaminative Strategy. Org. Chem. Front. 2018, 5, 3443–3446. 10.1039/C8QO01046C. [CrossRef] [Google Scholar]
  • Wu K.; Wang L.; Colón-Rodríguez S.; Flechsig G. U.; Wang T. Amidyl Radical Directed Remote Allylation of Unactivated sp3 C–H Bonds by Organic Photoredox Catalysis. Angew. Chem., Int. Ed. 2019, 58, 1774–1778. 10.1002/anie.201811004. [PubMed] [CrossRef] [Google Scholar]
  • Lang S. B.; Wiles R. J.; Kelly C. B.; Molander G. A. Photoredox Generation of Carbon-Centered Radicals Enables the Construction of 1,1-Difluoroalkene Carbonyl Mimics. Angew. Chem., Int. Ed. 2017, 56, 15073–15077. 10.1002/anie.201709487. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Matsui J. K.; Gutiérrez-Bonet Á.; Rotella M.; Alam R.; Gutierrez O.; Molander G. A. Photoredox/Nickel-Catalyzed Single-Electron Tsuji–Trost Reaction: Development and Mechanistic Insights. Angew. Chem., Int. Ed. 2018, 57, 15847–15851. 10.1002/anie.201809919. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Johnston C. P.; Smith R. T.; Allmendinger S.; MacMillan D. W. C. Metallaphotoredox-Catalysed sp3-sp3 Cross-Coupling of Carboxylic Acids with Alkyl Halides. Nature 2016, 536, 322–325. 10.1038/nature19056. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Lévêque C.; Corcé V.; Chenneberg L.; Ollivier C.; Fensterbank L. Photoredox/Nickel Dual Catalysis for the C(sp3)–C(sp3) Cross-Coupling of Alkylsilicates with Alkyl Halides. Eur. J. Org. Chem. 2017, 2017, 2118–2121. 10.1002/ejoc.201601571. [CrossRef] [Google Scholar]
  • Smith R. T.; Zhang X.; Rincón J. A.; Agejas J.; Mateos C.; Barberis M.; García-Cerrada S.; De Frutos O.; Macmillan D. W. C. Metallaphotoredox-Catalyzed Cross-Electrophile Csp3-Csp3 Coupling of Aliphatic Bromides. J. Am. Chem. Soc. 2018, 140, 17433–17438. 10.1021/jacs.8b12025. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kautzky J. A.; Wang T.; Evans R. W.; Macmillan D. W. C. Decarboxylative Trifluoromethylation of Aliphatic Carboxylic Acids. J. Am. Chem. Soc. 2018, 140, 6522–6526. 10.1021/jacs.8b02650. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Zhou W. J.; Cao G. M.; Shen G.; Zhu X. Y.; Gui Y. Y.; Ye J. H.; Sun L.; Liao L. L.; Li J.; Yu D. G. Visible-Light-Driven Palladium-Catalyzed Radical Alkylation of C–H Bonds with Unactivated Alkyl Bromides. Angew. Chem., Int. Ed. 2017, 56, 15683–15687. 10.1002/anie.201704513. [PubMed] [CrossRef] [Google Scholar]
  • Ren L.; Cong H. Visible-Light-Driven Decarboxylative Alkylation of C–H Bond Catalyzed by Dye-Sensitized Semiconductor. Org. Lett. 2018, 20, 3225–3228. 10.1021/acs.orglett.8b01077. [PubMed] [CrossRef] [Google Scholar]
  • Garrido-Castro A. F.; Choubane H.; Daaou M.; Maestro M. C.; Alemán J. Asymmetric Radical Alkylation of: N -Sulfinimines under Visible Light Photocatalytic Conditions. Chem. Commun. 2017, 53, 7764–7767. 10.1039/C7CC03724D. [PubMed] [CrossRef] [Google Scholar]
  • Patel N. R.; Kelly C. B.; Siegenfeld A. P.; Molander G. A. Mild, Redox-Neutral Alkylation of Imines Enabled by an Organic Photocatalyst. ACS Catal. 2017, 7, 1766–1770. 10.1021/acscatal.6b03665. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Plasko D. P.; Jordan C. J.; Ciesa B. E.; Merrill M. A.; Hanna J. M. Visible Light-Promoted Alkylation of Imines Using Potassium Organotrifluoroborates. Photochem. Photobiol. Sci. 2018, 17, 534–538. 10.1039/C8PP00061A. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Zidan M.; McCallum T.; Thai-Savard L.; Barriault L. Photoredox Meets Gold Lewis Acid Catalysis in the Alkylative Semipinacol Rearrangement: A Photocatalyst with a Dark Side. Org. Chem. Front. 2017, 4, 2092–2096. 10.1039/C7QO00590C. [CrossRef] [Google Scholar]
  • Yao S.; Zhang K.; Zhou Q. Q.; Zhao Y.; Shi D. Q.; Xiao W. J. Photoredox-Promoted Alkyl Radical Addition/Semipinacol Rearrangement Sequences of Alkenylcyclobutanols: Rapid Access to Cyclic Ketones. Chem. Commun. 2018, 54, 8096–8099. 10.1039/C8CC04503H. [PubMed] [CrossRef] [Google Scholar]
  • Huang H.; Jia K.; Chen Y. Hypervalent Iodine Reagents Enable Chemoselective Deboronative/Decarboxylative Alkenylation by Photoredox Catalysis. Angew. Chem., Int. Ed. 2015, 54, 1881–1884. 10.1002/anie.201410176. [PubMed] [CrossRef] [Google Scholar]
  • Wang C.; Lei Y.; Guo M.; Shang Q.; Liu H.; Xu Z.; Wang R. Photoinduced, Copper-Promoted Regio- and Stereoselective Decarboxylative Alkylation of α,β-Unsaturated Acids with Alkyl Iodides. Org. Lett. 2017, 19, 6412–6415. 10.1021/acs.orglett.7b03289. [PubMed] [CrossRef] [Google Scholar]
  • Zhang J. J.; Yang J. C.; Guo L. N.; Duan X. H. Visible-Light-Mediated Dual Decarboxylative Coupling of Redox-Active Esters with α,β-Unsaturated Carboxylic Acids. Chem. - Eur. J. 2017, 23, 10259–10263. 10.1002/chem.201702200. [PubMed] [CrossRef] [Google Scholar]
  • Xu K.; Tan Z.; Zhang H.; Liu J.; Zhang S.; Wang Z. Photoredox Catalysis Enabled Alkylation of Alkenyl Carboxylic Acids with N-(Acyloxy)Phthalimide via Dual Decarboxylation. Chem. Commun. 2017, 53, 10719–10722. 10.1039/C7CC05910H. [PubMed] [CrossRef] [Google Scholar]
  • Kurandina D.; Parasram M.; Gevorgyan V. Visible Light-Induced Room-Temperature Heck Reaction of Functionalized Alkyl Halides with Vinyl Arenes/Heteroarenes. Angew. Chem., Int. Ed. 2017, 56, 14212–14216. 10.1002/anie.201706554. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Koy M.; Sandfort F.; Tlahuext-Aca A.; Quach L.; Daniliuc C. G.; Glorius F. Palladium-Catalyzed Decarboxylative Heck-Type Coupling of Activated Aliphatic Carboxylic Acids Enabled by Visible Light. Chem. - Eur. J. 2018, 24, 4552–4555. 10.1002/chem.201800813. [PubMed] [CrossRef] [Google Scholar]
  • Zheng C.; Cheng W. M.; Li H. L.; Na R. S.; Shang R. Cis-Selective Decarboxylative Alkenylation of Aliphatic Carboxylic Acids with Vinyl Arenes Enabled by Photoredox/Palladium/Uphill Triple Catalysis. Org. Lett. 2018, 20, 2559–2563. 10.1021/acs.orglett.8b00712. [PubMed] [CrossRef] [Google Scholar]
  • Kurandina D.; Rivas M.; Radzhabov M.; Gevorgyan V. Heck Reaction of Electronically Diverse Tertiary Alkyl Halides. Org. Lett. 2018, 20, 357–360. 10.1021/acs.orglett.7b03591. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Xie J.; Li J.; Weingand V.; Rudolph M.; Hashmi A. S. K. Intermolecular Photocatalyzed Heck-like Coupling of Unactivated Alkyl Bromides by a Dinuclear Gold Complex. Chem. - Eur. J. 2016, 22, 12646–12650. 10.1002/chem.201602939. [PubMed] [CrossRef] [Google Scholar]
  • Cao H.; Jiang H.; Feng H.; Kwan J. M. C.; Liu X.; Wu J. Photo-Induced Decarboxylative Heck-Type Coupling of Unactivated Aliphatic Acids and Terminal Alkenes in the Absence of Sacrificial Hydrogen Acceptors. J. Am. Chem. Soc. 2018, 140, 16360–16367. 10.1021/jacs.8b11218. [PubMed] [CrossRef] [Google Scholar]
  • Patel N. R.; Kelly C. B.; Jouffroy M.; Molander G. A. Engaging Alkenyl Halides with Alkylsilicates via Photoredox Dual Catalysis. Org. Lett. 2016, 18, 764–767. 10.1021/acs.orglett.6b00024. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Sumino S.; Uno M.; Huang H. J.; Wu Y. K.; Ryu I. Palladium/Light Induced Radical Alkenylation and Allylation of Alkyl Iodides Using Alkenyl and Allylic Sulfones. Org. Lett. 2018, 20, 1078–1081. 10.1021/acs.orglett.7b04050. [PubMed] [CrossRef] [Google Scholar]
  • Zhou Q. Q.; Düsel S. J. S.; Lu L. Q.; König B.; Xiao W. J. Alkenylation of Unactivated Alkyl Bromides through Visible Light Photocatalysis. Chem. Commun. 2019, 55, 107–110. 10.1039/C8CC08362B. [PubMed] [CrossRef] [Google Scholar]
  • Raviola C.; Protti S.; Ravelli D.; Fagnoni M. Photogenerated Acyl/Alkoxycarbonyl/Carbamoyl Radicals for Sustainable Synthesis. Green Chem. 2019, 21, 748–764. 10.1039/C8GC03810D. [CrossRef] [Google Scholar]
  • Banerjee A.; Lei Z.; Ngai M.-Y. Acyl Radical Chemistry via Visible-Light Photoredox Catalysis. Synthesis 2019, 51, 303–333. 10.1055/s-0037-1610329. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Fusano A.; Fukuyama T.; Nishitani S.; Inouye T.; Ryu I. Synthesis of Alkyl Alkynyl Ketones by Pd/Light-Induced Three-Component Coupling Reactions of Iodoalkanes, CO, and 1-Alkynes. Org. Lett. 2010, 12, 2410–2413. 10.1021/ol1007668. [PubMed] [CrossRef] [Google Scholar]
  • Sumino S.; Ui T.; Hamada Y.; Fukuyama T.; Ryu I. Carbonylative Mizoroki-Heck Reaction of Alkyl Iodides with Arylalkenes Using a Pd/Photoirradiation System. Org. Lett. 2015, 17, 4952–4955. 10.1021/acs.orglett.5b02302. [PubMed] [CrossRef] [Google Scholar]
  • Ishiyama T.; Murata M.; Suzuki A.; Miyaura N. Synthesis of Ketones from Iodoalkenes, Carbon Monoxide and 9-Alkyl-9-Borabicyclo[3.3.1]Nonane Derivatives via a Radical Cyclization and Palladium-Catalysed Carbonylative Cross-Coupling Sequence. J. Chem. Soc., Chem. Commun. 1995, 3, 295. 10.1039/c39950000295. [CrossRef] [Google Scholar]
  • Jiang X.; Zhang M. M.; Xiong W.; Lu L. Q.; Xiao W. J. Deaminative (Carbonylative) Alkyl-Heck-Type Reactions Enabled by Photocatalytic C–N Bond Activation. Angew. Chem., Int. Ed. 2019, 58, 2402–2406. 10.1002/anie.201813689. [PubMed] [CrossRef] [Google Scholar]
  • Ryu I.; Tani A.; Fukuyama T.; Ravelli D.; Fagnoni M.; Albini A. Atom-Economical Synthesis of Unsymmetrical Ketones through Photocatalyzed C-H Activation of Alkanes and Coupling with CO and Electrophilic Alkenes. Angew. Chem., Int. Ed. 2011, 50, 1869–1872. 10.1002/anie.201004854. [PubMed] [CrossRef] [Google Scholar]
  • Okada M.; Fukuyama T.; Yamada K.; Ryu I.; Ravelli D.; Fagnoni M. Sunlight Photocatalyzed Regioselective β-Alkylation and Acylation of Cyclopentanones. Chem. Sci. 2014, 5, 2893–2898. 10.1039/C4SC01072H. [CrossRef] [Google Scholar]
  • Cartier A.; Levernier E.; Corcé V.; Fukuyama T.; Dhimane A. L.; Ollivier C.; Ryu I.; Fensterbank L. Carbonylation of Alkyl Radicals Derived from Organosilicates through Visible-Light Photoredox Catalysis. Angew. Chem., Int. Ed. 2019, 58, 1789–1793. 10.1002/anie.201811858. [PubMed] [CrossRef] [Google Scholar]
  • Amani J.; Molander G. A. Direct Conversion of Carboxylic Acids to Alkyl Ketones. Org. Lett. 2017, 19, 3612–3615. 10.1021/acs.orglett.7b01588. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Amani J.; Molander G. A. Synergistic Photoredox/Nickel Coupling of Acyl Chlorides with Secondary Alkyltrifluoroborates: Dialkyl Ketone Synthesis. J. Org. Chem. 2017, 82, 1856–1863. 10.1021/acs.joc.6b02897. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Amani J.; Alam R.; Badir S.; Molander G. A. Synergistic Visible-Light Photoredox/Nickel-Catalyzed Synthesis of Aliphatic Ketones via N-C Cleavage of Imides. Org. Lett. 2017, 19, 2426–2429. 10.1021/acs.orglett.7b00989. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Zheng S.; Primer D. N.; Molander G. A. Nickel/Photoredox-Catalyzed Amidation via Alkylsilicates and Isocyanates. ACS Catal. 2017, 7, 7957–7961. 10.1021/acscatal.7b02795. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ackerman L. K. G.; Martinez Alvarado J. I.; Doyle A. G. Direct C-C Bond Formation from Alkanes Using Ni-Photoredox Catalysis. J. Am. Chem. Soc. 2018, 140, 14059–14063. 10.1021/jacs.8b09191. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Sun A. C.; McAtee R. C.; McClain E. J.; Stephenson C. R. J. Advancements in Visible-Light-Enabled Radical C(sp)2-H Alkylation of (Hetero)Arenes. Synthesis 2019, 51, 1063–1072. 10.1055/s-0037-1611658. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Browne D. L.; Wright S.; Deadman B. J.; Dunnage S.; Baxendale I. R.; Turner R. M.; Ley S. V. Continuous Flow Reaction Monitoring Using an On-Line Miniature Mass Spectrometer. Rapid Commun. Mass Spectrom. 2012, 26, 1999–2010. 10.1002/rcm.6312. [PubMed] [CrossRef] [Google Scholar]
  • Matsui J. K.; Primer D. N.; Molander G. A. Metal-Free C-H Alkylation of Heteroarenes with Alkyltrifluoroborates: A General Protocol for 1°, 2° and 3° Alkylation. Chem. Sci. 2017, 8, 3512–3522. 10.1039/C7SC00283A. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Yan H.; Hou Z. W.; Xu H. C. Photoelectrochemical C–H Alkylation of Heteroarenes with Organotrifluoroborates. Angew. Chem., Int. Ed. 2019, 58, 4592–4595. 10.1002/anie.201814488. [PubMed] [CrossRef] [Google Scholar]
  • McCallum T.; Barriault L. Direct Alkylation of Heteroarenes with Unactivated Bromoalkanes Using Photoredox Gold Catalysis. Chem. Sci. 2016, 7, 4754–4758. 10.1039/C6SC00807K. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Nuhant P.; Oderinde M. S.; Genovino J.; Juneau A.; Gagné Y.; Allais C.; Chinigo G. M.; Choi C.; Sach N. W.; Bernier L.; et al. Visible-Light-Initiated Manganese Catalysis for C–H Alkylation of Heteroarenes: Applications and Mechanistic Studies. Angew. Chem., Int. Ed. 2017, 56, 15309–15313. 10.1002/anie.201707958. [PubMed] [CrossRef] [Google Scholar]
  • Dong J.; Lyu X.; Wang Z.; Wang X.; Song H.; Liu Y.; Wang Q. Visible-Light-Mediated Minisci C-H Alkylation of Heteroarenes with Unactivated Alkyl Halides Using O2 as an Oxidant. Chem. Sci. 2019, 10, 976–982. 10.1039/C8SC04892D. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Bissonnette N. B.; Boyd M. J.; May G. D.; Giroux S.; Nuhant P. C-H. Functionalization of Heteroarenes Using Unactivated Alkyl Halides through Visible-Light Photoredox Catalysis under Basic Conditions. J. Org. Chem. 2018, 83, 10933–10940. 10.1021/acs.joc.8b01589. [PubMed] [CrossRef] [Google Scholar]
  • Garza-Sanchez R. A.; Tlahuext-Aca A.; Tavakoli G.; Glorius F. Visible Light-Mediated Direct Decarboxylative C–H Functionalization of Heteroarenes. ACS Catal. 2017, 7, 4057–4061. 10.1021/acscatal.7b01133. [CrossRef] [Google Scholar]
  • Zhang X. Y.; Weng W. Z.; Liang H.; Yang H.; Zhang B. Visible-Light-Initiated, Photocatalyst-Free Decarboxylative Coupling of Carboxylic Acids with N-Heterocycles. Org. Lett. 2018, 20, 4686–4690. 10.1021/acs.orglett.8b02016. [PubMed] [CrossRef] [Google Scholar]
  • Genovino J.; Lian Y.; Zhang Y.; Hope T. O.; Juneau A.; Gagné Y.; Ingle G.; Frenette M. Metal-Free-Visible Light C-H Alkylation of Heteroaromatics via Hypervalent Iodine-Promoted Decarboxylation. Org. Lett. 2018, 20, 3229–3232. 10.1021/acs.orglett.8b01085. [PubMed] [CrossRef] [Google Scholar]
  • Koeller J.; Gandeepan P.; Ackermann L. Visible-Light-Induced Decarboxylative C-H Adamantylation of Azoles at Ambient Temperature. Synthesis 2019, 51, 1284–1292. 10.1055/s-0037-1611633. [CrossRef] [Google Scholar]
  • Wang B.; Li P.; Miao T.; Zou L.; Wang L. Visible-Light Induced Decarboxylative C2-Alkylation of Benzothiazoles with Carboxylic Acids under Metal-Free Conditions. Org. Biomol. Chem. 2019, 17, 115–121. 10.1039/C8OB02476F. [PubMed] [CrossRef] [Google Scholar]
  • Cheng W. M.; Shang R.; Fu M. C.; Fu Y. Photoredox-Catalysed Decarboxylative Alkylation of N-Heteroarenes with N-(Acyloxy)Phthalimides. Chem. - Eur. J. 2017, 23, 2537–2541. 10.1002/chem.201605640. [PubMed] [CrossRef] [Google Scholar]
  • Li G.-X.; Hu X.; He G.; Chen G. Photoredox-Mediated Minisci-Type Alkylation of N-Heteroarenes with Alkanes with High Methylene Selectivity. ACS Catal. 2018, 8, 11847–11853. 10.1021/acscatal.8b04079. [CrossRef] [Google Scholar]
  • Quattrini M. C.; Fujii S.; Yamada K.; Fukuyama T.; Ravelli D.; Fagnoni M.; Ryu I. Versatile Cross-Dehydrogenative Coupling of Heteroaromatics and Hydrogen Donors via Decatungstate Photocatalysis. Chem. Commun. 2017, 53, 2335–2338. 10.1039/C6CC09725A. [PubMed] [CrossRef] [Google Scholar]
  • 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. 10.1039/C8SC04134B. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Zhang W. M.; Dai J. J.; Xu J.; Xu H. J. Visible-Light-Induced C2 Alkylation of Pyridine N-Oxides. J. Org. Chem. 2017, 82, 2059–2066. 10.1021/acs.joc.6b02891. [PubMed] [CrossRef] [Google Scholar]
  • Perry I. B.; Brewer T. F.; Sarver P. J.; Schultz D. M.; DiRocco D. A.; MacMillan D. W. C. Direct Arylation of Strong Aliphatic C–H Bonds. Nature 2018, 560, 70–75. 10.1038/s41586-018-0366-x. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Yan D. M.; Xiao C.; Chen J. R. Strong C(sp3)-H Arylation by Synergistic Decatungstate Photo-HAT and Nickel Catalysis. Chem. 2018, 4, 2496–2498. 10.1016/j.chempr.2018.10.014. [CrossRef] [Google Scholar]
  • Zhang X.; MacMillan D. W. C. Alcohols as Latent Coupling Fragments for Metallaphotoredox Catalysis: sp3-sp2 Cross-Coupling of Oxalates with Aryl Halides. J. Am. Chem. Soc. 2016, 138, 13862–13865. 10.1021/jacs.6b09533. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Lévêque C.; Chenneberg L.; Corcé V.; Goddard J. P.; Ollivier C.; Fensterbank L. Primary Alkyl Bis-Catecholato Silicates in Dual Photoredox/Nickel Catalysis: Aryl- and Heteroaryl-Alkyl Cross Coupling Reactions. Org. Chem. Front. 2016, 3, 462–465. 10.1039/C6QO00014B. [CrossRef] [Google Scholar]
  • Jouffroy M.; Primer D. N.; Molander G. A. Base-Free Photoredox/Nickel Dual-Catalytic Cross-Coupling of Ammonium Alkylsilicates. J. Am. Chem. Soc. 2016, 138, 475–478. 10.1021/jacs.5b10963. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Lin K.; Wiles R. J.; Kelly C. B.; Davies G. H. M.; Molander G. A. Haloselective Cross-Coupling via Ni/Photoredox Dual Catalysis. ACS Catal. 2017, 7, 5129–5133. 10.1021/acscatal.7b01773. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Patel N. R.; Molander G. A. Phenol Derivatives as Coupling Partners with Alkylsilicates in Photoredox/Nickel Dual Catalysis. J. Org. Chem. 2016, 81, 7271–7275. 10.1021/acs.joc.6b00800. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Jouffroy M.; Davies G. H. M.; Molander G. A. Accessing Elaborated 2,1-Borazaronaphthalene Cores Using Photoredox/Nickel Dual-Catalytic Functionalization. Org. Lett. 2016, 18, 1606–1609. 10.1021/acs.orglett.6b00466. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Alpers D.; Cole K. P.; Stephenson C. R. J. Visible Light Mediated Aryl Migration by Homolytic C–N Cleavage of Aryl Amines. Angew. Chem., Int. Ed. 2018, 57, 12167–12170. 10.1002/anie.201806659. [PubMed] [CrossRef] [Google Scholar]
  • Tellis J. C.; Amani J.; Molander G. A. Single-Electron Transmetalation: Photoredox/Nickel Dual Catalytic Cross-Coupling of Secondary Alkyl β-Trifluoroboratoketones and -Esters with Aryl Bromides. Org. Lett. 2016, 18, 2994–2997. 10.1021/acs.orglett.6b01357. [PubMed] [CrossRef] [Google Scholar]
  • Primer D. N.; Karakaya I.; Tellis J. C.; Molander G. A. Single-Electron Transmetalation: An Enabling Technology for Secondary Alkylboron Cross-Coupling. J. Am. Chem. Soc. 2015, 137, 2195–2198. 10.1021/ja512946e. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Primer D. N.; Molander G. A. Enabling the Cross-Coupling of Tertiary Organoboron Nucleophiles through Radical-Mediated Alkyl Transfer. J. Am. Chem. Soc. 2017, 139, 9847–9850. 10.1021/jacs.7b06288. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Palaychuk N.; Delano T. J.; Boyd M. J.; Green J.; Bandarage U. K. Synthesis of Cycloalkyl Substituted 7-Azaindoles via Photoredox Nickel Dual Catalytic Cross-Coupling in Batch and Continuous Flow. Org. Lett. 2016, 18, 6180–6183. 10.1021/acs.orglett.6b03223. [PubMed] [CrossRef] [Google Scholar]
  • Buzzetti L.; Prieto A.; Roy S. R.; Melchiorre P. Radical-Based C–C Bond-Forming Processes Enabled by the Photoexcitation of 4-Alkyl-1,4-Dihydropyridines. Angew. Chem., Int. Ed. 2017, 56, 15039–15043. 10.1002/anie.201709571. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ratani T. S.; Bachman S.; Fu G. C.; Peters J. C. Photoinduced, Copper-Catalyzed Carbon-Carbon Bond Formation with Alkyl Electrophiles: Cyanation of Unactivated Secondary Alkyl Chlorides at Room Temperature. J. Am. Chem. Soc. 2015, 137, 13902–13907. 10.1021/jacs.5b08452. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Bao X.; Wang Q.; Zhu J. Dual Photoredox/Copper Catalysis for the Remote C(sp3)–H Functionalization of Alcohols and Alkyl Halides by N-Alkoxypyridinium Salts. Angew. Chem., Int. Ed. 2019, 58, 2139–2143. 10.1002/anie.201813356. [PubMed] [CrossRef] [Google Scholar]
  • Heitz D. R.; Rizwan K.; Molander G. A. Visible-Light-Mediated Alkenylation, Allylation, and Cyanation of Potassium Alkyltrifluoroborates with Organic Photoredox Catalysts. J. Org. Chem. 2016, 81, 7308–7313. 10.1021/acs.joc.6b01207. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ramirez N. P.; König B.; Gonzalez-Gomez J. C. Decarboxylative Cyanation of Aliphatic Carboxylic Acids via Visible-Light Flavin Photocatalysis. Org. Lett. 2019, 21, 1368–1373. 10.1021/acs.orglett.9b00064. [PubMed] [CrossRef] [Google Scholar]
  • Dai J. J.; Zhang W. M.; Shu Y. J.; Sun Y. Y.; Xu J.; Feng Y. S.; Xu H. J. Deboronative Cyanation of Potassium Alkyltrifluoroborates: Via Photoredox Catalysis. Chem. Commun. 2016, 52, 6793–6796. 10.1039/C6CC01530A. [PubMed] [CrossRef] [Google Scholar]
  • Wang M.; Huan L.; Zhu C. Cyanohydrin-Mediated Cyanation of Remote Unactivated C(sp3)-H Bonds. Org. Lett. 2019, 21, 821–825. 10.1021/acs.orglett.8b04104. [PubMed] [CrossRef] [Google Scholar]
  • Huang H.; Zhang G.; Gong L.; Zhang S.; Chen Y. Visible-Light-Induced Chemoselective Deboronative Alkynylation under Biomolecule-Compatible Conditions. J. Am. Chem. Soc. 2014, 136, 2280–2283. 10.1021/ja413208y. [PubMed] [CrossRef] [Google Scholar]
  • Pan Y.; Jia K.; Chen Y.; Chen Y. Investigations of Alkynylbenziodoxole Derivatives for Radical Alkynylations in Photoredox Catalysis. Beilstein J. Org. Chem. 2018, 14, 1215–1221. 10.3762/bjoc.14.103. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Le Vaillant F.; Courant T.; Waser J. Room-Temperature Decarboxylative Alkynylation of Carboxylic Acids Using Photoredox Catalysis and EBX Reagents. Angew. Chem., Int. Ed. 2015, 54, 11200–11204. 10.1002/anie.201505111. [PubMed] [CrossRef] [Google Scholar]
  • Zhou Q. Q.; Guo W.; Ding W.; Wu X.; Chen X.; Lu L. Q.; Xiao W. J. Decarboxylative Alkynylation and Carbonylative Alkynylation of Carboxylic Acids Enabled by Visible-Light Photoredox Catalysis. Angew. Chem., Int. Ed. 2015, 54, 11196–11199. 10.1002/anie.201504559. [PubMed] [CrossRef] [Google Scholar]
  • Yang J.; Zhang J.; Qi L.; Hu C.; Chen Y. Visible-Light-Induced Chemoselective Reductive Decarboxylative Alkynylation under Biomolecule-Compatible Conditions. Chem. Commun. 2015, 51, 5275–5278. 10.1039/C4CC06344A. [PubMed] [CrossRef] [Google Scholar]
  • Gao C.; Li J.; Yu J.; Yang H.; Fu H. Visible-Light Photoredox Synthesis of Internal Alkynes Containing Quaternary Carbons. Chem. Commun. 2016, 52, 7292–7294. 10.1039/C6CC01632D. [PubMed] [CrossRef] [Google Scholar]
  • Schwarz J.; König B. Decarboxylative Alkynylation of Biomass-Derived Compounds by Metal-Free Visible Light Photocatalysis. ChemPhotoChem. 2017, 1, 237–242. 10.1002/cptc.201700034. [CrossRef] [Google Scholar]
  • Ociepa M.; Turkowska J.; Gryko D. Redox-Activated Amines in C(sp3)-C(sp) and C(sp3)-C(sp2) Bond Formation Enabled by Metal-Free Photoredox Catalysis. ACS Catal. 2018, 8, 11362–11367. 10.1021/acscatal.8b03437. [CrossRef] [Google Scholar]
  • Li J.; Tian H.; Jiang M.; Yang H.; Zhao Y.; Fu H. Consecutive Visible-Light Photoredox Decarboxylative Couplings of Adipic Acid Active Esters with Alkynyl Sulfones Leading to Cyclic Compounds. Chem. Commun. 2016, 52, 8862–8864. 10.1039/C6CC04386K. [PubMed] [CrossRef] [Google Scholar]
  • Song Z. Y.; Zhang C. L.; Ye S. Visible Light Promoted Coupling of Alkynyl Bromides and Hantzsch Esters for the Synthesis of Internal Alkynes. Org. Biomol. Chem. 2019, 17, 181–185. 10.1039/C8OB02912A. [PubMed] [CrossRef] [Google Scholar]
  • Liu W.; Li L.; Li C.-J. Empowering a Transition-Metal-Free Coupling between Alkyne and Alkyl Iodide with Light in Water. Nat. Commun. 2015, 6, 6526. 10.1038/ncomms7526. [PubMed] [CrossRef] [Google Scholar]
  • Hazra A.; Lee M. T.; Chiu J. F.; Lalic G. Photoinduced Copper-Catalyzed Coupling of Terminal Alkynes and Alkyl Iodides. Angew. Chem., Int. Ed. 2018, 57, 5492–5496. 10.1002/anie.201801085. [PubMed] [CrossRef] [Google Scholar]
  • Hu D.; Wang L.; Li P. Decarboxylative Borylation of Aliphatic Esters under Visible-Light Photoredox Conditions. Org. Lett. 2017, 19, 2770–2773. 10.1021/acs.orglett.7b01181. [PubMed] [CrossRef] [Google Scholar]
  • Fawcett A.; Pradeilles J.; Wang Y.; Mutsuga T.; Myers E. L.; Aggarwal V. K. Photoinduced Decarboxylative Borylation of Carboxylic Acids. Science 2017, 357, 283–286. 10.1126/science.aan3679. [PubMed] [CrossRef] [Google Scholar]
  • Wu J.; He L.; Noble A.; Aggarwal V. K. Photoinduced Deaminative Borylation of Alkylamines. J. Am. Chem. Soc. 2018, 140, 10700–10704. 10.1021/jacs.8b07103. [PubMed] [CrossRef] [Google Scholar]
  • Sandfort F.; Strieth-Kalthoff F.; Klauck F. J. R.; James M. J.; Glorius F. Deaminative Borylation of Aliphatic Amines Enabled by Visible Light Excitation of an Electron Donor-Acceptor Complex. Chem. - Eur. J. 2018, 24, 17210–17214. 10.1002/chem.201804246. [PubMed] [CrossRef] [Google Scholar]
  • Wu J.; Grant P. S.; Li X.; Noble A.; Aggarwal V. K. Catalyst-Free Deaminative Functionalizations of Primary Amines by Photoinduced Single-Electron Transfer. Angew. Chem., Int. Ed. 2019, 58, 5697–5701. 10.1002/anie.201814452. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ryu I.; Tani A.; Fukuyama T.; Ravelli D.; Montanaro S.; Fagnoni M. Efficient C–H/C–N and C–H/C–CO–N Conversion via Decatungstate-Photoinduced Alkylation of Diisopropyl Azodicarboxylate. Org. Lett. 2013, 15, 2554–2557. 10.1021/ol401061v. [PubMed] [CrossRef] [Google Scholar]
  • Hu A.; Guo J.-J. J.; Pan H.; Zuo Z. Selective Functionalization of Methane, Ethane, and Higher Alkanes by Cerium Photocatalysis. Science 2018, 361, 668–672. 10.1126/science.aat9750. [PubMed] [CrossRef] [Google Scholar]
  • Lang S. B.; Cartwright K. C.; Welter R. S.; Locascio T. M.; Tunge J. A. Photocatalytic Aminodecarboxylation of Carboxylic Acids. Eur. J. Org. Chem. 2016, 2016, 3331–3334. 10.1002/ejoc.201600620. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Yatham V. R.; Bellotti P.; König B. Decarboxylative Hydrazination of Unactivated Carboxylic Acids by Cerium Photocatalysis. Chem. Commun. 2019, 55, 3489–3492. 10.1039/C9CC00492K. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Do H. Q.; Bachman S.; Bissember A. C.; Peters J. C.; Fu G. C. Photoinduced Copper-Catalyzed Alkylation of Amides with Unactivated Secondary Alkyl Halides at Room Temperature. J. Am. Chem. Soc. 2014, 136, 2162–2167. 10.1021/ja4126609. [PubMed] [CrossRef] [Google Scholar]
  • Ahn J. M.; Peters J. C.; Fu G. C. Design of a Photoredox Catalyst That Enables the Direct Synthesis of Carbamate-Protected Primary Amines via Photoinduced, Copper-Catalyzed N-Alkylation Reactions of Unactivated Secondary Halides. J. Am. Chem. Soc. 2017, 139, 18101–18106. 10.1021/jacs.7b10907. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Matier C. D.; Schwaben J.; Peters J. C.; Fu G. C. Copper-Catalyzed Alkylation of Aliphatic Amines Induced by Visible Light. J. Am. Chem. Soc. 2017, 139, 17707–17710. 10.1021/jacs.7b09582. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Wang Y.; Li G. X.; Yang G.; He G.; Chen G. A Visible-Light-Promoted Radical Reaction System for Azidation and Halogenation of Tertiary Aliphatic C-H Bonds. Chem. Sci. 2016, 7, 2679–2683. 10.1039/C5SC04169D. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Kamijo S.; Watanabe M.; Kamijo K.; Tao K.; Murafuji T. Synthesis of Aliphatic Azides by Photoinduced C(sp3)-H. Synthesis 2016, 48, 115–121. 10.1055/s-0035-1560705. [CrossRef] [Google Scholar]
  • Margrey K. A.; Czaplyski W. L.; Nicewicz D. A.; Alexanian E. J. A General Strategy for Aliphatic C–H Functionalization Enabled by Organic Photoredox Catalysis. J. Am. Chem. Soc. 2018, 140, 4213–4217. 10.1021/jacs.8b00592. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Zhao W.; Wurz R. P.; Peters J. C.; Fu G. C. Photoinduced, Copper-Catalyzed Decarboxylative C–N Coupling to Generate Protected Amines: An Alternative to the Curtius Rearrangement. J. Am. Chem. Soc. 2017, 139, 12153–12156. 10.1021/jacs.7b07546. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • 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, 10204–10207. 10.1021/jacs.7b05214. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Liang Y.; Zhang X.; MacMillan D. W. C. Decarboxylative sp3 C–N Coupling via Dual Copper and Photoredox Catalysis. Nature 2018, 559, 83–88. 10.1038/s41586-018-0234-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Mao R.; Frey A.; Balon J.; Hu X. Decarboxylative C(sp3)–N Cross-Coupling via Synergetic Photoredox and Copper Catalysis. Nat. Catal. 2018, 1, 120–126. 10.1038/s41929-017-0023-z. [CrossRef] [Google Scholar]
  • Mao R.; Balon J.; Hu X. Cross-Coupling of Alkyl Redox-Active Esters with Benzophenone Imines: Tandem Photoredox and Copper Catalysis. Angew. Chem., Int. Ed. 2018, 57, 9501–9504. 10.1002/anie.201804873. [PubMed] [CrossRef] [Google Scholar]
  • Davies J.; Angelini L.; Alkhalifah M. A.; Sanz L. M.; Sheikh N. S.; Leonori D. Photoredox Synthesis of Arylhydroxylamines from Carboxylic Acids and Nitrosoarenes. Synthesis 2018, 50, 821–830. 10.1055/s-0036-1591744. [CrossRef] [Google Scholar]
  • Zheng C.; Wang Y.; Xu Y.; Chen Z.; Chen G.; Liang S. H. Ru-Photoredox-Catalyzed Decarboxylative Oxygenation of Aliphatic Carboxylic Acids through N-(Acyloxy)Phthalimide. Org. Lett. 2018, 20, 4824–4827. 10.1021/acs.orglett.8b01885. [PubMed] [CrossRef] [Google Scholar]
  • Mao R.; Balon J.; Hu X. Decarboxylative C(sp3)–O Cross-Coupling. Angew. Chem., Int. Ed. 2018, 57, 13624–13628. 10.1002/anie.201808024. [PubMed] [CrossRef] [Google Scholar]
  • Song H.-T.; Ding W.; Zhou Q.-Q.; Liu J.; Lu L.-Q.; Xiao W.-J. Photocatalytic Decarboxylative Hydroxylation of Carboxylic Acids Driven by Visible Light and Using Molecular Oxygen. J. Org. Chem. 2016, 81, 7250–7255. 10.1021/acs.joc.6b01360. [PubMed] [CrossRef] [Google Scholar]
  • Asaba T.; Katoh Y.; Urabe D.; Inoue M. Total Synthesis of Crotophorbolone. Angew. Chem., Int. Ed. 2015, 54, 14457–14461. 10.1002/anie.201509160. [PubMed] [CrossRef] [Google Scholar]
  • Tateno H.; Iguchi S.; Miseki Y.; Sayama K. Photo-Electrochemical C–H Bond Activation of Cyclohexane Using a WO3 Photoanode and Visible Light. Angew. Chem., Int. Ed. 2018, 57, 11238–11241. 10.1002/anie.201805079. [PubMed] [CrossRef] [Google Scholar]
  • Laudadio G.; Govaerts S.; Wang Y.; Ravelli D.; Koolman H. F.; Fagnoni M.; Djuric S. W.; Noël T. Selective C(sp3)–H Aerobic Oxidation Enabled by Decatungstate Photocatalysis in Flow. Angew. Chem., Int. Ed. 2018, 57, 4078–4082. 10.1002/anie.201800818. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Manabe Y.; Kitawaki Y.; Nagasaki M.; Fukase K.; Matsubara H.; Hino Y.; Fukuyama T.; Ryu I. Revisiting the Bromination of C-H Bonds with Molecular Bromine by Using a Photo-Microflow System. Chem. - Eur. J. 2014, 20, 12750–12753. 10.1002/chem.201402303. [PubMed] [CrossRef] [Google Scholar]
  • Strauss F. J.; Cantillo D.; Guerra J.; Kappe C. O. A Laboratory-Scale Continuous Flow Chlorine Generator for Organic Synthesis. React. Chem. Eng. 2016, 1, 472–476. 10.1039/C6RE00135A. [CrossRef] [Google Scholar]
  • Fukuyama T.; Tokizane M.; Matsui A.; Ryu I. A Greener Process for Flow C-H Chlorination of Cyclic Alkanes Using: In Situ Generation and on-Site Consumption of Chlorine Gas. React. Chem. Eng. 2016, 1, 613–615. 10.1039/C6RE00159A. [CrossRef] [Google Scholar]
  • Zhao M.; Lu W. Visible Light-Induced Oxidative Chlorination of Alkyl sp3 C-H Bonds with NaCl/Oxone at Room Temperature. Org. Lett. 2017, 19, 4560–4563. 10.1021/acs.orglett.7b02153. [PubMed] [CrossRef] [Google Scholar]
  • Halperin S. D.; Fan H.; Chang S.; Martin R. E.; Britton R. A Convenient Photocatalytic Fluorination of Unactivated C-H Bonds. Angew. Chem., Int. Ed. 2014, 53, 4690–4693. 10.1002/anie.201400420. [PubMed] [CrossRef] [Google Scholar]
  • West J. G.; Bedell T. A.; Sorensen E. J. The Uranyl Cation as a Visible-Light Photocatalyst for C(sp3)–H Fluorination. Angew. Chem., Int. Ed. 2016, 55, 8923–8927. 10.1002/anie.201603149. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Xia J. B.; Zhu C.; Chen C. Visible Light-Promoted Metal-Free sp3-C—H Fluorination. Chem. Commun. 2014, 50, 11701–11704. 10.1039/C4CC05650G. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Egami H.; Masuda S.; Kawato Y.; Hamashima Y. Photofluorination of Aliphatic C-H Bonds Promoted by the Phthalimide Group. Org. Lett. 2018, 20, 1367–1370. 10.1021/acs.orglett.8b00133. [PubMed] [CrossRef] [Google Scholar]
  • Ventre S.; Petronijevic F. R.; Macmillan D. W. C. Decarboxylative Fluorination of Aliphatic Carboxylic Acids via Photoredox Catalysis. J. Am. Chem. Soc. 2015, 137, 5654–5657. 10.1021/jacs.5b02244. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Ma J.; Xu W.; Xie J. Predictable Site-Selective Radical Fluorination of Tertiary Ethers. Sci. China: Chem. 2020, 63, 187–191. 10.1007/s11426-019-9636-8. [CrossRef] [Google Scholar]
  • Tarantino G.; Hammond C. Catalytic Formation of C(sp3)–F Bonds via Heterogeneous Photocatalysis. ACS Catal. 2018, 8, 10321–10330. 10.1021/acscatal.8b02844. [CrossRef] [Google Scholar]
  • 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, 744–748. 10.1002/anie.201710790. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Su J. Y.; Grünenfelder D. C.; Takeuchi K.; Reisman S. E. Radical Deoxychlorination of Cesium Oxalates for the Synthesis of Alkyl Chlorides. Org. Lett. 2018, 20, 4912–4916. 10.1021/acs.orglett.8b02045. [PubMed] [CrossRef] [Google Scholar]
  • Candish L.; Standley E. A.; Gómez-Suárez A.; Mukherjee S.; Glorius F. Catalytic Access to Alkyl Bromides, Chlorides and Iodides via Visible Light-Promoted Decarboxylative Halogenation. Chem. - Eur. J. 2016, 22, 9971–9974. 10.1002/chem.201602251. [PubMed] [CrossRef] [Google Scholar]
  • Wappes E. A.; Vanitcha A.; Nagib D. A. β C-H Di-Halogenation via Iterative Hydrogen Atom Transfer. Chem. Sci. 2018, 9, 4500–4504. 10.1039/C8SC01214H. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Barata-Vallejo S.; Bonesi S.; Postigo A. Late Stage Trifluoromethylthiolation Strategies for Organic Compounds. Org. Biomol. Chem. 2016, 14, 7150–7182. 10.1039/C6OB00763E. [PubMed] [CrossRef] [Google Scholar]
  • Candish L.; Pitzer L.; Gómez-Suárez A.; Glorius F. Visible Light-Promoted Decarboxylative Di- And Trifluoromethylthiolation of Alkyl Carboxylic Acids. Chem. - Eur. J. 2016, 22, 4753–4756. 10.1002/chem.201600421. [PubMed] [CrossRef] [Google Scholar]
  • 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. 10.1021/jacs.6b09970. [PubMed] [CrossRef] [Google Scholar]
  • Jin Y.; Yang H.; Fu H. An N-(Acetoxy)Phthalimide Motif as a Visible-Light pro-Photosensitizer in Photoredox Decarboxylative Arylthiation. Chem. Commun. 2016, 52, 12909–12912. 10.1039/C6CC06994K. [PubMed] [CrossRef] [Google Scholar]
  • Qiu G.; Lai L.; Cheng J.; Wu J. Recent Advances in the Sulfonylation of Alkenes with the Insertion of Sulfur Dioxide via Radical Reactions. Chem. Commun. 2018, 54, 10405–10414. 10.1039/C8CC05847D. [PubMed] [CrossRef] [Google Scholar]
  • Liu T.; Li Y.; Lai L.; Cheng J.; Sun J.; Wu J. Photocatalytic Reaction of Potassium Alkyltrifluoroborates and Sulfur Dioxide with Alkenes. Org. Lett. 2018, 20, 3605–3608. 10.1021/acs.orglett.8b01385. [PubMed] [CrossRef] [Google Scholar]
  • Liu T.; Ding Y.; Fan X.; Wu J. Photoinduced Synthesis of (E)-Vinyl Sulfones through the Insertion of Sulfur Dioxide. Org. Chem. Front. 2018, 5, 3153–3157. 10.1039/C8QO00965A. [CrossRef] [Google Scholar]
  • Ye S.; Zheng D.; Wu J.; Qiu G. Photoredox-Catalyzed Sulfonylation of Alkyl Iodides, Sulfur Dioxide, and Electron-Deficient Alkenes. Chem. Commun. 2019, 55, 2214–2217. 10.1039/C9CC00347A. [PubMed] [CrossRef] [Google Scholar]
  • Wang X.; Li H.; Qiu G.; Wu J. Substituted Hantzsch Esters as Radical Reservoirs with the Insertion of Sulfur Dioxide under Photoredox Catalysis. Chem. Commun. 2019, 55, 2062–2065. 10.1039/C8CC10246E. [PubMed] [CrossRef] [Google Scholar]
  • Zheng M.; Li G.; Lu H. Photoredox- or Metal-Catalyzed in Situ SO2-Capture Reactions: Synthesis of β-Ketosulfones and Allylsulfones. Org. Lett. 2019, 21, 1216–1220. 10.1021/acs.orglett.9b00201. [PubMed] [CrossRef] [Google Scholar]
  • Nguyen J. D.; D’Amato E. M.; Narayanam J. M. R.; Stephenson C. R. J. Engaging Unactivated Alkyl, Alkenyl and Aryl Iodides in Visible-Light-Mediated Free Radical Reactions. Nat. Chem. 2012, 4, 854–859. 10.1038/nchem.1452. [PubMed] [CrossRef] [Google Scholar]
  • Zhang H.; Liu P. F.; Chen Q.; Wu Q. Y.; Seville A.; Gu Y. C.; Clough J.; Zhou S. L.; Yang G. F. Synthesis and Absolute Configuration Assignment of Albucidin: A Late-Stage Reductive Deiodination by Visible Light Photocatalysis. Org. Biomol. Chem. 2016, 14, 3482–3485. 10.1039/C6OB00371K. [PubMed] [CrossRef] [Google Scholar]
  • Devery J. J.; Nguyen J. D.; Dai C.; Stephenson C. R. J. Light-Mediated Reductive Debromination of Unactivated Alkyl and Aryl Bromides. ACS Catal. 2016, 6, 5962–5967. 10.1021/acscatal.6b01914. [CrossRef] [Google Scholar]
  • Matsubara R.; Shimada T.; Kobori Y.; Yabuta T.; Osakai T.; Hayashi M. Photoinduced Charge-Transfer State of 4-Carbazolyl-3-(Trifluoromethyl)Benzoic Acid: Photophysical Property and Application to Reduction of Carbon–Halogen Bonds as a Sensitizer. Chem. - Asian J. 2016, 11, 2006–2010. 10.1002/asia.201600538. [PubMed] [CrossRef] [Google Scholar]
  • Fukuyama T.; Fujita Y.; Miyoshi H.; Ryu I.; Kao S. C.; Wu Y. K. Electron Transfer-Induced Reduction of Organic Halides with Amines. Chem. Commun. 2018, 54, 5582–5585. 10.1039/C8CC02445F. [PubMed] [CrossRef] [Google Scholar]
  • Kawamoto T.; Ryu I. Radical Reactions of Borohydrides. Org. Biomol. Chem. 2014, 12, 9733–9742. 10.1039/C4OB01784F. [PubMed] [CrossRef] [Google Scholar]
  • Griffin J. D.; Zeller M. A.; Nicewicz D. A. Hydrodecarboxylation of Carboxylic and Malonic Acid Derivatives via Organic Photoredox Catalysis: Substrate Scope and Mechanistic Insight. J. Am. Chem. Soc. 2015, 137, 11340–11348. 10.1021/jacs.5b07770. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • McCallum T.; Slavko E.; Morin M.; Barriault L. Light-Mediated Deoxygenation of Alcohols with a Dimeric Gold Catalyst. Eur. J. Org. Chem. 2015, 2015, 81–85. 10.1002/ejoc.201403351. [CrossRef] [Google Scholar]
  • Milligan J. A.; Phelan J. P.; Polites V. C.; Kelly C. B.; Molander G. A. Radical/Polar Annulation Reactions (RPARs) Enable the Modular Construction of Cyclopropanes. Org. Lett. 2018, 20, 6840–6844. 10.1021/acs.orglett.8b02968. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Revol G.; McCallum T.; Morin M.; Gagosz F.; Barriault L. Photoredox Transformations with Dimeric Gold Complexes. Angew. Chem., Int. Ed. 2013, 52, 13342–13345. 10.1002/anie.201306727. [PubMed] [CrossRef] [Google Scholar]
  • Shen Y.; Cornella J.; Juliá-Hernández F.; Martin R. Visible-Light-Promoted Atom Transfer Radical Cyclization of Unactivated Alkyl Iodides. ACS Catal. 2017, 7, 409–412. 10.1021/acscatal.6b03205. [CrossRef] [Google Scholar]
  • Gu X.; Lu P.; Fan W.; Li P.; Yao Y. Visible Light Photoredox Atom Transfer Ueno-Stork Reaction. Org. Biomol. Chem. 2013, 11, 7088–7091. 10.1039/c3ob41600c. [PubMed] [CrossRef] [Google Scholar]
  • Fukuyama T.; Fujita Y.; Rashid M. A.; Ryu I. Flow Update for a Cossy Photocyclization. Org. Lett. 2016, 18, 5444–5446. 10.1021/acs.orglett.6b02727. [PubMed] [CrossRef] [Google Scholar]
  • McTiernan C. D.; Pitre S. P.; Ismaili H.; Scaiano J. C. Heterogeneous Light-Mediated Reductive Dehalogenations and Cyclizations Utilizing Platinum Nanoparticles on Titania (PtNP@TiO2). Adv. Synth. Catal. 2014, 356, 2819–2824. 10.1002/adsc.201400547. [CrossRef] [Google Scholar]
  • Rackl D.; Kreitmeier P.; Reiser O. Synthesis of a Polyisobutylene-Tagged fac-Ir(Ppy)3 Complex and Its Application as Recyclable Visible-Light Photocatalyst in a Continuous Flow Process. Green Chem. 2016, 18, 214–219. 10.1039/C5GC01792K. [CrossRef] [Google Scholar]
  • Sha W.; Ni S.; Han J.; Pan Y. Access to Alkyl-Substituted Lactone via Photoredox-Catalyzed Alkylation/Lactonization of Unsaturated Carboxylic Acids. Org. Lett. 2017, 19, 5900–5903. 10.1021/acs.orglett.7b02899. [PubMed] [CrossRef] [Google Scholar]
  • Tang Q.; Liu X.; Liu S.; Xie H.; Liu W.; Zeng J.; Cheng P. N-(Acyloxy)Phthalimides as Tertiary Alkyl Radical Precursors in the Visible Light Photocatalyzed Tandem Radical Cyclization of N-Arylacrylamides to 3,3-Dialkyl Substituted Oxindoles. RSC Adv. 2015, 5, 89009–89014. 10.1039/C5RA17292F. [CrossRef] [Google Scholar]
  • Liu L.; Dong J.; Yan Y.; Yin S. F.; Han L. B.; Zhou Y. Photoredox-Catalyzed Decarboxylative Alkylation/Cyclization of Alkynylphosphine Oxides: A Metal- and Oxidant-Free Method for Accessing Benzo[b]Phosphole Oxides. Chem. Commun. 2019, 55, 233–236. 10.1039/C8CC08689C. [PubMed] [CrossRef] [Google Scholar]
  • Qin Q.; Yu S. Visible-Light-Promoted Remote C(sp3)-H Amidation and Chlorination. Org. Lett. 2015, 17, 1894–1897. 10.1021/acs.orglett.5b00582. [PubMed] [CrossRef] [Google Scholar]
  • 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, 9974–9978. 10.1002/anie.201604704. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  • Richers J.; Heilmann M.; Drees M.; Tiefenbacher K. Synthesis of Lactones via C-H Functionalization of Nonactivated C(sp3)-H Bonds. Org. Lett. 2016, 18, 6472–6475. 10.1021/acs.orglett.6b03371. [PubMed] [CrossRef] [Google Scholar]
  • Sun L.; Ye J. H.; Zhou W. J.; Zeng X.; Yu D. G. Oxy-Alkylation of Allylamines with Unactivated Alkyl Bromides and CO2 via Visible-Light-Driven Palladium Catalysis. Org. Lett. 2018, 20, 3049–3052. 10.1021/acs.orglett.8b01079. [PubMed] [CrossRef] [Google Scholar]
  • Wang P. Z.; Yu X. Y.; Li C. Y.; He B. Q.; Chen J. R.; Xiao W. J. A Photocatalytic Iminyl Radical-Mediated C-C Bond Cleavage/Addition/Cyclization Cascade for the Synthesis of 1,2,3,4-Tetrahydrophenanthrenes. Chem. Commun. 2018, 54, 9925–9928. 10.1039/C8CC06145A. [PubMed] [CrossRef] [Google Scholar]
  • Nishikawa K.; Ando T.; Maeda K.; Morita T.; Yoshimi Y. Photoinduced Electron Transfer Promoted Radical Ring Expansion and Cyclization Reactions of α-(ω-Carboxyalkyl) β-Keto Esters. Org. Lett. 2013, 15, 636–638. 10.1021/ol303460u. [PubMed] [CrossRef] [Google Scholar]
  • Kaldas S. J.; Cannillo A.; McCallum T.; Varriault L. Indole Functionalization via Photoredox Gold Catalysis. Org. Lett. 2015, 17, 2864–2866. 10.1021/acs.orglett.5b01260. [PubMed] [CrossRef] [Google Scholar]
  • Mühmel S.; Alpers D.; Hoffmann F.; Brasholz M. Iridium(III) Photocatalysis: A Visible-Light-Induced Dearomatizative Tandem [4 + 2] Cyclization to Furnish Benzindolizidines. Chem. - Eur. J. 2015, 21, 12308–12312. 10.1002/chem.201502572. [PubMed] [CrossRef] [Google Scholar]
  • Yang J. C.; Zhang J. Y.; Zhang J. J.; Duan X. H.; Guo L. N. Metal-Free, Visible-Light-Promoted Decarboxylative Radical Cyclization of Vinyl Azides with N-Acyloxyphthalimides. J. Org. Chem. 2018, 83, 1598–1605. 10.1021/acs.joc.7b02861. [PubMed] [CrossRef] [Google Scholar]
  • Rohe S.; McCallum T.; Morris A. O.; Barriault L. Transformations of Isonitriles with Bromoalkanes Using Photoredox Gold Catalysis. J. Org. Chem. 2018, 83, 10015–10024. 10.1021/acs.joc.8b01380. [PubMed] [CrossRef] [Google Scholar]
  • He Z.; Bae M.; Wu J.; Jamison T. F. Synthesis of Highly Functionalized Polycyclic Quinoxaline Derivatives Using Visible-Light Photoredox Catalysis. Angew. Chem., Int. Ed. 2014, 53, 14451–14455. 10.1002/anie.201408522. [PubMed] [CrossRef] [Google Scholar]
  • Li Y.; Mao R.; Wu J. N-Radical Initiated Aminosulfonylation of Unactivated C(sp3)-H Bond through Insertion of Sulfur Dioxide. Org. Lett. 2017, 19, 4472–4475. 10.1021/acs.orglett.7b02010. [PubMed] [CrossRef] [Google Scholar]
  • Zhao Y.; Chen J. R.; Xiao W. J. Visible-Light Photocatalytic Decarboxylative Alkyl Radical Addition Cascade for Synthesis of Benzazepine Derivatives. Org. Lett. 2018, 20, 224–227. 10.1021/acs.orglett.7b03588. [PubMed] [CrossRef] [Google Scholar]
  • Ciamician G. THE PHOTOCHEMISTRY OF THE FUTURE. Science 1912, 36, 385–394. 10.1126/science.36.926.385. [PubMed] [CrossRef] [Google Scholar]
  • Syroeshkin M. A.; Kuriakose F.; Saverina E. A.; Timofeeva V. A.; Egorov M. P.; Alabugin I. V. Upconversion of Reductants. Angew. Chem., Int. Ed. 2019, 58, 5532–5550. 10.1002/anie.201807247. [PubMed] [CrossRef] [Google Scholar]
  • Capaldo L.; Quadri L. L.; Ravelli D. Merging Photocatalysis with Electrochemistry: The Dawn of a New Alliance in Organic Synthesis. Angew. Chem., Int. Ed. 2019, 58, 17508–17510. 10.1002/anie.201910348. [PubMed] [CrossRef] [Google Scholar]
  • Yan H.; Hou Z.; Xu H. Photoelectrochemical C–H Alkylation of Heteroarenes with Organotrifluoroborates. Angew. Chem., Int. Ed. 2019, 58, 4592–4595. 10.1002/anie.201814488. [PubMed] [CrossRef] [Google Scholar]
  • Huang H.; Strater Z. M.; Rauch M.; Shee J.; Sisto T. J.; Nuckolls C.; Lambert T. H. Electrophotocatalysis with a Trisaminocyclopropenium Radical Dication. Angew. Chem., Int. Ed. 2019, 58, 13318–13322. 10.1002/anie.201906381. [PMC free article] [PubMed] [CrossRef] [Google Scholar]

Articles from Chemical Reviews are provided here courtesy of American Chemical Society