Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Oct 27.
Published in final edited form as: Trends Chem. 2019 Feb 22;1(1):111–125. doi: 10.1016/j.trechm.2019.01.008

Illuminating Photoredox Catalysis

Rory C McAtee 1, Edward J McClain 1, Corey RJ Stephenson 1,*
PMCID: PMC9608853  NIHMSID: NIHMS1835768  PMID: 36313819

Abstract

Over the past decade, photoredox catalysis has risen to the forefront of synthetic organic chemistry as an indispensable tool for selective small-molecule activation and chemical-bond formation. This cutting-edge platform allows photosensitizers to convert visible light into chemical energy prompting generation of reactive radical intermediates. In this Review, we highlight some of the recent key contributions in the field, including: the impact of the chosen light arrays; promoting fundamental cross-coupling steps; selectively functionalizing aliphatic amines; engaging complementary mechanistic paradigms; and applications in industry. With such a wide breadth of reactivity already realized, the presence of photoredox catalysis in all sectors of organic chemistry is expected for years to come.

Keywords: photocatalysis, photoredox, chemical tool, synthesis, radicals, visible light

From Classical to Modern Radical Generation

Radical intermediates are molecules that are transiently generated during a reaction and contain an unpaired electron (Figure 1A, Key Figure). Classic chemical approaches to generate these intermediates rely on hazardous radical initiators (e.g., AIBN and BEt3), toxic reagents (e.g., Bu3SnH), and in many cases, high temperature or high energy UV irradiation [13]. These modes of traditional radical generation have partly led to radical intermediates being both underexploited and underappreciated in chemical synthesis.

Figure 1.

Figure 1.

Photoredox catalysis as a modern approach to radical intermediate generation.

(A) The classic (left) and modern (right; photoredox catalysis) approach to radical intermediates. (B) Commonly employed metal-centered and organic photocatalysts. (C) A general representation of the oxidative and reductive quenching cycle of Ru(bpy)32+. MLCT and ISC are metal-to-ligand charge transfer and intersystem coupling, respectively.

For nearly fifty years, photoredox catalysis has found widespread utility in the areas of carbon dioxide reduction [4], water splitting [5], and solar-cell materials [6]. Only recently, have these fundamental principles been translated to radical generation for chemical synthesis by recognizing that visible light can be converted to chemical energy for synthetic applicability in a controlled and mild fashion [713]. Common photocatalysts include Ru(II) (1) and Ir(III) (2-3) complexes, as well as organic dyes (4) (Figure 1B). Upon irradiation with visible light, metal-centered catalysts undergo a metal-to-ligand charge transfer (MLCT) (see Glossary) followed by intersystem crossing (ISC) (Box 1) revealing a relatively long-lived triplet excited-state species (e.g., for Ru(bpy)32+*, τ = 1100 ns) (Figure 1C) [1416]. From the excited state, these catalysts can engage in single-electron transfer (SET) events with organic substrates providing access to reactive open-shell intermediates. Notably, excited-state species may act as both a strong oxidant and strong reductant simultaneously (generating Ru(bpy)33+ or Ru(bpy)3+, respectively) allowing for exceptional operator flexibility. Additional catalyst tuning (e.g., metal center and ligand sphere) allows one to predictively modify the catalyst’s redox potential and thus the inherent catalyst properties. In contrast to classic chemical approaches to radical intermediate generation, methods in photoredox catalysis are exceptionally mild relying on easily accessible and bench-stable materials with ambient-temperature operation. Despite remarkable advances in classical UV radical photochemistry [17], the need for specialized equipment and lack of predictable product outcomes have ultimately skewed the perception and industrial adaptation of such methods. Fortunately, a modern era of radical chemistry dawns with photoredox catalysis (and electrocatalysis [18]), which promises to be a steadfast and translatable tool for years to come [19]. In this Review, we highlight important contributions over the last several years for visible light-mediated radical generation and synthetic utilization that supports the notion that photoredox catalysis is a lasting synthetic tool.

Text Box 1. Light Mediated Excitation of Polypyridyl Photocatalysts.

Visible light irradiation of the polypyridyl photocatalysts leads to an excitation event that ultimately furnishes a triplet excited state which is sufficiently long-lived (i.e., rate of relaxation is slower than the rate of diffusion) to allow for single electron transfer events to occur [20]. To access the triplet excited state, the photocatalyst first undergoes a metal-to-ligand charge-transfer (MLCT) event upon visible light irradiation. This photophysical process is characterized by the promotion of an electron from a metal-centered t2g orbital to a ligand-centered π* orbital, resulting in a singlet excited state (S1) of the photocatalyst. Following MLCT, the singlet excited state undergoes intersystem crossing (ISC), which is characterized as a configurational spin flip of the electron in the ligand-centered π* orbital, to give the lowest energy triplet excited state (T1). The resultant triplet excited state is a long-lived excited state, capable of undergoing single-electron transfer events with organic substrates. Importantly, these processes result in the oxidation of the metal center, reduction of the ligand and a configurational spin flip of the promoted electron; but the overall charge of the complex is unchanged.

Figure I.

Figure I.

Metal-to-ligand charge-transfer (MLCT) and intersystem crossing (ISC) events from the ground state to the triplet excited state for the prototypical transition-metal photocatalyst, Ru(bpy)32+, upon visible light irradiation.

Evolution of Light Sources for Enhanced Reactivity and Scalability

The diverse reaction profiles accessible via photochemical approaches have reinvigorated researchers in both academia and industry to solve a host of synthetic challenges [17, 21]. Significant effort has been expended on novel reaction invention; however, there has also been significant interest in decreasing reaction times, developing standardized operating protocols to improve reproducibility, and scaling photochemical reactions for industrial applications. As revealed by the Beer-Lambert-Bouguer law, photon flux decreases exponentially with increasing path length and concentration. Thus, in large batch reactors, incomplete irradiation of reaction solutions causes the photoexcited catalyst to exist only at the reactor surface [22]. This phenomenon leads to both long reaction times and poor reaction efficiencies for larger reaction volumes. It is clear that an increase in light intensity will proportionally lead to an increase in photon capture by the photocatalyst affording a higher concentration of the excited-state species [23].

With this in mind, the synthetic photoredox community has witnessed a rapid evolution of light sources employed to improve established photochemical transformations and aid in the discovery of novel bond-forming reactions (Figure 2A). Importantly, a suite of light setups of varying intensities and wavelengths allows for modular reaction tailoring to best fit the selected photocatalyst. For example, common household compact fluorescent lightbulbs (CFL, broadband emission) can be easily converted to light-emitting diode (LED) arrays of both specific wavelengths and varying intensities. One unique development toward improving reaction efficiencies and rates has been the design of a small-scale integrated photoreactor by MacMillan and co-workers (Figure 2B) [24]. The photoreactor was optimized for maximum power output from the chosen LED array, electronic interface for operating simplicity, and outcome reproducibility. Of note, calorimeter measurements revealed a 10× increase in total incident radiant power with their chosen high power 450 nm LED array (>1.1 W output per LED) system relative to a standard LED lamp apparatus. This reactor was shown to provide significantly shorter reaction times in eight photoredox transformations commonly employed in medicinal chemistry, thereby supporting its further utility.

Figure 2.

Figure 2.

Evolution of light array designs for enhanced reaction efficiency and scalability.

(A) Small library of commonly employed light array designs over the past decade. (B) The demonstration of enhanced reaction efficiency with the use of an integrated photoreactor. (C) The comparison of batch and flow processing for the trifluoromethylation of N-Boc pyrrole.

There is an increasing drive from industry to employ photoredox catalysis because it is well-suited to operate in continuous flow, allowing for more uniform light penetration, and therefore, efficient catalyst excitation relative to batch processes [2528]. Further, continuous flow processing enhances scalability and while simultaneously reducing occupational hazards and industrial waste streams. A successful demonstration of photoredox continuous flow processing was exemplified by Beatty and co-workers radical trifluoromethylation method seeking to address the limited number of scalable trifluoromethylation protocols (Figure 2C) [29]. In collaboration with Eli Lilly, the authors identified trifluoroacetic anhydride (TFAA) as a trifluoromethyl radical source, in conjunction with pyridine N-oxide (PNO) as a sacrificial redox auxiliary. This low cost and operationally simple procedure uses 0.1 mol% Ru(bpy)3Cl2 as the photocatalyst and is proposed to proceed via single-electron reduction of acylated pyridine N-oxide, followed by fragmentation to give the reactive trifluoromethyl radical, CO2, and pyridine. Comparison of the trifluormethylation of N-Boc pyrrole revealed that the reaction efficiency was significantly superior with flow processing compared with that of batch processing, generating 3.33 g (71% yield, Rt = 10 min) of product per hour (compared with 17.8 g, 57% yield over 15 hours in batch). Later, this method was efficiently performed on kilogram scale (0.95 kg isolated, 20 gh−1, 50% yield) [30]. Very recently, Harper and co-workers reported on the design of a continuous flow stirred-tank reactor equipped with a high intensity laser to achieve kg/day throughput for several commonly encountered photochemical coupling reactions [31]. Using this 100 mL reactor, the authors reported tremendous reaction-time acceleration and catalysts-loading optimization. Continued advancements in light array designs will ensure optimal photocatalyst loadings, reduced waste streams, and decreased reaction times ultimately leading to a more sustainable approach to chemical synthesis.

Impacting Elementary Cross-Coupling Steps with Photoredox Catalysis

Transition metal catalysis is a fundamental tool for organic chemists, allowing for simpler construction of C–C and C–X bonds [32]. Typically, these transformations operate via oxidative addition of an aryl halide, followed by transmetallation, and subsequent reductive elimination to furnish the new bond (Figure 3A). Photoredox catalysis has emerged as a tool to enable cross-coupling reactivity under exceptionally mild conditions by shuttling electrons to facilitate redox events, thereby promoting otherwise energetically unfavorable reaction steps [33]. In addition to facilitating reactivity, emerging methods have provided a platform to engage common chemical functionalities as suitable cross-coupling handles (e.g., carboxylic acids, alcohols, and C−H bonds). Dual photoredox/transition metal-catalyzed cross-coupling reactions have been reported with cobalt [34], copper [35], gold [36], nickel [37], palladium [38], ruthenium [39], and rhodium [40]; however, in this minireview, we focus on reaction manifolds utilizing only copper and nickel photoredox catalysis. Specifically, we highlight how light and photocatalysts have been strategically applied to alter the fundamental steps of typical cross-coupling reactions.

Figure 3.

Figure 3.

Impacting elementary cross-coupling steps with photoredox catalysis.

(A) A general transition metal-mediated cross-coupling catalytic cycle. (B) The use of photoredox catalysis to overcome the challenges of oxidative addition to a CuI center. (C) Photoredox catalysis has had the greatest influence on the elementary transmetallation step. (D) The demonstration of triplet-triplet energy transfer from an excited state photocatalyst to an organometallic intermediate.

Oxidative Addition

In a traditional cross-coupling reaction, oxidative addition is often implicated as the first fundamental step of the catalytic cycle, shown by an insertion of the metal center into a polarized (pseudo)halide–carbon bond coupled with a concerted two-electron oxidation of the metal center [32]. Methods developed for copper-catalyzed cross-coupling have been limited due to difficulties associated with oxidative addition to the CuI metal center, as the resultant CuIII species are unstable [41, 42]. Recent work by the Fu and co-workers employs photocatalysis to address the issues often encountered in the oxidative-addition step of copper-catalyzed cross-coupling reactions (Figure 3B). This report discloses the copper catalyzed C–N coupling of lithium carbazole and aryl halides under irradiation by a 13-W CFL lamp [35]. Catalysis is hypothesized to proceed through photoinduced electron transfer from a bis(phosphine)CuIcarbazole to the aryl halide, thereby generating an aryl radical species that may react with the copper catalyst to generate a CuIII intermediate and induce reductive elimination ultimately furnishing the new N-aryl carbazole.

Transmetallation

Transmetallation is the fundamental step that has undergone the largest advancements in photoredox- based methodologies, as a result of the contributions from the MacMillan, Doyle, and Molander groups (Figure 3C). Traditionally, transmetallation is the transfer of a ligand from one metal (e.g., -B(OH)2, -MgCl, or -SnBu3) to another metal [32]. For cross-couplings, the ligand is transferred to the active catalyst through a polar two-electron process. The merger of photoredox and transition-metal catalysis has been propelled by the ability to drive reactivity through formal single-electron transmetallation processes. Molander and co-workers reported use of benzylic potassium trifluoroborate salts for the preparation of bis(aryl)methanes via nickel/photoredox cooperative catalysis [43]. The reaction manifold described exploits the easily oxidized C–B bond of potassium trifluoroborate salts allowing for the generation of benzylic radicals that could be subsequently captured by the nickel catalyst. It is proposed that an aryl bromide undergoes oxidative addition with the alkyl nickel(I) species. The subsequent Ni(III) intermediate is then proposed to undergo reductive elimination furnishing the new C–C bond; reduction of the resultant Ni(I) organometallic species regenerates the active Ni(0) catalyst [44]. Since the seminal report, the scope of reactivity has been extensively evaluated [45], and the methodology has been extended to silicates for transmetallation [46].

Concurrent with the report from Molander and colleagues, the Doyle and MacMillan groups reported the coupling of aryl halides with sp3-hybridized carbon-centered radical intermediates derived from carboxylic acids [37]. Importantly, this is not a formal transmetallation, but rather the generation and capture of carbon-centered radicals. This mechanism is analogous to that of trifluoroborate salts, and thus will be discussed in the context of being an alternative coupling handle to achieve analogous transmetallation-like reactivity. In this work, the application of a photosensitizer allowed for the oxidative decarboxylation of carboxylic acids. This generates a free carbon-centered radical, which is proposed to be captured by a Ni(II) species to generate a Ni(III) intermediate capable of undergoing an analogous reductive elimination to furnish the cross-coupled product [37]. This seminal publication immediately highlighted the potential for dual metal photoredox systems to revolutionize cross-coupling methodologies, as the application of stable and widely available carboxylic acids can now be used as coupling partners for the formation of C–C bonds, circumventing the need for classically employed organometallic reagents.

Reductive Elimination

Reductive elimination is the final fundamental step to a cross-coupling reaction. This reaction step is highlighted by the formation of the new bond, with the corresponding two electron reduction of the metal center (Figure 3D) [32]. One way that dual photoredox transition-metal cross-coupling has been used to enable reductive elimination of the catalytic intermediates is through oxidatively induced reductive elimination. This step has been studied by Hillhouse [47] for aryl and alkyl nickel species and likely facilitates reactivity through single-electron transmetallation of many dual photoredox transition-metal mediated cross-couplings. While this has been the most utilized mechanism to promote reductive elimination, it is not the only means to impact reductive elimination. Triplet-triplet energy transfer from the excited state photocatalyst to an organometallic intermediate has also been documented to result in the reductive elimination from an organonickel(II) to give a new C–O bond and the regeneration of the active Ni(0) catalyst [48]. Importantly, the mechanism of this transformation was thoroughly studied, and the application of time-resolved spectroscopy allowed for the assignment of energy transfer.

Selective Oxidation of Aliphatic Amines with Photoredox Catalysis

Aliphatic amines represent a ubiquitous functionality in biologically active compounds and pharmaceuticals. As such, selective and efficient functionalization of aliphatic amines has represented a major point of emphasis for organic chemists [49, 50]. The direct oxidation of trisubstituted aliphatic amines has an astounding impact on the bond dissociation energy of α-C–R bonds (Figure 4A) [51]. Initial efforts in this area focused on in situ generation and subsequent functionalization of imines from N-aryl tetrahydroquinoline core structures (Figure 4B) [52, 53]. This work served as an important proof of concept, as the reductive quenching of the photocatalyst excited state (PC*) gave rise to the radical cation of tri-substituted amine; subsequent hydrogen-atom transfer led to the formation of imines that could be efficiently trapped upon the addition of a nucleophile.

Figure 4.

Figure 4.

Selective oxidation of aliphatic amines enabled by photoredox catalysis and subsequent synthetic applications.

(A) The resultant impact on the oxidation of aliphatic amines. (B) Initial efforts aimed at the direct oxidation of N-aryl tetrahydroquinoline and subsequent functionalization. (C) Amine oxidation to mimic the proposed biosynthesis of several catharanthine derived alkaloids. (D) The utility of photoredox catalysis in providing new tools for the synthesis of saturated building blocks of interest to the pharmaceutical sector.

As the breadth of reactivity enabled by photoredox catalysis expanded, Stephenson and colleagues applied these methods to novel bond disconnections in organic synthesis. Based on the proposed biosynthesis of several catharanthine derived alkaloids, it was hypothesized that (+)-catharanthine would be an ideal entry point to the selective modification of complex molecular scaffolds through amine oxidation. Following amine oxidation to intermediate 5, subsequent strain driven cleavage of the C16-C21 bond (α to the amine) led to the facile production of an imine (not shown) [54]. Trapping of the imine with an equivalent of cyanide followed by single-electron reduction and protonation of intermediate 6 led to α-amino nitrile 7; a common intermediate in the synthesis of several structurally related alkaloids (Figure 4C). The synthetic utility of α-amino nitrile 7 was demonstrated as it was readily converted to (−)-pseudotabersonine, (+)-coronaridine, and (−)-pseudovincadifformine in 90%, 48%, and 55% yields, respectively [54].

New and mild tools enabled by photoredox catalysis have afforded the preparation of new saturated compounds of interest to the pharmaceutical sector because they are less prone to adverse metabolic processing. Anilines represent a common structural alert motif known to predispose a potential drug candidate to metabolism-driven toxicities [55]. However, this functionality is commonly found in modern drug discovery screening libraries due to the cornucopia of methods for its preparation [56]. 1-Aminonorbornanes represent a class of molecules that are well suited to serve as bioisosteres for anilines, providing a core saturated structure likely less prone to adverse metabolic processing events (Figure 4D). Although historical examples of 1-aminonorbornanes applications exist, their application to drug discovery has been precluded by limited synthetic accessibility. Recently, Stephenson and co-workers provided a solution to this need by reporting the application of photoredox catalysis to access a formal (3+2) cycloaddition of aminocyclopropanes with tethered olefins to provide the corresponding 1-aminonorbornane products [57]. In this work, the oxidation of aminocyclopropanes allows for the strain-driven homolysis of the α-amino C–C bond, and subsequent serial 6-exo-trig and 5-exo-trig radical cyclizations to furnish the desired 1-aminonorbornane core. This methodology proved general as it allowed access to a variety of C2-, C3-, C4-, and C7- substituted 1-aminonorbornanes. Metabolic stability studies supported the initial hypothesis that the saturated 1-aminonorbornanes is less prone to metabolic processing than their aniline counterparts, as in all cases, the 1-aminonorbornanes were found to outperform, or were on par, with the stability of the most robust aniline compounds.

Complementary Reaction Paradigms for Anti-Markovnikov Additions to Alkenes

Photoredox catalysis offers unique opportunities for developing complementary mechanistic profiles depending on how a given substrate is initially activated. For example, photocatalytic anti-Markovnikov selective hydrofunctionalization of alkenes was recently demonstrated by both Nicewicz [58] and Knowles [59] through contrasting C–nucleophile bond forming strategies while using a common photoredox catalysis cycle (Figure 5). Despite the pervasiveness of Markovnikov-selective additions (H–nucleophiles to alkenes) in organic chemistry, methods to access the opposite selectivity with the same substrates remains challenging and is limited to forcing transition metal catalysis and monosubstituted activated alkenes [60, 61].

Figure 5.

Figure 5.

Complementary mechanistic paradigms for anti-Markovnikov additions to alkenes.

(A) Anti-Markovnikov selective hydrofunctionalization of alkenes via alkene radical cations. (B) Anti-Markovnikov selective hydrofunctionalization of alkenes via concerted proton-coupled electron transfer approach.

Over the past several years, Nicewicz and colleagues have developed powerful strategies that reverse selectivity of traditional hydrofunctionalization reactions of alkenes by taking advantage of the known reactivity of transiently generated radical cations (Figure 5A) [6265]. In these cases, polar nucleophiles selectively add to the least hindered site of radical cations generating a stabilized radical adduct. Key to the success of this single-electron alkene oxidation strategy is the judicious choice of a potent oxidizing photocatalyst [66, 67]. Notably, many terminal styrenes, as well as mono-, di-, and trisubstituted alkenes, have oxidation potentials outside the range of commonly employed transition-metal-based polypyridyl catalysts. Consequently, the group has spent tremendous effort on designing acridinium photocatalysts with potent redox behaviors [68]. Following mechanistic studies, the authors propose the following steps for their reported anti-Markovnikov hydrofunctionalization reactions of alkenes [62, 69]. Initial single-electron transfer from the alkene to an excited-state acridinium photocatalyst provides the reactive radical cation intermediate 8. Inter- or intramolecular nucleophilic addition to the radical cation followed by a hydrogen atom transfer (HAT) event between a suitable donor, such as 2-phenylmalonitrile, furnishes the functionalized anti-Markovnikov product. The authors have successfully extended this photoredox HAT strategy to accomplish anti-Markovnikov hydroetherifications [58], hydroaminations [63, 70], hydroacetoxylations [71], and hydrohalogentaions [72, 73] of alkenes. Recently, Stephenson and co-workers reported an intermolecular alkene aminoarylation method from transiently generated alkene radical cations and arylsulfonylacetamides as a bifunctional reagent [74]. This report illustrates the power of alkenes radical cations to provide access to valuable chemical motifs (such as arylethylamines) in a succinct manner.

Instead of alkene oxidation, nucleophile activation via non-covalent catalysis in conjunction with photoredox catalysis has enabled a wide breadth of fundamentally distinct transformations. Knowles and colleagues have pioneered the use of concerted proton-coupled electron transfer (PCET) in organic synthesis [75, 76]. PCETs are unconventional elementary redox processes resulting in the concomitant transfer of a proton and an electron to or from two independent donor/acceptor species. This strategy for homolytic activation of strong bonds, often in the presence of weaker ones, allows access to radical species that would be kinetically challenging to form via sequential proton and electron transfer steps. To showcase the synthetic utility of photoredox PCET, the group has developed anti-Markovnikov alkene functionalization methods through the oxidative generation of amidyl radicals (9) from the corresponding amides (Figure 5B) [59, 7779]. Following homolytic cleavage of a redox activated amide N−H bond (BDFE = 110 kcal mol−1), the generated amidyl radical is poised to cyclize onto pendent alkenes to provide a nucleophilic carbon−centered radical. Depending on the nature of the reaction conditions, the radical intermediate can (i) be trapped with a suitable Michael acceptor for a C−C bond forming event or (ii) abstract an H-atom from a HAT catalyst (such as thiophenol) to provide hydroamination products. Overall, given the exceptionally mild nature of both photoredox alkene oxidation and PCET methods, a wide range of functional groups are tolerated and will aid in accelerating the synthetic utility of these complementary approaches for substrate activation.

Recent Applications of Photoredox Catalysis

Since 2008, the field of photoredox catalysis has experienced exponential growth, providing synthetic chemists with novel bond-disconnection strategies and direct approaches to targeting native functionalities (including C−H bonds [80, 81]) under exceptionally mild conditions. Photoredox catalysis has also proven useful in the synthesis of congested quaternary centers through either oxidative [82] or reductive generation [8387] of radical intermediates. Given these qualities, it is unsurprising that photoredox catalysis has served as a key bond-forming strategy in the total synthesis of complex natural products (including (+)-gliocladin C [88], heitziamide A [89], and (−)-aplyviolene) [90, 91], as well as medicinally relevant compounds (Figure 6A) [21].

Figure 6.

Figure 6.

Recent applications of photoredox catalysis in total synthesis and the industrial sector.

(A) The use of photoredox catalysis as a key step in the synthesis of natural products. (B) The use of photoredox catalysis in the direct late-stage C−H methylation, ethylation, and cyclopropanation of pharmaceutical and agrochemical agents. (C) Photocatalytic indoline dehydrogenation as a key step in the sustainable synthesis of elbasvir. (D) A practical photoredox-mediated hydrogen atom transfer protocol to selectively deuterate and tritiate α-amino sp3 C−H bonds of 18 pharmaceutical compounds.

The feasibility of translating small-scale photoredox reactions to large-scale flow-platforms has attracted industrial chemists for applications in late-stage drug modifications and large-scale production of key synthetic intermediates [26, 92, 93]. For example, DiRocco and co-workers have reported a direct late-stage C−H methylation, ethylation, and cyclopropanation of pharmaceutical and agrochemical agents (Figure 6B) [94]. From high-throughput experimentation, it was found that photoredox catalysis can activate organic peroxides to be suitable radical alkylating agents. Given that the method exhibits exceptional functional group tolerance, it is ideally suited for drug discovery. In a subsequent report, the Merck team in collaboration with Knowles and colleagues, reported on a photocatalytic indoline dehydrogenation as a key step in the sustainable synthesis of elbasvir, a clinically investigated inhibitor of the hepatitis C virus (Figure 6C) [95]. The photocatalyst 2 could be used in combination with tert-butyl perbenzoate to provide good yield and excellent ee (85% yield, >99% ee) of the dehydrogenative product. Notably, the reaction could be scaled to 100 g and processed over 5 h with a residence time of 60 min using Merck’s in-house flow reactor [96]. More recently, MacMillan and colleagues reported a practical photoredox-mediated hydrogen atom transfer protocol to selectively deuterate and tritiate α-amino sp3 C−H bonds of 18 pharmaceutical compounds (Figure 6D) [97]. Isotopically labeled molecules are essential diagnostic tools in drug discovery as they provide information about compound metabolism and biological uptake [98100]. This single-step operation, which uses isotopically labeled water as the heavy atom source, is anticipated to be broadly enabling for interrogating the biological activity of novel drug candidates in the future.

Concluding Remarks

Modern advances in visible-light photoredox catalysis have led to myriad of novel synthetic methodologies. The ability of excited state photocatlysts to simultaneously act as both an oxidant and reductant and their ability to convert visible light into useful chemical energy have led to unprecedented reactivity, holding significant promise for enabling the continued discovery of valuable organic transformations. As the pharmaceutical sector continues to embrace photoredox catalysis, there is an ever-increasing opportunity for academic discoveries to be immediately translated to future technologies. Moreover, in an era when sustainable chemical practices are of crucial importance, further development of novel visible light-mediated methodologies and easily adaptable platforms for scaling reactions are needed for this field to continue to grow and thrive (see Outstanding Questions).

Acknowledgements

Financial support for this work was provided by the NIH NIGMS (R01-GM127774), the Camille Dreyfus Teacher-Scholar Award Program, and the University of Michigan. This work is supported by an NSF Graduate Research Fellowship for R.C.M. (grant DGE 1256260).

Glossary

Aliphatic

fully saturated organic compounds composed of carbon and hydrogen atoms (hydrocarbons)

anti-Markovnikov

a rule describing the regiochemical outcome of an alkene or alkyne addition reaction where a substituent becomes bonded to the less substituted carbon, rather than the more substituted carbon

Bioisostere

chemical groups with similar physical or chemical properties that produce similar biological properties to another chemical compound

Hydrogen Atom Transfer (HAT)

hydrogen atom abstraction; when a hydrogen free radical is abstracted from a substrate

Intersystem Crossing (ISC)

a radiationless process involving a transition between the two electronic states with different states spin multiplicity

Metal-to-Ligand Charge Transfer

promotion of an electron from the highest lying metal centered orbital to the lowest lying unoccupied ligand centered orbital. Typically, d→π*

Open-Shell Intermediates

a molecule that contains singly occupied molecular orbitals which tend to be highly reactive

Oxidative Addition

a process in which the oxidation state and coordination number of a metal center increase

Proton-coupled electron transfer (PCET)

a chemical reaction that involves the concerted transfer of an electron and a proton

Reductive Elimination

a process in which the oxidation state of the metal center decreases while forming a new covalent bond between two ligands

Single-Electron Transfer

an event where an electron relocates from an atom or molecule (donor) to another atom or molecule (acceptor) resulting in a change in oxidation state of the chemical entities

Transmetallation

the transfer of a ligand from one metal center to another metal center

Footnotes

Conflicts of Interest

The authors declare no conflicts of interest.

References

  • 1.O’Mahony G (2004) Triethylborane (Et3B). Synlett 2004 (03), 572–573. [Google Scholar]
  • 2.Vitaku E et al. (2014) Analysis of the structural diversity, substitution patterns, and frequency of nitrogen heterocycles among US FDA approved pharmaceuticals: miniperspective. Journal of medicinal chemistry 57 (24), 10257–10274. [DOI] [PubMed] [Google Scholar]
  • 3.Curran DP (1988) The Design and Application of Free Radical Chain Reactions in Organic Synthesis. Part 1. Synthesis 1988 (06), 417–439. [Google Scholar]
  • 4.Takeda H and Ishitani O (2010) Development of efficient photocatalytic systems for CO2 reduction using mononuclear and multinuclear metal complexes based on mechanistic studies. Coordination Chemistry Reviews 254 (3), 346–354. [Google Scholar]
  • 5.Graetzel M (1981) Artificial photosynthesis: water cleavage into hydrogen and oxygen by visible light. Accounts of Chemical Research 14 (12), 376–384. [Google Scholar]
  • 6.Kalyanasundaram K and Grätzel M (1998) Applications of functionalized transition metal complexes in photonic and optoelectronic devices. Coordination Chemistry Reviews 177 (1), 347–414. [Google Scholar]
  • 7.Narayanam JMR and Stephenson CRJ (2011) Visible light photoredox catalysis: applications in organic synthesis. Chemical Society Reviews 40 (1), 102–113. [DOI] [PubMed] [Google Scholar]
  • 8.Xuan J and Xiao W-J (2012) Visible-Light Photoredox Catalysis. Angewandte Chemie International Edition 51 (28), 6828–6838. [DOI] [PubMed] [Google Scholar]
  • 9.Reckenthäler M and Griesbeck AG (2013) Photoredox Catalysis for Organic Syntheses. Advanced Synthesis & Catalysis 355 (14‐15), 2727–2744. [Google Scholar]
  • 10.Prier CK et al. (2013) Visible Light Photoredox Catalysis with Transition Metal Complexes: Applications in Organic Synthesis. Chemical Reviews 113 (7), 5322–5363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Schultz DM and Yoon TP (2014) Solar Synthesis: Prospects in Visible Light Photocatalysis. Science 343 (6174). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Skubi KL et al. (2016) Dual Catalysis Strategies in Photochemical Synthesis. Chemical Reviews 116 (17), 10035–10074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Romero NA and Nicewicz DA (2016) Organic Photoredox Catalysis. Chemical Reviews 116 (17), 10075–10166. [DOI] [PubMed] [Google Scholar]
  • 14.Kalyanasundaram K (1982) Photophysics, photochemistry and solar energy conversion with tris(bipyridyl)ruthenium(II) and its analogues. Coordination Chemistry Reviews 46, 159–244. [Google Scholar]
  • 15.Campagna S et al. (2007) Photochemistry and photophysics of coordination compounds: ruthenium. In Photochemistry and Photophysics of Coordination Compounds I, pp. 117–214, Springer. [Google Scholar]
  • 16.McCusker JK (2003) Femtosecond Absorption Spectroscopy of Transition Metal Charge-Transfer Complexes. Accounts of Chemical Research 36 (12), 876–887. [DOI] [PubMed] [Google Scholar]
  • 17.Kärkäs MD et al. (2016) Photochemical Approaches to Complex Chemotypes: Applications in Natural Product Synthesis. Chemical Reviews 116 (17), 9683–9747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Yan M et al. (2017) Synthetic Organic Electrochemical Methods Since 2000: On the Verge of a Renaissance. Chemical Reviews 117 (21), 13230–13319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Yan M et al. (2016) Radicals: Reactive Intermediates with Translational Potential. Journal of the American Chemical Society 138 (39), 12692–12714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Arias-Rotondo DM and McCusker JK (2016) The photophysics of photoredox catalysis: a roadmap for catalyst design. Chemical Society Reviews 45 (21), 5803–5820. [DOI] [PubMed] [Google Scholar]
  • 21.Douglas JJ et al. (2016) Visible Light Photocatalysis: Applications and New Disconnections in the Synthesis of Pharmaceutical Agents. Organic Process Research & Development 20 (7), 1134–1147. [Google Scholar]
  • 22.Shvydkiv O et al. (2010) From Conventional to Microphotochemistry: Photodecarboxylation Reactions Involving Phthalimides. Organic Letters 12 (22), 5170–5173. [DOI] [PubMed] [Google Scholar]
  • 23.Elliott LD et al. (2014) Batch versus Flow Photochemistry: A Revealing Comparison of Yield and Productivity. Chemistry – A European Journal 20 (46), 15226–15232. [DOI] [PubMed] [Google Scholar]
  • 24.Le CC et al. (2017) A General Small-Scale Reactor To Enable Standardization and Acceleration of Photocatalytic Reactions. ACS Central Science 3 (6), 647–653. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Tucker JW et al. (2012) Visible-Light Photoredox Catalysis in Flow. Angewandte Chemie International Edition 51 (17), 4144–4147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Cambié D et al. (2016) Applications of Continuous-Flow Photochemistry in Organic Synthesis, Material Science, and Water Treatment. Chemical Reviews 116 (17), 10276–10341. [DOI] [PubMed] [Google Scholar]
  • 27.Su Y et al. (2014) Photochemical Transformations Accelerated in Continuous-Flow Reactors: Basic Concepts and Applications. Chemistry – A European Journal 20 (34), 10562–10589. [DOI] [PubMed] [Google Scholar]
  • 28.Pouliot M et al. (2009) Oxidation of catecholboron enolates with TEMPO. Angewandte Chemie International Edition 48 (33), 6037–6040. [DOI] [PubMed] [Google Scholar]
  • 29.Beatty JW et al. (2015) A scalable and operationally simple radical trifluoromethylation. Nature Communications 6, 7919. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Beatty Joel W. et al. (2016) Photochemical Perfluoroalkylation with Pyridine N-Oxides: Mechanistic Insights and Performance on a Kilogram Scale. Chem 1 (3), 456–472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Harper KC et al. (2019) A Laser Driven Flow Chemistry Platform for Scaling Photochemical Reactions with Visible Light. ACS Central Science. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Spessard GO and Miessler GL (2015) Organometallic Chemistry, Oxford University Press. [Google Scholar]
  • 33.Twilton J et al. (2017) The merger of transition metal and photocatalysis. Nature Reviews Chemistry 1, 0052. [Google Scholar]
  • 34.Ruhl KE and Rovis T (2016) Visible Light-Gated Cobalt Catalysis for a Spatially and Temporally Resolved [2+2+2] Cycloaddition. Journal of the American Chemical Society 138 (48), 15527–15530. [DOI] [PubMed] [Google Scholar]
  • 35.Creutz SE et al. (2012) Photoinduced Ullmann C–N Coupling: Demonstrating the Viability of a Radical Pathway. Science 338 (6107), 647. [DOI] [PubMed] [Google Scholar]
  • 36.Sahoo B et al. (2013) Combining Gold and Photoredox Catalysis: Visible Light-Mediated Oxy- and Aminoarylation of Alkenes. Journal of the American Chemical Society 135 (15), 5505–5508. [DOI] [PubMed] [Google Scholar]
  • 37.Zuo Z et al. (2014) Dual catalysis. Merging photoredox with nickel catalysis: coupling of alpha-carboxyl sp(3)-carbons with aryl halides. Science 345 (6195), 437–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Kalyani D et al. (2011) Room-Temperature C–H Arylation: Merger of Pd-Catalyzed C–H Functionalization and Visible-Light Photocatalysis. Journal of the American Chemical Society 133 (46), 18566–18569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Fabry DC et al. (2015) C–H Functionalization of Phenols Using Combined Ruthenium and Photoredox Catalysis: In Situ Generation of the Oxidant. Angewandte Chemie International Edition 54 (9), 2801–2805. [DOI] [PubMed] [Google Scholar]
  • 40.Fabry DC et al. (2014) Combining Rhodium and Photoredox Catalysis for C–H Functionalizations of Arenes: Oxidative Heck Reactions with Visible Light. Angewandte Chemie International Edition 53 (38), 10228–10231. [DOI] [PubMed] [Google Scholar]
  • 41.Jones GO et al. (2010) Computational Explorations of Mechanisms and Ligand-Directed Selectivities of Copper-Catalyzed Ullmann-Type Reactions. Journal of the American Chemical Society 132 (17), 6205–6213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Yu H-Z et al. (2010) Alternative Mechanistic Explanation for Ligand-Dependent Selectivities in Copper-Catalyzed N- and O-Arylation Reactions. Journal of the American Chemical Society 132 (51), 18078–18091. [DOI] [PubMed] [Google Scholar]
  • 43.Tellis JC et al. (2014) Single-electron transmetalation in organoboron cross-coupling by photoredox/nickel dual catalysis. Science 345 (6195), 433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Gutierrez O et al. (2015) Nickel-Catalyzed Cross-Coupling of Photoredox-Generated Radicals: Uncovering a General Manifold for Stereoconvergence in Nickel-Catalyzed Cross-Couplings. Journal of the American Chemical Society 137 (15), 4896–4899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Tellis JC et al. (2016) Single-Electron Transmetalation via Photoredox/Nickel Dual Catalysis: Unlocking a New Paradigm for sp3–sp2 Cross-Coupling. Accounts of Chemical Research 49 (7), 1429–1439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Jouffroy M et al. (2016) Base-Free Photoredox/Nickel Dual-Catalytic Cross-Coupling of Ammonium Alkylsilicates. Journal of the American Chemical Society 138 (2), 475–478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Koo K and Hillhouse GL (1995) Carbon-Nitrogen Bond Formation by Reductive Elimination from Nickel(II) Amido Alkyl Complexes. Organometallics 14 (9), 4421–4423. [Google Scholar]
  • 48.Welin ER et al. (2017) Photosensitized, energy transfer-mediated organometallic catalysis through electronically excited nickel(II). Science 355 (6323), 380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Chen J-R et al. (2016) Visible light photoredox-controlled reactions of N-radicals and radical ions. Chemical Society Reviews 45 (8), 2044–2056. [DOI] [PubMed] [Google Scholar]
  • 50.Davies J et al. (2018) Hydroxylamine Derivatives as Nitrogen-Radical Precursors in Visible-Light Photochemistry. Chemistry – A European Journal 24 (47), 12154–12163. [DOI] [PubMed] [Google Scholar]
  • 51.Beatty JW and Stephenson CRJ (2015) Amine Functionalization via Oxidative Photoredox Catalysis: Methodology Development and Complex Molecule Synthesis. Accounts of Chemical Research 48 (5), 1474–1484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Condie AG et al. (2010) Visible-Light Photoredox Catalysis: Aza-Henry Reactions via C−H Functionalization. Journal of the American Chemical Society 132 (5), 1464–1465. [DOI] [PubMed] [Google Scholar]
  • 53.Freeman DB et al. (2012) Functionally Diverse Nucleophilic Trapping of Iminium Intermediates Generated Utilizing Visible Light. Organic Letters 14 (1), 94–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Beatty JW and Stephenson CRJ (2014) Synthesis of (−)-Pseudotabersonine, (−)-Pseudovincadifformine, and (+)-Coronaridine Enabled by Photoredox Catalysis in Flow. Journal of the American Chemical Society 136 (29), 10270–10273. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Ritchie TJ et al. (2011) The impact of aromatic ring count on compound developability: further insights by examining carbo- and hetero-aromatic and -aliphatic ring types. Drug Discovery Today 16 (3), 164–171. [DOI] [PubMed] [Google Scholar]
  • 56.Brown DG and Boström J (2016) Analysis of Past and Present Synthetic Methodologies on Medicinal Chemistry: Where Have All the New Reactions Gone? Journal of Medicinal Chemistry 59 (10), 4443–4458. [DOI] [PubMed] [Google Scholar]
  • 57.Staveness D et al. (2018) Providing a New Aniline Bioisostere through the Photochemical Production of 1-Aminonorbornanes. Chem. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Hamilton DS and Nicewicz DA (2012) Direct Catalytic Anti-Markovnikov Hydroetherification of Alkenols. Journal of the American Chemical Society 134 (45), 18577–18580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Musacchio AJ et al. (2014) Catalytic Olefin Hydroamination with Aminium Radical Cations: A Photoredox Method for Direct C–N Bond Formation. Journal of the American Chemical Society 136 (35), 12217–12220. [DOI] [PubMed] [Google Scholar]
  • 60.Beller M et al. (2004) Catalytic Markovnikov and anti-Markovnikov Functionalization of Alkenes and Alkynes: Recent Developments and Trends. Angewandte Chemie International Edition 43 (26), 3368–3398. [DOI] [PubMed] [Google Scholar]
  • 61.Han L-B and Tanaka M (1999) Transition metal-catalysed addition reactions of H–heteroatom and inter-heteroatom bonds to carbon–carbon unsaturated linkages via oxidative additions. Chemical Communications (5), 395–402. [Google Scholar]
  • 62.Margrey KA and Nicewicz DA (2016) A General Approach to Catalytic Alkene Anti-Markovnikov Hydrofunctionalization Reactions via Acridinium Photoredox Catalysis. Accounts of Chemical Research 49 (9), 1997–2006. [DOI] [PubMed] [Google Scholar]
  • 63.Nguyen TM et al. (2014) anti-Markovnikov Hydroamination of Alkenes Catalyzed by a Two-Component Organic Photoredox System: Direct Access to Phenethylamine Derivatives. Angewandte Chemie International Edition 53 (24), 6198–6201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Gesmundo NJ et al. (2015) Amide and Amine Nucleophiles in Polar Radical Crossover Cycloadditions: Synthesis of γ-Lactams and Pyrrolidines. Organic Letters 17 (5), 1316–1319. [DOI] [PubMed] [Google Scholar]
  • 65.Miller DC et al. (2015) Catalytic Olefin Hydroamidation Enabled by Proton-Coupled Electron Transfer. Journal of the American Chemical Society 137 (42), 13492–13495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Fukuzumi S et al. (2004) Electron-Transfer State of 9-Mesityl-10-methylacridinium Ion with a Much Longer Lifetime and Higher Energy Than That of the Natural Photosynthetic Reaction Center. Journal of the American Chemical Society 126 (6), 1600–1601. [DOI] [PubMed] [Google Scholar]
  • 67.Fukuzumi S et al. (2014) Long-Lived Charge Separation and Applications in Artificial Photosynthesis. Accounts of Chemical Research 47 (5), 1455–1464. [DOI] [PubMed] [Google Scholar]
  • 68.Joshi-Pangu A et al. (2016) Acridinium-Based Photocatalysts: A Sustainable Option in Photoredox Catalysis. The Journal of Organic Chemistry 81 (16), 7244–7249. [DOI] [PubMed] [Google Scholar]
  • 69.Romero NA and Nicewicz DA (2014) Mechanistic Insight into the Photoredox Catalysis of Anti-Markovnikov Alkene Hydrofunctionalization Reactions. Journal of the American Chemical Society 136 (49), 17024–17035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Nguyen TM and Nicewicz DA (2013) Anti-Markovnikov Hydroamination of Alkenes Catalyzed by an Organic Photoredox System. Journal of the American Chemical Society 135 (26), 9588–9591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Perkowski AJ and Nicewicz DA (2013) Direct Catalytic Anti-Markovnikov Addition of Carboxylic Acids to Alkenes. Journal of the American Chemical Society 135 (28), 10334–10337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Wilger DJ et al. (2014) The direct anti-Markovnikov addition of mineral acids to styrenes. Nature Chemistry 6, 720. [DOI] [PubMed] [Google Scholar]
  • 73.Wilger DJ et al. (2013) Catalytic hydrotrifluoromethylation of styrenes and unactivated aliphatic alkenes via an organic photoredox system. Chemical Science 4 (8), 3160–3165. [Google Scholar]
  • 74.Monos TM et al. (2018) Arylsulfonylacetamides as bifunctional reagents for alkene aminoarylation. Science 361 (6409), 1369–1373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Gentry EC and Knowles RR (2016) Synthetic Applications of Proton-Coupled Electron Transfer. Accounts of Chemical Research 49 (8), 1546–1556. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Miller DC et al. (2016) Proton-Coupled Electron Transfer in Organic Synthesis: Fundamentals, Applications, and Opportunities. Top Curr Chem (Cham) 374 (3), 30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Choi GJ and Knowles RR (2015) Catalytic Alkene Carboaminations Enabled by Oxidative Proton-Coupled Electron Transfer. Journal of the American Chemical Society 137 (29), 9226–9229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Nguyen LQ and Knowles RR (2016) Catalytic C–N Bond-Forming Reactions Enabled by Proton-Coupled Electron Transfer Activation of Amide N–H Bonds. ACS Catalysis 6 (5), 2894–2903. [Google Scholar]
  • 79.Choi GJ et al. (2016) Catalytic alkylation of remote C–H bonds enabled by proton-coupled electron transfer. Nature 539, 268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Perry IB et al. (2018) Direct arylation of strong aliphatic C–H bonds. Nature 560 (7716), 70–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Schultz DM et al. (2017) Oxyfunctionalization of the Remote C-H Bonds of Aliphatic Amines by Decatungstate Photocatalysis. Angew Chem Int Ed Engl 56 (48), 15274–15278. [DOI] [PubMed] [Google Scholar]
  • 82.Nguyen JD et al. (2011) Intermolecular Atom Transfer Radical Addition to Olefins Mediated by Oxidative Quenching of Photoredox Catalysts. Journal of the American Chemical Society 133 (12), 4160–4163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Tucker JW et al. (2010) Electron Transfer Photoredox Catalysis: Intramolecular Radical Addition to Indoles and Pyrroles. Organic Letters 12 (2), 368–371. [DOI] [PubMed] [Google Scholar]
  • 84.Tucker JW et al. (2010) Tin-free radical cyclization reactions initiated by visible light photoredox catalysis. Chemical Communications 46 (27), 4985–4987. [DOI] [PubMed] [Google Scholar]
  • 85.Furst L et al. (2010) Visible Light-Mediated Intermolecular C−H Functionalization of Electron-Rich Heterocycles with Malonates. Organic Letters 12 (13), 3104–3107. [DOI] [PubMed] [Google Scholar]
  • 86.Swift EC et al. (2016) Intermolecular Photocatalytic C–H Functionalization of Electron-Rich Heterocycles with Tertiary Alkyl Halides. Synlett 27 (05), 754–758. [Google Scholar]
  • 87.Pratsch G et al. (2015) Constructing Quaternary Carbons from N-(Acyloxy)phthalimide Precursors of Tertiary Radicals Using Visible-Light Photocatalysis. The Journal of Organic Chemistry 80 (12), 6025–6036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Furst L et al. (2011) Total Synthesis of (+)-Gliocladin C Enabled by Visible-Light Photoredox Catalysis. Angewandte Chemie International Edition 50 (41), 9655–9659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Lin S et al. (2011) Radical Cation Diels–Alder Cycloadditions by Visible Light Photocatalysis. Journal of the American Chemical Society 133 (48), 19350–19353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Schnermann MJ and Overman LE (2011) Enantioselective Total Synthesis of Aplyviolene. Journal of the American Chemical Society 133 (41), 16425–16427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Schnermann MJ and Overman LE (2012) A Concise Synthesis of (−)-Aplyviolene Facilitated by a Strategic Tertiary Radical Conjugate Addition. Angewandte Chemie International Edition 51 (38), 9576–9580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.McAtee RC et al. (2018) Radical Chlorodifluoromethylation: Providing a Motif for (Hetero)arene Diversification. Organic Letters 20 (12), 3491–3495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Sun AC et al. (2018) Visible Light-Mediated Decarboxylative Alkylation of Pharmaceutically Relevant Heterocycles. Organic Letters 20 (12), 3487–3490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.DiRocco DA et al. (2014) Late-Stage Functionalization of Biologically Active Heterocycles Through Photoredox Catalysis. Angewandte Chemie International Edition 53 (19), 4802–4806. [DOI] [PubMed] [Google Scholar]
  • 95.Yayla HG et al. (2016) Discovery and mechanistic study of a photocatalytic indoline dehydrogenation for the synthesis of elbasvir. Chemical Science 7 (3), 2066–2073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Borman S, Chemists Go 100% Organic At 2015 NOS, AMER CHEMICAL SOC 1155 16TH ST, NW, WASHINGTON, DC 20036 USA, 2015, pp. 33–34. [Google Scholar]
  • 97.Loh YY et al. (2017) Photoredox-catalyzed deuteration and tritiation of pharmaceutical compounds. Science. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Isin EM et al. (2012) Use of Radiolabeled Compounds in Drug Metabolism and Pharmacokinetic Studies. Chemical Research in Toxicology 25 (3), 532–542. [DOI] [PubMed] [Google Scholar]
  • 99.Elmore CS and Bragg RA (2015) Isotope chemistry; a useful tool in the drug discovery arsenal. Bioorganic & Medicinal Chemistry Letters 25 (2), 167–171. [DOI] [PubMed] [Google Scholar]
  • 100.Elmore CS (2009) Chapter 25 The Use of Isotopically Labeled Compounds in Drug Discovery. In Annual Reports in Medicinal Chemistry (Macor JE ed), pp. 515–534, Academic Press. [Google Scholar]

RESOURCES