Next Article in Journal
Absorption Spectra of Protonated Corroles: Two Distinct Patterns Due to Peripheral Substitution Architecture
Previous Article in Journal
Stimuli-Sensitive Pyrenylated Hydrogels as Optical Sensing Platform for Multiple Metal Ions
Previous Article in Special Issue
Palladium Catalyzed Allylic C-H Oxidation Enabled by Bicyclic Sulfoxide Ligands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Application of Alkoxy Radical in Organic Synthesis

Department of Chemistry, University at Albany, State University of New York, 1400 Washington Ave., Albany, NY 12222, USA
*
Author to whom correspondence should be addressed.
Organics 2023, 4(4), 459-489; https://doi.org/10.3390/org4040033
Submission received: 2 May 2023 / Revised: 29 June 2023 / Accepted: 22 September 2023 / Published: 28 September 2023

Abstract

:
Alkoxy radicals have been identified as versatile intermediates in synthetic chemistry in the last few decades. Over the last decade, various catalytic processes for the in situ generation of alkoxy radicals have been explored, leading to the development of new synthetic methodologies based on alkoxy radicals. In this review, we provided a comprehensive review of recent developments in the utilization of alkoxy radicals in diverse organic transformations, natural product synthesis, and the late-stage modification of bioactive molecules through the implementation of the photoredox methodology.

1. Introduction

Alkoxy radicals are highly reactive structural frameworks in which radicals are localized at the oxygen atom and singularly bound to an alkyl group [1,2]. Alkoxy radicals, without stabilization by delocalization effects, are a highly energetic class of intermediates among heteroatom-centered radicals. Modern catalytic methods utilize oxygen-centered radicals as versatile intermediates [3,4,5,6]. Alkoxy radical intermediates deliver the most stabilized alkyl radical via the β-scission pathway, resulting in unsymmetrical substrates that are highly expected [7,8,9]. This radical chemistry has been well utilized for natural product synthesis in recent decades [10,11,12,13,14]. However, there are still many challenges in in situ generating oxygen radical species from free alcohols due to the lower reduction potential and strong bond dissociation energy (BDE, ~105 kcal mol−1) of the alcoholic bond [15]. These barriers hinder their application in synthetic organic transformations. The O-radicals can be indirectly generated from conventional sources such as lead (IV) alkoxides [16], peroxides [17,18], sulfonates [19,20], hypohalites nitrite esters [12,21], N-alkoxylpyridine-2-thiones [22], and N-alkoxy phthalimides [13,23,24].
The development of radical chemistry via photoredox catalysis has facilitated a wide range of applications in organic transformation over recent years [25]. This novel approach is gaining high interest because it produces highly energetic radical intermediates that are difficult to obtain through conventional catalytic methodologies. It has established remarkable progress in carbon radical generation, realizing new C-X, C-C, C-N, C-H, and C-S bond formation [26,27,28]. Nitrogen and oxygen-centered radicals have also been widely applied in organic synthesis [2,4,29,30]. The light-induced catalytic system has been an attractive field where synthetic chemists could harness solar energy to generate the desired radical species [31,32,33], especially using heteroatom-centered radicals [34,35,36,37,38]. (Scheme 1). In addition to the well-discovered MLCT of coordination complexes [39,40] and organic promoters [2,41,42], electron excitation processes supported by redox complexes [43], novel PCET process [44], and unexpected excitations of organometallic compounds were also applied as efficient catalytic models to trigger photoactivation [5,6,45]. Such photocatalysts have been used for challenging applications, such as getting clean energy fuel by the splitting of water molecules [46,47] and reducing carbon dioxide into methane [48,49].
The recent synthetic method development of alkoxy radicals has promoted various applications. This review focuses on the recent development of alkoxy radical generation and natural product syntheses by using alkoxy radicals as key intermediates. In the first part, the recent development of alkoxy radical generation methods will be summarized in chronological order; the second part will present the recent applications of alkoxy radicals in natural product syntheses.

2. Photocatalytic Generation of Alkoxy Radicals

In the year 2014, DiRocco and colleagues described a late-stage functionalization of heterocycles strategy through photoredox catalysis. The tert-Butoxy radical (t-BuO●) was produced by using an IrIII catalyst, 450 nm light, and trifluoracetic acid (TFA). The key species in the transformation, methyl radical, is achieved through β-scission of the t-BuO●. The addition of the methyl radical to heterocycles realized the direct Csp2−Hydrogen alkylation of heteroarenes. This unique reaction condition was used to activate bioactive molecules, including eszopiclone, diflufenican, fasudil, camptothecin, fenarimol, vernicline, caffeine, and voriconazole in moderate yields (40–77%, Scheme 2) [50]. The method would potentially expand the late-stage functionalization toolbox for medicinal chemists.
The utilization of fluorinated ketones presents a convenient and effective means of introducing fluorine atoms into intricate molecular structures. While the incorporation of α-fluorinated ketones is relatively straightforward, achieved by converting the acidic α-hydrogen of the ketone through various electrophilic fluorinating agents, the direct fluorination process for synthesizing fluorinated ketones remains infrequently documented. However, a notable breakthrough has been made by Zhu and colleagues, who devised an efficient methodology for producing β- and γ-fluorinated ketones. The Ag-catalyzed ring-opening approach involves a sequential sequence of bond cleavage and formation reactions, enabling the generation of a broad range of substrates with favorable regioselectivity. In relation to the potential mechanism, three pathways may be involved. Firstly, when the electrophilic fluorinating species is present, the four-membered ring can potentially undergo direct opening without requiring the Ag catalyst (Path A). Another possibility is the Ag-assisted ring-opening process (Path B). Similar to the RhI-catalyzed ring-opening reactions, AgI coordinates with the hydroxyl group and inserts itself into the neighboring C−C bond, resulting in the formation of a C−Ag bond. After reacting with SelectFluor, the resulting AgIII intermediate endures reductive elimination, leading to the formation of the desired product. The third pathway involves a radical mechanism (Path C) (Scheme 3) [51].
In 2016, Chen’s group reported another work on an effective protocol for the ring-opening chlorination process of cyclobutanols via alkoxy radical formation using a manganese catalyst. These mild reaction conditions were applied to various functionalities of acyclic ketones and cyclic chlorides and achieved excellent product yield. The efficient method could lead to an important intermediate, which is utilized for direct access to biologically active compounds such as Haloperidol, Fluanisone, and Fupailliduo (Scheme 4) [52].
Chen and coworkers described a process for the photocatalytic-promoted generation of oxygen-centered radicals using N-alkoxy-phthalimide as a radical precursor. The authors performed luminescence quenching experiments and showed that the Hantzsch ester demonstrates efficient quenching of the photoexcited fac-[Ir(ppy)3]*, indicating that the predominant reaction pathway for generating the IrII intermediate is through reductive quenching. This protocol was applied to achieve selective allylation and alkenylation on activated and inactivated Csp3-H bonds. Furthermore, this alkoxyl radical system was utilized in the functionalization of complex molecules or biomolecules (Scheme 5) [53].
In the year 2021, Xiao Shen and coworkers introduced 1,2-silyl transfer to functionalized allylic sulfones. The key factor of this approach is that when the silyl group attaches to the α-position of alcohol, it can avoid β-carbon elimination. This synthetic approach was used for the synthesis of L(-) Menthol, Epiandrosterone, and Diosgenin analogs in great yields (72–85%, Scheme 6) [54].
Moreover, Meggers’s research group presented a combination of photoredox systems with chiral Lewis acid catalysis for C-H activation through alkoxy radical chemistry. This protocol shows significant tolerance on a wide array of functionalities of C-H bonds and simultaneously introduces two stereocenters. This approach utilizes a photoredox-mediated transformation of an alkoxy phthalimide compound into an alkoxy radical that subsequently converts to a carbon radical through 1,5-HAT. The resulting carbon radical then undergoes an intermolecular addition to an alkene species. It uses recently discovered redox-active radical intermediate, namely N-alkoxyphthalimides and N-acylpyrazoles, as new acidic functionality (Scheme 7) [55]. The method provided not only good-to-great yields (54–86%) but also good-to-excellent enantioselectivity (86% ee–97% ee).
In the same year, Chen et al.’s other report showed a ruthenium photocatalytic system leads to the formation of structurally complex steroidals using visible light energy. Both alkenyl and alkynyl functionality were introduced in strained cycloalkanol with the assistance of benziodoxole reagents. In this report, visible light-induced alcohol oxidation generated alkoxy radical species via cyclic iodine intermediate (Scheme 8) [56].
In 2017, Chen and coworkers described the selective carbonyl-C(sp3) bond scission in various functional groups such as amides, esters, and alcohols in a regio- and chemoselective manner under modest conditions using a ruthenium photocatalyst. The reaction benefits from the catalytic properties of the cyclic iodine-III reagent, affording model compounds effectively. These findings support the hypothesis that the complex of β-carbonyl alcohol and CIR works as a crucial intermediate in the process. Within the photocatalytic system, this intermediate undergoes oxidation, leading to the β-carbonyl alkoxyl radical. In the photocatalytic process, it is hypothesized that the Ru(bpy)32+* species was obtained upon light irradiation and was subsequently oxidized to Ru(bpy)33+. Following this step, Ru(bpy)33+ oxidizes the in situ formed β-carbonyl alcohol/CIR intermediate, liberating the CIR cation (CIR+) and initiating another CIR catalytic cycle. The resulting alkoxyl radical undertakes a bond cleavage reaction at the β-carbonyl-C(sp3) position. Such a process can lead to the formation of carbamoyl, alkoxylcarbonyl, or acyl radicals, which subsequently undergo radical addition to the alkynyl benziodoxole. This method is applied for the synthesis of complex synthons such as ynamides, ynoates, and ynones (Scheme 9) [57].
Meanwhile, the Knowles group has successfully developed a unique ring-opening protocol that proceeds via β-scission, applying an iridium photocatalyst induced by a visible light source. This protocol achieved isomerization of cyclic alcohol that led to the synthesis of a hecogenin analog. Mechanistic studies reveal that the proton-coupled electron transfer approach triggered O-H bond homolysis and delivered a oxygen-centered radical. It is proposed that the radical cation species attracted the electron in coordination with a proton moves to a weak base (Scheme 10) [58].
In the same year, Zuo and coworkers published their work based on CeCl3-induced alcohol oxidation to generate alkoxyl radicals from unstrained cycloalkanols and achieve amination of the C-C bond of cycloalkanol. This strategy is utilized to rapidly structure complex biomolecules such as Talampanel. Mechanistic investigation disclosed that the cerium alkoxide could be easily photoexcited by visible light and later converted to the corresponding Ce (IV) intermediate. Then, this species could lead to the challenging β-fragmentation and produce an alkyl radical (Scheme 11) [59].
Achieving control over the specific location and type of C-C bond formation when utilizing unactivated C(sp3)-H bonds is a difficult task. However, the use of alkoxyl radicals presents an effective strategy for activating C-H bonds through 1,5-hydrogen atom transfer (1,5-HAT) reactions, mainly due to the higher energy of the O-H bond. When employing different alcohols, such as primary alcohols, in 1,5-HAT reactions, the choice of alcohol can influence both the reactivity and stability of the resulting alkoxy radicals. Primary alcohols generally give rise to more stable alkoxy radicals compared to secondary or tertiary alcohols. This enhanced stability can be attributed to the presence of a single alkyl group attached to the oxygen in primary alcohols, forming a less hindered radical center. The Zuo research group applied a photoinduced LMCT activation mode to activate the free O-H bond in alcohols and produced an alkoxy intermediate using a low-priced cerium photocatalyst via a 1,5 HAT process. This effective catalytic protocol offers selective C-H bond activation using a carbon-linked oxygen radical intermediate. This method delivers novel and simple reaction conditions to achieve a wide array of complex molecules from simple and plentiful alcohols (Scheme 12) [60].
The same group has described a simple and practical photocatalytic methodology to synthesize bridged lactones through the alkoxy radical approach. Additionally, this protocol was applied for bulk scale synthesis using flow reactors for polycyclic framework. After careful investigation, they hypothesized that upon irradiation, ligand metal charge transfer would lead to homolysis and produce the key carbon-linked oxygen intermediate. After that, this transient radical species initiates a β-fragmentation in cycloalkanol and delivers a carbon radical species. Then, the resultant intermediate is coupled with alkene and creates an additional carbon−carbon bond (Scheme 13) [61].
During the same period, Zhu’s group presented an advanced photocatalytic ring-opening strategy for various unstrained cyclic alcohols via cyclic C–C cleavage generating alkoxy radicals. This process delivers valuable feedstocks for the production of haloperidol analogs. This group postulated two pathways for alkoxy radical generation. Path a shows that alkoxy radical can be generated via the PCET process employing the Iridium complex and succinimide anion. Alternatively, path b presents that the interaction between cycloalkanol and PIDA leads to homolysis of the O-I bond and produces challenging transient O-radical species (Scheme 14) [62].
The Knowles group presented the homolytic activation of free alcohol in cycloalkanols via the PCET strategy. This innovative approach could be applied in the hydrogenation of structural derivatives of cholesterol, botulin, pentose, and hexose in great yields (63–93%). This protocol certainly leads as an instance for investigating a new chemical space by activating the small bio-molecules (Scheme 15) [63].
Rueping and coworkers developed a PCET-enabled nickel photocatalytic methodology for site-specific functionalization of ketones from readily available 3° alcohols by using a strongly oxidizing acridinium salt, which gained high reaction efficiency. This process displays that the lower bond-dissociation free energy of alcohol caused rapid O-radical generation and led to the formation of C-radicals. Then, alkyl radical species were captured by nickel(0), formed a nickel intermediate, and led to the structuring of the provocative C(sp3)−C(sp2) bond. Thus, a wide range of linear, cyclic, land-bridged alcohols can be successfully explored in this cross-coupling procedure and achieved late-stage modification on Celecoxib and Diacetone-D-galactose derivatives (Scheme 16) [64].
Shi’s group has successfully delivered the halogenation of various cycloalkanols using visible light energy via the ring-opening process. This method used a combination of tetrabutylammonium halide and phthaloyl peroxide under moderate conditions with blue LEDs, achieved a high tolerability for the sensitive functional group, and delivered a wide array of halogenated ketones. This strategy was adopted for the modification of natural product structural units and gained excellent yields. The authors proposed a mechanism according to both experimental results and density functional theory (DFT) calculations. The crucial role of intermediate B, which is generated by the blend of PPO and TBAB, was identified in the step of photolytic O−Br bond cleavage. This cleavage process resulted in the generation of two species: the intermediate C and the Br radical. Using cyclobutanol as the substrate, a simultaneous homolysis of the O−H bond and the strained C−C bond occurred. This concerted process directly yielded radical intermediate G and the byproduct F. On the other hand, cyclic alcohols with five- to eight-membered rings were proposed to go through a stepwise mechanism, initially undergoing hydrogen atom transfer with intermediate C to generate alkoxy radical E. Finally, carbon radical intermediate G reacted with the Br radical, resulting in the formation of the brominated ketones (Scheme 17) [65].
The highly reactive nature of the alkoxyl radicals makes it very useful in organic synthesis. Serving as a reactive intermediate, it allows for the wide range of functionalization of unactivated C-H bonds. When there are δ-C-H bonds in the molecule, the 1,5-HAT reaction is preferred, resulting in the abstraction of the δ-C-H bond. In other cases, when intramolecular δ-C-H bonds are absent, intermolecular HAT reactions become more prominent. It is important to note, however, that intramolecular C-H abstraction by alkoxyl radicals at positions other than the δ-position is less frequently observed because of the presence of unfavorable transition states and high activation energies. So far, there are limited studies on the 1,2-HAT reactivity of alkoxyl radicals, and the practical application of 1,2-HAT for the formation of new C-C bonds is still not well understood or explored. In the year 2020, Chen’s group demonstrated a selective allylation of α-trifluoromethyl, benzylic N-alkoxyl phthalimides, α-carbonyl, and α-cyano using the 1,2-HAT process. The density functional theory calculations, electron paramagnetic resonance results, and mechanistic experiment confirmed that the HAT process is led by the alkoxyl radicals and proved that the activation energy can be minimized by using proton donor solvents in order to accelerate an allylation reaction. This study allowed the rapid synthesis of structurally complex steroid molecules (Scheme 18) [66].
Shortly after, Hong et al. presented the visible light-enhanced phosphorylation of natural analogs such as coumarins and quinolinones through an alkoxy intermediate-promoted intermolecular hydrogen atom transfer process. They used diphenyl phosphine as a phosphor donor functionality in this reaction. In this methodology, the N-alkoxy-pyridinium precursor was used as an ethoxy radical donor and worked as the oxidant. This method could be applied for the mono-phosphonation of various functionalized quinolinones in good yields (55–87%, Scheme 19) [67]. This regioselective phosphonation strategy would allow a rapid preparation of 3-phosphonylated derivates, which are an essential structural motif in many bioactive compounds.
One year later, Zhu et al. showed alcohol-directed remote C-H activation. The formation of carbon-linked oxygen radical species from the free alcohol O-H bond under a photoredox system leads to successive hydrogen transfer and a heteroaryl moiety. This methodology delivers high tolerability for sensitive functionality to achieve late-stage activation from challenging alcohols (Scheme 20) [68].
Moreover, the same group developed a facile, economic, and metal-free methodology for non-hypohalite-mediated alkoxy radical generation under 100 W blue LED irradiation conditions, using iodine (III) reagents such as phenyl iodine bis(trifluoroacetate) (PIFA) as the alkoxy radical precursor. This process successfully delivered the Minisci-type products via intermolecular heteroarylation of alcohols. This reaction has displayed that all kinds (1°, 2°, and 3°) of alcohols can generate the alkoxy radical species and can deliver highly regioselective heteroarylation of C(sp3)–H bonds. This protocol structured analogs of voriconazole and eszopiclone (Scheme 21) [69].
In 2018, Hong et al. reported a simple and facile photocatalytic radical distal translocation strategy for Minisic-type heteroarylation and phosphorylation. This method enabled mild reaction conditions, and the quinolinone organic ligand works as a photoredox catalyst. The most important key factor is that the N-alkoxy-pyridinium precursor acts as an alkoxy radical donor and works as a hetero-arylating reagent. This protocol successfully accomplished late-stage modification of cholesterol, pentoxifylline, pyriproxyfen, and vismodegib in yields of 70%, 70%, 66%, and 86%, respectively (Scheme 22A) [70]. Shortly after, a similar strategy was further explored for the functionalization of pyridines, providing a phosphinoylation and carbonylation of pyridines containing bio-active molecules, such as bisacodyl, vismodegib, and pyriproxyfen, in moderate yields (Scheme 22B) [71].
Moreover, the Zuo group designed a simple and convenient visible light-induced strategy for the activation of alkanols via radical chemistry. This method effectively delivered hydroboration–oxidation transformations using blue LEDs and low-priced cerium photocatalysts. This operationally facile protocol shows that simple alcohol can lead to the rapid construction of challenging nucleoside structures via an alkoxy radical mechanism (Scheme 23) [72].
In addition, Zhu and coworkers reported a successful merging of visible light-induced functional group migration. The reactions were performed under moderate conditions by using an Ir photocatalyst and blue LEDs. Interestingly, BF3OEt2 was used as an additive for promoting the transformation. The reaction provided an unsymmetrical 1,8-dicarbonyl structural motif and utilized the synthesis of SAHA derivatives (Scheme 24) [73].
The Liu group reported a visible light-induced selective activation in challenging primary alcohols in 2019. This highly selective cleavage was performed under moderate conditions, and [bis(trifluoroacetoxy)iodo]-benzene (PIFA) was used as an oxidant and an activation reagent. Remarkably, a wide variety of natural structural products like sugars and steroid derivatives from ribofuranoside, glycerol, and xylitol were synthesized using this approach. In order to provide a more comprehensive mechanistic understanding, a series of experiments were devised by the author. Initially, the combination of compound (a) and PIFA in CDCl3 promptly yielded a transient intermediate (b), which was deduced based on the analysis of crude 1H-NMR and 13C-NMR. However, when PIDA was reacted with alcohol (a), intermediate (b) was not formed. These findings provide an explanation for the lack of reactivity exhibited by PIDA in cleaving the C(sp3)-C(sp3) bond in alcohols. Subsequently, the reaction of intermediate (b) with lepidine proceeded smoothly, affording the desired product in 73% yield. A radical clock experiment was executed to prove the radical process. The PIFA-promoted reaction between lepidine and 2-cyclopropylethan-1-ol (d) resulted in the ring-opening product (e) in a yield of 65%. These results strongly support the involvement of a free radical pathway, specifically the β-scission of the alkoxyl radical. Based on the experimental results, the authors hypothesized a mechanism and showed that, firstly, PIFA performed nucleophilic substitution with alcohol, resulting in intermediate A. Light-promoted homolytic cleavage of intermediate A delivered a carbon-linked oxygen-centered radical. Subsequently, the alkoxy species experienced β-fragmentation to offer carbon radical species and formaldehyde. Then, the carbon radical reacted with heteroarene and led to the formation of cation B. After that, the radical cation led to single electron oxidation and resulted in the final product (Scheme 25) [74].
Shortly after, González’s group designed an approach using an organic photocatalyst PhIO and I2 combined system and applied it for the synthesis of guanidine derivatives such as (+) Saxitoxin, Crambescin, Tetrodotoxin, and Monanchorin. A key factor in this protocol is that the labile functional group rapidly generates the alkoxy radical intermediate via homolytic scission, which triggers β-fragmentation on the C1-C2 bond and leads to the formation of a new carbon-centered radical intermediate at position C-2 (Scheme 26) [75]. Although the yields is not ideal (5–49%), the method indeed opened a new avenue to the formation of medium-sized guanidine-containing heterocycles.
The Zhu group reported a metal-free coupling reaction between heteroarenes and simple alkenes via amidyl/alkoxy radical chemistry. This approach delivered mild and neutral reaction conditions and excellent yield for the Minisci reaction products. The interaction between amide/alcohol and PIFA plays a lead role in producing the amidyl/alkoxy radical intermediates. The author presents a proposed mechanism wherein the interaction between a PIFA (PhI(OCOCF3)2) and an amide or alcohol substrate initiates the formation of intermediate I or II. Upon exposure to visible light, these intermediates undergo homolysis, forming the corresponding amidyl or alkoxy radicals (III or IV). Simultaneously, an iodanyl radical is generated at the same time. The N-radical (or O-radical) then abstracts a hydrogen atom from 2 intermolecularly, leading to C-radical V (path a). Instead, the iodanyl radical can engage in another potential hydrogen atom transfer (HAT) process (path b). It is important to note that the presence of TFA (trifluoroacetic acid) in the reaction mixture activates the N-heteroarene, eliminating the need for additional acid. The addition of alkyl radical intermediate V to the activated heteroarene produces intermediate VI, which would be oxidized to the final product 3. This novel strategy is applied to the green synthetic process in natural product derivatives (Scheme 27) [76].
Zuo and coworkers have developed a practical and comprehensive strategy for selectively breaking and modifying C-C bonds within ketones. They have successfully harnessed the ligand-to-metal charge transfer (LMCT) excitation mode to specifically cleave C-C bonds in ketones while simultaneously introducing diverse functional groups to the substrates. This process is carried out under straightforward conditions, utilizing cost-effective cerium catalysts and simple blue LED light. Remarkably, this method demonstrates broad applicability to a wide range of substrates. Both cyclic ketones and acyclic ones work well under this condition. In addition, the reaction condition could be applied to not only simple ketones but also complex functionalized androsterone, thereby enabling their transformation into versatile chemical building blocks. It is noteworthy that this photocatalytic process achieves exceptional regioselectivity in all cases. Such an innovative photocatalytic approach offers a great alternative to the Norrish type I reaction with more selectivity, paving the way to exciting opportunities in more C-C bond cleavage synthetic strategies (Scheme 28) [77].
At the same time, the Li group demonstrated the alkene functionalization via an in situ-produced oxygen radical species-promoted intermolecular hydrogen atom transfer process by using the organo-photocatalyst 4CzIPN. This method can be applied to realize the carbonylation of natural structural units such as bexarotene, menthol, and estrone (Scheme 29) [78].
In the same year, Liu’s group demonstrated a copper-mediated enantioselective cyanation for the synthesis of enantioenriched β-carbonyl nitriles in outstanding yield. Additionally, the chiral β-cyano propanoic ester formed from the ring opening of cyclo-propanone acetals. The strategy can be applied in the synthesis of GABA receptor agonist (R)-baclofen, which acts as an inhibitory neurotransmitter (Scheme 30) [79].
Recently, the Rueping group used the photoredox PCET and cobalt synergistic catalytic combination for generating alkoxy radicals from a range of alcohols. This process delivered selective ring-opening of various cyclic alcohols and produced dehydrogenated ketones that are challenging to deliver with current methodologies. Both secondary and tertiary alcohols are well accommodated in this reaction. The strategy can be applied to structures of natural product derivatives such as (-)-Bomeol, Norcamphor, and Pregenenolone (Scheme 31) [80].
In addition, the Hong group showed a protocol for site-selective pyridine modification via an alkoxy-centered radical strategy. This approach enabled the concurrent inclusion of pyridyl moiety in architecturally distinct β-carbonyl functionality with outstanding C4 selectivity. This advanced process has been further explored for the selective functionalization of naturally structured designs such as lithocholic acid, ibuprofen, vismodegib, and pyriproxyfen (Scheme 32) [81].

3. Application of Alkoxy Radicals in Natural Product Synthesis

In addition to the development of new synthetic methods, alkoxy radicals have been explored in several elegant syntheses of natural products. In the year 2016, Gilmour’s group showed a novel and practical approach for Z/E configuration alteration and cyclic lactonization of natural product coumarins by employing the organo-catalyst (-)-riboflavin under 402 LED light irradiation, forming in situ alkoxy radical species (Scheme 33) [82].
In 2019, the Wang group successfully performed a total synthesis of the hitorins using alkoxyl radical chemistry. The interaction between alkoxy intermediate and monoterpene (+)-sabinene provided the corresponding tetrahydrofuran ring of hitorins. The possible mechanism involves generating an alkoxy radical from hydroperoxide using Fe(II) as a reducing agent, which then reacts with (+)-sabinene to produce a carbon-centered radical. The syntheses of hitorin A and B were realized by either radical oxidation by Cu(II) or β-H elimination of intermediate (Scheme 34) [83].
In the same year, the Smith group designed a novel photoredox approach in order to produce dithianyl and dioxolanyl radical species via the HAT process. The resultant radical intermediates are involved in conjugate additions in order to reach the formal alkylation and allylation. This radical reaction pathway offers an effective way to structure a variety of 1,4,7-polyols or spiro-ketals, which have been applied in the synthesis of danshenspiroketallactones (Scheme 35) [84].
A recent example of employing a visible photoredox-promoted strategy for the synthesis of (−)- and (+)-polyoxamic acid from sugar molecules was delivered by the Ohno group. This methodology involves blue light-induced β-fragmentation and a 1,5-HAT strategy to afford the alditol synthesis using modest reaction parameters (Scheme 36) [85].

4. Conclusions and Outlook

The effective organic transformation for complex structures demands both a practical approach and the utilization of powerful bond-forming reactions. The effective approaches diminish the unnecessary steps in synthesis and rapidly form new bonds of the target structure. The chemical transformations via radical chemistry deliver an ideal platform for lowering the step for natural product synthesis. This concise review concludes the latest discoveries in the alkoxy radical facilitated organic transformations. Recently, the growing interest in photoredox catalysis broadened the domain of alkoxy radical-promoted organic conversion at an extraordinary pace. These novel and efficient strategies avoid the pre-functionalization of the substrate and stoichiometric reagent and are further explored for late-stage modification in structurally elaborate molecules by alkoxy radical species, including selective activation of the C-H bond, C-N and C-C bond cleavage, and C-O bond shaping. Although alkoxy radical utilization in these activations has broadened the interest in novel chemical transformation, several challenges remain. We anticipate that the discovery of novel alkoxy-radical precursors and practically applicable approaches will further innovate in chemical transformations and lead to an advanced level of natural product synthesis.

Funding

National Institute of Health (1R15GM134479).

Acknowledgments

We are grateful to the National Institute of Health (1R15GM134479) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

BDEBond dissociation energy
BIBenziodoxole
CIRCyclic iodine(III) reagents
DBADDi-tert-butyl azodicarboxylate
DPA(9,10-diphenylanthracene)
EPRElectron paramagnetic resonance
HATHydrogen atom transfer
HPLCHigh-performance liquid chromatography
HEHantzsch ester
LMCTLigand-to-metal charge transfer
LEDLight-emitting diode
MLCTMetal-to-ligand charge transfer
PCETProton-coupled electron transfer
PMPp-methoxyphenyl
PIDAIodobenzene diacetate
PIFA[bis(trifluoroacetoxy)iodo]benzene
PPOPhthaloyl peroxide
PTProton Transfer
TBABTetrabutylammonium bromide
4CzIPN2,4,5,6-tetra(carbazol-9-yl)benzene-1,3-dicarbonitrile

References

  1. Tsui, E.; Wang, H.; Knowles, R.R. Catalytic Generation of Alkoxy Radicals from Unfunctionalized Alcohols. Chem. Sci. 2020, 11, 11124–11141. [Google Scholar] [CrossRef] [PubMed]
  2. Chang, L.; An, Q.; Duan, L.; Feng, K.; Zuo, Z. Alkoxy Radicals See the Light: New Paradigms of Photochemical Synthesis. Chem. Rev. 2022, 122, 2429–2486. [Google Scholar] [CrossRef] [PubMed]
  3. Barton, D.H.R.; Beaton, J.M. A Synthesis of Aldosterone Acetate. J. Am. Chem. Soc. 2002, 82, 2641. [Google Scholar] [CrossRef]
  4. Barton, D.R.; Beaton, J.M.; Geller, L.E.; Pechet, M.M. A New Photochemical Reaction1. J. Am. Chem. Soc. 2002, 83, 4076–4083. [Google Scholar] [CrossRef]
  5. Jia, K.; Chen, Y. Visible-Light-Induced Alkoxyl Radical Generation for Inert Chemical Bond Cleavage/Functionalization. Chem. Commun. 2018, 54, 6105–6112. [Google Scholar] [CrossRef] [PubMed]
  6. Wu, X.; Zhu, C. Recent Advances in Alkoxy Radical-Promoted C–C and C–H Bond Functionalization Starting from Free Alcohols. Chem. Commun. 2019, 55, 9747–9756. [Google Scholar] [CrossRef]
  7. Bietti, M.; Lanzalunga, O.; Salamone, M. Structural Effects on the β-Scission Reaction of Alkoxyl Radicals. Direct Measurement of the Absolute Rate Constants for Ring Opening of Benzocycloalken-1-Oxyl Radicals. J. Org. Chem. 2005, 70, 1417–1422. [Google Scholar] [CrossRef]
  8. Morcillo, S.P. Radical-Promoted C−C Bond Cleavage: A Deconstructive Approach for Selective Functionalization. Angew. Chem. Int. Ed. 2019, 58, 14044–14054. [Google Scholar] [CrossRef]
  9. Yu, X.Y.; Chen, J.R.; Xiao, W.J. Visible Light-Driven Radical-Mediated C-C Bond Cleavage/Functionalization in Organic Synthesis. Chem. Rev. 2021, 121, 506–561. [Google Scholar] [CrossRef]
  10. Hartung, J. Stereoselective Construction of the Tetrahydrofuran Nucleus by Alkoxyl Radical Cyclizations. Eur. J. Org. Chem. 2001, 2001, 619–632. [Google Scholar] [CrossRef]
  11. de Armas, P.; Francisco, C.G.; Suárez, E. Fragmentation of Carbohydrate Anomeric Alkoxy Radicals. Tandem β-Fragmentation-Cyclization of Alcohols. J. Am. Chem. Soc. 1993, 115, 8865–8866. [Google Scholar] [CrossRef]
  12. Martin, A.; Salazar, J.A.; Suárez, E. Synthesis of Chiral Spiroacetals from Carbohydrates. J. Org. Chem. 1996, 61, 3999–4006. [Google Scholar] [CrossRef] [PubMed]
  13. Zhu, H.; Wickenden, J.G.; Campbell, N.E.; Leung, J.C.T.; Johnson, K.M.; Sammis, G.M. Construction of Carbo- and Heterocycles Using Radical Relay Cyclizations Initiated by Alkoxy Radicals. Org. Lett. 2009, 11, 2019–2022. [Google Scholar] [CrossRef] [PubMed]
  14. Yamashita, S.; Naruko, A.; Nakazawa, Y.; Zhao, L.; Hayashi, Y.; Hirama, M. Total Synthesis of Limonin. Angew. Chem. Int. Ed. 2015, 54, 8538–8541. [Google Scholar] [CrossRef] [PubMed]
  15. Xue, X.S.; Ji, P.; Zhou, B.; Cheng, J.P. The Essential Role of Bond Energetics in C–H Activation/Functionalization. Chem. Rev. 2017, 117, 8622–8648. [Google Scholar] [CrossRef] [PubMed]
  16. Pirts, J.N.; Tolberg, R.S.; Martin, T.W.; Thompson, D.D.; Woolfolk, R.W. Intramolecular Free-Radical Reactions. Angew. Chem. Int. Ed. English 1964, 3, 525–538. [Google Scholar]
  17. Abarca, R.M. Global manual on surveillance of adverse events following immunizaion. WHO Nuevos Sist. Comun. E Inf. 2021, 4, 2013–2015. [Google Scholar]
  18. Kundu, R.; Ball, Z.T. Copper-Catalyzed Remote Sp3 C-H Chlorination of Alkyl Hydroperoxides. Org. Lett. 2010, 12, 2460–2463. [Google Scholar] [CrossRef]
  19. Petrović, G.; Čeković, Ž. Alkylation of Remote Non-Activated δ-Carbon Atoms: Addition of δ-Carbon Radicals, Generated by 1,5-Hydrogen Transfer in Alkoxy Radical Intermediates, to Activated Olefins. Tetrahedron 1999, 55, 1377–1390. [Google Scholar] [CrossRef]
  20. Petrović, G.; Saičić, R.N.; Čeković, Ž. Synthesis of Acetyl Scopine. Intramolecular Reactions of N-Carbethoxy Nortropine-3α-Benzene-Sulfenate. Synlett 1999, 1999, 635–637. [Google Scholar] [CrossRef]
  21. Martín, A.; Pérez-Martín, I.; Suárez, E. Intramolecular Hydrogen Abstraction Promoted by Amidyl Radicals. Evidence for Electronic Factors in the Nucleophilic Cyclization of Ambident Amides to Oxocarbenium Ions. Org. Lett. 2005, 7, 2027–2030. [Google Scholar] [CrossRef] [PubMed]
  22. Hartung, J.; Gallon, F. Ring Closure Reactions of Substituted 4-Pentenyl-l-Oxy Radicals. J. Org. Chem 1995, 60, 6706–6716. [Google Scholar] [CrossRef]
  23. Kim, S.; Lee, T.A.; Song, Y. Facile Generation of Alkoxy Radicals from N-Alkoxyphthalimides. Synlett 1998, 5, 471–472. [Google Scholar] [CrossRef]
  24. Zlotorzynska, M.; Sammis, G.M. Photoinduced Electron-Transfer-Promoted Redox Fragmentation of N-Alkoxyphthalimides. Org. Lett. 2011, 13, 6264–6267. [Google Scholar] [CrossRef]
  25. Hoffmann, N. Photochemical Reactions as Key Steps in Organic Synthesis. Chem. Rev. 2008, 108, 1052–1103. [Google Scholar] [CrossRef]
  26. Souillart, L.; Cramer, N. Catalytic C–C Bond Activations via Oxidative Addition to Transition Metals. Chem. Rev. 2015, 115, 9410–9464. [Google Scholar] [CrossRef]
  27. Xuan, J.; Zhang, Z.G.; Xiao, W.J. Visible-Light-Induced Decarboxylative Functionalization of Carboxylic Acids and Their Derivatives. Angew. Chem. Int. Ed. 2015, 54, 15632–15641. [Google Scholar] [CrossRef]
  28. Yamaguchi, J.; Yamaguchi, A.D.; Itami, K. C-H Bond Functionalization: Emerging Synthetic Tools for Natural Products and Pharmaceuticals. Angew. Chem. Int. Ed. 2012, 51, 8960–9009. [Google Scholar] [CrossRef]
  29. R Barton, D.H.; Gellbr Cambridge, L.E.; Pechet, M.; Simpson, S.A.; Tait, T.F.; Wettstein, A.; Neher, R.; Euw, J.V.; Schindler, O.; Barton, H.R.; et al. A New Photochemical Reaction. J. Am. Chem. Soc. 2002, 82, 2640–2641. [Google Scholar] [CrossRef]
  30. Rivero, A.R.; Fodran, P.; Ondrejková, A.; Wallentin, C.J. Alcohol Etherification via Alkoxy Radicals Generated by Visible-Light Photoredox Catalysis. Org. Lett. 2020, 22, 8436–8440. [Google Scholar] [CrossRef]
  31. Pitre, S.P.; Overman, L.E. Strategic Use of Visible-Light Photoredox Catalysis in Natural Product Synthesis. Chem. Rev. 2021, 122, 1717–1751. [Google Scholar] [CrossRef] [PubMed]
  32. Genzink, M.J.; Kidd, J.B.; Swords, W.B.; Yoon, T.P. Chiral Photocatalyst Structures in Asymmetric Photochemical Synthesis. Chem. Rev. 2022, 122, 1654–1716. [Google Scholar] [CrossRef] [PubMed]
  33. Lechner, V.M.; Nappi, M.; Deneny, P.J.; Folliet, S.; Chu, J.C.K.; Gaunt, M.J. Visible-Light-Mediated Modification and Manipulation of Biomacromolecules. Chem. Rev. 2022, 122, 1752–1829. [Google Scholar] [CrossRef] [PubMed]
  34. Rossi-Ashton, J.A.; Clarke, A.K.; Unsworth, W.P.; Taylor, R.J.K. Phosphoranyl Radical Fragmentation Reactions Driven by Photoredox Catalysis. ACS Catal. 2020, 10, 7250–7261. [Google Scholar] [CrossRef]
  35. Stache, E.E.; Ertel, A.B.; Rovis, T.; Doyle, A.G. Generation of Phosphoranyl Radicals via Photoredox Catalysis Enables Voltage-Independent Activation of Strong C-O Bonds. ACS Catal. 2018, 8, 11134–11139. [Google Scholar] [CrossRef] [PubMed]
  36. Xiong, T.; Zhang, Q. New Amination Strategies Based on Nitrogen-Centered Radical Chemistry. Chem. Soc. Rev. 2016, 45, 3069–3087. [Google Scholar] [CrossRef]
  37. Yu, X.Y.; Zhao, Q.Q.; Chen, J.; Xiao, W.J.; Chen, J.R. When Light Meets Nitrogen-Centered Radicals: From Reagents to Catalysts. Acc. Chem. Res. 2020, 53, 1066–1083. [Google Scholar] [CrossRef]
  38. Bloom, S.; Zhou, L.; Lu, B.; Xiao, W.-J.; Chen, J.-R. Recent Advances in Visible-Light-Mediated Amide Synthesis. Molecules 2022, 27, 517. [Google Scholar]
  39. Prier, C.K.; Rankic, D.A.; MacMillan, D.W.C. Visible Light Photoredox Catalysis with Transition Metal Complexes: Applications in Organic Synthesis. Chem. Rev. 2013, 113, 5322–5363. [Google Scholar] [CrossRef]
  40. Twilton, J.; Le, C.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. [Google Scholar] [CrossRef]
  41. Romero, N.A.; Nicewicz, D.A. Organic Photoredox Catalysis. Chem. Rev. 2016, 116, 10075–10166. [Google Scholar] [CrossRef] [PubMed]
  42. Speckmeier, E.; Fischer, T.G.; Zeitler, K. A Toolbox Approach to Construct Broadly Applicable Metal-Free Catalysts for Photoredox Chemistry: Deliberate Tuning of Redox Potentials and Importance of Halogens in Donor-Acceptor Cyanoarenes. J. Am. Chem. Soc. 2018, 140, 15353–15365. [Google Scholar] [CrossRef] [PubMed]
  43. Crisenza, G.E.M.; Mazzarella, D.; Melchiorre, P. Synthetic Methods Driven by the Photoactivity of Electron Donor–Acceptor Complexes. J. Am. Chem. Soc. 2020, 142, 5461–5476. [Google Scholar] [CrossRef] [PubMed]
  44. Gentry, E.C.; Knowles, R.R. Synthetic Applications of Proton-Coupled Electron Transfer. Acc. Chem. Res. 2016, 49, 1546–1556. [Google Scholar] [CrossRef] [PubMed]
  45. Murakami, M.; Ishida, N. β-Scission of Alkoxy Radicals in Synthetic Transformations. Chem. Lett. 2017, 46, 1692–1700. [Google Scholar] [CrossRef]
  46. Gratzel, M. Artificial Photosynthesis: Water Cleavage into Hydrogen and Oxygen by Visible Light. Acc. Chem. Res. 2002, 14, 376–384. [Google Scholar] [CrossRef]
  47. Meyer, T.J. Chemical Approaches to Artificial Photosynthesis. Acc. Chem. Res. 2002, 22, 163–170. [Google Scholar] [CrossRef]
  48. Takeda, H.; Ishitani, O. Development of Efficient Photocatalytic Systems for CO2 Reduction Using Mononuclear and Multinuclear Metal Complexes Based on Mechanistic Studies. Coord. Chem. Rev. 2010, 254, 346–354. [Google Scholar] [CrossRef]
  49. Yamazaki, Y.; Takeda, H.; Ishitani, O. Photocatalytic Reduction of CO2 Using Metal Complexes. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 106–137. [Google Scholar] [CrossRef]
  50. 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. [Google Scholar] [CrossRef]
  51. Zhao, H.; Fan, X.; Yu, J.; Zhu, C. Silver-Catalyzed Ring-Opening Strategy for the Synthesis of β- and γ-Fluorinated Ketones. J. Am. Chem. Soc. 2015, 137, 3490–3493. [Google Scholar] [CrossRef] [PubMed]
  52. Huan, L.; Zhu, C. Manganese-Catalyzed Ring-Opening Chlorination of Cyclobutanols: Regiospecific Synthesis of γ-Chloroketones †. Org. Chem. Front. 2016, 3, 1467. [Google Scholar] [CrossRef]
  53. Zhang, J.; Li, Y.; Zhang, F.; Hu, C.; Chen, Y. Generation of Alkoxyl Radicals by Photoredox Catalysis Enables Selective C(Sp3)−H Functionalization under Mild Reaction Conditions. Angew. Chem. Int. Ed. 2016, 55, 1872–1875. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, Y.; Zhang, Y.; Shen, X. Alkoxyl-Radical-Mediated Synthesis of Functionalized Allyl Tert-(Hetero)Cyclobutanols and Their Ring-Opening and Ring-Expansion Functionalizations. Chem. Catal. 2021, 1, 423–436. [Google Scholar] [CrossRef]
  55. Wang, C.; Harms, K.; Meggers, E. Catalytic Asymmetric C−H Functionalization under Photoredox Conditions by Radical Translocation and Stereocontrolled Alkene Addition. Angew. Chem. Int. Ed. 2016, 55, 13495–13498. [Google Scholar] [CrossRef]
  56. Jia, K.; Zhang, F.; Huang, H.; Chen, Y. Visible-Light-Induced Alkoxyl Radical Generation Enables Selective C(Sp3)-C(Sp3) Bond Cleavage and Functionalizations. J. Am. Chem. Soc. 2016, 138, 1514–1517. [Google Scholar] [CrossRef]
  57. Jia, K.; Pan, Y.; Chen, Y. Selective Carbonyl-C(Sp 3 ) Bond Cleavage To Construct Ynamides, Ynoates, and Ynones by Photoredox Catalysis. Angew. Chem. Int. Ed. 2017, 56, 2478–2481. [Google Scholar] [CrossRef]
  58. Yayla, H.G.; Wang, H.; Tarantino, K.T.; Orbe, H.S.; Knowles, R.R. Catalytic Ring-Opening of Cyclic Alcohols Enabled by PCET Activation of Strong O-H Bonds. J. Am. Chem. Soc. 2016, 138, 10794–10797. [Google Scholar] [CrossRef]
  59. Guo, J.-J.; Nhua Hu, A.; Chen, I.; Sun, J.; Tang, H.; Zuo, Z.; Guo, J.; Hu, A.; Chen, Y.; Sun, J.; et al. Photocatalytic C−C Bond Cleavage and Amination of Cycloalkanols by Cerium(III) Chloride Complex. Angew. Chem. Int. Ed. 2016, 55, 15319–15322. [Google Scholar] [CrossRef]
  60. Hu, A.; Guo, J.J.; Pan, H.; Tang, H.; Gao, Z.; Zuo, Z. δ-Selective Functionalization of Alkanols Enabled by Visible-Light-Induced Ligand-to-Metal Charge Transfer. J. Am. Chem. Soc. 2018, 140, 1612–1616. [Google Scholar] [CrossRef]
  61. Hu, A.; Chen, Y.; Guo, J.J.; Yu, N.; An, Q.; Zuo, Z. Cerium-Catalyzed Formal Cycloaddition of Cycloalkanols with Alkenes through Dual Photoexcitation. J. Am. Chem. Soc. 2018, 140, 13580–13585. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, D.; Mao, J.; Zhu, C. Visible Light-Promoted Ring-Opening Functionalization of Unstrained Cycloalkanols via Inert C–C Bond Scission. Chem. Sci. 2018, 9, 5805–5809. [Google Scholar] [CrossRef] [PubMed]
  63. Ota, E.; Wang, H.; Frye, N.L.; Knowles, R.R. A Redox Strategy for Light-Driven, Out-of-Equilibrium Isomerizations and Application to Catalytic C-C Bond Cleavage Reactions. J. Am. Chem. Soc. 2019, 141, 1457–1462. [Google Scholar] [CrossRef]
  64. Huang, L.; Ji, T.; Rueping, M. Remote Nickel-Catalyzed Cross-Coupling Arylation via Proton-Coupled Electron Transfer-Enabled C-C Bond Cleavage. J. Am. Chem. Soc. 2020, 142, 3532–3539. [Google Scholar] [CrossRef] [PubMed]
  65. Zhao, R.; Yao, Y.; Zhu, D.; Chang, D.; Liu, Y.; Shi, L. Visible-Light-Enhanced Ring Opening of Cycloalkanols Enabled by Brønsted Base-Tethered Acyloxy Radical Induced Hydrogen Atom Transfer-Electron Transfer. Org. Lett. 2018, 20, 1228–1231. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, J.; Liu, D.; Liu, S.; Ge, Y.; Lan, Y.; Chen, Y. Visible-Light-Induced Alkoxyl Radicals Enable α-C(Sp3)-H Bond Allylation. Iscience 2020, 23, 100755. [Google Scholar]
  67. Kim, I.; Min, M.; Kang, D.; Kim, K.; Hong, S. Direct Phosphonation of Quinolinones and Coumarins Driven by the Photochemical Activity of Substrates and Products. Org. Lett. 2017, 19, 1394–1397. [Google Scholar] [CrossRef]
  68. Wu, X.; Wang, M.; Huan, L.; Wang, D.; Wang, J.; Zhu, C. Tertiary-Alcohol-Directed Functionalization of Remote C(Sp3)−H Bonds by Sequential Hydrogen Atom and Heteroaryl Migrations. Angew. Chem. Int. Ed. 2018, 57, 1640–1644. [Google Scholar] [CrossRef]
  69. Wu, X.; Zhang, H.; Tang, N.; Wu, Z.; Wang, D.; Ji, M.; Xu, Y.; Wang, M.; Zhu, C. Metal-Free Alcohol-Directed Regioselective Heteroarylation of Remote Unactivated C(Sp3)–H Bonds. Nat. Commun. 2018, 9, 3343. [Google Scholar] [CrossRef]
  70. Kim, I.; Park, B.; Kang, G.; Kim, J.; Jung, H.; Lee, H.; Baik, M.H.; Hong, S. Visible-Light-Induced Pyridylation of Remote C(Sp3)−H Bonds by Radical Translocation of N-Alkoxypyridinium Salts. Angew. Chem. Int. Ed. 2018, 57, 15517–15522. [Google Scholar] [CrossRef]
  71. Kim, I.; Kang, G.; Lee, K.; Park, B.; Kang, D.; Jung, H.; He, Y.T.; Baik, M.H.; Hong, S. Site-Selective Functionalization of Pyridinium Derivatives via Visible-Light-Driven Photocatalysis with Quinolinone. J. Am. Chem. Soc. 2019, 141, 9239–9248. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, K.; Chang, L.; An, Q.; Wang, X.; Zuo, Z. Dehydroxymethylation of Alcohols Enabled by Cerium Photocatalysis. J. Am. Chem. Soc. 2019, 141, 10556–10564. [Google Scholar] [CrossRef] [PubMed]
  73. 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. [Google Scholar] [CrossRef]
  74. Wang, Y.; Yang, L.; Liu, S.; Huang, L.; Liu, Z.Q. Surgical Cleavage of Unstrained C(Sp3)−C(Sp3) Bonds in General Alcohols for Heteroaryl C−H Alkylation and Acylation. Adv. Synth. Catal. 2019, 361, 4568–4574. [Google Scholar] [CrossRef]
  75. Santana, A.G.; González, C.C. Tandem Radical Fragmentation/Cyclization of Guanidinylated Monosaccharides Grants Access to Medium-Sized Polyhydroxylated Heterocycles. Org. Lett. 2020, 22, 8492–8495. [Google Scholar] [CrossRef]
  76. Zhu, C.; Shao, X.; Wu, X.; Wu, S. Metal-Free Radical-Mediated C(Sp3)-H Heteroarylation of Alkanes. Org. Lett. 2020, 22, 7450–7454. [Google Scholar]
  77. Chen, Y.; Du, J.; Zuo, Z. Selective C-C Bond Scission of Ketones via Visible-Light-Mediated Cerium Catalysis. Chem 2020, 6, 266–279. [Google Scholar] [CrossRef]
  78. Zheng, M.; Hou, J.; Zhan, L.W.; Huang, Y.; Chen, L.; Hua, L.L.; Li, Y.; Tang, W.Y.; Li, B.D. Visible-Light-Driven, Metal-Free Divergent Difunctionalization of Alkenes Using Alkyl Formates. ACS Catal. 2021, 11, 542–553. [Google Scholar] [CrossRef]
  79. Wu, L.; Wang, L.; Chen, P.; Guo, Y.L.; Liu, G. Enantioselective Copper-Catalyzed Radical Ring-Opening Cyanation of Cyclopropanols and Cyclopropanone Acetals. Adv. Synth. Catal. 2020, 362, 2189–2194. [Google Scholar] [CrossRef]
  80. Huang, L.; Ji, T.; Zhu, C.; Yue, H.; Zhumabay, N.; Rueping, M. Bioinspired Desaturation of Alcohols Enabled by Photoredox Proton-Coupled Electron Transfer and Cobalt Dual Catalysis. Nat. Commun. 2022, 13, 809. [Google Scholar] [CrossRef]
  81. Vellakkaran, M.; Kim, T.; Hong, S. Visible-Light-Induced C4-Selective Functionalization of Pyridinium Salts with Cyclopropanols. Angew. Chem. Int. Ed. 2022, 61, e202113658. [Google Scholar] [CrossRef] [PubMed]
  82. Metternich, J.B.; Gilmour, R. One Photocatalyst, n Activation Modes Strategy for Cascade Catalysis: Emulating Coumarin Biosynthesis with (-)-Riboflavin. J. Am. Chem. Soc. 2016, 138, 1040–1045. [Google Scholar] [CrossRef]
  83. Li, X.; Zhai, T.; He, P.; Wei, Z.; Wang, Z. Biomimetic Synthesis of Hitorins A and B via an Intermolecular Alkoxy Radical-Olefin Coupling Cascade. In ChemRxiv; Cambridge Open Engage: Cambridge, UK, 2019. [Google Scholar] [CrossRef]
  84. Deng, Y.; Nguyen, M.D.; Zou, Y.; Houk, K.N.; Smith, A.B. Generation of Dithianyl and Dioxolanyl Radicals Using Photoredox Catalysis: Application in the Total Synthesis of the Danshenspiroketallactones via Radical Relay Chemistry. Org. Lett. 2019, 21, 1708–1712. [Google Scholar] [CrossRef] [PubMed]
  85. Matsuoka, T.; Inuki, S.; Miyagawa, T.; Oishi, S.; Ohno, H. Total Synthesis of (+)-Polyoxamic Acid via Visible-Light-Mediated Photocatalytic β-Scission and 1,5-Hydrogen Atom Transfer of Glucose Derivative. J. Org. Chem. 2020, 85, 8271–8278. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Alkoxy radical promoted organic transformation for natural motifs.
Scheme 1. Alkoxy radical promoted organic transformation for natural motifs.
Organics 04 00033 sch001
Scheme 2. Late-stage functionalization of complex heterocycles.
Scheme 2. Late-stage functionalization of complex heterocycles.
Organics 04 00033 sch002
Scheme 3. Synthesis of ɣ- and β-fluorinated ketones and possible mechanistic pathways.
Scheme 3. Synthesis of ɣ- and β-fluorinated ketones and possible mechanistic pathways.
Organics 04 00033 sch003
Scheme 4. Ring-opening transformation for preparation of bioactive molecules.
Scheme 4. Ring-opening transformation for preparation of bioactive molecules.
Organics 04 00033 sch004
Scheme 5. Alkoxyl radical enabled Csp3−H functionalization.
Scheme 5. Alkoxyl radical enabled Csp3−H functionalization.
Organics 04 00033 sch005
Scheme 6. Allylic sulfone activation via 1,2-silyl transfer. a = [Ir(dF(CF3)ppy)2(5,50 -d(CF3)bpy)]PF6 (2 mol %) was used as the catalyst instead of Eosin Y. b = No desilylation step.
Scheme 6. Allylic sulfone activation via 1,2-silyl transfer. a = [Ir(dF(CF3)ppy)2(5,50 -d(CF3)bpy)]PF6 (2 mol %) was used as the catalyst instead of Eosin Y. b = No desilylation step.
Organics 04 00033 sch006
Scheme 7. Photoredox catalytic C-H activation via radical approach.
Scheme 7. Photoredox catalytic C-H activation via radical approach.
Organics 04 00033 sch007
Scheme 8. Visible light-promoted carbon–carbon bond cleavage and functionalization.
Scheme 8. Visible light-promoted carbon–carbon bond cleavage and functionalization.
Organics 04 00033 sch008
Scheme 9. Visible light-mediated alkynylation of β-carbonyl alcohols.
Scheme 9. Visible light-mediated alkynylation of β-carbonyl alcohols.
Organics 04 00033 sch009
Scheme 10. PCET-enabled activation on unstrained cyclic alcohols. PMP = para-methoxyphenyl.
Scheme 10. PCET-enabled activation on unstrained cyclic alcohols. PMP = para-methoxyphenyl.
Organics 04 00033 sch010
Scheme 11. Cerium photocatalyzed reaction of cycloalkanols.
Scheme 11. Cerium photocatalyzed reaction of cycloalkanols.
Organics 04 00033 sch011
Scheme 12. δ selective-C–H functionalization of alcohols.
Scheme 12. δ selective-C–H functionalization of alcohols.
Organics 04 00033 sch012
Scheme 13. Visible photoredox cerium catalyst promoted β-fragmentation and ring-enlargement of cyclic alcohols for the structure complex bridged lactone.
Scheme 13. Visible photoredox cerium catalyst promoted β-fragmentation and ring-enlargement of cyclic alcohols for the structure complex bridged lactone.
Organics 04 00033 sch013
Scheme 14. Photoredox ring opening of unstrained cycloalkanols.
Scheme 14. Photoredox ring opening of unstrained cycloalkanols.
Organics 04 00033 sch014
Scheme 15. Hydrogenation of complex cyclic alcohol by PCET.
Scheme 15. Hydrogenation of complex cyclic alcohol by PCET.
Organics 04 00033 sch015
Scheme 16. Nickel-catalyzed arylation for late-stage modification.
Scheme 16. Nickel-catalyzed arylation for late-stage modification.
Organics 04 00033 sch016
Scheme 17. Alkoxy radical induced halogenation of cycloalkanols.
Scheme 17. Alkoxy radical induced halogenation of cycloalkanols.
Organics 04 00033 sch017
Scheme 18. Chemoselective allylation via 1,2-HAT process.
Scheme 18. Chemoselective allylation via 1,2-HAT process.
Organics 04 00033 sch018
Scheme 19. Phosphonation of quinolinones and coumarins.
Scheme 19. Phosphonation of quinolinones and coumarins.
Organics 04 00033 sch019
Scheme 20. Ir photocatalyst-induced heteroarylation and cyanation.
Scheme 20. Ir photocatalyst-induced heteroarylation and cyanation.
Organics 04 00033 sch020
Scheme 21. PIFA-mediated photocatalytic generation of alkoxy radicals.
Scheme 21. PIFA-mediated photocatalytic generation of alkoxy radicals.
Organics 04 00033 sch021
Scheme 22. Alkoxy radical mediated Minisic-type reaction and phosphorylation.
Scheme 22. Alkoxy radical mediated Minisic-type reaction and phosphorylation.
Organics 04 00033 sch022
Scheme 23. Hydroboration−oxidation and dehydroxymethylation.
Scheme 23. Hydroboration−oxidation and dehydroxymethylation.
Organics 04 00033 sch023
Scheme 24. Synthesis of SAHA derivatives.
Scheme 24. Synthesis of SAHA derivatives.
Organics 04 00033 sch024
Scheme 25. Blue light-induced selective activation in primary alcohols.
Scheme 25. Blue light-induced selective activation in primary alcohols.
Organics 04 00033 sch025
Scheme 26. Synthesis of cyclic guanidines.
Scheme 26. Synthesis of cyclic guanidines.
Organics 04 00033 sch026
Scheme 27. C−H functionalization of simple alkanes in metal-free condition.
Scheme 27. C−H functionalization of simple alkanes in metal-free condition.
Organics 04 00033 sch027
Scheme 28. Cerium catalyzed LMCT mode for activation of ketone.
Scheme 28. Cerium catalyzed LMCT mode for activation of ketone.
Organics 04 00033 sch028
Scheme 29. Metal-free divergent alkene functionalization.
Scheme 29. Metal-free divergent alkene functionalization.
Organics 04 00033 sch029
Scheme 30. Cu-mediated enantioselective cyanation of cyclopropanols.
Scheme 30. Cu-mediated enantioselective cyanation of cyclopropanols.
Organics 04 00033 sch030
Scheme 31. Bioinspired desaturation of alcohols via PCET and cobalt dual catalysis.
Scheme 31. Bioinspired desaturation of alcohols via PCET and cobalt dual catalysis.
Organics 04 00033 sch031
Scheme 32. Photoredox promoted selective activation of pyridinium salts.
Scheme 32. Photoredox promoted selective activation of pyridinium salts.
Organics 04 00033 sch032
Scheme 33. Photocatalysis route to coumarins catalyzed by (−)-riboflavin.
Scheme 33. Photocatalysis route to coumarins catalyzed by (−)-riboflavin.
Organics 04 00033 sch033
Scheme 34. Biomimetic synthesis of hitorins A and B.
Scheme 34. Biomimetic synthesis of hitorins A and B.
Organics 04 00033 sch034
Scheme 35. Total synthesis of danshenspiroketallactones.
Scheme 35. Total synthesis of danshenspiroketallactones.
Organics 04 00033 sch035
Scheme 36. Stepwise synthesis of polyoxamic acid.
Scheme 36. Stepwise synthesis of polyoxamic acid.
Organics 04 00033 sch036
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ali, M.; Sewell, S.; Li, J.; Wang, T. Recent Advances in Application of Alkoxy Radical in Organic Synthesis. Organics 2023, 4, 459-489. https://doi.org/10.3390/org4040033

AMA Style

Ali M, Sewell S, Li J, Wang T. Recent Advances in Application of Alkoxy Radical in Organic Synthesis. Organics. 2023; 4(4):459-489. https://doi.org/10.3390/org4040033

Chicago/Turabian Style

Ali, Munsaf, Shi Sewell, Juncheng Li, and Ting Wang. 2023. "Recent Advances in Application of Alkoxy Radical in Organic Synthesis" Organics 4, no. 4: 459-489. https://doi.org/10.3390/org4040033

Article Metrics

Back to TopTop