Next Article in Journal
Photochemical Implications of Changes in the Spectral Properties of Chromophoric Dissolved Organic Matter: A Model Assessment for Surface Waters
Next Article in Special Issue
Recent Advances in the Synthesis of Di- and Trisubstituted Hydroxylamines
Previous Article in Journal
Advances in Semiconductor-Based Nanocomposite Photo(electro)catalysts for Nitrogen Reduction to Ammonia
Previous Article in Special Issue
Recent Advances in N-O Bond Cleavage of Oximes and Hydroxylamines to Construct N-Heterocycle
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Molecule Synthesis Involving C-C Bond Cleavage of Ketoxime Esters

1
Key Laboratory for Green Organic Synthesis and Application of Hunan Province, Key Laboratory of Environmentally Friendly Chemistry and Application of Ministry of Education, College of Chemistry, Xiangtan University, Xiangtan 411105, China
2
Hunan Provincial Key Laboratory for Synthetic Biology of Traditional Chinese Medicine, School of Pharmaceutical Sciences, Hunan University of Medicine, Huaihua 418000, China
3
School of Chemistry and Chemical Engineering, Henan Normal University, Xinxiang 453007, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(6), 2667; https://doi.org/10.3390/molecules28062667
Submission received: 18 February 2023 / Revised: 8 March 2023 / Accepted: 10 March 2023 / Published: 15 March 2023
(This article belongs to the Special Issue Advances on the Application of N-O Bond Compounds)

Abstract

:
The synthetic strategies of oxime derivatives participating in radical-type reactions have been rapidly developed in the last few decades. Among them, the N–O bond cleavage of oxime esters leading to formation of nitrogen-centered radicals triggers adjacent C–C bond cleavage to produce carbon-centered free radicals, which has been virtually used in organic synthesis in recent years. Herein, we summarized the radical reactions involving oxime N–O bond and C–C bond cleavage through this special reaction form, including those from acyl oxime ester derivatives and cyclic ketoxime ester derivatives. These contents were systematically classified according to different reaction types. In this review, the free radical reactions involving acyl oxime esters and cyclic ketoxime esters after 2021 were included, with emphasis on the substrate scope and reaction mechanism.

1. Introduction

The nitrogen-containing skeleton widely exists in a large number of natural products, active molecules and functional materials, and various commodities composed of them, such as food, drugs and various other materials that play crucial roles in human life [1,2,3,4,5,6]. The strategy involving nitrogen-centered radicals (NCRs), as a pragmatic strategy to construct a C–N bond rapidly, has always been one of the research hotspots of chemists in the last decades [7,8]. Although a large number of scientists have been committed to developing simpler and more effective paths for C–N bond construction, the development of a nitrogen-centered free radical strategy is still very limited, which is, as concluded by Zard [8], mainly hampered by “a dearth of convenient access to these species and a lack of awareness pertaining to their reactivity” [9].
Ketone-derivated oxime esters have been widely used in the field of organic synthesis [10,11,12,13]. Among them, one of the most important aspects is the application in the construction of nitrogen heterocyclic compounds [14,15]. In the past few decades, the implementation of transition metal catalysis with palladium, rhodium, ruthenium, cobalt, etc. for oxime N–O bond activation has been intensively reported, which exhibited excellent site selectivity and provided effective and convenient methods for the synthesis of aza- compounds [16]. For example, in 2022, the review of O-acyl ketoximes summarized by Huang’s group detailed the cases of the synthesis of nitrogen heterocyclic compounds under various transition-metal-based catalytic systems [16]. In recent years, free radical chemistry has been developed rapidly and has showed featured advantages, among which oxime esters were also widely used in free-radical-type reactions [17,18,19,20,21,22]. Typically, oxime ester derivatives generate imino radicals under the action of photocatalysts or transition metals, and then further convert into carbon-centered radicals (CCRs). The conversion from NCRs to CCRs mainly involves the following methods (Scheme 1a): (1) 1,3-hydrogen transfer process to give α-imidoyl carbon radicals (path I) [23,24]; (2) 1,5-hydrogen transfer process to generate a γ-imidoyl carbon radicals (path II) [25,26]; (3) intramolecular addition of an iminyl radical to a tethered alkene or alkyne in a 5-exo-trig or 5-exo-dig fashion (path III) [27,28,29]; and (4) intramolecular carbon–carbon bond cleavage to generate carbon radicals (path IV) [30,31,32].
Recently, there were many detailed reviews on these topics. In 2019, the Wang group disclosed a review on imino radical-involved reactions [33]. Subsequently, path II and III on photocatalytic cross-couplings via the cleavage of N–O bonds were described in the review by Fu and Zhu [34]. Recently, both the Zhou [31] and Samanta [32] groups presented detailed summaries of the mechanisms, as well as reaction forms, of the various reactions in which cyclic ketoximes are involved. In addition, numerous chemistry groups, including Zard [8], Chen [35], Xu [36], etc., have more or less summarized strategies for the above forms of nitrogen radical centers. However, the use of oxime esters as acylation reagents has only emerged in recent few years, and to our knowledge, there are no related reviews reported; therefore, we mainly focus on this topic to summarize related works. Since acyl oxime esters also belong to the path IV type, we have also added a summary of the work on cyclic ketoximes after 2021. The oxime ester compounds used in this reaction mode all have special structures, one with an o-diketone derivative and the other with a ring structure, which can be obtained by two-step conversion from the corresponding ketones. Meanwhile, they have a similar reaction mode (Scheme 1b) in that oximes undergo single electron transfer with a transition metal or photosensitizer to achieve N–O cleavage and afford imino radicals, which subsequently undergo β-C-C bond cleavage to generate acyl radicals or cyanoalkyl. The nitriles by-products are formed when using acyl oxime esters in reactions. The acyl radicals and cyanoalkyl radicals can be captured by various free radical receptors and participate in various forms of reactions [31,32,37,38,39]. In this review, various free radical reactions involving acyl esters and cyclic ketoxime esters are summarized, and they are classified according to different types of reactions, mainly including radical addition, cross-coupling and radical coupling.

2. Acyl Oxime Ester

2.1. Radical Addition

2.1.1. Acyl Addition to Alkenes

In 2019, the Wu and Tung group demonstrated a groundbreaking example of acylation/cyclization of Michael acceptors with acyl radicals derived from acyl oxime esters 1 with C–C bond cleavage. Hence, cyclobutyl ketone derivatives 4 and acylated oxindoles 6 were obtained using fac-Ir(ppy)3 as a photosensitizer at room temperature (Scheme 2) [40]. The mechanism studies show that the acyl oxime esters 1 are first reduced by a IrIII catalyst under 3 W blue-light conditions, and then the acyl radicals 7 are produced by C–C bond cleavage. On the one hand, the acyl radical 7 is captured by styrene 3 and oxidized by IrIV to remove a proton to produce enone 8, which is sensitized by IrIII* to further react with another styrene 3 to lead to cross [2 + 2] cyclization and obtain acylated cyclobutanones derivatives 4. Alternatively, acyl radicals 7 could also add to the C–C double bond of the acrylamides 5 to afford structurally useful acylation oxindole skeletons 6.
In the same year, this group reported a novel three-component difunctionalization reaction of acyl oxime esters, olefins and alkane nitriles, which generated a series of β-carbonyl imides 12 with excellent yields (Scheme 3) [41]. The isotope tracking experiment and control experiments revealed the efficient intermolecular reorganization of oxime esters into styrene with the aid of solvent exchange. A possible reaction mechanism is shown in Scheme 3. First, the acyloxime esters 1 are cleaved to nitrile, carboxylate anions 13 and acyl radicals 7 by a single electron transfer (SET) process with the excited state IrIII*, which is generated from IrIII under blue-light irradiation. Then, a SET occurs between intermediate 14, which is produced by the addition of an acyl radical 7 to the olefin 3 and IrIV to complete the cycle of catalyst Ir and produce the carbocation 15. The intermediate 16, produced by the nucleophilic attack of nitrile to cation 15, undergoes a Mumm rearrangement with carboxylate anions 13 to obtain the amide products 12.
A photocatalytic hydroacylation of alkenes with acyl oximes for the synthesis of valuable ketones was developed by Yang et al. (Scheme 4) [42]. When CF3-attached styrenes were used, a E1cB-type fluoride elimination pathway was proposed for the obtainment of the 1,1-difluoroolefin products 19. In this difunctionalization reaction, the triphenylphosphine radical cation generated by the reaction of the excited catalyst IrIII* with triphenyl phosphine combines with the nucleophile acyl oximes 1 to obtain the phosphorus radical intermediate 20. The N–O bond cleavage of the intermediate 20 gives the iminyl radical 21 via β-scission. The imino radical 21 undergoes C–C bond scission to give acyl radical 7 and a molecule of acetonitrile. The acyl radical is added to alkene 17 and reduced by IrII to get the carbanion intermediate 23, with the following protonation to generate the desired product 18.
In 2021, Sun and co-workers developed a distinctive carbonylation approach using Ir and DIPEA to access γ-keto acids 24 and cyanocarboxylic acids 25 (Scheme 5) [43]. Mechanistically, this reaction involves the photocatalytic radical addition of acyl radicals and cyanoalkyl radicals to aromatic olefins and then carbon dioxide capture. The valuable dicarbofunctionalization tolerates a wide range of cyclic ketoxime substrates; however, only aliphatic acyl oxime esters proceed smoothly to afford corresponding γ-keto ester products.
Larionov’s group described a N-heterocyclic, carbene-photocatalyzed, three-component regioselective 1,2-diacylation of alkenes, using acyl oxime esters and aldehydes as two different acylating agents (Scheme 6) [44]. In particular, the authors demonstrated the mechanism by density functional theory (DFT) and time-dependent-DFT (TD-DFT), where an EDA intermediate is probably formed in the diacylation, which initiates photoexcitation to mediate charge transfer. The EDA intermediate 33, formed by the complexation of the Breslow intermediate 32 with the acyl oxime ester 1, is excited under light conditions and subsequently cleaved to remove BzO- and acetonitrile to give the intermediate 34 and acyl radical 7. The intermediate 14 produced by the addition of the acyl radical 7 to the olefin 3 couples with the intermediate 34 and then decomposes to a diketone product 30 and carbene 31.
Liu and Huang’s group reported an intriguing route to 3-acyl-substituted chroman-4-one derivatives 36, which involves SET reduction of acyl oxime esters 1 by fac-Ir(ppy)3 under thermal and light conditions (Scheme 7) [45]. In this functionalization reaction, the acyl radical generated from the acyl oxime ester first adds to the carbon double bond of the 2-allyloxy benzaldehydes 35, followed by intramolecular cyclization and 1,2-HAT (hydrogen atom transfer) to obtain the alcohol radical intermediate 40. Finally, SET oxidation by IrIV and deprotonation occur to obtain the target products 36.
Another acylation/cyclization strategy mediated by a photocatalytic nitrogen-centered radical of acyl oxime esters 1 with activated acrylamides 42 was reported by Liu and Huang’s group (Scheme 8) [46]. The authors screened a variety of acyl oximes with different leaving groups, such as 4-CF3C6H4CO, 4-NO2C6H4CO, C6F5CO, CF3CO, etc., and achieved up to 95% yields with 58 synthetic examples of acylated transformations. The fragmentation of one C–C bond and one N–O bond and the formulation of two new C–C bonds were carried out in one-pot conditions with a low catalyst loading (1 mol%). This protocol represents a simple and green road for the synthesis of acylated indolo/benzimidazo-[2,1,a]isoquinolin-6(5H) ones.
Most recently, the same group uncovered a new acylation reaction of acyl oxime esters with activated N-sulfonyl acrylamides 48, with a broad substrate scope and good functional group compatibility (Scheme 9) [47]. This method provides an effective nitrogen center-mediated approach for the generation of acyl radical intermediates to access acylated oxindole 6. This acylation transformation proceeds through a normal Smiles rearrangement process that involves cascade intramolecular ispo-cyclization (50), de-SO2 (51) and re-cyclization (52).

2.1.2. Acyl Addition to Alkynes

The heterocyclic skeleton presents in commercial drugs and many naturally occurring compounds; thus, the development of convenient and environmentally friendly methods for the construction of this structural motif is an important and promising research field [48,49,50]. Alkynes with specific substituents, such as alkyne amides [51,52,53], N-propylindoles [54,55,56], alkyne esters [57,58], alkyne amines [59,60,61], etc., are often used as radical acceptors to construct potential heterocyclic compounds by tandem cyclization. Recently, Liu and Zhou have independently developed a multitude of innovative photocatalytic, nitrogen-centered, radical-mediated acylation strategies of reactive alkynes using acyl oxime esters as acyl sources (Scheme 10) [37,39,62,63].
Liu’s group pioneered a radical addition strategy of acyl oxime esters 1 with activated alkynes 54 to construct a series of 3-acylated spiro [4,5]trienones derivatives 55 [37]. The acyl radical generated by the C–C bond breakage of acyl oxime esters attacks the carbon–carbon triple bond of propiolamides 54 to give the radical intermediate 62. The intermediate 62 undergoes intramolecular ispo-cyclization and is oxidized to carbocations 64 by IrIV, followed by the combination with hydroxide anion produced in water to obtain the intermediate 65. Finally, a methoxy anion and hydrogen ion are removed to obtain the expected trienone product 55.
Subsequently, Liu’s group developed two pragmatic acylation strategies using N-propynylindoles 56 [63] and alkynoates 58 [62] under photoexcitation conditions. A large quantity of acylated pyrrolo[1,2-a]indole 57 and coumarin 59 derivatives were constructed, and interestingly, both the alkanoyl and aryl groups were well adapted. These studies further developed the application of acyl oxime esters and were of great importance for the development of free radical synthetic chemistry.
Comparably, Zhou’s group has achieved a series of constructions of 3-acyl quinoline skeleton 61 under photocatalytic conditions with N-propargyl aromatic amines 60 as radical acceptors [39]. The mechanism of this acylation reaction states that acyl radicals add to the carbon–carbon triple bond of N-propargyl aromatic amines 60 for intramolecular cyclization and then proceed to dehydroarylation to give the final quinoline product 61.

2.2. Radical Cross-Couple

Recently, the development of the photoredox/palladium-catalyzed C-H acylation of 2-arylpyridines 78 with acyl oxime esters 1 using fac-Ir(ppy)3 as a photosensitizer was reported by the Chen group (Scheme 11) [64]. In order to thoroughly study the reaction mechanism, the author carried out some control experiments, including the radical clock reaction using benzocycloketoxime ester 80 with ethene-1,1-diyldibenzene under standard conditions and TEMPO as a radical inhibitor to test the activity of the reaction. The generation of the target product 83 and the detection of the free radical capture product 84 by the HRMS indicate that this acylation reaction contains a free radical mechanism.
In addition to the above acyl oxime esters, Wu’s group reported the benzyl oxime esters 85 with the same reaction mode, i.e., a single electron transfer followed by carbon–carbon bond cleavage to produce the corresponding radical (Scheme 12) [38]. The resulting benzyl radicals 91 are further oxidized to benzyl carbocation 92, which are then coupled with O and N-nucleophilic reagents to access the target benzyl ethers 88 and benzylamines 89. This benzylation strategy well tolerates various functional groups under mild conditions, where a wide range of nucleophilic substrates, such as primary and secondary alcohols, amines and even H2O, worked smoothly.

3. Cyclic Ketoxime Esters

3.1. Radical Addition

3.1.1. Ring Opening and Addition to Alkenes

In 2021, Chen and co-workers reported a difunctionalization synthesis of cyanoalkylfluorinated products that exploits a 10-phenyl-10H-phenothiazine (Ph-PTZ) as a photosensitizer approach (Scheme 13) [65]. The three-component cyanoalkylfluorination is initiated by the excitation of Ph-PTZ by 40 W purple LEDs. The cyanoalkyl radical was obtained by the C–C bond cleavage of the imino radical 94 generated by the single electron transfer between the excited-state Ph-PTZ* and the cycloketoxime ester. Oxidation of the intermediate 96 by Ph-PTZ•+ delivers the carbocation intermediate 97 and Ph-PTZ. The resulting intermediate is then attacked by nucleophilic substances, such as F and alcohols, to produce corresponding products (93 and 98). This strategy provides a transition-metal-free nitrogen-centered radical pathway for the construction of fluorocyanoalkylation derivatives.
In the same year, Du [66] and Guo [67] independently reported N-heterocyclic, carbene-catalyzed, three-component acylation/cyanoalkylation of alkenes with cyclic ketoxime esters and aldehydes (Scheme 14). The Du group suggested that this radical relay process relied on the SET achieved by the EDA complex 105 bound by Mg, the Breslow intermediate and oxime esters. The EDA complex 105 was subjected to thermally controlled SET and cleaved off the ester group to give the cyanoalkyl radical 95 and the intermediate 103. However, Guo’s group believed that it was the acyl oxime ester that directly undergoes SET with the intermediate bound by the aldehyde and N-heterocyclic carbene to remove the ester group to form the butyl cyano radical and intermediate 103. Subsequently, the addition of cyanoalkyl radicals 95 to olefins 3, followed by the radical-coupling with the intermediate 103, occurred to obtain the intermediate 104, which was further transformed to the final product 99 with the regeneration of the NHC for the next catalytic cycle.
In addition, Du’s group developed N-heterocyclic carbene-catalyzed 1,4-alkylcarbonylation of 1,3-enynes 106 with aldehydes 29 and different alkyl radical precursors, such as CF3I, alkyl halides, cycloketone oxime esters and redox-active esters derived from aliphatic carboxylic acids, affording a series of tetra-substituted allenyl ketones with excellent yields [68]. Among them, a variety of cyanoalkylated and acylated bianenone derivatives 107 were obtained in yields of 44–94% when cyclic ketoxime esters 2 were involved in this N-heterocyclic carbene-catalyzed system (Scheme 15).
Guo and co-workers reported 1,4-cyanoalkylarylations of 1,3-enynes 106 and aryl boronic acids 108 with a range of cyclobutanone oxime esters 2 to furnish multiple functionalizatized allenes 109. In the presence of [Ir(ppy)2(bpy)]PF6 and Cu(CH3CN)4PF6, the reaction completed within 7 minutes to reach 80% yield (Scheme 16) [69]. Cyclic ketoxime esters 2 were smoothly disintegrated and ring-opened to cyanoalkyl radicals with the help of the IrIII* catalyst, followed by the addition to the olefin of 1,3-enynes to obtain the radical intermediate 110, which underwent a radical shift to give the allene radical species 111. The oxidative addition of Ar-Cu(II) in situ formed by the transmetallation of the Cu(II) species with aryl boronic acids 108 to the diolefin radical 111 gives the intermediate 113, which undergoes reductive elimination to provide the target tetra-substituted allene 109 and the Cu(I) species.
Jiang et al. have established the 100-percent atom utilization, radical, tandem difunctionalization of 1,6-enynes 114 with cycloketone oxime esters 2 for the synthesis of 1-indanones bearing an all-carbon quaternary center by using CuBr as the catalyst, 1,10-phen as the ligand and Na2CO3 as the base (Scheme 17) [70]. The alkenyl radical intermediate 118 obtained by cyclization after attacking the alkyne 114 by the cyanoalkyl radical 95 reacted with the carboxylate anions 116 and Cu(II) species, producing all-atom utilization target products 115.
In 2021, a splendent method toward cyanoalkylsulfonylated oxindoles 120 and cyanoalkyl amides 121 was developed by Liang and co-workers (Scheme 18) [71]. Interestingly, the properties of substituent groups on the nitrogen atom in activated alkenes 48 resulted in different products. In particular, when the R4 = Me, oxindoles 120 were accessed. More surprisingly, the cyanoalkyl radical in the cyclization process could capture SO2 from the N-sulfonylated acrylamides 48, thus realizing cyanoalkylation/sulfonylation without an additional sulfur source. On the other hand, when N-arylacrylamides (R4 = aryl) were subjected to the system, hydrogenated cyanoalkyl amides were afforded.
A cascade cyanoalkylation/sulfonylation/cyclization of N-arylacrylamides 5 was reported by Gui’s group, which accessed a range of cyanoalkyl sulfonylated oxindole products 120 in good yields (Scheme 19) [72]. Compared to other strategies, this ring-opening sulfonylation transformation avoids the usage of catalysts and oxidants and requires only a 390 nm purple light source to proceed smoothly. The light irradiation was proposed to enable the N–O bond cleavage of ketoximes in the initial step.
Nitrogen-containing heterocycles, especially phenanthridine and quinoxaline derivatives, have been applied extensively in drugs, materials and organic synthesis [73,74]. In 2021, Zhou et al. developed nickel-catalyzed, difunctionalization, oxidative annulation reactions of vinyl azides 128 and cycloketone oxime esters 2 to prepare cyanoalkyl containing quinoxalin-2(1H)-ones 129 (Scheme 20) [73]. One of the highlights of the strategy is that five- and six-membered cyclic ketoxime esters could also react to obtain ring-opening products in moderate yields. In the same year, a similar transition metal-catalyzed system for cyanoalkylation and cyclization of activated alkenes and cyclic ketoxime esters 2 was developed by Xu, Kong and co-workers (Scheme 20) [74]. Compared to Zhou’s work, this reaction replaces the nickel catalyst with a metallic copper catalyst and extends it to 2-(1-azidovinyl)-1,1′-biphenyl derivatives 131. Further, spirocyclic ketone products 132 were obtained when 2-azido-N-(4-methoxyphenyl)acrylamides were used as the substrate. While the origin of the ketone oxygen atom has not been further verified by the authors, it usually comes from the water in the solvent in such a reaction.
Guo, Qiu and co-workers reported a convenient room-temperature synthesis of cyanoalkyl-containing β-enamino ketones from heterocyclic-substituted azidyl homoallylic alcohol precursors under visible-light irradiation (Scheme 21) [75]. The remarkable feature is the short reaction time within 9.3 min; yet, the yields of the substrates are generally low, especially for the cyclic ketoxime substrates, which generally have only 20–50% yields. Mechanistically, the addition of the cyanoalkyl radicals 95 to azide olefins 135, with removal of one molecule of nitrogen, generates the imino radicals 137, which undergo a 5-exo intramolecular radical process to give the key short-lived spiro radical 138. Subsequently, the spirocyclic radical undergoes C–C bond breaking to obtain the alcohol radical intermediate 139, which is oxidized by the IrIV to lose a proton to obtain the final product 136.
The unpaired catalytic synthesis of chiral ligands with transition metals is a hot topic, but also a difficult area in the field of organic synthesis. Chen’s group has developed a purple-light-irradiated enantioselective difunctionalization of 1,3-butadienes using cyclic ketoxime ester as a bifunctional reagent under the action of copper catalysis and ligand (R, R)-L1 to generate diversely substituted allylic esters with the ee up to 95% (Scheme 22) [76]. As proposed by the authors, the copper catalyst is combined with the ligand (R, R)-L1 to obtain the [L1Cu(I)X] species, which is subsequently excited by purple light into the excited state [L1Cu(I)X]*. The excited state [L1Cu(I)X]* undergoes electron transfer with the cyclic ketoxime ester 2 to generate an imino radical 94, carboxylate anions and [L1Cu(II)X] species. At the same time, the carboxylate anions combine with [L1Cu(II)X] to obtain a copper–nucleophile complex 143. The imino radical generates the cyanoalkyl radical 95 and then attacks the terminal double bond of 1,3-butadiene 141 to obtain the radical intermediate 144, which is cross-coupled with the above copper–nucleophile complex 143 to provide the intermediate 145. Finally, the enantioselective generation of sp3 C–O bonds provides the expected products 142 and [L1Cu(I)X] species, which then participate in the next catalytic cycle.
The strategy for visible light-mediated synthesis of the 3-cyanoalkyl coumarin skeleton 147 without an external photocatalyst was developed by Yang and co-workers (Scheme 23) [77]. The ortho-hydroxycinnamic esters 146 in the strategy served as both the reactant and photosensitizer. This metal-free cyanoalkylation of cycloketone oxime esters 2 with 2-hydroxyalkenyl esters 146 features a broad substrate scope and good functional group compatibility; especially, the 5-membered oxygen-heterocyclic ketoxime ester could also deliver the target product with a 31% yield.
In 2021, He’s group reported a straightforward approach to synthesize β-lactam derivatives 155 by using cycloketone oxime esters 2 as the cyanoalkyl radical source via Cu-mediated cyclization of inactivated alkenes containing aminoquinoline 154 (Scheme 24) [78]. In particular, the copper catalyst in the cyclization reaction needs the assistance of the aminoquinoline ligand. Mechanistically, as described in Scheme 24, the Cu(II) catalyst chelated with the substrate 154 to obtain the chelated Cu(II) alkene complex 156, which will be nucleophilically added by the cyanoalkyl radical 95 to obtain the five-membered Cu(III)-metallacycle 157. Finally, the β-lactam product and Cu(I) are generated through intramolecular reductive elimination. In addition, the authors further transformed the obtained β-lactam derivatives 155, such as by removing the substituent group on the N atom to obtain β-lactam 158 and opening the quaternary ring to obtain a long-chain amino ester compound 159.

3.1.2. Ring Opening and Addition to Alkynes

The process involving the ring-opening of cyclic ketoxime esters to afford cyanoalkyl radicals and then capturing sulfur dioxide is considered as an effective sulfonylation strategy. Compared with the traditional sulfonylation strategy, it has the advantage of mild and environmentally benign conditions [79,80]. In 2021, Liu’s group demonstrated a cyanoalkyl-radical-triggered cyanoalkylsulfonylation and cyclization transformation of phenyl alkynoates 58 under visible-light conditions to access 3-cyanoalkylsulfonyl-substituted coumarins 160 (Scheme 25) [81]. In this strategy, cyanoalkyl radicals capture sulfur dioxide produced by potassium sulfite to the cyanoalkylsulfonyl radical 125, and then attack the C–C triple bond of the alkynoates 58 to obtain the intermediate 162. Generally, there are two different reaction sequences after the spirane structure 163 is obtained by the ispo-cyclization of the intermediate 164. One is that the helix intermediates will be oxidized directly into carbocations before 1,2-ester migration; the other is that the helix intermediates will be oxidized into carbocation after 1,2-ester migration in the form of free radicals. Both approaches are feasible. It is worth mentioning that in the following year, Liao and co-workers developed a SO2-free strategy with the same raw material, and the reaction mechanism was believed to be the oxidation after 1,2-ester migration [82].
A visible-light-induced, multi-component cascade cyanoalkylsulfonylation of N-propargyl aromatic amines 60 with cycloketone oxime esters 2 was discovered by Zhou’s group to synthesize cyanoalkylsulfonylated quinolines 166 (Scheme 26) [83]. This methodology provides a facile and mild route for the synthesis of a large number of cyanoalkylsulfonylated quinoline derivatives, with excellent isolated yields ranging from 53% to 99%.
In 2021, Wu et al. established the four-component, free radical, tandem selenosulfonylation of alkynes with cyclobutanone oxime esters 2, DABSO and diphenyl diselenides 172 for the synthesis of β-cyanoalkylsulfonylated vinyl selenides 173 under mild reaction conditions (Scheme 27) [84]. The radical capture experiments demonstrated that the reaction involved a radical course, where cyanoalkylsulfonyl radicals were probably produced. The multicomponent tandem reaction sequentially undergoes the generation of imino radicals 94, C–C bond breaking to produce cyanoalkyl radicals 95, sulfur dioxide capture, addition of cyanoalkylsulfonyl radicals 125 to alkynes 171 and coupling of aryl selenyl radicals 176 with alkenyl radicals 175.
In the following year, Zhou’s group developed a novel radical tandem cyanoalkylsulfonylation reaction of 3-arylethynyl- [1,10-biphenyl]-2-carbonitriles 177 by using DABSO as the sulfur dioxide source for the synthesis of cyanoalkylsulfonyl-substituted cyclopenta [gh]phenanthridines 178. (Scheme 28) [85]. Mechanistic studies showed that the alkenyl radical intermediate 179, obtained by the addition of a cyanoalkylsulfonyl radical 125 to the carbon–carbon triple bond of substrates 177, undergoes intramolecular cyanogenic cyclization to give the imino radical species 180, which then adds to the benzene ring to get the conjugated radical 181. Finally, the radical 181 could undergo two different pathways to provide the final products, i.e., base-promoted homogeneous aromatic substitution and organometallic-based β-hydride elimination.
Isocyanic compounds are an excellent class of radical acceptors. Some pre-functionalized isocyanic substrates are often used to construct various polycyclic nitrogen-containing heterocyclic compounds. Shao’s group achieved facile construction of a series of 1-cyanoalkyl isoquinolines 183 and 6-cyanoalkyl phenanthridines 185 by cyanoalkylation under iron catalysis using various isocyanines (184 and 186) as substrates and cyclic ketoxime esters 2 as radical precursors (Scheme 29) [86]. As shown in the proposed mechanism, the radical intermediate 188 has two pathways to obtain the final product: oxidation by Fe3+ species followed by deprotonation by 1,8-Diazabicyclo [5.4.0]undec-7-ene (DBU) or otherwise direct oxidation by oxime ester 2 after deprotonation.

3.2. Radical Cross-Coupling

In 2021, Jiang and co-workers reported an asymmetric deconstructive alkynylation reaction that is facilitated by copper/chiral cinchona alkaloid-based N,N,P-ligand catalysts L1 (Scheme 30) [87]. A variety of chiral molecules with cyanoalkyl groups 191 were successfully obtained under mild conditions, including δ-alkynyl amides, ε-alkenyl carbamate and ζ-keto nitrile, with ee up to 99%. The strategy has good functional group tolerance and substrate adaptability, and a range of cyclic oximes, including five- and six-membered reactants, coupled smoothly in moderate to excellent yields. However, this protocol only applies to aryl-substituted alkynes; other alkynes, including ether-substituted and alkyl alkynes, did not react. In addition, Wu’s group also realized a cyanoalkylation strategy with olefins instead of alkynes under copper catalysis [88]. A range of diverse (E)-cyanoalkylsulfonyl alkenes 192 were accessed with excellent yields by using copper iodide as the catalyst. According to the available literature and control experiments, the cyanoalkyl radical and cyanoalkylsulfonyl radical were believed to be the key intermediates. Also, the protocol has some limitations in that cyclic olefins, alkyl-substituted olefins, 1,2-disubstituted olefins and 1,3-butadiene olefins failed to give the target products, and five- and six-membered cyclic ketoxime esters did not work in the conversion.
In 2022, a radical tandem functionalization of 1,4-quinones 198 through cyanoalkylation was revealed by Akondi and Mainkar’s group to afford a series of cyanoalkylated quinone derivatives 199 at room temperature in the absence of a transition metal catalyst (Scheme 31) [89]. A range of substituted 1,4-quinones 198 smoothly served as the coupling partners with cycloketone oxime esters 2, in the presence of 2 mol% Rose Bengal, affording cyanoalkylated quinone derivatives 199 in moderate yields.
The electrochemical technique has long been touted as an environmentally beneficial method due to its inherent ability to achieve redox reactions in a sustainable manner [90,91]. Very recently, Wang’s group developed a pioneering case of electrocatalytic, quinoxalin-2(1H)-ones 201, C-H direct cyanoalkylation that uses cyclic ketoximes 2 as a cyanoalkyl source to successfully construct a series of 3-cyanoalkyl-substituted quinolones 202 (Scheme 32) [92]. Similar to thermal and photochemistry, cyclic ketoxime esters 2 first undergo a single electron transfer at the anode to remove the ester group, and then β–C–C bond cleavage occurs to give the cyanoalkyl radical, which then attacks the C–N double bond of the quinoxalin-2(1H)-ones to produce the nitrogen radical intermediate 203. Finally, the intermediate 203 undergoes a single electron oxidation at the cathode, followed by deprotonation, to produce the final cyanoalkylated product.
An interesting iron-catalyzed selective functionalization process for the direct C(sp3)-H cyanoalkylation of glycine derivatives promoted by pyridine-oxazoline ligands was developed by Gong and co-workers (Scheme 33) [93]. A range of substituted glycine derivatives 204 served as cross-coupling partners with cyclobutanone oxime esters 2, in the presence of 1 mol% Fe(NTf2)2 and 10 mol% pyridine-oxazoline ligands (L3 or L4), to access cyanoalkylated amino acids and peptides 205 in moderate to high yields. Importantly, this is the first example of C(sp3)-H cyanoalkylation with cyclic ketoximes, providing a new pathway for the synthesis of unnatural amino acids. The authors demonstrated experimentally that the interaction between the iron-catalyst and glycinate substrates could greatly improve the catalytic efficiency, and the in situ-formed imine intermediates are a key for high chemo-selectivity.
Guo and co-workers reported an alkoxy, radical-triggered, ring-opening halogenation of cyclopentyl hydroperoxides under a simple and effective copper catalytic system involving redox-neutral conditions and green halogen sources [94]. In addition, the authors applied this halogenation reaction system to the ring-opening halogenation of cyclic ketoximes 2, and the structurally diverse halogenated cyanides 209 were constructed in moderate yields (Scheme 34).
Recently, Qin and co-workers utilized transition metal-catalyzed denitrogenative cyanoalkylation of nitroalkenes by a reductive C–C bond formation process (Scheme 35) [95]. The mild method provides convenient catalytic access to cyanoalkylated styrene derivatives with good functional group tolerance. A variety of cyanoalkylated styrenes were synthesized in moderate to excellent yields.
In 2021, Lu and co-workers reported the decarboxylative cross-coupling of α,β-unsaturated carboxylic acids 212 with cycloketone oxime esters 2, exploiting a photo/nickel dual catalysis (Scheme 36) [96]. Different from the previous decarboxylation mechanism of α,β-unsaturated carboxylic acids, this reaction involves the nickelacyclopropane intermediate 215, which then undergoes decarboxylative elimination to access the final product. This method not only constructs a large number of cyanoalkyl olefins with good functional group tolerance, but also has an important role in the modification of complex molecules.

3.3. Radical Coupling

Elements of the group of sulfur, such as S, Se and Te, are often introduced into the skeletons of various organic compounds as a class of heteroatoms, because organic compounds containing such heteroatoms have a wide distribution and application in pharmaceuticals, drug candidates, agrochemicals, catalysis and functional materials [97]. In 2021, Zhao’s group developed two different chalcogenation reactions of cyclic ketoxime esters 2 (Scheme 37). The ring-opening radical coupling of cyclic ketoxime esters 2 with dichalcogenides 217 occurs under additive-free conditions [98]. The mild photocatalysis enabled its coupling with potassium selenocyanate 219 [99].

4. Conclusions

With the increasing requirements for environmental protection, great efforts have been made into advancing green technologies and sustainable chemistry [100,101]. The strategies involving free radical chemistry have been demonstrated as a powerful tool, with features regarding ease of operation, high efficiency, mild conditions and environmental friendliness [21,100,101,102]. Among them, nitrogen-centered radicals are highly versatile reactive intermediates, thus widely exploited for synthetic utility in C–N bond-forming reactions [35]. Readily accessible oxime ester derivatives have gained fame due to their specific N-O structures and their tendency to generate nitrogen-centered radicals with the aid of transition metals or/and photocatalysis.
This paper summarizes related radical reactions of oxime esters through imino radicals, which experience carbon–carbon bond rupture for conversion to carbon-centered radicals (CCRs). Acyl oxime esters and cyclic ketoxime esters are representative substrates of these reactions, which formally include radical addition reactions, cross-couplings, radical couplings, etc. Although the research on oxime esters has been developed rapidly and significant results have been achieved, there are still demands to develop new models of reactivities and expand their applications in related fields. Specifically, the ester group fraction of ketoximes was generally released to form acid by-products, which leads to significant issues of atom economy. Moreover, only a few examples of asymmetric synthesis were disclosed, mostly limited to transition metal catalysis or photocatalysis. In the context of sustainable catalysis, we look forward to further progress in the application of oxime esters with novel reactions and as a multipurpose development in academic and industrial settings.

Author Contributions

Conceptualization, P.C. and H.H.; investigation, Q.T.; writing—original draft preparation, P.C. and H.H.; writing—review and editing, X.J. and F.Z.; supervision, F.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (21801076, 22071211) and the Open Research Fund of the School of Chemistry and Chemical Engineering, Henan Normal University (2022C02).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Homer, J.A.; Sperry, J. Mushroom-Derived Indole Alkaloids. J. Nat. Prod. 2017, 80, 2178–2187. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, J.; Zhang, C.; Zheng, Z.; Zhou, P.; Liu, W. Research Progress of Sulfoxonium Ylides in the Construction of Five/Six-Membered Nitrogen-Containing Heterocycles. Chin. J. Org. Chem. 2022, 42, 2745–2759. [Google Scholar] [CrossRef]
  3. Showalter, H.D.H. Progress in the Synthesis of Canthine Alkaloids and Ring-Truncated Congeners. J. Nat. Prod. 2013, 76, 455–467. [Google Scholar] [CrossRef] [PubMed]
  4. Stepek, I.A.; Itami, K. Recent Advances in C–H Activation for the Synthesis of Π-Extended Materials. ACS Mater. Lett. 2020, 2, 951–974. [Google Scholar] [CrossRef]
  5. Qu, Z.; Wang, P.; Chen, X.; Deng, G.-J.; Huang, H. Visible-Light-Driven Cadogan Reaction. Chin. Chem. Lett. 2021, 32, 2582–2586. [Google Scholar] [CrossRef]
  6. Chen, X.; Zhou, X.; Ji, X.; Huang, H. Visible Light-Induced Aerobic Oxidative Dehydrogenative Coupling of Thiophenols. Chin. J. Org. Chem. 2021, 41, 4704–4711. [Google Scholar] [CrossRef]
  7. Yang, S.; Song, J.; Dong, D.; Yang, H.; Zhou, M.; Zhang, H.; Wang, Z. Progress of N-Amino Pyridinium Salts as Nitrogen Radical Precursors in Visible Light Induced C-N Bond Formation Reactions. Chin. J. Org. Chem. 2022, 42, 4099–4110. [Google Scholar] [CrossRef]
  8. Zard, S.Z. Recent Progress in the Generation and Use of Nitrogen-Centred Radicals. Chem. Soc. Rev. 2008, 37, 1603–1618. [Google Scholar] [CrossRef]
  9. Xiong, T.; Zhang, Q. New Amination Strategies Based on Nitrogen-Centered Radical Chemistry. Chem. Soc. Rev. 2016, 45, 3069–3087. [Google Scholar] [CrossRef]
  10. Bolotin, D.S.; Bokach, N.A.; Demakova, M.Y.; Kukushkin, V.Y. Metal-Involving Synthesis and Reactions of Oximes. Chem. Rev. 2017, 117, 13039–13122. [Google Scholar] [CrossRef]
  11. Kitamura, M.; Narasaka, K. Synthesis of Aza-Heterocycles from Oximes by Amino-Heck Reaction. Chem. Rec. 2002, 2, 268–277. [Google Scholar] [CrossRef]
  12. Li, W.; Towne, T.B.; Liang, G.; Haight, A.R.; Barnes, D.M. Practical Large-Scale Synthesis of 6-Bromo-2-Naphthylmethanesulfonamide Using Semmler–Wolff Reaction. Org. Process Res. Dev. 2014, 18, 1696–1701. [Google Scholar] [CrossRef]
  13. Tabolin, A.A.; Ioffe, S.L. Rearrangement of N-Oxyenamines and Related Reactions. Chem. Rev. 2014, 114, 5426–5476. [Google Scholar] [CrossRef]
  14. Huang, H.; Cai, J.; Deng, G.-J. O-Acyl Oximes: Versatile Building Blocks for N-Heterocycle Formation in Recent Transition Metal Catalysis. Org. Biomol. Chem. 2016, 14, 1519–1530. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, C.; Zhao, J.; Shi, X.; Liu, L.; Zhu, Y.-P.; Sun, W.; Zhu, B. Recent Advances in Cyclization Reactions of Unsaturated Oxime Esters (Ethers): Synthesis of Versatile Functionalized Nitrogen-Containing Scaffolds. Org. Chem. Front. 2020, 7, 1948–1969. [Google Scholar] [CrossRef]
  16. Qu, Z.; Tian, T.; Deng, G.-J.; Huang, H. Diverse Catalytic Systems for Nitrogen-Heterocycle Formation from O-Acyl Ketoximes. Chin. Chem. Lett. 2023, 34, 107565. [Google Scholar] [CrossRef]
  17. Yi, H.; Zhang, G.; Wang, H.; Huang, Z.; Wang, J.; Singh, A.K.; Lei, A. Recent Advances in Radical C–H Activation/Radical Cross-Coupling. Chem. Rev. 2017, 117, 9016–9085. [Google Scholar] [CrossRef] [PubMed]
  18. Yu, X.-Y.; Chen, J.-R.; Wang, P.-Z.; Yang, M.-N.; Liang, D.; Xiao, W.-J. A Visible-Light-Driven Iminyl Radical-Mediated C−C Single Bond Cleavage/Radical Addition Cascade of Oxime Esters. Angew. Chem. Int. Ed. 2018, 57, 738–743. [Google Scholar] [CrossRef]
  19. Wang, Z.; Liu, Q.; Ji, X.; Deng, G.-J.; Huang, H. Bromide-Promoted Visible-Light-Induced Reductive Minisci Reaction with Aldehydes. ACS Catal. 2020, 10, 154–159. [Google Scholar] [CrossRef]
  20. Bhunia, A.; Studer, A. Recent Advances in Radical Chemistry Proceeding through Pro-Aromatic Radicals. Chem. 2021, 7, 2060–2100. [Google Scholar] [CrossRef]
  21. Wu, X.; Ma, Z.; Feng, T.; Zhu, C. Radical-Mediated Rearrangements: Past, Present, and Future. Chem. Soc. Rev. 2021, 50, 11577–11613. [Google Scholar] [CrossRef] [PubMed]
  22. Qu, Z.; Tian, T.; Tan, Y.; Ji, X.; Deng, G.-J.; Huang, H. Redox-Neutral Ketyl Radical Coupling/Cyclization of Carbonyls with N-Aryl Acrylamides through Consecutive Photoinduced Electron Transfer. Green Chem. 2022, 24, 7403–7409. [Google Scholar] [CrossRef]
  23. Ran, L.; Ren, Z.-H.; Wang, Y.-Y.; Guan, Z.-H. Copper-Catalyzed Homocoupling of Ketoxime Carboxylates for Synthesis of Symmetrical Pyrroles. Green Chem. 2014, 16, 112–115. [Google Scholar] [CrossRef]
  24. Ke, J.; Tang, Y.; Yi, H.; Li, Y.; Cheng, Y.; Liu, C.; Lei, A. Copper-Catalyzed Radical/Radical C-H/P-H Cross-Coupling: A-Phosphorylation of Aryl Ketone O-Acetyloximes. Angew. Chem. Int. Ed. 2015, 54, 6604–6607. [Google Scholar] [CrossRef] [PubMed]
  25. Davies, J.; Booth, S.G.; Essafi, S.; Dryfe, R.A.W.; Leonori, D. Visible-Light-Mediated Generation of Nitrogen-Centered Radicals: Metal-Free Hydroimination and Iminohydroxylation Cyclization Reactions. Angew. Chem. Int. Ed. 2015, 54, 14017–14021. [Google Scholar] [CrossRef] [Green Version]
  26. Jiang, H.; Studer, A. Iminyl-Radicals by Oxidation of α-Imino-Oxy Acids: Photoredox-Neutral Alkene Carboimination for the Synthesis of Pyrrolines. Angew. Chem. Int. Ed. 2017, 56, 12273–12276. [Google Scholar] [CrossRef]
  27. Li, J.; Zhang, P.; Jiang, M.; Yang, H.; Zhao, Y.; Fu, H. Visible Light as a Sole Requirement for Intramolecular C(sp3)–H Imination. Org. Lett. 2017, 19, 1994–1997. [Google Scholar] [CrossRef]
  28. Shu, W.; Nevado, C. Visible-Light-Mediated Remote Aliphatic C−H Functionalizations through a 1,5-Hydrogen Transfer Cascade. Angew. Chem. Int. Ed. 2017, 56, 1881–1884. [Google Scholar] [CrossRef]
  29. Gu, Y.-R.; Duan, X.-H.; Chen, L.; Ma, Z.-Y.; Gao, P.; Guo, L.-N. Iminyl Radical-Triggered Intermolecular Distal C(sp3)–H Heteroarylation Via 1,5-Hydrogen-Atom Transfer (Hat) Cascade. Org. Lett. 2019, 21, 917–920. [Google Scholar] [CrossRef]
  30. Wu, X.; Zhu, C. Recent Advances in Ring-Opening Functionalization of Cycloalkanols by C–C σ-Bond Cleavage. Chem. Rec. 2018, 18, 587–598. [Google Scholar] [CrossRef]
  31. Xiao, T.; Huang, H.; Anand, D.; Zhou, L. Iminyl-Radical-Triggered C–C Bond Cleavage of Cycloketone Oxime Derivatives: Generation of Distal Cyano-Substituted Alkyl Radicals and Their Functionalization. Synthesis 2020, 52, 1585–1601. [Google Scholar] [CrossRef]
  32. Guin, S.; Majee, D.; Samanta, S. Recent Advances in Visible-Light-Driven Photocatalyzed Γ-Cyanoalkylation Reactions. Asian J. Org. Chem. 2021, 10, 1595–1618. [Google Scholar] [CrossRef]
  33. Yin, W.; Wang, X. Recent Advances in Iminyl Radical-Mediated Catalytic Cyclizations and Ring-Opening Reactions. New J. Chem. 2019, 43, 3254–3264. [Google Scholar] [CrossRef]
  34. Zhu, X.; Fu, H. Photocatalytic Cross-Couplings Via the Cleavage of N–O Bonds. Chem. Commun. 2021, 57, 9656–9671. [Google Scholar] [CrossRef] [PubMed]
  35. 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] [PubMed]
  36. Xiong, P.; Xu, H.-C. Chemistry with Electrochemically Generated N-Centered Radicals. Acc. Chem. Res. 2019, 52, 3339–3350. [Google Scholar] [CrossRef] [PubMed]
  37. Chen, P.; Xie, J.; Chen, Z.; Xiong, B.-Q.; Liu, Y.; Yang, C.-A.; Tang, K.-W. Visible-Light-Mediated Nitrogen-Centered Radical Strategy: Preparation of 3-Acylated Spiro[4,5]Trienones. Adv. Synth. Catal. 2021, 363, 4440–4446. [Google Scholar] [CrossRef]
  38. Fan, X.; Lei, T.; Liu, Z.; Yang, X.-L.; Cheng, Y.-Y.; Liang, G.; Chen, B.; Tung, C.-H.; Wu, L.-Z. Benzyl C–O and C–N Bond Construction Via C–C Bond Dissociation of Oxime Ester under Visible Light Irradiation. Eur. J. Org. Chem. 2020, 2020, 1551–1558. [Google Scholar] [CrossRef] [Green Version]
  39. Zhang, M.; Wu, S.; Wang, L.; Xia, Z.; Kuang, K.; Xu, Q.; Zhao, F.; Zhou, N. Visible-Light-Induced Cascade Cyclization of N-Propargyl Aromatic Amines and Acyl Oxime Esters: Rapid Access to 3-Acylated Quinolines. J. Org. Chem. 2022, 87, 10277–10284. [Google Scholar] [CrossRef]
  40. Fan, X.; Lei, T.; Chen, B.; Tung, C.-H.; Wu, L.-Z. Photocatalytic C–C Bond Activation of Oxime Ester for Acyl Radical Generation and Application. Org. Lett. 2019, 21, 4153–4158. [Google Scholar] [CrossRef]
  41. Cheng, Y.-Y.; Lei, T.; Su, L.; Fan, X.; Chen, B.; Tung, C.-H.; Wu, L.-Z. Visible Light Irradiation of Acyl Oxime Esters and Styrenes Efficiently Constructs β-Carbonyl Imides by a Scission and Four-Component Reassembly Process. Org. Lett. 2019, 21, 8789–8794. [Google Scholar] [CrossRef]
  42. Zheng, L.; Xia, P.-J.; Zhao, Q.-L.; Qian, Y.-E.; Jiang, W.-N.; Xiang, H.-Y.; Yang, H. Photocatalytic Hydroacylation of Alkenes by Directly Using Acyl Oximes. J. Org. Chem. 2020, 85, 11989–11996. [Google Scholar] [CrossRef]
  43. Bai, J.; Li, M.; Zhou, C.; Sha, Y.; Cheng, J.; Sun, J.; Sun, S. Visible-Light Photoredox-Catalyzed Dicarbofunctionalization of Styrenes with Oxime Esters and CO2: Multicomponent Reactions toward Cyanocarboxylic Acids and Γ-Keto Acids. Org. Lett. 2021, 23, 9654–9658. [Google Scholar] [CrossRef]
  44. Jin, S.; Sui, X.; Haug, G.C.; Nguyen, V.D.; Dang, H.T.; Arman, H.D.; Larionov, O.V. N-Heterocyclic Carbene-Photocatalyzed Tricomponent Regioselective 1,2-Diacylation of Alkenes Illuminates the Mechanistic Details of the Electron Donor–Acceptor Complex-Mediated Radical Relay Processes. ACS Catal. 2022, 12, 285–294. [Google Scholar] [CrossRef]
  45. Liu, Y.-C.; Chen, P.; Li, X.-J.; Xiong, B.-Q.; Liu, Y.; Tang, K.-W.; Huang, P.-F. Visible-Light-Induced Dual Acylation of Alkenes for the Construction of 3-Substituted Chroman-4-Ones. J. Org. Chem. 2022, 87, 4263–4272. [Google Scholar] [CrossRef]
  46. Guo, Y.; Huang, P.-F.; Liu, Y.; He, B.-H. Visible-Light-Induced Acylation/Cyclization of Active Alkenes: Facile Access to Acylated Isoquinolinones. Org. Biomol. Chem. 2022, 20, 3767–3778. [Google Scholar] [CrossRef]
  47. Luo, Z.-T.; Fan, J.-H.; Xiong, B.-Q.; Liu, Y.; Tang, K.-W.; Huang, P.-F. Visible-Light-Induced Acylation/Arylation of Alkenes Via Aryl Migration/Desulfonylation. Eur. J. Org. Chem. 2022, 2022, e202200793. [Google Scholar] [CrossRef]
  48. Michael, J.P. Quinoline, Quinazoline and Acridone Alkaloids. Nat. Prod. Rep. 2002, 19, 742–760. [Google Scholar] [CrossRef] [PubMed]
  49. Meng, S.; Tang, G.-L.; Pan, H.-X. Enzymatic Formation of Oxygen-Containing Heterocycles in Natural Product Biosynthesis. Chembiochem. 2018, 19, 2002–2022. [Google Scholar] [CrossRef] [PubMed]
  50. Pawlowski, R.; Stanek, F.; Stodulski, M. Recent Advances on Metal-Free, Visible-Light- Induced Catalysis for Assembling Nitrogen- and Oxygen-Based Heterocyclic Scaffolds. Molecules 2019, 24, 1533. [Google Scholar] [CrossRef] [Green Version]
  51. Liu, Y.; Wang, Q.-L.; Zhou, C.-S.; Xiong, B.-Q.; Zhang, P.-L.; Yang, C.-A.; Tang, K.-W. Visible-Light-Mediated Ipso-Carboacylation of Alkynes: Synthesis of 3-Acylspiro[4,5]Trienones from N-(p-Methoxyaryl)Propiolamides and Acyl Chlorides. J. Org. Chem. 2018, 83, 2210–2218. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, Y.; Wang, Q.-L.; Chen, Z.; Zhou, Q.; Xiong, B.-Q.; Zhang, P.-L.; Tang, K.-W. Visible-Light Promoted One-Pot Synthesis of Sulfonated Spiro[4,5]Trienones from Propiolamides, Anilines and Sulfur Dioxide under Transition Metal-Free Conditions. Chem. Commun. 2019, 55, 12212–12215. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, P.; Fan, J.-H.; Yu, W.-Q.; Xiong, B.-Q.; Liu, Y.; Tang, K.-W.; Xie, J. Alkylation/Ipso-Cyclization of Active Alkynes Leading to 3-Alkylated Aza- and Oxa-Spiro[4,5]-Trienones. J. Org. Chem. 2022, 87, 5643–5659. [Google Scholar] [CrossRef] [PubMed]
  54. Zhu, J.; Sun, S.; Xia, M.; Gu, N.; Cheng, J. Copper-Catalyzed Radical 1,2-Cyclization of Indoles with Arylsulfonyl Hydrazides: Access to 2-Thiolated 3h-Pyrrolo[1,2-a]Indoles. Org. Chem. Front. 2017, 4, 2153–2155. [Google Scholar] [CrossRef]
  55. Liu, Y.; Wang, Q.-L.; Chen, Z.; Chen, P.; Tang, K.-W.; Zhou, Q.; Xie, J. Visible-Light-Induced Cascade Sulfonylation/Cyclization of N-Propargylindoles with Aryldiazonium Tetrafluoroborates via the Insertion of Sulfur Dioxide. Org. Biomol. Chem. 2019, 17, 10020–10029. [Google Scholar] [CrossRef]
  56. Liu, Y.; Chen, Z.; Wang, Q.-L.; Chen, P.; Xie, J.; Xiong, B.-Q.; Zhang, P.-L.; Tang, K.-W. Visible Light-Catalyzed Cascade Radical Cyclization of N-Propargylindoles with Acyl Chlorides for the Synthesis of 2-Acyl-9h-Pyrrolo[1,2-a]Indoles. J. Org. Chem. 2020, 85, 2385–2394. [Google Scholar] [CrossRef]
  57. Liu, T.; Ding, Q.; Zong, Q.; Qiu, G. Radical 5-exo Cyclization of Alkynoates with 2-Oxoacetic Acids for Synthesis of 3-Acylcoumarins. Org. Chem. Front. 2015, 2, 670–673. [Google Scholar] [CrossRef]
  58. Mi, X.; Wang, C.; Huang, M.; Wu, Y.; Wu, Y. Preparation of 3-Acyl-4-Arylcoumarins via Metal-Free Tandem Oxidative Acylation/Cyclization between Alkynoates with Aldehydes. J. Org. Chem. 2015, 80, 148–155. [Google Scholar] [CrossRef]
  59. Zhang, P.; Zhang, L.; Gao, Y.; Tang, G.; Zhao, Y. Synthesis of 3-Phosphinoylquinolines via a Phosphinoylation–Cyclization–Aromatization Process Mediated by tert-Butyl Hydroperoxide. RSC Adv. 2016, 6, 60922–60925. [Google Scholar] [CrossRef]
  60. Sun, D.; Yin, K.; Zhang, R. Visible-Light-Induced Multicomponent Cascade Cycloaddition Involving N-Propargyl Aromatic Amines, Diaryliodonium Salts and Sulfur Dioxide: Rapid Access to 3-Arylsulfonylquinolines. Chem. Commun. 2018, 54, 1335–1338. [Google Scholar] [CrossRef]
  61. Liu, J.; Wang, M.; Li, L.; Wang, L. Electrooxidative Tandem Cyclization of N-Propargylanilines with Sulfinic Acids for Rapid Access to 3-Arylsulfonylquinoline Derivatives. Green Chem. 2021, 23, 4733–4740. [Google Scholar] [CrossRef]
  62. Zhou, Q.; Xiong, F.-T.; Chen, P.; Xiong, B.-Q.; Tang, K.-W.; Liu, Y. The Visible-Light-Induced Acylation/Cyclization of Alkynoates with Acyl Oximes for the Construction of 3-Acylcoumarins. Org. Biomol. Chem. 2021, 19, 9012–9020. [Google Scholar] [CrossRef]
  63. Yu, W.-Q.; Xie, J.; Chen, Z.; Xiong, B.-Q.; Liu, Y.; Tang, K.-W. Visible-Light-Induced Transition-Metal-Free Nitrogen-Centered Radical Strategy for the Synthesis of 2-Acylated 9H-Pyrrolo[1,2-a]Indoles. J. Org. Chem. 2021, 86, 13720–13733. [Google Scholar] [CrossRef]
  64. He, B.-Q.; Gao, Y.; Wang, P.-Z.; Wu, H.; Zhou, H.-B.; Liu, X.-P.; Chen, J.-R. Dual Photoredox/Palladium-Catalyzed C–H Acylation of 2-Arylpyridines with Oxime Esters. Synlett 2021, 32, 373–377. [Google Scholar]
  65. Qian, H.; Chen, J.; Zhang, B.; Cheng, Y.; Xiao, W.-J.; Chen, J.-R. Visible-Light-Driven Photoredox-Catalyzed Three-Component Radical Cyanoalkylfluorination of Alkenes with Oxime Esters and a Fluoride Ion. Org. Lett. 2021, 23, 6987–6992. [Google Scholar] [CrossRef]
  66. Chen, L.; Jin, S.; Gao, J.; Liu, T.; Shao, Y.; Feng, J.; Wang, K.; Lu, T.; Du, D. N-Heterocyclic Carbene/Magnesium Cocatalyzed Radical Relay Assembly of Aliphatic Keto Nitriles. Org. Lett. 2021, 23, 394–399. [Google Scholar] [CrossRef]
  67. Gao, Y.; Quan, Y.; Li, Z.; Gao, L.; Zhang, Z.; Zou, X.; Yan, R.; Qu, Y.; Guo, K. Organocatalytic Three-Component 1,2-Cyanoalkylacylation of Alkenes Via Radical Relay. Org. Lett. 2021, 23, 183–189. [Google Scholar] [CrossRef] [PubMed]
  68. Chen, L.; Lin, C.; Zhang, S.; Zhang, X.; Zhang, J.; Xing, L.; Guo, Y.; Feng, J.; Gao, J.; Du, D. 1,4-Alkylcarbonylation of 1,3-Enynes to Access Tetra-Substituted Allenyl Ketones Via an NHC-Catalyzed Radical Relay. ACS Catal. 2021, 11, 13363–13373. [Google Scholar] [CrossRef]
  69. Sun, Q.; Zhang, X.-P.; Duan, X.; Qin, L.-Z.; Yuan, X.; Wu, M.-Y.; Liu, J.; Zhu, S.-S.; Qiu, J.-K.; Guo, K. Photoinduced Merging with Copper- or Nickel-Catalyzed 1,4-Cyanoalkylarylation of 1,3-Enynes to Access Multiple Functionalizatized Allenes in Batch and Continuous Flow. Chin. J. Chem. 2022, 40, 1537–1545. [Google Scholar] [CrossRef]
  70. Wang, S.-C.; Shen, Y.-T.; Zhang, T.-S.; Hao, W.-J.; Tu, S.-J.; Jiang, B. Cyclic Oxime Esters as Deconstructive Bifunctional Reagents for Cyanoalkyl Esterification of 1,6-Enynes. J. Org. Chem. 2021, 86, 15488–15497. [Google Scholar] [CrossRef] [PubMed]
  71. Li, M.; Wang, C.-T.; Bao, Q.-F.; Qiu, Y.-F.; Wei, W.-X.; Li, X.-S.; Wang, Y.-Z.; Zhang, Z.; Wang, J.-L.; Liang, Y.-M. Copper-Catalyzed Radical Aryl Migration Approach for the Preparation of Cyanoalkylsulfonylated Oxindoles/Cyanoalkyl Amides. Org. Lett. 2021, 23, 751–756. [Google Scholar] [CrossRef] [PubMed]
  72. Teng, F.; Du, J.; Xun, C.; Zhu, M.; Lu, Z.; Jiang, H.; Chen, Y.; Li, Y.; Gui, Q.-W. Photoinduced Efficient Synthesis of Cyanoalkylsulfonylated Oxindoles Via Sulfur Dioxide Insertion. Org. Biomol. Chem. 2021, 19, 8929–8933. [Google Scholar] [CrossRef] [PubMed]
  73. Zhou, N.; Wu, S.; Kuang, K.; Wu, M.; Zhang, M. Ni-Catalyzed Radical Cyclization of Vinyl Azides with Cyclobutanone Oxime Esters to Access Cyanoalkyl Containing Quinoxalin-2(1H)-Ones. Org. Biomol. Chem. 2021, 19, 4697–4700. [Google Scholar] [CrossRef] [PubMed]
  74. Liang, Q.; Lin, L.; Li, G.; Kong, X.; Xu, B. Synthesis of Phenanthridine and Quinoxaline Derivatives via Copper-Catalyzed Radical Cyanoalkylation of Cyclobutanone Oxime Esters and Vinyl Azides. Chin. J. Chem. 2021, 39, 1948–1952. [Google Scholar] [CrossRef]
  75. Duan, X.; Sun, Q.; Yuan, X.; Qin, L.-Z.; Zhang, X.-P.; Liu, J.; Wu, M.-Y.; Zhu, S.-S.; Ma, C.-L.; Qiu, J.-K.; et al. Photoinduced Remote Heteroaryl Migration Accompanied by Cyanoalkylacylation in Continuous Flow. Green Chem. 2021, 23, 8916–8921. [Google Scholar] [CrossRef]
  76. Chen, J.; Liang, Y.-J.; Wang, P.-Z.; Li, G.-Q.; Zhang, B.; Qian, H.; Huan, X.-D.; Guan, W.; Xiao, W.-J.; Chen, J.-R. Photoinduced Copper-Catalyzed Asymmetric C–O Cross-Coupling. J. Am. Chem. Soc. 2021, 143, 13382–13392. [Google Scholar] [CrossRef]
  77. Hu, Y.-Z.; Ye, Z.-P.; Xia, P.-J.; Song, D.; Li, X.-J.; Liu, Z.-L.; Liu, F.; Chen, K.; Xiang, H.-Y.; Yang, H. Visible-Light-Driven, Photocatalyst-Free Cascade to Access 3-Cyanoalkyl Coumarins from Ortho-Hydroxycinnamic Esters. J. Org. Chem. 2021, 86, 4245–4253. [Google Scholar] [CrossRef] [PubMed]
  78. Zhang, H.; Lv, X.; Yu, H.; Bai, Z.; Chen, G.; He, G. Β-Lactam Synthesis Via Copper-Catalyzed Directed Aminoalkylation of Unactivated Alkenes with Cyclobutanone O-Benzoyloximes. Org. Lett. 2021, 23, 3620–3625. [Google Scholar] [CrossRef]
  79. Qiu, G.; Lai, L.; Cheng, J.; Wu, J. Recent Advances in the Sulfonylation of Alkenes with the Insertion of Sulfur Dioxide Via Radical Reactions. Chem. Commun. 2018, 54, 10405–10414. [Google Scholar] [CrossRef]
  80. Qiu, G.; Zhou, K.; Wu, J. Recent Advances in the Sulfonylation of C–H Bonds with the Insertion of Sulfur Dioxide. Chem. Commun. 2018, 54, 12561–12569. [Google Scholar] [CrossRef]
  81. Chen, P.; Chen, Z.; Xiong, B.-Q.; Liang, Y.; Tang, K.-W.; Xie, J.; Liu, Y. Visible-Light-Mediated Cascade Cyanoalkylsulfonylation/Cyclization of Alkynoates Leading to Coumarins Via SO2 Insertion. Org. Biomol. Chem. 2021, 19, 3181–3190. [Google Scholar] [CrossRef] [PubMed]
  82. Yang, X.; Xia, Y.; Tong, J.; Ouyang, L.; Lai, Y.; Luo, R.; Liao, J. Photoinduced Radical Cascade Cyclization of Acetylenic Acid Esters with Oxime Esters to Access Cyanalkylated Coumarins. Org. Biomol. Chem. 2022, 20, 5239–5244. [Google Scholar] [CrossRef] [PubMed]
  83. Zhou, N.; Xia, Z.; Wu, S.; Kuang, K.; Xu, Q.; Zhang, M. Visible-Light-Induced Multicomponent Cascade Cycloaddition of N-Propargyl Aromatic Amines, Cyclobutanone Oxime Esters, and K2S2O5: Access to Cyanoalkylsulfonylated Quinolines. J. Org. Chem. 2021, 86, 15253–15262. [Google Scholar] [CrossRef]
  84. He, F.-S.; Yao, Y.; Tang, Z.; Qiu, Y.; Xie, W.; Wu, J. Synthesis of β-Cyanoalkylsulfonylated Vinyl Selenides through a Four-Component Reaction. Chem. Commun. 2021, 57, 12603–12606. [Google Scholar] [CrossRef]
  85. Zhou, N.; Xu, Q.; Xia, Z.; Kuang, K.; Wu, S.; Li, W.; Zhang, M. Palladium-Catalyzed Radical Cascade Cyanoalkylsulfonylation/Cyclization of 3-Arylethynyl-[1,1′-Biphenyl]-2-Carbonitriles with Cyclobutanone Oxime Esters and Dabso. Chem. Commun. 2022, 58, 2335–2338. [Google Scholar] [CrossRef]
  86. Xue, D.; Liu, R.; Zhang, D.; Li, N.; Xue, Y.; Ge, Q.; Shao, L. Iron-Catalyzed Radical Cascade Cyclization of Oxime Esters with Isocyanides: Synthesis of 1-Cyanoalkyl Isoquinolines and 6-Cyanoalkyl Phenanthridines. Org. Biomol. Chem. 2021, 19, 8597–8606. [Google Scholar] [CrossRef]
  87. Zuo, H.-D.; Zhu, S.-S.; Hao, W.-J.; Wang, S.-C.; Tu, S.-J.; Jiang, B. Copper-Catalyzed Asymmetric Deconstructive Alkynylation of Cyclic Oximes. ACS Catal. 2021, 11, 6010–6019. [Google Scholar] [CrossRef]
  88. Liu, Y.; Wang, L.; Zeng, L.-H.; Zhao, Y.; Zhu, T.; Wu, J. A Copper-Catalyzed Three-Component Reaction of Alkenes, Cycloketone Oximes and DABCO·(SO2)2: Direct C(sp2)-H Cyanoalkylsulfonylation. Chin. Chem. Lett. 2022, 33, 2383–2386. [Google Scholar] [CrossRef]
  89. Kulthe, A.-D.; Jaiswal, S.; Golagani, D.; Mainkar, P.-S.; Akondi, S.-M. Organophotoredox-Catalyzed Cyanoalkylation of 1,4-Quinones. Org. Biomol. Chem. 2022, 20, 4534–4538. [Google Scholar] [CrossRef]
  90. Ackermann, L. Metalla-Electrocatalyzed C–H Activation by Earth-Abundant 3d Metals and Beyond. Acc. Chem. Res. 2020, 53, 84–104. [Google Scholar] [CrossRef]
  91. Yuan, Y.; Yang, J.; Lei, A. Recent Advances in Electrochemical Oxidative Cross-Coupling with Hydrogen Evolution Involving Radicals. Chem. Soc. Rev. 2021, 50, 10058–10086. [Google Scholar] [CrossRef] [PubMed]
  92. Ding, L.; Niu, K.; Liu, Y.; Wang, Q. Electro-Reductive C-H Cyanoalkylation of Quinoxalin-2(1H)-Ones. Chin. Chem. Lett. 2022, 33, 4057–4060. [Google Scholar] [CrossRef]
  93. Lu, D.; Cui, J.; Yang, S.; Gong, Y. Iron-Catalyzed Cyanoalkylation of Glycine Derivatives Promoted by Pyridine-Oxazoline Ligands. ACS Catal. 2021, 11, 4288–4293. [Google Scholar] [CrossRef]
  94. Liu, S.; Bai, M.; Xu, P.-F.; Sun, Q.-X.; Duan, X.-H.; Guo, L.-N. Copper-Catalyzed Radical Ring-Opening Halogenation with HX. Chem. Commun. 2021, 57, 8652–8655. [Google Scholar] [CrossRef] [PubMed]
  95. Wu, Z.; Li, X.; Li, T.; Xiao, T.; Jiang, Y.; Qin, G. Fe-Catalyzed Denitrative Cyanoalkylation of Nitroalkenes with Cycloketone Oxime Esters Via Reductive C–C Bond Formation. Org. Chem. Front. 2022, 9, 6319–6324. [Google Scholar] [CrossRef]
  96. Lu, X.-Y.; Xia, Z.-J.; Gao, A.; Liu, Q.-L.; Jiang, R.-C.; Liu, C.-C. Dual Nickel/Ruthenium Strategy for Photoinduced Decarboxylative Cross-Coupling of α,β-Unsaturated Carboxylic Acids with Cycloketone Oxime Esters. J. Org. Chem. 2021, 86, 8829–8842. [Google Scholar] [CrossRef] [PubMed]
  97. Mugesh, G.; Du Mont, W.-W.; Sies, H. Chemistry of Biologically Important Synthetic Organoselenium Compounds. Chem. Rev. 2001, 101, 2125–2180. [Google Scholar] [CrossRef]
  98. Ji, L.; Qiao, J.; Liu, J.; Tian, M.; Lu, K.; Zhao, X. Metal-Free Chalcogenation of Cycloketone Oxime Esters with Dichalcogenides. Tetrahedron Lett. 2021, 75, 153202. [Google Scholar] [CrossRef]
  99. Zhao, X.; Ji, L.; Gao, Y.; Sun, T.; Qiao, J.; Li, A.; Lu, K. Visible-Light-Promoted Selenocyanation of Cyclobutanone Oxime Esters Using Potassium Selenocyanate. J. Org. Chem. 2021, 86, 11399–11406. [Google Scholar] [CrossRef]
  100. Ren, L.-Q.; Li, N.; Ke, J.; He, C. Recent Advances in Photo- and Electro-Enabled Radical Silylation. Org. Chem. Front. 2022, 9, 6400–6415. [Google Scholar] [CrossRef]
  101. Li, N.; Li, Y.; Wu, X.; Zhu, C.; Xie, J. Radical Deuteration. Chem. Soc. Rev. 2022, 51, 6291–6306. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, Y.; Cai, Z.; Warratz, S.; Ma, C.; Ackermann, L. Recent Advances in Electrooxidative Radical Transformations of Alkynes. Sci. Chin. Chem. 2022, 66, 703–724. [Google Scholar] [CrossRef]
Scheme 1. Main pathway switching for NCRs to CCRs.
Scheme 1. Main pathway switching for NCRs to CCRs.
Molecules 28 02667 sch001
Scheme 2. Visible-light-induced radical acylation/cyclization of acyl oxime esters with olefins.
Scheme 2. Visible-light-induced radical acylation/cyclization of acyl oxime esters with olefins.
Molecules 28 02667 sch002
Scheme 3. Visible-light-induced radical difunctionalization of acyl oxime esters with styrenes and nitriles.
Scheme 3. Visible-light-induced radical difunctionalization of acyl oxime esters with styrenes and nitriles.
Molecules 28 02667 sch003
Scheme 4. Visible-light-induced hydroacylation of acyl oximes with alkenes.
Scheme 4. Visible-light-induced hydroacylation of acyl oximes with alkenes.
Molecules 28 02667 sch004
Scheme 5. Visible-light photoredox-catalyzed dicarbofunctionalization of styrenes with oxime esters and CO2.
Scheme 5. Visible-light photoredox-catalyzed dicarbofunctionalization of styrenes with oxime esters and CO2.
Molecules 28 02667 sch005
Scheme 6. N-heterocyclic carbene-photocatalyzed 1,2-diacylation of alkenes with acyl oxime esters and aldehydes.
Scheme 6. N-heterocyclic carbene-photocatalyzed 1,2-diacylation of alkenes with acyl oxime esters and aldehydes.
Molecules 28 02667 sch006
Scheme 7. Visible-light-induced acylation of acyl oxime esters with 2-allyloxy benzaldehydes.
Scheme 7. Visible-light-induced acylation of acyl oxime esters with 2-allyloxy benzaldehydes.
Molecules 28 02667 sch007
Scheme 8. Visible-light-induced acylation of acyl oxime esters with reactive olefins.
Scheme 8. Visible-light-induced acylation of acyl oxime esters with reactive olefins.
Molecules 28 02667 sch008
Scheme 9. Visible-light-induced acylation of acyl oxime esters with N-(arylsulfonyl)acrylamides.
Scheme 9. Visible-light-induced acylation of acyl oxime esters with N-(arylsulfonyl)acrylamides.
Molecules 28 02667 sch009
Scheme 10. Visible-light photoredox-catalyzed acylation of alkynes with acyl oxime esters.
Scheme 10. Visible-light photoredox-catalyzed acylation of alkynes with acyl oxime esters.
Molecules 28 02667 sch010
Scheme 11. Photoredox/palladium-catalyzed C-H acylation of 2-arylpyridines with acyl oxime esters.
Scheme 11. Photoredox/palladium-catalyzed C-H acylation of 2-arylpyridines with acyl oxime esters.
Molecules 28 02667 sch011
Scheme 12. Visible-light-induced coupling of benzyl oxime esters with O- and N- nucleophiles.
Scheme 12. Visible-light-induced coupling of benzyl oxime esters with O- and N- nucleophiles.
Molecules 28 02667 sch012
Scheme 13. Visible-light-induced cyanoalkylfluorination of alkenes with cycloketone oxime esters and fluoride.
Scheme 13. Visible-light-induced cyanoalkylfluorination of alkenes with cycloketone oxime esters and fluoride.
Molecules 28 02667 sch013
Scheme 14. N-heterocyclic carbene-catalyzed radical acylation/cyanoalkylation of alkenes with cycloketone oxime esters and aldehydes.
Scheme 14. N-heterocyclic carbene-catalyzed radical acylation/cyanoalkylation of alkenes with cycloketone oxime esters and aldehydes.
Molecules 28 02667 sch014
Scheme 15. N-heterocyclic carbene-catalyzed 1,4-alkylcarbonylation of 1,3-enynes with cycloketone oxime esters and aldehydes.
Scheme 15. N-heterocyclic carbene-catalyzed 1,4-alkylcarbonylation of 1,3-enynes with cycloketone oxime esters and aldehydes.
Molecules 28 02667 sch015
Scheme 16. Photo/Cu co-catalyzed 1,4-cyanoalkylarylation of 1,3-enynes with cycloketone oxime esters and boronic acids.
Scheme 16. Photo/Cu co-catalyzed 1,4-cyanoalkylarylation of 1,3-enynes with cycloketone oxime esters and boronic acids.
Molecules 28 02667 sch016
Scheme 17. Cu-catalyzed difunctionalization of 1,6-enynes with cycloketone oxime esters.
Scheme 17. Cu-catalyzed difunctionalization of 1,6-enynes with cycloketone oxime esters.
Molecules 28 02667 sch017
Scheme 18. Cu-catalyzed difunctionalization of N-sulfonylated acrylamides with cycloketone oxime esters.
Scheme 18. Cu-catalyzed difunctionalization of N-sulfonylated acrylamides with cycloketone oxime esters.
Molecules 28 02667 sch018
Scheme 19. Photoinduced cyanoalkylsulfonylation of N-arylacrylamides with cycloketone oxime esters.
Scheme 19. Photoinduced cyanoalkylsulfonylation of N-arylacrylamides with cycloketone oxime esters.
Molecules 28 02667 sch019
Scheme 20. Nickel- or copper-catalyzed cyanoalkylation of vinyl azides with cycloketone oxime esters.
Scheme 20. Nickel- or copper-catalyzed cyanoalkylation of vinyl azides with cycloketone oxime esters.
Molecules 28 02667 sch020
Scheme 21. Photoinduced cyanoalkylation of azidyl homoallylic alcohol with cycloketone oxime esters.
Scheme 21. Photoinduced cyanoalkylation of azidyl homoallylic alcohol with cycloketone oxime esters.
Molecules 28 02667 sch021
Scheme 22. Photoinduced copper-catalyzed asymmetric difunctionalization of 1,3-dienes with cycloketone oxime esters.
Scheme 22. Photoinduced copper-catalyzed asymmetric difunctionalization of 1,3-dienes with cycloketone oxime esters.
Molecules 28 02667 sch022
Scheme 23. Visible-light-driven cyanoalkylation of 2-hydroxyalkenyl esters with cycloketone oxime esters.
Scheme 23. Visible-light-driven cyanoalkylation of 2-hydroxyalkenyl esters with cycloketone oxime esters.
Molecules 28 02667 sch023
Scheme 24. Cu-mediated cyanoalkylation/cyclization of inactivated alkenes with cycloketone oxime esters.
Scheme 24. Cu-mediated cyanoalkylation/cyclization of inactivated alkenes with cycloketone oxime esters.
Molecules 28 02667 sch024
Scheme 25. Visible-light-induced cyanoalkylation of alkynoates with cycloketone oxime esters.
Scheme 25. Visible-light-induced cyanoalkylation of alkynoates with cycloketone oxime esters.
Molecules 28 02667 sch025
Scheme 26. Visible-light-induced cyanoalkylsulfonylation of N-propargyl amines with cycloketone oxime esters.
Scheme 26. Visible-light-induced cyanoalkylsulfonylation of N-propargyl amines with cycloketone oxime esters.
Molecules 28 02667 sch026
Scheme 27. Copper-catalyzed four-component selenosulfonylation of alkynes, cycloketone oxime esters, DABCO and diselenides.
Scheme 27. Copper-catalyzed four-component selenosulfonylation of alkynes, cycloketone oxime esters, DABCO and diselenides.
Molecules 28 02667 sch027
Scheme 28. Palladium-catalyzed cyanoalkylsulfonylation of alkynes with cycloketone oxime esters and DABSO.
Scheme 28. Palladium-catalyzed cyanoalkylsulfonylation of alkynes with cycloketone oxime esters and DABSO.
Molecules 28 02667 sch028
Scheme 29. Iron-catalyzed radical cascade cyclization of isocyaniness with cycloketone oxime esters.
Scheme 29. Iron-catalyzed radical cascade cyclization of isocyaniness with cycloketone oxime esters.
Molecules 28 02667 sch029
Scheme 30. Copper-catalyzed radical cascade of inactivated alkynes or alkenes with cycloketone oxime esters.
Scheme 30. Copper-catalyzed radical cascade of inactivated alkynes or alkenes with cycloketone oxime esters.
Molecules 28 02667 sch030
Scheme 31. Visible-light-induced cross-coupling of 1,4-quinones with cycloketone oxime esters.
Scheme 31. Visible-light-induced cross-coupling of 1,4-quinones with cycloketone oxime esters.
Molecules 28 02667 sch031
Scheme 32. Electrocatalytic C-H cyanoalkylation of quinoxalin-2(1H)-ones with cycloketone oxime esters.
Scheme 32. Electrocatalytic C-H cyanoalkylation of quinoxalin-2(1H)-ones with cycloketone oxime esters.
Molecules 28 02667 sch032
Scheme 33. Iron-catalyzed cyanoalkylation of glycine derivatives with cycloketone oxime esters.
Scheme 33. Iron-catalyzed cyanoalkylation of glycine derivatives with cycloketone oxime esters.
Molecules 28 02667 sch033
Scheme 34. Copper-catalyzed halogenation of cycloketone oxime esters.
Scheme 34. Copper-catalyzed halogenation of cycloketone oxime esters.
Molecules 28 02667 sch034
Scheme 35. Iron-catalyzed denitrative cyanoalkylation of nitroalkenes with cycloketone oxime esters.
Scheme 35. Iron-catalyzed denitrative cyanoalkylation of nitroalkenes with cycloketone oxime esters.
Molecules 28 02667 sch035
Scheme 36. Photo-nickel dual-catalyzed cyanoalkylation of α, β-unsaturated carboxylic acids with cycloketone oxime ester.
Scheme 36. Photo-nickel dual-catalyzed cyanoalkylation of α, β-unsaturated carboxylic acids with cycloketone oxime ester.
Molecules 28 02667 sch036
Scheme 37. Free radical coupling of cycloketone oxime esters with elements of the group of sulfur.
Scheme 37. Free radical coupling of cycloketone oxime esters with elements of the group of sulfur.
Molecules 28 02667 sch037
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

Chen, P.; Huang, H.; Tan, Q.; Ji, X.; Zhao, F. Recent Advances in Molecule Synthesis Involving C-C Bond Cleavage of Ketoxime Esters. Molecules 2023, 28, 2667. https://doi.org/10.3390/molecules28062667

AMA Style

Chen P, Huang H, Tan Q, Ji X, Zhao F. Recent Advances in Molecule Synthesis Involving C-C Bond Cleavage of Ketoxime Esters. Molecules. 2023; 28(6):2667. https://doi.org/10.3390/molecules28062667

Chicago/Turabian Style

Chen, Pu, Huawen Huang, Qi Tan, Xiaochen Ji, and Feng Zhao. 2023. "Recent Advances in Molecule Synthesis Involving C-C Bond Cleavage of Ketoxime Esters" Molecules 28, no. 6: 2667. https://doi.org/10.3390/molecules28062667

Article Metrics

Back to TopTop