Yu-Yang Xie
a and
Ying-Ming Pan
*b
aJiangsu Key Laboratory of New Drug Research and Clinical Pharmacy, School of Pharmacy of Xuzhou Medical University, Xuzhou 221004, People's Republic of China
bState Key Laboratory for Chemistry and Molecular Engineering of Medicinal Resources, School of Chemistry and Pharmaceutical Sciences of Guangxi Normal University, Guilin 541004, People's Republic of China. E-mail: panym@mailbox.gxnu.edu.cn
First published on 12th August 2024
The primary greenhouse gas is carbon dioxide (CO2), which also serves as an excellent C1 building block for the synthesis of heterocyclic compounds. The preparation and transformation of these structures have long been a focus in organic chemistry, making the utilization of CO2 highly advantageous. This paper reviews progress in the photochemical/electrochemical construction of heterocyclic compounds using CO2, differentiating between various reaction types and elucidating their underlying mechanisms as well as potential applications.
However, the activation of linear CO2 presents a significant challenge due to its inherent thermodynamic and kinetic stability.9 As a result, conventional CO2 conversion reactions require the use of highly active reagents or harsh reaction conditions,10 such as carboxylation with air-sensitive organometallic reagents that have high-energy substrates. Additionally, the incorporation of a metal catalyst11 to activate CO2 or the starting material can expedite the reaction by reducing the energy barrier associated with the transition state. However, there are still several key issues that need to be addressed in these transformations, such as the utilization of precious metal catalysts and stoichiometric oxidants. Therefore, it is imperative to develop green, efficient, and convenient methods for CO2 utilization. Furthermore, the significance of environmentally sustainable synthetic chemistry has been increasingly emphasized in recent years. In particular, photochemical/electrochemical organic synthesis has gained significant attention due to its ability to produce highly active intermediates such as free radicals, free radical ions, and electron transfer complexes that are difficult to obtain through conventional methods.12 Additionally, using electricity instead of various redox reagents in electrochemical synthesis aligns with the current trend and demand for reducing low carbon emissions.13 Under photochemical reaction conditions, light irradiation facilitates the conversion of reactants into active free radical intermediates under mild conditions, leading to the formation of final products.14 Considering these aspects, photoelectrochemical CO2 fixation appears more appealing and convenient compared to traditional methods.
Although numerous reviews15 are available in published studies on CO2 carboxylation reactions,16 there is a lack of comprehensive summaries regarding the synthesis of heterocyclic compounds17 using CO2. This prompts us to further explore the potential of photo/electrochemical synthesis for CO2 reactions. Here, we will provide a comprehensive overview of the research progress in the synthesis of heterocyclic compounds involving CO2 participation, with a focus on distinct reaction types within photochemistry and electrochemistry. Additionally, we will elucidate the practicality of these reactions and their resulting products while delving into the underlying reaction mechanisms.
Entry | Photocatalyst | E1/2 (M+/M*) | E1/2 (M*/M−) | E1/2 (M+/M) | E1/2 (M/M−) | Ref. |
---|---|---|---|---|---|---|
a Redox potentials are reported in V vs. SCE. | ||||||
1 | Ru(bpy)32+ | −0.81 | +0.77 | +1.29 | −1.33 | 23 |
2 | Ir[dF(CF3)ppy]2(dtb-bpy)+ | −0.89 | +1.21 | +1.69 | −1.37 | 23 |
3 | fac-Ir(ppy)3 | −1.73 | +0.31 | +0.77 | −2.19 | 23 |
4 | [Ir(ppy)2(dtbbpy)]+ | −0.96 | +0.66 | +1.21 | −1.51 | 23 |
5 | 4CzIPN | −1.04 | +1.35 | +1.52 | −1.21 | 24 |
According to the literature, photochemical fixation of CO2 into heterocycles can be classified into four modes: (i) single electron transfer (SET) processes; (ii) cooperative/dual photocatalysis; (iii) visible light-induced transition metal catalyzed transformations; and (iv) photoexcitation.
In recent years, chemists have developed various methods for generating different types of radicals through visible-light photoredox catalysis, which offers distinct advantages such as high efficiency,26 mild conditions, and easy operation. In this regard, Yu and his colleagues demonstrated a strategy using allylamines 1, difluoroalkyl reagents 2, and CO2 via visible-light photoredox catalysis to synthesize functionalized 2-oxazolidinones 3 (Scheme 3).27 The experiment demonstrates that allylamine with aryl substituents is suitable as a viable substrate for the reaction. Furthermore, various difluoroalkyl reagents were investigated and found to be effective in obtaining the desired compound. Additionally, by increasing the amount of substrate to 4 mmol, a product yield of 84% (1.31 g) was successfully achieved. Subsequent functional group transformations were efficiently conducted by hydrolyzing the ester group CF2COOEt into carboxylic acid 4 while leaving the 2-oxazolidinone skeleton unaffected. Moreover, LiAlH4 and NaBH4 served as efficient and selective reducing agents to obtain reducers 5 and 6, respectively. All of these factors indicate great potential for widespread use of this conversion in organic synthesis and medicinal chemistry.
This difluoroalkylation/carboxylative cyclization can use BrCF2COOEt as the radical precursor (Scheme 4). Mechanistically, DABCO reduction quenches the excited photocatalyst to produce RuI, reduces BrCF2COO to produce the RuI catalyst, and reduces BrCF2COOEt 2 to produce CF2COOEt radicals B. Carbamate A is formed in situ using 1 and CO2, followed by the addition of CF2COOEt radical B to the double bond of carbamate to generate benzylic radical C. The subsequent step involves the oxidation of C by excited *RuII species and intramolecular cyclization of D affording product 3.
Carboxylic bifunctionalization of propargyl amines to obtain functionalized vinyloxazolidinones has been well studied.28 The previous methods, although possessing potential advantages, rely on high loadings of metal catalysts, oxidants and/or elevated temperatures and are restricted to propargyl amines. Recently, McGarrigle and co-workers reported a visible light-assisted carboxylative sulfonylation by employing sodium arylsulfinates 8 as radical precursors, which afforded oxazolidinones 9 and vinyloxazinones 10 (Scheme 5).29 α-Ethyl-substituted bulky and aromatic heterocyclic furan-3-ylmethyl and propyl-substituted propargyl amines were also well tolerated in this protocol and yielded the corresponding oxalidones (9b–9d). In contrast, N-phenyl and N-tosyl substituted propargyl amines failed to generate the corresponding sulfonylated oxazolidinones, possibly due to the weak nucleophilicity of the amines. Interestingly, N-substituted homopropargyl amines were also found to be compatible under the reaction conditions. This is the first report on carboxylative functionalization of homopropargyl amines for synthesizing six-membered rings. N-Cyclohexyl homopropargyl amine, halogen (4-F) bearing sulfinates, and simple benzenesulfinate reacted smoothly under this protocol yielding the corresponding compounds 10b–10d.
Scheme 5 Photocatalyst-promoted synthesis of oxazolidinones and vinyloxazinones from propargylamines with CO2 and sodium arylsulfinates. |
A plausible mechanism was proposed (Scheme 6). The RuII complex is initially excited by the absorption of blue light, which then oxidizes sulfinate 8 to form sulfinyl radical B and RuI species. Propargyl amine 7 reacts with CO2 and a base to form carbamate intermediate A. The addition of carbamate A to sulfinyl radical B produces vinyl radical intermediate C. Intermediate C then undergoes SET with Ru, followed by a cyclization reaction to generate cyclic intermediate D. Intermediate D might undergo SET with RuI, followed by the addition of H+ to generate a radical intermediate G and regenerate the RuII complex. The reaction of intermediate G in the presence of a base and photocatalyst leads to the elimination of H+ and affords product 9. Alternatively, in the presence of a base, intermediate D can form anion E/F, which is subsequently protonated with H+ to afford the final product 9.
The production of CO2 as a by-product is commonly observed in nearly all decarboxylation-involving reactions, and attempts to effectively utilize this crucial C1 building block for subsequent conversions have been met with limited success. In 2021, Sun and co-workers reported a new visible light-induced photoredox-catalyzed α-aminomethyl carboxylation of styrenes 11, leading to diverse γ-lactams 13 (Scheme 7).30 The fact that the CO2 released during the decarboxylation step can be recycled for subsequent carboxylation reactions is noteworthy. The present work extends the reaction substrate to aliphatic carboxylic acids, which is unheard of for transition metal-catalyzed decarboxylative coupling reactions. The mechanism suggested for the formation of γ-lactams is proposed. The decarboxylation reaction, under the influence of photocatalysis, generates aminomethyl radical A, which subsequently undergoes addition to the olefin double bond resulting in the formation of carbon radical B. The photocatalyst reduction and subsequent single-electron transfer result in the generation of carbanion C. Next, sequestration of the released CO2 with this anion, leading to protonation, forms γ-amino acid D. The presence of a free N–H bond in the amino group facilitates intramolecular lactamidation of intermediate D, resulting in the formation of various γ-lactams 13.
Scheme 7 Visible-light photoredox-catalyzed reductive α-aminomethyl carboxylation of styrenes with sodium glycinates and CO2. |
In 2024, the same group reported the synthesis of lactams through visible-light photoredox-catalyzed intermolecular sequential α-aminomethyl/carboxylative dearomatization of indoles 14 with CO2 and α-aminoalkyl radical precursors (Scheme 8).31 The source of CO2 differed from that in previous reports.
Scheme 8 Visible-light photoredox-catalyzed intermolecular sequential carboxylative dearomatization of indoles with CO2 and α-aminoalkyl radical precursors. |
In 2020, Yu and colleagues reported a strategy involving visible-light photoredox-catalyzed successive single electron transfer to enable the dearomative arylcarboxylation of indoles 17 with CO2 (Scheme 9).32 The present method allows for selective avoidance of ipso-carboxylation of aryl halides and β-hydride elimination, which are the primary by-products associated with conventional aryl halide double bond functionalization processes. This approach offers a comparatively more straightforward means of accessing indoline-3-carboxylic acids. Despite the challenging steric hindrance, the researchers successfully achieved a 50% yield in obtaining biquaternary carbon products 18b at the C2 and C3 positions. Moreover, this strategy extends beyond the formation of five-membered rings, enabling the synthesis of the corresponding six-membered (18c) and seven-membered (18d) products with separation yields of 42% and 14%, respectively, under standard reaction conditions. To further demonstrate the utility of this method, a gram-scale reaction was conducted under standard conditions resulting in an impressive yield of compound 18a at 80%. Additionally, valuable structures can be easily obtained through simple and selective derivatizations such as bromination or reduction of carboxylic acids.
Scheme 9 The dearomative arylcarboxylation of indoles with CO2 via visible-light photoredox catalysis. |
Based on the Stern–Volmer, isotope-labeling, and radical trapping experiments, it has been demonstrated that benzylic radicals and anions can be generated as key intermediates. This provides a pathway for reductive couplings with other electrophiles such as D2O and aldehyde (Scheme 10). Mechanistically, the photo-activated 1,2,3,5-tetrakis(carbazol-9-yl)-4,6-dicyanobenzene (4CzIPN) undergoes reductive quenching by DIPEA leading to the formation of the radical anion 4CzIPN˙− and the radical cation DIPEA˙+ in subsequent steps. The reduced 4CzIPN˙− then generates the [ArBr]˙− radical anion A through single-electron reduction while rejuvenating the catalyst 4CzIPN. The generated [ArBr]˙− radical anion subsequently undergoes fragmentation liberating the bromide anion and aryl radical B. Subsequently, aryl radical B readily undergoes intramolecular radical addition to indole's C2–C3 double bond forming benzylic radical C. In the next step, C reacts with 4CzIPN˙− generating carboanion D which further undergoes nucleophilic addition to CO2, followed by protonation resulting in the dearomatized arylcarboxylation product 18a.
Later, the same group developed the visible-light photoredox-catalyzed arylcarboxylation of unactivated alkenes 22 with CO2, leading to a variety of highly valued polycyclic acids 23 in moderate to good yields (Scheme 11).33 The reaction exhibits high chemo- and regio-selectivities, mild reaction conditions, good compatibility with functional groups, and broad substrate scope, allowing for easy scalability and facile derivatization of products. A wide range of electron-donating groups and electron-withdrawing groups were tolerated at the ortho/meta/para-positions of the arene moiety, providing the desired products in moderate to good yields. Substrates with different substituents on the aliphatic chain and 1,1-disubstituted unactivated alkenes were also suitable for this transformation. Additionally, indolin-3-ylacetic acid derivatives, chroman-4-ylacetic acid derivatives, and thiochromane-4-ylacetic acid derivatives can be prepared using this strategy.
The reaction produced an impressive 84% yield of product 23a, demonstrating its scalability. Furthermore, derivatization of 23a was conducted to showcase its potential synthetic applications. Selective reduction of product 23a using NaBH4 resulted in a remarkable yield of 92%. The condensation between compound 23a and glycine methyl ester led to the high-yield formation of cycloamide 25. Additionally, the successful decarboxylation of the primary carboxylic acid in compound 23a was enabled by synergistic photoredox and HAT catalysis. Moreover, compound 23a exhibited moderate reactivity in decarboxyl-trifluoromethylation, resulting in the formation of compound 27. It is worth noting that compound 23a readily converted into oxidation-active ester 28, facilitating C–P bond formation through decarboxylphosphorylation as well as C–S bond formation via aryl thioylation (Scheme 12).
A possible mechanism for the overall transformation of unactivated alkenes is proposed (Scheme 13). The irradiated photocatalyst IrIII generates excited *IrIII, which can be reductively quenched through a catalytic thiolate to produce IrII and a thiyl radical. The IrII species may then participate in the reduction of CO2 through a SET event, resulting in the generation of CO2˙− and the regeneration of IrIII, thus completing the photoredox catalytic cycle. The in situ generated CO2˙− undergoes radical addition to the CC double bond of an unactivated alkene 22 to form an alkyl carbon radical A, which is expected to rapidly undergo cyclization and form the radical intermediate B. Finally, carboxylate B can be obtained through a hydrogen atom transfer (HAT) process involving radical intermediate C and the thiyl radical while simultaneously regenerating the thiol catalyst. The final product 23 is then formed during workup by protonation. Furthermore, it is proposed by the authors that incorporating silane as an additive can effectively facilitate alternative pathways for generating CO2˙−.
In recent years, transition metal-catalyzed functionalized carboxylation of alkynes with CO2 has been extensively studied by many groups.34 However, the excessive use of metal reagents and the presence of transition metal residues have prevented their wide applications in industry. Moreover, the majority of these reactions proceed through two primary pathways: either the cyclometallation of CO2 with a low-valent transition-metal catalyst and alkyne or the carboxylation of the in situ generated alkenyl metallic intermediate with CO2. Both pathways involve two-electron activation processes. Visible-light photochemistry has emerged as a captivating strategy for achieving innovative chemical transformations in an environmentally friendly manner. The hydrocarboxylation of alkynes driven by visible light poses greater challenges compared to that of alkenes due to the difficulty in achieving high levels of chemical, regional, and Z/E selectivity, which may undergo mutual changes under light irradiation. The issue was resolved by Wu and Zhao in 2018 through the utilization of the visible light photooxidation/cobalt dual catalysis technique.35 Subsequently, difunctionalizing carboxylation of alkynes 32 with CO2 was further extended by Yu's group in 2023 (Scheme 14).36 Various substituted thiophenols 31 and aryl acetylenes 32 could be directly converted into important target thiochromones 33 with high chemoselectivity and good regioselectivity. The involvement of single electron activation of CO2 in thiocarboxylation has been demonstrated by mechanistic and computational studies, leading to unique β-carboxylation. To demonstrate the synthetic potential of this one-pot scheme, gram-scale reactions and several product derivatization experiments were carried out. The gram-scale reaction succeeded in obtaining the target compound at a yield rate of 75%. Additionally, thiochromones can be derived into CMV protease inhibitors, insecticidal compounds, and antimalarial compounds.
In 2018, Wu, Zhao, and their colleagues developed a one-pot hydrocarboxylation/isomerization/cyclization protocol that utilizes iridium/cobalt dual catalysis to efficiently produce bicyclic scaffolds from simple alkyne precursors 34 (Scheme 16).35 This approach not only enables facile access to coumarins but also allows for the generation of 2-quinolones and 2-benzoxepinones. A proposed mechanism is outlined in Scheme 16. The visible-light-excited *IrIII undergoes reductive quenching by iPr2NEt to form an amine radical cation and a reduced IrII species which is then oxidized by the CoII complex to regenerate IrIII. The active CoI species generated in this step reacts with CO2 and alkyne substrates 34, leading to the formation of five-membered cobaltacycle intermediate A. Protonolysis of the C–CoIII bond by cationic amine radicals results in carboxylatocobalt intermediate C formation. The transient complex C may undergo photoredox-catalyzed reduction facilitated by ZnBr2 through transmetalation, resulting in carboxylate D formation. Finally, aryl-substituted compound D can undergo reversible isomerization via an Ir-mediated triplet–triplet energy transfer process under visible-light irradiation producing product 37 through acid-mediated intramolecular cyclization. Terminal alkynes will insert another alkyne molecule into five-membered cobaltacycle A forming seven-membered cobaltacycle B, followed by reductive elimination yielding 2-pyrone 36 and regenerating CoI. In the case of the substrate ortho-ester substituted aryl alkyne, the in situ generated C–CoIII bond undergoes intramolecular addition to the ester substituent, resulting in the formation of carbo-carboxylation intermediate G. Complex G can potentially undergo photoredox-catalyzed reduction and transmetalation with ZnBr2, leading to the formation of zinc complex H while simultaneously regenerating CoI species. The tautomerization of H yields γ-keto acrylic zinc species I, which further undergoes another tautomerization to achieve the synthesis of γ-hydroxybutenolide 38.
Scheme 16 The direct synthesis of various pharmaceutically important heterocycles from alkynes and CO2. |
In 2018, Yu's group developed oxy-alkylation of allylamines 39 with CO2 and unactivated alkyl bromides via visible light-driven palladium catalysis (Scheme 18a).38 Exclusively utilizing commercially available Pd(PPh3)4 as the photocatalyst, blue LED light is employed for irradiation purposes. A variety of primary, secondary and tertiary alkyl bromides are applied in this reaction to give 2-oxazolidinones 41 under 1 atm of CO2. In 2019, Cheng and co-workers adapted a similar catalytic system to the radical oxy-alkylation of 2-(1-arylvinyl)anilines 42 with unactivated alkyl bromides and CO2, which further expanded the substrate to 1,4-dihydro-2H-3,1-benzoxazin-2-ones 43 (Scheme 18b).39
The visible light-induced transition metal-catalyzed reactions, in which the catalytic cycle is initiated by photoexcitation of the Pd0 catalyst, involve several steps (Scheme 19). First, the excited Pd0L* undergoes oxidation by alkyl bromide 40, resulting in the formation of active PdI species C and an alkyl radical. Then, substrate 39a rapidly reacts with CO2 to form carbamate A. The introduction of the radical into the alkene within carbamate A leads to the formation of a more stabilized benzylic radical B. Subsequently, a SET occurs between intermediate B and PdI species, regenerating the Pd0 catalyst and forming the stabilized cation D. This cation is then quickly trapped intramolecularly by a nucleophilic carbamate, resulting in the desired product 41a.
In 1972, Giezendanner and his colleagues initially demonstrated the pioneering application of photochemistry in the fixation of CO2 to heterocycles. They successfully achieved the catalyst-free synthesis of substituted 2H-oxazol-5-ones 45 through the reaction between various arylazirines 44 and CO2 under mercury lamp irradiation (Scheme 21a).41 The tunable structures of arylazirines 46 can lead to an enhanced isolated yield of 2H-oxazol-5-ones 47 in the catalyst-free system, thereby improving overall efficiency (Scheme 21b).42 The reaction of 2,3-diphenyl-2H-arizine 44a resulted in the formation of an equimolar mixture comprising two inseparable isomers.
The Minakata group has reported the conversion of CO2 into 2-oxazolidinone for the first time.43 Subsequent advancements in this reaction model primarily focused on expanding active iodinating reagents or starting materials. Expanding the application of photochemical synthesis to CO2 fixation represents a highly promising field. In 2017, He et al. demonstrated the utilization of perfluoroalkyl iodides as radical and fluorine sources in a photoinduced radical-initiated protocol for the carboxylative cyclization of allyl amines 48 with CO2 (Scheme 22).44 The perfluoroalkylated cyclic carbonate 51 is successfully obtained via this protocol from allyl alcohol and CO2, which is particularly noteworthy.
A plausible mechanism for the metal-free and visible light-driven carboxylative cyclization of allyl amines 48 with CO2 is illustrated in Scheme 22 through a set of control experiments. The carbamate A is readily formed from 48 and CO2, facilitated by TBD in the initial stage. Under the initiation of visible light, the carbon–iodine bond in nC4F9I undergoes homolytic cleavage, resulting in the generation of ˙nC4F9 and I˙ radical species. The addition of the ˙nC4F9 radical to the π-bond of carbamate A yields a perfluoroalkylated secondary carbon radical B, which is subsequently captured by the I˙ radical to form the iodine perfluoroalkylated carbamate intermediate C (Path I). Alternatively, the intermediate C could potentially be generated through a radical propagation of B with nC4F9I (Path II). Finally, the perfluoroalkylated 2-oxazolidinone 49 is obtained through intramolecular nucleophilic cyclization of C, accompanied by the release of [TBDH]+I−.
Shortly thereafter, the same group demonstrated a unique strategy for synthesizing exo-iodomethylene 2-oxazolidinones using propargylic amines 52 as substrates and CO2 as a C1 synthon, and initiating the carboxylative cyclization of propargylic amines with CO2 through molecular iodine (Scheme 23).45 The carbamate A was formed through TMG-mediated coupling of amine 52 and CO2 in the initial stage, according to the proposed mechanism illustrated in Scheme 23. Concurrently, under visible light irradiation, the generation of iodine radical I˙ from I2 was achieved. Subsequently, insertion of I˙ species into the triple bond of intermediate A occurred to form radical intermediate B. The interaction between B and another radical I˙ ultimately resulted in liberation of [TMGH]+I−, along with carboxylative annulation, thereby yielding the desired target product 53.
Scheme 23 The visible light-mediated synthesis of exo-iodomethylene 2-oxazolidinones from CO2, propargylic amines, and iodine. |
To our satisfaction, the desired product 53a was conveniently obtained with a separation yield of 88% through an amplification reaction using 4 mmol of the starting material (Scheme 24).
Furthermore, the researchers explored various transformations of iodinated 2-oxazolidinones in order to demonstrate the versatility of this visible light-driven metal-free scheme in the carboxylative cyclization synthesis of propylamine with CO2. Fortunately, a successful Buchwald–Hartwig reaction, Suzuki reaction, and Sonogashira reaction could be carried out using 53a as the substrate, resulting in good yields of products 54, 55, and 56, respectively. In addition, 53a is capable of photoreducing and photocatalyzing the “ene Reaction”, resulting in the formation of compounds 57 and 58, respectively. The extensive array of transformations observed in this study serves to illustrate the versatility of exo-iodomethyl 2-oxazolidinone as a chemical substrate. Consequently, it can be regarded as a foundational compound that readily lends itself to the synthesis of numerous valuable derivatives.
Scheme 26 The derivatives of maleic anhydride were constructed by electrochemical carboxylation/cyclization of CO2 and alkynes. |
However, when catalytic amounts of CuI were introduced into the system (with 4 MPa CO2), tricarboxylation product 61 was furnished instead.49 It is speculated that the copper salt can coordinate to the double bond of the dicarboxylate intermediate 59, facilitating the attack of CO2 radical anions on the double bond, followed by the attack of the CO2 radical anions to produce F. Furthermore, F is reduced to G. After being acidified and dehydrated, the product 61 is obtained.
Buckley, Wijayantha, and co-workers reported the electrosynthesis of cyclic carbonates from epoxides and atmospheric CO2 in the absence of expensive catalysts (Scheme 27).50 The reaction was carried out in an undivided cell equipped with acetonitrile as the solvent, n-Bu4NBr as the electrolyte, and a 60 mA current. Excellent conversion into cyclic carbonate was achieved using a copper cathode/magnesium anode combination. In addition, the use of tetrabutylammonium bromide is essential for the reaction; only 17% conversion to the cyclic carbonate was observed when the bromide counterion was replaced by tetrafluoroborate. Without passing a current, the reaction did not occur. A wide range of substrates are suitable for the approach; excellent yields of cyclic carbonate 63 were produced, including electron-rich and electron-poor aromatic and aliphatic epoxides 62. The approach is easy to set up, is reliable, requires no expensive catalyst, and runs under atmospheric CO2 pressure and at ambient temperature.
The electrochemical production of enantiomerically pure ring carbonates was achieved by employing chiral epoxides as starting materials in a similar fashion. In this reaction, stainless steel served as the cathode and magnesium as the anode.51 Scheme 28 provides a description of potential mechanisms involved. Sacrificial anodization of magnesium rods resulted in the formation of Mg2+. Simultaneously, CO2 underwent reduction to form anion radicals which were stabilized by Mg2+ ions, leading to the formation of complex A. A subsequent reaction with epoxides involved a nucleophilic attack by bromide anions on the β-C atom, resulting in the formation of compound C. Following an intramolecular nucleophilic attack by oxygen atoms on the epoxide and electron transfer with another molecule of CO2, compound E was formed and the stable complex A was regenerated. Carboxyl oxide anions acted as nucleophilic substitutes for bromine ions within the molecule, ultimately yielding ring carbonate 63. This proposed mechanism elucidates the crucial roles played by Mg2+ and bromine ions.
In 2013, Jiang, Yuan and co-workers reported the electrochemical synthesis of alkylidene lactones 65 from CO2 and 1,4-diarylbutyl-1,3-diacetylene 64 using copper as a catalyst (Scheme 29).52 The reaction proceeded smoothly in an undivided electrochemical cell equipped with a nickel cathode and an aluminum anode, employing an n-Bu4NBr-DMF electrolyte under constant current conditions at a CO2 pressure of 4 MPa. In the presence of a CuI catalyst, the desired alkylidene lactones were obtained in satisfactory yields.
The electrochemical reactions can occur not only through direct cathode reduction but also via anodic oxidation. In 2023, Liang, Pan and co-workers developed a sophisticated electrochemically mediated three-component reaction involving proparamide 66, CO2, and diselenides for the synthesis of selenium-containing oxazolidin-2,4-dione 67 (Scheme 30).53 The oxazolidinone derivatives were obtained through electrolysis using a platinum plate as the cathode material, tetrabutylammonium hexafluorophosphate as the electrolyte, 20 mol% Cu(OAc)2 as the catalyst, K2CO3 as the alkaline additive, and DMSO as the reaction solvent under constant current conditions of 20 mA for 5 hours. The utilization of the precious metal palladium is circumvented by this method, in contrast to the conventional synthesis approach. In the investigation of substrate universality, the authors not only explored the impact of electrical and spatial factors on the yield by incorporating various substituents but also extended the applicability of the substrate by introducing a range of natural products or drug molecular groups (67a–67f). Among them, mosapride, dopamine, tryptamine, and abietic acid derivatives were obtained with yields ranging from 50% to 63%. Aminoester derivatives were obtained with a yield of 75%. Methetine derivatives were obtained in 36% yield.
Scheme 30 Three-component cyclization reaction catalyzed by electrochemistry and copper for the synthesis of oxazolidine-2,4-diones. |
The diphenyldiene can undergo anodization to generate Se ions in this reaction (Scheme 31). Alternatively, the combination of selenides and [Cu] catalysts forms the copper–selenium complex, which enhances the electrophilicity of selenium atoms. Subsequently, the Se species undergoes an electrophilic addition reaction with the carbamate anion intermediate A through path A/B to obtain the intermediate B. Finally, an intramolecular nucleophilic attack takes place to yield the corresponding product 67.
However, the C–Te bond in tellurium is prone to cleavage due to its low bond energy. Moreover, the strong coordination ability of tellurium atoms often hinders transition metal-based catalysts during coupling reactions. These factors have resulted in synthetic methods for preparing organic tellurium molecules lagging far behind those used for sulfur-containing compounds. Therefore, Pan et al. developed a straightforward and efficient method for synthesizing tellurium-containing 2-oxazolidinones 69 from CO2, ditelluride, and propylpargyne 68 (Scheme 32).54 The reaction exhibits a high degree of substrate universality, enabling the synthesis of target products with moderate to excellent yields across diverse functional groups, natural product scaffolds, and drug-like molecular structures.
The bioactivity of these tellurium-containing oxazolidinones was investigated by evaluating the antitumor activity of all products in vitro using the MTT assay. The experimental results demonstrated that the oxazolidinones containing tellurium exhibited superior inhibitory activity against tumor cells compared to 5-fluorouracil (5-FU), which served as a positive control (Table 2). Among them, compound 69f displayed potent cytotoxicity against T-24, MDA-MB-231, MGC-803, and MIA PaCa-2 cells with IC50 values of 0.4 ± 0.7, 0.6 ± 0.9, 0.9 ± 1.1, and 0.5 ± 2.0 μM, respectively. These findings present a novel approach for the synthesis of organotellurium compounds and offer new insights into exploring the biological activity of sulfur-containing organic compounds.
Scheme 33 Electrochemically mediated cyclization-carboxylation of unsaturated haloaryl ethers and CO2. |
The subsequent investigation involved the examination of diverse metallic nickel complexes and various reactants, including epoxides, resulting in the production of cyclic carbonates and other heterocyclic compounds.56 In addition, the use of metal nanoparticle electrodes as cathodes avoids any additional catalysts for the electrosynthesis of ring carbonates from CO2 and epoxides.57
Scheme 34 Electrocatalyzed iodine-mediated carboxylation-cyclization CO2 with olefins to cyclic carbonates. |
In 2016, Yuan et al. successfully synthesized 1,3,4-oxadiazol-2(3H)-1 derivatives 78 in a one-pot reaction using NaI as the medium (Scheme 35).59 This was achieved through a three-component coupling reaction of CO2 with arylhydrazine and paraformaldehyde. Electrolysis was conducted at room temperature in a non-splitting battery setup with a nickel foil cathode and graphite rod anode. By avoiding the use of toxic phosgene or CO commonly employed in traditional methods, this process offers an environmentally efficient route for obtaining 1,3,4-oxadiazol-2(3H)-1 derivatives. The proposed reaction mechanism involves the initial formation of an intermediate product A by reacting phenylhydrazine 76 with paraformaldehyde 77, followed by anodic oxidation of I− ions to form I2 which then reacts to produce the key intermediate N-(alpha-iodoethylene)-N′-phenylhydrazine B. CH3OH is electroreduced at the cathode to generate CH3O− base species while N-(alpha-iodoethylene)-N′-phenylhydrazine B is converted into its corresponding anion form C through electrogenic base CH3O− and additional t-BuOK base treatment. Finally, an intermolecular nucleophilic attack occurs between N-(alpha-iodoethylene)-N′-phenylhydrazine B and CO2 resulting in the formation of product 78 while eliminating I− ions from the system.
Scheme 35 Electrochemical conversion of CO2 with arylhydrazine and paraformaldehyde to oxadiazole derivatives. |
In 2021, an efficient method for the electrocatalytic synthesis of new carbamate compounds based on CO2, N-(2-vinylphenyl)-sulfonamide 79 and amines has been developed by Liang, Tang and co-workers (Scheme 36).60 The reaction conditions were a current density of 25 mA cm−2, a temperature of −10 °C, platinum electrodes as the cathode and anode, NH4I as the electrolyte and CH3CN as the solvent. Researchers tested multiple amines coupled with N-(2-vinylphenyl) toluenesulfonamide and CO2. Various amines, such as pyrrolidine, dialkylamine, primary amine, etc., can produce the desired product with satisfactory yield (80a–80d). It is worth noting that some amino esters (80i, 80j) that can participate in the reaction are also used. The cyclization of N-(2-vinylphenyl) sulfonamide derivatives containing additional sulfonyl groups can obtain high yields (80e). However, the presence of electron-withdrawing groups in substrates with chlorine, bromine, and cyanide substituents on the benzene ring leads to reduced yields (80f–80h). Additionally, the synthesized compounds were subjected to pharmacological investigation. The antitumor activity of these compounds was assessed using the MTT assay. Interestingly, carbamate derivatives exhibited relatively superior antitumor activity compared to both the raw materials and products obtained from the reaction between raw materials and amines. These findings suggest that this novel carbamate demonstrates enhanced potential for anti-tumor effects through CO2 fixation with amines and N-alkenylsulfonamide as substrates.
Under the condition of constant current electrolysis, NH4I serves as both an electrolyte and a catalyst (Scheme 37). Initially, iodide is oxidized to form iodine molecules. With the assistance of iodine, substrate 79 undergoes cyclization to generate intermediate A. Subsequently, intramolecular cyclization of intermediate A leads to the formation of crucial intermediate B.
The reaction between CO2 and amines yields carbamate anion C. Intermediate B reacts with carbamate anion C to yield the final product 80. This reaction process employs paired electrolysis without requiring additional oxidants or alkalis, making it environmentally friendly and pollution-free.
In recent years, the prevailing synthetic approach for 2-oxazolidinone has involved the reaction of β-amino alcohols with various carbonylation reagents. However, these methods require the use of phosgene or CO, and their application is limited due to the toxicity associated with these substances. Pan et al. introduced a method for synthesizing pharmacologically active 2-oxazolidinone 83 by reacting CO2 with allylamine 81 (Scheme 38).61 This methodology effectively avoids the need for toxic reagents while preserving the unsaturated double bond in the resulting 2-oxazolidinone. Consequently, the resulting product exhibits enhanced malleability and facilitates subsequent structural modifications. Furthermore, this methodology demonstrates its practicality as it allows for the synthesis of six-membered heterocyclic rings (1,3-oxazinone-2-ones) and can accommodate low concentrations of CO2 in the reaction. Notably, moderate yields were achieved when utilizing three natural products or drug molecules (83e–83g): dehydroabietylamine derivative, amino ester derivative, and dopamine derivative.
Scheme 38 Electrochemically mediated carboxylative cyclization of allylic/homoallylic amines with CO2. |
The metal-free electrochemical coupling of aromatic hydrocarbons 84 with NH3, CO2 or CS2 to form 2-oxazolidinones or 2-thiazolidinones 86 in the condensation sequence was reported by Vos and co-workers (Scheme 39).62 The reaction proceeds through N–H azopyridine, followed by ring enlargement using CO2 or CS2. This two-step reaction is catalyzed by a simple iodide catalyst that exhibits excellent regional selectivity compared to olefins, resulting in a total yield of up to 91%. The sustainable metal-free approach demonstrates exceptional atomic efficiency, facilitating the direct synthesis of these unprotected heterocycles from simple aromatic alkenes using renewable energy sources and readily available abundant materials.
In 2002, Deng and coworkers reported the synthesis of cyclic carbonates 63 from CO2 and epoxides 62 in pure ionic liquids without the need for additional supporting electrolytes and catalysts (Scheme 40a).65 The experiments were conducted under mild conditions, employing copper as the working electrode and Al or Mg rods as the anode, in an undivided battery setup. The CO2 cycloaddition properties of various epoxides 62 such as propylene oxide, epichlorohydrin, and styrene epoxide were examined. When [BMIm][BF4] was used as the reaction medium with propylene oxide as the substrate, a yield of 92% was achieved. By utilizing inert electrodes in a split cell, Lu et al. successfully achieved the electrochemical reduction of [BMIm][BF4] to produce NHC, while also accomplishing the conversion of CO2 and diol 87 into ring carbonates 88 (Scheme 40b).66
In 2011, Senboku and co-workers developed benzoate-mediated electrochemical reactions for the reductive radical cyclization of aryl halides with olefins followed by tandem carboxylation (Scheme 41).67 Cyclic voltammetry studies showed that 4-tert-butyl benzoate has a reversible peak relative to Ag/Ag+ at −2.9 V, which is higher than that of the aryl bromo substrate (3.2 V), thus supporting the benzoate-mediated substrate reduction pathway. The probable reaction pathway is that 4-tert-butyl benzoate undergoes a one-electron reduction to the radical anion, reducing aryl bromide 72a to the aryl radical A. Upon addition of the intramolecular radical to the pendant olefinic portion of the substrate, the resulting carbon-centered radical B is reduced to carbon C, which is then captured by CO2 to complete the reaction. On the other hand, dissolution of an Mg anode as the magnesium ion proceeds results in the prevention of any species from oxidizing at the anode.
Scheme 41 Electrochemical aryl radical cyclization-carboxylation using methyl 4-tert-butylbenzoate as an electron-transfer mediator. |
Later, Senboku et al. reported constant-current electrolysis of 2-(2-propyl oxygen) bromobenzene 89 in DMF using an undivided cell equipped with a Pt cathode and a Mg anode in the presence of CO2 and the electron transfer medium methyl 4-tert-butylbenzoate (Scheme 42).68 This leads to aryl cyclization with a carbon–carbon triple bond and then immobilizes two molecules of CO2 to obtain 2,2-ring fused succinic acid derivatives 90, in moderate to good yields. Through aryl cyclization, the framework of dihydrobenzofuran, indoline, dihydrobenzothiophene, indenane and tetrahydropyrane was successfully constructed, and succinic acid was successively produced by a unique series carboxylation reaction.
Scheme 42 Electrochemical cyclization/carboxylation of 2-(2-propyl oxygen) bromobenzene aryl by methyl 4-tert-butylbenzoate. |
They also reported that vinyl radicals can also be produced by electrochemical reduction in the presence of CO2 (Scheme 43).69 Continuous radical cyclization and carboxylation of vinyl radicals were carried out successfully, and cyclized unsaturated carboxylic acid 92 was obtained in medium and high yield. Summarizing the data from the authors’ study, we find that methyl 4-tert-butylbenzoate is not a necessary electron transfer mediator in this study.
Inesi et al. reported a method for the electrochemical synthesis of chiral oxazolidin-2-ketones 94 from chiral amino alcohols 93 and CO2 (Scheme 45).71 Oxazolidin-2-ketones were obtained with good yield under mild conditions, without the use of toxic, polluting or harmful chemicals and without the addition of any base or precursor base.
The Inesi team also reported the efficient electrochemical synthesis of 5-methylene-1,3-oxazolidin-2-one 95 (Scheme 46).72 The reaction uses ethylamine 52 and CO2 as raw materials by direct electrolysis of MeCN and Et4NPF6 solutions containing ethylamine, followed by CO2 bubbling and heating. The yield ranges from medium to excellent, the conditions are mild, and the use of toxic and harmful chemicals and catalysts is avoided.
In 2010, Yuan's group reported a convenient and effective electrochemical method for the cyclization reaction of CO2 and terminal propargyl alcohol 96 at room temperature (Scheme 47).73 In an integrated electrolytic cell containing n-Bu4NBr-MeCN electrolyte, propargyl alcohol can be successfully electrosynthesized with CO2 at a constant current under a CO2 pressure of 3 MPa, and α-alkylene cyclic carbonates 97 can be obtained with good to excellent separation yield. The experimental results show that under the synergistic action of acetonitrile and n-Bu4NBr, the Cu+ ions and strong bases generated by electrolysis can effectively catalyze or promote the coupling reaction, and then the cyclization reaction occurs. In the electrolysis process, the tetrabutylammonium ion R4N+ on the surface of the nickel cathode initially acquires electrons to form tributylamine R3N and an alkyl radical. Subsequently, the alkyl radical further gains electrons to generate the alkyl anion R−. Notably, the alkyl anion R− exhibits sufficient basicity to deprotonate propargyl alcohol's hydroxyl group, resulting in the formation of intermediate A. This intermediate A then undergoes a reaction with CO2 to yield carbamate anion intermediate B. Simultaneously, continuous oxidation of a sacrificial copper anode leads to CuI formation which can activate intramolecular ring closure reaction on acetylene leading to production of compound D, subsequently releasing the CuI catalyst and forming the corresponding ring carbonate 97.
Although significant advancements have been achieved in utilizing CO2 as a C1 source in photochemistry/electrochemistry, there still exist certain limitations. (1) Currently, photocatalysts primarily consist of precious metal complexes such as Ir and Ru, along with organic dyes. The catalytic system operates by absorbing visible light to induce electron transition from the ground state to the excited state, followed by a catalytic cycle through SET with the substrate. This intramolecular charge transfer induced by visible light necessitates a large π delocalization structure or conjugation in metal–ligand complexes to create a band gap for absorption in the low energy range of visible light. Consequently, complex molecular structures are required to achieve visible light-excited electron transition, inevitably increasing the cost of photocatalysts. In view of this, many chemists have endeavored to utilize Earth's abundant first-row transition metals74 as cost-effective alternatives to rare elements like Ru and Ir that are scarcely found in the Earth's crust. Alternatively, chemists have explored intermolecular charge transfer for photoredox catalysis and developed non-metallic anion complex photocatalytic systems75 capable of achieving decarboxylation coupling reactions under mild conditions; these systems also hold promise for CO2 synthesis and conversion. (2) The main catalysts employed in photoredox catalysis include Ru, Ir, and other metals or small molecule photocatalysts. However, their homogeneity renders them non-reusable and economically impractical. Consequently, researchers have been exploring more cost-effective and reusable alternatives to these photocatalysts, aiming to develop protocols that can be readily implemented in the industrial sector. In this context, scientists have proposed various nanomaterials as sustainable and economical substitutes. For instance, [Ru(bpy)3]2+ can be incorporated into metal–organic frameworks (MOFs) to facilitate the separation and reusability of precious metals. Additionally, in electrochemical reactions, scientists fabricate metal catalysts into nanoelectrodes which exhibit high catalytic efficiency while maintaining reusability without any loss of activity. Furthermore, by employing ionic liquids as reaction media for CO2 electrocatalytic cycloaddition reactions, smooth operation can be achieved without the need for additional supporting electrolytes or catalysts. (3) In the field of electrocatalytic reactions, the development of novel electrocatalysts to enhance chemical conversion efficiency and accuracy remains a prominent research topic in organic chemistry. Due to the fact that most electrochemical transformations solely occur at either the anode or cathode, the complementary half-reaction at the other electrode is often disregarded or wasted within an electrochemical system. For instance, during an electrochemical reduction reaction, a sacrificial anode is typically employed. Consequently, exploring paired electrolytic reactions utilizing both cathodes and anodes has emerged as another focal point in electrochemical synthesis. (4) From the perspective of green chemistry, asymmetric photoelectrochemical synthesis offers a novel approach for the construction of chiral compounds. Despite the development of numerous sophisticated transformations in asymmetric photoelectrochemical synthesis by chemists, there remains a dearth of reactions utilizing small molecules like CO2 for performing high-value asymmetric photoelectrochemical transformations. The exploration and advancement in this field pose significant and formidable challenges. (5) The photocatalytic and electrocatalytic pathways provide a cost-effective means of converting CO2 into valuable organic compounds. However, practical applications necessitate the development of an efficient and durable reusable catalyst capable of withstanding impurities in CO2 feed during industrial processes, while maintaining high activity and selectivity. Therefore, scientists continue to strive towards designing efficient pathways for photocatalytic and electrocatalytic cycloaddition reactions suitable for industrial applications.
This journal is © The Royal Society of Chemistry 2024 |