Azar Rostampoor and
Abdolali Alizadeh*
Department of Chemistry, Tarbiat Modares University, P. O. Box 14115-175, Tehran, Iran. E-mail: aalizadeh@modares.ac.ir; abdol_alizad@yahoo.com
First published on 6th May 2025
This review discusses the diverse applications of β-ketonitriles in organic synthesis, highlighting protocols that generate various compounds, including cyclic hydrocarbons, aromatic compounds, heterocycles, spirocycles, and fused heterocycles. These compounds serve as valuable building blocks for biologically active scaffolds like chromenes, quinolines, and natural products. It provides an overview of 69 research articles published from 2014 to 2023, focusing on reactions involving benzoyl acetonitrile derivatives and other β-ketonitriles. The methodologies include cascade, domino, and sequential reactions facilitated by different catalysts, showcasing the versatility of β-ketonitriles in organic chemistry.
These compounds are used as precursors for the production of anti-cancer,2 anti-inflammatory,3 and antimalarial drugs,4 and anti-HIV agents.5 Additionally, β-ketonitriles have been studied extensively in medicinal chemistry due to their potential therapeutic applications. For example, some derivatives have shown promising activity against cancer cells, while others have been studied for their potential as anti-inflammatory agents or as inhibitors of enzymes involved in various disease pathways.6 Ongoing research in this area may lead to the development of new drugs with improved efficacy and reduced side effects. Various methods have been presented for synthesizing β-ketonitriles.7 Kiyokawa et al. completely summarized the synthesis of these valuable compounds in a review.8 One well-known β-ketonitrile is benzoylacetonitrile, utilized as a starting material for synthesizing various heterocyclic compounds. It can undergo different reactions to produce various types of heterocycles such as pyridines,9 pyrimidines,10 and pyrazoles.11 One common reaction is the condensation of benzoylacetonitrile with aldehydes or ketones to form pyridine. This reaction is known as the Hantzsch synthesis and involves the formation of a dihydropyridine intermediate, which is then oxidized to the pyridine product.9 Benzoylacetonitrile can also undergo cyclization reactions with different reagents to form various heterocyclic compounds. For example, a reaction with hydrazine derivatives can lead to the formation of pyrazoles,11 while a reaction with urea or thiourea can lead to the formation of pyrimidines.10 In addition, benzoylacetonitrile can be used as a starting material for the synthesis of benzonitriles.12 Overall, benzoylacetonitrile is a versatile starting material for synthesizing a wide range of heterocyclic compounds with potential applications in pharmaceuticals and agrochemicals.13 In 2013, Bakr F. and colleagues successfully synthesized various heterocycles such as pyran, pyridazine, pyrimidine, pyrazine, and triazine compounds using benzoylacetonitrile as a precursor.14 They conducted a review of the use of benzoyl acetonitrile from 1985 to 2013. In the last decade, many efforts have been made in this field, leading to numerous interesting results. Benzoylacetonitrile was used as a precursor in several multicomponent and one-pot reactions that led to the construction of functionalized cyclic molecules. In addition to some common carbocyclic and heterocyclic compounds, other unusual compounds, such as propellanes, carbazoles, quinoxalines, benzofuropyrroles, N-fused bicyclic systems, and other polycyclic heterocycles, were also reported. To provide a better understanding of the topic, this review is mainly organized according to different product structures.
The following year, Cheng Guo et al. reported a new method for esterification that does not require transition metals or photocatalysts. This process involves using visible light to facilitate the decyanation of aryl ketonitriles 1 in the presence of dioxygen and alcohols 4 at room temperature, forming α-ketoesters 5 (53–93% yield). This reaction enables C–H bond functionalization, C–C σ-bond cleavage, and dioxygen activation to be achieved in a single step, cleavage, and dioxygen activation to be achieved in a single step (Scheme 2).17
In a different study, Hocaoglu and colleagues utilized the radical addition of β-ketonitriles 1 to conjugated dienes 20 to produce 5-ethenyl-4,5-dihydrofuran-3-carbonitriles 21 in yields ranging from low to good. In this method, Mn(OAc)3 and CAN were used as radical oxidants in the reactions. Based on these results, the authors concluded that Mn(OAc)3 is effective in reactions of 3-oxopropanenitriles with dienes that contain a thiophen-2-yl group. Conversely, CAN is more efficient with dienes that do not contain thiophen-2-yl (Scheme 9).24 Wang and their team have reported on a solvent/base-switchable platform that allows for the selective conversion of β-ketonitriles 1 to multi-substituted 2,3-dihydrofuran-2-carbonitriles 22 and 4,5-dihydrofuran-3-carbonitriles 23 without the use of metals. By simply changing the solvent and reaction base, the same oxidant (TBHP) can produce the desired products through two distinct pathways. Through mechanistic exploration, this method has shown that the reaction of β-ketonitriles under non-polar and polar solvents leads to unexpected rearrangement of the –CN group migration and hydroxylation, respectively, resulting in different products of hydrofurans (Scheme 10).25
![]() | ||
Scheme 10 Synthesis of multi-substituted 2,3-dihydrofuran-2-carbonitriles 22 and 4,5-dihydrofuran-3-carbonitriles 23. |
In 2016, Feng and coworkers obtained regio-diastereo and enantioselective 2,3-dihydro-furans 25 in 93–92% yield from the reaction of substituted β-ketonitriles 1 with bromonitrostyrenes 24 in toluene solvent at 60 °C for 72 h by using achiral N,N′-dioxide as organocatalyst. In the mechanical exploration of this reaction, bromonitrostyrenes are first added to substituted β-ketonitrile from Michael's process (C–C bond formation). Then, by adjusting the reaction conditions, the O-alkylation path is preferred to the C-alkylation path. With this strategy, the desired chiral 2,3-dihydrofurans 25 were obtained in up to 95% yield with 95:
5 dr and 93% ee (Scheme 11).26
In a different research, Weng et al. discovered that by using two different bases, Na2CO3 and NaOAc, in the reaction of β-ketonitriles 1 with 3-bromo-1,1,1-trifluoroacetone 27, it becomes possible to selectively synthesize two trifluoromethylated furans 28 and dihydrofuranols 29 without using any metals (Scheme 12).27
Zhang and colleagues have developed a new method for synthesizing chiral bicyclic dihydrofurans 31 with two vicinal carbon stereocenters. The method involves a Pd-catalyzed asymmetric allylic substitution cascade reaction of allylic dicarbonates 30 with β-ketonitriles 1. The products are obtained in high yields and with up to 97% ee (Scheme 13).28
The following year, Xinwei He et al., synthesized a series of polysubstituted furans 33 by the reactions of propargylamines 32 with β-ketonitriles 1 in the presence of Na2CO3 under heating at 90 °C for 24 h. Further exploration of the mechanism of this reaction indicates that it proceeds through a sequence of steps. First, there is a 1,4-conjugate addition of β-ketonitriles 1 to propargylamines 32, followed by a 5-exo-dig annulation/isomerization to form the fully substituted furans 33 (Scheme 14).29
In 2019, Jeong and coworkers synthesized highly substituted furan-linked biheteroaryls 36 by a three-component cascade reaction of aryl glyoxals 34, isonitriles 35, and aroyl or heteroaryl β-ketonitriles 1 in the presence of p-TsOH and EtOH under reflux condition for 4 h. The reaction sequence involves a Knoevenagel condensation of aryl glyoxals 34 with β-ketonitriles 1 followed by an isocyanide 35 insertion via formal [4 + 1] cycloaddition, and then a rapid [1,3]-H shift to afford novel bi-heterocycles with unique decorations (Scheme 15).30
In 2020, Zhong and colleagues achieved moderate to high yields of tetrasubstituted furan derivatives 38 through [3 + 2] annulation of ethynyl benzoxazinanones 37 with β-ketonitriles 1. This was done in the presence of Cu(acac)2 as a catalyst and i-Pr2NEt in CH2Cl2 at room temperature over 24 hours. It was found that intermediate copper allenylidines acted as a C2 synthon during the cycloaddition reaction (Scheme 16).31
In 2021, Wan et al. reported the synthesis of substituted-furo[2,3-b]furans 40 by oxidizing malononitrile 39 to 2-hydroxy-malononitrile I, which can then react with β-ketonitriles 1 in CH3CN in the presence of 0.5 equiv. of benzoic acid at room temperature for 12 h. In the mechanistic exploration of this reaction, it was hypothesized that SeO2 oxidizes malononitrile to 2-hydroxy-malononitrile, which can then react with β-ketonitriles 1 and during the two processes of Michael addition and intramolecular cyclization, it provides target compounds with 27–56% efficiency. The authors have highlighted that the synthesized compounds exhibit interesting photoluminescence properties in both solution and solid-state (Scheme 17).32
In a recent study, Gnanasambandam and his colleagues synthesized heterocyclic furo(2,3-b)furan derivatives 43 with an efficiency rate of 60–80%. They achieved this through a three-component reaction that involved methylglyoxal 41, benzothiazole acetonitrile 42, and β-ketonitriles 1, with the use of DABCO (1,4-diazabicyclo[2.2.2]octane) as a base and EtOH/H2O solvent mixture at room temperature for two hours. The reaction involved Knoevenagel condensation, Michael addition, intramolecular cyclization, and tautomerization processes (Scheme 18).33
Followed by Jeong et al. have disclosed a highly efficient one-step protocol for constructing highly functionalized dihydrofuro[2,3-b]furans 44 and substituted phenyl furan 3-hydroxy-3-phenylacrylamides 45. This is achieved through a three-component cascade reaction between aromatic or aliphatic glyoxals 41, β-ketonitriles 1, and two different equivalent amounts of triethylamine. The reaction can be performed in two different conditions: at room temperature or under reflux, with excellent yields obtained in both cases. The protocol involves a Knoevenagel and Michael adduct via Paal–Knorr cyclization with aromatic or aliphatic glyoxal and β-ketonitriles under mild heating conditions. The final product can be easily obtained through a simple filtration method (Scheme 19).34
![]() | ||
Scheme 19 Synthesis of functionalized dihydrofuro[2,3-b]furans 44 and substituted phenyl furan 3-hydroxy-3-phenylacrylamides 45. |
Also, Fang and their team have successfully developed a novel method to create a large variety of cyclic molecules, such as dihydrofurofurans 47 dihydrocyclopentafuranols 48 and, by performing regioselectivity-switchable reactions between alkynyl α-diketones 46 and β-ketonitriles 1. This approach provides a platform for generating a broad range of structurally diverse heterocycles with new dihydrofuran-cyclopentenone skeletons and dihydrofurofurans.
The study also revealed that base and solvent play a critical role in modulating the regioselectivity of this reaction (Scheme 20).35
In a separate study, Liu and colleagues used NIS to accomplish a quasi-three-component reaction of β-ketonitriles 1 and aryl sulfonyl hydrazides 56 in ethanol solvent under reflux conditions, yielding 3-aryl-4-(arylthio)-1H-pyrazol-5-amine derivatives 59 and 3-phenyl-1-(phenylsulfonyl)-4-(phenylthio)-1H-pyrazol-5-amines 60. In the mechanistic investigation of this method, it was discovered that the process involves a sequence of cyclization, sulfonylation, and removal of aryl sulfonyl in the presence of NIS (Scheme 23).39
![]() | ||
Scheme 23 Synthesis of 3-aryl-4-(arylthio)-1H-pyrazol-5-amine derivatives 59 and 3-phenyl-1-(phenylsulfonyl)-4-(phenylthio)-1H-pyrazol-5-amines 60. |
Kachanov et al. produced 1,3-oxaselenole heterocycles 63 and 64 in good yield with a cyano group using aroyl acetonitrile 1 and selenium(IV) oxide. The resulting products were found to react with ammonia, hydrazine, or primary amines, during which an aryl rearrangement was observed (Scheme 25).41
In another study, a series of 2-oxazolines (67 and 68) were produced using a simple one-pot method under inert and non-moisture conditions from β-ketonitrile 1 and β-amino alcohols 65 and 66 with 115–172 mol% ZnCl2 (Scheme 26).42
An effective oxidative system, which does not require the use of metals, has been developed using TBHP and AIBN. This system has allowed for the successful synthesis of substituted 2-aminothioazoles 70 by reacting β-ketonitrile 1 with thiourea 69. Mechanistic studies have shown that the reaction proceeds through the formation of a C–S bond via a radical process, followed by the formation of a C–N bond through an intramolecular condensation reaction (Scheme 27).43
Krasavin et al. have prepared a set of isoxazole-5-amines 71 by reacting readily available β-ketonitriles 1 with hydroxylamine in 15% aqueous NaOH solution at reflux for 14 hours (Scheme 28).44
The desired compounds were acquired with exceptional yields (up to 91%) and enantioselectivities (up to 98% ee) (Scheme 29).45
In 2016, Tong et al. found that by using 6′-deoxy-6′-[(L)-N,N-(2,2′-oxydiethyl)-valine amido]quinine 77 as the catalyst, the formation of 4H-pyrans 78 through (3 + 3) annulations of β′-acetoxy allenoates 76 with β-ketonitriles 1 can occur quickly and with excellent enantioselectivity. Catalyst 77 has three functions, including Lewis base (quinuclidine N), H-bond donor (amide NH), and Brønsted base (morpholine N), each playing crucial roles in the chemo- and enantio-selectivity for the construction of 4H-pyrans 78 (Scheme 30).46
![]() | ||
Scheme 30 Synthesis of 4H-pyran derivatives 78 by using 6′-deoxy-6′-[(L)-N,N-(2,2′-oxydiethyl)-valine amido]quinine 77 as a catalyst. |
The following year, Zhou et al. discovered that using 6′-deoxy-6′ perfluorobenzamido-quinine 80 as a catalyst can result in the quick formation of 4H-pyrans 81 through [3 + 3] annulations of δ-acetoxy allenoates 79 and β-ketonitriles 1 with excellent enantioselectivity. The researchers found that the amide NH of 80 plays a crucial role as an H-bond donor, facilitating the formation of cationic intermediate I and increasing the electrophilicity of its δ-position (Scheme 31).47
![]() | ||
Scheme 31 Synthesis of 4H-pyrans 81 by using 6′-deoxy-6′ perfluorobenzamido-quinine 80 as a catalyst. |
In 2022, Zhang et al. realized that using Cu(OAc)2 as the catalyst can result in the quick formation of polysubstituted 4H-pyran derivatives 84 with a quaternary CF3-containing center through (3 + 3) annulations of alkynyl ketimines 82 and β-ketonitriles 1 with excellent yields (86–99%) and good enantioselectivities (71–78% ee). The researchers found that the reaction involves a process with a base-catalyzed or chiral thiourea-catalyzed Mannich-type reaction followed by a highly regioselective copper-catalyzed ring-closing reaction on the alkynyl moiety in a 6-endo-dig fashion (Scheme 32).48
In 2019, the same group synthesized derivatives of 2-amino-4H-pyran-3,5-dicarbonitrile 85 through a three-component reaction. This reaction involved aldehydes 9, malononitrile 39, and β-ketonitriles 1 and was carried out in a mixture of ethanol and water at room temperature. The resulting products had a wide range of functional groups and were obtained in high yields (Scheme 33).49
Bharkavi et al. reported a series of new compounds called 2,6-diaryl-4-(1H-indol-3-yl)-3-cyanopyridines 91. These compounds were produced in good yields through the domino reactions of β-ketonitriles 1, 4,4,4-trifluoro-1-phenylbutane-1,3-dione 90, and aromatic aldehydes 9 in the presence of ammonium acetate without the use of solvents. This process creates two C–C and two C–N bonds, leading to the formation of a six-membered ring in a single operation (Scheme 35).51
In the mechanistic study of this reaction, it is explained that the hydroxyl group of chitosan activates the carbonyl group of aldehyde 9 at first, which then undergoes condensation with guanidine 92 leading to the formation of intermediate I. The amine group present on the surface of chitosan facilitates the formation of III. The final step involves the addition of intermediates I and III to form a new intermediate IV, which undergoes intramolecular cyclization followed by aromatization V to produce the desired product 93 in good yield (Scheme 36).52
In 2016, Dhiman used copper as a catalyst for the three-component domino reactions of 2-bromobenzaldehydes 100, active methylene nitriles (1 and 101), and sodium azide 102. The reactions were performed in DMSO under reflux at 150 °C in the presence of L-proline and K2CO3 leading to the synthesis of 2-aminoquinolines 103 and 2-arylquinoline-3 carbonitriles 104. Mechanistic exploration of this reaction shows that the formation of substituted quinolines involves Knoevenagel condensation of ortho-bromobenzaldehyde with active methylene nitriles followed by copper-catalyzed reductive amination and intramolecular cyclization (Scheme 39).55
Recently, our group reported an efficient synthesis of functionalized pyrano[2,3-b]quinoline and benzo[h]pyrano[2,3-b]quinoline derivatives 107a and 107b by using 2-chloroquinoline-3-carbaldehyde or 2-chlorobenzo[h]quinoline-3-carbaldehyd 105a and 105b, 1-aryl-2-(1,1,1-triphenyl-λ5-phosphanylidene)ethan-1-one (Wittig reagent) 106 and β-ketonitriles 1 in EtOH under reflux at 80 °C in the presence of Et3N in good to excellent yields. The mechanical of this reaction involves the two processes of C–C bond formation (Michael addition) and intramolecular cyclization (by attacking the oxygen atom of active methylene compounds) (Scheme 40).56
![]() | ||
Scheme 40 Synthesis of functionalized pyrano[2,3-b]quinoline and benzo[h]pyrano[2,3-b]quinoline derivatives. |
Hussain et al. have developed an efficient method to produce 3-cyano-4-quinolone derivatives 109. This is achieved through decarboxylative cyclization of isatoic anhydrides 108 with β-ketonitriles 1 using DABCO as a non-toxic, eco-friendly, and inexpensive reagent under microwave conditions. The reaction is performed in CH3CN under reflux at 80 °C for 30 minutes. This approach provides an easy pathway for making this class of compounds in good to high yield from readily available starting materials in short reaction times (Scheme 41).57
Yang et al. conducted a study on synthesizing 3-trifluoromethyl-isoquinolines 111, which are important compounds in biology, using an Rh(III)-catalyzed annulation of β-ketonitriles 1 and CF3-substituted imidoyl sulfonium ylides (TFISYs) 110. The reaction was conducted in THF solvent with NaOAc at 60 °C for 24 hours. The transformation involved removing a C–H bond and an unusual C–C bond from benzoyl acetonitrile, while also removing a molecule of dimethyl sulfoxide (DMSO) and acetonitrile (CH3CN). This reaction has several advantages, including low catalyst loadings, mild conditions, wide substrate scope, high efficiency, and scalability (Scheme 42).58
In 2018, He et al. developed a concise and efficient method for synthesizing functionalized 2-aryl-4H-chromenes 116 using propargylamines 115 and β-ketonitriles 1 in the presence of FeCl3 as a catalyst. This method is environmentally friendly and utilizes CH3CN solvent for 24 hours under reflux conditions. The reaction features a highly efficient tandem sequence of 1,4-conjugate addition, 6-endo-dig cyclization, and oxidation. This protocol accommodates various functional groups, rendering it a practical and effective method for synthesizing 2-aryl-4H-chromene skeletons 116 (Scheme 44).60
The Pan group recently developed an organocatalytic method for the synthesis of chiral 3,4-dihydrocoumarins 119 and tetrasubstituted chromas 120. This process involves Michael addition and intramolecular cyclization of β-ketonitriles 1 and 2-sulfonyl methyl phenols 117. The reaction takes place in dichloromethane with sodium bicarbonate at room temperature. In the reaction mechanism, the amino group of catalyst 118 activates the carbonyl groups and acts as a Lewis acid to facilitate the reaction (Scheme 45).61
Jiang and his team conducted research on the domino reaction of two molecules of β-ketonitriles 1 and one molecule of 2-aryl-3-nitro chromene 121, in the presence of triethylamine. They performed this reaction in tetrahydrofuran solvent to synthesize various derivatives of dihydrofuro[2,3-c]chromenes 122, and by using polar protic solvents such as ethanol and methanol, to synthesize chromeno[3,4-b] substituted pyridines 123. These products were obtained under mild conditions. One of the most significant advantages of this process is the wide range of products produced with satisfactory efficiency (Scheme 46).62
![]() | ||
Scheme 46 Synthesis of dihydrofuro[2,3-c]chromene 122 and chromeno[3,4-b] substituted pyridines 123. |
Also, our research group a general and efficient method for the chemoselective synthesis of benzo[c]chromen-6-ones 125 has been developed. The reactions were accomplished in the presence of Et3N in EtOH under reflux conditions to afford functionalized benzo[c]chromen-6-ones 125 in 70–91% yield. The mechanism of this reaction includes a base-promoted nucleophilic substitution/deprotonation/intramolecular aldol condensation/carboxylic acid or alkyl hydrogen carbonate elimination/aromatization reaction of β-ketonitriles 1 and α,β-unsaturated coumarins 124 (Scheme 47).63
Bhuyan and colleagues have developed an effective and versatile method for the selective synthesis of tetrazole-fused pyrido[3,2-c]coumarin derivatives 127.
This is achieved through a one-pot three-component reaction of 4-chloro-3-formylcoumarin 126 via intramolecular 1,3-cycloaddition reaction of azides to β-ketonitriles 1. The reactions were carried out in DMF under mild conditions with the addition of one drop of Et3N. The resulting products were obtained with yields ranging between 75-86% (Scheme 48).64
In another study Alizadeh and Rostampoor developed a method for synthesizing dihydro-6H-chromeno[4,3-d]pyrazolo[1,5-a]pyrimidin-6-ones 128 through a three-component reaction involving α,β-unsaturated coumarins 124, β-ketonitriles 1, and hydrazine hydrate 52. This process includes 1,4-addition and aza-Michael addition, forming two carbon–nitrogen (C–N) bonds. The reactions were accomplished in the presence of Et3N in EtOH under reflux conditions to afford products up to 77–92% yield (Scheme 49).65
Choudhury and his team developed a concise method using molecular iodine to synthesize dihydrochromeno[4,3-b]pyrazolo[4,3-e]pyridin-6(7H)-ones 130 through a four-component reaction involving aromatic aldehydes 9, 4-hydroxycoumarin 129, β-ketonitriles 1, and hydrazine hydrate 52 in EtOH under reflux conditions. The synthesis process involves Knoevenagel condensation, Michael addition, and intramolecular condensation (Scheme 50).66
Wu and colleagues developed a method using just 2 mol% of a chiral organocatalyst 134, producing chiral spiro[4H-pyran-oxindole] derivatives 135 with 97–99% yields and 76–97% enantioselectivities. This is achieved through a one-pot reaction of β-ketonitriles 1 and isatylidene malononitriles 133 in CH2Cl2 at −10 °C and adding 1 mol% morpholine. The addition of 2% mol of tertiary amine catalyst 134, which contains many hydrogen bond donors, activates the carbonyl β-ketonitrile 1 group. This activation promotes the Michael addition process, resulting in the formation of spirooxindole through intramolecular cyclization (Scheme 52).68
Jeong and his team developed a simple and concise method catalyzed by DBU for synthesizing new substituted heterocyclic[4.3.3]propellane 141. This was done through a one-pot reaction of a-diketones 140 with variously substituted β-ketonitriles 1 in MetOH under reflux conditions for 8–12 h. The synthesis process involves Knoevenagel, Michael, and intramolecular/Paal–Knorr cyclization (Scheme 54).70
In 2022, the same group synthesized fused (epoxyetheno)indeno-furans 143 by reacting acenaphthoquinolidene 142 and β-ketonitriles 1. The reaction was carried out in the presence of 30 mol% morpholine in DCM under mild conditions for 3 hours. The reaction mechanism is similar to Scheme 54 and involves three processes: Knoevenagel, Michael, and intramolecular/Paal–Knorr cyclization (Scheme 55).71
![]() | ||
Scheme 56 Synthesis of functionalized pyrazolo[1,5-a]pyrimidine-4-iome sulfonates 144 and pyrazolo[1,5-a]pyrimidines 145. |
In 2021, Kim and his colleagues developed a reaction between β-ketonitriles 1 and N-substituted pyrrole-2-carboxaldehyde 146 in CH3CN, in the presence of piperidinium acetate, which allows selective access to 5-acylindolizine-7-carbonitrile 147 through a Knoevenagel condensation–intramolecular aldol cyclization sequence (Scheme 57).73
In a recent study, Khosropour and his colleagues developed a novel catalyst named 3-(propylthio)propane-1-sulfonic acid-functionalized MCM-41 (PTPSA@MCM-41) 148 which was found to be effective and reusable. This catalyst was used for the multicomponent synthesis of dihydro-1H-pyrazolo-[3,4-b]pyridines 149 and 1H-pyrazolo[3,4-b]pyridines 150 through the reaction of aldehydes 9, β-ketonitriles 1, and 1H-pyrazol-5-amines 53 at 80 °C under solvent-free conditions. One significant advantage of this catalytic system is its high catalytic activity and reusability, mild reaction conditions, simple operation, and benign environmental impact (Scheme 58).74
A selective protocol has been developed for the synthesis of quinoxaline derivatives 152 under the same reaction conditions simply from the reaction of phenylenediamine 151 with various derivatives of β-ketonitriles 1 in the presence of visible light in THF solvent and room temperature. This is a novel protocol that accommodates various β-ketonitriles 1 and binary aromatic amines via visible light-induced electron transfer and oxidative coupling. This metal-free method works at room temperature with various substrates and does not require extra oxidants. It typically gives moderate to good yields (Scheme 59).75
Singh and his team synthesized a series of pyrazolo[1,5-c]quinazolines 159 that function as inhibitors for EGFR. They used a highly efficient multicomponent route involving a palladium-catalyzed four-component one-pot tandem reaction.
In this one-pot process, they first formed azomethine 158 by reacting azidobenzaldehyde derivatives 155, isocyanides 156, and arylsulfonylhydrazides 157 in the presence of Pd(OAc)2 (7.5 mol%) in the toluene solvent. Then, they added β-ketonitriles 1 to the reaction mixture in the presence of DABCO, which resulted in pyrazolo[1,5-c]quinazoline formation. The target compounds were screened against MDA-MB-231, A549, and H1299 cancer cell lines (Scheme 61).77
In 2017, Rong and co-workers developed a simple and efficient method for the synthesis of pyrimido[1,2-b]indazol-3-carbonitrile derivatives 161 from aromatic aldehydes 9, 1H-indazol-3-amine (4-chloro-1H-indazol-3-amine) 160 and β-ketonitriles 1 (3-(1H-indol-3-yl)-3-oxopropanenitrile or 3-oxo-3-arylpro-panenitrile) under metal-free conditions. This was a very successful technique for making pyrimido[1,2-b]indazole compounds 161 using only ethanol and triethylamine in a typical laboratory setting. This process has many advantages, including simple operation, high efficiency, easy isolation, and a wide range of substrates (Scheme 62).78
In another study, Jeong et al. have developed a one-pot coupling method that involves three components: 2-hydroxybenzaldehydes 168, β-ketonitriles 1, and isonitriles 156. This method results in the construction of a new tricyclic 2-phenyl-1H-benzofuro[2,3-b]pyrrole ring 169. The reaction sequence begins with a Knoevenagel condensation of 2-hydroxybenzaldehydes 168 with β-ketonitriles 1, followed by the nucleophilic addition of the divalent isocyanic carbon 156. This reaction produces a reactive nitrilium carbon that can be easily trapped by a nearby phenolic group of 2-hydroxybenzaldehydes 168, yielding diverse benzofuro[2,3-b]pyrroles 169 in a single step (Scheme 65).81
According to Wang et al., the use of an iridium catalyst facilitates the cascade annulation reactions of β-ketonitriles 1 with diazo compounds 170 and 171, leading to the formation of substituted naphtho[1,8-bc]pyrans 172 and 173. The reactions involve sequential cleavage of C(sp2)–H/C(sp3)–H and C(sp2)–H/O–H bonds, resulting in the production of different types of naphtho[1,8-bc]pyrans 172 and 173 depending on whether cyclic or open-chain diazo compounds are used. The researchers found that the reactions yield most products in moderate to good amounts and work well with a wide range of substrates (Scheme 66).82
Continuing research in this field, Zhang and colleagues have developed an efficient approach for synthesizing functionalized naphtho[1,8-bc]pyrans 172 through Rh(III)-catalyzed cascade reactions of β-ketonitriles 1 with cyclic 2-diazo-1,3-dicarbonyl compounds 170. The formation of the title compounds involves a cascade process that goes through two steps. The process begins with the cleavage of C(sp2)AH/C(sp3)AH bonds, followed by metalation and carbenoid insertion of 170 with I, resulting in intramolecular annulation to yield substituted 1-naphthol V as a key intermediate. In the second step, the in situ formed 1-naphthol intermediate V undergoes C(sp2)AH/OAH bonds cleavage, metalation, carbenoid insertion with 170, and an intramolecular cyclization to give the naphtho[1,8-bc]pyran product. According to the authors, this new method for synthesizing naphtho[1,8-bc]pyran derivatives 172 offers several advantages over previously reported methods, including a simple operating method, readily available substrates, high efficiency, and excellent atom economy (Scheme 67).83
In a recent report, Wang and colleagues demonstrated the successful application of rhodium(III) catalysis in the synthesis of naphthols 175 and 2,3-dihydronaphtho[1,8-bc]pyrans 176.
The process is involved the cascade activation of β-ketonitriles 1 and annulation with sulfoxonium ylides 174. This study shows that this method is very efficient, selective, and versatile, and a wide range of β-ketonitrile 1 and sulfoxonium ylide 174 have been successfully employed. The redox-neutral conditions and broad substrate scope make this approach suitable for the synthesis of complex structures that are otherwise challenging to access (Scheme 68).84
Bai et al. have reported the development of three-component reactions that involve 4-hydroxyindole 177, aldehydes 9, and β-ketonitriles 1 for the synthesis of 3,4-fused tricyclic indoles 178. These reactions utilize either potassium fluoride or diethylamine as a catalyst, offering straightforward access to 3,4-fused tricyclic indoles with good to excellent yields (Scheme 69).85
The dual reactivity of β-ketonitriles, serving as both electrophiles from the CN moiety and nucleophiles from hydroxyl and CH groups, enables the design of novel tandem reactions. Their high reactivity and ability to coordinate with catalysts through hydrogen bonding and acid-base interactions make them valuable substrates in synthetic organic chemistry, particularly for synthesizing chiral scaffolds.
This journal is © The Royal Society of Chemistry 2025 |