首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
In this paper, the base-mediated cascade reactions of 4-chloro-3-vinyl coumarins with β-ketodinitriles were demonstrated, allowing the efficient synthesis of coumarin-based cyclopenta[c]pyran-7-carbonitriles with interesting chemoselectivity. These transformations include the domino-style formation of C–C/C–C/C–O bonds through a base-mediated nucleophilic substitution, Michael addition, tautomerization, O-cyclization, elimination, and aromatization. The presented synthetic strategy has many advantages such as simple and readily available starting materials, green solvent, highly chemoselective route, synthetically useful yields, and easy purification of products by washing them with EtOH (96%), described as GAP (Group-Assistant-Purification) chemistry.

In this paper, the base-mediated cascade reactions of 4-chloro-3-vinyl coumarins with β-ketodinitriles were demonstrated, allowing the efficient synthesis of coumarin-based cyclopenta[c]pyran-7-carbonitriles with interesting chemoselectivity.  相似文献   

2.
An efficient [4 + 2] benzannulation of α-cyano-β-methylenones and α,β-unsaturated aldehydes was achieved under metal-free reaction conditions selectively delivering a wide range of polyfunctional benzenes in high yields respectively (up to 94% yield).

An efficient [4 + 2] benzannulation of α-cyano-β-methylenones and α,β-unsaturated aldehydes was achieved under metal-free reaction conditions selectively delivering a wide range of polyfunctional benzenes in high yields respectively (up to 94% yield).

Multi-substituted benzenes are privileged structural units ubiquitous in pharmaceuticals,1 natural products2 and advanced functional materials.3 Various excellent methodologies have been investigated for the construction of functionalized aromatics including nucleophilic or electrophilic substitution,4 transition metal-catalyzed coupling reactions5 and directed metalation.6 However, the widespread application of these strategies established thus far suffer from the limitations of functional groups introduced on the pre-existing benzene and regioselectivity issues. Among various synthetic methods, tandem benzannulation reactions arguably represent an attractive alternative to classical methods for rapid construction of polysubstituted benzenes in an atom-economical fashion.7 This protocol featuring an efficient transformation of acyclic building blocks into structurally valuable benzene skeletons. In this context, α-cyano-β-methylenones has been employed as substrates to format six-membered ring in tandem cyclization reactions due to the activation of the pronucleophile methyl group. In 2015, Tong and co-workers developed a phosphine-catalyzed addition/cycloaddition domino reactions of β′-acetoxy allenoate with 2-acyl-3-methyl-acrylonitriles to give 2-oxabicyclo[3.3.1]nonanes (Scheme 1a).8 Soon after that, the construction of benzonitrile derivatives and 1,3,5-trisubstituted benzenes via N-heterocyclic carbene catalysis has been reported by the groups of Wang and Ye independently (Scheme 1b).9 Then the synthesis of 1,3,5-trisubstituted benzenes by 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU)-mediated annulation of α-cyano-β-methylenones and α,β-unsaturated carboxylic acids was also developed by Ye and co-workers (Scheme 1c).10 Shi et al. reported a base-promoted tandem cyclization reaction of α-cyano-β-methylenones and α,β-unsaturated enones, which have electron-withdrawing group (EWG), accessing to a wide range of benzonitriles in a different C–C bond formation process (Scheme 1d).11 As part of our ongoing interest in harnessing enones for developing new methodologies for the construction of functionalized benzenes, we have recently demonstrated NHC-catalyzed convenient benzonitrile assembly in the presence of oxidant.9a While the same reaction of enals and α-cyano-β-methylenones was conducted in the basic condition without NHC, a novel polyfunctionalized benzene product was obtained (Scheme 1e). The result inspired us to extend the synthetic potential of benzannulation strategy to access diverse benzonitriles, particularly from simpler, abundantly available starting materials.Open in a separate windowScheme 1α-Cyano-β-methylenones in cycloaddition domino reactions.At the outset, model reaction of 2-benzoyl-3-phenylbut-2-enenitrile 1a and cinnamaldehyde 2a was used to evaluate reaction parameters. Key results of condition optimization are summarized in 12 The configuration of products were assigned unambiguously by X-ray analysis of the product 3a. A quick solvent screening demonstrated that chloroform is the best choice to produce the benzannulation product 3a in a desirable yield (entries 10–13, ). Reducing the loading of the cinnamaldehyde or NaOH to 1.2 equivalence led to dramatical loss of the yield (entries 14 &15,
EntryBaseSolventTime (h)Yieldb (%)
1Cs2CO3Toluene2470
2Na2CO3Toluene2442
3K2CO3Toluene2438
4NaOHToluene1278
5NaOAcToluene2452
6KOHToluene1274
7K3PO4Toluene2458
8DBUToluene2433
9Et3NToluene4846
10NaOHDCM1288
11 NaOH CHCI 3 12 94
12NaOHDCE1284
13NaOHH2O480
14cNaOHCHCI31285
15dNaOHCHCI31284
16eNaOHCHCI31280
Open in a separate windowaReaction conditions: 1a (0.1 mmol, 1.0 equiv.), 2a (0.15 mmol, 1.5 equiv.), base (0.2 mmol, 2.0 equiv.), and solvent (1 mL) for 12 h.bIsolated yields.c1a : 2a = 1 : 1.2.dNaOH used 1.2 equiv.e50 °C.Finally, the standard reaction conditions for the base-promoted synthesis of the multi-functionalized benzene derivatives identified as follows: 1.5 equivalence of NaOH and CHCl3 as the solvent under an atmosphere of air for 12 hours at room temperature.With the optimized reaction conditions in hand, we explored the scope of the reaction. A series of enones were examined, variation of the electronic nature of the aromatic ring (R1, including the substituted phenyl or thienyl) has little influence on the reaction efficiency (3b–f, 86–93% yields, Open in a separate windowaReaction conditions: 1a (0.1 mmol, 1.0 equiv.), 2a (0.15 mmol, 1.5 equiv.), NaOH (0.2 mmol, 2.0 equiv.), and CHCl3 (1 mL) for 12 h.We next turned our attention to examine the scope of enals. Different substituents on the phenyl ring of cinnamaldehydes were tolerated even disregarding the position and properties, giving 4a–g in satisfying yields (82–92% yields, Open in a separate windowaReaction conditions: 1a (0.1 mmol, 1.0 equiv.), 2a (0.15 mmol, 1.5 equiv.), NaOH (0.2 mmol, 2.0 equiv.), and CHCl3 (1 mL) for 12 h.To highlight the practicality of this mild and efficient method, the reaction of 2-benzoyl-3-phenylbut-2-enenitrile 1a at 4.0 mmol scale proceed well under the standard conditions to generate the desired product in 88% yield (Scheme 2).Open in a separate windowScheme 2Gram-Scale Synthesis of 3a.The formyl group could be easily reduced by using LiAlH4 in THF at reflux, leading to the formation of the benzyl alcohol product 5 in 95% yield while keeping the CN group intact. Suzuki coupling of 3o with phenylboronic acid furnished derivative 6 in 90% yield13 (Scheme 3).Open in a separate windowScheme 3Synthetic transformation.To gain insight into the role of air in this reaction, a control experiment was designed and investigated (Scheme 4). When the reaction of 1a and 2a was carried out under an argon atmosphere, the desired product 3a was obtained in 10% yield and product 7 could be isolated in 82% yield. The results indicate that oxygen is necessary for the oxidation process and played a key role in this reaction.Open in a separate windowScheme 4Control experiment.A postulated reaction course is illustrated in Scheme 5. Briefly, α-deprotonation of enone 1a in the presence of bases, subsequent 1,4-addition of deprotonated enone I to enal 2a generates intermediate II, which undergoes an intramolecular aldol reaction to yield the adduct 7.14 Lastly, dehydration of 7 followed by spontaneous oxidative aromatization affords the polysubstituted benzonitrile 3a.Open in a separate windowScheme 5The proposed mechanism.  相似文献   

3.
Enantioselective conjugate hydrosilylation of α,β-unsaturated ketones     
Huan Yang  Guanglin Weng  Dongmei Fang  Changjiang Peng  Yuanyuan Zhang  Xiaomei Zhang  Zhouyu Wang 《RSC advances》2019,9(21):11627
Enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones was realized. In the presence of a chiral picolinamide–sulfonate Lewis base catalyst, the reactions provided various chiral ketones bearing a chiral center at the β-position in up to quantitative yields with moderate enantioselectivities.

Enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones was realized.

Chiral ketones are important intermediates for the synthesis of natural products or chiral drugs, and some themselves are useful chiral drugs.1 Asymmetric conjugate reduction of α,β-unsaturated carbonyl compounds is an attractive and challenging transformation for the construction of chiral ketones. During the past few decades, many groups have devoted considerable efforts to this area and made great improvement, for instance, transition metal catalyzed asymmetric hydrogenation of α,β-unsaturated carbonyl compounds, including palladium,2 iridium,3 copper,4 ruthenium,5 rhodium6 and cobalt.7 Besides, some organocatalyzed asymmetric conjugate transfer hydrogenations using Hantzsch ester8 or pinacolborane9 as the hydride source have also been reported. However, very few examples about chiral Lewis base catalyzed conjugate hydrosilylation of α,β-unsaturated carbonyl compounds have been reported and only chiral phosphine oxide Lewis base catalysts were used in these reactions (Scheme 1).10 Therefore, exploration of the application of other kinds of chiral Lewis base catalysts in this reaction is still highly desirable.Open in a separate windowScheme 1Enantioselective conjugate hydrosilylation of α,β-unsaturated ketones by chiral Lewis base catalysts.Recently, we developed a kind of easily accessible chiral picolinamide–sulfonate Lewis base catalysts and used them in asymmetric hydrosilylation of α-acyloxy-β-enamino esters,11 one of them exhibiting excellent reactivity, diastereoselectivity and enantioselectivity. Herein we present the enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones using this kind of Lewis base as catalysts, leading to various chiral ketones bearing a chiral center at β-position (Fig. 1).Open in a separate windowFig. 1Three typical chiral drugs containing chiral ketone moiety.First, various chiral Lewis base catalysts 2 were screened in the enantioselective hydrosilylation of (E)-1,3-diphenylbut-2-en-1-one 1a in acetonitrile at 0 °C. As shown in Fig. 2, l-piperazine-2-carboxylic acid derived N-formamide 2a,12aR-(+)-tert-butylsulfinamide derived catalyst 2b12b and picolinamide 2c12c were found to be totally inactive for the reaction. Meanwhile picolinamide–tosylate catalyst 2d was highly active to afford the product with excellent yield in moderate ee value. Afterwards, several picolinamide–tosylate catalysts 2f–2h bearing electron-withdrawing group in 4-position of pyridine were employed in the reaction and 4-bromo picolinamide 2f delivered a slightly higher ee value. 5-Methoxy picolinamide 2i gave the product with the same ee value as that of 2d but in much lower yield. Lower enantioselectivities were observed with 4-phenyl picolinamide 2j and 3-methyl picolinamide 2k. When (R)-1-(2-aminonaphthalen-1-yl)naphthalen-2-ol derived catalyst 2l was used, no product was observed. When (1S,2R)-1-amino-2,3-dihydro-1H-inden-2-ol derived catalyst 2m gave the product in both poor yield and ee value. Moreover, two picolinamide–sulfonamide catalysts 2n and 2o were also used in the reaction and almost racemic products were obtained. Hence, 2f was determined as the optimal catalyst and was used through out our study.Open in a separate windowFig. 2Evaluation of the chiral Lewis base catalysts 2 in conjugate hydrosilylation of (E)-1,3-diphenylbut-2-en-1-one 1a. Unless otherwise specified, the reactions were carried out with 1a (0.1 mmol), trichlorosilane (0.2 mmol) and catalyst 2 (0.02 mmol) in 1 mL of acetonitrile at 0 °C for 24 hours. Isolated yield based on 1a. The ee values were determined by using chiral HPLC.Subsequently, the other reaction conditions were optimized. The results are summarized in EntryaSolvent T (°C)Time [h]Yieldb [%]eec,d [%]1CH3CN0247751 (R)2CH3CH2CN0248450 (R)3C6H5CN0248324 (R)4THF0249211 (R)51,4-Dioxane024635 (R)6CHCl30249225 (S)7CH2Cl20249638ClCH2CH2Cl0249010 (R)9CCl40249535 (S)10Toluene0249555 (S)11Xylene024N.R.—12Mesitylene024N.R.—13C6H5CF30249531 (S)14Toluene−10489764 (S)15Toluene−20609265 (S)16Toluene−4072N.R.—Open in a separate windowaUnless otherwise specified, the reactions were carried out with 1a (0.1 mmol), trichlorosilane (0.2 mmol) and catalyst 2f (0.02 mmol) in 1 mL of solvent.bIsolated yield based on 1a.cThe ee values were determined by using chiral HPLC.dThe absolute configuration of 3a was determined by comparison of the retention times of the two enantiomers on the stationary phase with those in the literatures.With the optimized conditions in hand, the scope and limitations of the reaction were explored. The results are summarized in Fig. 3. In the presence of 20 mol% of chiral Lewis base catalyst 2f and 2 equivalents of trichlorosilane, various α,β-unsaturated ketones were hydrosilylated. We first tested the effect of various 3-aryl groups of 1. 3-(4-Methoxy-phenyl) substrate 3e (Fig. 3) and 3-(naphthalen-2-yl) substrate 3i (Fig. 3) underwent the reaction to give the products with good yields in enantioselectivities close to 3a, while lower ee values were observed with 3b–3d, 3h, 3k and 3l (Fig. 3). No product was obtained with 3-(2-chloro-phenyl) substrate 3f and 3-(naphthalen-1-yl) substrate 3j, perhaps due to the high steric hindrance (Fig. 3). When 3-(2-methoxy-phenyl) substrate 3g was used, by prolonging the reaction time to 60 hours, the product was obtained with good yield but very poor enantioselection (Fig. 3). Trace amount of the product was detected with 3-(pyridin-2-yl) substrate 3m (Fig. 3). Next, some 3-phenyl-but-2-en-1-ones with different 1-aryl groups were also employed in the reaction. The 4-substituted or 3-substituted substrates delivered good yields of the products with similar or slightly lower enantioselectivities (Fig. 3), while much lower ee value was observed with 2-substituted substrate 3s (Fig. 3). For some other 3-alkyl chalcones, moderate to good yields and moderate ee values were obtained (Fig. 3), except for the bulkier tertiary butyl substituted 3w that gave trace amount of the product (Fig. 3). Reaction of 3y with two different 3,3-aryl groups afforded the product with moderate yield in very low enantioselectivity (Fig. 3). Finally, cyclic substrate 3z was subjected in the reaction and provided the product with excellent yield but in poor ee value (Fig. 3).Open in a separate windowFig. 3Substrate scope of the reaction. Unless otherwise specified, the reactions were carried out with 1 (0.1 mmol), trichlorosilane (0.2 mmol) and catalyst 2f (0.02 mmol) in 1 mL of toluene at −10 °C for 48 hours. Isolated yield based on 1. The ee values were determined by using chiral HPLC. The absolute configuration of 3a was determined by comparison of the retention times of the two enantiomers on the stationary phase with the literatures. The absolute configurations of other products were determined in analogy. aThe reaction time was 60 hours. bThe reaction time was 72 hours.Although detailed structural and mechanistic studies remain to be carried out, based on the absolute configuration of the product 3a, we propose a mechanism shown in Scheme 2. First, the nitrogen atom of the pyridine ring and the carbonyl oxygen atom of catalyst 2f are coordinated to Cl3SiH to create an activated hydrosilylation species. Substrate 1a may approache the Cl3SiH-catalyst complex to generate two transition states A and B. In transition state A, the N–H of catalyst 2f activates the carbonyl group of 1a through H-bonding. In addition, there could be π–π stackings between the two aromatic systems of the catalyst and the substrate. Then Si-face conjugate attack of the hydride to 1a generate (S)-product 3a. On the contrary, in transition state B through which (R)-product will be obtained, the carbonyl group of 1a can not be connected with the N–H of catalyst 2f. Thus the fact that (S)-enriched product 3a was obtained is consistent with the suggestion that the hydrosilylation predominantly proceeds through the pathway involving transition state A rather than transition state B.Open in a separate windowScheme 2A plausible reaction mechanism for hydrosilylation of 1a catalyzed by 2f.In conclusion, we have developed a facile, metal-free and mild enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones. By using chiral picolinamide–sulfonate Lewis base as catalyst, the reactions provided various optically active ketones bearing a chiral center at β-position with moderate to good yields in moderate enantioselectivities. Comparing with the chiral phosphine oxide Lewis base catalysts, the chiral picolinamide–sulfonate is cheaper and easier accessible. The absolute configuration of one product was determined by comparison of the retention times of the two enantiomers on the stationary phase with those in the literature.  相似文献   

4.
Phosphine-catalyzed [3 + 2] annulation of β-sulfonamido-substituted enones with trans-α-cyano-α,β-unsaturated ketones for the synthesis of highly substituted pyrrolidines     
Zhenzhen Gao  Lei Xie  Lusha Ji  Xin Ma  Xiaojing Li  Honglei Liu  Hongchao Guo 《RSC advances》2021,11(63):40136
To synthesize highly substituted pyrrolidines, we developed a phosphine-catalyzed [3 + 2] annulation of β-sulfonamido-substituted enones with trans-α-cyano-α,β-unsaturated ketones. We prepared a series of pyrrolidines under mild conditions with high yields and moderate-to-good diastereoselectivities. A catalytic mechanism for this reaction is suggested.

To synthesize highly substituted pyrrolidines, we developed a phosphine-catalyzed [3 + 2] annulation of β-sulfonamido-substituted enones with trans-α-cyano-α,β-unsaturated ketones.

Nucleophilic phosphine catalysis is a practical and powerful synthetic approach to obtain heterocyclic compounds using various annulation reactions, the advantages of which are it being mild and metal-free, ecologically friendly, and inexpensive.1 Phosphine-catalyzed intermolecular [3 + 2],2 [4 + 1],3 [2 + 2 + 1]4 and intramolecular annulations are often used to obtain pyrrole derivatives. Intermolecular [3 + 2] annulations of imines and phosphorus ylides formed in situ from allenoates, alkynes, or Morita–Baylis–Hillman carbonates under the presence of phosphine catalysts are especially the most widely used approach to synthesize pyrrolidine derivates. In these reactions, phosphorus ylides act as C–C–C synthons for the [3 + 2] annulations with a C Created by potrace 1.16, written by Peter Selinger 2001-2019 N bond converting to a pyrrolidine ring (Scheme 1). However, literature reports on exploring new activation modes, namely, phosphorus ylides acting as C–C–N synthons for the [3 + 2] annulations, are rare.Open in a separate windowScheme 1Pyrrolidine ring formation through reaction of phosphorus ylides act as C–C–C and C–C–N synthons.β-Sulfonamido-substituted enones could be used as C–C–N synthons to form various N-based heterocycles. Catalytically activated (by amines) β-sulfonamido-substituted enones act as nucleophiles towards electron-deficient olefins or imines during [3 + 2] annulation reactions. Du''s5 and Pan''s groups6 have made outstanding contributions to this field.7 In 2018, Guo''s group developed a Bu3P-catalyzed [5 + 1] annulation of γ-sulfonamido-substituted enones with N-sulfonyl-imines to obtain chiral 2,4-di-substituted imidazolidines. They also synthesized γ-sulfonamido-substituted enones attacked by phosphine catalyst and acting as C–C–C–C–N synthon (see Scheme 2).8 Recently, Guo et al.9 used β-sulfonamido-substituted enone as a phosphine acceptor as well as a C–C–N synthon for the [3 + 2] annulation with sulfamate-derived cyclic imines (see Scheme 2). Using of β-sulfonamido-substituted enone as a novel phosphine acceptor is very promising for phosphine-catalyzed reactions. Inspired by Guo''s work, we further extended the substrate scope of this reaction from sulfamate-derived cyclic imines to unsaturated ketones for the construction of pyrrolidine rings. Therefore, in this work, we report phosphine-catalyzed [3 + 2] annulation of β-sulfonamido-substituted enones and trans-α-cyano-α,β-unsaturated ketones, to synthesize highly substituted pyrrolidines (see Scheme 2), which are among the primary building blocks and the core structures of natural and bioactive compounds.10Open in a separate windowScheme 2Phosphine-catalyzed annulation of γ-sulfonamido-substituted enones and β-sulfonamido-substituted enones.We first used trans-α-cyano-α,β-unsaturated ketone 1a and β-sulfonamido-substituted enone 2a as model substrates to obtain optimum reaction conditions. Tertiary phosphine catalysts were screened with 1,2-dichloroethane (DCE) as solvent at room temperature (see Thus, the optimum reaction conditions were determined as follows: using 20 mol% of PMe3 as catalyst, CHCl3 as solvent at room temperature.Optimization of reaction conditionsa
EntryPR3Solvent t/hCon./mol L−1Yieldb (%)drc
1MePPh2DCE80.1855 : 1
2EtPPh2DCE80.1744 : 1
3 n-PrPPh2DCE80.1764 : 1
4Me2PPhDCE80.1823 : 1
5PBu3DCE80.1783 : 1
6PMe3DCE80.1846 : 1
7PMe3THF80.1857 : 1
8PMe3Toluene80.1757 : 1
9PMe3EtOAc80.1785 : 1
10PMe3CHCl380.1888 : 1
11dPMe3CHCl380.1848 : 1
12ePMe3CHCl380.1868 : 1
13PMe3CHCl3240.05859.5 : 1
14PMe3CHCl3480.0338511 : 1
15PMe3CHCl3240.026514 : 1
16PMe3CHCl3720.028614 : 1
Open in a separate windowaUnless otherwise indicated, all reactions were carried out at room temperature using 0.12 mmol of 1aa and 0.1 mmol of 2aa in a solvent containing 20 mol% of the catalyst.bIsolated yield.cDetermined by 1H NMR.d100 mg 3 Å molecular sieves were used.e100 mg 4 Å molecular sieves were used.Under the optimum conditions, the performance of various trans-α-cyano-α,β-unsaturated ketone 1 with β-sulfonamido-substituted enones 2a in the cycloaddition reactions was analyzed (see 11Screening of various trans-α-cyano-α,β-unsaturated ketones as substratesa
EntryR13Yieldb (%)drc
1Ph (1a)3aa8614 : 1
22-MeC6H4 (1b)3ba7510.5 : 1
33-MeC6H4(1c)3ca7712.5 : 1
44-MeC6H4 (1d)3da7810.5 : 1
54-OMeC6H4 (1e)3ea8014 : 1
64-CF3-C6H4 (1f)3fa6610.5 : 1
72-FC6H4 (1g)3ga729.5 : 1
83-FC6H4 (1h)3ha746 : 1
94-FC6H4 (1i)3ia765 : 1
102-ClC6H4 (1j)3ja748 : 1
113-ClC6H4(1k)3k7610 : 1
124-ClC6H4 (1l)3la825 : 1
134-BrC6H4 (1m)3ma856 : 1
141-Naphthyl (1n)3na8114 : 1
152-Naphthyl (1o)3oa808 : 1
162-thienyl (1p)3pa787 : 1
172-furyl (1q)3qa8014 : 1
Open in a separate windowaUnless otherwise indicated, all reactions were conducted at room temperature for 3 days using 0.12 mmol of compound 1 and 0.1 mmol of compound 2 in 5 ml CHCl3 in the presence of 20 mol% of PMe3.bIsolated yield.cDetermined by 1H NMR.We also tested various substituted enones containing different R groups under the optimal reaction conditions (see EntryR2/R33Yieldb (%)drc1Ph/Ts (2a)3aa8614 : 12Ph/Bs (2b)3ab8410 : 13Ph/Ns (2c)3ac814.5 : 142-FC6H4/Ts (2d)3ad778 : 153-FC6H4/Ts (2e)3ae799 : 162-ClC6H4/Ts (2f)3af828 : 173-BrC6H4/Ts(2g)3ag749 : 184-BrC6H4/Ts (2h)3ah858 : 193,4-Cl2C6H3/Ts (2i)3ai7410 : 1104-CNC6H4/Ts (2j)3aj8611 : 1113-OMeC6H4/Ts (2k)3ak7910 : 1124-OMeC6H4/Ts (2l)3al808.5 : 1134-PhC6H4/Ts (2m)3am8612.5 : 1142-naphthyl/Ts (2n)3an818 : 1Open in a separate windowaUnless otherwise noted, all reactions were performed at room temperature for 3 days using 0.12 mmol of compound 1 and 0.10 mmol of compound 2 in 5 ml CHCl3 under the presence of 20 mol% PMe3.bIsolated yield.cDetermined by 1H NMR.To demonstrate the synthetic potential of the cycloaddition reaction, a scale-up preparation of 3aa and the derivatization of 3am were performed (Scheme 3). The unsaturated ketone 1a (699 mg, 3.0 mmol) reacted with substituted enone 2a (788 mg, 2.5 mmol) under the standard condition to give 3aa in 81% yield with 13 : 1 dr. In comparison with the reaction at 0.1 mmol of scale, no significant loss of yield and diastereoselectivity was observed. Reduction of the carbonyl group of 3ma with NaBH4 in MeOH/CH2Cl2 led to the formation of compound 4 in 85% yield and 5.5 : 1 dr.Open in a separate windowScheme 3The reaction on the gram-scale and further transformations.The asymmetric variant of the present reaction had also been investigated ( EntryCatalyst t/hYieldb (%)drceec1P172Trace——2P27220>20 : 153P37250>20 : 1314P472NRd——5P572NRd——Open in a separate windowaUnless otherwise indicated, all reactions were carried out at room temperature using 0.06 mmol of 1aa and 0.05 mmol of 2aa in a solvent containing 20 mol% of the catalyst in 2.5 ml of CHCl3.bIsolated yield.cDetermined by HPLC on chiral column.dNo reaction.All these results allowed us to propose a catalytic cycle (see Scheme 4). Nucleophilic addition of the phosphine-based catalysts to β-sulfonamido-substituted enones yields phosphonium intermediate A, which converts into an intermediate B by proton transferation. The intermediate B undergoes intramolecular aza-Michael addition to an alkene yielding an intermediate compound C, followed by intramolecular nucleophilic substitution and the producing of product 3, during which the phosphine regenerates.Open in a separate windowScheme 4Proposed mechanism.In conclusion, we developed a synthesis method (under mild conditions) for highly substituted pyrrolidines through phosphine-catalyzed [3 + 2] annulation of β-sulfonamido-substituted enones with trans-α-cyano-α,β-unsaturated ketones. A series of pyrrolidine derivates were obtained in good yields with moderate-to-good diastereoselectivities. In this reactions, using of β-sulfonamido-substituted enone as a novel phosphine acceptor, the formed phosphorus ylides act as C–C–N synthons for annulations. Further investigations on the application of β-sulfonamido-substituted enones in the asymmetric phosphine-catalyzed reactions are in progress in our laboratory.  相似文献   

5.
Metal-free hydrosulfonylation of α,β-unsaturated ketones: synthesis and application of γ-keto sulfones     
Xiufang Cheng  Shuo Wang  Yibo Wei  Huamin Wang  Ying-Wu Lin 《RSC advances》2022,12(55):35649
γ-Keto sulfones are versatile building blocks and valuable intermediates in organic synthesis and pharmaceutical chemistry. Motivated by their excellent properties, we herein report a green, convenient, metal-free hydrosulfonylation method for a variety of ynones, vinyl ketones, and sodium sulfinates in the absence of stoichiometric oxidants. This operationally simple protocol provides straightforward and practical access to a wide range of γ-keto sulfones with broad functional group tolerance from easily available starting materials. Moreover, the β,γ-unsaturated keto sulfones could further react with 2,3-butadienoate to generate cyclopentenes in phosphine-mediated [3 + 2] cycloaddition.

The methodology features a convenient, mild, efficient, C–S sulfonylation approach without the use of any metal catalysts and stoichiometric oxidants.

As a useful common structural fragment in a broad number of pharmaceuticals1 and functional materials,2 keto sulfones are usually present in promising biologically active molecules such as Casodex,3VCAM-1 (ref. 4) and anti-HIV-1 (ref. 5) (Fig. 1). Furthermore, a valuable synthetic impression is associated with the role of reactive intermediates in various high-demand synthetic transformations,6 including total synthesis.7 Owing to their excellent properties, and efficient and practical synthesis methods keto sulfones are in high demand.Open in a separate windowFig. 1Representative biologically active γ-keto sulfones.In the past decades, a variety of protocols have been developed to construct β-keto sulfones.8 Whereas succinct synthetic routes toward structurally related γ-keto sulfones are scarce,9 traditionally, γ-keto sulfones were synthesized via the nucleophilic substitution of sodium sulfinates by 2-chlorovinyl ketones,10 the elimination of the bromo derivatives of saturated keto sulfones11 and the oxidation of the corresponding sulfides or sulfoxides.12 However, the principal drawback is that these procedures were strongly limited by multiple steps, narrow substrate scope, or poor stereoselectivity.Indeed, several streamlined strategies for the preparation of γ-keto sulfones involves addition reaction of alkenes or alkynes have been developed.13 Li''s group14 reported the synthesis of (E)-vinyl sulfones through Pd-catalyzed conjugate additions of alkynes with 1,2-bis(phenylsulfonyl)ethane. In 2013 Jiang and co-workers15 showed that a Pd-catalyzed sulfonylation of alkynoates with sodium sulfinates affords γ-keto sulfones (Scheme 1a). Li and coworkers16 reported that BPO triggered the hydrosulfonylation of chalcones with arylsulfonyl hydrazides producing γ-keto sulfones. Subsequently, Bi''s group17 developed a Ag2CO3-promoted sulfonylation of allyl/propargyl alcohols with sodium sulfinates for the preparation of γ-keto sulfones (Scheme 1b). Nevertheless, most cases still have to use large excess oxidants, noble metal catalysts, or require high temperatures. Accordingly, an efficient, mild and practical method to furnish γ-keto sulfones is worthwhile studying.Open in a separate windowScheme 1Methods for the synthesis of γ-keto sulfones.With growing demand for sustainable chemistry, an “ideal” reaction system for such transformations would be “metal-free” due to cost efficiency and possible advantages regarding toxicity, as well as selectivity. With this intent, we herein describe a simple and efficient acid-mediated sulfonylation of sodium sulfinates and α,β-unsaturated ketones for the selective synthesis of γ-keto sulfones (Scheme 1c). The significant advantages of this method are high efficiency, metal-free and mild reaction conditions, thus providing a potential application in natural product synthesis and medicinal chemistry.Further studies were commenced with the optimization of the conditions for the hydrosulfonylation of the ynone 1f with sodium benzosulfonate 2a (13e and acetyl chloride/H2O,19 as used in the previous study, were completely ineffective due to several unknown complex products being formed (entry 1). Gratifyingly, the desired γ-keto sulfone 3fa was isolated in a 49% yield (E/Z = 90 : 10) as the major product for the reaction mediated by AcOH (entry 3). Encouraged by this initial result, we screened an array of acids. The results showed that 4-chlorobenzoic acid (PCBA) gave the best result, leading to the isolation of γ-keto sulfone 3fa in a yield of 80% (E/Z = 98 : 02) (entries 4–15). Solvent screening indicated that mesitylene could improve the yield to 85% (E/Z = 95 : 05) (entry 21). Further investigations on the reduced usage of PCBA to 2.0 equivalents, the yield of 3fa was slightly reduced (entry 22, 83% yield, E/Z = 95 : 05). The amounts of sodium benzosulfonate 2a and the reaction temperature have deleterious effects on the reaction yields (entries 23–27). Thus, the optimized reaction conditions were successfully established as 1f (1.0 equiv.), 2a (2.5 equiv.), PCBA (2.0 equiv.), and mesitylene (2.0 mL) at 30 °C in this process.Optimization of the reaction conditionsa
EntryAcid (x equiv.)SolventYieldb (%) E/Zc
1Buffer (pH = 3.5)DMFNR
2Acetyl chloride/H2OCHCl3NR
3AcOH (3.0)Toluene4990 : 10
4HCO2H (3.0)Toluene2385 : 15
5HCl (3.0)Toluene3680 : 20
6HNO3 (3.0)Toluene3687 : 13
7Benzoic acid (3.0)Toluene6696 : 04
8 p-Toluic acid (3.0)Toluene5296 : 04
94-Acetylbenzoic acid (3.0)Toluene5796 : 04
104-Fluorobenzoic acid (3.0)Toluene7398 : 02
11PCBA (3.0)Toluene8098 : 02
124-Bromobenzoic acid (3.0)Toluene7295 : 05
13PNBA (3.0)Toluene5788 : 12
142-Naphthoic acid (3.0)Toluene4894 : 06
152-Nitrobenzoic acid (3.0)Toluene2994 : 06
16PCBA (3.0) o-Xylene7392 : 08
17PCBA (3.0) p-Xylene7090 : 10
18PCBA (3.0) m-Xylene7296 : 04
19PCBA (3.0)DMFNR
20PCBA (3.0)MeOH6889 : 11
21PCBA (3.0)Mesitylene8595 : 05
22PCBA (2.0)Mesitylene8395 : 05
23PCBA (1.2)Mesitylene7691 : 09
24PCBA (0.5)Mesitylene4473 : 27
25dPCBA (2.0)Mesitylene7990 : 10
26ePCBA (2.0)Mesitylene8095 : 05
27fPCBA (2.0)Mesitylene5397 : 03
Open in a separate windowaReaction conditions: 1f (0.1 mmol), 2a (0.25 mmol), acid (x equiv.), solvent (1.0 mL), 30 °C, 48 h.bIsolated yields.cDetermined by RP-HPLC.dWith 2.0 equiv. 2a.e50 °C.f80 °C. PCBA = 4-chlorobenzoic acid. PNBA = p-nitrobenzoic acid.We then sought to explore the generality of the method for the synthesis of α,β-unsaturated γ-keto sulfones, using various ynones in reactions with 2a under the optimized conditions (Scheme 2). The reaction of the 1-phenylprop-2-yn-1-one 1a with 2a proceeded reasonably to provide an excellent yield of the corresponding γ-keto sulfone 3aa (97% yield, E/Z = 98 : 02). To our delight, the reaction worked successfully with a range of ynones 1 bearing various substituents on the aromatic ring. Substituents such as methyl, thiomethylmethoxy, phenyl, halogen and dimethylamino atoms could be tolerated and gave the corresponding products 3ba–3ja with high to excellent yields (71–98% yield) and stereoselectivity (E/Z = 82 : 18 to 98 : 02). Trifluoromethyl and nitro substituents on the aromatic ring were also compatible and products 3ka and 3la were afforded 95% and 71% yields, respectively. 9-Anthracenee-derived ynone successfully afforded 3ma in a 94% yield (E/Z = 98 : 02). The methyl group in the ortho or meta positions of the aromatic ring gave the desired γ-keto sulfones in 82% and 94% yield, respectively. The desired product 3pa bearing a pitavastatin unit could be readily prepared in a yield of 87%. When alkyl terminal alkynone 1q was subjected to the reaction, affording the desired product 3qa in 65% yield (E/Z = 97 : 03).1Open in a separate windowScheme 2Sulfonylation reaction of various terminal alkynones with 2a. All reactions were carried on 0.2 mmol scale in mesitylene (2.0 mL) and used 2.5 equiv. of 2a, 2.0 equiv. PCBA, at 30 °C. Yields of isolated products are reported. E/Z ratios were determined by RP-HPLC.Inspired by the above results, the nonterminal alkyne was used as the substrate to react with PhSO2Na at 30 °C for 36 h. The reaction provided E and Z-β-sulfonyl-α,β-unsaturated carbonyl mixed compounds 3qa13e (86% yield, E/Z = 1 : 1).The results of ynone 1a reacting with a number of sodium sulfinates under the optimized condition are depicted in Scheme 3. Gratifyingly, no matter whether the phenyl ring of sodium sulfinate was substituted with either a sterically hindered, electron-donating, or electron-withdrawing group, all of them smoothly furnished the corresponding products in moderate to excellent yields with a high range of E/Z ratios from 52 : 48 to 97 : 03 (3ab–3an). Likewise, 2-napthyl and cyclopropyl substituted sodium sulfinates were both effective in this reaction with a yield of 87% and 85%, respectively (3ao and 3ap). Additionally, l-10-camphorsulfonyl sulfinate 2q was also suitable for this reaction.Open in a separate windowScheme 3Sulfonylation reaction of terminal alkynone (1a) with sodium Sulfinates. All reactions were carried on 0.2 mmol scale in mesitylene (2.0 mL) and used 2.5 equiv. of 2, 2.0 equiv. PCBA, at 30 °C. Yields of isolated products are reported. E/Z ratios were determined by RP-HPLC.Interestingly, the treatment of the vinyl ketone 4a with PhSO2Na (2a) under the standard conditions furnished sulfone 5aa (Scheme 4). The substrate scope was also explored in Scheme 4. Delightfully, it was perfectly tolerable to introduce both electron-donating (OCH3 and Ph) and electron-withdrawing (F, Cl, and CN) groups at the para position of the phenyl ring, affording the corresponding products (5ba–5fa) in excellent yields. 4-Toluene sulfonate and cyclopropane sulfonate also reacted well with substrate 2a to form γ-keto sulfone in excellent yields. We were pleased to find that the β-trifluoromethylated enone 4h and trans-chalcone (4i–4j) could be successfully employed to give desired products (5ha–5ja, 55–61% yields). Unfortunately, no reaction occurred for 2-cyclopentenone.Open in a separate windowScheme 4Sulfonylation reaction of vinyl ketone with sodium sulfinates. All reactions were carried on 0.2 mmol scale in mesitylene (2.0 mL) and used 2.0 equiv. of 2a, 2.0 equiv. PCBA, at 30 °C. Yields of isolated products are reported. aAt 80 °C for 72 h.Additionally, the synthetic utility of the γ-keto sulfones obtained by the present method was explored (Scheme 5). Gram-scale ynone 1a was reacted with sodium benzosulfonate 2a to form product 3aa with an excellent E/Z ratio (A). Lu''s [3 + 2] cycloaddition of 2,3-butadienoate with α,β-unsaturated γ-keto sulfones 3 mediated by phosphine produced cycloadducts 6 (ref. 18) in good yields (B). Moreover, pyrazole derivative 6b could be efficiently obtained from 3aa under ultrasound (US) irradiation conditions (C). Next, γ-keto sulfone 3pa derived from the biologically active pitavastatin could also react with hydrazine to give a high yield of 6c (D).Open in a separate windowScheme 5Gram-scale preparation and further synthetic utilization.To understand the reaction mechanism, control reactions of 1a with 2a were examined (Scheme 6a). When 1a and 2a was subjected to the standard reaction conditions except using deuterated 4-chlorobenzoic acid system, the 3a were detected with 80% yield. An attempt to run the reaction of 1a and 2a in a anhydrous solvent system under an N2 atmosphere also successfully delivered 3a in 97% yield.20 The results unambiguously disclosed that the incorporated hydrogen atoms in 3a originated from acid rather than water. The reaction using 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) and 2,4-di-tert-butyl-4-methylphenol (BHT) as the radical scavengers showed no observable radical intermediates and unaffected desired products formation, which suggests that the radical process could be ruled out.20 On the basis of the results presented above and previous reports, we propose the following mechanism in Scheme 6b. The 4-chlorobenzoic acid activates the carbonyl group in α,β-unsaturated ketones 1 (4) to afford intermediate I or tautomerize to intermediate II. Finally, sulfonyl anion can add to the unsaturated bond of intermediate II to afford the products 3 (5).Open in a separate windowScheme 6Mechanistic studies.  相似文献   

6.
Catalyst-free chemoselective α-sulfenylation/β-thiolation for α,β-unsaturated carbonyl compounds     
Xi Huang  Juan Li  Xiang Li  Jiayi Wang  Yanqing Peng  Gonghua Song 《RSC advances》2019,9(45):26419
A novel, efficient, catalyst-free and product-controllable strategy has been developed for the chemoselective α-sulfenylation/β-thiolation of α,β-unsaturated carbonyl compounds. An aromatic sulfur group could be chemoselectively introduced at α- or β-position of carbonyls with different sulfur reagents under slightly changed reaction conditions. A series of desired products were obtained in moderate to excellent yields. Mechanistic studies revealed that B2pin2 played the key role in activating the transformation towards the β-thiolation of α,β-unsaturated carbonyl compounds. This transition-metal-catalyst-free method provides a convenient and efficient tool for the highly chemoselective preparation of α-thiolation or β-sulfenylation products of α,β-unsaturated carbonyl compounds.

This catalyst-free method provides a useful and efficient tool for the highly chemoselective preparation of α-thiolation or β-sulfenylation products of α,β-unsaturated carbonyl compounds.  相似文献   

7.
Hydrophilic and organophilic pervaporation of industrially important α,β and α,ω-diols     
Shivshankar Chaudhari  HyeonTae Shin  SeoungYong Choi  KieYong Cho  MinYoung Shon  SeungEun Nam  YouIn Park 《RSC advances》2021,11(16):9274
The distillation-based purification of α,β and α,ω-diols is energy and resource intensive, as well as time consuming. Pervaporation separation is considered to be a remarkable energy efficient membrane technology for purification of diols. Thus, as a core pervaporation process, hydrophilic polyvinyl alcohol (PVA) membranes for the removal of water from 1,2-hexanediol (1,2-HDO) and organophilic polydimethylsiloxane–polysulfone (PDMS–PSF) membranes for the removal of isopropanol from 1,5 pentanediol (1,5-PDO) were employed. For 1,2-HDO/water separation using a feed having a 1 : 4 weight ratio of 1,2-HDO/water, the membrane prepared using 4 vol% glutaraldehyde (GA4) showed the best performance, yielding a flux of 0.59 kg m−2 h−1 and a separation factor of 175 at 40 °C. In the organophilic pervaporation separation of the 1,5-PDO/IPA feed having a 9 : 1 weight ratio of components, the PDMS membrane prepared with a molar ratio of TEOS alkoxy groups to PDMS hydroxyl groups of 70 yielded a flux of 0.12 kg m−2 h−1 and separation factor of 17 638 at 40 °C. Long term stability analysis found that both hydrophilic (PVA) and organophilic (PDMS) membranes retained excellent pervaporation output over 18 days'' continuous exposure to the feed. Both the hydrophilic and organophilic membranes exhibited promising separation performance at elevated operating conditions, showing their great potential for purification of α,β and α,ω-diols.

The distillation-based purification of α,β and α,ω-diols is energy and resource intensive, as well as time consuming.  相似文献   

8.
The synthesis of HMF-based α-amino phosphonates via one-pot Kabachnik–Fields reaction     
Weigang Fan  Yves Queneau  Florence Popowycz 《RSC advances》2018,8(55):31496
The first use of biomass-derived HMF in the one-pot Kabachnik–Fields reaction is reported here. A wide range of furan-based α-amino phosphonates were prepared in moderate to excellent yields under mild, effective and environmentally-benign conditions: iodine as a non-metal catalyst, biobased 2-MeTHF as the solvent and room or moderate temperature. The hydroxymethyl group of HMF persists in the Kabachnik–Fields products, widening the scope of further modification and derivatization compared to those arising from furfural. Issues involving the diastereoselectivity and double Kabachnik–Fields condensation were also faced.

A mild and efficient one-pot protocol for the synthesis of α-amino phosphonates directly from 5-HMF was described.

Recently, the production of chemicals from renewable biomass has attracted growing interests due to the dwindling reserves of fossil resources and the increasing awareness of environmental concerns.1 5-Hydroxymethylfurfural (HMF), a promising primary biomass-derived platform chemical readily obtained from acid-catalyzed dehydration of six-carbon carbohydrates, displays a strong potential in organic synthesis.2 Besides the well-developed conversions of HMF towards monomers and biofuels via oxidation or reduction reactions,3 some remarkable strategies converting HMF to high value-added fine chemicals have been disclosed.4 Nevertheless, the specific reactivity and reduced stability of HMF, in comparison with the pentose-derived furfural homolog, have limited its use in synthetic applications.5 Furthermore, its commercial availability, though not anymore a barrier nowadays, has limited the number of investigations in the past. In this regard, developing efficient and economic routes to existing or novel fine chemicals from HMF, with its different reactivity compared to simpler aldehydes, is still a challenge.Multi-component reactions (MCRs) are extremely convenient and efficient strategies to prepare highly functionalized compounds from simple starting materials by one-pot procedures. Since they have many advantages, such as high atom economy, high convergence, time and energy saving, MCRs have gained much attention in modern synthetic organic chemistry.6 Nevertheless, to our knowledge, the direct utilization of HMF in MCR processes has been only rarely explored.α-Amino phosphonates, due to the structural analogy to natural α-amino acids and their significant biological activities, such as antitumor, antitubercular, cytotoxic activities, and so on,7 have been the subject of considerable interest in the past decades both in synthetic organic and medicinal chemistry.8 Among several methods for preparing α-amino phosphonates, the Kabachnik–Fields reaction, a one-pot condensation of an aldehyde, an amine and a dialkyl phosphite is the most effective and convenient strategy.9 A large number of conditions have been reported for the acid-catalyzed (Lewis/Brønsted)10 and catalyst-free11 Kabachnik–Fields reaction, affording the α-amino phosphonates from various aldehydes. However, none of studies have included HMF as a substrate in their scope although the products from HMF can offer more possibilities for further functionalization, thanks to its CH2OH appendage. The sole synthesis of α-amino phosphonates from HMF was reported by Cottier and Skowroński in a two-step reaction strategy, consisting in pre-formation of the imine which upon isolation reacted with dialkyl phosphites as nucleophilic species at high temperature using trifluoroacetic acid as catalyst.12 Thus, a milder and more direct procedure for the synthesis of α-amino phosphonates from HMF is still to be developed.As part of our on-going interest on the application of HMF towards fine chemicals and on green and sustainable chemistry,13 we explored the possibility to synthesize furan-based α-amino phosphonates via the one-pot Kabachnik–Fields condensation, directly from HMF.For this study, we selected molecular iodine as a mild and effective Lewis acid catalyst, as often used in multicomponent synthesis because of its operational simplicity, low cost and toxicity and likely to be compatible with HMF sensitivity to acidic conditions.14 Wu and co-workers have confirmed its efficiency in Kabachnik–Fields reactions of simple aldehydes such as benzaldehyde and furfural.15 The primary set of experimental conditions has been fixed as 5 mol% iodine in ethanol [0.5 M] with equimolar stoichiometric ratio of all partners (HMF, aniline, diethyl phosphite). The corresponding Kabachnik–Fields product 4a was obtained in 71% isolated yield after 24 h, together with around 11% of unreacted HMF and 6% of the intermediate imine ( EntryCat. loadingSolvent [0.5 M]Temp.Ratio 1a/2a/3aTimeIsolated yield15 mol%EtOH25 °C1 : 1 : 124 h71%25 mol%MeCN25 °C1 : 1 : 124 h60%35 mol%DCM25 °C1 : 1 : 124 h31%45 mol%THF25 °C1 : 1 : 124 h84%55 mol%2-MeTHF25 °C1 : 1 : 124 h74%65 mol%THF25 °C1 : 1 : 1.524 h90% 7 5 mol% 2-MeTHF 25 °C 1 : 1 : 1.5 8 h 91% 82.5 mol%2-MeTHF25 °C1 : 1 : 1.58 h77%91 mol%2-MeTHF25 °C1 : 1 : 1.58 h61%10—2-MeTHF25 °C1 : 1 : 1.58 h54% (80%)b115 mol%2-MeTHF50 °C1 : 1 : 1.54 h83%125 mol%2-MeTHF78 °C1 : 1 : 13 h71%Open in a separate windowaThe reaction was carried out in a sealed tube with HMF, aniline, diethyl phosphite, solvent and iodine, stirred at corresponding temperature for indicated time.b24 h.Based on this preliminary result, the reaction conditions were optimized, first by studying the influence of the solvent. THF was found to provide a better yield than EtOH, MeCN and DCM, affording product 4a in 84% yield (16 we decided to continue the investigation with 2-MeTHF as the solvent.Decreasing the catalyst loading to 2.5 mol% and 1 mol% led to slightly slower reactions (77% and 61% respectively) (Scheme 1 is depicted the scope of amines used in the reaction.Open in a separate windowScheme 1The Kabachnik–Fields reaction of HMF and different amines.a,b aThe reaction was carried out with HMF (1 mmol), amine (1 mmol), diethyl phosphite (1.5 mmol) with I2 (5 mol%) in 2-MeTHF (2 mL), stirred at 25 °C for indicated time. bIsolated yield. cAt 50 °C.Whatever the electron-donating or electron-withdrawing nature of the para substituent (methoxy-, chloro-, bromo-, iodo- and nitro-) on the aniline, the corresponding α-amino phosphonates 4b–4f were obtained in good to excellent yields (71–90%). An exception was observed for p-iodo-aniline requiring a 50 °C temperature for producing 4e in 77% yield. The same tendency was noticed for meta-substituted anilines, presumed to display low electronic influence on the reactivity (yields of 93% for 4g and 87% for 4h), and in a more unexpected way for 2-chloroaniline (82% for 4i). These results revealed that the substituted group on phenyl ring of aniline has globally a low impact on the reaction. Compared to anilines, aliphatic amines were found consistently as less reactive. In the case of aliphatic amines, elevated temperature (50 °C) was required to promote the reaction. Benzylamine and furfurylamine provided the corresponding α-amino phosphonates 4j and 4k in moderate yields, respectively 71% and 70%. Similar results were obtained for n-butylamine, cyclohexylamine and allylamine (4l–4n). Non-protected tryptamine afforded compound 4o in 57% yield. The product possibly arising from the reaction of the pyrrolic amine of tryptamine was not observed. tert-Butyl glycinate also worked under the conditions but gave a poor yield of 4p (31%). N-Methyl aniline, as an example of secondary amine, was also less reactive than aniline, giving 4q in 58% yield. When the chiral amine (R)-α-methylbenzylamine was used, the mixture of products was obtained 4r in 72% yield, from which the two isomers could not be separated entirely by column chromatography. A moderate diastereoselectivity was observed, with a 3.3 : 1 ratio of two diastereoisomers observed on the base of 31P NMR spectra. Similarly, (S)-α-methylbenzylamine gave the products 4s as a 3.5 : 1 mixture of two diastereoisomers in 70% yield.The nature of the dialkyl phosphite was also examined but to a minor extent due to the low diversity of commercially available phosphite reagents (Scheme 2). Dimethyl-, diisopropyl- and dibenzyl-phosphites afforded the corresponding products (4t–4v) in a range of yields of 86–89%. Surprisingly, phosphite with two strongly electron-withdrawing CF3– groups could also be used affording 4w in a modest 54% yield under the optimized conditions.Open in a separate windowScheme 2The Kabachnik–Fields reaction of HMF and commercially available phosphites.In order to expand the application of HMF, a pre-prepared 5,5′-[oxybis(methylene)]bis-2-furfural via self-etherification of HMF was subjected to the optimized conditions, resulting into the expected double Kabachnik–Fields product 4x in 86% yield. Alternatively, using p-phenylenediamine instead of aniline led to the other type of double Kabachnik–Fields product 4y. The results above undoubtedly indicate the possible application of the strategy towards highly functional polymers via Kabachnik–Fields polycondensation of 5,5′-[oxybis(methylene)]bis-2-furfural and suitable diamines (Scheme 3).17Open in a separate windowScheme 3The double Kabachnik–Fields reaction.A couple of derivatizations on hydroxyl group of the Kabachnik–Fields product were investigated using 4t as model substrate (Scheme 4). The aldehyde 4aa could be prepared in 87% yield by oxidation of 4t using Dess–Martin periodinane (DMP). The hydroxyl group of 4t could be also converted into an azido group after treatment with diphenylphosphoryl azide in the presence of DBU in 42% yield (compound 4ab). The acrylate 4ac was also easily obtained in a good yield. Diversification and optimization of these reactions are now in progress in the lab for further illustrating the usefulness of the hydroxymethyl appendage and providing a library of new α-amino phosphonates.Open in a separate windowScheme 4The derivatizations on hydroxyl group.Usually in a multicomponent reaction, the mechanism is not distinct because the reaction may undergo different pathways depending on which reactants react at first step. In order to gain insight into the mechanism of the Kabachnik–Fields reaction in our case, a series of control stepwise experiments were carried out (Scheme 5 and for more details see ESI). Mixing HMF (1 mmol) and aniline (1 mmol) in 2-MeTHF yielded the imine rapidly with and without iodine, with around 90% conversion observed in the crude NMR after 40 min in both cases (Exp. A and B). Subsequent addition of diethyl phosphite (1.5 mmol) and I2 (5 mol%) to the solution of the in situ formed imine (HMF, aniline, 1 h) afforded cleanly 4a after 8 h as seen by NMR (Exp. G). On the other hand, no reaction occurred when HMF (1 mmol) and diethyl phosphite (1.5 mmol) were mixed, either in the presence or absence of iodine (Exp. C and D). The above results indicate that the reaction likely undergoes the imine pathway, followed by nucleophilic attack by the phosphite to afford the α-amino phosphonate.8e,18 It is also known that iodine, as a Lewis acid, can activate imines in nucleophilic addition reactions.8e,19 This imine pathway was corroborated by the observation of the imine in the crude NMR of the three-component reaction mixture in the absence of iodine (Exp. E). In the presence of iodine, the proton of CH Created by potrace 1.16, written by Peter Selinger 2001-2019 N is shifted from 8.17 ppm to 8.42 ppm which made it difficult to identify, but the imine component was clearly detected in the crude reaction mixture by MS (imine plus H+: 202.0), thus also supporting the imine pathway (Exp. F).Open in a separate windowScheme 5Control experiments.  相似文献   

9.
Chemoselective and metal-free reduction of α,β-unsaturated ketones by in situ produced benzeneselenol from O-(tert-butyl) Se-phenyl selenocarbonate     
Andrea Temperini  Marco Ballarotto  Carlo Siciliano 《RSC advances》2020,10(56):33706
The carbon–carbon double bond of arylidene acetones and chalcones can be selectively reduced with benzeneselenol generated in situ by reacting O-(tert-butyl) Se-phenyl selenocarbonate with hydrochloric acid in ethanol. This mild, metal-free and experimentally simple reduction procedure displays considerable functional-group compatibility, products are obtained in good to excellent yields, and the use of toxic Se/CO mixture and NaSeH, or the smelly and air-sensitive benzeneselenol, is avoided.

The carbon–carbon double bond of arylidene acetones and chalcones can be selectively reduced with benzeneselenol generated in situ by reacting O-(tert-butyl) Se-phenyl selenocarbonate with hydrochloric acid in ethanol.  相似文献   

10.
Cationic palladium(ii)-catalyzed synthesis of substituted pyridines from α,β-unsaturated oxime ethers     
Takahiro Yamada  Yoshimitsu Hashimoto  Kosaku Tanaka  III  Nobuyoshi Morita  Osamu Tamura 《RSC advances》2022,12(33):21548
An efficient method for the synthesis of multi-substituted pyridines from β-aryl-substituted α,β-unsaturated oxime ethers and alkenes via Pd-catalyzed C–H activation has been developed. The method, using Pd(OAc)2 and a sterically hindered pyridine ligand, provides access to various multi-substituted pyridines with complete regioselectivity. Mechanistic studies suggest that the pyridine products are formed by Pd-catalyzed electrophilic C–H alkenylation of α,β-unsaturated oxime followed by aza-6π-electrocyclization. The utility of this method is showcased by the synthesis of 4-aryl-substituted pyridine derivatives, which are difficult to synthesize efficiently using previously reported Rh-catalyzed strategies with alkenes.

An efficient method for the synthesis of multi-substituted pyridines from α,β-unsaturated oxime ethers via cationic Pd(ii)-catalyzed C–H activation has been developed.  相似文献   

11.
Direct β-selectivity of α,β-unsaturated γ-butyrolactam for asymmetric conjugate additions in an organocatalytic manner     
Yuan Zhong  Sihua Hong  Zhengjun Cai  Shixiong Ma  Xianxing Jiang 《RSC advances》2018,8(51):28874
The β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by a bifunctional chiral tertiary amine has been developed, which provides an efficient access to optically active β-position functionalized pyrrolidin-2-one derivatives in both high yield and enantioselectivity (up to 78% yield and 95 : 5 er). This is the first catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach.

The asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by (DHQD)2AQN has been developed, which provides an access to β-position functionalized pyrrolidin-2-one derivatives in high levels yield and enantioselectivity.

Metal-free organocatalytic asymmetric transformations have successfully captured considerable enthusiasm of chemists as powerful methods for the synthesis of various kinds of useful chiral compounds ranging from the preparation of biologically important molecules through to novel materials.1 Chiral pyrrolidin-2-ones have been recognized as important structural motifs that are frequently encountered in a variety of biologically active natural and synthetic compounds.2 In particular, the β-position functionalized pyrrolidin-2-one backbones, which can serve as key synthetic precursors for inhibitory neurotransmitters γ-aminobutyric acids (GABA),3 selective GABAB receptor agonists4 as well as antidepressant rolipram analogues,5 have attracted a great deal of attention. Therefore, the development of highly efficient, environmentally friendly and convenient asymmetric synthetic methods to access these versatile frameworks is particularly appealing.As a direct precursor to pyrrolidin-2-one derivatives, recently, α,β-unsaturated γ-butyrolactam has emerged as the most attractive reactant in asymmetric organometallic or organocatalytic reactions for the synthesis of chiral γ-position functionalized pyrrolidin-2-ones (Scheme 1). These elegant developments have been achieved in the research area of catalytic asymmetric vinylogous aldol,6 Mannich,7 Michael8 and annulation reactions9 in the presence of either metal catalysts or organocatalysts (a, Scheme 1). These well-developed catalytic asymmetric methods have been related to the γ-functionalized α,β-unsaturated γ-butyrolactam to date. However, in sharp contrast, the approaches toward introducing C-3 chirality at the β-position of butyrolactam through a direct catalytic manner are underdeveloped (b, Scheme 1)10 in spite of the fact that β-selective chiral functionalization of butyrolactam can directly build up α,β-functionalized pyrrolidin-2-one frameworks.Open in a separate windowScheme 1Different reactive position of α,β-unsaturated γ-butyrolactam in catalytic asymmetric reactions.So far, only a few metal-catalytic enantioselective β-selective functionalized reactions have been reported. For examples, a rhodium/diene complex catalyzed efficient asymmetric β-selective arylation10a and alkenylation10b have been reported by Lin group (a, Scheme 2). Procter and co-workers reported an efficient Cu(i)–NHC-catalyzed asymmetric silylation of unsaturated lactams (b, Scheme 2).10c Despite these creative works, considerable challenges still exist in the catalytic asymmetric β-selective functionalization of γ-butyrolactam. First, the scope of nucleophiles is limited to arylboronic acids, potassium alkenyltrifluoroborates and PhMe2SiBpin reagents. Second, the catalytic system and activation mode is restricted to metal/chiral ligands. To our knowledge, an efficient catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach has not yet been established. Therefore, the development of organocatalytic asymmetric β-selective functionalization of γ-butyrolactam are highly desirable. In conjunction with our continuing efforts in building upon chiral precedents by using chiral tertiary amine catalytic system,11 we rationalized that the activated α,β-unsaturated γ-butyrolactam might serve as a β-position electron-deficient electrophile. This γ-butyrolactam may react with a properly designed electron-rich nucleophile to conduct an expected β-selective functionalized reaction of γ-butyrolactam under a bifunctional organocatalytic fashion, while avoiding the direct γ-selective vinylogous addition reaction or β,γ-selective annulation as outlined in Scheme 2. Herein we report the β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters12 catalyzed by a bifunctional chiral tertiary amine, which provides an efficient and facile access to optically active β-position functionalized pyrrolidin-2-one derivatives with both high diastereoselectivity and enantioselectivity.Open in a separate windowScheme 2β-Selective functionalization of γ-butyrolactam via metal- (previous work) or organo- (this work) catalytic approach.To begin our initial investigation, several bifunctional organocatalysts13 were firstly screened to evaluate their ability to promote the β-selective asymmetric addition of γ-butyrolactam 2a with cyclic imino ester 3a in the presence of 15 mol% of catalyst loading at room temperature in CH2Cl2 (entries 1–6, EntryCat.SolventYieldeerf11aCH2Cl270%40 : 6021bCH2Cl2<5%57 : 4331cCH2Cl270%65 : 3541dCH2Cl268%70 : 3051eCH2Cl258%63 : 4761fCH2Cl271%77 : 2371fDCE72%80 : 2081fCHCl370%80 : 2091fMTBE68%79 : 21101fToluene63%78 : 22111fTHF45%76 : 24121fMeOH32%62 : 3813b1fDCE : MTBE75%87 : 1314c1fDCE : MTBE72%87 : 1315d1fDCE : MTBE70%85 : 15Open in a separate windowaReaction conditions: unless specified, a mixture of 2a (0.2 mmol), 3a (0.3 mmol) and a catalyst (15 mmol%) in a solvent (2.0 mL) was stirred at rt. for 48 h.bThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1).cThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) for 24 h.dThe reaction was carried out in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) and 10 mol% of catalyst was used.eIsolated yields.fDetermined by chiral HPLC, the product was observed with >99 : 1 dr by 1H NMR and HPLC. Configuration was assigned by X-ray crystal data of 4a.The results of experiments under the optimized conditions that probed the scope of the reaction are summarized in Scheme 3. The catalytic β-selective asymmetric addition of γ-butyrolactam 2a with cyclic imino esters 3a in the presence of 15 mol% (DHQD)2AQN 1f was performed. A variety of phenyl-substituted cyclic imino esters including those bearing electron-withdrawing and electron-donating substituents on the aryl ring, heterocyclic were also examined. The electron-neutral, electron-rich, or electron-deficient groups on the para-position of phenyl ring of the cyclic imino esters afforded the products 4a–4m in 57–75% yields and 82 : 18 to 95 : 5 er values. It appears that either an electron-withdrawing or an electron-donating at the meta- or ortho-position of the aromatic ring had little influence on the yield and stereoselectivity. Similar results on the yield and enantioselectivities were obtained with 3,5-dimethoxyl substituted cyclic imino ester (71% yield and 91 : 9 er). It was notable that the system also demonstrated a good tolerance to naphthyl substituted imino ester (78% yield and 92 : 8 er value). The 2-thienyl substituted cyclic imino ester proceeded smoothly under standard conditions as well, which gave the desired product 4p in good enantioselectivity (88 : 12 er), although yield was slightly lower. However, attempts to extend this methodology to aliphatic-substituted product proved unsuccessful due to the low reactivity of the substrate 3q. It is worth noting that the replacement of Boc group with 9-fluorenylmethyl, tosyl or benzyl group as the protection, no reaction occurred. The absolute and relative configurations of the products were unambiguously determined by X-ray crystallography (4a, see the ESI).Open in a separate windowScheme 3Substrate scope of the asymmetric reaction of α,β-unsaturated γ-butyrolactam 2 to cyclic imino esters 3.a aReaction conditions: unless specified, a mixture of 2 (0.2 mmol), 3 (0.3 mmol) and 1f (15.0 mmol%) in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) was stirred at rt. bIsolated yields. cDetermined by chiral HPLC, all products were observed with >99 : 1 dr by 1H NMR and HPLC. Configuration was assigned by comparison of HPLC data and X-ray crystal data of 4a.We then examined the substrate scope of the imide derivatives (Scheme 4). Investigations with maleimides 4r–4u gave 48–61% yield of corresponding products as lower er and dr values than most of γ-butyrolactams. As for methyl substituted maleimides, the reaction failed to give any product.Open in a separate windowScheme 4Substrate scope of the asymmetric reaction of maleimides to cyclic imino esters.a aReaction conditions: unless specified, a mixture of 2 (0.2 mmol), 3 (0.3 mmol) and 1f (15.0 mmol%) in 2.2 mL a mixture of dichloroethane and methyl tert-butyl ether (volume ratio = 10 : 1) was stirred at rt. bIsolated yields. cDetermined by 1H NMR and chiral HPLC.The chloride product 4a ((R)-tert-butyl 4-((R)-3-((E)-(4-chlorobenzylidene)amino)-2-oxotetra hydrofuran-3-yl)-2-oxopyrrolidine-1-carboxylate) was recrystallized and the corresponding single crystal was subjected to X-ray analysis to determine the absolute structure. Based on this result and our previous work, a plausible catalytic mechanism involving multisite interactions was assumed to explain the high stereoselectivity of this process (Fig. 1). Similar to the conformation reported for the dihydroxylation and the asymmetric direct aldol reaction, the transition state structure of the substrate/catalyst complexes might be presumably in the open conformation. The acidic α-carbon atom of cyclic imino ester 3a could be activated by interaction between the tertiary amine moiety of the catalyst and the enol of 3avia a hydrogen bonding. Moreover, the enolate of 3a in the transition state might be in part stabilized through the π–π stacking between the phenyl ring of 3a and the quinoline moiety. Consequently, the Re-face of the enolate is blocked by the left half of the quinidine moiety. The steric hindrance between the Boc group of 2a and the right half of the quinidine moiety make the Re-face of 2a face to the enolate of 3a. Subsequently, the attack of the incoming nucleophiles forms the Si-face of enolate of 3a to Re-face of 2a takes place, which is consistent with the experimental results.Open in a separate windowFig. 1Proposed transition state for the reaction.In conclusion, we have disclosed the β-selective asymmetric addition of γ-butyrolactam with cyclic imino esters catalyzed by a bifunctional chiral tertiary amine, which provides an efficient and facile access to optically active β-position functionalized pyrrolidin-2-one derivatives with high diastereoselectivity and enantioselectivity. To our knowledge, this is the first catalytic method to access chiral β-functionalized pyrrolidin-2-one via a direct organocatalytic approach. Current efforts are in progress to apply this new methodology to synthesize biologically active products.  相似文献   

12.
Asymmetric vinylogous aldol addition of alkylidene oxindoles on trifluoromethyl-α,β-unsaturated ketones     
Simone Crotti  Giada Belletti  Nicola Di Iorio  Emanuela Marotta  Andrea Mazzanti  Paolo Righi  Giorgio Bencivenni 《RSC advances》2018,8(58):33451
A novel vinylogous aldol addition of alkylidene oxindole with 1-trifluoromethyl-3-alkylidene-propan-2-ones is presented. The reaction, catalyzed by a bifunctional tertiary amine, provides an efficient application of the vinylogous reactivity of oxindoles for the preparation of enantioenriched trifluoromethylated allylic alcohols.

The vinylogous aldol addition of alkylidene oxindole with 1-trifluoromethyl-3-alkylidene-propan-2-ones was developed. The reaction, provides straightforward access to enantioenriched trifluoromethylated allylic alcohols.  相似文献   

13.
Cs2CO3 catalyzed direct aza-Michael addition of azoles to α,β-unsaturated malonates     
Zi-Yu Jiang  Zhe-Yao Huang  Hong Yang  Lin Zhou  Qing-Han Li  Zhi-Gang Zhao 《RSC advances》2022,12(30):19265
A highly efficient method for the synthesis of azole derivatives via a direct aza-Michael addition of azoles to α,β-unsaturated malonates using Cs2CO3 as a catalyst, has been successfully developed. A series of azole derivatives have been obtained in up to 94% yield and the reaction could be amplified to gram scale in excellent yield in the presence of 10 mol% of Cs2CO3.

A highly efficient method for the synthesis of azole derivatives via a direct aza-Michael addition of azoles to α,β-unsaturated malonates has been successfully developed. A series of azole derivatives have been obtained in up to 94% yield.  相似文献   

14.
Formation of trisubstituted buta-1,3-dienes and α,β-unsaturated ketones via the reaction of functionalized vinyl phosphates and vinyl phosphordiamidates with organometallic reagents     
Petr Oeser  Jakub Koudelka  Hana Dvo&#x;kov  Tom&#x; Tobrman 《RSC advances》2020,10(58):35109
We studied the reactions of vinyl phosphates and vinyl phosphordiamidates containing an ester functional group with organometallic reagents. We found that the functionalized vinyl phosphates were smoothly converted into tri- and tetrasubstituted buta-1,3-dienes via the reaction with aryllithium reagents. Moreover, the vinyl phosphordiamidates were converted into α,β-unsaturated ketones using Grignard reagents. Based on the performed experiments, we proposed a reaction mechanism, which was confirmed by means of the isolation of key intermediates.

We studied the reactions of vinyl phosphates and vinyl phosphordiamidates containing an ester functional group with organometallic reagents.  相似文献   

15.
Enantioselective Michael addition of malonates to α,β-unsaturated ketones catalyzed by 1,2-diphenylethanediamine     
Wei Wang  Ling Ye  Zhichuan Shi  Zhigang Zhao  Xuefeng Li 《RSC advances》2018,8(73):41699
A general and highly enantioselective Michael addition of malonates to cinnamones and chalcones has been developed. The commercially available 1,2-diphenylethanediamine could be directly utilized as the organocatalyst to furnish the desired adducts in satisfactory yield (61–99%) and moderate to excellent enantiopurity (65 to >99% ee). β-Ketoester was also a competent donor and was employed to construct densely functionalized cyclohexenones via a tandem Michael-aldol condensation process.

1,2-Diphenylethanediamine could be directly utilized to promote the Michael addition of malonates and β-ketoesters to various α,β-unsaturated ketones.  相似文献   

16.
Correction: Synthesis of α,β-unsaturated esters of perfluoropolyalkylethers (PFPAEs) based on hexafluoropropylene oxide units for photopolymerization     
Cline Bonneaud  Mlanie Decostanzi  Julia Burgess  Giuseppe Trusiano  Trevor Burgess  Roberta Bongiovanni  Christine Joly-Duhamel  Chadron M. Friesen 《RSC advances》2020,10(30):17416
Correction for ‘Synthesis of α,β-unsaturated esters of perfluoropolyalkylethers (PFPAEs) based on hexafluoropropylene oxide units for photopolymerization’ by Céline Bonneaud et al., RSC Adv., 2018, 8, 32664–32671.

The authors regret that a consistent structure error appears in Schemes 4–6 and the graphical abstract of the original article. The definition shown for RF is incorrect, and the correct definition is RF = –CF(CF3)(OCF2CF(CF3))nOCF2CF2CF3. The correct versions of Schemes 4–6 are shown below, and the graphical abstract has been updated in the online version of the original article.Open in a separate windowScheme 4Synthesis scheme of maleate oligo(HFPO) using DCC and DMAP.Open in a separate windowScheme 5Reaction scheme of Steglich esterification with oligo(HFPO) methylene alcohol Mw ∼ 2000 g mol−1.Open in a separate windowScheme 6Reaction scheme of Steglich esterification on oligo(HFPO) alcohol Mw ∼ 1250 g mol−1.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

17.
Kinetics and thermodynamics of enzymatic decarboxylation of α,β-unsaturated acid: a theoretical study     
Phorntep Promma  Charoensak Lao-ngam  Rung-Yi Lai  Kritsana Sagarik 《RSC advances》2022,12(22):14223
Enzymatic decarboxylation of α,β-unsaturated acid through ferulic acid decarboxylase (FDC1) has been of interest because this reaction has been anticipated to be a promising, environmentally friendly industrial process for producing styrene and its derivatives from natural resources. Because the local dielectric constant at the active site is not exactly known, enzymatic decarboxylation to generate β-methylstyrene (β-MeSt) was studied under two extreme conditions (ε = 1 and 78 in the gas phase and aqueous solution, respectively) using the B3LYP/DZP method and transition state theory (TST). The model molecular clusters consisted of an α-methylcinnamate (Cin) substrate, a prenylated flavin mononucleotide (PrFMN) cofactor and all relevant residues of FDC1. Analysis of the equilibrium structures showed that the FDC1 backbone does not play the most important role in the decarboxylation process. The potential energy profiles confirmed that the increase in the polarity of the solvent could lead to significant changes in the energy barriers, especially for the transition states that involve proton transfer. Analysis of the rate constants confirmed the low/no quantum mechanical tunneling effect in the studied temperature range and that inclusion of the fluctuation of the local dielectric environment in the mechanistic model was essential. Because the computed rate constants are not compatible with the time resolution of the stopped-flow spectrophotometric experiment, the direct route for generating β-MeSt after CO2 elimination (acid catalyst (2)) is unlikely to be utilized, thereby confirming that indirect cycloelimination in a low local dielectric environment is the rate determining step. The thermodynamic results showed that the elementary reactions that involve charge (proton) transfer are affected by solvent polarity, thereby leading to the conclusion that overall, the enzymatic decarboxylation of α,β-unsaturated acid is thermodynamically controlled at high ε. The entropy changes due to the generation of molecules in the active site appeared more pronounced than that due to only covalent bond breaking/formation or structural reorientation. This work examined in detail for the first time the scenarios in each elementary reaction and provided insight into the effect of the fluctuations in the local dielectric environment on the enzymatic decarboxylation of α,β-unsaturated acids. These results could be used as guidelines for further theoretical and experimental studies on the same and similar systems.

The kinetically controlled path for enzymatic decarboxylation of α,β-unsaturated acid is proposed based on DFT and TST methods. The mechanism involves fluctuation of the local dielectric environment in the active site of the FDC1 enzyme.  相似文献   

18.
Lanthanide complexes combined with chiral salen ligands: application in the enantioselective epoxidation reaction of α,β-unsaturated ketones     
Xuexiu Xia  Chengrong Lu  Bei Zhao  Yingming Yao 《RSC advances》2019,9(24):13749
Readily available lanthanide amides Ln[N(SiMe3)2]3 (Ln = Nd (1), Sm (2), Eu (3), Yb (4), La (5)), combined with chiral salen ligands H2La ((S,S)-N,N′-di-(3,5-disubstituted-salicylidene)-1,2-cyclohexanediamine) and H2Lb ((S,S)-N,N′-di-(3,5-disubstituted-salicylidene)-1,2-diphenyl-1,2-ethanediamine) were employed in the enantioselective epoxidation of α,β-unsaturated ketones. It was found that the salen–La complex shows the highest efficiency and enantioselectivity. A relatively broad scope of α,β-unsaturated ketones was investigated, and excellent yields (up to 99%) and moderate to good enantioselectivities (37–87%) of the target molecules were achieved.

The enantioselective epoxidation of α,β-unsaturated ketones was catalysed by readily available lanthanide amides La[N(SiMe3)2]3 combined with chiral salen ligands.  相似文献   

19.
A facile metal-free one-flask synthesis of multi-substituted furans via a BF3·Et2O mediated formal [4 + 1] reaction of 3-chloro-3-phenyldiazirines and α,β-alkenyl ketones     
Zixin Zhang  Aimin Huang  Lin Ma  Jian-hua Xu  Min Zhang 《RSC advances》2022,12(24):15190
A facile, efficient and metal free one-flask approach to diversely substituted furans from easily accessible 3-chloro-3-phenyldiazirines and α,β-alkenyl ketones is reported. This protocol integrates three steps of cyclopropanation, Cloke–Wilson rearrangement and elimination of HCl in one-flask to give products in moderate to good yields. It provides a metal and oxidant free approach to multi-substituted furans with the advantages of easy operation, mild reaction conditions and a broad scope of substrates.

A metal and oxidant free approach to multi-substituted furans from easily accessible 3-chloro-3-phenyldiazirines and α,β-alkenyl ketones in one flask.  相似文献   

20.
Iodine/water-mediated deprotective oxidation of allylic ethers to access α,β-unsaturated ketones and aldehydes     
Yuntian Xue  Yaolong Yan  Kezhi Jiang  Weifeng Chen  Lei Yang 《RSC advances》2020,10(25):14720
The first iodine/water-mediated deprotective oxidation of allylic ethers to access α,β-unsaturated ketones and aldehydes was achieved. The reaction tolerates a wide range of functionalities. Furthermore, this protocol was found to be applicable to the oxidative transformation of allylic acetates. The proposed mechanism involves an oxygen transfer from solvent water to the carbonyl products.

The first iodine/water-mediated deprotective oxidation of allylic ethers to access α,β-unsaturated ketones and aldehydes was effectively achieved.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号