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 93 毫秒
1.
Solvent-free and solvent-less slipping-on of the dibenzo-24-crown-8 (DB24C8) over the N-hydroxysuccinimide end of an ammonium-containing thread has been studied and compared to the same reaction operated in solution. Slippage proved to be possible in solvent-free conditions, but the fastest slippage was obtained under heating when preliminary Liquid-Assisted Grinding (LAG) conditions were applied to the reactants followed by aging under an atmosphere of acetonitrile.

Very efficient slipping-on of the dibenzo-24-crown-8 over the NHS end of an ammonium-containing molecular axle was carried out through a solvent-less procedure.

The recently awarded 2016 Nobel Prize in chemistry1 has put a light on molecular machines.2 Some of these machines benefit from their interlocked molecular architecture3 so that the relative displacement of one interlaced element among others becomes possible and controllable. Hence, the straightforward synthesis of interlocked molecules is appealing in order to access novel molecular machines. Using the slippage strategy,4 we recently reported the preparation of an insulated and storable, albeit activated, N-hydroxysuccinimide (NHS) ester-containing [2]rotaxane building block (Scheme 1 and entry 1 of 5 This compound is a valuable activated building block for post-interlocking elongation of the encircled axle using bulky amino compounds.6 As the mechanism of such aminolysis reactions preserves the mechanical bond, it allowed the efficient and straightforward preparation of more sophisticated interlocked compounds such as [2] and [3]rotaxane molecular shuttles.5,6 Improving the access to the NHS ester-containing [2]rotaxane building block 2 is therefore of real interest. This is particularly justified since in acetonitrile solution, the slipping-on of the DB24C8 (3 equiv.) over the NHS extremity of an ammonium-containing thread (at a concentration of 3 × 10−2 M) is very slow and necessitates heating (13 days and 333 K, respectively). In this paper, we wondered if this slipping-on process could be possible, nay improved, by drastically reducing the amount of solvent. Since solvent-free/solvent-less conditions are highly prone to induce mass transfer limitations, utilisation of ball-milling was envisaged. Indeed it was previously shown that ball-milling could improve the speed of inherently slow reactions.7 A few examples of solvent-free/solvent-less synthesis of rotaxanes have been reported to date,8 and to the best of our knowledge, only three of these examples are related to slippage process through a co-melting process9 or an immediate solvent evaporation method.10 Herein, different experimental procedures were considered to yield the activated [2]rotaxane 2: solvent-free grinding,11 Liquid-Assisted Grinding (LAG),12 and aging by heating with or without an acetonitrile atmosphere.13 LAG is defined as the use of small amounts of a non-reactive liquid during grinding.14 It has been shown by us and by other research groups to have a considerable effect on the course of reactions run under mechanical forces.15 Besides, aging is the action of letting the reaction take place in the absence of any mechanical agitation. This reactivity is based on the inherent mobility of molecules and can be accelerated by the presence of vapours and/or a slight heating in the presence, or not, of a catalyst.16 Both LAG and aging are processes that are using generally much lower quantities of solvents compared to traditional solvent-based syntheses, and therefore are generally qualified as “solvent-less” processes. The results of these slipping-on experiments are shown in Fig. 1 and Open in a separate windowScheme 1Slippage process of the NHS ester-containing molecular axle 1 by the DB24C8.Experimental conditions relative to the slippage of thread 1
EntryProcedurePreliminary grindingLAGAging under an atm. of T (K) t (h)Conversion in 2 (%)
1In solutionNoNo33331272
2Solvent-freeNoNo2985053
3Solvent-lessYesYes29815
4Solvent-freeNoNoAr3335045
5Solvent-freeYesNoAr33350410
6Solvent-lessYesYesAr33350410
7Solvent-lessNoNoMeCN3337258
8Solvent-lessYesYesMeCN33314989
Open in a separate windowOpen in a separate windowFig. 1Kinetics comparison of the slipping-on relative to the formation of rotaxane 2 from 1. Corresponding experimental procedures are given in 17 the reaction proceeded slowly (312 h – 13 days – to reach the equilibrium, corresponding to a conversion of 72%) because of the bulkiness of the NHS ester with respect to the size of the inner diameter of the DB24C8 (entry 1, Fig. 1). The slippage was then attempted at room temperature under solvent-free conditions: after 1 equiv. of thread 1 have been mixed with a spatula for few minutes with 3 equiv. of DB24C8 then let reacting without any mixing under ambient conditions for 505 hours (21 days) (entry 2, Fig. 1).18 No significant improvement was noticed for the conversion yield in [2]rotaxane 2 nor for the time to reach the equilibrium of the reaction. Adding to this experiment a preliminary 1 h grinding period in the vibrating ball-mill afforded a slight improvement probably due to the more intimate mixing of the reactants (entry 5, Fig. 1). The same ratio of rotaxane 2 was observed if 0.2 μL of acetonitrile (per mg of reactants) is added to the ball-mill as a liquid-grinding assistant (entry 6, Fig. 1). For all these solvent-free/solvent-less experiments that were carried out at 333 K, the time to reach the equilibrium of the slippage reaction proved to be much longer than that realized in solution (entries 1 and 4–6, Fig. 1). However, it became lower if the solid mixture was heated at 333 K under an acetonitrile atmosphere (entries 7–8, Fig. 1). This procedure, called “aging”, has been recently described by other groups as a valuable, low-energy demanding and solvent-less alternative to classical synthesis in solution.19 In the absence of preliminary grinding, 58% of rotaxane 2 could be obtained after 72 h under an acetonitrile atmosphere (entry 7, 20 By directly comparing with the reaction operated in solution (entry 1), the ratio of rotaxane 2 was higher (89% vs. 72%) and most of all, the time to reach the equilibrium was more than twice lower (149 h vs. 312 h).In conclusion, slippage process of thread 1 by the DB24C8 was proved possible in the absence of solvent although producing [2]rotaxane 2 in very low yields (3–10%). In the absence of any acetonitrile atmosphere, the rate of the rotaxane formation was much slower than in solution and preliminary grinding did not afford significant improvement. On the opposite, the ratio of rotaxane as well as the time to reach the equilibrium of its formation could be tremendously enhanced when combining preliminary grinding with heating at 333 K under an acetonitrile atmosphere. Noteworthy, the aging-based solvent-less slippage, which occurred in the presence of a 216 times lesser quantity of acetonitrile than in solution, resulted in a time to reach the equilibrium more than twice shorter and an improved yield ratio of 24%. It demonstrates that combining ball-milling with accelerated aging is an easy and efficient protocol that should be envisaged to facilitate the inherently slow formation of other sophisticated interlocked molecules.  相似文献   

2.
A radical trifluoromethylation reaction of tertiary enamides was investigated and trifluoromethyl-containing isoindolinones were prepared under mild conditions. Using TMSCF3 as a radical source, PhI(OAc)2 as an oxidant and KHF2 as an additive, tertiary enamides were converted to isoindolinones via a cascade addition and cyclization process in moderate to good yields.

Radical trifluoromethylation and cyclization of tertiary enamides was developed and trifluoromethyl-containing isoindolinones were obtained in moderate to good yields.

In recent years, trifluoromethyl-containing azaheterocycles have attracted much attention for their potential application in the fields of pharmaceutical and agricultural chemistry.1 Thus, lots of efforts have been devoted to the synthesis of trifluoromethyl azaheterocycles,2 and among these developed methods, radical cascade addition and cyclization has emerged as a remarkable strategy due to its unique properties such as economy and high efficiency. Unsaturated amides are commonly used substrates for this type of transformation, which could be attacked by a CF3 radical followed by intramolecular C–O, C–N, or C–C bond formation to give different kinds of trifluoromethyl azaheterocycles. Fu reported a metal-free trifluoromethylation of N-allyamides with CF3SO2Na for the synthesis of trifluoromethyl-containing oxazolines via oxytrifluoromethylation.3 In the presence of copper salts, N-acyl-2-allylaniline could be converted to trifluoromethylated indolines in moderate to good yields via aminotrifluoromethylation process.4 With Togni''s reagent,5 TMSCF3,6 CF3SO2Na,7 CF3SO2Cl8 and other reagents9 as the CF3 source, α, β-unsaturated amides, tosyl amides, or imides underwent a tandem conversion to give trifluoromethyl-containing oxindoles or isoquinoline-1,3-diones by trifluoromethylation/arylation reaction. On the other hand, as a special type of unsaturated amide containing an active double bond, enamide also exhibited excellent reactivity in radical reactions.10 In fact, trifluoromethylation of enamides has already been investigated, and in most cases trifluoromethylated alkenes were obtained as the main products.11 To the best of our knowledge, the radical trifluoromethylation and cyclization of enamide still remains undeveloped.Isoindolinones are important N-heterocyclic compounds necessary in organic and pharmaceutical chemistry, and these compounds are used widely as anticoagulants and tranquilizers such as aristolactam, pagoclone, and zopiclone.12 To introduce a CF3 group into isoindolinones, Wang and co-workers explored a convenient way to the synthesis of trifluoromethyl-containing isoindolinones by radical aminotrifluoromethylation (Scheme 1a),13 but this transformation only occurred for N-methoxylbenzamides, and in case of N-alkylbenzamides trifluoromethylated alkenes were obtained as the major products. 1,1-disubstituted terminal alkenes were also not suitable substrates because of the competition between O-trapping and N-trapping process. Thus, development a new method for the synthesis of trifluoromethyl-containing isoindolinones is still in demand. Here in, as a continuation of our efforts on the radical modification of amide derivatives,14 we wish to present our work on the synthesis of trifluoromethyl-containing isoindolinones using enamides as the start materials by radical carbon trifluoromethylation (Scheme 1b).Open in a separate windowScheme 1Synthesis of trifluoromethyl-containing isoindolinones.Initially, N-n-butyl-N-(2-propenyl) benzamide 1a was chosen as the model substrate to optimize the reaction conditions of this radical carbontrifluoromethylation process. As shown in
EntryAdditive (0.3 equiv.)Solvent (2 mL)Temp. (°C)Yield of 2ab (%)
1NaFEtOAc8015
2KFEtOAc8038
3CsFEtOAc8035
4NaHF2EtOAc8052
5KHF2EtOAc8075
6NH5F2EtOAc8040
7KHF2CH3CN8021
8KHF2CH2Cl28032
9KHF2Toluene80Trace
10KHF2EtOAc10061
11KHF2EtOAc6043
12KHF2EtOAcr.t.NR
13KHF2EtOAc8037c
14KHF2EtOAc8062d
15KHF2EtOAc8058e, 47f
16KHF2EtOAc8073g
Open in a separate windowaThe reaction was carried out on 0.2 mmol scale in a sealed tube under N2.bIsolated yield.cPhI(OCOCF3)2 (4.0 equiv.) was used as the oxidant.dReaction carried out with PhI(OAc)2 (3.0 equiv.) and TMSCF3 (3.0 equiv.).eWith 1.5 equiv. KHF2.fWith 0.5 equiv. KHF2.g1.0 equiv. NaOAc was added.Under the optimized reaction conditions, the scope of substrates was investigated with results summarized in Open in a separate windowaThe reaction was performed with 1 (0.2 mmol), KHF2 (0.2 mmol), TMSCF3 (0.8 mmol), PhI(OAc)2 (0.8 mmol) in EtOAc (2.0 mL) under N2 at 80 °C for 12 h in a sealed tube.bIsolated yields.When N-n-butyl-N-(2-vinyl) benzamide 3a was subjected to the reaction conditions, no isoindolinone was observed, and the main product was trifluoromethylated alkene ( Open in a separate windowaThe reaction was performed with 3 (0.2 mmol), KHF2 (0.2 mmol), TMSCF3 (0.8 mmol), PhI(OAc)2 (0.8 mmol) in EtOAc (2.0 mL) under N2 at 80 °C for 12 h in a sealed tube.bIsolated yields.Heterocyclic substrate such as 5a and 5b was also examined, but no cyclization product could be found and trifluoromethylated alkene 6a and 6b was obtained as the only product (Scheme 2).Open in a separate windowScheme 2. Results of heterocyclic substrate 5a and 5b.To gain insights into the reaction mechanism, a control experiment was carried out to elucidate the mechanism. When 1.0 equiv. TEMPO was added to the reaction, the yield of 2a decreased significantly to 15%, which indicated the possibility of a radical pathway. Based on the control experimental result and the previous investigation on aryltrifluoromethylation of alkenes, plausible mechanism for our methodology is proposed in Scheme 2. In the presence of KHF2, TMSCF3 reacted with PhI(OAc)2 to generate CF3 radical, then the CF3 radical attacked enamide 1 or 3 affording radical intermediate A. Depending on the structure of the substrate, intermediate A would be converted to trifluoromethyl-containing isoindolinone or trifluoromethylated alkene according to different pathways as followed: (path a) intramolecular cyclization of A gave the resulting radical B with an aryl ring, which was oxidized to intermediate C then underwent deprotonation to give rise to the final product 2a–r or 4; (path b) A was oxidized to intermediate D then underwent elimination to give trifluoromethylated alkene 2s, 4′ or 6 (Scheme 3).Open in a separate windowScheme 3Possible mechanism.  相似文献   

3.
N-Heterocyclic carbene (NHC)-catalyzed oxidation of unactivated aldimines to amides via imine umpolung under aerobic conditions     
Jakkula Ramarao  Sanjay Yadav  Killari Satyam  Surisetti Suresh 《RSC advances》2022,12(13):7621
Herein, we disclose an NHC-catalyzed aerobic oxidation of unactivated aldimines for the synthesis of amides via umpolung of imines proceeding through an aza-Breslow intermediate. We have developed an eco-friendly method for the conversion of imines to amides by using molecular oxygen in air as the sole oxidant and dimethyl carbonate (DMC) as a green solvent under mild reaction conditions. Broad substrate scope, high yields and gram scale syntheses expand the practicality of the developed method.

A general NHC-catalyzed conversion of imines to amides proceeding through umpolung–oxidation under aerobic conditions in green solvent is reported.

NHCs have emerged as an important class of organocatalysts for unconventional organic transformations due to their unique property of umpolung i.e. essentially reversing the polarity of electrophilic carbon centers. NHC-catalyzed umpolung reactivity has been extensively exploited over the past two decades for the construction of carbon–carbon and carbon–heteroatom bonds.1 NHC-catalytic transformations proceeding through umpolung reactivity, via Breslow intermediate, have been well established using aldehydes.1,2 Despite imines being considered to be potentially important building blocks, the obvious use of imines as reactants for the same is still an under developed area, probably due their lower reactivity. The umpolung of imine came into existence after the isolation of a nitrogen containing Breslow intermediate (known as aza-Breslow intermediate) by the Douthwaite and Rovis groups from the reaction of stoichiometric amounts of NHC and imine or iminium salt, respectively.3 For the first time, our group4 and the Biju group,5 independently, reported the NHC-catalyzed imine umpolung transformations. Recently, a few other groups including our group,6 the Wei–Fu–Huang,7 Biju,8 Lupton9 and Tian–Zhang–Chi10 groups reported NHC-catalyzed imine C–H functionalization or oxidation of imines. In continuation of our ongoing research on NHC-catalyzed umpolung transformations,4,6,11 herein, we present the development of NHC-catalyzed aerobic oxidation of aldimines to amides, via imine umpolung without using any additive. We became interested to access the amide functionality because amide is a very crucial functional group due its ubiquitous presence in life-processes in the form of peptide bond in protein molecules as well as its appearance in several of the drug molecules.12 For example, imatinib13a and ponatinib13a are used in the chemotherapy treatment of chronic myelogenous leukemia (CML) in cancer disease (Fig. 1). Betrixaban containing two amide functional groups is an oral anticoagulant drug (Fig. 1).13bOpen in a separate windowFig. 1Selected amide containing therapeutic molecules.The scientific community has been showing great interest to develop new and efficient methods for the construction of amide bond. Coupling of carboxylic acid with amine is one of the most common methods used for the construction of amide molecule.14 However, this method requires stoichiometric amounts of peptide coupling reagents such as carbodiimides and 1-hydroxy benzotriazoles or activated carboxylic acid derivatives.15 The Schmidt reaction16 and Beckmann rearrangement17 are classical examples for the synthesis of amides. However, there are considerably a few reports available for the oxidation of imine to amide. Palladium-catalyzed oxidation of imines to amides was reported by using excess tert-butyl hydroperoxide (TBHP) as an oxidant.18 Methods for the oxidation of imines to amides were reported by using peroxy acids in the presence of strong Lewis acid and Brønsted acid, which generate stoichiometric amount of by-products.19 Cheon and co-workers reported the oxidation of imines to amides by using sodium cyanide (NaCN) in stoichiometric amounts. High toxic nature of NaCN is the limitation of this methodology (Scheme 1a).20a Recently, Fu and Huang group reported the oxidation of imines, limited to the imines derived from heteroaryl amines, to amides by using NHC catalysis with the assistance of excess lithium chloride as Lewis acid (Scheme 1b).7a Recently, we reported an NHC-catalyzed tandem aza-Michael oxidation of β-carboline cyclic imines in the presence of external Michael acceptors (Scheme 1c).6 Besides the limitations associated with the above transformations, those were performed in non-green solvent media. However, to the best of our knowledge, there are no reports for the conversion of imines to amides under NHC-catalysis without using an external additive/assistance.21 On the other hand, the development of new methods to access amide functionality inclusive of green chemistry principles such as organocatalysis, air as the sole oxidant and use of green solvent medium under ambient conditions is highly desirable. Herein, we report an NHC-catalyzed conversion of aldimines to amides, proceeding through imine umpolung–oxidation, in the presence of air in a green solvent such as DMC22 under mild conditions.Open in a separate windowScheme 1Prior work and this work.Initially, we began our investigation with the reaction of aldimine 3a, derived from benzaldehyde 1a and aniline 2a, in the presence of NHC A1 catalyst under open air conditions at room temperature in a green solvent such as DMC. Gratifyingly, we observed the formation of the corresponding amide 4a in 53% yield ( for an extensive optimization survey). We conducted an experiment in presence of LiCl (1.2 equiv.), and it did not help to improve the yield of the product.Optimization study
EntryaNHC precatalystBaseSolventYield of 4ab
1A1Cs2CO3DMC53
2B1Cs2CO3DMC
3 C1 Cs 2 CO 3 DMC 87
4D1Cs2CO3DMC
5C1DBUDMC71
6C1DABCODMC65
7C1NaHDMC72
8C1K2CO3DMC55
9C1Cs2CO3THF75
10C1Cs2CO3EtOAc65
11C1Cs2CO3DMSO63
12C1Cs2CO3EtOH
13C1Cs2CO3DMC70c
14C1Cs2CO3DMC72d
15Cs2CO3DMC
16C1DMC
Open in a separate windowaReaction conditions: 3a (0.5 mmol), NHC precatalyst (0.1 mmol), base (0.6 mmol), solvent (4 mL).bYields are of pure compounds after crystallization.cWith 0.075 mmol of C1.dWith 0.5 mmol of Cs2CO3; Mes: 2,4,6-trimethylphenyl; DBU: 1,8-diazabicyclo[5.4.0]undec-7-ene; DABCO: 1,4-diazabicyclo[2.2.2]octane; DMSO = dimethyl sulfoxide.By choosing the acceptable optimized conditions from Scheme 2).Open in a separate windowScheme 2Sequential imine formation–NHC-catalyzed aerobic oxidation to access amide 4a.We then examined the scope of the NHC-catalyzed imine umpolung–oxidation to access amides under aerobic conditions in DMC at room temperature. Firstly, the reaction of variously substituted aromatic, heteroaromatic and vinyl aldehydes 1 were converted to the corresponding aldimines 3 with aniline 2a. Subsequently the aldimine 3, without further purification, was subjected to optimized NHC catalysis conditions (Scheme 3). Imines derived from aromatic aldehydes bearing either electron-withdrawing or electron-donating groups smoothly afforded the corresponding substituted amides 4 in high yields. The imines derived from ortho-/para-halo-substituted benzaldehydes provided the amides 4b–f in high yields. It was interesting to note that sterically hindered 2,6-dichlorobenzaldehyde derived imine also provided the corresponding amide 4g in 70% yield. Aldimines bearing mono-/di-substituted electron-donating groups provided the respective amides 4h–k in good yields. The aldimines containing both electron-donating and halogen substituents furnished the corresponding amides 4l and 4m in 72% and 74%, yields, respectively. Imines having electron-withdrawing functional groups such as NO2 or CN also shown tolerance to afford their amides 4n–p in 73–79% yields. The naphthaldehyde imine provided its amide 4q in 78% yield. We also tested the imines derived from heterocyclic aldehyde such as 2-furaldehyde and α,β-unsaturated aldehyde such as cinnamaldehyde in this transformation to produce the corresponding amides 4r and 4s in 64% and 55%, yields, respectively. The yield of 4s was only slightly increased with 40 mol% NHC C1.Open in a separate windowScheme 3Scope of the reaction starting with different aldehydes.To further study the substrate scope of the NHC-catalyzed imines to amides, benzaldehyde imines derived from variously substituted aromatic/heteroaromatic amines were tested in this transformation (Scheme 4). The imines bearing mono-/di-substituted halogen groups on the aniline side gave the corresponding amides 4t–y in high yields. The imines containing electron-donating and electron-withdrawing groups on the aniline side were also smoothly converted to their respective amides 4z-4ab in good yields. Imines derived from heteroaromatic amines such as 3-aminoquinoline and 2-aminobenzothiazole furnished the corresponding amides 4ac and 4ad in 77% and 80% yields, respectively.Open in a separate windowScheme 4Scope of the reaction starting with different amines.We also tested the imines derived from substituted benzaldehyde and substituted aniline to give the desired amide 4ae in 72% yield (Scheme 5). Later, an imine derived from heteroaromatic aldehyde and heteroaromatic amine was subjected to NHC-catalyzed oxidation to afford the respective amide 4af in 70% yield (Scheme 5). We also conducted the reactions with imines derived from aliphatic aldehyde/aliphatic amine, however, the corresponding amide formation was not observed. The reason may be attributed to the less reactivity of these imines.Open in a separate windowScheme 5Additional scope of the reaction.The practicality of this transformation was tested by gram-scale syntheses on 10 mmol scales; the NHC-catalyzed aerobic oxidation of imines 3a, 3n, 3v proceeded smoothly to afford the corresponding amides 4a, 4n, 4v in 70%, 65%, 66% yields, respectively (Scheme 6).Open in a separate windowScheme 6Gram-scale syntheses of 4a, 4n and 4v.We have performed a few control experiments to know the requirement of molecular oxygen for the NHC-catalyzed oxidation of imine to amide. Accordingly, we conducted a reaction under inert conditions and observed a drastic decrease in the yield of the product (Scheme 7a).6 This result indicate that molecular oxygen in air is responsible and acting as the sole oxidant in this transformation. We also conducted the NHC-catalyzed imine 3a oxidation in the presence of pure oxygen to give the desired amide 4a in 84% yield (Scheme 7b). We further conducted a direct reaction of 1a and 2a under optimized NHC-catalyzed conditions to know whether oxidation of 1a provide NHC-azolium intermediate and add to the amine, akin to the NHC-catalyzed ester formation from the reaction of aldehydes and alcohols.23 However, in this reaction we did not observed the formation of amide but obtained the compound 520b resulted from the benzoin condensation-acylation (Scheme 7c).Open in a separate windowScheme 7Control experiments.Based on the previous literature reports,6,7,24 a possible mechanism for the NHC-catalyzed aerobic oxidation of imine to amide is depicted in Scheme 8. Initially, the free NHC would add to the imine 3 to form intermediate I. Aza-Breslow intermediate II would generate from intermediate I upon proton shift. The intermediate II would react with molecular oxygen and undergo single electron transfer of the intermediate II with dioxygen followed by radical recombination to give intermediate III.7,24 Thereafter one more molecule of aza-Breslow intermediate II would react with intermediate III to produce two molecules of intermediate IV. Then from intermediate IV NHC would regenerate and produce the amide 4. This mechanism suggests that one molecule of oxygen is sufficient to produce two molecules of the amide 4.Open in a separate windowScheme 8Plausible mechanism.  相似文献   

4.
Direct access to multi-functionalized benzenes via [4 + 2] annulation of α-cyano-β-methylenones and α,β-unsaturated aldehydes     
Qianfa Jia  Yunfei Lan  Xin Ye  Yinhe Lin  Qiao Ren 《RSC advances》2020,10(49):29171
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 (%)1Cs2CO3Toluene24702Na2CO3Toluene24423K2CO3Toluene24384NaOHToluene12785NaOAcToluene24526KOHToluene12747K3PO4Toluene24588DBUToluene24339Et3NToluene484610NaOHDCM1288 11 NaOH CHCI 3 12 94 12NaOHDCE128413NaOHH2O48014cNaOHCHCI3128515dNaOHCHCI3128416eNaOHCHCI31280Open 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   

5.
Visible-light synthesis of 4-substituted-chroman-2-ones and 2-substituted-chroman-4-ones via doubly decarboxylative Giese reaction     
Marek Moczulski  Ewelina Kowalska  El bieta Ku mierek  &#x;ukasz Albrecht  Anna Albrecht 《RSC advances》2021,11(45):27782
  相似文献   

6.
Palladium-catalyzed cross-coupling reaction of alkenyl aluminums with 2-bromobenzo[b]furans     
Chang Wen  Xin Jiang  Kun Wu  Ruiqiang Luo  Qinghan Li 《RSC advances》2020,10(33):19610
Highly efficient and simple cross-coupling reactions of 2-bromobenzo[b]furans with alkenylaluminum reagents for the synthesis of 2-alkenylbenzo[b]furan derivatives using PdCl2 (3 mol%)/XantPhos (6 mol%) as catalyst are reported. Excellent yields (up to 97%) were obtained for a wide range of substrates at 80 °C for 4 h in DCE.

PdCl2 (3 mol%)/XantPhos (6 mol%) complexes was found to be a highly efficient catalyst for the synthesis of 2-alkenylbenzo[b]furans from 2-bromobenzo[b]furans and alkenylaluminums. The reaction was also found to be effective in gram-scale synthesis.

2-Substituted benzo[b]furans are important structural scaffolds found in many natural products and pharmaceutical products.1 Some of these compounds have been known to exhibit anti-inflammatory,2 antitumor,3 anticancer,4 and anti-fungal,5 antiplasmodial,6 antioxidant,7 anti-HIV,8 and estrogenic activities.9 In addition, they serve as building blocks for many organic transformations.10 Thus, their synthesis and applications have attracted considerable attention in the chemical and pharmaceutical industries over the past decades.11 Until now numerous effective synthetic methodologies of synthesis 2-substituted benzo[b]furans have been reported.12,13 Among these methods hitherto developed, the metal-catalyzed 2-halobenzo[b]furans coupling with organometallic nucleophiles is one of the most effective methods (Scheme 1).13 However, in most cases generally suffer from one or more drawbacks such as requirement co-catalyst like Cu salts, limited substrate scope, high catalyst loading, high temperature and poor chemoselectivity etc. Therefore, the development of more efficient and atom economical approaches for the preparation of 2-substituted benzo[b]furans remains as desirable work. Previous studies show that organoaluminum reagents are highly efficient nucleophiles for cross-coupling reactions with aromatic halides14 or benzylic halides,15 and the investigations have demonstrated that palladium is a good catalytic metal.16Open in a separate windowScheme 1Palladium-catalyzed cross-coupling reactions of 2-halobenzo[b]furans derivatives with organometallic nucleophiles.At present, a variety of methods have been developed to prepare compounds containing olefin functional groups through hydrocarbon activation of olefins.17 To continue our effort to develop coupling reactions using reactive organoaluminum reagents,18 we herein report a palladium(ii)-catalyzed, base free cross-coupling reactions of 2-bromo benzo[b]furans with alkenylaluminum reagents at 80 °C in short reaction time with good to excellent isolated yields for 2-alkenyl benzo[b]furans. The process was simple and easily performed, and it provides an efficient method for the synthesis of 2-alkenyl benzo[b]furans derivatives. Notably, in our procedure palladium is used as the single catalyst and base free.Our initial studies used 2-bromo-6-methoxybenzo[b]furan (2e) with diethyl(oct-1-enyl)aluminum (1a) as model substrates. Treatment of compound 2e with the alkenylaluminum (1a) using PdCl2 (3 mol%)/XantPhos (6 mol%) as catalyst in toluene at 60 °C for 4 h, the coupled product 6-methoxy-2-(oct-1-enyl)benzo[b]furan (3ae) was obtained in 46% isolated yield ( EntryPd salt.1a (equiv.)Base (x equiv.)Solvent3aeb yield (%)1PdCl21.0—Toluene462Pd(OAc)21.0—Toluene193Pd(acac)21.0—Toluene104PdCl21.0Et3N (2.0)Toluene275PdCl21.0K2CO3 (2.0)TolueneNR6PdCl21.0TMEDA (2.0)TolueneNR7PdCl21.0—Hexane478PdCl21.0—THF519PdCl21.0—DCE7410cPdCl21.0—DCE8511cPdCl20.8—DCE8412cPdCl20.6—DCE5313c,dPdCl20.8—DCE49Open in a separate windowa1a/2a/PdCl2/XantPhos = 1.0/0.5/0.03/0.06 mmol, 60 °C, 3 mL solvent, 4 h, Ar2.bIsolated yield of 3ae.c80 °C.d1a/2a/PdCl2/XantPhos = 0.8/0.5/0.02/0.04 mmol.Under the optimized conditions, coupling reactions of aliphatic alkenylaluminum reagents, such as di-sec-butyl(oct-1-enyl)aluminum (1a) and di-sec-butyl(dec-1-enyl)aluminum (1b), proceed with electron-neutral, electron rich and electron-deficient 2-bromo benzo[b]furans derivatives affording the products in good yields ( Open in a separate windowa1/2/PdCl2/XantPhos = 0.8/0.5/0.03/0.06 mmol, 80 °C, 4 h. Isolated yield of 3, two runs.The reaction was also found to be effective in gram-scale synthesis, which indicated its potential for practical application (Scheme 2). 2-Substituted benzo[b]furans derivatives 3ah was synthesized in 1.31 gram using this methodology.Open in a separate windowScheme 2Preparative scale synthesis of compound 3ah.In order to further explore the reaction mechanism, control experiments were carried out (see the ESI). We performed the reaction between 2-bromobenzo[b]furan (2a, 0.5 mmol) with di-sec-butyl(oct-1-enyl)aluminum (1a, 0.8 mmol) in the presence of PdCl2 (3 mol%)/XantPhos (6 mol%) in DCE at 80 °C for 4 h. The reaction mixture was analyzed by 31P NMR, it was found that the characteristic peak of 31P NMR appeared around at 22.98 ppm and 30.98 ppm. However, 31P NMR peak of pure XantPhos is −18.03 ppm. The results show that XantPhos work as a ligand of the palladium center. Thus, a proposed possible reaction mechanism for the cross-coupling reaction is shown in Scheme 3. The first step is the oxidative addition of 2-bromobenzo[b]furans (2) to Pd(0) phosphine complex (4) (which in turn from PdCl2 and RAlMe2 (1) reagents) to form the organopalladium(ii) bromide intermediate (5). Transmetalation of RAlMe2 (1) with complex 5 gives R''PdR(ii) intermediate (6) and Me2AlBr. Finally, complex 6 under goes reductive elimination to afford the desired coupling product of 2-alkenylbenzo[b]furans (3) and regenerate the active Pd(0) species for the next catalytic cycle.Open in a separate windowScheme 3The proposed mechanism for the formation of coupled product 3.  相似文献   

7.
Highly dispersed Ni nanoparticles on mesoporous silica nanospheres by melt infiltration for transfer hydrogenation of aryl ketones     
Hyemin Kweon  Sanha Jang  Akerke Bereketova  Ji Chan Park  Kang Hyun Park 《RSC advances》2019,9(25):14154
Nickel-based catalysts have been applied to the catalytic reactions for transfer hydrogenation of carbonyl compounds. In the present work, highly dispersed nickel particles located at the pores of mesoporous silica spheres (Ni@mSiO2) were prepared via an optimized melt infiltration route. The nickel nanoparticles of 10 wt% in the Ni@mSiO2 catalyst could be uniformly loaded with high dispersion of 36.3%, resulting excellent performance for catalytic transfer hydrogenation of aryl ketones.

Nickel-based catalysts have been applied to the catalytic reactions for transfer hydrogenation of aryl ketones.

The reduction of carbonyl compounds has received steady interest as a common transformation in organic synthetic chemistry. Corresponding alcohols are important building blocks for the manufacture of chemicals, pharmaceuticals, and cosmetics.1,2 The transfer hydrogenation (TH) reaction of ketones has lots of advantages owing to its low cost, simple process, easy handling, and mild conditions.3 Isopropyl alcohol both as a solvent and hydrogen donor provides safe reaction conditions.4 Until now, many catalysts have been developed for TH reactions.5Among the catalysts for TH reactions, nickel-based catalysts are representative, due to the metallic nickel surface can absorbs hydrogen and easily activates hydrogen in the atomic state.6–11 In addition, nickel compounds could be the significant catalyst candidate because of the low price, accessibility, and high reactivity.12,13 Therefore, several works used nickel nanocatalysts for TH reactions.14–17,44–51 However, developing the new supported catalyst with the high metal dispersion and narrow particle size distribution is still required, because it can enhance the reactant conversion and improve product yield.18 Recently, melt infiltration process has been exploited to prepare supported nanocatalysts as a fast and convenient route with no solvent use.19–21The method which uses highly porous supports with regular porosity and hydrated metal salts with mild melting temperatures, allowed the metal salts to permeate inside the porous support via capillary forces. High performance nanocatalysts based on well-dispersed nanoparticles could be obtained through uniform infiltration of metal precursors and sequential thermal treatment.22 The obtained nanoparticles have narrow size distributions as well as small sizes, enlarging active sites.23,24Mesoporous silica nanosphere (mSiO2) as a porous support can be an outstanding platform for the synthesis of metal nanoparticle, owing to the good thermal stability, high surface area, and large pore volume as well as uniform pore structures.25–29 Herein, we report a new catalyst with highly dispersed Ni nanoparticles at mesoporous silica sphere nanostructures (Ni@mSiO2) for catalytic TH reactions, which was synthesized through a melt infiltration of hydrated nickel salts and a sequential thermal reduction (Scheme 1). First, mSiO2 supports could be prepared by a sol–gel method based on a modified Stöber process. More details were described in the ESI. The mesopores were formed in the silica nanospheres using the cationic surfactant C16TAB as a self-assembly template. The TEM images show mesoporous silica spheres with an average diameter of 417 ± 20 nm (Fig. 1a and b). Ordered mesoporous channels were generated by removing the long carbon chains of C16TAB by calcination at 500 °C. The corresponding Fourier-transform (FT) pattern showed a ring pattern of polycrystalline (inset of Fig. 1c). The inverse fast Fourier-transform (IFFT) image, calculated by Micrograph TM Gatan software, also demonstrated the existence of hexagonal lattice planes (Fig. 1c).Open in a separate windowScheme 1A brief synthetic scheme of Ni@mSiO2 nanocatalyst.Open in a separate windowFig. 1(a and b) TEM images, (c) FT pattern for (b) (inset of c) and IFFT image, (d) HRTEM image, (e) N2 adsorption/desorption isotherms, and (f) pore size distribution diagrams of mSiO2. The bars represent 1 mm (a), 100 nm (b), and 20 nm (d).The brightest of the rings originated from a two-dimensional hexagonal (p6mm) structure with d100 spacing. The interplanar spacing computed from the brightest FFT diffraction ring was approximately 3 nm, corresponding to the pore sizes shown in the TEM image (Fig. 1d).N2 sorption experiments for the mSiO2 nanosphere exhibited type IV adsorption–desorption hysteresis with H4 type hysteresis loop (Fig. 1e). The Brunauer–Emmett–Teller (BET) surface area and pore volume of the mSiO2 nanosphere were 1501 m2 g−1 and 0.70 cm3 g−1, respectively. The average pore size was estimated to be approximately 2 nm from the adsorption/desorption branches of N2 isotherms by using the Barrett–Joyner-Halenda (BJH) method (Fig. 1f). Scheme 1 illustrated the simple procedure for the synthesis of the Ni@mSiO2 nanocatalyst. Based on the small pore size (3 nm) of mSiO2 nanospheres, very tiny nickel nanoparticles (∼2 nm) could be generated via a melt-infiltration process and thermal treatment under hydrogen flow. Using a 0.55 gnickel salt/gsilica support ratio, the hydrated nickel salt (melting point = 56.7 °C) was successfully incorporated in the mSiO2 nanosphere during the melt infiltration process, driven by the capillary forces. The Ni-loading content after the final thermal treatment was calculated to be nominally 10 wt% on the basis of Ni converted from the nickel nitrate salt. First, the hydrated nickel salt was melt-infiltrated into the mesoporous silica pores of the mSiO2 nanospheres by physical mixing at room temperature with subsequent aging at 60 °C for 24 h in a tumbling oven. Then, tiny nickel nanoparticles were generated by thermal reduction of the confined hydrated Ni salt in the small pores at 500 °C under hydrogen flow.The low-resolution TEM images of the Ni@mSiO2 nanostructure show Ni nanoparticles as black dots (Fig. 2a and b). The high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image showed bright dots of a very small size (average 2.0 nm), which indicates uniform incorporation of the Ni nanoparticles in the porous silica (Fig. 2c and S1 in the ESI). In the elemental mapping of silica (green color) and nickel (red color), high dispersion of nickel nanoparticles was observed (Fig. 2d). HRTEM analysis showed lattices of a single Ni nanoparticles in the mSiO2 pores (Fig. 2e). The inner particle size was observed to be around 2 nm. The Fourier-transform pattern represented a single crystal of metallic nickel with a distance of 0.2 nm between neighboring fringes, which corresponded to the (111) planes of face centered-cubic nickel (Fig. 2e).Open in a separate windowFig. 2(a and b) TEM and (c) HAADF TEM images, (d) scanning TEM image with elemental mapping and (e) HRTEM images, and (f) XRD spectrum of Ni@mSiO2 nanostructure. The bars represent 200 nm (a), 20 nm (b–d), and 2 nm (e).The X-ray diffraction (XRD) spectrum of Ni@mSiO2 nanocatalyst has a broad peak at 2θ = 44.5°, which is assigned to the reflections of the (111) plane in the fcc-nickel phase (Fig. 2f, JCPDS No. 04-0850). The average crystal size of the nickel particle was estimated to be 1.8 nm, from the broadness of the (111) peak using the Debye–Scherrer equation, which is well-matched with that observed in the TEM images. The other intense peak around 23° of the Ni@mSiO2 nanocatalyst indicates the presence of amorphous silica.Using a H2 chemisorption experiment, the active nickel surface area and average nickel particle size of the Ni@mSiO2 nanocatalyst could be analyzed, and measured 242 m2 g−1 and 2.78 nm, respectively. The obtained Ni dispersion was very high, and was calculated to be 36.3%.The XPS spectrum analysis was carried out to probe the chemical states of Ni@mSiO2 (Fig. S2). The Ni 2p peak of Ni@mSiO2 showed NiO and Ni peak. The four peaks at 853.9, 855.6, 860.7, and 863.5 eV are assigned to NiO, and the other peak at 852.1 eV is assigned metallic Ni peak. The metallic Ni surface of the small nanoparticles (∼2 nm) was easily oxidized under an ambient condition.N2 sorption experiments for the Ni@mSiO2 nanocatalyst showed a type IV isotherm with type H4 hysteresis (Fig. 3a). The BET surface area was calculated to be 636 m2 g−1. The total pore volume was found to be 0.3 cm3 g−1, which is about 43% of the initial mSiO2 nanosphere (0.7 cm3 g−1). The significant decrease in pore volume was attributed to inner nickel nanoparticle occupancy. Applying the BJH method, small pore sizes were obtained by the adsorption/desorption branches (Fig. 3b). Because of the occupancy of tiny nickel particles in the pristine silica pores, the pore size of the Ni@mSiO2 nanocatalyst was also slightly decreased, and was observed to be 1.8 nm.Open in a separate windowFig. 3(a) N2 adsorption/desorption isotherms and (b) pore size distribution diagrams of Ni@mSiO2 nanocatalyst.The Ni@mSiO2 nanocatalyst was applied to the hydrogen-transfer reaction of acetophenone. Acetophenone is an ideal substrate in the hydrogen transfer reaction, and is generally used as a hydrogen acceptor.30 Among the various products of acetophenone reduction, only 1-phenylethanol resulted from the catalyzed reaction. The metallic Ni on the catalyst surface is the active species and the reaction was promoted by base.17 The dihydride species referred in this catalyst system to make alcohol from the transfer of the two hydrogen atoms of the donor to the surface of the metal.31 However, small amount of various byproduct including hemiacetals and condensation products which is occurred between ketone and alcohol were produced due to basic conditions.32 For optimization, the reaction parameters, such as amount of catalyst and base, temperature, and time, are adjusted. The TH reaction results of acetophenone are summarized in EntryCat. (mol%)Temp. (°C)Time (min)Base (eq.)Conv.b (%)Yieldb (%)TON1111045110095952111060176737331100602100999941100601100929251100600.562595961906011009696718045158555581806017268689180751100979710180901676565110.58075110096192120.25807511095380130.1807516260603140.05100602806262Open in a separate windowaRxn. condition: acetophenone (2 mmol), i-PrOH (solvent, 10 mL), base (NaOH).bDetermined by GC-MS spectroscopy.cnickel-aluminium alloy purchased from Lancaster (10034177) was applied.First, we studied the effect of the base and determined the amount of base required (Entries 3–6, 33,34 Although no reaction was observed without bases, a small amount of base was sufficient to trigger the reaction (Entry 5, 6,15,35 The optimized conditions were applied to extend the scope.To verify the general applicability of Ni@mSiO2, various aromatic ketones were tested ().36–43Catalytic transfer hydrogenation reactions of various aromatic carbonyl compounds with Ni@mSiO2a
EntryCompoundConv. (%)Sel. (%)Yield (%)
1 78b, 100c, 10033b, 75c, 6526b, 75c, 65
2 54, 68d, 44e92, 97d, 100e50, 66d, 44e
3 55, 64d57, 55d31, 35d
4 738864
5 8, 35f88, 97f7, 34f
6 40, 45e97, 93e39, 42e
7 34, 56d93, 93d32, 52d
8 55, 67e76, 57e42, 38e
9 556134
10 78, 99e96, 97e75, 96e
11 69, 93e95, 100e66, 96e
12 75, 95e64, 84e48, 79e
13 71, 95e78, 97e56, 92e
14 63, 91e64, 53e40, 48e
Open in a separate windowaCat. (0.25 mol%), base NaOH (1 eq.), rxn. temp. 80 °C, rxn. time 75 min, determined by GC-MS spectroscopy.bRxn. time 15 min.cRxn. time 30 min.dRxn. time 2 h.eRxn. time 3 h.fRxn. temp. 100 °C.In the TEM image of the recovered Ni@mSiO2 nanocatalyst shown totally collapsed silica structure and sintering nickel nanoparticles (Fig. S5a and b). In the stability of structure, silica was occurred procreated due to the steam during the reaction. In the XRD data of the recovered Ni@mSiO2 nanocatalyst showed sharp peaks (Fig. S5c). It means reflected the increased crystal size. In addition, metallic nickel phase changed nickel silicide and nickel carbide during the reaction.The recycle reaction of the recovered Ni@mSiO2 catalyst showed a low conversion (79%) compared high conversion (100%) of fresh Ni@mSiO2 for 1 h. In the recycle reaction, the catalyst exhibited lower performance due to metallic nickel was changed no active site catalyst such as nickel carbide and nickel silicide.  相似文献   

8.
Catalytic N-methyl amidation of carboxylic acids under cooperative conditions     
Li Yingxian  Chen Wei  Zhao Linchun  Zhang Ji-Quan  Zhao Yonglong  Li Chun  Guo Bing  Tang Lei  Yang Yuan-Yong 《RSC advances》2022,12(32):20550
Amide is a fundamental group that is present in molecular structures of all domains of organic chemistry and the construction of this motif with high atom economy is the focus of the current research. Specifically, N-methyl amides are valuable building blocks in natural products and pharmaceutical science. Due to the volatile nature of methyl amine, the generation of N-methyl amides using simple acids with high atom economy is rare. Herein, we disclose an atom economic protocol to prepare this valuable motif under DABCO/Fe3O4 cooperative catalysis. This protocol is operationally simple and compatible with a range of aliphatic and (hetero)aromatic acids with very good yields (60–99%). Moreover, the Fe3O4 can be easily recovered and high efficiency is maintained for up to ten cycles.

The generation of N-methyl amides using simple acids with high atom economy is rare owning to the volatile nature of methyl amine. Herein, an atom economic protocol was disclosed to prepare this valuable motif under DABCO/Fe3O4 cooperative catalysis.

Amide is a fundamental group that is present in molecular structures of all domains of organic chemistry.1 It is widely distributed in natural products, synthetic drugs and functional polymers, and is also the key chemical connection in proteins.2 It has been shown that amide bond formation alone accounts for 65% of all preliminary screening reactions in the pharmaceutical industry.3 This means the generation of amide bonds with high atom efficiency is of high practical importance. And not surprisingly, ‘amide formation avoiding poor atom economy reagents’ was voted as the top challenge for organic chemistry by the ACS Green Chemistry Institute in 2007.3From synthetic point of view, the ideal way to produce amide bonds would be the direct coupling of readily available carboxylic acids and amines, but this process is thermodynamically unfavourable due to the formation of the corresponding carboxylate-ammonium salt,4 therefore, stoichiometric amount of coupling reagents, such as DCC, DIC, EDCI, HATU, HBTU, HCTU, SOCl2, BOP, acid chloride etc, are generally required to sidestep thermal conditions for amide bond formation.5 These reagents are highly successful, but the process generally suffers from poor atom economy and side products removal issue especially in the large-scale applications.5 To overcome these drawbacks, “nonclassical” amide bonds formation routes were investigated.6 In these processes, the catalyst takes the role of a coupling reagent in generating an active ester suitable for amidation in a waste-free manner. However, these processes have not been applied in the preparation of N-methyl amides, probably because the methyl amine was delivered in its hydrochloride salt, alcoholic or aqueous form due to its volatile nature.On a different note, N-methyl amides are extensively presented in numerous natural products and pharmaceutical molecules, as shown in Fig. 1,7 and the methylation of amides is a promising way to improve the pharmacological property of molecules.8 However, the synthesis of N-methyl amides compounds relies heavily on non-catalytic approaches.5,9 Catalytic approaches were also investigated by Hisaeda,10 Kundu,11 Li,12 Guo,13 Yu,14 Maruoka,15 Wang,16 Chen,17 Lamaty18 and their co-workers starting from nitriles, primiary amides, aldoximes, aldehydes, lignin, carbamoylsilane and alcohols. Until recently, Thakur,19 Marce,20 Sadeghzadeh21 and their co-workers developed elegant N-methyl amidation approach starting from carboxylic acids under nano-MgO, diatomite Earth@IL/ZrCl4 and Mg(NO3)2·6H2O catalysis respectively, while limitations like poor substrate scope or sophisticated tailored catalyst still persist. Mindful of all the above issues, developing an N-methyl amidation process of simple carboxylic acids, which is still of great challenge in synthesis, and establishing a broad (hetero)aryl scope with high atom economy from commercial available reagents and catalysts were critical considerations in this study. Moreover, the significance of N-methyl amides combined with our interests in the development of green synthetic approaches motivated us to explore the direct coupling of the carboxylic acids and isothiocyanates. To the best of our knowledge, this is the first successful work using isothiocyanatomethane to prepare N-methyl amides.Open in a separate windowFig. 1Marketed drugs bearing N-methyl amide group.Our initial investigation begins with phenylacetic acid and isothiocyanatomethane as model substrate for condition optimization. Using acetonitrile as solvent, only trace amount of product was detected under catalyst free or p-toluenesulfonic acid (PTSA) catalysis conditions ( EntryAdditiveTime (h)CatalystYield (%)1—24—52—24PTSA—3—48TEA174—48DBU455—48DMAP436—48DBN517—48DABCO658LiBr48DABCO719Mn(OAc)248DABCO7510MnO48DABCO7911MgO48DABCO8812Al2O348DABCO8513Fe3O448DABCO9814Fe3O424DABCO7515bFe3O448DABCO80Open in a separate windowaReactions were run on 1 mmol 1a and 1.1 mmol 2a with 10 mol% catalyst and 10 mol% additive in 1 mL of MeCN at 85 °C for 48 hours unless otherwise noted.bReaction was conducted at 60 °C.Firstly, different acids were employed to react with isothiocyanatomethane and the results were summarized in Open in a separate windowaReactions were run on 1 mmol 1 and 1.1 mmol 2 with 10 mol% DABCO and 10 mol% Fe3O4 in 1 mL of MeCN for 48 hours at 85 °C unless otherwise noted.Subsequently, aromatic and heteroaromatic acids were tested for their compatibility with current reaction conditions and the results were summarized in Fig. 1) respectively, and all could be convenient prepared using current procedure with excellent yields.Substrate scope for the amidation reactiona
Open in a separate windowaReactions were run on 1 mmol 1 and 1.1 mmol 2 with 10 mol% DABCO and 10 mol% Fe3O4 in 1 mL of MeCN for 48 hours at 85 °C unless otherwise noted.Furthermore, to demonstrate the synthetic utilization of our methodology, the preparation of bioactive compounds was demonstrated Scheme 1. Compounds 3as is a patent HDAC4 inhibitor.22 Under the standard reaction conditions, 3as could be obtained from commercial available 1as in 92% yield. Our procedure is much more atom economy as it excluded the usage of activating reagent and excess amount of base. In a recent report, Yang group reported their pilot-scale synthesis of substituted phenylacetamides to tetrahydroisoquinoline-2-ones.23 In their practice, corrosive thionyl chloride was applied as activating reagent and large excess amount of methyl amine was required, however, moderate yield of 3a was obtained, while our method can achieve better yield along with the exclusion of corrosive thionyl chloride. Following this report, tetrahydroisoquinoline-2-one 4a could be obtained in 82% yield, which could be used in the preparation of various bioactive 4-aryl-tetrahydroisoquinolines 5a with known procedure.24Open in a separate windowScheme 1Application of N-methyl amide.Finally, owing to the magnetic nature of Fe3O4, we try to recover the Fe3O4 from the reaction system and test its efficiency. As the Fe3O4 is always stick to the magnetic stir bar, after the termination of the reaction, the reaction solution was pour out and the tub along with the magnetic stir bar was rinsed with MeCN three times, oven dried and used for the next cycle. The results shown that the Fe3O4 could be used 10 times and still maintained very good efficiency (Fig. 2).Open in a separate windowFig. 2The efficiency of recovered Fe3O4.Combined with the literature reports and experimantal observation,25 a plausable mechanism was proposed in Scheme 2. Firstly, the carboxylic acid reacts with the Fe3O4 to get iron (II and III) carboxylate A, which will coordinate to the intermediate B generated from DABCO and isothiocyanate to get intermediate C. Then, one of the carboxylate attack intermediate B to release DABCO and generates intermediate D. Intermediate D go through an intramolecular addition to generate intermediate E, which go through a rearrangement reaction to get intermediate F with the release of carbonyl sulfide. Finally, the protonation of F with carboxylic acid to get the final product and regenerate the iron (II and III) carboxylate A.Open in a separate windowScheme 2Proposed reaction mechanism.  相似文献   

9.
Synthesis of tetrahydrochromenes and dihydronaphthofurans via a cascade process of [3 + 3] and [3 + 2] annulation reactions: mechanistic insight for 6-endo-trig and 5-exo-trig cyclisation     
Yeruva Pavankumar Reddy  V. Srinivasadesikan  Rengarajan Balamurugan  M. C. Lin  Shaik Anwar 《RSC advances》2023,13(9):5796
Substituted tetrahydrochromenes and dihydronaphthofurans are easily accessible by the treatment of β-tetralone with trans-β-nitro styrene derived Morita–Baylis–Hillman (MBH) acetates through a formal [3 + 3]/[3 + 2] annulation. The reaction proceeds through a cascade Michael/oxa-Michael pathway with moderate to good yields. A DFT study was carried out to account for the formation of the corresponding six and five-membered heterocycles via 6-endo-trig and 5-exo-trig cyclization.

A [3 + 3] and [3 + 2] annulation strategy using nitrostyrene derived MBH primary and secondary nitro allylic acetate for the construction of tetrahydrochromenes and dihydronaphthofurans at room temperature.

The ability to synthesize diverse molecules utilizing nitro allylic MBH acetates in various cascade reactions has received considerable interest.1 A few molecules synthesized using nitro allylic acetates have shown promising cytotoxic, trypanocidal and AchE inhibition2 activity in pharmaceutical and medicinal chemistry. Nitro allylic MBH acetates have been used as main precursors in organocatalysis3 and heterocyclic chemistry,4 and as bicyclic skeletons5 for the construction of elegant building blocks like tetrahydro-pyranoquinolinones,6 sulfonyl furans,7 pyranonaphthoquinones,8 arenopyrans/arenylsulfanes,9 triazoles,10a tetrasubstituted furans,10b fused furans,10c,d tetrasubstituted pyrroles,11a benzofuranones,11b and tetrahydropyrano scaffolds/pyranocoumarins.12 The nitro allylic MBH-acetates can also undergo asymmetric benzylic13a and allylic alkylation13b reactions as well as kinetic resolution [KR]13c,d under normal conditions. These acetates undergo a range of cascade [2 + 3],14 [3 + 2]15 and [3 + 3]16 ring annulation reactions using different substrates. They have been widely utilized in [3 + 2]17a and [3 + 3]17b annulation reactions due to their unique nature of 1,2-/1,3-biselectrophilic reactivity to form either five or six membered rings depending on the nature of nucleophiles employed in the reaction17c,d These adducts are also stable under NHC catalytic conditions to yield cyclopentanes.18Peng-Fei Xu et al. (Scheme 1, eqn (a)) synthesised tetrahydropyranoindoles through organocatalytic asymmetric C–H functionalization of indoles via [3 + 3] annulation through 6-endo trig cyclization.19 The Namboothiri group recently developed a metal free regioselective synthesis of α-carbolines via [3 + 3] annulation involving secondary MBH acetate (Scheme 1, eqn (b)).20 Previously, our group carried out a [3 + 3] cyclization reaction of β-naphthol with primary MBH acetate to study the scope of SN2′ vs. SN2 reaction.21 With our ongoing interest in using nitro styrene derived MBH adducts22 explored the reactivity of primary and secondary MBH acetate with β-tetralone 1 as our model reaction. Initially, the reaction carried out using β-tetralone 1 with primary MBH-acetate 2, predominantly gave a tetrahydrochromene 3via [3 + 3] annulation involving 6-endo trig cyclization through Michael/oxa-Michael cascade process. The possible dihydronaphthofuran product was not observed under the present conditions as primary MBH acetate 2 acts as 1,3-biselectrophile instead of 1,2-biselectrophile (Scheme 1, eqn (c)).Open in a separate windowScheme 1[3 + 3] and [3 + 2] annulation reactions using 1°- and 2°-nitro allylic MBH acetate.On the other hand, the reaction of β-tetralone 1 with secondary MBH acetate 4 gave dihydronaphthofuran instead of the possible tetrahydrochromene product due to the 1,2-biselectrophile nature of secondary MBH acetate (Scheme 1, eqn (d)). The formation of dihydronaphthofuran 5 occurs in an SN2′ fashion via [3 + 2] annulation involving 5-exo-trig cyclization through Michael followed by intramolecular oxa-Michael reaction with the elimination of HNO2. Subsequently, we have carried out a DFT calculation to prove the formation of tetrahydrochromene 3 using primary MBH acetate 2 and dihydronaphthofuran 5 in the case of secondary MBH acetate 4.Initially, we carried out the optimization conditions for constructing tetrahydrochromenes 3a using β-tetralone 1 with MBH nitro allylic primary acetate 2a with different bases and solvents. Reaction with organic base, i.e. DABCO using a polar aprotic solvent such as acetonitrile at room temperature gave the desired product in 27% (23 (CCDC-2149875) ( EntryBaseSolventTime (h)Yield (%)drb1DABCOCH3CN52799 : 12DABCOCH2Cl252299 : 13DABCOCHCl351999 : 14DMAPTHF54099 : 15TEATHF54599 : 16PPh3THF54499 : 17K2CO3THF46099 : 1 8 Cs 2 CO 3 THF 4 77 99 : 19cCs2CO3THF85099 : 110dCs2CO3THF56199 : 111eCs2CO3THF46599 : 112fCs2CO3THF460n.dOpen in a separate windowaUnless otherwise noted, reactions were carried out by and (0.11 mmol) of 1 with (0.11 mmol) of 2a using 0.22 mmol of a base in 1 ml of THF solvent.bDetermined by 1H-NMR analysis of crude reaction mixture.cReaction was carried out using 0.5 equiv. of Cs2CO3.dReaction was carried out using 1.0 equiv. of Cs2CO3.eReaction was carried out using 1.5 equiv. of Cs2CO3.fReaction was carried out using 3.0 equiv. of Cs2CO3.Substrate scope for tetrahydrochromenes 3a–f
Open in a separate windowBased on the best optimized conditions, we studied the scope of different nitro allylic MBH primary acetates (2a–e) with β-tetralone 1. The reaction accommodates various electron rich substituents on the primary MBH acetates (2a–e). The electron rich substituent contaning 2b gave 76% yield for the benzyloxy product 3b. The substrate having meta-OMe and para–OMe gave the desired product 3c and 3d with 71 & 68% of yield, respectively. Furthermore, using fluoro substituent at para position of the MBH adduct gave the product 3e with 72% yield. Reaction carried out using 6-bromo tetralone 1b gave the corresponding product 3f in 68% of yield. Notably, remarkable diastereoselectivity of 99 : 1 dr was observed in all the cases of base screening and substrate scope of MBH primary acetate.Encouraged, by the high diastereoselectivity for various tetrahydrochromenes derivatives 3a–e, we pursued our studies towards asymmetric synthesis of 3a using different chiral catalysts (I–IV). We observed the poor ee for the product formation in the presence of cinchona based squaramide catalyst I & BINAM based urea catalyst II ( EntryCatalystTime (h)Yield (%)ee (%)1I459<22II460<23III46010 4 IV 4 63 49 Open in a separate windowaAll the reactions were carried out with (0.11 mmol) of 1, (0.11 mmol) of 2a, (0.22 mmol) of base and 10 mol% in 1 ml of THF solvent.We next focused our studies on understand the reactivity of secondary MBH acetate 4a using β-tetralone 1. Interestingly, the reaction followed an SN2′ Michael/intramolecular oxa-Michael pathway to form dihydronaphthofuran 5avia [3 + 2] annulation instead of an alternate path resulting in the formation of chromene product via 3 + 3 annulation (i.e., Scheme 1; eqn (d)). To recognize the optimal reaction condition, we carried out the reaction in the presence of Cs2CO3 in THF solvent to get the desired dihydronaphthofuran 5a in 35% yield ( EntryBaseSolventTime (h)Yield (%)1Cs2CO3THF6352Cs2CO3CH2Cl24.5313Cs2CO3CHCl34.5274Cs2CO3CCl44.5165TEATHF5466DABCOCH3CN6417PPh3CH3CN527 8 K 2 CO 3 CH 3 CN 4 72 9bK2CO3CH3CN75910cK2CO3CH3CN66311dK2CO3CH3CN56812eK2CO3CH3CN465Open in a separate windowaUnless otherwise noted, reactions were carried out with (0.11 mmol) of 1 with (0.11 mmol) of 4a using 0.22 mmol% of base in 1 ml of acetonitrile solvent.bReaction was carried out using 0.5 equiv. of K2CO3.cReaction was carried out using 1.0 equiv. of K2CO3.dReaction was carried out using 1.5 equiv. of K2CO3.eReaction was carried out using 3.0 equiv. of K2CO3.The reaction with reduced base equivalents, led to reduced yields confirming that 2.0 equiv. of K2CO3 is desirable to yield dihydronaphthofuran 5a at room temperature within 4 h ( Open in a separate windowTo further demonstrate our protocol''s practical and scalable utility, we have carried out the gram scale preparation of tetrahydrochromene 3a and dihydronaphthofuran 5a in 66 and 70% of yield. We observed the retention of diastereoselectivity i.e., 99 : 1 of tetrahydrochromene 3a, even at the gram scale condition (Scheme 2).Open in a separate windowScheme 2Gram scale synthesis of tetrahydrochromene 3a and dihydronaphthofuran 5a.We have successfully applied the synthetic utility for dihydronaphthofuran 5a. Reduction of the ester group in 5a was feasible using LAH in THF to afford the desired alcohol product 6 with 60% of yield. Using KOH, the ester group in dihydronaphthofuran 5a was hydrolysed to the corresponding acid derivative 7 in 70% yield. The amidation reaction of 7 with aniline accomplished oxidation of the tetralone ring providing the N-phenyl-2-(1-phenylnaphtho[2,1-b]furan-2-yl)acetamide product 8 in 67% of yield (Scheme 3).Open in a separate windowScheme 3Synthetic utility for the dihydronaphthofuran 5a.  相似文献   

10.
DIPEA-induced activation of OH− for the synthesis of amides via photocatalysis     
Mei Wu  Sheng Huang  Huiqing Hou  Jie Lin  Mei Lin  Sunying Zhou  Zhiqiang Zheng  Weiming Sun  Fang Ke 《RSC advances》2022,12(23):14724
The development of green protocols for photocatalysis where water acts as a nucleophile, induced by a weak organic base, is difficult to achieve in organic chemistry. Herein, an efficient light-mediated strategy for the synthesis of amides in which a weak organic base acts as a reductant to induce the formation of OH– from water under metal-free conditions is reported. A mechanistic study reveals that the generation of an N,N-diisopropylethylamine (DIPEA) radical via single electron transfer (SET), with the assistance of photocatalyst, that increases the nucleophilicity of the water molecules with respect to the cyanides is essential. Moreover, the removal rate of nitrile in wastewater can be as high as 83%, indicating that this strategy has excellent potential for nitrile degradation.

Under weak organic base condition DIPEA as a reductant to increase the nucleophilicity of H2O an excellent potential system for nitrile degradation.

The synthesis of amides is a subject of continuous interest and great importance because of their important applications in detergents, agrochemicals, polymers and pharmaceuticals.1 Traditional methods for their synthesis require the transformation of an acid into the corresponding acyl chloride, facilitated by the use of the Schotten–Baumann reaction.2 Although these methods produce amides in good yields, stoichiometric amounts of an activating reagent are required, making these poorly atom economic processes. Various strategies for carboxamide synthesis, such as oxidative alcohol–amine and aldehyde–amine coupling reactions, amine dehydrogenation or oxidation reactions, and C–N coupling reactions, have been developed in recent years.3 Despite this, one of the most straightforward and atom-economical ways to synthesize amides remains the hydration of organonitriles.4 Conventional strategies mostly use strong inorganic bases to generate strongly nucleophilic hydroxide ions, require tough conditions, and are sensitive, especially when using bioactive molecules,5 to the substrate. Moreover, water molecules are usually used as nucleophiles in the hydration of nitriles. Thus, compared to the use of strong basic conditions, the direct nucleophilic addition of water to the cyano group is kinetically slow due to the high energy of the carbon–nitrogen triple bond.6 To circumvent these problems, transition metal (TM) catalytic procedures, where the metal center of the catalyst acts as a Lewis acid to activate the nitrile and the ligand acts as a Lewis base-activated nucleophile, have been developed in recent years.7 But these protocols are associated with certain debilitating disadvantages that include the presence of toxic transition metal cations within the molecular structure of the reagents and difficulties in preventing over-hydrolysis to the corresponding carboxylic acids.Recently, there have been some reports that reductants have been used to change the morphology of water to increase its nucleophilicity.8 Organoamine bases, such as N,N-diisopropylethylamine (DIPEA), have acted in the role of both base and nitrogen radical intermediate and are considered to be reductants.9 However, DIPEA does not lose electrons easily and therefore has a low reductive activity, which means the nitrogen center has to cross a higher energetic barrier. Recently, it was shown that DIPEA could reductively quench many excited photocatalysts by single electron transfer (SET) to generate nitrogen-centered radicals.10 For example, Xu and coworkers11 proposed a new approach using DIPEA to construct difluoroalkylated diarylmethane compounds via visible light photocatalytic radical–radical cross-coupling reactions, in which DIPEA can carry out electron transfer due to the induction of the photocatalyst. It indicates that photooxidation–reduction and organic amine reduction are, when exposed to sufficient light intensity, co-catalytic processes and can generate nitrogen-centered radicals so that the downstream reaction process can continue.12 Herein, an efficient light-mediated strategy for the synthesis of amides in which a weak organic base acts as a reductant to induce the formation of OH– from water under metal-free conditions is reported.Initially, the reaction of the benzonitrile (1a) was selected for the screening of the reaction conditions (Fig. 1, ,22 and and33.Optimization of the reaction conditionsa
EntryVariations from the standard conditionsYieldb (%)
1None89
2Eosin B instead of eosin Y41
3Rose bengal instead of eosin Y73
4Rhodamine B instead of eosin YTrace
5Erythrosin B instead of eosin Y78
6TEA instead of DIPEA64
7DABCO instead of DIPEA13
8DMSO/H2O (1/2) instead of H2O 3 mL54
9DMF/H2O (1/2) instead of H2O 3 mL27
1012 h instead of 24 h47
1128 h instead of 24 h90
12Blue light 5 W instead of blue light 12 W63
13White light instead of blue light43
14DIPEA 1.0 equiv. instead of 2.0 equiv.59
15cGram-scale experiment72
Open in a separate windowaStandard conditions: 1a (0.5 mmol), DIPEA (2.0 equiv.), eosin Y (0.1 equiv.), blue light, 12 W, H2O 3 mL, 24 h, rt.bIsolated yield.c10 mmol 1a, 15 equiv. DIPEA and 0.1 equiv. eosin Y, blue light 12 W, H2O 25 mL 36 h.Open in a separate windowFig. 1General methods for the hydration of organonitriles.Open in a separate windowFig. 2Pharmaceuticals and biomolecules containing a primary amide functional group.Open in a separate windowFig. 3Variation of removal rate of nitrile at different reaction times.With the optimized conditions in hand, the substrate scope was investigated ( Open in a separate windowaStandard conditions: 1a (0.5 mmol), DIPEA (2.0 equiv.), eosin Y (0.1 equiv.), blue light 12 W, H2O 3 mL, 24 h, rt.bIsolated yield.c10 mmol 1ac, 1.5 equiv. DIPEA and 0.1 equiv. eosin Y, blue light 12 W, H2O 25 mL, 36 h.Nitrile wastewater is a big threat to the environment, especially to aquatic organisms. With the optimal reaction conditions established, the effect of the reaction time on the removal rate of nitrile in wastewater was investigated (the nitrile concentration of wastewater was 200 mg L−1 as determined using HPLC). It was shown that the removal rate of nitrile increased up to 83% as the reaction time increased to 24 h and remained stable.To explore the reaction mechanism, a series of control experiments were performed (Scheme 1). These reactions were essentially carried out under conditions in which only one reaction parameter was changed. The control experiments revealed that no reaction occurred in the absence of either the visible light or the photocatalyst, indicating that these two components are essential to the reaction (Scheme 1a, 1b). In addition, no products are formed in the absence of DIPEA, which indicates that the organic base is key to this reaction system (Scheme 1c). Upon conducting the nitrile hydration under an H218O atmosphere (Scheme 1d), we obtained an 18O-labeled product, demonstrating that H2O rather than molecule oxygen serves as the oxygen source.Open in a separate windowScheme 1Control reactions.On the basis of the mechanistic studies above and the literature, a plausible mechanism is outlined in Scheme 2. Initially, the photocatalyst eosin Y is irradiated to give an activated species eosin Y* from which oxygen abstracts an electron to form an O2˙− radical. Then, the oxidation state of the photocatalyst is reduced by the reductive quencher A.10,11,13a Subsequently, the O2˙− radical acquires an electron and H+ from radical B to form HO2. Next, a water molecule and HO2 instantaneously form OH and H2O2 (as determined using HPLC).13b,c Then, a nucleophilic addition of OH to the electrophilic carbon atom of the nitrile generates intermediate D, which is further hydrated to form the product 2a.13dOpen in a separate windowScheme 2Proposed mechanism for this transformation.To verify the above proposed mechanism, density functional theory (DFT) calculations were performed and are shown in Fig. 4. First, the generated eosin Y free radical can easily attack DIPEA to generate radical B along with the release of 27.70 kcal mol−1 of energy. Afterwards, the resulting radical B will spontaneously react with the radical O2˙− and an H2O molecule to generate OH. Then, the obtained OH further reacts with nitrile 1a to form intermediate D, a process with a very small energy barrier of 11.27 kcal mol−1. Finally, intermediate D is rapidly oxidized to the target product 2a.Open in a separate windowFig. 4DFT study of the hydration of nitrile.  相似文献   

11.
[3 + 2] cycloaddition of nonstabilized azomethine ylides and 2-benzothiazolamines to access imidazolidine derivatives     
Kai-Kai Wang  Yan-Li Li  Ming-Yue Wang  Jun Jing  Zhan-Yong Wang  Rongxiang Chen 《RSC advances》2022,12(44):28295
A simple and practical method for the construction of 1,3,5-trisubstituted imidazolidine derivatives via [3 + 2] cycloaddition reaction has been developed. This reaction could smoothly proceed between nonstabilized azomethine ylides generated in situ and 2-benzothiazolamines to deliver a wide scope of differently substituted imidazolidines in high yields (up to 98%). The structure of one example was confirmed by X-ray single-crystal structure analysis.

An effective method for the synthesis of functionalized imidazolidine derivatives via a [3 + 2] cycloaddition reaction from nonstabilized azomethine ylides generated in situ from 2-benzothiazolamines in high yields (up to 98%) has been developed.

Heterocyclic compounds are important structural motifs that are frequently discovered in natural products and biologically active molecules, with wide applications and potency in the field of medicinal chemistry.1 Among them, the 2-aminobenzothiazole scaffolds, which contain a core of isothiourea motifs as a privileged scaffold, are ubiquitous in natural products, drugs and bioactive compounds.2 Some examples of biologically active molecules and pharmaceuticals containing benzothiazoles are shown in Fig. 1. These compounds display a variety of important biological activities, such as anti-HIV, anticancer, antineoplastic and anticonvulsant activity, and so on.3Open in a separate windowFig. 1Selected examples of biologically active compounds containing a 2-aminobenzothiazole scaffolds.Due to their momentous and potent biological activities and structural diversifications, more organic and medicinal chemists have been significantly attracted to developing effective and advanced methodologies for the construction of benzazole scaffolds.4 The 2-benzothiazolamines could serve as alternative C4 synthons with various reaction partners in [4 + 2] cycloadditions for the straightforward and convenient access to 2-aminobenzothiazole scaffolds, and this type of reaction has been well established (Scheme 1a).5 By contrast, the 2-benzothiazolamines which acted as C2 synthons in cycloaddition reactions have been very limited.6 Therefore, the development of efficient cycloaddition between the 2-benzothiazolamines as C2 synthons and suitable reaction partners to provide diverse functionalized heteroarenes molecules is always in great demand.Open in a separate windowScheme 1Previous reports and our protocol.On the other hand, the 1,3-dipole cycloaddition is recognized one of the most efficient and powerful methods for building five membered heterocyclic rings from simple starting materials.7 In particular, the nonstabilized azomethine ylides generated in situ from N-benzyl-substituted compounds were a highly reactive intermediate with unsaturated compounds, such as activated alkenes,8 aromatic ketones,9 aromatic aldehydes,10 phthalic anhydrides,11 imines,12 cyano compounds13 or stable dipoles,14 to build various N-containing heterocycles via [3 + 2] or [3 + 3] cycloaddition reactions (Scheme 1b). Despite the progress were well developed via 1,3-dipolar cycloadditions employing nonstabilized azomethine ylides as the substrates, challenges still remain. Herein, we would like to present an effective process to furnish functionalized 1,3,5-trisubstituted imidazolidine derivatives via [3 + 2] cycloaddition strategy from 2-benzothiazolamines with nonstabilized azomethine ylides generated in situ (Scheme 1c). In contrast, the desired product was not obtained via [4 + 3] cycloaddition reaction.Initially, we used 2-benzothiazolimine 1a and N-(methoxymethyl)-N-(trimethylsilyl-methyl)-benzyl amine 2a which could in situ form azomethine ylide in the presence of acid as model substrates to optimize the reaction conditions. The results were shown in EntryAdditiveSolventTimeYield of 3ab (%)1AcOHCHCl312482TfOHCHCl312513TFACHCl33904HClCHCl312205TFACH2Cl21986TFADCE3827TFAEtOAc6758TFAToluene12709TFAMeOH123110TFACH3CN37111TFADioxane125912TFAEt2O124613TFATHF1262Open in a separate windowaReaction conditions: 2-benzothiazolimine 1a (0.1 mmol, 1.0 equiv.), N-(methoxymethyl)-N-(trimethylsilyl-methyl)-benzyl-amine 2a (0.12 mmol, 1.2 equiv.), addition (0.01 mmol, 0.1 equiv.) and solvent (1.0 mL) in a test tube at room temperature.bYield of the isolated product.After establishing the optimal reaction conditions, the scope and limitations of this [3 + 2] cycloaddition reaction for formation of 1,3,5-trisubstituted imidazolidines were examined. The results were shown in )15 was unequivocally confirmed by X-ray crystallographic analysis (see ESI). On the other hand, when the R2 groups were heteroaromatic substituted group, such as 2-furanyl, 3-thienyl, 1-naphthyl and 2-naphthyl, the cycloaddition reaction were well tolerated and could also proceed smoothly without obvious interference to produce corresponding pyrazoles 3k–3n in 92–96% yields. Meanwhile, substituted substrates at the C4 or C6-position of the 2-aminobenzothiazole, such as Me and Cl were also reacting smoothly to afford the corresponding products in 95%, 96% yields (3o–3p), respectively. Notably, the 2-benzoxazolimine as substrate also reacted smoothly with azomethine ylides 2a to deliver the corresponding product 3q in high yield (94%). Especially, when using (R)-nonstabilized azomethine ylide as the chiral substrate in the cycloaddition reaction, the reaction could proceed smoothly under the standard conditions to afford the corresponding chiral product 3r in 95% yield as single diastereomer only. The chiral center was determined by NOESY analysis (see ESI). Nevertheless, as for the R2 group, when it was changed from an aryl group to an alkyl group, the reaction did not take place (3s). The possible reason may be due to the low activity of alkyl substituted substrate.Substrate scopes for the formation of 1,3,5-trisubstituted imidazolidine derivatives 3a
Open in a separate windowaReaction conditions: 2-benzothiazolimine 1 (0.1 mmol, 1.0 equiv.), N-alkyl amine 2 (0.12 mmol, 1.2 equiv.), TFA (0.01 mmol, 0.1 equiv.) and CH2Cl2 (1.0 mL) in a test tube at room temperature for 1 h. Yield of the isolated product.To further illustration and highlight the practical utility of the method for 1,3,5-trisubstituted imidazolidines, the gram scale experiments were performed. The reaction of 3 mmol of 1a with 3.6 mmol 2a proceeded smoothly under optimized condition for 1 h, producing the 1,3,5-trisubstituted imidazolidine 3a (1.057 g) in 95% yield without an obvious loss of efficiency (Scheme 2).Open in a separate windowScheme 2Gram-scale synthesis of product 3a.On the other hand, we developed conditions to convert the bromine atom into phenyl group via Suzuki coupling of product 3h with phenylboronic acid, which obtained product 4 in 65% yield (Scheme 3).Open in a separate windowScheme 3Transformations of product 3h.Based on the results presented in Scheme 4. First, the nonstabilized azomethine ylide from N-(methoxymethyl)-N-(trimethylsilyl-methyl)-benzyl amine 2a is generated in the presence of TFA. Then, this nonstabilized azomethine ylide could react with 2-benzothiazolimine 1a to obtain the desired product 3avia [3 + 2] cycloaddition reaction.Open in a separate windowScheme 4Proposed mechanism.In conclusion, we have developed a mild [3 + 2] cycloaddition reaction from nonstabilized azomethine ylides generated in situ with 2-benzothiazolamines. In this cycloaddition, 2-benzothiazolamines were used as C2 synthon. The method could offer a broad range of functionalized 1,3,5-trisubstituted imidazolidines in high yields (up to 98%) with high regioselectivity at room temperature. Additionally, the merits of our method are cheap starting materials, wide substrate scope, without metal catalyst. The synthetic utility and practicality were also highlighted by the gram-scale experiment and the synthetic transformation. The potential application of these 1,3,5-trisubstituted imidazolidines is under investigation in this laboratory.  相似文献   

12.
A formal intermolecular [4 + 1] cycloaddition reaction of 3-chlorooxindole and o-quinone methides: a facile synthesis of spirocyclic oxindole scaffolds     
Chao Lin  Qi Xing  Honglei Xie 《RSC advances》2021,11(30):18576
Herein, we developed an efficient and straightforward method for the rapid synthesis of spirocyclic oxindole scaffolds via the [4 + 1] cyclization reaction of 3-chlorooxindole with o-quinone methides (o-QMs), which were generated under mild conditions. The products could be obtained in excellent yields with numerous types of 3-chlorooxindole. This methodology features mild reaction conditions, high atom-economy and broad substrate scope.

Herein, we developed an efficient and straightforward method for the rapid synthesis of spirocyclic oxindole scaffolds via the [4 + 1] cyclization reaction of 3-chlorooxindole with o-quinone methides (o-QMs), which were generated under mild conditions.

The structural diversity of spirocyclic oxindole scaffolds is a reason for their frequent occurrence in many relevant natural products and medicinal agents (Fig. 1).1 In particular, natural spirocyclic-2-oxindole scaffolds have been proven to exhibit a broad range of biological activities and have attracted increasing attention in the synthetic field. For instance, XEN 907 is a novel pentacyclic spirooxindole with excellent activities as sodium channel blockers.2 Due to their unique structure and intriguing biological activity, numerous methodologies have been developed for the construction of these privileged frameworks.3 For example, in the past few years, transition-metal catalyzed or organocatalytic [3 + 2] cycloaddition reactions have been developed for the synthesis of spirocyclic oxindole scaffolds.4 Despite the emergence of these elegant approaches, exploiting new strategies for the construction of spirocyclic oxindole derivatives is still highly desirable.Open in a separate windowFig. 1Examples of biologically active spirocyclic oxindole scaffolds. Ortho-quinone methides (o-QMs) as highly reactive versatile intermediates have been of great interest to the chemical and biological community.5o-QMs react with various classes of reagents by three typical reaction pathways: 1,4-addition of nucleophiles, [4 + 2] cycloaddition with dienophiles and oxa-6π-electrocyclization.6 Because most o-QMs are unstable, these reactions generally depend on the reaction conditions used for the generation of o-QMs in situ. Rokita et al. reported that o-silylated phenols when exposed to fluoride could also produce o-QMs under mild reaction conditions.7Because of the dual nature (nucleophilic/electrophilic) of the C-3 position, 3-chlorooxindole serves as a highly reactive starting material in the synthesis of spirocyclic oxindole scaffolds. The introduction of a chloro group at the C-3 position of indoles serves as an excellent leaving group in favour of the subsequent cyclization. In addition, this also increases the acidity of the C–H bond for directly entering the C-3 quaternary centers.8 Inspired by this reactivity profile, 3-chlorooxindoles have been successfully utilized for [2 + 1]9 and [4 + 1]10 cyclization to synthesize spirocyclic oxindole scaffolds (Fig. 2).Open in a separate windowFig. 2Representation of the synthesis and applications of 3-chlorooxindoles.We designed an efficient and straightforward method for the rapid synthesis of spirocyclic oxindoles via the [4 + 1] cyclization reaction of 3-chlorooxindole with o-QMs, which were generated under mild conditions. In this study, using TBAF as the fluoride source and base ensures that the one-pot domino reaction will occur in mild reaction conditions, with high atom-economy and broad substrate scope.Initially, we carried out optimization studies by examining the reaction between O-silylated phenol 2a and 3-chlorooxindole 1a. Indeed, when TBAF was employed as the fluoride source, a smooth [4 + 1] cyclization reaction occurred, affording the spirocyclic oxindole product 3a with 75% yield (entry 1, EntryF source X Y SolventTemp (°C)Yieldb1TBAFc1.24.0DCMrt75%2TBAF1.54.0DCMrt87%3TBAF2.04.0DCMrt85%4CsF1.54.0DCMrt11%5TBAF1.53.0DCMrt80%6TBAF1.54.0CHCl3rt83%7TBAF1.54.0THFrt94%8TBAF1.54.0Toluenert90%9TBAF1.54.0DMFrt72%10TBAF1.54.0MeCNrt84%11TBAF1.54.0MeOHrtNDOpen in a separate windowaReaction conditions: 1a (0.3 mmol), solvent (3.0 mL), 6 h.bIsolated yield.cTBAF (1 M in THF solution).With the optimal conditions known, we next investigated the substrate scope of substituted 3-chlorooxindole 1 using O-silylated phenol 2a as a representative ( Open in a separate windowaReaction conditions: 1 (0.3 mmol), 2a (1.5 eq.), TBAF (4.0 eq.), THF (3.0 mL), 6 h.bIsolated yields.Next, we also explored the substrate scope of substituted 3-chlorooxindole 1 using O-silylated phenols 2a′ ( Open in a separate windowaReaction conditions: 1 (0.3 mmol), 2a′ (1.5 eq.), TBAF (4.0 eq.), THF (3.0 mL), 6 h.bIsolated yields.On the basis of above-mentioned results, a plausible mechanism for this formal [4 + 1] cycloaddition reaction is depicted in Scheme 1. Initially, the highly reactive o-QMs are generated via the desilylation/elimination reaction. Then, 3-chlorooxindole 1a as a nucleophile attacks the external carbon of o-QMs, affording zwitterion Int-1. Finally, the zwitterion Int-1 loses one molecular HCl through a nucleophilic attack, yielding the spirocyclic oxindole product 3a.Open in a separate windowScheme 1Possible mechanism.  相似文献   

13.
Catalyst- and solvent-free approach to 2-arylated quinolines via [5 + 1] annulation of 2-methylquinolines with diynones     
Hai-Yuan Zhao  Fu-Song Wu  Li Yang  Ying Liang  Xiao-Lin Cao  Heng-Shan Wang  Ying-Ming Pan 《RSC advances》2018,8(9):4584
A novel route for the synthesis of 2-arylated quinolines through a [5 + 1] annulation directly from 2-methylquinolines and diynones under catalyst-free and solvent-free conditions was disclosed. This synthetic process was atom-economic, with good tolerance of a broad range of functional groups, and with great practical worth.

A novel route for the synthesis of 2-arylated quinolines through a [5 + 1] annulation directly from 2-methylquinolines and diynones under catalyst-free and solvent-free conditions was disclosed.

Nitrogen-containing heterocyclic compounds are ubiquitous in natural molecules and exhibit a wide array of biological activities.1 Among various N-heterocycles, quinoline nuclei are privileged scaffolds that occupy an important role in many medicinally relevant compounds.2 2-Arylated quinolines are found in many medicinal compounds including etoricoxib,3 rosuvastatin,4 and gleevec,5 as well as molecules designed for other purposes including P, N ligands, such as QUINAP.6 Because of their unique biological activity and wide application, the functionalized 2-arylated quinoline elicited considerable synthetic interest, and a variety of synthetic routes have been established.7 In addition, some classical synthetic methods such as Kumada,8 Suzuki,9 Negishi,10 or Stille9b are usually used to efficiently prepare these compounds, but these methods require the preparation of cross-coupling reagents such as Grignard reagents, boronic acids, organozincs, and organostannanes in advance and these cross-coupling reagents are unstable or toxic or can''t be isolated as solids.11 More recently, much research has been directed toward the synthesis of 2-arylated quinolones and their derivatives via transition-metal-catalyzed C–H arylation12 and many other methods also have been developed by transitional-metal-catalyzed cross-couplings.13 These transition metals included Co, Cu or Pd. Moreover, 2-arylated quinolines synthesis via direct C–H arylation of quinolones with various aryl bromides, arylboronic acid, or arylzinc reagents catalyzed by transition metal catalyst, such as Rh, Fe, or Ni, have been investigated by Bergman,14 Maiti,15 and Tobisu.16 Ru17 or Mn18 also was used to catalyze indirect Friedländer synthesis to obtain 2-arylated quinolines that involved oxidative cyclization of 2-aminobenzyl alcohol with either ketones or alcohols. Although these processes were highly efficient and significance, none of these procedures could directly provide the final products with a low content of the heavy metal impurities which are strict restrictions in drugs and pharmaceuticals.19 Thus, an alternative, general, solvent-free, and environmentally sustainable procedure is urgently required for the quick synthesis of 2-arylated quinolines.Recently, as our continuous study on the C(sp3)–H activation of 2-methylquinolines, which provided a facile synthetic approach to access substituted pyrrolo[1,2-a]quinolones (Scheme 1),20 we had found that the methyl of 2-methylquinolines has very high reactivity. Based on this, herein we reported the catalyst-free and solvent-free [5 + 1] annulation of 2-methylquinolines and diynones to access 4-(quinolin-2-yl)-phenols. To the best of our knowledge, there were none of the group that reported direct construction of six-member aromatic-ring at the methyl of 2-methylquinolines with diynones to give 4-(quinolin-2-yl)-phenols. The present novel construction protocol for 4-(quinolin-2-yl)-phenols had several significant advantages: (1) this chemistry provided a novel and simple strategy for the synthesis of highly valuable 4-(quinolin-2-yl)-phenols under very simple conditions; (2) according to the atom economy concept, this protocol was carried out under catalyst-free and solvent-free conditions, without the addition of any acid, base, or other reagents, which provided the final products without heavy metal impurities and improved its potential utility; (3) the method featured high functional group tolerance, high yields, and broad substrate scopes. Particularly, this route could directly introduce two different substituent groups on the newly formed of six-member aromatic-ring (see Open in a separate windowScheme 1C–H bond activation of 2-methylquinolines and 2-methylpyridine.Synthesis of 4-(quinolin-2-yl)phenol derivativesa,b,c
Open in a separate windowaYield of the isolated product, calculated from 2.bThe reaction completed under solvent-free (see ESI).c3d, 3g, 3h, 3j completed in PhCl (see ESI).To examine the feasibility of our proposed protocol, 2-methylquinoline (1a) and 1,5-diphenylpenta-1,4-diyn-3-one (2a) were chosen as the model substrates and diverse reaction conditions were screened as shown in Fig. 1).21 To improve the efficiency, we used Sm(OTf)3 as catalyst, but the result provided 3a in less than 60% ( EntryCat.SolventTemp. (°C)Yieldb (%)1—PhCl120652Sm(OTf)3PhCl120553cSm(OTf)3PhCl12004Cs2CO3PhCl12005Et3NPhCl12006AcOHPhCl12007—Toluene120358—DMF120639—DMSO1204510—1,4-Dioxane120Trace11d——1207312d——11075 13 d —— 100 85 14d——906015d,e——10086Open in a separate windowaReactions conditions: 1a (0.6 mmol), 2a (0.5 mmol), catalyst (10 mol% to 2a), solvent (1 mL), sealed tube, 10 h.bIsolated yield of pure product based on 2a.cCs2CO3 as base was added to the reaction.d1a (1.5 mmol), 2a (0.5 mmol), sealed tube, 10 h.eReaction time was 15 h.Open in a separate windowFig. 1X-ray crystal structure of 3a (CCDC 1534893).With the optimized conditions in hand, a series of diynones and 2-methylquinolines were subjected to the reaction to investigate the scope and the results were shown in Scheme 2. It was observed that the presence of 2 equiv. of 1,1-diphenylethylene or BHT (2,6-di-tert-butyl-4-methylphenol) didn''t suppress the synthesis of 3a. These results suggested that a radical mechanism wasn''t likely involved.Open in a separate windowScheme 2Control experiments.A possible mechanism is proposed in Scheme 3. At first, the enamine intermediate A was formed from 1avia tautomerization,22 followed by, Michael addition to diynones, giving the intermediate B. And then, the enamine intermediate C was generated from Bvia the requisite disruption of aromaticity. Subsequently, intermediate C was transformed into intermediate D by intramolecular annulation reaction and the intermediate D was rapidly aromatized to form the stable product 3a.Open in a separate windowScheme 3Plausible mechanism.In summary, we have developed a rapid, simple, efficient, catalyst-free, and solvent-free reaction to access 4-(quinolin-2-yl)-phenols through a [5 + 1] annulation directly from 2-methylquinolines and diynones. The synthetic process was atom-economic, applicable to wide range of substrates, and has functional group tolerance, and these features would render this method attractive for academic and industrial use.  相似文献   

14.
1,3-Dipolar cycloaddition of isatin N,N′-cyclic azomethine imines with α,β-unsaturated aldehydes catalyzed by DBU in water     
Zhan-Yong Wang  Ting Yang  Rongxiang Chen  Xueji Ma  Huan Liu  Kai-Kai Wang 《RSC advances》2020,10(41):24288
A simple and green procedure was established by [3 + 3] cycloaddition reaction of isatin derived cyclic imine 1,3-dipoles with α,β-unsaturated aldehydes, giving the desired spiro heterocyclic oxindoles with aza-quaternary centers in good yields and diastereoselectivities. It should be noted that water can be employed as a suitable solvent for the improvement of diastereoselectivity.

A simple and green procedure was established by [3 + 3] cycloaddition reaction of isatin derived cyclic imine 1,3-dipoles with α,β-unsaturated aldehydes, giving spirooxindoles with aza-quaternary center in good yields and diastereoselectivities.

Aza-quaternary centers are pivotal structural units, which exist in a variety of bioactive molecules and natural products.1 In particular, spirooxindoles at the C3 position bearing a quaternarized N-heterocycle have attracted considerable attention because of their privileged structural units with attractive bioactivities,2 for example, antimalarial,3 anti-HIV,4 antitumor,5 anticancer,6 inhibitor at the vanilloid receptor,7 antituberculosis,8etc. (Fig. 1). Due to their remarkable biological importance, great efforts have been made to access spiro heterocyclic oxindoles with aza-quaternary centers. These methods include cycloaddition of imines,9 1,3-dipolar cycloaddition,10 multicomponent cyclization reaction11 and metal-catalyzed cycloaddition.12 Among them, 1,3-dipolar cycloaddition is one of the most powerful tools for the construction of diverse spirooxindole fused N-heterocyclic scaffolds. Of these, N,N′-cyclic azomethine imines were widely studied for constructing various types of N-heterocyclic skeletons with spirooxindole as a stable and easily accessed 1,3-dipoles. In 2013, Wang''s group reported their pioneering studies on Et3N-catalyzed diastereoselective [3 + 3] annulation of N,N′-cyclic azomethine imines with isothiocyanatooxindoles to build 3,3′-triazinylspirooxindoles. In 2017, Wang et al. developed a new isatin-derived N,N′-cyclic azomethine imine 1,3-dipoles, and successfully applied in the [3 + 2] cycloaddition reaction for the construction of spirooxindoles bearing N-heterocycles (Scheme 1a).10c Very recently, Jin''s group reported a Cs2CO3-catalyzed [3 + 4] annulation of isatin-derived 1,3-dipole with aza-oQMs (Scheme 1b).10d Furthermore, Moghaddam and coworkers developed an efficient method for the synthesis of pyridazine-fused spirooxindole scaffolds by 1,3-dipolar [3 + 3] cycloadditions (Scheme 1c).10e On the other hand, α,β-unsaturated aldehydes and their analogs as readily available substrates are also important building blocks in the synthesis of heterocyclic compounds which are widely applied in N-heterocyclic carbenes catalysis and other organocatalysis.13 Inspired by these great works and our continuing efforts towards green synthesis of spirooxindole skeletons. We envisioned a quick and efficient way of [3 + 3] cyclization reaction of α,β-unsaturated aldehydes with the new isatin N,N′-cyclic azomethine imine 1,3-dipoles via oxindole C3 umpolung. We wish to disclose herein that a green and practical access to synthesize pharmacologically interesting spirooxindole derivatives by involving isatin N,N′-cyclic azomethine imine 1,3-dipole as nucleophiles and various α,β-unsaturated aldehydes in water using DBU as organocatalyst. Our initial examinations were carried out using isatin derivated cyclic imine 1,3-dipole 1a (0.1 mmol) and α,β-unsaturated aldehyde 2a (0.12 mmol) as the model substrates, the results of condition optimization are shown in Open in a separate windowFig. 1Selected bioactive products of C3-spirooxindoles with aza-quaternary centers.Open in a separate windowScheme 1Isatin-derived N,N′-cyclic azomethine imine 1,3-dipoles participated in the construction of N-heterocyclic skeletons with C3-spirooxindole.Optimization of the reaction conditionsa
EntryCatalystSolventTime (h)Yieldb (%)3a : 4ac
1DCM24
2DABCODCM24Trace
3DMAPDCM24Trace
4NEt3DCM3251 : 3.4
5DIPEADCM24Trace
6DBUDCM0.1751.7 : 1
7Cs2CO3DCM1341 : 2.4
8KOButDCM0.1251 : 2.4
9PPh3DCM24Trace
10PyrrolidineDCM24591.6 : 1
11dDBUDCM24611.8 : 1
12DBUToluene0.2751.3 : 1
13DBUCH3CN0.1202.3 : 1
14DBUTHF0.1851.2 : 1
15DBUCHCl30.1861.7 : 1
16eDBUCHCl30.1861.5 : 1
17DBUEtOH24621.8 : 1
18Na2CO3EtOH24461 : 1.6
19DBUH2O24618 : 1
Open in a separate windowaOtherwise specified, all reactions were carried out using 1a (0.1 mmol), 2a (0.12 mmol), catalyst (0.1 mmol), solvent (1 ml).bIsolated yields of diastereoisomeric mixture.cDetermined by 1H NMR.dCatalyst (0.01 mmol).ePerformed at reflux.Under the optimal reaction conditions, the generality of this reaction was next investigated. As can be seen from Open in a separate windowaAll reactions were carried out using 1 (0.1 mmol), 2 (0.12 mmol), DBU (1.0 equiv.) in water (1.0 ml) at room temperature.bIsolated yields were diastereoisomeric mixture.cdr was determined by 1H NMR in the crude products.Based on our results and previous studies, a plausible catalytic cycle is proposed in Scheme 2. 1a was promoted by a base to form more stable intermediate I. After this, intermediate I underwent 1,4-Michael addition with α,β-unsaturated aldehyde 2a to form II. Next, keto–enol tautomerism occurred to form intermediate III. To avoid the steric hindrance, the intermediate III attack preferentially to the Re-face of aldehyde, leading to the formation of the major product 3a.Open in a separate windowScheme 2A plausible catalytic cycle.In conclusion, we have disclosed a novel metal-free DBU-catalyzed [3 + 3] cycloaddition reaction via C3 umpolung strategy of oxindole. Varieties of isatin derivated cyclic imine 1,3-dipoles and α,β-unsaturated aldehydes were compatible with this protocol under mild conditions, and afforded spiro heterocyclic oxindoles with aza-quaternary center in good yields with good to high diastereoselectivities. Notably, water as a green solvent had positive effect on the diastereoselectivities.  相似文献   

15.
Iodine-catalyzed sulfonylation of sulfonyl hydrazides with tert-amines: a green and efficient protocol for the synthesis of sulfonamides     
Jinyang Chen  Xiaoran Han  Lan Mei  Jinchuan Liu  Kui Du  Tuanwu Cao  Qiang Li 《RSC advances》2019,9(54):31212
This study provides a direct, sustainable and eco-friendly method for the synthesis of various sulfonamides via the sulfonylation of sulfonyl hydrazides with tert-amines. The method utilizes sulfonyl hydrazides to oxidize and couple with tertiary amines through selective cleavage of C–N bonds. In this reaction, molecular iodine was used as the catalyst and t-butyl hydroperoxide was used as the oxidant.

Sustainable and eco-friendly method for the synthesis of various sulfonamides.

Sulfonamides commonly serve as synthetic intermediates to produce various drugs and industrial compounds (Scheme 1).1 They also commonly act as a N-sulfonyl protecting group for easy removal under mild conditions.2 There have been many efforts devoted to synthesizing these compounds. Among the methods developed, nucleophilic substitution of an amine with a sulfonyl chloride or sulfonamides with organic halides in the presence of a base is frequently utilized.3 Over the past few years, transition metal catalysis has been proved a powerful tool to synthesize sulfonamides. For example, a cross-coupling reaction of primary sulfonamides with aryl halide or boronic acids,4 a Chan-Lam type coupling reaction of sulfonyl azides with boronic acids,5 or an oxidative coupling reaction of sulfinate salts with amines were developed.6 However, the use of non-stable, hazardous and mutagenic starting materials and toxic high boiling polar solvents in these reactions resulted in a larger amount of toxic waste. Furthermore, the use of stoichiometric amounts of bases or transition metals makes reactions with a slow reactivity and poor functional group tolerability. Therefore, a novel, sustainable, efficient, and eco-friendly method is desired to synthesize sulfonamides.Open in a separate windowScheme 1Examples of important sulfonamide drugs in top 200 pharmaceuticals of 2018.Iodine and its salts have been reported as very efficient catalysts in CDC reactions in which transition metals are used as catalysts.7 In recent years, many methods have been reported for synthesis of sulfonamides under metal-free conditions.8 As a source of sulfonyl groups, sulfonyl hydrazides are readily accessible solid. They are stable in air and under moisture conditions, and can be easily prepared and stored. Most importantly, only water and nitrogen were obtained as by-products during the reactions using sulfonyl hydrazides as starting materials. Iodine catalyzed oxidative coupling of sulfonyl hydrazides with secondary amines have been developed (Scheme 2(a)).9Tert-amines can donate an amine group via a C–N cleavage in place of primary or secondary amines. Compared to the high reactivity of primary or secondary amines, tert-amines are less nucleophilic and non-destructive for some amine-sensitive functional groups. Recently, Yuan et al. and Gui''s have developed a new method to sulfonamides using I2-mediated or catalyzed C–N bond cleavage of tert-amines (Scheme 2(b)).10 Meanwhile, Sheykhan et al. reported a novel electrochemical oxidative sulfonylation11 of tert-amines (Scheme 2(c)).12 Moreover, catalytic reactions in the aqueous phase have also been recently developed,13 and sulfonylation of sulfonyl hydrazides has caused wide interest recently.14 In this study, we will report a new method to synthesize sulfonamides using iodine-catalyzed oxidative coupling of sulfonyl hydrazides with tert-amines (Scheme 2(d)). This approach avoids use of metal catalysts and hazardous regents; the materials, sulfonyl hydrazides and tert-amines, are versatile intermediates in commercial.Open in a separate windowScheme 2Methods for the synthesis of sulfonamides.Commercially available sulfonyl hydrazide 1a was selected as a sulfonyl source to synthesize sulfonamides. When 1a was mixed with 1 equiv. of N-ethyl pyrrolidine under various solvents at 80 °C, an 80% yield of the corresponding sulfonamide 3a was obtained after 4 hours in the aqueous phase (Scheme 3).Optimization of the reaction conditionsa
EntryI2 (mol%)Oxidant (equiv.)SolventTemp. (°C)Yieldsb (3a/4a%)
1I2 (20 mol%)TBHP (2.0 equiv.)CH3CN800/50
2I2 (20 mol%)TBHP (2.0 equiv.)THF805/29
3 I 2 (20 mol%) TBHP (2.0 equiv.) H 2 O 80 80/0
4I2 (20 mol%)TBHP (2.0 equiv.)DMF80NR
5I2 (20 mol%)TBHP (2.0 equiv.)Toluene80NR
6I2 (20 mol%)TBHP (2.0 equiv.)EtOH8031/27
7TBHP (2.0 equiv.)H2O800
8I2 (20 mol%)H2O80NR
9I2 (20 mol%)H2O2 (2.0 equiv.)H2O80NR
10I2 (20 mol%)Oxone (2.0 equiv.)H2O80NR
11I2 (20 mol%)O2H2O80NR
12I2 (10 mol%)TBHP (2.0 equiv.)H2O8056/0
13I2 (5 mol%)TBHP (2.0 equiv.)H2O8043/0
14I2 (20 mol%)TBHP (1.0 equiv.)H2O8062/0
15I2 (20 mol%)TBHP (2.0 equiv.)H2O10061/0
16I2 (20 mol%)TBHP (2.0 equiv.)H2O6040/0
17NH4ITBHP (2.0 equiv.)H2O80Trace
18TBAITBHP (2.0 equiv.)H2O80NR
19cI2 (20 mol%)TBHP (2.0 equiv.)H2O8080/0
Open in a separate windowaReaction conditions: 1a (0.3 mmol), 2a (0.3 mmol), I2 (20 mol%), H2O (3 mL), 8 h, 80 °C. TBHP: tert-butyl hydroperoxide, 5.0–6.0 M in decane.bIsolated yield.cThe reaction was performed under nitrogen atmosphere.Open in a separate windowScheme 3The standard process of the sulfonylation reactions.In sulfonylation of amines, it is essential to maintain the process intact by methyl, methoxy, tertiary butyl, phenyl, halo, cyan and nitro substituents presenting at the aromatic ring. In all cases, the yields of the corresponding sulfonamides 3b–3k were in the range of 65–85%. In this series, the lowest yields were obtained for electron-deficient substituted products, which could be attributed to electron factors affecting the catalysis. Notably, arylsulfonyl hydrazides bearing substituents at the metal position could be converted to the sulfonamides in yields of 78% (3l). The yield of 2-naphthalene sulfonamides was 83% (3m). The reaction of 3-chloro-4-(trifluoromethyl)benzene-1-sulfonyl hydrazide with amine produced 3p in a good yield, which could be leveraged for consequent coupling reaction. Because most sulfonamides with relevance for crop protection or medical application contain heterocycles, we included such substrates in our work. Fortunately, heteroaromatic sulfonyl hydrazides could be tolerated in this reaction, achieving the desired product in better yields (3o and 3p). Thus, this method has been demonstrated as a practical and efficient way to synthesize sulfonamides ( Open in a separate windowaConditions: 1 (0.3 mmol), 2a (0.3 mmol), I2 (20 mol%), H2O (3 mL), 8 h, 80 °C. TBHP (0.6 mmol).bIsolated yields.In subsequent studies, we examined the sulfonylation of sulfonyl hydrazides with various tertiary amines under the optimal conditions, and results were summarized in Open in a separate windowaConditions: 1c (0.3 mmol), 2 (0.3 mmol), I2 (20 mol%), H2O (3 mL), 8 h, 80 °C. TBHP (0.6 mmol).bIsolated yields.The sulfonylation can also be carried out on a larger scale reaction, and the desired product (3a) was obtained in the yield of 80%, when 6 mmol of benzenesulfonohydrazide (1a) was treated with 6 mmol of 1-ethylpyrrolidine (2a) under the standard conditions (Scheme 4(a)). To shed light on the mechanism of the reaction, benzenesulfonohydrazide (1a) was treated with 1-ethylpyrrolidine (2a) under standard conditions by using 2.0 equiv. of TEMPO or BHT as radical scavengers (Scheme 4(b)), and desired product 3a was obtained in trace yields, suggesting that a single-electron transfer process was involved through the whole reaction.Open in a separate windowScheme 4(a) Larger-scale synthesis of 3a. (b) Control experiments for mechanism study.On the basis of the above experimental results and previous works,9a,10b a possible mechanism has been depicted in Scheme 5. Firstly, the transformation presumably involves an initial reaction of I2 with TBHP to create a reactive tert-butoxyl or tert-butyl peroxy radical. Then, tert-butoxyl or tert-butyl peroxy radical abstract a hydrogen from N-ethyl pyrrolidine to form radical 4a. After an electron transformation, it is converted to an intermediate aninium ion 4b. This intermediate 4b was then hydrolyzed by the elimination of an aldehyde to result in a secondary amine 4c. Finally, 4c reacts with a sulfonyl radical to generate the desired sulfonamide product 3a.Open in a separate windowScheme 5Proposed mechanism of the sulfonylation.  相似文献   

16.
Oxidative bridgehead functionalization of (4 + 3) cycloadducts obtained from oxidopyridinium ions     
Wanna Sungnoi  Michael Harmata 《RSC advances》2022,12(44):28572
Treatment of selected (4 + 3) cycloadducts derived from oxidopyridinium ions with N-iodosuccinimide (NIS) in hexafluoroisopropanol (HFIP) resulted in the formation of bridgehead ethers via a net oxidative C–H activation.

Selected azabicyclo[4.3.1]deca-3,8-dienes react with NIS in HFIP to form bridgehead ethers via a net oxidative C–H activation.

Since their beginnings in the mid-1970s, the (4 + 3) cycloaddition reactions of N-substituted oxidopyridinium ions have provided an attractive and facile method for the construction of nitrogenous, heterocyclic seven-membered rings (Scheme 1).1,2 While N-aryl and N-alkenyl substitution of the pyridinium nitrogen have largely dominated the literature, to the best of our knowledge, there had been only one example of a (4 + 3) cycloaddition reaction of N-alkyl oxidopyridinium ions reported prior to 2017.2a Our recent reports of the reaction of N-methyl oxidopyridinium ions with conjugated dienes expanded the scope of the (4 + 3) process, due to the incorporation of an ester functional group at the 5-position of the hydroxypyridine starting material.3 Cycloadducts are formed in good to excellent yields, and the reactions of select dienes proceed in high regioselectivity. Furthermore, the cycloaddition process can be steered towards a high preference for the endo diastereomer when the starting diene bears a bulky trialkylsilyl groups at C2, as we have recently reported (Scheme 2).4Open in a separate windowScheme 1Example of a (4 + 3) cycloaddition of an oxidopyridinium ion.Open in a separate windowScheme 2 Endo selective (4 + 3) cycloaddition of an oxidopyridinium ion.Upon our initial exploration into the chemistry of the latter (4 + 3) cycloadduct products, our intentions were the replacement of the trialkylsilyl group with a halogen atom such as iodine or bromine. We were particularly inspired by a report from the Zakarian group, who showed that vinylsilanes react with N-iodosuccinimide in hexafluoroisopropanol to afford iodoalkenes stereospecifically and in good yield (Scheme 3).5 However, when our (4 + 3) cycloadducts were exposed to such reaction conditions, the replacement of the C–Si bond by a C–I bond was not observed. Thus, when a mixture of 3a–c was treated with NIS in hexafluoroisopropanol (HFIP),6 a vinyl iodide was not formed. Instead, functionalization of the bridgehead carbon as a hexafluoroisopropyl ether was observed, a formal oxidative C–H activation process, affording 4a (Scheme 4).4 This communication reports further examples of this process and other results that provide some insights for future studies.Open in a separate windowScheme 3Zakarian''s vinyl iodide synthesis.Open in a separate windowScheme 4First example of the bridgehead oxidation of a (4 + 3) cycloadduct.The substrates for the process are known compounds, produced in our laboratories using methodology we developed for the (4 + 3) cycloaddition of oxidopyridinium ions with dienes.4 From the very first example of the process, we noted an interesting phenomenon. For example, entry 1 of Scheme 4) makes use of an inseparable mixture of diastereomers as the starting material. All three of the diastereomers are detectable by 1H NMR in the mixture. Nevertheless, the product of the oxidation is a single diastereomer. The structure of 4a was supported by 1H NMR data. The fate of the minor diastereomers of the starting material is unknown at this time, as nothing else could be isolated from the reaction mixture. This is also true for similar entries in EntrySubstrateaTimea (h)ProductYieldb (%)1 1 522 23 543 1.5 524 22 515 21 446 23 387 23 168 5.5 339c 2 3510c 3 5911 4 3112 23 64Open in a separate windowaThe major diastereomer/regioisomer of the mixture is shown.bYields are based on the entire mass of the starting material, including isomers that may not have given rise to the product observed.cThe substitution pattern of the major isomer is given in parentheses.In general, the reactions were conducted with 0.25 mmol of substrate dissolved in 4 mL of hexafluoroisopropanol (HFIP). The stirred solution was cooled to 0 °C in an ice bath. After addition of N-iodosuccinimide (1.5 equiv.), the mixture was allowed to slowly warm to the room temperature. The reaction was monitored by TLC until starting material was completely consumed.The substrates that contained an alkyl substituent at C-2 and a trialkylsilyl group at C-4 afforded product in about 50% yield (entries 1–4, 8Not all substrates bearing an alkyl substituent at C-2 and a trialkylsilyl group at C-4 afforded product in the 50% range. The lower yield of 24a from 23a–b, is perhaps due to the phenyl group on the silicon being sufficiently reactive to cause side product formation (entry 11, Fig. 1 shows substrates that did not afford bridgehead oxidation product and led to complex reaction mixtures when treated with NIS. It does appear that the silyl group provides protection to the alkene, but the effect is clearly not universal.Open in a separate windowFig. 1Substrates that did not lead to oxidation.We attempted to modify the reaction conditions in the hope of obtaining different products or limit the need for HFIP as solvent using 3a–c as a model substrate. When LiI, NaI, or NaBr was added to the reaction mixture, oxidation products were obtained in low yield. No halogen incorporation was observed, though such products might be expected to be labile in any case, as they would likely be readily solvolyzed. Attempts to reduce the amount of HFIP used in the process (2–30 equiv. in MeCN) gave bridgehead substitution products in 5–46% yield with recovered starting material isolated in yields ranging from 29–64%. Addition of small amounts of trifluoroacetic acid to the standard reaction mixture led to decomposition. Excess NIS (15 equiv.) gave decomposition and NBS was not effective at all in producing the oxidation product under the standard reaction conditions.Our working mechanism for this reaction is shown in Scheme 5. Reaction of 3a with NIS activated by HFIP through hydrogen bonding produces the iminium ion 27a.7 Deprotonation of this intermediate with succinimide anion affords the bridgehead iminium species 28a, which is trapped with HFIP to produce the product 4a (Scheme 5). Inclusion of allyltrimethylsilane in the reaction to trap 28a led to a complex product mixture.Open in a separate windowScheme 5Proposed mechanism of the oxidation.In conclusion, we report a unique bridgehead oxidation process of selected (4 + 3) cycloadducts derived from oxidopyridinium ions. While this process will certainly possess limitations, it raises questions about mechanism and what other reagents might be used to effect such an oxidation in a more general way, and whether more general bridgehead functionalization might be possible through such a mechanism. We plan on addressing these questions. Results will be reported in due course.  相似文献   

17.
Simple and efficient synthesis of bicyclic enol-carbamates: access to brabantamides and their analogues     
Ondrej Zborský   udmila Petrovi ov  Jana Doh&#x;o&#x;ov  Jn Moncol  Rbert Fischer 《RSC advances》2020,10(12):6790
A novel synthetic approach towards the formation of the unusual bicyclic enol-carbamates, as found in brabantamides A–C, is reported. The bicyclic oxazolidinone framework was obtained in very good yield and with high E/Z selectivity from a readily available β-ketoester under mild reaction conditions using Tf2O and 2-chloropyridine tandem. The major E isomer was used in the synthesis of the brabantamide A analogue.

A simple and short synthesis of bicyclic enol-carbamates with high E/Z selectivity and the synthesis of brabantamide A analogue are presented.

Brabantamides A–C (1–3) (Fig. 1) were first isolated in 2000 from the culture extracts of Pseudomonas fluorescens.1 They displayed nanomolar inhibitory activity towards lipoprotein-associated phospholipase A2 (Lp-PLA2) and therefore they could be used in the treatment of the inflammatory diseases such as atherosclerosis.1–3 It was found that the sugar moiety is not necessary for the biological activity against Lp-PLA2. In contrast, deglycosylated brabantamides A and C showed improved inhibitory activity compared to their natural counterparts.3 Moreover, simplified brabantamide analogues with amide functional group and long alkyl side chains showed even higher inhibitory effect.4 Brabantamides A–C also exhibit significant activity against Gram-positive bacteria, fungi, and oomycetes.5,6 A recent study confirmed that the bicyclic scaffold and the long lipophilic side chain are essential for the antibacterial activity.7Open in a separate windowFig. 1Structures of brabantamides A–C (1–3). Rha = rhamnose.Surprisingly, only five reports concerning the synthesis of the bicyclic oxazolidinone with an exocyclic double bond have been reported to date.In 2000, Pinto et al. prepared a series of analogues of brabantamide A where the enol-carbamate 5 was first obtained by direct iodocyclization from acetylene derivative 4 and then converted into key ester 6 by carbonylation of the corresponding vinyl iodide in the presence of 2-(trimethylsilyl)ethanol and PdCl2 (Scheme 1a).4 Another synthesis of the related esters 9 has been reported by Snider et al. in 2006, employing Wittig reaction between stabilized ylides and bicyclic oxazolidindione 8 synthesized from l-proline 7 (Scheme 1b).8 Shortly after, bicyclic enol-carbamate 10 with an exocyclic methylene group was synthesized in one step from acetylene derivative 4 using gold(i) catalysts (Scheme 1c).9 Very recently, Witte et al. reported the synthesis of series of Z-analogues of brabantamides 12 by cyclization of β-ketoamides 11 using CDI (Scheme 1d).7 As a part of our ongoing research program aimed at the utilization of trifluoromethanesulfonic anhydride (Tf2O)/2-halopyridine (2-XPy) tandem in the synthesis of bioactive natural products and their analogues, we envisioned that similar reaction conditions could be effectively used to form the bicyclic oxazolidinone framework as found in brabantamides.Open in a separate windowScheme 1Literature syntheses of bicyclic enol-carbamates and method proposed herein.The combination of Tf2O/2-XPy was extensively and successfully used in the amide activation10,11 as well as in generating isocyanate species from N-Boc and N-Cbz protected amines.12 We anticipated that the N-Boc-protected β-ketoester 13 could react in its enolate form with in situ generated isocyanate ion by intramolecular 5-endo-dig cyclization to give bicyclic enol-cyclocarbamate 14 (Scheme 1e).At the start of our investigation, model β-ketoester 15, prepared from N-Boc-l-proline,7 was chosen as the model substrate to identify optimal reaction conditions (12c the initial using of 1.5 equivalents of Tf2O and 3 equivalents of 2-ClPy led to a full conversion of the substrate 15 in 15 minutes (monitored by TLC) (13 Any variation of the amount of 2-ClPy did not have any positive impact on the reaction ( EntryTf2OBaseTime (min)16a : 16baYieldb (%)11.5—60—–c21.5Et3N60—–c31.5DMAP60—–c41.5Pyridine60—–c51.5d2,6-Lutidine60—–c61.5dDBU6090 : 1041e,f71.52-ClPy1593 : 753e 8 1.1 2-ClPy 15 85 : 15 80 e 91.12-ClPy (1.5 equiv.)4085 : 1575e101.12-ClPy (5 equiv.)1587 : 1364e111.12-FPy1589 : 1171e121.12-BrPy7086 : 1473e131.12-IPy9086 : 1468e14Witte''s protocolgOvernight50 : 5036eOpen in a separate windowaRatio determined by 1H NMR of the crude reaction mixture.bIsolated combined yield.cTraces of products.dReactions performed with 1.1 equiv. of Tf2O did not lead to full conversion of ester 15.eReactions were performed on 1 mmol of ester 15.fReaction mixture contained a large amount of unidentified by-products.gReaction conditions: (1) TMSOTf (2 equiv.), CH2Cl2, 0 °C, 1 h. (2) CDI (1.5 equiv.), CH2Cl2, 0 °C – rt, overnight.7It is noteworthy that the reaction can be performed on a gram scale without affecting the yield and both isomers are easily separable by FCC (see the ESI).The 1H and 13C NMR data of the major E isomer 16a were consistent with those published previously.8 Possible racemization in the course of the reaction was dismissed based on the comparing specific optical rotation with the published data for 16a ([α]22D = −261.1 (c 1.01, MeOH); ref. 8: [α]22D = −207 (c 1.0, MeOH)). Most importantly, X-ray crystallographic analysis of 16a (Fig. 2; see the ESI for further details)14 confirmed its absolute configuration on the C-7a carbon atom.Open in a separate windowFig. 2Molecular structure of the enol-carbamate E-16a confirmed by X-ray crystallographic analysis.The minor Z isomer 16b was isolated for the first time as the pure compound and was fully characterized. Its structure was assigned on the basis of its 1H, 13C, COSY, HSQC, and HMBC NMR spectra.A plausible mechanism of the cyclization of β-ketoester 15 was based upon previous works12ac and it is depicted in Scheme 2. Isocyanate cation II, as a key intermediate, can be formed directly from iminium triflate I (path A) or through the formation of carbamoyl triflate III with subsequent elimination of triflate ion spontaneously (path B). Ester enolate moiety IV then reacts as O-nucleophile via 5-endo-dig cyclization and leads predominantly to the formation of the enol-carbamate 16a.Open in a separate windowScheme 2Plausible mechanism of the cyclization β-ketoester 15.Next, the optimized conditions were briefly applied in the synthesis of the brabantamide A analogue 21 (Scheme 3). Starting β-ketoester 18 was synthetized in two steps in a 70% yield using both commercially available N-Boc-d-proline 17 and 2-(trimethylsilyl)ethanol. It ought to be mentioned that previously examined hydrolysis of the corresponding methyl ester 16a under acidic as well as basic conditions failed due to the instability of the bicyclic enol-carbamate.3,8Open in a separate windowScheme 3Synthesis of the brabantamide A analogue 21.Subsequent cyclization of ester 18 using optimized reaction conditions afforded enol-carbamate 19 in 76% yield as a mixture of E and Z isomers in a ratio of 89 : 11. After isolation of the major isomer E-19a, it was treated with TBAF, providing free acid 20 in 89% yield. Finally, an amidation of 20 with tetradecylamine in the presence of EDCI gave amide 21 in moderate 40% yield. Both free acid 20 and amide 21 were fully characterized for the first time and their structures were assigned on the basis of its 1H, 13C, COSY, HSQC, and HMBC NMR spectra. Moreover, their structures were unambiguously confirmed by X-ray crystallographic analysis (Fig. 3; see ESI for further details).14Open in a separate windowFig. 3Molecular structures of acid 20 (top) and amide 21 (bottom) confirmed by X-ray crystallographic analysis.  相似文献   

18.
Metal-free synthesis of 1,4-benzodiazepines and quinazolinones from hexafluoroisopropyl 2-aminobenzoates at room temperature     
Jiewen Chen  En Liang  Jie Shi  Yinrong Wu  Kangmei Wen  Xingang Yao  Xiaodong Tang 《RSC advances》2021,11(9):4966
Herein, we describe the novel reactivity of hexafluoroisopropyl 2-aminobenzoates. The metal-free synthesis of 1,4-benzodiazepines and quinazolinones from hexafluoroisopropyl 2-aminobenzoates has been developed at room temperature. These procedures feature good functional group tolerance, mild reaction conditions, and excellent yields. The newly formed products can readily be converted to other useful N-heterocycles. Moreover, the products and their derivatives showed potent anticancer activities in vitro by MTT assay.

A metal-free synthesis of 1,4-benzodiazepines and quinazolinones from hexafluoroisopropyl 2-aminobenzoates has been developed at room temperature.

Benzodiazepines (BDZs), especially 1,4-benzodiazepines, are privileged motifs in pharmaceuticals.1 For examples, oxazepam is used to treat anxiety disorders or alcohol withdrawal symptoms; triazolam is used to treat insomnia (Scheme 1). Until now, some synthetic approaches to 1,4-benzodiazepine skeletons have been developed include isocyanide-based multicomponent reactions,2 cycloadditions,3 metal-catalyzed tandem reactions,4 and redox-neutral [5+2] annulation with o-aminobenzaldehydes.5 However, these procedures have some limitations involving unavailable materials, several steps, harsh reaction conditions and low yields. α-Haloamides are widely used to synthesize N-heterocycles.6 Recently, some groups reported the synthesis of 1,4-benzodiazepines with α-haloamides.7 Kim and coworkers developed a [4+3]-annulation reaction between α-haloamides and isatoic anhydrides for 1,4-benzodiazepines, but the reaction required 80 °C reaction temperature and provided an unsatisfactory yield.7a Singh''s group reported a two-step method to construct 1,4-benzodiazepines from α-haloamides and anthranils, but the anthranils are not readily available substrates and the second step also required 80 °C reaction temperature.7b Quinazolinones are a significant class of heterocycles that widely occur in natural products and pharmaceuticals (Scheme 1).8 These compounds exhibit a range of biological activities including anticancer, antibacterial, antiinflammatory, antifungal, etc. Due to their significant value, the synthesis of quinazolinones has attracted considerable attention. The reported synthetic methods can be summarized as: (i) condensation of 2-aminobenzamides with carbonyl compounds;9 (ii) oxidative cyclization of primary alcohols with 2-aminobenzamides or 2-aminobenzonitriles;10 (iii) metal-catalyzed coupling/cyclization reactions;11 and (iv) palladium-catalyzed carbonylation reactions.12 But these synthetic methods also had some disadvantages. Thus, it is highly desirable to develop new available reagents for synthesizing the useful N-heterocycles such as benzodiazepines and quinazolinones with good yields under mild reaction conditions.Open in a separate windowScheme 1Structures of representative 1,4-benzodiazepine and quinazolinone drugs.In the past few years, 2-aminobenzoates have been used for the synthesis of N-heterocycles via [4+n] cyclization (Scheme 2a).13 However, harsh reaction conditions such as high reaction temperatures and strong bases or acids were required to effect alkoxy leaving. When we tried to synthesize benzodiazepines or quinazolinones with methyl or tert-butyl 2-aminobenzoates, we failed. Recently, hexafluoroisopropanol (HFIP) has attracted a lot of attention when used as solvent or substrate, due to its special properties.14 When we used isatoic anhydrides as substrates and NEt3 as base in HFIP at room temperature, we unexpectedly discovered that hexafluoroisopropyl 2-aminobenzoates were completely formed. We supposed hexafluoroisopropyl 2-aminobenzoates were good synthons for the synthesis of N-heterocycles. Herein, we report metal-free procedures for the synthesis of 1,4-benzodiazepines and quinazolinones from hexafluoroisopropyl 2-aminobenzoates at room temperature with excellent yields (Scheme 2b).Open in a separate windowScheme 2Synthesis of N-heterocycles from 2-aminobenzoates.We examined the annulation reaction with hexafluoroisopropyl 2-aminobenzoate (1a) and α-bromoamide (2a) as the model substrates. Initially, when the reaction was performed with 1 equiv. of Et3N in HFIP at room temperature for 0.5 h, 3a was formed, but cyclization product 4a was not obtained. We thought the transformation from 3a to the product 4a needing to release one molecule of HFIP, and the transformation maybe be inhibited when HFIP was used as solvent. So we removed the solvent HFIP under vacuum and added 2.0 mL DMF to react for 0.5 h. Pleasingly, the desired product 4a was obtained in 68% yield ( EntryBaseSolvent ASolvent BYield (%)1Et3NHFIPDMF682Cs2CO3HFIPDMF973NaHCO3HFIPDMFn.d.4K2CO3HFIPDMFn.d.5DBUHFIPDMF616NaOHHFIPDMF937DIPEAHFIPDMF638—HFIPDMF09Cs2CO3DMSO—010Cs2CO3DMA—011Cs2CO3MeCN—012Cs2CO3Toluene—013Cs2CO3HFIPDMA7114Cs2CO3HFIPDMSO4115Cs2CO3HFIPMeCN4016Cs2CO3HFIPDioxane3817Cs2CO3HFIPNMPn.d.18Cs2CO3HFIPToluenen.d.19bCs2CO3HFIPDMF92Open in a separate windowaReaction conditions: unless otherwise noted, all reactions were performed with 1a (0.3 mmol), 2a (0.3 mmol), and base (0.3 mmol) in solvent A (3.0 mL) at room temperature for 0.5 h, then the solvent A was removed under vacuum and solvent B (2.0 mL) added to react for 0.5 h. Isolated yield.bYield on a 3.0 mmol scale.After determining the optimized reaction conditions, the scope of the cyclization reaction for 1,4-benzodiazepines was examined ( Open in a separate windowaReaction conditions: 1 (0.3 mmol), 2 (0.3 mmol), and Cs2CO3 (0.3 mmol) in solvent HFIP (3.0 mL) at room temperature for 0.5 h, then the HFIP was removed under vacuum and added DMF (2.0 mL) to continue to react for 0.5 h. Isolated yield.When hexafluoroisopropyl 2-aminobenzoates reacted with amidines hydrochloride in the presence of base, quinazolinones were produced. Then, we optimized the reaction conditions to enhance the yields of the quinazolinones (see the ESI for more details). With optimum conditions in hand, substrate scope for the synthesis of quinazolinones was next investigated ( Open in a separate windowaReaction conditions: 1 (0.3 mmol), 5 (0.36 mmol), K3PO4 (0.45 mmol), DMF (2.0 mL) at room temperature for 10 h. Isolated yields.To probe the reaction mechanism, several preliminary experiments were conducted (Scheme 3). Under standard reaction conditions, methyl 2-aminobenzoates 7 reacted with 2a to provide compound 8 in 92% yield, and 4a was not detected. This control experiment indicated the importance of hexafluoroisopropyl (Scheme 3, eqn (1)). Treatment of methyl 2-aminobenzoates 7 with 5a in the presence of K3PO4 did not furnish any product 6a, and 3a did not convert at all (Scheme 3, eqn (2)). On the basis of the control experiments and previous reports, we propose a possible mechanism. First, aza-oxyallyl cation A is formed from α-bromoamide with Cs2CO3.15 Whereafter, aza-oxyallyl cation A combines with 1a to produce compound 3a.15b The product 4a is obtained via intramolecular nucleophilic substitution, releasing a molecule of hexafluoroisopropanol. The nucleophilic attack of 5a onto 1a provides the intermediate B. Subsequently, product 6a is formed by intramolecular nucleophilic addition/deamination cyclization.Open in a separate windowScheme 3Control experiments and possible reaction mechanism.In order to address the potential synthetic application of our methods, the transformations of the obtained 1,4-benzodiazepines and quinazolinones were performed (Scheme 4). Compound 9 was formed from 4a through cleavage of the N–O bond with Mo(CO)6 (Scheme 4, eqn (1)). The quinazolinones can be transformed into substituted quinazolines with anilines or phenols as nucleophilic reagents in the presence of BOP and DBU (Scheme 4, eqn (2) and (3)).Open in a separate windowScheme 4Derivatization of products.We next investigated the cytotoxicity of the products and their derivatives against cancer cell lines (A549, HCT116 and MCF7) by MTT assay, with 5-fluorouracil (5-FU) as the positive control. To our delight, some products and their derivatives exhibited potent inhibitory activities, and some of them showed better inhibitory activities than 5-Fu (CompoundsIC50 (μM)A549HCT116MCF74e64.69 ± 7.3533.27 ± 5.8440.32 ± 0.496c35.11 ± 3.4026.61 ± 1.2658.12 ± 3.456d52.32 ± 2.8523.58 ± 1.5081.32 ± 2.806f82.89 ± 10.3459.59 ± 1.6038.52 ± 1.836g67.00 ± 8.2432.90 ± 0.6042.54 ± 3.796l19.56 ± 1.1617.73 ± 2.3225.00 ± 5.3010a14.79 ± 1.1526.31 ± 3.9529.70 ± 0.0910b5.98 ± 0.4215.41 ± 4.4121.12 ± 1.0611a68.54 ± 3.7017.84 ± 3.1375.84 ± 2.5011b94.76 ± 1.1425.14 ± 5.3167.13 ± 3.655-Fu>10013.03 ± 2.8029.58 ± 12.86Open in a separate windowIn summary, we have developed novel and simple approaches for the synthesis of 1,4-benzodiazepines and quinazolinones from hexafluoroisopropyl 2-aminobenzoates with α-bromoamides or amidines hydrochloride. These protocols feature readily available starting materials, mild reaction conditions, good functional group tolerance, and excellent yields. In addition, the newly obtained products and their derivatives showed potent anticancer activities in vitro by MTT assay. Further studies on the synthesis of other N-heterocycles from hexafluoroisopropyl 2-aminobenzoates are in progress.  相似文献   

19.
Enantioselective amination of 4-alkylisoquinoline-1,3(2H,4H)-dione derivatives     
Cheng Cheng  Ying-Xian Li  Xue-Min Jia  Ji-Quan Zhang  Yong-Long Zhao  Wei Feng  Lei Tang  Yuan-Yong Yang 《RSC advances》2020,10(70):42912
A mild and efficient enantioselective amination of 4-alkylisoquinoline-1,3(2H,4H)-dione derivatives was established, which is compatible with a broad range of substrates and delivers the final products in excellent yields (up to 99%) and ee values (up to 99%) with low catalyst loading (down to 1 mol%). The synthetic potential of this methodology was also demonstrated in the gram scale level.

A mild and efficient enantioselective amination of 4-alkylisoquinoline-1,3(2H,4H)-dione derivatives was established and a broad range of amination products were prepared in excellent yields and ee values with low catalyst loading.

Isoquinolinedione, bearing the carbon skeleton of tetrahydroisoquinoline (THIQ), is an important structural motif present in bioactive compounds and natural products with a broad array of biological properties.1 However, the construction of isoquinolinedione, particularly the chiral version, is currently underdeveloped,2 and the reported methods heavily rely on the radical-initiated addition–cyclization of activated alkenes to prepare this structural motif that hard to be further diversified.3 From a pharmaceutical point of view, the presence of heteroatoms (such as nitrogen) is essential for their biological activity (Fig. 1).4 Therefore, the introduction of other functional groups or heteroatoms into this framework is a pressing issue to be addressed.Open in a separate windowFig. 1Bioactive compounds bearing isoquinolinedione or tetrahydroisoquinoline core structure.On a different note, amine attached to a stereogenic center is a ubiquitous structure in natural products and bioactive compounds and becomes impetus for continuous exploration.5 Using azodicarboxylates or nitrosoarenes as electrophilic amine sources,6 activated substrates such as 1,3-dicarbonyl compounds and pyrazolones could be readily transformed into the corresponding amination products in high ee and yields.7 With the pioneering work of List and Jørgensen, the α-amination of aldehydes could be realized through enamine activation.8 The α-amination of less activated substrates such as nitroisoxazole derivatives could be realized via phase-transfer catalysis.9 Recently, cyclic ketones or vinyl ketones were transformed into the corresponding amination products via organo- or metal catalysis.10 Surprisingly, reports on the amination of heterocyclic compounds are very limited, and they majorly focus on the oxindole scaffold.11 Therefore, the construction of other pharmaceutical relevant α-amination heterocyclic compounds would be a meaningful work.12 In addition, the organo-catalyzed asymmetric amination reactions generally require relatively high catalyst loading to achieve the optimal yields and enantioselectivities; for this reason, the development of an efficient amination protocol would still be highly desirable.Recently, our group reported the amination of 4-arylisoquinolinedione via organo-catalysis.13 However, due to the attenuated reactivity at low temperatures, high catalyst loading is required for satisfactory yields and enantioselectivities, and the substrate scope is limited to 4-aryl substituents. To further expand the scope of this reaction, we tried to extend this amination methodology to 4-alkylisoquinolinedione derivatives.Our study commenced with 2-benzyl-4-butylisoquinoline-1,3(2H,4H)-dione 4a and di-tert-butyl azodicarboxylate 5 as model substrates for condition optimization ( Entrya7 (mol%)SolventYield 6ab (%)eec (%)1d10CHCl35098210CHCl37197310Toluene8274410Ether9580510THF9921610DCM9990710Chlorobenzene8594810CH2ClCH2Cl999895CH2ClCH2Cl979710e2CH2ClCH2Cl999711f,g1CH2ClCH2Cl839512f,g1CH2ClCH2Cl8893Open in a separate windowaAll reaction was conducted with 0.2 mmol compound 4a, 0.44 mmol compound 5, in 0.5 mL solvent and reacted at 25 °C for 24 h.bIsolated yield.cThe ee was determined by HPLC analysis.dReaction was conducted at 5 °C and reacted for 24 h.eReaction was run for 35 h.fReaction was reacted for 72 h.gReaction was conducted at 40 °C.Further solvent optimization reveals that DCM gives the best yield along with very good ee (14Substrate scope for the amination reaction
EntryaRProductYieldb (%)eec (%)
1 n-Propyl6a9997
2 i-Butyl6b9694
3 i-Propyl6c9981
4PhCH2CH26d9493
5Ph6e9997
64-MeC6H46f9992
74-OMeC6H46g9097
84-FC6H46h9696
94-ClC6H46i9999
104-BrC6H46j9999
113-MeC6H46k9976
123-BrC6H46l9989
132-OmeC6H46m9993
142-MeC6H46n9293
152-ClC6H46o9598
163,4,5-OmeC6H26p9987
172-Naphthyl6q9999
182-Indolyl6r9082
193-Indolyl6s9999
202-Me-3-indolyl6t9996
212-Fural6u9997
Open in a separate windowaReactions were run on a 0.03 mmol 1 and 0.036 mmol 2 with the 2 mol% catalyst in 500 μL solvent at 25 °C for 48 h.bYield was based on the isolated product of 3.cThe ee was determined via HPLC analysis.To demonstrate the practical synthetic application of current protocol, the gram scale synthesis of chiral 6i has been demonstrated (Scheme 1). The product was produced in excellent yield and ee value at the 2 mmol scale. Moreover, a synthetically desirable amino product could be obtained from the cleavage of the N–N bond and deprotection of the Boc group in two steps with very good yield and ee value (Scheme 2).15Open in a separate windowScheme 1Gram scale preparation of 6i.Open in a separate windowScheme 2Transform the product into amino product.In an effort to account for the observed stereocontrol of the reaction, a plausible reaction mechanism is proposed in Scheme 3. With the previously established bifunctional catalyst by Rawal et al.,16 the isoquinolinedione was activated by the alkyl amine moiety to attack the azodicarboxylate that was activated by the squaramide moiety via hydrogen bonding interactions in a well-defined manner to deliver the final product in S configuration.17 The outcome in this study is in accordance with our previous reports13 as the benzyl group alleviates the steric hindrance of the substituted phenyl ring from the reaction center and delivers the product in a high ee value (Open in a separate windowScheme 3Proposed mechanism for the amination reaction.To summarize, a highly enantioselective amination methodology with low catalyst loading was established (down to 1 mol%), which is compatible with a broad range of substrates and delivers the final products in excellent yields (up to 99%) and ee values (up to 99%). Moreover, the maintaining of yield and ee in up-scale preparation clearly demonstrates the synthetic potential of this methodology. Most importantly, this reaction is mild and operationally simple and could be performed without the exclusion of air or moisture at room temperature.  相似文献   

20.
Solvent-free synthesis of 3,5-isoxazoles via 1,3-dipolar cycloaddition of terminal alkynes and hydroxyimidoyl chlorides over Cu/Al2O3 surface under ball-milling conditions     
Rafael A. Hernandez R.  Kelly Burchell-Reyes  Arthur P. C. A. Braga  Jennifer Keough Lopez  Pat Forgione 《RSC advances》2022,12(11):6396
Scalable, solvent-free synthesis of 3,5-isoxazoles under ball-milling conditions has been developed. The proposed methodology allows the synthesis of 3,5-isoxazoles in moderate to excellent yields from terminal alkynes and hydroxyimidoyl chlorides, using a recyclable Cu/Al2O3 nanocomposite catalyst. Furthermore, the proposed conditions are reproducible to a 1.0-gram scale without further milling time variations.

A practical and scalable mechanochemical 1,3-dipolar cycloaddition between hydroxyimidoyl chlorides and terminal alkynes catalyzed by Cu/Al2O3 allows a quick access to 3,5-isoxazole derivatives.

The addition of oxygen or nitrogen-containing heterocycles in drug candidates has become a common feature of the recently approved drugs by the FDA.1,2 In particular, isoxazoles are common molecular scaffolds employed in medicinal chemistry due to the non-covalent interactions such as hydrogen bonding (through the N) and π–π stacking (by the unsaturated 5-membered ring).3–6 Within the isoxazole family, 3,5-isoxazoles (1) are regularly utilized as pharmacophores in medicinal chemistry.2,5,6 Selected examples including muscimol (GABAa agonist), isocarboxazib (antidepressant), isoxicam (anti-inflammatory), berzosertib (ATR kinase inhibitor), and sulfamethoxazole (antibiotic) are highlighted in Fig. 1.7–10Open in a separate windowFig. 1Examples of isoxazoles with pharmacological activity.Various methodologies to synthesize 3,5-isoxazoles have been developed over the years.7,9,11–13 Specifically, 1,3-dipolar cycloaddition between terminal alkynes (2) and nitrile oxides (4) formed in situ by deprotonation of hydroxyimidoyl chlorides (3) is a standard route to access 3,5-isoxazoles (1) (Fig. 2).7,9,14 Recent reports have sought to mitigate the environmental impact of this reaction by performing 1,3-dipolar cycloaddition under solvent-free conditions, using green solvents such as water or ionic liquids, under metal-free conditions, or using mild oxidants.14–28 However, these methodologies have a low atom economy, have a higher hazardous waste production, and are less energy efficient. Therefore, developing a greener methodology that enables rapid and efficient access to these scaffolds is highly desirable.Open in a separate windowFig. 21,3-Dipolar cycloaddition of terminal alkynes and nitrile oxides.Mechanochemistry has been recognized as an environmentally friendly technique as reactions can be performed under solvent-free conditions. Additionally, in some instances, work-up and purification are simplified or absent from procedures, and the process consumes less energy than other solution-based techniques.29–33 The use of mechanochemical techniques to synthesize isoxazoles is limited. Sherin et al. reported a synthesis of 3,5-isoxazoles (7) by grinding in a mortar and pestle curcumin derivatives (5), hydroxylamine (6), and sub-stoichiometric amounts of acetic acid to form the 3,5-isoxazole (7) in short times and excellent yields (Fig. 3a).34 Likewise, Xu et al. studied the synthesis of trisubstituted isoxazoles (10) via 1,3-dipolar cycloaddition of N-hydroxybenzimidoyl chlorides (8) and N-substituted β-enamino carbonyl (9) compounds by ball-milling (Fig. 3b) in high yields, short reaction times, in the absence of catalyst and liquid additives.35 To our knowledge, mechanochemical synthesis of 3,5-isoxazoles (1) from terminal alkynes (2) and hydroxyimidoyl chloride (3) has not been reported (Fig. 3c). The proposed methodology employs a planetary ball-milling technique that provides a route to access in large scale, short reaction times, and high atom economy the corresponding 3,5-isoxazoles (1). Additionally, it utilizes synthetically accessible or commercially available motifs such as terminal alkynes (2) and hydroxyimidoyl chlorides (3) that are recurrent or easily installed in many substrates. Herein, we report a mechanochemical 1,3-dipolar cycloaddition using the planetary ball-mill to synthesize a wide range 3,5-isoxazoles from a broad library of alkynes and (E,Z)-N-hydroxy-4-nitrobenzimidoyl chloride (3a), ethyl (E,Z)-2-chloro-2-(hydroxyimino)acetate (3b), hydroxycarbonimidic dibromide (3c), or (E,Z)-N-hydroxy-4-methoxybenzimidoyl chloride (3d) in moderate to excellent yields, in short reaction time, and with less waste production than in solution based reactions (Fig. 3c).Open in a separate windowFig. 3Previously reported synthesis of isoxazoles.We began our investigation by performing an optimisation of the 1,3-dipolar cycloaddition reaction between alkyne 2a and hydroxyimidoyl chlorides 3a by milling the selected substrates in a stainless-steel (SS) jar in the planetary ball-mill to obtain 3,5-isoxazole 1a (36,37Optimization of reaction conditionsa
EntryChanges from optimized conditionsYieldb (%) of 1a
1None72
2Milling for 10 min, 7 SS balls59
3Milling for 15 min 7 SS balls64
4Milling for 30 min58
5Milling for 40 min60
6Using 1.0 equiv. of 3a65
7Using 2.0 equiv. of 3a57
8Using K2CO371
9Using Cs2CO371
10Using CaCO344
11Using Ag2CO318
12Using NEt3N.R.
Open in a separate windowaReaction Conditions: 0.166 mmol of 2a, 0.250 mmol of 3a, 0.332 mmol of Na2CO3, SS beaker (50 mL capacity), 8 × SS milling balls (10 mm diameter), 20 min milling, 60 Hz.b 1H-NMR yields were measured using 1,3,5-trimethoxybenzene as an internal standard.Having optimized the milling time, we next attempted to improve the yield by varying the equivalents of hydroxyimidoyl chlorides 3a since reaction stoichiometry has been shown to impact the product formed during mechanochemical reactions.38,39 3,5-isoxazole 1a was obtained in lower yields when using equimolar amounts alkyne 2a to hydroxyimidoyl chlorides 3a (entry 6, 40–45 Likewise, increasing the equivalents of 18a from 1.0 to 2.0 equivalents lowered the yield of the reaction (entry 7, 46,47 Using triethylamine (NEt3) proved impractical as the addition of NEt3 to hydroxyimidoyl chlorides was highly exothermic in the absence of solvent (entry 12, 35,48 Therefore, a milling time optimization for other alkyne and hydroxyimidoyl chloride combinations revealed that the most optimal milling time was determined to be between 10 and 30 minutes (see ESI for milling time optimizations).As shown in Fig. 4, stannanyl isoxazole 1a and 1b, silyl isoxazole 1c, and phenyl isoxazole 1d were synthesized with satisfactory yields under the proposed conditions. To explain these results, we suggest an electronic argument. The electron-withdrawing character of the metal substituents, stannyl or silyl of alkyne 2a and 2b, respectively, accelerates the reaction by deactivating the alkyne moiety.49–51 It is observed that alkyne 2a bearing the alkylstannane substituent has a more pronounced effect than the alkyne with the silyl substituent (2b). Therefore, alkyne 2a was the most reactive as it reacted with hydroxyimidoyl chlorides 3a and 3b to synthesize 3,5-isoxazole 1a and 1b respectively, in short times and excellent yields (Fig. 4). On the other hand, ethynyltrimethylsilane (2b) was less reactive as it could only react with a more labile hydroxyimidoyl chlorides 3b to form 3,5-isoxazole 1c (Fig. 4). Comparably, we suggest that the phenyl substituent of alkyne 1c increases the polarizability of the molecule, resulting in deactivating the alkyne moiety. As a result, phenylacetylene (1c) reacted in excellent yields with hydroxyimidoyl chlorides 1b.42 In addition, we observed that the electronic nature of the hydroxyimidoyl chloride substituent affects the reactivity of the nitrile oxide dipole. Hydroxyimidoyl chlorides 3a containing an aromatic substituent with strong electron-withdrawing groups decreased the reactivity of the nitrile oxide.52,53 Consequently, the nitrile oxide synthesized in situ from hydroxyimidoyl chlorides 3a could only react with tributyl(ethynyl)stannane (2a). On the other hand, hydroxyimidoyl chlorides 3b was the most reactive due to the bearing of a weaker electron-withdrawing group such as the ester functional group.43,52 Unfortunately, other alkynes containing substituents such as esters, pyridines, or substituted arenes were not tolerated under these conditions. Previous reports demonstrated the effect of copper catalyst or copper additives to accelerate the reaction and obtain the 3,5-isoxazoles in a regioselective manner.15,53–59 Therefore, we aimed to investigate the effect of copper additives or catalysts on this reaction.Open in a separate windowFig. 4Catalyst-free mechanochemical synthesis of 3,5-isoxazoles.Although the mechanochemical synthesis of 3,5-isoxazoles using copper(ii) catalyst is unprecedented, 1,2,3-triazoles have been synthesized in this way with copper(ii) salts and copper(ii) ions in alumina nanocomposites (Cu/Al2O3).60,61 We investigated the effect of Cu/Al2O3 (see ESI for XPS spectrum) and copper salts using methyl propiolate (2d) and (E,Z)-2-chloro-2-(hydroxyimino)acetate (3b) as model substrates (ii) in the synthesis of 3,5-isoxazolesac
EntryCu(ii)EquivalentsTime (min)Yieldb (%) 19e
1Cu/Al2O30.14 of Cu(ii)1073
2076
3079
4064
5056
2Cu(NO3)2·2.5H2O0.13078
3Cu(NO3)2·2.5H2O1.03084
4Cu(OAc)2·H2O1.03088
5Cu(OTf)21.03076
6CuCl2·H2O1.03076
7Cu2CO3(OH)22.03036
Open in a separate windowaReaction conditions: 0.220 mmol of 2d, 0.330 mmol of 3b, 0.220 mmol of Na2CO3, 0.440 mmol (14 mol%) of Cu/Al2O3, SS beaker (50 mL capacity), 8 × SS milling balls (10 mm diameter), 60 Hz.b 1H NMR yields were measured using 1,3,5-trimethoxybenzene as an internal standard.cSee ESI for solid-state characterization by FT-IR and MALDI-TOF-MS of reaction crude 1e.We observed a significant increase in yield and regioselective control when using sub-stoichiometric amounts of copper (0.14 equivalents or 14 mol%) of Cu/Al2O3 or 10 mol% of Cu(NO3)2·2.5H2O while milling the reagents for 30 minutes (entries 1 and 2, ii), it was observed that Cu(OAc)2·H2O performs similarly to Cu(NO3)2·2.5H2O (entry 4, 62–64We decided to continue our investigations using Cu/Al2O3 as the catalyst can be filtered and washed with solvent, thereby facilitating catalyst recovery and recycling (Fig. 5).60Open in a separate windowFig. 5(a) Filtration of the Cu/Al2O3 catalyst after the first run. (b) Colour change of the Cu/Al2O3 catalyst after recycling. From left to right. (left) Fresh catalyst: blue. (middle) First recycle: green. (right) Second recycle: brown.The Cu/Al2O3 catalyst effect was not exclusively beneficial for the cycloaddition with methyl propiolate (2d) (3,5-isoxazole 1e, Fig. 6). This system improves the reactivity of hydroxyimidoyl chlorides 3a, 3b, 3c, and 3d and other alkynes inaccessible under copper-free conditions, thus allowing access to a broader library of 3,5-isoxazoles (Fig. 6). Moreover, the presence of Cu/Al2O3 nanocomposite as part of the reaction conditions is not impaired by the presence of labile substituents such as silanes (1c, f–h), alkyl halides (1i–j), and boronic esters (1n) (Fig. 6). However, the presence of alkyl stannane substituents in the dipolarophile (2a) was not tolerated with Cu/Al2O3 catalyst, and no product was observed. Furthermore, Cu/Al2O3 enhances the reactivity of dipolarophiles bearing arenes with electron-donating substituents (EDG) (1o–q) and electron-withdrawing groups (EWG) (1n, 1r-2) when coupled with hydroxyimidoyl chlorides 3a and 3b. Additionally, pyridine substituents were more reactive towards the more reactive hydroxyimidoyl chlorides (3b) (3,5-isoxazole 1s, Fig. 6). Ethynyltrimethylsilane (1c) reacted efficiently with hydroxyimidoyl chlorides bearing EWG (3a, 3b, and 3c) to form the respective isoxazoles 1c,1f, and 1h, where silyl isoxazole 1c is obtained in higher yields compared to copper-free conditions (1c, Fig. 4). Hydroxyimidoil chloride bearing EDG (3d); resulted incompatible with terminal alkyne 2b and silyl isoxazole 1g was obtained in lower yields than with EWG in the hydroxyimidoyl chloride. However, terminal alkynes having an aliphatic substituent (2e and 2f) showed greater reactivity towards hydroxyimidoil chloride (3d) bearing EDG; consequently, aliphatic isoxazole 1j was obtained in higher yields than 1i. Then, we evaluated the impact of our conditions in the synthesis of 3,5-isoxazole 1f on a 1.0-gram scale (10.18 mmol). We were pleased to observe that the optimized Cu/Al2O3 conditions can be translated with excellent reproducibility from a 100 mg scale to a 1.0-gram scale without extending the milling time of the reagents (Fig. 6).Open in a separate windowFig. 6Mechanochemical synthesis of 3,5-isoxazoles reaction scope. aAll shown yields are isolated yields. bReaction performed in 1.0-gram scale.The practicality of the proposed methodology allows the recovery of Cu/Al2O3 nanocomposite catalyst directly after the milling of the reagents. In addition, the catalyst recovery allowed investigating the reusability of the recovered catalyst. The Cu/Al2O3 was reused on four occasions, and it was observed that 3,5-isoxazole 1f was obtained successfully with only a minimal drop in yield with each subsequent use for the first two recycling cycles (Fig. 7). The decrease in yield is explained by the decrease in the concentration of active Cu species in the Cu/Al2O3 nanocomposite (see ESI). ICP-MS analysis demonstrates that the Cu concentration of the first recycling represents a decrease of 1.24-fold (with respect to the fresh catalyst); thus, similar yields are obtained compared to the fresh catalyst (Fig. 7). However, the decrease in Cu concentration becomes more substantial for the second and third reuse with a decrease of 2.42 and 6.48-fold, respectively. Therefore, a considerable decrease in the yield of isoxazole 1f is observed. Furthermore, a change in the oxidation state and the bonding of the supported Cu(ii) ions. X-ray Photoelectron Spectroscopy (XPS) analysis of the first and second recycled catalyst reveals that the characteristic satellite signals of Cu(ii) found at about 942.8 eV are weak while the satellite signal at 963.2 eV is absent. Additionally, the 2p3/2 signal at about 933–934 eV is wider than in the fresh sample (see ESI for XPS spectra of the fresh, Fig. S3 for first recycling and Fig. S4 for second recycling). These observations suggest that the supported Cu(ii) is reduced to Cu(0) and CuO is formed with each subsequent recycling.65–67Open in a separate windowFig. 7Cu/Al2O3 efficiency study in the synthesis of 3,5-isoxazole 1f.Lastly, we evaluated the sustainability of the proposed mechanochemical 1,3-dipolar cycloaddition conditions by comparing E-factor for the synthesis of 3,5 isoxazoles 1d and 1f to previously reported solution-based conditions (Fig. 8).54,68 Using E-factor, the values calculated for the planetary ball milling conditions (pathway a and c, Fig. 8) demonstrate the sustainability of this methodology compared to solution-based reactions (pathway b and d) (see ESI for calculations). With our conditions, the absence of organic solvent is the most significant factor contributing to lowering the E-factor.69 Time differences were also another factor of comparison with previously reported solution-based conditions. Our mechanochemical conditions did not surpass 60 minutes, contrary to the reported solution-based conditions that require at least two hours to synthesize the desired 3,5-isoxazoles. Furthermore, our conditions did not show any sensitivity to oxygen or moisture present in the air as all reactions were performed in an open atmosphere.Open in a separate windowFig. 8Comparative green metrics of the proposed methodology to previously reported solution-based methodologies.  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