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 15 毫秒
1.
The Mn–oxygen species have been implicated as key intermediates in various Mn-mediated oxidation reactions. However, artificial oxidants were often used for the synthesis of the Mn–oxygen intermediates. Remarkably, the Mn(v)–oxo and Mn(iv)–peroxo species have been observed in the activation of O2 by Mn(iii) corroles in the presence of base (OH) and hydrogen donors. In this work, density functional theory methods were used to get insight into the mechanism of dioxygen activation and formation of Mn(v)–oxo. The results demonstrated that the dioxygen cannot bind to Mn without the axial OH ligand. Upon the addition of the axial OH ligand, the dioxygen can bind to Mn in an end-on fashion to give the Mn(iv)–superoxo species. The hydrogen atom transfer from the hydrogen donor (substrate) to the Mn(iv)–superoxo species is the rate-limiting step, having a high reaction barrier and a large endothermicity. Subsequently, the O–C bond formation is concerted with an electron transfer from the substrate radical to the Mn and a proton transfer from the hydroperoxo moiety to the nearby N atom of the corrole ring, generating an alkylperoxo Mn(iii) complex. The alkylperoxo O–O bond cleavage affords a Mn(v)–oxo complex and a hydroxylated substrate. This novel mechanism for the Mn(v)–oxo formation via an alkylperoxo Mn(iii) intermediate gives insight into the O–O bond activation by manganese complexes.

DFT calculations revealed a novel mechanism for the formation of Mn(v)–oxo in the dioxygen activation by a Mn(iii) corrole complex involving a Mn(iii)–alkylperoxo intermediate.  相似文献   

2.
The catalytic C–H alkylation with alkenes is of much interest and importance, as it offers a 100% atom efficient route for C–C bond construction. In the past decade, great progress in rare-earth catalysed C–H alkylation of various heteroatom-containing substrates with alkenes has been made. However, whether or how a heteroatom-containing substrate would influence the coordination or insertion of an alkene at the catalyst metal center remained elusive. In this work, the mechanism of Sc-catalysed C–H alkylation of sulfides with alkenes and dienes has been carefully examined by DFT calculations, which revealed that the alkene insertion could proceed via a sulfide-facilitated mechanism. It has been found that a similar mechanism may also work for the C–H alkylation of other heteroatom-containing substrates such as pyridine and anisole. Moreover, the substrate-facilitated alkene insertion mechanism and a substrate-free one could be switched by fine-tuning the sterics of catalysts and substrates. This work provides new insights into the role of heteroatom-containing substrates in alkene-insertion-involved reactions, and may help guide designing new catalysis systems.

The alkene insertion via the heteroatom-containing substrate facilitated mechanism were computationally revealed in rare-earth-catalyzed C–H alkylation of sulfides and other heteroatom-containing substrates such as pyridines and anisoles.  相似文献   

3.
The reaction of the trimetallic clusters [H2Os3(CO)10] and [Ru3(CO)10L2] (L = CO, MeCN) with 2-ethynylpyridine has been investigated. Treatment of [H2Os3(CO)10] with excess 2-ethynylpyridine affords [HOs3(CO)10(μ-C5H4NCH=CH)] (1), [HOs3(CO)93-C5H4NC Created by potrace 1.16, written by Peter Selinger 2001-2019 CH2)] (2), [HOs3(CO)93-C5H4NC Created by potrace 1.16, written by Peter Selinger 2001-2019 CCO2)] (3), and [HOs3(CO)10(μ-CH Created by potrace 1.16, written by Peter Selinger 2001-2019 CHC5H4N)] (4) formed through either the direct addition of the Os–H bond across the C Created by potrace 1.16, written by Peter Selinger 2001-2019 C bond or acetylenic C–H bond activation of the 2-ethynylpyridine substrate. In contrast, the dominant pathway for the reaction between [Ru3(CO)12] and 2-ethynylpyridine is C–C bond coupling of the alkyne moiety to furnish the triruthenium clusters [Ru3(CO)7(μ-CO){μ3-C5H4NC Created by potrace 1.16, written by Peter Selinger 2001-2019 CHC(C5H4N) Created by potrace 1.16, written by Peter Selinger 2001-2019 CH}] (5) and [Ru3(CO)7(μ-CO){μ3-C5H4NCCHC(C5H4N)CHCHC(C5H4N)}] (6). Cluster 5 contains a metalated 2-pyridyl-substituted diene while 6 exhibits a metalated 2-pyridyl-substituted triene moiety. The functionalized pyridyl ligands in 5 and 6 derive via the formal C–C bond coupling of two and three 2-ethynylpyridine molecules, respectively, and 5 and 6 provide evidence for facile alkyne insertion at ruthenium clusters. The solid-state structures of 1–3, 5, and 6 have been determined by single-crystal X-ray diffraction analyses, and the bonding in the product clusters has been investigated by DFT. In the case of 1, the computational results reveal a rare thermodynamic preference for a terminal hydride ligand as opposed to a hydride-bridged Os–Os bond (3c,2e Os–Os–H bond).

The reactivity of 2-ethynylpyridine at low-valent triosmium and triruthenium centers has been investigated.  相似文献   

4.
A Ru or Rh-catalyzed efficient and atom-economic C7 allylation of indolines with vinylcyclopropanes was developed via sequential C–H and C–C activation. A wide range of substrates were well tolerated to afford the corresponding allylated indolines in high yields and E/Z selectivities under microwave irradiation. The obtained allylated indolines could further undergo transformations to afford various value-added chemicals. Importantly, this reaction proceeded at room temperature under solvent-free conditions.

A Ru or Rh-catalyzed direct C7 allylation of indolines with vinylcyclopropanes via sequential C–H/C–C activation under microwave irradiation has been disclosed.

The development of sustainable methodologies is attractive for access to complex molecular architectures in organic chemistry.1 In recent years, various non-conventional techniques, such as microwave irradiation, sonochemistry, mechanical grinding and photochemistry, have achieved remarkable success.2 In particular, microwaves have shown unique advantages with regards to reaction times, energy efficiency, temperature, and reaction media.3 On the other hand, transition-metal-catalyzed activation of C–H4 and C–C5 bonds has been considered as an ideal method for the formation of C–C and C–X bonds. Nevertheless, transition-metal-catalyzed C–H or C–C bond activation under the above non-conventional techniques remains to be explored. It is thus highly imperative to develop a practical strategy in combination of C–H or C–C activation and microwave irradiation.6Recently, there have significant advances in C–H activation technology by merging C–H functionalization with challenging C–C cleavage strategies.7 Since the pioneering work by Bergman and co-workers8 on the sequential C–H and C–C bond activation, many research groups, including Dong,9 Ackermann,10 Li,11 Cramer,12 and others13 have contributed to C–H/C–C activation. In this content, certain small strained rings are often utilized as an effective synthons to undergo ring-opening reactions driven by strain-release energy.14 Very recently, VCPs (vinylcyclopanes) have been reported as allyl reagents to access various (hetero)aromatic derivatives through sequential C–H and C–C activation (Scheme 1a–d).15Open in a separate windowScheme 1Sequential C–H/C–C activations using VCPs.As a continuation of our interest in chelation-directed reactions and novel methods for C–H functionalization,16 we herein report a Ru or Rh-catalyzed C-7 allylation of indolines under microwave irradiation using VCPs as the allylating agents (Scheme 1e). This transformation possesses great synthetic potential from the viewpoint of green and sustainable chemistry. Notable features of our protocol include (1) C–H/C–C activation with VCPs by microwave irradiation, (2) broad substrate scope with good regio- and E/Z selectivities, (3) high atom economy, and (4) high efficiency (2 h) at room temperature under solvent-free conditions.We initiated our investigation by choosing indoline 1a and VCP 2a as model substrates under microwave irradiation conditions (
EntryCatalyst (mol%)Additive (mol%) T (°C)Yield (%)
1[Ru(p-cymene)Cl2]2AdCOOH9040
2RuCl3·3H2OAdCOOH90N.R
3[Cp*RuCl2]2AdCOOH90N.R
4[Ru(p-cymene)Cl2]2MesCOOH9047
5[Ru(p-cymene)Cl2]2AcOH9040
6[Ru(p-cymene)Cl2]2NaOAc9020
7[Ru(p-cymene)Cl2]2PivONa·H2O9021
8[Ru(p-cymene)Cl2]2DABCO90Trace
9b[Ru(p-cymene)Cl2]2MesCOOH9057
10b[Ru(p-cymene)Cl2]2MesCOOH7068
11b[Ru(p-cymene)Cl2]2MesCOOH5083
12b[Ru(p-cymene)Cl2]2MesCOOH2565
13b,c[Ru(p-cymene)Cl2]2MesCOOH2587 (>20 : 1)e
14c,d[Cp*Rh(CH3CN)3](SbF6)2AdCOOH8078 (10 : 1)e
Open in a separate windowaReaction conditions: 1a (0.2 mmol), 2a (0.4 mmol), [Ru(p-cymene)Cl2]2 (5 mol%), additive (30 mol%), MW, 1 h, 90 °C.bMesCOOH (50 mol%).c t = 2 h.d[Cp*Rh(CH3CN)3](SbF6)2 (8 mol%).eThe E : Z ratio was determined by 1H NMR analysis. MW = microwave irradiation.Under the optimized reaction conditions with the ruthenium catalyst, the substrate scope of indolines 1 was investigated (). While 3b′a was obtained in 84% yield, others failed to give the coupling products. Overall, indolines with a variety of functionalities ranging from C2 to C6 positions could react with VCP 2a to afford the allylated products in good yields with high E/Z selectivities under the [Ru(p-cymene)Cl2]2 catalytic system. On the other hand, when [RhCp*(CH3CN)3](SbF6)2 was employed as the catalyst instead of Ru analogue, decreased yields accompanied with lower E/Z selectivities was observed in most cases.Substrate scope of indolinesa,b
Open in a separate windowaReaction conditions [Ru]: 1 (0.2 mmol), 2a (0.4 mmol), [Ru(p-cymene)Cl2]2 (5 mol%), MesCOOH (50 mol%), AgSbF6 (25 mol%), MW, 25 °C, 2 h; [Rh]: 1 (0.2 mmol), 2a (0.4 mmol), [Cp*Rh(CH3CN)3](SbF6)2 (8 mol%), AdCOOH (30 mol%), AgSbF6 (20 mol%), MW, 80 °C, 2 h.bThe E : Z ratio was determined by 1H NMR analysis. MW = microwave.Encouraged by the above results, the scope of VCPs 2 was also explored to further examine the generality of the current C–C/C–H activation strategy with the ruthenium catalyst ( Open in a separate windowaReaction conditions [Ru]: 1 (0.2 mmol), 2a (0.4 mmol), [Ru(p-cymene)Cl2]2 (5 mol%), MesCOOH (50 mol%), AgSbF6 (25 mol%), MW, 25 °C, 2 h; [Rh]: 1 (0.2 mmol), 2a (0.4 mmol), [Cp*Rh(CH3CN)3](SbF6)2 (8 mol%), AdCOOH (30 mol%), AgSbF6 (20 mol%), MW, 80 °C, 2 h.bThe E : Z ratio was determined by 1H NMR analysis.c40 °C.d[Ru] CH2Cl2 (0.5 mL), 40 °C; [Rh] : MeOH (0.5 mL). MW = microwave.To evaluate the practical utility of the current methodology, a gram-scale reaction between 1a (6.0 mmol) and 2a (12.0 mmol) was performed under standard conditions, which delivered the allylated product 3aa in 74% yield (Scheme 2). Meanwhile, the derivatizations of 3aa were also conducted to highlight the synthetic importance of allylated indolines. Firstly, 3aa could undergo decarboxylation in the presence of NaOEt in DMSO to afford compound 4 in 86% yield. Under the condition of KOH in EtOH, the hydrolysis product 5 was obtained in 90% yield. Secondly, in the presence of DDQ in toluene, product 3aa could be oxidized to compound 6 in 38% yield. Thirdly, hydrogenation of 3aa under Pd–C/H2 conditions would afford a reduced product 7, which could further undergo oxidation and decarboxylation transformations to give indole derivatives 8 and 9 in 90% and 82% yields, respectively.Open in a separate windowScheme 2Gram-scale reaction and further derivatization of product 3aa.To gain insights into the reaction mechanism, a series of control experiments were conducted (Scheme 3). In the H/D exchange experiment, the deuterated [D]-1a was prepared and subjected to the standard conditions. It was found that only negligible amount of deuterium was detected for retrieved 1a. When 2a reacted with [D]-1a for 30 min, a similar result with a distinct D/H exchange was also observed (Scheme 3a). Next, the intermolecular competition experiment between substituted indolines 1o and 1t was conducted. Methoxyl-substituted allylated indoline 3oa was isolated as the major product, indicating that indolines bearing electron-donating groups are more reactive (Scheme 3b). When a radical quencher, such as 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO), 2,6-di-tert-butyl-4-methylphenol (BHT), or benzoquinone (BQ), was added, the allylation reaction was suppressed with product 3aa obtained in a decreased yield (Scheme 3c). These results could''t indicate that a non-radical mediated reaction pathway. Finally, the KIE values observed in parallel reactions suggest that the C–H bond cleavage is not involved in the limiting step (Scheme 3d).Open in a separate windowScheme 3Control experiments.On the basis of above discussion and related reports,17 a plausible catalytic cycle was proposed (Scheme 4). A pyrimidine-directed C–H activation between indoline 1a and an in situ-generated Ru cationic complex I gives a six-membered ruthenacycle II. Coordination of intermediate II with VCP 2a provides intermediate III, which undergoes 1,2-migratory insertion to afford an ruthenium intermediate IV. Then oxygen coordinates to Ru metal center to form intermediate V. After C–C cleavage of VCP 2a, intermediate VI will be generated, followed by protonation to deliver the desired product 3aa and regenerate the active catalyst species I.Open in a separate windowScheme 4Proposed reaction mechanism.  相似文献   

5.
A theoretical study based on DFT calculations on the different influence of functional groups on the C–H activation process via Pd-catalysed β-X elimination     
Xin Xiang  Zeng-Xia Zhao  Hong-Xing Zhang 《RSC advances》2022,12(40):26116
We have performed a series of theoretical calculations for palladium-catalyzed β-X elimination reactions. The DFT calculation combined with energy decomposition analysis shows the determining factors of reactivity. Such as, the elemental composition, the structure of different functional groups and the stronger steric repulsions contribution.

The energy decomposition analysis (EDA) results show the quantitative contribution of nucleophile groups in intramolecular interactions.  相似文献   

6.
Comparative DFT study of metal-free Lewis acid-catalyzed C–H and N–H silylation of (hetero)arenes: mechanistic studies and expansion of catalyst and substrate scope     
Pan Du  Jiyang Zhao 《RSC advances》2019,9(64):37675
Direct selective dehydrogenative silylation of thiophenes, pyridines, indoles and anilines to synthesize silyl-substituted aromatic compounds catalyzed by metal-free Lewis acids was achieved recently. However, there is still insufficient mechanistic data for these transformations. Using density functional theory calculations, we conducted a detailed investigation of the mechanism of the B(C6F5)3-catalyzed dehydrogenative silylation of N-methylindole, N,N-dimethylaniline and N-methylaniline. We successfully located the most favourable reaction pathways that can explain the experimental observations notably well. The most favourable pathway for B(C6F5)3-catalyzed C–H silylation of N-methylindole includes nucleophilic attack, proton abstraction and hydride migration. The C–H silylation of N,N-dimethylaniline follows a similar pathway to N-methylindole rather than that proposed by Hou''s group. Our mechanism successfully explains that the transformations of N-methylindoline to N-methylindole produce different products at different temperatures. For N-methylaniline bearing both N–H and para-phenyl C–H bonds, the N–H silylation reaction is more facile than the C–H silylation reaction. Our proposed mechanism of N–H silylation of N-methylaniline is different from that proposed by the groups of Paradies and Stephan. Lewis acids Al(C6F5)3, Ga(C6F5)3 and B(2,6-Cl2C6H3)(p-HC6F4)2 can also catalyze the C–H silylation of N-methylindole like B(C6F5)3, but the most favourable pathways are those promoted by N-methylindoline. Furthermore, we also found several other types of substrates that would undergo C–H or N–H silylation reactions under moderate conditions. These findings may facilitate the design of new catalysts for the dehydrogenative silylation of inactivated (hetero)arenes.

We investigated the mechanism of the dehydrosilylation of (hetero)arenes and extended the scope of the silylation catalysts and substrates.  相似文献   

7.
One-pot regioselective C–H activation iodination–cyanation of 2,4-diarylquinazolines using malononitrile as a cyano source     
Ziqiao Yan  Banlai Ouyang  Xunchun Mao  Wei Gao  Zhihong Deng  Yiyuan Peng 《RSC advances》2019,9(32):18256
A one-pot cyanation of 2,4-arylquinazoline with NIS and malononitrile has been developed. The one-pot reaction includes two steps. The Rh-catalyzed selective C–H activation/iodization of 2,4-diarylquinazoline with NIS, and then Cu-catalyzed cyanation of the corresponding iodinated intermediate with malononitrile to selectively give 2-(2-cyanoaryl)-4-arylquinazolines or 2-(2,6-dicyanoaryl)-4-arylquinazolines in good to excellent yields.

A one-pot cyanation of 2,4-arylquinazoline with NIS and malononitrile has been developed.  相似文献   

8.
Palladium-catalyzed intramolecular C–H arylation of 2-halo-N-Boc-N-arylbenzamides for the synthesis of N–H phenanthridinones     
Quan-Fang Hu  Tian-Tao Gao  Yao-Jie Shi  Qian Lei  Luo-Ting Yu 《RSC advances》2018,8(25):13879
A palladium catalyzed synthesis of N–H phenanthridinones was developed via C–H arylation. The protocol gives phenanthridinones regioselectively by one-pot reaction without deprotection. It exhibits broad substrate scope and affords targets in up to 95% yields. Importantly, it could be applied for the less reactive o-chlorobenzamides.

Pd(t-Bu3P)2/KOAc proved to be a good combination for one-pot synthesis of N–H phenanthridinones with up to 95% yield.  相似文献   

9.
Ruthenium-catalyzed,site-selective C–H activation: access to C5-substituted azaflavanone     
Manickam Bakthadoss  Tadiparthi Thirupathi Reddy  Duddu S. Sharada 《RSC advances》2020,10(52):31570
A site-selective ruthenium-catalyzed keto group assisted C–H bond activation of 2-aryl tetrahydroquinoline (azaflavanone) derivatives has been achieved with a variety of alkenes for the first time. A wide range of substrates was utilized for the synthesis of a wide variety of alkenylated azaflavanones. This simple and efficient protocol provides the C5-substituted azaflavanone derivatives in high yields with a broad range of functional group tolerance. Further, the C5-alkenylated products were converted into substituted 2-aryl quinoline derivatives in good yields.

A site-selective ruthenium-catalyzed keto group assisted C–H bond activation of 2-aryl tetrahydroquinoline (azaflavanone) derivatives has been achieved with a variety of alkenes for the first time.  相似文献   

10.
O2 activation by core–shell Ru13@Pt42 particles in comparison with Pt55 particles: a DFT study     
Jing Lu  Bo Zhu  Shigeyoshi Sakaki 《RSC advances》2020,10(59):36090
The reaction of O2 with a Ru13@Pt42 core–shell particle consisting of a Ru13 core and a Pt42 shell was theoretically investigated in comparison with Pt55. The O2 binding energy with Pt55 is larger than that with Ru13@Pt42, and O–O bond cleavage occurs more easily with a smaller activation barrier (Ea) on Pt55 than on Ru13@Pt42. Protonation to the Pt42 surface followed by one-electron reduction leads to the formation of an H atom on the surface with considerable exothermicity. The H atom reacts with the adsorbed O2 molecule to afford an OOH species with a larger Ea value on Pt55 than on Ru13@Pt42. An OOH species is also formed by protonation of the adsorbed O2 molecule, followed by one-electron reduction, with a large exothermicity in both Pt55 and Ru13@Pt42. O–OH bond cleavage occurs with a smaller Ea on Pt55 than on Ru13@Pt42. The lower reactivity of Ru13@Pt42 than that of Pt55 on the O–O and O–OH bond cleavages arises from the presence of lower energy in the d-valence band-top and d-band center in Ru13@Pt42 than in Pt55. The smaller Ea for OOH formation on Ru13@Pt42 than on Pt55 arises from weaker Ru13@Pt42–O2 and Ru13@Pt42–H bonds than the Pt55–O2 and Pt55–H bonds, respectively. The low-energy d-valence band-top is responsible for the weak Ru13@Pt42–O and Ru13@Pt42–OH bonds. Thus, the low-energy d-valence band-top and d-band center are important properties of the Ru13@Pt42 particle.

In this theoretical study by DFT computations, characteristic features of the Ru13@Pt42 core–shell particle in O2 activation are clearly discussed in comparison with Pt55.  相似文献   

11.
Computational insights into Ir(iii)-catalyzed allylic C–H amination of terminal alkenes: mechanism,regioselectivity, and catalytic activity     
Deng Pan  Gen Luo  Yang Yu  Jimin Yang  Yi Luo 《RSC advances》2021,11(31):19113
Computational studies on Ir(iii)-catalyzed intermolecular branch-selective allylic C–H amination of terminal olefins with methyl dioxazolone have been carried out to investigate the mechanism, including the origins of regioselectivity and catalytic activity difference. The result suggests that the reaction proceeds through generation of active species, alkene coordination, allylic C–H activation, decarboxylation, migratory insertion, and protodemetalation. The presence of AgNTf2 could thermodynamically promote the formation of catalytically active species [Cp*Ir(OAc)]+. Both the weaker Ir–C(internal) bond and the closer interatomic distance of N⋯C(internal) in the key allyl-Ir(v)-nitrenoid intermediate make the migratory insertion into Ir–C(internal) bond easier than into the Ir–C(terminal) bond, leading to branch-selective allylic C–H amidation. The high energy barrier for allylic C–H activation in the Co system could account for the observed sluggishness, which is mainly ascribed to the weaker coordination capacity of alkenes to the triplet Cp*Co(OAc)+ and the deficient metal⋯H interaction to assist hydrogen transfer.

DFT studies on Ir(iii)-catalyzed branch-selective allylic C–H amination of terminal olefins with methyl dioxazolone have been carried out to investigate the mechanism, including the origins of regioselectivity and catalytic activity difference.  相似文献   

12.
Copper-catalyzed one-pot domino reactions via C–H bond activation: synthesis of 3-aroylquinolines from 2-aminobenzylalcohols and propiophenones under metal–organic framework catalysis     
Ha V. Dang  Hoang T. B. Le  Loan T. B. Tran  Hiep Q. Ha  Ha V. Le  Nam T. S. Phan 《RSC advances》2018,8(55):31455
A Cu2(OBA)2(BPY) metal–organic framework was utilized as a productive heterogeneous catalyst for the synthesis of 3-aroylquinolines via one-pot domino reactions of 2-aminobenzylalcohols with propiophenones. This Cu-MOF was considerably more active towards the one-pot domino reaction than a series of transition metal salts, as well as nano oxide and MOF-based catalysts. The MOF-based catalyst was reusable without a significant decline in catalytic efficiency. To the best of our knowledge, the transformation of 2-aminobenzylalcohols to 3-aroylquinolines was not previously reported in the literature, and this protocol would be complementary to previous strategies for the synthesis of these valuable heterocycles.

Cu2(OBA)2(BPY) metal–organic framework was utilized as a productive heterogeneous catalyst for the synthesis of 3-aroylquinolines via one-pot domino reactions of 2-aminobenzylalcohols with propiophenones.  相似文献   

13.
The concise synthesis and resolution of planar chiral [2.2]paracyclophane oxazolines by C–H activation     
Shashank Tewari  Maulik N. Mungalpara  Suraj Patel  Gareth J. Rowlands 《RSC advances》2022,12(14):8588
Planar chiral [2.2]paracyclophanes are resolved through the direct C–H arylation of enantiopure oxazolines, providing a convenient route to ligands and chiral materials. Preliminary results show that hydrolysis followed by decarboxylative phosphorylation leads to enantiopure [2.2]paracyclophane derivatives that are otherwise challenging to prepare.

Racemic bromo[2.2]paracyclophanes are directly transformed into enantiomerically pure planar chiral oxazolines in one step with simultaneous resolution of planar chirality.

[2.2]Paracyclophane 1 (X = H; Scheme 1) has garnered considerable attention as the core for planar chiral pre-ligands,1,2 catalysts,3,4 bioactive entities5,6 and chiroptical materials.7,8 The rigidity of the [2.2]paracyclophane provides an almost unrivalled opportunity to control the relative spatial position of various functional groups through a series of regioisomers.4,9,10 The major roadblock hampering progress in [2.2]paracyclophane chemistry is the resolution of planar chirality, and there are only a limited number of commonly used enantiopure precursors, some of which are hard to obtain on a large scale.11,12Open in a separate windowScheme 1Previous routes to oxazolines and proposed chemistry.Oxazolines form a host of ‘privileged’ pre-ligands,13–15 and it is no surprise that numerous oxazoline-substituted [2.2]paracyclophanes 5 have shown potential in asymmetric catalysis.16–20 The oxazoline moiety aids the synthesis of these compounds, either by permitting resolution of the planar chirality,16–19,21,22 or by directing functionalisation of the cyclophane backbone.16,19,21,23 Every one of these oxazoline-containing [2.2]paracyclophanes 5 was synthesised from a bromo derivative 2via the carboxylic acid 3, then the amide 4, and finally cyclisation (Scheme 1). This adds steps to the synthesis and limits the functionality found in the molecule as the acid is invariably formed by halogen–metal exchange.Direct addition of an oxazoline 6 would allow simultaneous functionalisation and resolution of the planar chirality of [2.2]paracyclophane. Such a strategy would permit a more concise synthesis of [2.2]paracyclophane derivatives with a wider range of substituents. Previously, we have used C–H activation chemistry to synthesise planar chiral N-oxides,24,25 and believed that the oxazoline C–H activation chemistry of Ackermann26,27 and Lu28 could be applied to [2.2]paracyclophane. Herein, we report the successful realisation of this strategy. Planar chiral oxazolines are accessed in a single step furnishing molecules that can be used as pre-ligands or as precursors to carboxylic acids that can be further derivatised by traditional means or decarboxylative couplings.The coupling of oxazolines 6 and 4-bromo[2.2]paracyclophane 7 required little optimisation with the chemistry of Ackermann26 transferring to paracyclophane without incident. A range of (heteroatom-substituted) secondary phosphine pre-ligands [(HA)SPO] were screened, and di-tert-butylphosphine oxide (t-Bu2SPO) gave a catalyst that showed complete conversion of simple [2.2]paracyclophane derivatives, such as 7 (8a 82%; 8c 60% Scheme 2). More complex derivatives required the di-1-adamantylphosphine oxide (Ad2SPO) otherwise returned unreacted starting material (see below).Open in a separate windowScheme 2Synthesis of 4-oxazolinyl[2.2]paracyclophanes 8a and 8c.Benzyloxazoline 8a was formed as a mixture of diastereoisomers that are separable by column chromatography, but the similarity of the Rf values make resolution extremely challenging, with only the fractions at the beginning and end of a collection containing pure diastereoisomers. This led to screening a range of oxazolines. All coupled but with less satisfactory yields. The phenyloxazoline 6b appears to couple with concomitant aromatisation furnishing oxazole 9. For mono(oxazolines) 8, tert-butyl derivative 6c gave the best results in terms of resolution of planar chirality, with the diastereoisomers 8c being readily separable.21Next we established the scope of coupling reaction. Both the pseudo-para- (4,16)- (11a) and pseudo-ortho- (4,12)- bis(oxazolines) 11b were readily prepared from the corresponding dibromo[2.2]paracyclophanes (17,20,29 but only one diastereoisomer of the tert-butyl derivative 11c could be isolated. The other diastereoisomer co-ran with the two diastereomers of the product of protodebromination (8c). The synthesis of pseudo-ortho-bis(benzyloxazoline) 11b has been performed on >1 g scale. The yields are lower but the separation easier. If a single equivalent of 2H-oxazoline 6a is used in the coupling, it is possible to isolate the bromo mono(oxazoline) 11d, but the reaction is non-selective, with a mixture of mono-along and bis(oxazoline) always formed.16,18,30Scope of oxazoline coupling (only single diastereoisomer of product shown)
Open in a separate windowaConditions A: 6 (1.1 eq), Pd(OAc)2 (5 mol%), (t-Bu)2P(O)H (10 mol%), LiOt-Bu (2.5 eq); conditions B: 6 (2.2 eq), Pd(OAc)2 (10 mol%), (t-Bu)2P(O)H (20 mol%), LiOt-Bu (5.0 eq); conditions C: 6 (4.4 eq), Pd(OAc)2 (40 mol%), (t-Bu)2P(O)H (80 mol%), LiOt-Bu (10.0 eq); conditions D: 6 (1.1 eq), Pd(OAc)2 (5 mol%), (Ad)2P(O)H (10 mol%), LiOt-Bu (2.5 eq); conditions E: 6 (4.4 eq), Pd(OAc)2 (40 mol%), (Ad)2P(O)H (80 mol%), LiOt-Bu (10.0 eq); conditions F: 6 (2.2 eq), Pd(OAc)2 (10 mol%), (Ad)2P(O)H (20 mol%), LiOt-Bu (5.0 eq).bOnly single diastereoisomer shown in table. Diastereoisomers in red cannot be separated. Diastereoisomers in blue are partially resolved.cYield of separate diastereomers (diastereomer1)%/(diastereomer2)%.dCombined yield of all stereoisomers (includes mixed fractions).As is frequently observed with [2.2]paracyclophane derivatives, each regioisomer behaves differently, and it is hard to predict the influence of the substituents.31,32 Using t-Bu2SPO as pre-ligand, both primary and tertiary pseudo-gem amines gave the desired oxazolines as separable diastereoisomers 11e and 11f. The normally unreactive ortho substituted [2.2]paracyclophane derivative, 4-amino-5-bromo[2.2]paracyclophane,33 appeared to give the amino-oxazoline 11g as a pair of separable diastereomers in 34% yield, although these compounds proved to be unstable. All other amino[2.2]paracyclophanes gave unsatisfactory yields. The pseudo-ortho-amino-bromo[2.2]paracyclophane failed to couple and only furnished the product of protodebromination. On changing to Ad2SPO, the meta11h and pseudo-para11i, were synthesised as separable diastereomeric amino oxazolines in moderate to good yields while the pseudo-meta-amino-bromo [2.2]paracyclophane gave 11j as an inseparable mixture of diastereomers. It is unclear why this ligand is superior with these amino-bromo[2.2]paracyclophanes. Nitro-substituted [2.2]paracyclophanes (pseudo-gem and pseudo-meta) are unreactive. Esters proved problematic, giving low yields (11k and 11m) and/or the product of protodebromination (12). We assume that the strongly basic conditions are incompatible with this functionality. This is supported by the fact the free acid reacts in good to moderate yields, 44% for pseudo-gem (11l) and 61% for pseudo-ortho (11n) (34 Initial reaction with the t-Bu2SPO ligand gave a complex mixture of regio- and stereoisomers. The resolved tetra(oxazoline) ligand 11o was isolated in 10%. In addition, the para-bis(oxazoline)-dibromo[2.2]paracyclophane 11p was isolated as separable diastereoisomers along with an unidentified regioisomer of the diastereomeric bis(oxazoline)-dibromo[2.2]paracyclophanes (not included in the yield). Finally, the two diastereoisomers of the protodebromo tris(oxazoline) 11q were isolated. Altering the ligand to Ad2SPO simplified the mixture and gave just the tetra(oxazoline) 11o and tris(oxazoline) 11q.There are numerous reports describing the utility of [2.2]paracyclophane oxazolines. They have been used as pre-ligands,16,18–20,29 or have directed further functionalisation of the [2.2]paracyclophane by either bromination16,23 or metalation.19,21 We believe straightforward elaboration of the amine or acid functionality will deliver modular pre-ligands, but here we want to increase the versatility of the oxazoline moiety and present preliminary results on the hydrolysis and decarboxylative phosphorylation of the oxazolines.Oxazolines are robust, and hydrolysis is not straightforward. The hydrolysis of 11e by heating to reflux in 6 M HCl has been reported to give 13b in good yield (18 and these conditions worked for mono(benzyloxazoline) 8a but failed to hydrolyse the mono(tert-butyloxazoline) 8c or the bis(oxazoline) 11b. More forcing conditions using concentrated sulfuric acid in dioxane at 110 °C lead to the clean hydrolysis of mono(oxazolines) while the pseudo-ortho bis(oxazoline) gives a mixture of the 4-acid-12-oxazoline 11n and the desired 4,12-diacid 13c. Fortunately, the products are easily separated based on solubility and recycling 11n permits good overall yields of 13c to be achieved. No sign of racemisation can be detected by optical rotation. While the conditions are harsh, the rapid synthesis of the oxazolines with concomitant resolution still make this methodology attractive if conducted early in a synthetic sequence.Hydrolysis of oxazolinesa
Open in a separate windowaConditions A: 6 M HCl(aq), reflux; conditions B: conc. H2SO4, dioxane, 110 °C.bStarting from a mixture of diastereomers.cUsing pure diastereomer of 8.dHydrolysis reported in ref. 18.eReacted for 12 hours.fReacted for 24 hours.It is possible to subject the enantiomerically enriched 4-carboxylic acid 13a to a decarboxylative phosphorylation to give the phosphine oxide 14 in 8–20% (along with 50% unreacted starting material with no erosion of enantiopurity (Scheme 3).35–37 So far, we have been unable to react the pseudo-ortho-diacid under the same conditions. Although the preliminary results are low yielding, they demonstrate that it is possible to isolate enantiomerically enriched phosphine oxide in just four steps. Only our own sulfoxide chemistry38,39 is as efficient and that methodology was limited by the need for a sulfoxide–lithium exchange with tert-butyllithium.Open in a separate windowScheme 3Decarboxylative phosphorylation.[2.2]Paracyclophane offers a useful planar chiral skeleton but its widespread adoption has been frustrated by a lack of simple methods to resolve the planar chirality. In this paper, we have shown how oxazolines can be coupled to bromo-derivatives to give separable diastereomers. The methodology permits a range of disubstituted [2.2]paracyclophanes to be rapidly prepared in enantiomerically pure form. We report the first example of a decarboxylative coupling on [2.2]paracyclophane. Optimisation of the decarboxylative coupling offers an exciting new route to enantiomerically enriched planar chiral molecules that are challenging to synthesise by conventional methods. We intend to extend the utility of this reaction to other derivatives, and will publish these studies in due course. Simpler, more rapid access to enantiomerically pure [2.2]paracyclophane derivatives will facilitate their wider use in catalysts, bioactive compounds, and materials.  相似文献   

14.
Copper-catalyzed C–H acyloxylation of 2-phenylpyridine using oxygen as the oxidant     
Feifan Wang  Zhiyang Lin  Weisheng Yu  Qingdong Hu  Chao Shu  Wu Zhang 《RSC advances》2018,8(29):16378
An efficient copper-catalyzed direct o-acyloxylation of 2-phenylpyridine with carboxylic acids using oxygen as the oxidant has been developed. Various acids including aromatic acids, cinnamic acids and aliphatic acids are effective acyloxylation reagents and provide the desired products in moderate to excellent yields. The reaction proceeds well under an oxygen atmosphere, making this method potentially practical.

A copper catalyzed direct o-acyloxylation of 2-phenylpyridine with various carboxylic acids using oxygen as oxidant has been developed.  相似文献   

15.
TBAI-assisted direct C–H activation of indoles with β-E-styrene sulfonyl hydrazides: a stereoselective access to 3-styryl thioindoles     
Saira Hafeez  Aamer Saeed 《RSC advances》2021,11(26):15608
The current work describes the challenging introduction of a vinyl sulfide group by simple C–H activation on a variety of substrates. The direct C–H activation of indoles with β-(E)-styrene sulfonyl hydrazides under the sulfenylation conditions, assisted by the iodic catalyst tert-butyl ammonium iodide (TBAI), afforded a series of (E)-styrylthioindoles. Accordingly, β-(E)-styrene sulfonyl hydrazides undergo radical cross-coupling reactions with a variety of substituted indoles to afford structurally diverse indole vinyl thioethers in moderate to high yields with E-stereoselectivity. This method is metal-catalyst-free and is valuable not only because of its novelty, but also for providing a convenient synthetic pathway to a variety of (E)-styrylthioindoles with retention of the configuration. The current study paves the way for the use of β-(E)-styrene sulfonyl hydrazides as a unique styryl mercaptan source in chemical synthesis.

Direct C–H activation of indoles with beta-(E)-styrene sulfonyl hydrazides assisted by TBAI afforded structurally diverse indole vinyl thioethers in moderate to high yields with E-stereoselectivity.  相似文献   

16.
A DFT study on the C–H oxidation reactivity of Fe(iv)–oxo species with N4/N5 ligands derived from l-proline     
Jin Lin  Qiangsheng Sun  Wei Sun 《RSC advances》2021,11(4):2293
The hydroxylation of hexane by two FeIVO complexes bearing a pentadentate ligand (N5, Pro3Py) and a tetradentate ligand (N4, Pro2PyBn) derived from l-proline was studied by DFT calculations. Theoretical results predict that both FeIVO complexes hold triplet ground states. The hydrogen atom abstraction (HAA) processes by both FeIVO species proceed through a two-state reactivity, thus indicating that HAA occurs via a low-barrier quintet surface. Beyond the conventional rebound step, the dissociation path is also calculated and is found to potentially occur after HAA.

The hydroxylation of hexane by two FeIVO complexes bearing a pentadentate ligand (N5, Pro3Py) and a tetradentate ligand (N4, Pro2PyBn) derived from l-proline was studied by DFT calculations.  相似文献   

17.
Auxiliary-directed etherification of sp2 C–H bonds under heterogeneous metal–organic framework catalysis: synthesis of ethenzamide     
Chau B. Tran  Xuan N. T. Duong  Huy D. Lu  Thu T. V. Cao  Thanh Truong 《RSC advances》2018,8(5):2829
An efficient protocol for 8-aminoquinoline assisted alkoxylation and phenoxylation of sp2 C–H bonds under heterogeneous catalysis was developed. The optimal conditions employed Cu-MOF-74 (20%), K2CO3 base, pyridine ligand or dimethyl formamide solvent, and O2 oxidant at 80 °C or 100 °C for 24 hours. Cu-MOF-74 revealed remarkably higher activity when compared with other previously commonly used Cu-MOFs in cross coupling reactions, supported copper catalysts, and homogeneous copper salts. The reaction scope with respect to coupling partners included a wide range of various substrates. Interestingly, the developed conditions are applicable for the synthesis of high-profile relevant biological agents from easily accessible starting materials. Furthermore, a leaching test confirmed the reaction heterogeneity and the catalyst was reused and recycled at least 8 times with trivial degradation in activity.

An efficient protocol for 8-aminoquinoline assisted alkoxylation and phenoxylation of sp2 C–H bonds under heterogeneous catalysis was developed.  相似文献   

18.
New insights into the surface plasmon resonance (SPR) driven photocatalytic H2 production of Au–TiO2     
Jinlin Nie  Jenny Schneider  Fabian Sieland  Long Zhou  Shuwei Xia  Detlef W. Bahnemann 《RSC advances》2018,8(46):25881
The Surface Plasmon Resonance (SPR) driven photocatalytic H2 production upon visible light illumination (≥500 nm) was investigated on gold-loaded TiO2 (Au–TiO2). It has been clearly shown that the Au-SPR can directly lead to photocatalytic H2 evolution under illumination (≥500 nm). However, there are still some open issues about the underlying mechanism for the SPR-driven photocatalytic H2 production, especially the explanation of the resonance energy transfer (RET) theory and the direct electron transfer (DET) theory. In this contribution, by means of the EPR and laser flash photolysis spectroscopy, we clearly showed the signals for different species formed by trapped electrons and holes in TiO2 upon visible light illumination (≥500 nm). However, the energy of the Au-SPR is insufficient to overcome the bandgap of TiO2. The signals of the trapped electrons and holes originate from two distinct processes, rather than the simple electron–hole pair excitation. Results obtained by Laser Flash Photolysis spectroscopy evidenced that, due to the Au-SPR effect, Au NPs can inject electrons to the conduction band of TiO2 and the Au-SPR can also initiate e/h+ pair generation (interfacial charge transfer process) upon visible light illumination (≥500 nm). Moreover, the Density Functional Theory (DFT) calculation provided direct evidence that, due to the Au-SPR, new impurity energy levels occurred, thus further theoretically elaborating the proposed mechanisms.

The Surface Plasmon Resonance (SPR) driven photocatalytic H2 production upon visible light illumination (≥500 nm) was investigated on gold-loaded TiO2 (Au–TiO2).  相似文献   

19.
DMSO-mediated palladium-catalyzed cyclization of two isothiocyanates via C–H sulfurization: a new route to 2-aminobenzothiazoles     
Guangkai Yao  Bing-Feng Wang  Shuai Yang  Zhi-Xiang Zhang  Han-Hong Xu  Ri-Yuan Tang 《RSC advances》2019,9(6):3403
DMSO was found to activate arylisothiocyanates for self-nucleophilic addition. A subsequent intramolecular C–H sulfurization catalyzed by PdBr2 enables access to a wide range of 2-aminobenzothiazole derivatives in moderate to good yields. This is the first example of a DMSO-mediated Pd-catalyzed synthesis of 2-aminobenzothiazoles through cyclization/C–H sulfurization of two isothiocyanates.

DMSO-initiated coupling of two isothiocyanates; Pd-catalyzed C–H sulfurization.

2-Aminobenzothiazole derivatives are privileged scaffolds which have been widely used in medicine and agricultural chemicals (Scheme 1a).1 Significant effort has been devoted to the development of synthetic methodologies for the preparation of such species.2 Thioureas,3,4 isothiocyanates,5,6 and benzothiazoles7 are effective reaction partners for the synthesis of 2-aminobenzothiazole derivatives via cross-coupling reactions or C–H functionalization. Intramolecular C–H sulfurization of N-arylthioureas is an effective strategy for the synthesis of 2-aminobenzothiazoles. Such a transformation can be achieved through metal-catalyzed C–H activation3 or via radical reactions.6 For example, Doi and coworkers reported a Pd-catalyzed intramolecular C–H sulfurization of arylthiourea using O2 as an oxidant.3c Metal-free iodine-catalyzed intramolecular C–H sulfurization also provides an effective pathway to 2-aminobenzothiazoles.6e Zhu et al. developed an efficient oxidative radical strategy for the synthesis of 2-aminobenzothiazoles via an aminyl radical addition to aryl isothiocyanates followed by C–H sulfurization in the presence of n-Bu4NI and TBHP.6c Interestingly, Lei et al. reported a “green” synthesis of 2-aminobenzothiazoles by employing electro-catalysis technology for an external oxidant-free intramolecular C–H sulfurization.6d In terms of metal-catalyzed C–H sulfurization,8 palladium-catalyzed C–H sulfurization has not been widely reported because thioureas are often strong metal-chelating species that may lead to poisoning of palladium catalysts. Thus, only a few examples of Pd-catalyzed protocols have been reported (Scheme 1b).3b,c,e Therefore, the development of novel palladium-catalyzed C–H sulfurization reaction is of great interest. As a continuation of our interest in sulfur chemistry and C–H sulfurization,9 we attempted to transform isothiocyanates into sulfur-containing compounds.10 Unexpectedly, a newly formed 2-aminobenzothiazole was observed for the reaction of a phenylisothiocyanate. Encouraged by this discovery, we developed a new, facile and efficient DMSO-mediated palladium-catalyzed C–H sulfurization/cyclization of two isothiocyanates for the synthesis of 2-aminobenzothiazoles (Scheme 1c).Open in a separate windowScheme 1Biological application and synthetic methods for the preparation of 2-aminobenzothiazoles.Our study began with the reaction of phenyl isothiocyanate (1a) with PdCl2 in different solvents ( EntryCatalystSolventYieldb (%)1PdCl2Toluene02PdCl2Chlorobenzene03PdCl2DCE04PdCl2CH3CN05PdCl2DMF126PdCl2NMP167PdCl2DMSO728cPdCl2DMSO329dPdCl2DMSO5810PdBr2DMSO8211Pd(OAc)2DMSOTrace12Pd(PPh3)2Cl2DMSOTrace13ePdBr2DMSO7014—DMSO015fPdBr2DMSO/CH3CN5516gPdBr2DMSO/CH3CN2317hPdBr2DMSO/H2O6118iPdBr2DMSO/H2O75Open in a separate windowaReaction conditions: 1a (0.4 mmol), [Pd] (10 mol%), stirred at 80 °C in solvent (2 mL) for 24 h.bIsolated yield.cAt 100 °C.dAt 60 °C.ePdBr2 (5 mol%).fDMSO/CH3CN = 1 : 1.gDMSO/CH3CN = 1 : 4.hDMSO/H2O = 10 : 1.iDMSO/H2O = 20 : 1.With the optimized reaction conditions in hand, the reaction scope was investigated ( Open in a separate windowaReaction conditions: 1 (0.4 mmol), PdBr2 (10 mol%), stirred at 80 °C in DMSO (2 mL) for 24 h. Ortho-substituents, such as methyl, methoxy, methylthio, and phenyl, which increase steric hindrance, still performed well to give their corresponding products in moderate to good yields (products 12–20). For reactions involving ortho-substituted substrates, a small amount of brominated by-product similar to compound 19 was observed; however, the majority of these by-products were obtained in far lower yield than 19. To our delight, a naphthyl isothiocyanate also underwent reaction smoothly to give product 22 in 77% yield. 3-Isothiocyanatopyridine was also subjected to the reaction conditions. However, no cyclization/C–H sulfurization occurred; the electron-deficient pyridine ring is not reactive enough for C–H sulfurization. The reaction of meta-substituted methyl phenylisothiocyanates provided two isomers which could not be easily purified.To prove the feasibility of the cyclization of two different aryl isothiocyanates, a cross-reaction between 4-methyl-phenylisothiocyanate and 4-nitro-phenyliso-thiocyanate was conducted. The cross-reaction product 23, was obtained in 27% yield with compound 3 being formed as the main product in 52% yield (Scheme 2). The strong electron-withdrawing nitro group greatly reduces the electron density of the benzene ring, thus C–H sulfurization is disfavored; C–H sulfurization only occurred on the methyl-substituted benzene ring.Open in a separate windowScheme 2PdBr2-catalyzed cross-reaction of two different isothiocyanates.To gain insight into the reaction mechanism, control experiments were conducted, as shown in Scheme 3. In the absence of PdBr2, thiourea 24 was obtained in 35% yield for the reaction of 4-methyl phenylisothiocyanate in DMSO at 80 °C over 24 hours (Scheme 3, eqn (1)). This suggested that the thiourea may be the reaction intermediate; however, when thiourea 24 was subjected to reaction with PdBr2, product 3 was obtained in only 16% yield (Scheme 3, eqn (2)). Aryl isothiocyanate may decompose to aniline, and react with isothiocyanate for C–H sulfurization. However, only a trace amount of product 25 was observed in the reaction between 4-methylphenyl isothiocyanate and aniline (Scheme 3, eqn (3)). None of the target product 26 was observed for the reaction of benzothiazole with 4-methyl phenylisothiocyanate, suggesting that benzothiazole is not a reaction intermediate (Scheme 3, eqn (4)). To our delight, a mass peak corresponding to intermediate 27 was observed with real-time ESI-MS when phenylisothiocyanate was stirred in DMSO at 60 °C (Scheme 3, eqn (5)) (see the ESI, Fig. S1). Based on theoretical calculations,11 the dipole moment of the DMSO molecule is 3.9274 debye (gas) or 5.1281 debye in solvent form. Natural bond orbital analysis shows the differences in the charge distribution on the O and S atoms. Obviously, the polarity of the S–O bond increases and a greater distribution of the negative charge on the O atom leads to an improvement in its nucleophilic character. The oxygen atom attacks the carbon atom with the most positive NBO charge in PhNCS (see the ESI, Fig. S2 and S3). These results indicate that DMSO acts as a nucleophilic agent for the activation of arylisothiocyanates for C–H sulfurization.Open in a separate windowScheme 3Control experiments.Based on the above results and previous work,3 a possible reaction mechanism has been proposed and is shown in Scheme 4. Firstly, the polarized DMSO undergoes a nucleophilic addition with phenylisothiocyanate to afford the active intermediate A, which then reacts with another equivalent of phenylisothiocyanate to produce an intermediate B. Intermediate B chelates with PdBr2 to form an intermediate C. Subsequent elimination of HBr leads to the formation of intermediate D, which undergoes reductive elimination to afford the desired product 2. The Pd(0) is oxidized by DMSO to PdBr2 for the next catalytic cycle.Open in a separate windowScheme 4A possible mechanism.  相似文献   

20.
Dual C–H activation: Rh(iii)-catalyzed cascade π-extended annulation of 2-arylindole with benzoquinone     
Qijing Zhang  Qianrong Li  Chengming Wang 《RSC advances》2021,11(21):13030
A rhodium-catalyzed, N–H free indole directed cyclization reaction of benzoquinone via a dual C–H activation strategy is disclosed. This protocol has a good functional group tolerance and affords useful indole-fused heterocylces. Besides, it is insensitive to moisture, commercially available solvent can be directly used and work quite well for this transformation.

A Rh-catalyzed cascade annulation of N–H free 2-arylindole with benzoquinone via dual C–H activation strategy was reported.

Quinones are widely distributed in nature, and commonly occur in bacteria, flowering plants and arthropods (Fig. 1). They have a wide range of applications, including diverse important pharmacological properties, involvement in redox reactions and development for advanced electrochemical energy storage.1 Among varied reported quinones, benzoquinone (BQ) is the simplest and most important one. It has been well reported that BQ has a significant and unique role in oxidative palladium(ii)-catalyzed coupling reactions.2 The chemistry of benzoquinone has been extensively explored in detail, including nucleophilic addition and cycloaddition reactions, photochemistry and oxidative coupling.1b,c,2 Although great achievements have been obtained, only a few examples are disclosed about BQ as a reactant applying to transition-metal catalyzed C–H functionalization.1e Among the examples reported, cyclization or BQ direct functionalization products were mainly afforded (Scheme 1a).Open in a separate windowFig. 1Selected examples of bioactive molecules containing the benzoquinone moiety.Open in a separate windowScheme 1Transition-metal catalyzed C–H functionalization of BQ.Transition-metal catalyzed C–H functionalization has undergone great progresses in the past two decades.3 In order to get a better reactivity and controlled selectivity, a directing group is usually needed for this process. Therefore, various directing groups have been developed.4 However, many of them (e.g. various nitrogen-containing heterocycles) remained parts of products after reaction, therefore increasing the procedures and difficulty for structure further modification and manipulation.5 As a result, it is highly demanded to explore traceless or easily removable directing groups.6 In this context, N–H free indole moiety has gradually emerged as a versatile functionalizable directing group in transition-metal catalyzed cyclization reaction.7On the other hand, although the above-mentioned great breakthrough obtained in C–H functionalization, there are few examples reported for dual C–H activation reactions.8 During our research program exploring transition-metal catalysis and heterocyclic synthesis,9 we intended to prepare the indole-containing heterocycles based on the consideration of their potential biological activity. Herein, we report a rhodium-catalyzed N–H free indole directed annulation reaction with BQ through dual C–H activation strategy (Scheme 1b).Our initial study was carried out by examining 2-phenyl indole 1a and benzoquinone 2a in the presence of [{Cp*RhCl2}2] and Cu(OAc)2·H2O in commercial available N,N-dimethylformamide under argon atmosphere. To our delight, the desired 9H-dibenzo[a,c]carbazol-3-ol product 3a was isolated in 55% yield ( EntrySolventCatalystAdditiveYield1DMF[Cp*RhCl2]2Cu(OAc)2·H2O55%2DMF[Cp*RhCl2]2—<5%3DMF—Cu(OAc)2·H2O—4 t-Amyl-OH[Cp*RhCl2]2Cu(OAc)2·H2O—5DMAc[Cp*RhCl2]2Cu(OAc)2·H2O<5%6DMSO[Cp*RhCl2]2Cu(OAc)2·H2OTrace7DMF[RuCl2(p-cymene)]2Cu(OAc)2·H2O—8DMFPd(OAc)2Cu(OAc)2·H2O—9DMFRhCl(PPh3)3Cu(OAc)2·H2O<5%10DMF[Cp*RhCl2]2AgOAc—11DMF[Cp*RhCl2]2Ag2O—12DMF[Cp*RhCl2]2Cu(acac)2Trace 13 bb , cc DMF/DCE[Cp*RhCl2]2Cu(OAc)2·H2O 84% Open in a separate windowaReaction on a 0.2 mmol scale, using 1a (1.0 equiv.), 2a (1.0 equiv.), additive (2.0 equiv.), CsOAc (2.0 equiv.), [TM] (5 mol%), solvent (1.0 mL), under N2, isolated yield.b1a (1.5 equiv.), solvent (0.3 M).cNaOAc was used instead of CsOAc.With the optimized conditions in hand, we next tend to examine the substrates scope of this reaction. Various 2-aryl indoles with electron-rich substituted groups were tested and worked well for this reaction (10 Finally, an interesting S, N-fused heterocycle 3m was obtained when 2-thienyl indole was employed. Other derivatives of benzoquinone such as 1,4-naphthaquinone or methyl-p-benzoquinone currently failed to produce the related cyclization products with proper yields.Substrates scopea
Open in a separate windowaCondition A: 2-aryl indole (1.5 equiv.), BQ (1.0 equiv.), [Rh] (5 mol%), Cu(OAc)2·H2O (2.1 equiv.), NaOAc (2.0 equiv.), DMF/DCE(1.5 mL, 2 : 1), 100 °C.bCondiiton B: 2-aryl indole (1.0 equiv.), BQ (2.0 equiv.), [Rh] (5 mol%), Cu(OAc)2·H2O (2.1 equiv.), NaOAc (2.0 equiv.), DMF/DCE(1.5 mL, 2 : 1), 60 °C.In addition, this method allows quick access to a number of functional heterocycles (Scheme 2).7g,11 For example, the hydroxyl group can be easily removed to afford 9H-dibenzo[a,c]carbazole 4a which can be further converted into organic electroluminescent element 5avia reported methods.11Open in a separate windowScheme 2Diversity of the product.Finally, we proposed a mechanism for this transformation (Scheme 3) based on reported literatures.7,9ac,12 First, [{Cp*RhCl2}2] dissociates and delivers the active catalyst monomer [Cp*Rh(OAc)2] with the assistance of copper acetate and sodium acetate.9ac C–H activation of 2-phenyl indole by Rh(iii) produces rhodacyclic intermediate A,7 followed by insertion of benzoquinone affording intermediate B, which can be transformed into Cvia two folds protonation and fulfills the catalytic cycle. The final product 3a can be easily accessed via intramolecular condensation of C.7gOpen in a separate windowScheme 3Proposed mechanism.In conclusion, we have developed a Rh(iii)-catalyzed traceless directed dual C–H activation of 2-aryl indole and annulation with benzoquinone affording indole-fused heterocycles. The protocol is applicable to a wide range of indole derivatives, affording related products in middle to good yields. Further exploration of the synthetic utilities of this chemistry and detailed mechanistic study are currently in progress in our lab and will be reported in due course.  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