首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A simple and efficient metal-free arylation of imidazo[1,2-a]pyridines at the C-3 position with arylhydrazine has been achieved at room temperature under ambient air conditions. Various 2,3-disubstituted imidazopyridines and imidazothiazoles were synthesized with high yields. The present methodology demonstrates the usefulness of commercially available aryl hydrazine as an arylating agent.

Metal-free C–H arylation of imidazo[1,2-a]pyridines at C-3 position with arylhydrazines in presence of DBU has been developed at room temperature under ambient air.  相似文献   

2.
In this study, Cp*Co(iii)-catalyzed site-selective amidation of pyridones and isoquinolones using oxazolones as the amidation reagent is reported. This approach features mild conditions, high efficiency and good functional tolerance. Furthermore, gram-scale preparation and preliminary mechanism experiments were carried out. It provides a straightforward approach for the direct modification of pyridone derivatives.

In this study, Cp*Co(iii)-catalyzed site-selective amidation of pyridones and isoquinolones using oxazolones as the amidation reagent is reported.  相似文献   

3.
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   

4.
A visible-light-mediated method for the construction of N-monoalkylated products from easily available benzamides and benzyl alcohol in the presence of eosin Y has been developed. The reaction proceeded smoothly, for a wide range of derivatives of benzamides and benzyl alcohols, to give the desired products in good to excellent yields. Biological studies, such as those on drug-likeness and molecular docking, are carried out on the molecules.

A visible-light-mediated method for the construction of N-monoalkylated products from easily available benzamides and benzyl alcohol in the presence of eosin Y has been developed.  相似文献   

5.
An efficient Fe-catalyzed esterification of primary, secondary, and tertiary amides with various alcohols for the preparation of esters was performed. The esterification process was accomplished with FeCl3·6H2O, which is a stable, inexpensive, environmentally friendly catalyst with high functional group tolerance.

An efficient Fe-catalyzed esterification of primary, secondary, and tertiary amides with various alcohols was performed. Esterification was accomplished with inexpensive, environmentally friendly FeCl3·6H2O, and with high functional group tolerance

The amide bond is not only employed as an important natural peptide skeleton in biological systems, but also as a versatile functional group in organic transformations.1,2 Among amide transformations,3,4 transamidation and esterification of amides have attracted widespread attention from organic chemists.5–11 In contrast to transamidation of amides, esterification of amides is relatively difficult because of the low nucleophilicity of alcohols compared with amines.12 To overcome this shortcoming, various processes for the esterification of amides have been developed, such as using an activating agent,6 employing twisted amides,7 and forming intramolecular assisted groups.8 For increasing the synthetic flexibility of esterification of amides, new catalytic systems Zn(OTf)2 and Sc(OTf)3 were developed by Mashima and Williams for the esterification of amides.9 Shimizu et al. reported CeO2 catalyzed esterification of amides.10 Off late, progress of the significant Ni-catalyzed esterification of amides has gained precedence.11 Although this approach produces satisfactory yield, the drawbacks such as the use of an expensive catalyst, generation of reagent waste, high temperature and the limited scope of the substrate still persists. Herein, under mild reaction conditions, we report a novel and efficient Fe-catalyzed esterification of amides for the synthesis of esters (eqn (1)). Compared to conventional methods, this esterification procedure is distinguished by using a stable, inexpensive, environment-friendly catalyst, i.e., FeCl3·6H2O with a low toxicity solvent. The catalytic system has wide functional group compatibility. The reactions of several primary, secondary, and tertiary amides with various alcohols have been well tolerated in this process. Moreover, esters were also as effective as esterification reagents for the esterification of amides by the acyl–acyl exchange process. In the course of our present study, a Co-catalyzed amide C–N cleavage to form esters was reported by Danoun et al.13 However, an additional additive (Bipy) and Mn (3 equiv.) were needed and the scope of the reaction was limited to only tertiary amides; primary or secondary amides were not compatible with this catalytic oxidation system.1We initially selected benzamide 1a and ethanol 2a as model substrates to screen the reaction conditions, and the results are summarized in
EntryCatalyst (20%)Additive (40%)Yieldb
1FeCl231
2FeCl331
3FeBr325
4FeSO4·7H2O20
5Fe(NO3)3·9H2O20
6FeCl3·6H2O35
7CuCl2Trace
8PdCl2Trace
9Ni2Cl·6H2OTrace
10
12FeCl3·6H2OPyridineTrace
13FeCl3·6H2O1,10-PhenanthrolineTrace
14FeCl3·6H2OFormic acidTrace
15FeCl3·6H2O l-ProlineTrace
16FeCl3·6H2OH2SO422
17FeCl3·6H2OHNO325
18FeCl3·6H2OHCl91
19HCl35
20cFeCl3·6H2OHClTrace
21dFeCl3·6H2OHCl36
22eFeCl3·6H2OHCl28
23fFeCl3·6H2OHCl40
Open in a separate windowaReaction conditions: 1a (0.2 mmol), 2a (0.1 mL), catalyst (0.04 mmol, 20 mol%), H2SO4 (concentrated, 0.24 mmol), HNO3 (concentrated, 0.24 mmol), HCl (0.24 mmol, 36–38%), n-hexane (1.0 mL), 80 °C, 14 h.bGC yield using hexadecane as internal standard.cDMF was used as a solvent.d1,4-Dioxane was used as solvent.eToluene was used as solvent.fCH3CN was used as solvent.The substrate scope of esterification with alcohols was investigated under the optimized reaction conditions. As shown in EntryAlcoholProductsYieldb1Ethanol 2a 3a, 85%2 n-Propanol 2b 3b, 88%3 n-Octanol 2c 3c, 82%4 i-Propanol 2d 3d, 91%5 t-Butyl alcohol 2e 3e, 81%6Cyclohexanol 2f 3f, 80%7Phenethanol 2g 3g, 83%8Phenethanol 2h 3h, 81%9Trifluoroethanol 2i 3i, 81%10Ethane-1,2-diol 2j 3j, 85%Open in a separate windowaReaction conditions: 1a (0.2 mmol), 2a (0.1 mL), 2b–2j (0.24 mmol), FeCl3·6H2O (0.04 mmol, 20 mol%), HCl (0.24 mmol, 36–38%), n-hexane (1.0 mL), 80 °C, 14 h.bIsolated yield.We then examined the generality of amides and the results are compiled in EntryAmidesProducts and yieldb1 3k, 86%2 3l, 85%3 3m, 78%4 3n, 75%5 3o, 86%6 3p, 84%7 3q, 90%8 3r, 86%9 3s, 84%10 3t, 55%11 3u, 88%12 3v, 85%13 3w, 81%Open in a separate windowaReaction conditions: 1b–1n (0.2 mmol), 2a (0.1 mL), FeCl3·6H2O (0.04 mmol, 20 mol%), HCl (0.24 mmol, 36–38%), n-hexane (1.0 mL), 80 °C, 14 h.bIsolated yield.In addition to alcohols, esters were also used as substrates for esterification of amides through the acyl–acyl exchange process under the optimal reaction conditions.14 As shown in EntryAmidesEstersProducts and yieldb11a 21a 31a 41a 51b 61c 71h Open in a separate windowaReaction conditions: 1a–1h (0.2 mmol), 2k, 2l (0.1 mL), 2m–2o (0.24 mmol), FeCl3·6H2O (0.04 mmol, 20 mol%), HCl (0.24 mmol, 36–38%), n-hexane (1.0 mL), 80 °C, 14 h.bIsolated yield.The substrate scope of this iron-catalyzed esterification of secondary and tertiary amides with alcohols was investigated. Remarkably, when a secondary amide, viz., N-methylbenzamide (1o) and a tertiary amide, viz., N,N-diethylbenzamide (1p) were used as substrates, the corresponding esters 3a and 3v were afforded in high yield (Scheme 1).Open in a separate windowScheme 1Reaction of secondary and tertiary amides 1o and 1p with ethanol.In addition to aromatic primary, secondary and tertiary amides, the esterification of aliphatic primary, secondary and tertiary amides was also investigated under the optimal reaction conditions, the corresponding esters were afforded in good to excellent yield and the results are shown in Scheme 2. 2-Phenylacetamide (1q) participated in this catalytic reaction to form the ester 3y in 86% yield. N-Acetylaniline 1r and N,N-dimethylformamide 1s were also applicable to this catalytic system, and transformed into the corresponding esters 3z and 3z1 in 79% and 82% yield respectively.Open in a separate windowScheme 2Reaction of aliphatic amides 1q, 1r and 1s with different alcohols.To demonstrate the effect of iron salt in the current catalytic system, several control experiments were carried out as shown in Scheme 3. When 1 equivalent FeCl3·6H2O was used as the catalyst in the absence of HCl, 3a was obtained in 90% yield. This result indicated that the catalytic cycle of the iron salt was hindered in the present reaction conditions. When the stronger bidentate ligands such as 2,2-bipyridine was used, 3a was not obtained and trace amounts of 3a were detected on using ferrocene as the catalyst, indicating that the iron salt catalyst showed low efficiency in the presence of the stronger ligand. Thus, it was deduced that free Fe(ii or iii) was effective for this reaction.Open in a separate windowScheme 3Control experiments.According to the reported literature5d and our experimental results, the catalytic reaction pathways for the Fe-catalyzed esterification of amides by alcohols are proposed as shown in Scheme 4. The first step is the generation of the amidate complex A, which can be formed from free Fe(iii) and amide 1. The complex A reacts with alcohol 2 to produce an unstable intermediate B. The interaction between the alcohol oxygen and the carbonyl results in cyclic intermediate C, which is in equilibrium with its isomer D. Intermediate E can also be produced from Dvia C–N bond cleavage. Through the reaction of HCl and a new molecule amide, intermediate E produces the desired ester 3, ammonium chloride (colorless crystal, which was detected after the reaction), and the amidate complex A.Open in a separate windowScheme 4Plausible reaction mechanism for Fe-catalyzed esterification of amides by alcohols.  相似文献   

6.
Transition-metal-catalyzed remote C–H functionalization of thioethers     
Xiao-Qing Feng  He-Cheng Wang  Zhi Li  Long Tang  Xiaoqiang Sun  Ke Yang 《RSC advances》2022,12(17):10835
In the last decade, transition-metal-catalyzed direct C–H bond functionalization has been recognized as one of most efficient approaches for the derivatization of thioethers. Within this category, both mono- and bidentate-directing group strategies achieved the remote C(sp2)–H and C(sp3)–H functionalization of thioethers, respectively. This review systematically introduces the major advances and their mechanisms in the field of transition-metal-catalyzed remote C–H functionalization of thioethers from 2010 to 2021.

This minireview systematically introduces the major advances and their mechanisms in the field of transition-metal-catalyzed remote C–H functionalization of thioethers.  相似文献   

7.
Progress and prospects in copper-catalyzed C–H functionalization     
Thaipparambil Aneeja  Mohan Neetha  C. M. A. Afsina  Gopinathan Anilkumar 《RSC advances》2020,10(57):34429
Copper-catalyzed C–H functionalization is becoming a significant area in organic chemistry. Copper is now widely used as a catalyst in organic synthesis as it is inexpensive and not very toxic. Functionalization of C–H bonds to construct wide varieties of organic compounds has received much attention in recent times. This review focuses on the recent advances in Cu-catalyzed C–H functionalization and covers literature from 2018–2020.

Copper-catalyzed C–H functionalization has gone through some major progress in recent times. These efficient, selective and cost-effective reactions offer new avenues towards the synthesis of complex organic compounds.  相似文献   

8.
Ligand-free iridium-catalyzed regioselective C–H borylation of indoles     
Zilong Pan  Luhua Liu  Senmiao Xu  Zhenlu Shen 《RSC advances》2021,11(10):5487
We herein report a ligand-free Ir-catalyzed C–H borylation of N-acyl protected indoles. This simple protocol could tolerate a variety of functional groups, affording C3 borylated indoles in good yields with excellent regioselectivities. We also demonstrated that the current method is amenable to gram-scale borylation and the C–B bonds could be easily converted to C–C and C-heteroatom bonds.

First example of ligand-free Iridium-catalyzed Regioselective C–H borylation of indoles under mild reaction conditions.

Indoles are not only widespread subunits but also useful building blocks in drug discovery and synthetic chemistry.1 Thus, their functionalization and transformation have gained attention. In particular, borylated indoles are of significant importance because they can serve as useful synthons by converting C–B bonds into many other functionalities.2 As a result, a number of regioselective C–H borylation methods have been developed. In this context, regioselective transition-metal-catalyzed C–H borylation has emerged as a powerful tool for preparing borylated indoles in an atom- and step-economic way under mild reaction condition.3,4 Because the C–H borylation at the C2 positions of indoles are electronically more favorable,4fi the reactions at the other positions are more challenging. In general, directing groups are usually required to realize regioselective C–H borylation reactions. For example, bulky directing groups at nitrogen atoms such as Boc and Si(i-Pr)3 could result in C3-selective C–H borylation enabled by dtbpy/Ir catalysis (dtbpy = 4,4′-di-tert-butyl-2,2′-bipyridine).5 The N-Bpin moiety could serve as a traceless directing group for the dtbpy/Ir- and Ni(IMes)2-catalyzed C–H borylation at C3 positions.6 The use of C2 substituted indoles enables nitrogen-directed dtbpy/Ir-catalyzed C7-selective C–H borylation.7 Hartwig and co-workers used N–SiEt2H as the directing group to achieve relay directed dtbpy/Ir-catalyzed C7-selective C–H borylation.8 Another attractive and simple approach is the metal-free C–H borylation, which can regioselectively provide borylated indoles at C2, C3, C4, and C7 positions under mild reaction conditions.9 Despite the fact, the compatibility of functional groups is still narrow and usually limited to halogens, alkyl, and alkoxy substituted indoles. Thus, it is still appealing to develop complementary methods in this area.Ligand-free Ir-catalyzed regioselective C–H borylation of arenes have received growing interest. In this context, a judicious choice of directing group (DG) is crucial to promote the reaction. Usually, strong coordinating DGs are required. In this context, dithioacetal,10 pyrazorylaniline-modified boronic acid,11 phosphine,12 pyridine13 are commonly used for these transformations. It should be noted that all of the above-mentioned methods result in ortho-borylated products (Scheme 1A). In this work, we disclose the first example of ligand-free Ir-catalyzed C3-selective C–H borylation of N-acyl protected indoles. The current simple method could tolerate a variety of functionalities, including ester, ketone, nitro, and cyanide.Open in a separate windowScheme 1Ligand-free Ir-catalyzed C–H borylation of arenes.Our research commenced with the optimization of the reaction conditions. We chose N-isobutyryl indole 1a as our pilot substrate.14 The reaction 1a (0.20 mmol) with 1.5 equivalents of HBpin (pinacolborane) in the presence of a catalytic amount of [IrCl(cod)]2 (5.0 mol%) (cod: 1,5-cyclooctadiene) in n-hexane (1.0 mL) at 80 °C for 12 h affords C–H borylated product 2a in 79% yield with 97% C3 selectivity (15 Replacement of [IrCl(cod)]2 to either [Ir(OMe)(cod)]2 or [IrCl(coe)2]2 (coe: cyclooctene) results in significantly decreased yields (16 The solvent effect showed that the reaction in THF yield and regioselectivity (17Optimization of reaction conditions for the Ir-catalyzed distal hydroboration of 1aa
EntryVariation from standard conditions2a/2a′bYield of 2ac (%)
1None97 : 379
25 mol% [Ir(OMe)(cod)]2in lieu of [IrCl(cod)]297 : 343
35 mol% [IrCl(coe)]2in lieu of [IrCl(cod)]295 : 565
410 mol% P(C6F5)3<1 : 99
5THF in lieu of n-hexane89 : 1147
670 °C instead of 80 °C95 : 573
760 °C instead of 80 °C90 : 1031
Open in a separate windowaUnless otherwise noted, all the reactions were carried out with 1a (0.20 mmol), HBpin (0.30 mmol) in n-hexane (1.0 mL) at 80 °C for 12 h.bThe ratio of 2a/2a′ was determined by GC analysis.cIsolated yield of 2a.With optimized reaction conditions in hand ( Open in a separate windowaUnless otherwise noted, all the reactions were carried out with 1 (0.20 mmol), HBpin (0.30 mmol) in n-hexane (1.0 mL) at 80 °C for 12–24 h. The regioselectivity was determined by GC analysis.bThe regioselectivity was determined by 1H NMR of crude product.Interestingly, the reaction of C2-methyl substituted indole 3a under standard conditions could give C3-borylated product 4a exclusively, albeit with 40% isolated yield (eqn (1)). In contrast, no reaction was observed when C3-methyl or C2,C3-dimethyl substituted indoles 3b and 3c was employed (eqn (2)).In order to demonstrate the synthetic utility of the current protocol, a gram-scale reaction of 1a and several transformations of 2a were conducted as shown in Fig. 1. The reaction of 1a (1.00 g) with 2.5 mol% [IrCl(cod)]2 for 24 h afforded 2a in 79% isolated yield (1.32 g) with 97% C3 selectivity, which is almost identical with that obtained from the small-scale reaction. The C–B bond could be transferred to other functional groups bearing C–Br (5), C–I (6), C–CN (8), and C–Ph (10) bonds under various reaction conditions in good to excellent yields (70–90%).18–21 Interestingly, prolonging the reaction time of iodination and cyanation could ultimately result in deacylation products 7 and 9 in both 80% yields.Open in a separate windowFig. 1Gram-scale C–H borylation of 1a and synthetic application of borylated product 2a (DEMEDA = N,N′-dimethylenediamine).  相似文献   

9.
Solvent-free and room temperature microwave-assisted direct C7 allylation of indolines via sequential C–H and C–C activation     
Qiuling Wang  Linlin Shi  Shuang Liu  Changlei Zhi  Lian-Rong Fu  Xinju Zhu  Xin-Qi Hao  Mao-Ping Song 《RSC advances》2020,10(18):10883
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]2AdCOOH90402RuCl3·3H2OAdCOOH90N.R3[Cp*RuCl2]2AdCOOH90N.R4[Ru(p-cymene)Cl2]2MesCOOH90475[Ru(p-cymene)Cl2]2AcOH90406[Ru(p-cymene)Cl2]2NaOAc90207[Ru(p-cymene)Cl2]2PivONa·H2O90218[Ru(p-cymene)Cl2]2DABCO90Trace9b[Ru(p-cymene)Cl2]2MesCOOH905710b[Ru(p-cymene)Cl2]2MesCOOH706811b[Ru(p-cymene)Cl2]2MesCOOH508312b[Ru(p-cymene)Cl2]2MesCOOH256513b,c[Ru(p-cymene)Cl2]2MesCOOH2587 (>20 : 1)e14c,d[Cp*Rh(CH3CN)3](SbF6)2AdCOOH8078 (10 : 1)eOpen 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   

10.
C–H activation-annulation on the N-heterocyclic carbene platform     
Champak Dutta  Joyanta Choudhury 《RSC advances》2018,8(49):27881
Ring-fused cationic N-heterocycles are an important class of organic compounds recognized to be of significant interest in diverse research areas including bioactivity, materials chemistry, supramolecular chemistry, etc. Toward the synthesis of such molecules, recently unique chemistry has been explored utilizing a novel conjugative action of NHC ligands as a functionalizable directing group in rhodium(iii)-catalyzed aromatic/heteroaromatic/non-aromatic C–H activation and subsequent annulation of various imidazolium salts with internal alkynes. This review highlights the initial development and underscores the potential of this chemistry.

This review highlights the initial development of a new C–H activation–annulation chemistry accessible on the metal–N-heterocyclic carbene platform.  相似文献   

11.
Palladium-catalyzed dehydrogenative C–H cyclization for isoindolinone synthesis     
Masahiro Abe  Kaho Ueta  Saki Tanaka  Tetsutaro Kimachi  Kiyofumi Inamoto 《RSC advances》2021,11(43):26988
In this paper Pd-catalyzed intramolecular dehydrogenative C(sp3)–H amidation for the synthesis of isoindolinones is described. This method features the use of a Pd/C catalyst and the addition of a stoichiometric amount of oxidant is not necessary. A mechanistic study suggested the possible formation of H2 gas during the reaction.

Pd-catalyzed intramolecular dehydrogenative C(sp3)–H amidation for the synthesis of isoindolinones was developed. Use of Pd/C as a catalyst enables the desired cyclization to proceed smoothly without adding any stoichiometric oxidants.

The isoindolinone scaffold occurs frequently in numerous biologically active compounds ranging from designed medicinal agents to natural products, thus constituting an extremely important class of heterocycles.1,2 Although a number of synthetic methods exist for the preparation of isoindolinones,3 the development of more general, versatile, and efficient procedures to construct the isoindolinone framework is still an ongoing, intensive research area.Among a variety of synthetic strategies for isoindolinones, methods that make use of transition metal-catalyzed C–H functionalization represent a remarkable approach.4 In this research area, catalytic oxidative annulation of N-substituted benzamides and appropriate coupling partners (e.g. alkenes, alkynes, isocyanides, diazo compounds) has been extensively studied, most of which employ transition metals such as rhodium,4d,g,h,n,o ruthenium,4l cobalt,4a,i,j,k and palladium.4b,m These precedents often require the use of stoichiometric oxidants or prefunctionalized coupling partners to make the process catalytic.On the other hand, an approach that involves intramolecular C–H cyclization of benzamide derivatives exploiting a transition metal catalyst could provide another efficient, facile, and direct access to isoindolinones, although less attention has been paid to such a process (Scheme 1). Kondo et al. previously reported C(sp3)–H aminative cyclization of 2-methyl-N-arylbenzamides in the presence of a copper catalyst along with a stoichiometric amount of di-tert-butyl peroxide as an oxidant, which resulted in the formation of 3-unsubstituted isoindolinones (Scheme 1a).4c Bedford''s group also developed a copper-based catalytic system composed of 20 mol% of Cu(OTf)2 and 2 equiv. of PhI(OAc)2 that successfully effected the intramolecular benzylic C–H sulfamidation of 2-benzyl-N-tosylbenzamides for the synthesis of isoindolinones (Scheme 1b).4f In both cases, the choice of an oxidant should be crucial for the successful construction of the isoindolinone ring.Open in a separate windowScheme 1C–H cyclization strategies for isoindolinone synthesis.Herein, we describe a catalytic system that enables cyclization of 2-benzyl-N-mesylbenzamides leading to isoindolinone derivatives, in which benzylic C(sp3)–H functionalization in the presence of a palladium catalyst smoothly occurs. It is particularly noteworthy to mention that any stoichiometric oxidants are not necessary for our isoindolinone synthesis. The key to success is the use of Pd/C as a catalyst. In this dehydrogenative process, an only detectable by-product was H2, which should also be an appealing feature of the process.Based on our fruitful results of the heterocycles synthesis via transition metal-catalyzed C–H cyclization,5 our investigation began by examining the conversion of N-tosyl-protected benzamide 1a to the corresponding isoindolinone 2a. During the screening studies utilizing a range of transition metal catalysts as well as various oxidants, it was found that 2a did produce even in the absence of an oxidant when Pd/C was used as a catalyst. Indeed, the use of 10 mol% of Pd/C along with 20 mol% of KOAc enabled the desired cyclization process, providing 2a in 42% yield (6Effect of protecting groupa,b
Open in a separate windowaReactions were run on a 0.25 mmol scale.bIsolated yields (yield determined by 1H NMR using an internal standard in parentheses).Settled in –Ms for the protecting group, further examination of the reaction parameters was performed employing 1f (7,8Optimization of reaction parametersa
EntryVariation from “standard conditions"Yieldb,c (%)
1None(75)
2Na2HPO4 instead of KOAc86 (80)
3NaOAc, LiOAc, K2CO3 or Cs2CO3 instead of KOAc61–70
4DMA, DMI or DMSO instead of p-xylene0–28
50.25 M instead of 0.05 M67
6dO2 atmosphere instead of Ar atmosphere4
7dDegassed (bubbling with Ar)92 (86)
8dIn the absence of KOAc72
Open in a separate windowaReactions were run on a 0.25 mmol scale.bYields were determined by 1H NMR using an internal standard.cIsolated yields in parentheses.dNa2HPO4 was used instead of KOAc.Our new method for isoindolinone synthesis can be applied to cyclization of a wide range of substrates (9 An attempt to obtain 3-alkylisoindolinone 2r unfortunately failed. On the other hand, the reaction of 1f can be easily scaled up: in this case, 2f was obtained in 92% yield.10Substrate scopea,b
Open in a separate windowaReactions were run on a 0.25 mmol scale.bIsolated yields.c1 mmol scale.To further demonstrate the synthetic utility of the method we have developed, several transformations of isoindolinone 2f obtained were carried out (Scheme 2). Deprotection of a mesyl group with AIBN and Bu3SnH afforded the N-free isoindolinone 3 in high yield. Reduction of the carbonyl group of 2f using LiAlH4 also successfully preceded, giving rise to isoindoline 4 in excellent yield (97%).Open in a separate windowScheme 2Further transformation of isoindolinone 2f.To gain insight into the reaction mechanism of the process, C–H cyclization of 1f in the presence of 0.5 equiv. of methyl cinnamate (5) was performed under the optimized conditions (Scheme 3). In addition to isoindolinone 2f, methyl 3-phenylpropanoate (6) was observed in 32% NMR yield, implying that the formation of H2 gas is possibly occurring during the reaction.11Open in a separate windowScheme 3Mechanistic studies.On the basis of the finding above, a tentative reaction pathway shown in Scheme 4 is proposed. The reaction is likely initiated by the coordination of the nitrogen atom of an amide moiety to give complex A. Subsequently, the insertion of Pd(0) into the benzylic C(sp3)–H bond leading to the formation of six-membered palladacycle B accompanied with the evolution of H2 gas and the following reductive elimination process affords the desired isoindolinone 2.Open in a separate windowScheme 4Plausible reaction mechanism.In summary, we have developed intramolecular Pd-catalyzed dehydrogenative C(sp3)–H amidation for isoindolinone synthesis. The addition of oxidants is not necessary and the Pd/C catalyst along with a catalytic amount (20 mol%) of base is the only reagents required for this C–H cyclization. The method developed provides a simple, facile, and efficient access to a biologically important isoindolinone nucleus. Further studies to broaden the substrate scope of the process as well as to improve the catalytic efficiency are vigorously underway in our laboratory.  相似文献   

12.
Step-saving synthesis of star-shaped hole-transporting materials with carbazole or phenothiazine cores via optimized C–H/C–Br coupling reactions     
Jui-Heng Chen  Kun-Mu Lee  Chang-Chieh Ting  Ching-Yuan Liu 《RSC advances》2021,11(15):8879
In most research papers, synthesis of organic hole-transporting materials relies on a key-reaction: Stille cross-couplings. This requires tedious prefunctionalizations including the preparation and treatment of unstable organolithium and toxicity-concern organotin reagents. In contrast to traditional multistep synthesis, this work describes that a series of star-shaped small molecules with a carbazole or phenothiazine core can be efficiently synthesized through a shortcut using optimized direct C–H/C–Br cross-couplings as the key step, thus avoiding dealing with the highly reactive organolithium or the toxic organotin species. Device fabrication of perovskite solar cells employing these molecules (6–13) as hole-transporting layers exhibit promising power conversion efficiencies of up to 17.57%.

Carbazole or phenothiazine core-based hole-transport materials are facilely accessed by an optimized synthesis-shortcut. Perovskite solar cell devices with 6–13 demonstrate PCEs of up to 17.57%.  相似文献   

13.
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   

14.
Visible light-induced photocatalytic C–H ethoxycarbonylmethylation of imidazoheterocycles with ethyl diazoacetate     
Suvam Bhattacharjee  Sudip Laru  Sadhanendu Samanta  Mukta Singsardar  Alakananda Hajra 《RSC advances》2020,10(47):27984
A visible light-mediated regioselective C3-ethoxycarbonylmethylation of imidazopyridines with ethyl diazoacetate (EDA) was achieved under mild reaction conditions. In contrast to the carbene precursors from α-diazoester a first C3-ethoxycarbonylmethylation of imidazopyridines via a radical intermediate has been established. The present methodology provides a concise route to access pharmacologically useful esters with wide functional group tolerance in high yields.

A visible-light-promoted regioselective ethoxycarbonylmethylation of imidazoheterocycles has been developed using an α-diazoester via a radical pathway.

Ester groups are a most powerful and versatile synthetic building block in natural products as well as in organic synthesis since they can be derivatized to diverse functional groups such as carbonyl, hydroxymethyl, and amide, etc.1 Beside the traditional methods of using xanthates for ethoxycarbonylmethylation,2 the introduction of diazoesters represents a very important strategy for the synthesis of acetate derivatives.3 Generally, thermal,4a photoinduced,4b or metal catalyzed4c reactions of diazo compounds produce reactive carbene precursors which further undergo an addition reaction, insertion reaction, cyclopropanation, Wolff rearrangement, etc. (Scheme 1a).4 In contrast to the carbene precursors from α-diazoester, a new methodology is always highly desirable especially with a significantly different mode of activation under a distinct mechanism.5 In this context, a new strategy for installation of alkyl radicals from diazo compounds under photoredox-catalysis is very important.6 Recently, visible light-mediated photoredox catalyzed direct C–H bond functionalization reactions have come to the forefront in synthetic organic chemistry7 where Ru(ii)-polypyridyl complexes have been extensively used as photoredox-catalysts due to their unique photo-physical properties.8Open in a separate windowScheme 1Visible light-mediated ethoxycarbonylmethylation. (a) Previous reports through carbene pathway. (b) This work: visible-light mediated radical reaction of diazoacetate with imidazopyridine.Imidazo[1,2-a]pyridines, an important class of nitrogen-based privileged structural motifs widely found in a variety of natural products and alkaloids.9 In pharmaceutical chemistry, it possesses large number of biological activities like antitumor, antibacterial, antifungal, antipyretic, analgesic, antiinflammatory, etc.10a It is also important in material sciences due to its unique excited state intramolecular proton transfer phenomena (ESIPT).10b In particular, ethoxycarbonylmethylated imidazopyridines are the core structure of several marketed drugs such as alpidem, zolpidem, saripidem, etc.11 Therefore, practical methodologies for the synthesis of alkylated ester substituted imidazoheterocycles are highly desirable and will be of great interest to the synthetic chemists.11 However, to the best of our knowledge there is no such method for the direct photo-initiated ethoxycarbonylmethylation reaction of imidazoheterocycles with diazo compounds through distinct radical pathway.11–13 As part of our ongoing research on photoredox-catalyzed C–H functionalization reactions,13 herein we describe an environmentally benign visible light-promoted ethoxycarbonylmethylation for the synthesis of alkyl substituted imidazopyridines by the coupling between imidazopyridines and ethyl diazoacetate using Ru(bpy)3Cl2 as photoredox-catalyst under argon atmosphere at room temperature (Scheme 1b).To begin, the reaction was carried out using 2-phenylimidazo[1,2-a]pyridine (1a) and ethyl diazoacetate (EDA) (2) (2.0 equiv.) in presence of 0.2 mol% of Ru(bpy)3Cl2 as a photoredox-catalyst under the irradiation of 34 W blue LED lamp in methanol at room temperature. To our delight, 49% yield of the ethyl 2-(2-phenylimidazo[1,2-a]pyridin-3-yl)acetate (3a) was obtained after 36 h under argon atmosphere. However, no desired product was formed in aerobic conditions ( EntryPhotocatalyst (0.2 mol%)SolventYield (%)1Ru(bpy)3Cl2MeOH49, NRb2Ru(bpy)3Cl2EtOH213Ru(bpy)3Cl2DCM434Ru(bpy)3Cl2MeCN475Ru(bpy)3Cl2TolueneNR6Ru(bpy)3Cl2MeOH : H2O (1 : 1)767Ru(bpy)3Cl2MeOH : H2O (2 : 1)898Ru(bpy)3Cl2MeOH : H2O (1 : 2)Trace9Ru(bpy)3Cl2H2ONR10Ru(bpy)3Cl2EtOH : H2O (2 : 1)6811Ir(ppy)3MeOH : H2O (2 : 1)Trace12Rose bengalMeOH : H2O (2 : 1)NR13Eosin YMeOH : H2O (2 : 1)NR14Eosin BMeOH : H2O (2 : 1)NR15Rhodamine BMeOH : H2O (2 : 1)NR16—MeOH : H2O (2 : 1)NRc17Ru(bpy)3Cl2MeOH : H2O (2 : 1)NRd18Ru(bpy)3Cl2MeOH : H2O (2 : 1)NRe19Ru(bpy)3Cl2MeOH : H2O (2 : 1)45f, 84g20Ru(bpy)3Cl2MeOH : H2O (2 : 1)56h, NRiOpen in a separate windowaReaction conditions: all reactions were carried out with 0.25 mmol of 1a, 0.5 mmol of 2, and 0.2 mol% of photocatalyst in 2.0 mL of solvent for 36 h at room temperature under argon atmosphere and irradiation with 34 W blue LED. NR = no reaction.bIn aerobic condition.cAbsence of Ru(bpy)3Cl2.dWithout light source at rt.eReaction at 60 °C without light scource.f1.5 equiv. EDA.g2.5 equiv. EDA.hIrradiation with 20 W blue LED.i30 W white LED.With the optimized reaction conditions in hand we explored the substrate scope of this methodology with various substituted imidazopyridines. At first we checked the effect of various substituents on the pyridine ring of imidazopyridine derivatives and the results are summarized in Scheme 2. Imidazopyridine containing various electron-donating groups (8-Me and 7-OMe) efficiently reacted with ethyl diazoacetate to provide the desired products in good yields (3b and 3c). It is important to note that substrates bearing electron-withdrawing groups did not afford alkylated products under the optimized reaction condition. However, addition of 10 mol% N,N-dimethyl-m-toluidine as a redox-active additive gave a satisfactory yield of the desired products (see ESI). Therefore in modified reaction conditions, halogens containing imidazopyridines like –F, –Br, –I and strong electron-withdrawing group such as –CF3, –CN containing substrates were successfully afforded the desired products in 67–90% yields (3d–3h). Next, we examined the substrate scope with various substitutions in phenyl part of imidazo[1,2-a]pyridine. Imidazopyridine with electron-releasing (4-Me, and 4-OMe), halogens (4-F, 4-Cl, and 3-Br) and strong electron-withdrawing group (4-CF3, and 4-CN) containing substrates produced the desired products with good to excellent yields (3i–3p). 2-Hydroxy substituted imidazo[1,2-a]pyridine was also worked well in this transformation (3k). In addition, disubstituted imdazopyridines with different substituents at both the phenyl and pyridine ring underwent the reaction very smoothly (3q, and 3r). Heteroaryl and naphthyl substituted imidazopyridines produced the desired products without any difficulties (3s–3u). However, unsubstituted imidazo[1,2-a]pyridine did not produce the desired product in this protocol. To highlight the applicability of the present protocol a gram-scale reaction was carried out taking 1a in 6.0 mmol scale under normal laboratory setup. The yield of ethyl 2-(2-phenylimidazo[1,2-a]pyridin-3-yl)acetate (3a) was slightly diminished to 70% after 36 h along with the recovery of excess starting material.Open in a separate windowScheme 2Substrate scope of imidazopyridines.aaReaction conditions: 0.25 mmol of 1, 0.5 mmol of 2 in presence of 0.2 mol% of Ru(bpy)3Cl2 in 2.0 mL of MeOH : H2O (2 : 1) at room temperature under 34 W blue LED and argon atmosphere for 36 h. b10 mol% N,N-dimethyl-m-toluidine was added as an additive. c6.0 mmol scale.Next, to show the generality of this present methodology, we investigated the reaction taking other imidazoheterocycles like imidazo[2,1-b]thiazole and benzo[d]imidazo[2,1-b]thiazoles (Scheme 3). 6-Phenylimidazo-[2,1-b]thiazole reacted smoothly to afford 5a in 66% yield. Benzo[d]imidazo[2,1-b]thiazole containing –Me, –OMe, and –Cl substituted derivatives were well tolerable under the present reaction conditions (5b–5e). The structure of ethyl 2-(7-methoxy-2-phenylbenzo[d]imidazo[2,1-b]thiazol-3-yl)acetate (5c) was confirmed by X-ray crystallography.14 In addition, thiophene substituted benzo[d]imidazo[2,1-b]thiazole furnished the corresponding alkylated product in good yield (5f).Open in a separate windowScheme 3Substrate scope.aaReaction conditions: 0.25 mmol of 4, 0.5 mmol of 2 in presence of 0.2 mol% of Ru(bpy)3Cl2 in 2.0 mL of MeOH : H2O (2 : 1) at room temperature under 34 W blue LED and argon atmosphere for 36 h. b10 mol% N,N-dimethyl-m-toluidine was added as an additive.After that, we have directly modified the synthesized ethoxycarbonylmethylated product (3q) to amide derivative (6q) in 56% yield using benzylamine in the presence of La(OTf)3 as a catalyst under ambient air at rt to 70 °C as shown in Fig. 1.Open in a separate windowFig. 1Modification of ester group via direct amidation process.A few control experiments were performed to understand the mechanistic insights of the present protocol (Scheme 4). The alkylation reaction did not proceed in the presence of 3.0 equiv. radical scavengers like 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO), 2,6-di-tert-butyl-4-methyl phenol (BHT), 1,1-diphenylethylene (DPE), and p-benzoquinone (BQ) which suggest the reaction probably proceeds through radical pathway (Scheme 4, eqn (A)). Moreover, 3-phenylimidazo[1,2-a]pyridine (1v) did not produce the C-2 alkylated product that signifies the regioselectivity of the present protocol (Scheme 4, eqn (B)). Taking 2-(2-(vinyloxy)phenyl)imidazo[1,2-a]pyridine we also performed this reaction with ethyl diazoacetate under this standard reaction conditions. Desired C-3 alkylated product (8a) was formed in 35% yield (Scheme 4, eqn (C)). But cyclopropanation product was not formed in the alkene part of the allyl group via carbene pathway. From this experimental result strongly imply that the present strategy only proceeds through radical mechanistic pathway.Open in a separate windowScheme 4Control experiments.Based on the above mechanistic studies and previous literature reports,7,8,13 a possible radical mechanism was proposed for the formation of ethyl 2-(2-phenylimidazo[1,2-a]pyridin-3-yl)acetate (3a) as presented in Scheme 5. Initially, photo-catalyst Ru(bpy)32+ is transformed to its excited-state *Ru(bpy)32+ under blue LED irradiation. In the presence of MeOH/H2O, diazo compounds are in equilibrium with their protonated form (A). Then cationic diazoester involves single-electron transfer (SET) process with excited photocatalyst *Ru(bpy)32+ leading to the formation of alkyl radical intermediate (B) and Ru(bpy)33+.6b,c Next, alkyl radical intermediate (B) effectively reacts with imidazopyridine (1a) to produce imidazole radical intermediate (C) which is successively transformed to imidazolium radical cation intermediate (D) along with the generation of ground state photocatalyst. Finally, C-3 alkylated imidazopyridine (3a) is obtained through deprotonation of intermediate D. It is worthy to mention that Meggers et al.7h did not observe quenching of *Ru(bpy)32+ whereas Gryko et al.6c did observe the quenching of *Ru(bpy)32+ by ethyl diazoacetate that was accelerated in the presence of protic sources.Open in a separate windowScheme 5Proposed mechanistic pathway.In case of halogen or electron-withdrawing substituted imidazoheterocycles a possible mechanism for the radical alkylation reaction is illustrated in Scheme 6.6b,c Reductive quenching of *Ru(bpy)32+ by SET oxidation of aniline derivatives affords radical cation E and Ru(bpy)3+. Then reduced photocatalyst Ru(bpy)3+ comes back to ground state with the generation of radical intermediate Bvia the elimination of N2. Subsequently, radical intermediate B reacts with imidazopyridine 1 to produce radical intermediate C′. Finally, the desired product 3′ is produced through the hydrogen atom transfer (HAT) process between radical cation E and radical intermediate C′.Open in a separate windowScheme 6Mechanistic pathway for electon-deficient substrates.  相似文献   

15.
Metal catalyzed C–H functionalization on triazole rings     
Anushka Koranne  Khushboo Kurrey  Prashant Kumar  Sangeeta Gupta  Vikesh Kumar Jha  Rangnath Ravi  Prasanta Kumar Sahu  Anamika  Abadh Kishor Jha 《RSC advances》2022,12(42):27534
The present review covers advancement in the area of C–H functionalization on triazole rings, by utilizing various substrates with palladium or copper as catalysts, and resulting in the development of various substituted 1,2,3- and 1,2,4-triazoles. Synthesis of these substituted compounds is necessary from the perspective of pharmaceutical, medicinal, and materials chemistry.

The present review covers advancement in the area of C–H functionalization on triazole rings, by utilizing various substrates with palladium or copper as catalysts, and resulting in the development of various substituted 1,2,3- and 1,2,4-triazoles.  相似文献   

16.
Direct phosphorylation of benzylic C–H bonds under transition metal-free conditions forming sp3C–P bonds     
Qiang Li  Chang-Qiu Zhao  Tieqiao Chen  Li-Biao Han 《RSC advances》2022,12(29):18441
Direct phosphorylation of benzylic C–H bonds was achieved in a biphasic system under transition metal-free conditions. A selective radical/radical sp3C–H/P(O)–H cross coupling was proposed, and various substituted toluenes were applicable. The transformation provided a promising method for constructing sp3C–P bonds.

Direct phosphorylation of benzylic C–H bonds with secondary phosphine oxides was first realized. The reaction was performed in organo/aqueous biphasic system and under transition metal-free conditions, proceeding via the cross dehydrogenative coupling.

To construct C–P bonds is of great significance in modern organic synthesis,1 because organophosphorus compounds play varied roles in medical,2,3 materials,4 and synthetic chemistry fields.5 Traditionally, the C–P bonds were formed from P–Cl species via nucleophilic substitutions with organometallic reagents,6 P–OR species via Michaelis–Arbusov reactions,7 or P–H species via the alkylation in the present of a base or a transition metal.8Over the past decades, cross dehydrogenative coupling reactions (CDC reactions) have become a powerful and atom-economic methodology for constructing chemical bonds.9 By using this strategy, C–H bonds can couple with Z–H bonds without prefunctionalization and thus short-cut the synthetic procedures (Scheme 1a).Open in a separate windowScheme 1Cross dehydrogenative coupling reactions and direct phosphorylation of benzylic C–H bonds.A similar construction of C–P bonds via CDC was also realized.10,11 Among these methods, the phosphorylation of sp3C–H having an adjacent N or O atom, or the carbonyl group was well-developed.11 Relatively, the formation of benzylic sp3C–P bond was less reported,11w which was mainly limited to the sp3C–H of xanthene or 8-methylquinoline (Scheme 1b).12 In these reported processes, transition metal catalysts or photo-, electro-catalysts were usually involved,11v and an excess of P(O)–H compounds was usually employed.11,13 To the best of our knowledge, the phosphorylation of non-active benzylic C–H bonds has scarcely been reported.Considering both benzylic and phosphorus radicals could be generated by oxidation,14 which might subsequently couple, the phosphorylation of benzylic sp3C–H bonds would be achieved (Scheme 1c). Herein, we disclosed the construction of benzylic C–P bonds from toluene and P–H species. The reaction was carried out under transition metal-free reaction conditions,15 and exhibited high regio-selectivity. The aromatic C–H remained intact during the reaction.We began our investigation by exploring the reaction of toluene 1a and diphenylphosphine oxide 2a in the presence of an oxidant (16 Other persulfates could also be used as an oxidant albeit with decreasing yields ( EntryOxidantToluene/water (v/v)Additive3a yieldb (%)3a/4c1K2S2O81 : 0—Trace—2K2S2O81 : 1—1216 : 843K2S2O81 : 1SDS3726 : 744Na2S2O81 : 1SDS3127 : 735(NH4)2S2O81 : 1SDS2723 : 776Oxone1 : 1SDSNone—7—1 : 1SDSNone—8—1 : 1—None—9K2S2O81 : 1SDBS4136 : 6410K2S2O81 : 2SDBS4628 : 7211K2S2O81 : 3SDBS3524 : 7612dK2S2O81 : 2SDBS4632 : 6813eK2S2O81 : 2SDBS2621 : 7914fK2S2O81 : 2SDBS2926 : 7415gK2S2O81 : 2SDBS4838 : 6216g,hK2S2O81 : 2SDBS0—Open in a separate windowaReaction condition: 1a (1 mL), 2a (0.2 mmol), oxidant (2 equiv.), additive (1 equiv.) and H2O, 120 °C, 3 h. under N2.bGC yields using n-dodecane as an internal standard.cThe ratio of 3a/4 was determined by GC analysis.d50 mol% SDBS was used.e20 mol% SDBS was used.fAt 100 °C.g1 (0.8 mL) and H2O (1.6 mL), 3 equiv. K2S2O8 was used, 120 °C for 15 min.h3 equiv. TEMPO was added.Based on the above results, we can easily find that a serious amount of 1,2-diphenylethane 4 was formed. These results suggest that the homocoupling rate of 1a was very quick. Thus, the choice of the phase transfer reagent and oxidant is the key to cross-coupling of toluene and diphenylphosphine oxide in this biphasic solvent system.Excessive P–H species were usually employed in reported CDC reactions to form C–P bonds, because of their facile oxidation.13 In our procedure, toluene was excessive, thus the yields were calculated based on 2a. The yields looked like low, which did not indicate the poorer conversion rate of P–H species. With the optimized conditions in hand, the substrate scope of the CDC reactions was explored ( Open in a separate windowaReaction conditions: 1 (0.8 mL), 2 (0.2 mmol), K2S2O8 (3 equiv.), SDBS (50%) and H2O (1.6 mL), 120 °C, under N2, the reactions were monitored by TLC and/or GC until 2 work out.b1 mmol scale, 30 min.c130 °C.d100 °C.In addition to toluene, o-xylene, m-xylene, p-xylene, mesitylene, and 1,2,4,5-tetramethylbenzene all coupled with 2a to give the expected 3b–f in moderate yields. Methoxy substituted toluene gave relatively lower yields of 3g and 3h. para-Halo substituted toluene exhibited good reactivity, furnishing the coupling products 3i and 3j in moderate yields. Comparing to toluene, a decreasing order of reactivity was observed for ethyl benzene (3l, 30% yield), isobutyl benzene (3m, 27% yield), isopropyl benzene (3n, <10% yield), and diphenyl methane (3o, trace). The order was probably controlled by the steric hindrance around the benzylic carbon. 1-Methylnaphthalene and 2-methylnaphthalene also gave low yields (3p and 3q). However, 2-methylquinoline served well and coupled with 2a, affording the product 3r in 49% yield, which could be ascribed to the activation of the nitrogen atom. Besides of 2a, diaryl phosphine oxides having methyl, F, and Cl substituents could also be employed as the substrate, producing 3s–3u in moderate yields under similar reaction conditions.Although the mechanism of the direct phosphorylation of benzyl C–H bond in aqueous solution is not quite clear, some aspects could be grasped based on experimental results. Firstly, the desired product 3a was not detected when the 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) was added, implying that a radical pathway might be possible in this reaction (entry 16). Secondly, the reaction only occurred in aqueous phase, and relied on the presence of PTC (phase transfer catalyst), as seen in that 3a was difficultly formed in entries 1 and 2 of 17We supposed 1 is converted to benzyl radical 5 by SO4 radical that is generated via hemolytic cleavage of potassium persulfates.18 Meanwhile, phosphorus radical 6 is similarly formed from 2. Because potassium persulfate was water soluble, both 1 and 2 had to be transferred into aqueous solution to react with potassium persulfates.19 This proposal is also in accord with the experimental results that no products of sp2C–H phosphorylation are detected, which are the main products in the previous radical systems.20 Finally, cross couplings between 5 and 6 produce 3 (Scheme 2).Open in a separate windowScheme 2Proposed mechanism for the direct phosphorylation of benzylic C–H bonds under transition metal-free reaction conditions.  相似文献   

17.
Evaluation and comparison of antioxidant abilities of five bioactive molecules with C–H and O–H bonds in thermodynamics and kinetics     
Yan-Hua Fu  Zhen Wang  Kai Wang  Guang-Bin Shen  Xiao-Qing Zhu 《RSC advances》2022,12(42):27389
In this work, the antioxidant abilities of NADH coenzyme analogue BNAH, F420 reduction prototype analogue F420H, vitamin C analogue iAscH, caffeic acid, and (+)-catechin in acetonitrile in chemical reactions were studied and discussed. Three physical parameters of the antioxidant XH, homolytic bond dissociation free energy ΔG°(XH), self-exchange HAT reaction activation free energy ΔGXH/X, and thermo-kinetic parameter ΔG≠°(XH), were used to evaluate the antioxidant ability of XH in thermodynamics, kinetics, and thermo-kinetics. By comparing ΔG°(XH), ΔGXH/X and ΔG≠°(XH) of these five bioactive antioxidants to release hydrogen atoms, it is easy to find that iAscH is the best hydrogen atom donor both thermodynamically and kinetically among these antioxidants. Caffeic acid is the worst hydrogen atom donor thermodynamically, and F420H is the worst hydrogen atom donor kinetically. In addition, the thermodynamic hydride donating abilities of BNAH, F420H, and iAscH were also discussed, and the order of thermodynamic hydride donating abilities was BNAH > F420H > iAscH. Four HAT reactions BNAH/DPPH˙, (+)-catechin/DPPH˙, F420H/DPPH˙, and caffeic acid/DPPH˙ in acetonitrile at 298 K were studied by the stopped-flow method. The actual order of H-donating abilities of these four antioxidants in the HAT reactions is consistent with the order predicted by thermo-kinetic parameters. It is feasible to predict accurately the antioxidant abilities of antioxidants using thermo-kinetic parameters.

In this work, the antioxidant abilities of NADH coenzyme analogue BNAH, F420 reduction prototype analogue F420H, vitamin C analogue iAscH, caffeic acid, and (+)-catechin in acetonitrile in chemical reactions were studied and discussed.  相似文献   

18.
Reactions of triosmium and triruthenium clusters with 2-ethynylpyridine: new modes for alkyne C–C bond coupling and C–H bond activation     
Md. Tuhinur R. Joy  Roknuzzaman  Md. Emdad Hossain  Shishir Ghosh  Derek A. Tocher  Michael G. Richmond  Shariff E. Kabir 《RSC advances》2020,10(51):30671
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   

19.
An efficient nickel/silver co-catalyzed remote C–H amination of 8-aminoquinolines with azodicarboxylates at room temperature     
Ruinan Zhao  Yaocheng Yang  Xia Wang  Peng Ren  Qian Zhang  Dong Li 《RSC advances》2018,8(65):37064
A highly efficient nickel/silver co-catalyzed C–H amination at the C5 position of 8-aminoquinolines with azodicarboxylates at room temperature is reported. The reaction undergoes a self-redox process without the necessity of external oxidant. It proceeded under simple and mild conditions without any additional ligand, base or oxidant and provided the desired products in good to excellent yields. This method also possessed the merits of good functional group compatibility and air and moisture tolerance. It provides an efficient strategy for the synthesis of useful quinoline derivatives.

C–H amination at the C5 position of 8-aminoquinolines with azodicarboxylates proceeded efficiently using a nickel/silver co-catalyst at room temperature without any additional ligand, base or oxidant.  相似文献   

20.
Direct C–H photoarylation of diazines using aryldiazonium salts and visible-light     
Rodrigo C. Silva  Lucas F. Villela  Timothy J. Brocksom  Kleber T. de Oliveira 《RSC advances》2020,10(52):31115
In this study, direct C–H photoarylation of pyrazine with aryldiazonium salts under visible-light irradiation (blue-LEDs) is described, and additional examples including photoarylations of pyrimidine and pyridazine are also covered. The corresponding aryl-diazines were prepared in yields up to 84% only by mixing and irradiating the reaction with no need for an additional photocatalyst. We demonstrate the efficacy of this protocol by the scope with electron-donor, -neutral, and -withdrawing groups attached at the ortho, meta, and para positions of the aryldiazonium salts; the results are better than those reported for ruthenium-complex mediated photoarylations. Additionally, we demonstrate the robustness of this methodology with a 5 mmol scaled-up experiment. Mechanistic studies were carried out giving support to the proposal of a photocatalyzed approach by an electron donor–acceptor (EDA) complex, also highlighting the crucial role that solvents play in the formation of the EDA complex.

An electron donor–acceptor (EDA) approach for the direct C–H photoarylation of diazines using aryldiazonium salts and visible-light is described.  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