Search results

Search for "C(sp3)–H bond" in Full Text gives 33 result(s) in Beilstein Journal of Organic Chemistry.

Non-noble metal-catalyzed cross-dehydrogenation coupling (CDC) involving ether α-C(sp3)–H to construct C–C bonds

  • Hui Yu and
  • Feng Xu

Beilstein J. Org. Chem. 2023, 19, 1259–1288, doi:10.3762/bjoc.19.94

Graphical Abstract
  • bonds The possible mechanism of the CDC reaction involving ether α-C(sp3)–H bonds mainly follows the two pathways outlined in Scheme 2. Route a: First, the C(sp3)–H bond at the α-position of the oxygen atom undergoes a single-electron transfer under the combined action of the transition metal and an
  • ][48][49][50]: 1) CuII → CuI → CuII; 2) CuI → CuIII → CuI; 3) CuII → CuIII → CuI → CuII. In 2006, Li et al. demonstrated that the CDC reaction of the C(sp3)–H bond of malonate diesters or other active methylene compounds with the C(sp3)–H bond adjacent to the oxygen atom of cyclic and open-chain
  • extract a hydrogen from the ether C (sp3)–H bond to form radicals. Subsequently, a single electron transfer (SET) leads to the oxonium species. Then, the enamine generated in situ from methyl aryl ketone and pyrrolidine undergoes a nucleophilic reaction with the oxonium species followed by hydrolysis to
PDF
Album
Review
Published 06 Sep 2023

Photoredox catalysis harvesting multiple photon or electrochemical energies

  • Mattia Lepori,
  • Simon Schmid and
  • Joshua P. Barham

Beilstein J. Org. Chem. 2023, 19, 1055–1145, doi:10.3762/bjoc.19.81

Graphical Abstract
PDF
Album
Review
Published 28 Jul 2023

Transition-metal-catalyzed C–H bond activation as a sustainable strategy for the synthesis of fluorinated molecules: an overview

  • Louis Monsigny,
  • Floriane Doche and
  • Tatiana Besset

Beilstein J. Org. Chem. 2023, 19, 448–473, doi:10.3762/bjoc.19.35

Graphical Abstract
  • trifluoromethylthiolation persist. In particular, the functionalization of a C(sp3)–H bond with a trifluoromethylthiolated moiety by transition-metal-catalyzed C–H activation remains a challenging task both in terms of reactivity and selectivity. In 2015, Besset reported the first C(sp3)–SCF3 bond formation of unactivated
  • -C(sp3)–H bond (21 examples, up to 53% yield). The methodology was applied to the functionalization of a series of amides having an α-quaternary center (α,α-dialkyl (31a), α-alkyl,α-benzyl derivatives 31c–f) as well as to an amide with an α-tertiary center (31b) and pleasingly, the presence of α-C–H
PDF
Album
Review
Published 17 Apr 2023

Iron-catalyzed domino coupling reactions of π-systems

  • Austin Pounder and
  • William Tam

Beilstein J. Org. Chem. 2021, 17, 2848–2893, doi:10.3762/bjoc.17.196

Graphical Abstract
  • isotope studies revealed the cleavage of the C(sp3)–H bond may be involved in the rate-determining step of this transformation. Mechanistically, prototypical homolysis of the peroxide in the presence of the Fe(II) catalyst will generate the alkyl radical 78 formed via hydrogen abstraction. The
  • dihydrofurans [102]. In 2019, the Li group studied the selective activation of the α-C(sp3)–H of ketones and esters 103 for the tandem addition/cyclization of o-vinylanilides 102 (Scheme 20) [103]. Through a series of mechanistic experiments, it was noted the cleavage of the C(sp3)–H bond may be involved in the
PDF
Album
Review
Published 07 Dec 2021

Direct C(sp3)–H allylation of 2-alkylpyridines with Morita–Baylis–Hillman carbonates via a tandem nucleophilic substitution/aza-Cope rearrangement

  • Siyu Wang,
  • Lianyou Zheng,
  • Shutao Wang,
  • Shulin Ning,
  • Zhuoqi Zhang and
  • Jinbao Xiang

Beilstein J. Org. Chem. 2021, 17, 2505–2510, doi:10.3762/bjoc.17.167

Graphical Abstract
  • pyridylic C(sp3)–H bond (Scheme 1b). For examples, Tunge et al. developed a Pd-catalyzed intramolecular decarboxylative coupling of heterocyclic ally esters via a tandem allylation/Cope rearrangement strategy [18]; Hartwig and co-workers reported a stereo-divergent allylic substitution with azaarene
PDF
Album
Supp Info
Letter
Published 01 Oct 2021

Photoredox catalysis in nickel-catalyzed C–H functionalization

  • Lusina Mantry,
  • Rajaram Maayuri,
  • Vikash Kumar and
  • Parthasarathy Gandeepan

Beilstein J. Org. Chem. 2021, 17, 2209–2259, doi:10.3762/bjoc.17.143

Graphical Abstract
  • yields were observed for carboxylic acid substrates 81 with different steric properties. Similarly, amine substrates 80 with diverse substitution patterns and functional groups were well tolerated to provide the desired products in optimal yields. The proposed mechanism involves the cleavage of the C(sp3
  • )–H bond by a photo-generated bromine radical to give the carbon-centered alkyl radical, which subsequently engages in the nickel-catalyzed enantioselective acylation. Amides were also found to be competent acyl surrogates in the photoredox nickel-catalyzed direct C(sp3)–H acylation reactions as
PDF
Album
Review
Published 31 Aug 2021

On the application of 3d metals for C–H activation toward bioactive compounds: The key step for the synthesis of silver bullets

  • Renato L. Carvalho,
  • Amanda S. de Miranda,
  • Mateus P. Nunes,
  • Roberto S. Gomes,
  • Guilherme A. M. Jardim and
  • Eufrânio N. da Silva Júnior

Beilstein J. Org. Chem. 2021, 17, 1849–1938, doi:10.3762/bjoc.17.126

Graphical Abstract
  • applied in some very sophisticated protocols including several examples reported by the White group. In 2015, White and co-workers reported a new catalytic method using manganese tert-butylphthalocyanine [Mn(t-BuPc)Cl] for the chemoselective intramolecular amination of various C(sp3)–H bond types like
  • ′-bipyrrolidine) controls the site- and chemoselectivity in the hydroxylating step of the methylene C(sp3)–H bond while milder oxidation conditions help to increase the chemoselectivity. The methylation step is accomplished using a modestly nucleophilic organoaluminium reagent (AlMe3) to activate the hemiaminal
  • site-selective C(sp3)−H bond functionalization strategy (Scheme 28B) [160]. Starting from the abundant feedstock chemical cedrol, oxidation of the gem-dimethyl group was achieved on a gram scale, with the formation of a strained tetrahydrofuran ring. The latter was methylated and eliminated via the
PDF
Album
Review
Published 30 Jul 2021

Methodologies for the synthesis of quaternary carbon centers via hydroalkylation of unactivated olefins: twenty years of advances

  • Thiago S. Silva and
  • Fernando Coelho

Beilstein J. Org. Chem. 2021, 17, 1565–1590, doi:10.3762/bjoc.17.112

Graphical Abstract
  • methylcycloalkanes (94a and 94b) and the terminal olefin 93 (Scheme 37A). Both examples showed perfect regioselectivity for the functionalization of the tertiary C(sp3)–H bond. Mechanistic studies carried out by the authors revealed a strong kinetic isotope effect (KIE = 11.5:1) when a competitive reaction was
PDF
Album
Review
Published 07 Jul 2021

Heterogeneous photocatalytic cyanomethylarylation of alkenes with acetonitrile: synthesis of diverse nitrogenous heterocyclic compounds

  • Guanglong Pan,
  • Qian Yang,
  • Wentao Wang,
  • Yurong Tang and
  • Yunfei Cai

Beilstein J. Org. Chem. 2021, 17, 1171–1180, doi:10.3762/bjoc.17.89

Graphical Abstract
  • (Scheme 7d). These results indicate a large primary isotope effect, which suggest that the C(sp3)–H bond cleavage of acetonitrile contributed to the rate-determining step. Based on the present results and the literature [18][19][20][21][22][23][24][25][26][31][60], the reaction pathway is proposed as
PDF
Album
Supp Info
Full Research Paper
Published 17 May 2021

Manganese/bipyridine-catalyzed non-directed C(sp3)–H bromination using NBS and TMSN3

  • Kumar Sneh,
  • Takeru Torigoe and
  • Yoichiro Kuninobu

Beilstein J. Org. Chem. 2021, 17, 885–890, doi:10.3762/bjoc.17.74

Graphical Abstract
  • ]. The regioselectivity is controlled by the formation of a six-membered cyclic intermediate. Directing-group-assisted C(sp3)−H halogenation reactions are efficient for promoting regioselective C(sp3)−H halogenations (Scheme 1c) [24][25][26][27][28]. In these reactions, the C(sp3)–H bond at the β
  • reaction proceeded regioselectively at the methine C(sp3)–H bond of isoamyl benzoate (1b) to give 2b in 64% yield. Isoamyl benzoates bearing halogen atoms, such as fluorine, chlorine, or bromine, on the phenyl ring were also suitable substrates and gave C(sp3)–H brominated products 2c–e in 49–60% yields
  • , without any loss of the halogen substituents. Although the C(sp3)–H bromination of isobutyl benzoate 1f did not proceed at 60 °C, the corresponding C(sp3)–H brominated compound 2f was produced at higher temperature (80 °C). The C(sp3)–H bond in acetal 1g was efficiently brominated to give 2g in 79% yield
PDF
Album
Supp Info
Letter
Published 22 Apr 2021

An atom-economical addition of methyl azaarenes with aromatic aldehydes via benzylic C(sp3)–H bond functionalization under solvent- and catalyst-free conditions

  • Divya Rohini Yennamaneni,
  • Vasu Amrutham,
  • Krishna Sai Gajula,
  • Rammurthy Banothu,
  • Murali Boosa and
  • Narender Nama

Beilstein J. Org. Chem. 2020, 16, 3093–3103, doi:10.3762/bjoc.16.259

Graphical Abstract
  • , Yaragorla et al. published a review on C(sp3)–H bond functionalization of 2-methylazaarenes [39]. These strategies are proficient, but due to the involvement of drastic reaction conditions, the use of expensive reagents, toxic metals, harmful solvents, and tedious workup procedures, they need to be
PDF
Album
Supp Info
Letter
Published 23 Dec 2020

Photosensitized direct C–H fluorination and trifluoromethylation in organic synthesis

  • Shahboz Yakubov and
  • Joshua P. Barham

Beilstein J. Org. Chem. 2020, 16, 2151–2192, doi:10.3762/bjoc.16.183

Graphical Abstract
  • ][51][52][53][54][55][56][57]. Accordingly, the reactivity and selectivity correlate with the homolytic C–H bond dissociation enthalpy and polarity matching principles [58][59][60][61][62][63][64]. One of the advantages of polarity matching is that it enables C(sp3)–H bond activations that are not
  • ”hydridic” C(sp3)–H bond that forms a secondary radical is predominantly fluorinated. This indicates that these reactions operate by HAT, which is directed by polarity matching effects [58][59][60]. For example, n-hexane is mainly fluorinated at its C2 position (67%) and only 4% fluorination at C1 occurs
  • [201], due to the instability of the primary radical (Figure 10). The selective fluorination of C2 over C3 may be rationalized by steric effects (rather than polarity matching), which are reported for quinuclidinium radical cations [203]. A similar selectivity for the most hydridic C(sp3)–H bond was
PDF
Album
Review
Published 03 Sep 2020

When metal-catalyzed C–H functionalization meets visible-light photocatalysis

  • Lucas Guillemard and
  • Joanna Wencel-Delord

Beilstein J. Org. Chem. 2020, 16, 1754–1804, doi:10.3762/bjoc.16.147

Graphical Abstract
  • features when working with biological systems. Thus, a variety of functionalized purines nucleosides that are potentially of great importance in medicinal chemistry were obtained in good yields. In contrast to C(sp2)–H bonds, C(sp3)–H bond activation is much more challenging. Inspired by the potential of
  • C(sp3)–H bond activation reactions, was effective at room temperature. The coupling displayed high selectivity for β-methyl C(sp3)–H bonds. Of note is that a substitution at the C5-position of the 8-aminoquinoline ring was beneficial for the C–H activation step as the functionalization of 5-chloro-8
PDF
Album
Review
Published 21 Jul 2020

Aldehydes as powerful initiators for photochemical transformations

  • Maria A. Theodoropoulou,
  • Nikolaos F. Nikitas and
  • Christoforos G. Kokotos

Beilstein J. Org. Chem. 2020, 16, 833–857, doi:10.3762/bjoc.16.76

Graphical Abstract
  • desired products 140n–t in moderate to good yields (Scheme 30). A few months later, the same research group extended this methodology to an α-C(sp3)–H bond functionalization of nitrogen-containing molecules and thioethers [58]. In order to achieve the highest yield, an optimum concentration of acetone (4
PDF
Album
Review
Published 23 Apr 2020

Recent advances on the transition-metal-catalyzed synthesis of imidazopyridines: an updated coverage

  • Gagandeep Kour Reen,
  • Ashok Kumar and
  • Pratibha Sharma

Beilstein J. Org. Chem. 2019, 15, 1612–1704, doi:10.3762/bjoc.15.165

Graphical Abstract
  • 2-benzoylpyridine (145) and different benzylamines was carried out to synthesize 1,3-diarylated imidazo[1,5-a]pyridines (Scheme 50). The reaction took place under aerobic conditions utilizing Cu-MOF-74 as a catalyst [130]. The reaction was unprecedented in terms of oxidative amination of C(sp3)–H
  • bond catalyzed by Cu-MOF-74. The group of Nguyen have deeply investigated the reaction and found 10 mol % loading of the catalyst with 3 equiv of benzylamine and 0.2 M 2-benzoylpyridine to be optimal for appreciable yield (Scheme 51). The reaction was optimized with different salts and Cu(OAc)2 was
PDF
Album
Review
Published 19 Jul 2019

Cobalt- and rhodium-catalyzed carboxylation using carbon dioxide as the C1 source

  • Tetsuaki Fujihara and
  • Yasushi Tsuji

Beilstein J. Org. Chem. 2018, 14, 2435–2460, doi:10.3762/bjoc.14.221

Graphical Abstract
  • formation of products 10c-Me and 10d-Me. With regard to other amide groups, morpholides 9g–i could be used and benzylmethylamide- and diethylamide-bearing substrates, which afforded the corresponding products 10j-Me and 10k-Me, albeit with moderate yields. Allylic C(sp3)–H bond carboxylation The development
PDF
Album
Review
Published 19 Sep 2018

Cobalt-catalyzed nucleophilic addition of the allylic C(sp3)–H bond of simple alkenes to ketones

  • Tsuyoshi Mita,
  • Masashi Uchiyama,
  • Kenichi Michigami and
  • Yoshihiro Sato

Beilstein J. Org. Chem. 2018, 14, 2012–2017, doi:10.3762/bjoc.14.176

Graphical Abstract
  • , affording branched homoallylic alcohols in high yields with perfect branch selectivities. The intermediate of the reaction would be a nucleophilic allylcobalt(I) species generated via cleavage of the low reactive allylic C(sp3)–H bond of simple terminal alkenes. Keywords: alkenes; C–H activation; C(sp3)–H
  • involving catalytic C–C bond construction with the double bond of terminal alkenes (e.g., Heck reaction, hydrometalation followed by functionalization, carbometalation, and olefin metathesis) [10][11][12][13]. However, direct C–C bond formation of the allylic C(sp3)–H bond adjacent to double bonds has
  • 1-undecene (1a) but also 1-octadecene (1b) and 6-phenyl-1-hexene (1c) were tolerable to afford the corresponding products in around 70% yield with perfect branch selectivity (Figure 3). Although the allylic C(sp3)–H bond of α-olefins is weakly acidic (pKa value of 1-propene = 43), it is noteworthy
PDF
Album
Supp Info
Letter
Published 02 Aug 2018

Hypervalent organoiodine compounds: from reagents to valuable building blocks in synthesis

  • Gwendal Grelier,
  • Benjamin Darses and
  • Philippe Dauban

Beilstein J. Org. Chem. 2018, 14, 1508–1528, doi:10.3762/bjoc.14.128

Graphical Abstract
  • carbamates [92], sulfamates [94][95][96][97], ureas and guanidines [98], sulfamides [99], hydroxylamine-derived sulfamates [100], carbamimidates [101], and sulfonimidamides [102][103][104][105][106][107]. These reactions involve the formation of a metal-bound nitrene that can insert into a C(sp3)–H bond or a
PDF
Album
Review
Published 21 Jun 2018

Atom-economical group-transfer reactions with hypervalent iodine compounds

  • Andreas Boelke,
  • Peter Finkbeiner and
  • Boris J. Nachtsheim

Beilstein J. Org. Chem. 2018, 14, 1263–1280, doi:10.3762/bjoc.14.108

Graphical Abstract
  • –Hagihara reaction was developed by Dauban and co-workers (Scheme 16) [47]. The first step of this sequence includes an iodine(III)-mediated rhodium-catalysed enantioselective amination of an unactivated C(sp3)–H bond with a chiral sulfonimidamide 31. Afterwards, the iodoarene byproduct of the first step is
  • the sulfonimidamide 31 in the presence of the chiral rhodium(II) catalyst A. Hereby, one equivalent of iodoarene 2 is released. The insertion of the nitrene species into the C(sp3)–H bond affords the amination product C, which is the final product of the first reaction step. In the following step, C
PDF
Album
Review
Published 30 May 2018

Selective carboxylation of reactive benzylic C–H bonds by a hypervalent iodine(III)/inorganic bromide oxidation system

  • Toshifumi Dohi,
  • Shohei Ueda,
  • Kosuke Iwasaki,
  • Yusuke Tsunoda,
  • Koji Morimoto and
  • Yasuyuki Kita

Beilstein J. Org. Chem. 2018, 14, 1087–1094, doi:10.3762/bjoc.14.94

Graphical Abstract
  • ; C–H activation; iodine; oxygenation; radicals; Introduction The oxidative activation of a C(sp3)–H bond in organic molecules to directly install various functional groups and new carbon–carbon networks is a topic of interest for researchers engaged in modern synthetic chemistry [1][2][3][4][5][6][7
  • single-electron-transfer (SET) reactivities [33][34][35][36][37] allow selective activation of the benzylic C(sp3)–H bond for oxidative functionalization and coupling reactions. Initially, the SET oxidation ability of pentavalent iodine reagents, especially o-iodoxybenzoic acid (IBX), in benzylic
  • facilitate the activation of the benzylic C(sp3)–H bond are rare [21][57][58][59][60][61][62]. Furthermore, only a limited number of transition-metal-free methods have been reported; successful examples include the Wohl–Ziegler-type conditions [21], the sodium bromate system [57] for the conversion of
PDF
Album
Supp Info
Letter
Published 16 May 2018

Preparation, structure, and reactivity of bicyclic benziodazole: a new hypervalent iodine heterocycle

  • Akira Yoshimura,
  • Michael T. Shea,
  • Cody L. Makitalo,
  • Melissa E. Jarvi,
  • Gregory T. Rohde,
  • Akio Saito,
  • Mekhman S. Yusubov and
  • Viktor V. Zhdankin

Beilstein J. Org. Chem. 2018, 14, 1016–1020, doi:10.3762/bjoc.14.87

Graphical Abstract
  • example, azidobenziodazole was used as an efficient azidation reagent with a reactivity similar to azidobenziodoxoles [33]. Recently, the Wang group reported a rhenium catalyst-mediated oxidative dehydrogenative olefination of a C(sp3)–H bond using acetoxybenziodazole reagents [36]. To the best of our
PDF
Album
Supp Info
Full Research Paper
Published 08 May 2018

Progress in copper-catalyzed trifluoromethylation

  • Guan-bao Li,
  • Chao Zhang,
  • Chun Song and
  • Yu-dao Ma

Beilstein J. Org. Chem. 2018, 14, 155–181, doi:10.3762/bjoc.14.11

Graphical Abstract
  • allylic CF3 bonds from olefins. Following examples described the straightforward trifluoromethylation of terminal alkenes via allylic C(sp3)–H bond activation generating allylic trifluoromethylated compounds. In 2011, the group of Fu and Liu [49] described an unprecedented type of a Cu-catalyzed
PDF
Album
Review
Published 17 Jan 2018

Oxidative dehydrogenation of C–C and C–N bonds: A convenient approach to access diverse (dihydro)heteroaromatic compounds

  • Santanu Hati,
  • Ulrike Holzgrabe and
  • Subhabrata Sen

Beilstein J. Org. Chem. 2017, 13, 1670–1692, doi:10.3762/bjoc.13.162

Graphical Abstract
  • reacted with o-aminobenzylamine and substituted o-aminobenzylamines to provide the desired products in decent yields (Scheme 8b). Kumar et al. demonstrated transition metal-free α-C(sp3)–H bond functionalization of amines via an oxidative cross-dehydrogenative coupling reaction [44]. They reported a one
PDF
Album
Review
Published 15 Aug 2017

Synthesis of β-arylated alkylamides via Pd-catalyzed one-pot installation of a directing group and C(sp3)–H arylation

  • Yunyun Liu,
  • Yi Zhang,
  • Xiaoji Cao and
  • Jie-Ping Wan

Beilstein J. Org. Chem. 2016, 12, 1122–1126, doi:10.3762/bjoc.12.108

Graphical Abstract
  • functionalization chemistry [35][36][37][38][39][40], we have executed efforts to the AQ-assisted β-C–H functionalization reactions of alkylamides via activation of the C(sp3)–H bond, a classical protocol toward β-arylamide synthesis. While the known examples, including C(sp3)–H alkylation, arylation or oxygenation
PDF
Album
Supp Info
Full Research Paper
Published 03 Jun 2016

Enantioselective carbenoid insertion into C(sp3)–H bonds

  • J. V. Santiago and
  • A. H. L. Machado

Beilstein J. Org. Chem. 2016, 12, 882–902, doi:10.3762/bjoc.12.87

Graphical Abstract
  • modulation of its reactivity, and controls the chemo-, regio- and stereoselectivity in reactions. The activation of the C(sp3)–H bond needs an appropriate interaction between the carbenoid intermediate and the carbon atom of the C(sp3)–H. Depending on the electronic demand of the substituent attached to the
  • -cuparenone (8) through the construction of a five-membered ring prepared by an enantioselective carbenoid insertion into a C(sp3)–H bond (Scheme 3) [34]. To carry out the cyclization, the carbenoid was formed by the action of Rh2(OAc)4 on the diazo compound 6. That intermediate intramolecularly inserted into
  • the C(sp3)–H bond of the asymmetric carbon to yield ketoester 7 in 67% yield. This latter compound was converted to (+)-α-cuparenone (8) in 26% yield and 96% enantiomeric excess. In the late 1980s, many studies have been published by Taber [35], Sonawane [36], Doyle [37] and their respective coworkers
PDF
Album
Review
Published 04 May 2016
Other Beilstein-Institut Open Science Activities