Search results

Search for "palladium catalyst" in Full Text gives 82 result(s) in Beilstein Journal of Organic Chemistry.

Recent advances in hypervalent iodine(III)-catalyzed functionalization of alkenes

  • Xiang Li,
  • Pinhong Chen and
  • Guosheng Liu

Beilstein J. Org. Chem. 2018, 14, 1813–1825, doi:10.3762/bjoc.14.154

Graphical Abstract
  • intermediate 75. Subsequently, this intermediate is attacked by the palladium catalyst under a CO atmosphere to form the alkyl palladium species 76. Finally, the reductive elimination at the iodine(III) center and CO insertion into the newly formed C–Pd bond, affords the oxycarbonylation products 74 (Scheme 21
PDF
Album
Review
Published 18 Jul 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
  • –I fails to furnish the N-arylation products under the same reaction conditions, leading the authors to conclude that this suggested intermediate remains in the coordination sphere of the palladium catalyst. In 2015, the group of Greaney has described a tandem copper-catalyzed C3–H/N–H arylation of
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
  • was a reaction of N-arylimines 38 with ethynylbenziodoxolones 36a (EBX), under the influence of a palladium catalyst, developed by Yoshikai and co-workers [55]. Instead of the expected α-alkynylated product, highly substituted furans 39 were observed (Scheme 21). Both, the electrophilic alkyne and the
PDF
Album
Review
Published 30 May 2018

Nanoreactors for green catalysis

  • M. Teresa De Martino,
  • Loai K. E. A. Abdelmohsen,
  • Floris P. J. T. Rutjes and
  • Jan C. M. van Hest

Beilstein J. Org. Chem. 2018, 14, 716–733, doi:10.3762/bjoc.14.61

Graphical Abstract
  • nanoparticles (DSN) have been shown to be highly efficient in the catalysis of olefin hydrogenation and in Suzuki coupling reactions [98][99]. Ornelas et al. entrapped a palladium catalyst with dendrimers containing triazole groups (DSN) (Figure 6) [100]. The aim here was to provide a platform to perform
PDF
Album
Review
Published 29 Mar 2018

Reactivity of bromoselenophenes in palladium-catalyzed direct arylations

  • Aymen Skhiri,
  • Ridha Ben Salem,
  • Jean-François Soulé and
  • Henri Doucet

Beilstein J. Org. Chem. 2017, 13, 2862–2868, doi:10.3762/bjoc.13.278

Graphical Abstract
  • heteroarylations with a variety of heteroarenes using a phosphine-free palladium catalyst. Results and Discussion First, we examined the influence of the reaction temperature, using DMA as solvent, KOAc as base and 2 mol % Pd(OAc)2 as catalyst (Table 1). We had previously observed that these reaction conditions
PDF
Album
Supp Info
Full Research Paper
Published 22 Dec 2017

Total synthesis of TMG-chitotriomycin based on an automated electrochemical assembly of a disaccharide building block

  • Yuta Isoda,
  • Norihiko Sasaki,
  • Kei Kitamura,
  • Shuji Takahashi,
  • Sujit Manmode,
  • Naoko Takeda-Okuda,
  • Jun-ichi Tamura,
  • Toshiki Nokami and
  • Toshiyuki Itoh

Beilstein J. Org. Chem. 2017, 13, 919–924, doi:10.3762/bjoc.13.93

Graphical Abstract
  • -diisopropylethylamine (iPr2NEt) to prepare the TMG part of tetrasaccharide 9. Deprotection of acetyl groups at the 3-O-positions and the subsequent global deprotection of the benzyl groups and the anomeric thioaryl group of the tetrasaccharide by hydrogenation with hydrogen gas in the presence of a palladium catalyst
PDF
Album
Supp Info
Full Research Paper
Published 16 May 2017

Molecular-level architectural design using benzothiadiazole-based polymers for photovoltaic applications

  • Vinila N. Viswanathan,
  • Arun D. Rao,
  • Upendra K. Pandey,
  • Arul Varman Kesavan and
  • Praveen C. Ramamurthy

Beilstein J. Org. Chem. 2017, 13, 863–873, doi:10.3762/bjoc.13.87

Graphical Abstract
  • compound was then coupled with trimethyl(thiophene-2-yl)stannane through a Stille reaction using tris(dibenzylideneacetone)dipalladium(0) and tri-o-tolylphosphine as the catalyst system. The dithiophenylated product 3 was washed several times with methanol to remove the palladium catalyst and other
PDF
Album
Supp Info
Full Research Paper
Published 10 May 2017

Contribution of microreactor technology and flow chemistry to the development of green and sustainable synthesis

  • Flavio Fanelli,
  • Giovanna Parisi,
  • Leonardo Degennaro and
  • Renzo Luisi

Beilstein J. Org. Chem. 2017, 13, 520–542, doi:10.3762/bjoc.13.51

Graphical Abstract
  • the desired products 5a–j (Scheme 9). Noteworthy, the homogeneous catalysis requires only 1% of the XPhos-based palladium catalyst. A sonication bath was employed to prevent clogging and the reaction required a residence time of 15 min. Next, they turned their attention to the arylation of fluoro
PDF
Album
Review
Published 14 Mar 2017

Synthesis of 1-indanones with a broad range of biological activity

  • Marika Turek,
  • Dorota Szczęsna,
  • Marek Koprowski and
  • Piotr Bałczewski

Beilstein J. Org. Chem. 2017, 13, 451–494, doi:10.3762/bjoc.13.48

Graphical Abstract
  • ketoesters 59. The base-mediated cyclization of the latter in the presence of sodium ethoxide led to the formation of the corresponding 1-indanone anions α to carbonyl, which next were alkylated to give 2-substituted 1-indanones 60 (Scheme 20). This one-pot process utilizing a multi-task palladium catalyst
  • 90, it has been proved that this reaction proceeds with participation of Ni-complex 91 isolated in 83% yield which next was converted to 1-indanone 90a via the monomeric complex 92 or its dimer. o-Bromobenzaldehyde 93, in the presence of a palladium catalyst, underwent intermolecular carbopalladation
  • of palladium catalyst and gaseous oxygen as the terminal oxidant (Scheme 65). 1.8 From other compounds In 2016, Shi et al. have developed an unique, conditions-controlled [Rh2(esp)2] (esp = α,α,α’,α’-tetramethyl-1,3-benzenedipropionic acid)-catalyzed reaction of N-sulfonyl-1,2,3-triazoles 235 leading
PDF
Album
Review
Published 09 Mar 2017

Decarboxylative and dehydrative coupling of dienoic acids and pentadienyl alcohols to form 1,3,6,8-tetraenes

  • Ghina’a I. Abu Deiab,
  • Mohammed H. Al-Huniti,
  • I. F. Dempsey Hyatt,
  • Emma E. Nagy,
  • Kristen E. Gettys,
  • Sommayah S. Sayed,
  • Christine M. Joliat,
  • Paige E. Daniel,
  • Rupa M. Vummalaneni,
  • Andrew T. Morehead Jr,
  • Andrew L. Sargent and
  • Mitchell P. Croatt

Beilstein J. Org. Chem. 2017, 13, 384–392, doi:10.3762/bjoc.13.41

Graphical Abstract
  • ). As a control reaction, it was found that no reaction occurred in the absence of palladium catalyst (Table 1, entry 21). As shown earlier, bis-allylic sorbate 13 (Scheme 2) was found to be low yielding for the decarboxylative coupling reaction. Reactions of sorbate 13 monitored by 1H NMR showed nearly
  • products, it is hypothesized that the product may be sequestering the palladium catalyst. Two cyclic dienyl acetates were also studied (Table 3, entries 5 and 6) and they yielded tetraenes 8f and 8g. The dienes of entries 5 and 6 could have formed additional isomers by coupling to the other end of the
  • 1,3,6,8-tetraene. It is proposed for pathway B that the dienoate is required so that the α-carbon is not blocked by the bulky phosphine group since it can add in a 1,6- or 1,4-manner, both reversibly. Alternatively, pathway A has the palladium catalyst coordinate to one of the alkenes of the dienoate
PDF
Album
Supp Info
Full Research Paper
Published 28 Feb 2017

Highly bulky and stable geometry-constrained iminopyridines: Synthesis, structure and application in Pd-catalyzed Suzuki coupling of aryl chlorides

  • Yi Lai,
  • Zhijian Zong,
  • Yujie Tang,
  • Weimin Mo,
  • Nan Sun,
  • Baoxiang Hu,
  • Zhenlu Shen,
  • Liqun Jin,
  • Wen-hua Sun and
  • Xinquan Hu

Beilstein J. Org. Chem. 2017, 13, 213–221, doi:10.3762/bjoc.13.24

Graphical Abstract
  • under high temperature, if the palladium catalyst is stable enough. The formation of palladium black was observed in reactions of aryl chlorides using the reported palladium catalyst system, which inspired us to improve the thermal stability of the palladium complex. The steric environment adjacent to
PDF
Album
Supp Info
Full Research Paper
Published 03 Feb 2017

Diastereoselective anodic hetero- and homo-coupling of menthol-, 8-methylmenthol- and 8-phenylmenthol-2-alkylmalonates

  • Matthias C. Letzel,
  • Hans J. Schäfer and
  • Roland Fröhlich

Beilstein J. Org. Chem. 2017, 13, 33–42, doi:10.3762/bjoc.13.5

Graphical Abstract
  • for the hydrogenation of the benzyl esters: To a solution of 6.4 mmol benzyl ester in 100 mL of methanol 800 mg of the palladium catalyst (10% Pd on carbon) were added. Hydrogenation was performed at atmospheric pressure until no further hydrogen was consumed. After filtration through celite the
PDF
Album
Supp Info
Full Research Paper
Published 05 Jan 2017

Et3B-mediated and palladium-catalyzed direct allylation of β-dicarbonyl compounds with Morita–Baylis–Hillman alcohols

  • Ahlem Abidi,
  • Yosra Oueslati and
  • Farhat Rezgui

Beilstein J. Org. Chem. 2016, 12, 2402–2409, doi:10.3762/bjoc.12.234

Graphical Abstract
  • -enone (1a) and acyclic ethyl 2-(hydroxymethyl)acrylate (1b), under the action of 1,3-dicarbonyl compounds 2, in the presence of an appropriate palladium catalyst and Et3B as a Lewis acid promoter, into the allylation compounds 3–8 with the formation of only water as a byproduct. These derivatives can be
  • products 7e–i [41][44] in 60–70% yields (Table 5, entries 1–4). Conclusion In summary, we have developed a mild and direct process for the C–C bond formation from the reaction of the MBH alcohols 1a and 1b with β-dicarbonyl compounds 2 in the presence of a palladium catalyst and Et3B (promoter) with the
PDF
Album
Supp Info
Full Research Paper
Published 15 Nov 2016

Palladium-catalyzed ring-opening reactions of cyclopropanated 7-oxabenzonorbornadiene with alcohols

  • Katrina Tait,
  • Oday Alrifai,
  • Rebecca Boutin,
  • Jamie Haner and
  • William Tam

Beilstein J. Org. Chem. 2016, 12, 2189–2196, doi:10.3762/bjoc.12.209

Graphical Abstract
  • nucleophile at the internal cyclopropane carbon C, which could induce ring expansion to form seven-membered ring 12 . In this paper, we aim to explore the use of a palladium catalyst with an alcohol nucleophile on the ring opening of cyclopropanated oxabenzonorbornadiene with the goal of determining which
  • (Table 1, entry 13) along with the use of a platinum(II) catalyst (Table 1, entry 14). Using an anionic platinum(II) catalyst yielded substituted naphthalene 11a in a 22% yield, though this was considerably lower when compared to the optimized palladium catalyst. A variety of solvents were next screened
PDF
Album
Supp Info
Full Research Paper
Published 14 Oct 2016

Rapid regio- and multi-coupling reactivity of 2,3-dibromobenzofurans with atom-economic triarylbismuths under palladium catalysis

  • Maddali L. N. Rao,
  • Jalindar B. Talode and
  • Venneti N. Murty

Beilstein J. Org. Chem. 2016, 12, 2065–2076, doi:10.3762/bjoc.12.195

Graphical Abstract
  • -dibromobenzofuran (1.1) and 1 equiv of bismuth reagent gave 86% yield (Table 1, entry 10). A few control reactions without base or palladium catalyst showed inferior or no cross-coupling reactivity (Table 1, entries 11 and 12). This investigation results that the desired regio-selective cross-coupling reactivity
PDF
Album
Supp Info
Full Research Paper
Published 22 Sep 2016

Ionic liquids as transesterification catalysts: applications for the synthesis of linear and cyclic organic carbonates

  • Maurizio Selva,
  • Alvise Perosa,
  • Sandro Guidi and
  • Lisa Cattelan

Beilstein J. Org. Chem. 2016, 12, 1911–1924, doi:10.3762/bjoc.12.181

Graphical Abstract
  • of oxygen and a palladium catalyst (Scheme 6, middle). Though safer than the phosgenation of methanol, these synthetic routes still involved poisonous carbon monoxide and methyl nitrite, and chlorine-based catalysts. Carbon dioxide is the natural green alternative carbonyl source to these undesirable
PDF
Album
Review
Published 26 Aug 2016

Cationic Pd(II)-catalyzed C–H activation/cross-coupling reactions at room temperature: synthetic and mechanistic studies

  • Takashi Nishikata,
  • Alexander R. Abela,
  • Shenlin Huang and
  • Bruce H. Lipshutz

Beilstein J. Org. Chem. 2016, 12, 1040–1064, doi:10.3762/bjoc.12.99

Graphical Abstract
  • palladacycle; (2) reaction of the cationic palladacycle with an aryl iodide, arylboronic acid or acrylate, and (3) regeneration of the active cationic palladium catalyst. The reaction between a cationic palladium(II) complex and arylurea allowed the formation and isolation of the corresponding palladacycle
  • ”), there was a noticeable drop in the extent of conversion. In addition, lower loadings of HBF4, silver salt, or the palladium catalyst also gave inferior results. In the case of C–H activation/Suzuki–Miyaura coupling reactions, the commercially available, pre-formed cationic Pd(II) catalyst [Pd(MeCN)4
  • reactions to be run in water at room temperature using the cationic palladium catalyst [Pd(MeCN)4](BF4)2 (Figure 4, 5a–c, conditions A). While this reaction proceeded with a number of alkyl anilide derivatives, as well as ureas as directing groups (5c), the substrate scope was otherwise somewhat limited
PDF
Album
Supp Info
Full Research Paper
Published 20 May 2016

Catalytic asymmetric synthesis of biologically important 3-hydroxyoxindoles: an update

  • Bin Yu,
  • Hui Xing,
  • De-Quan Yu and
  • Hong-Min Liu

Beilstein J. Org. Chem. 2016, 12, 1000–1039, doi:10.3762/bjoc.12.98

Graphical Abstract
  • metal-catalyzed synthesis The chiral ligand/metal complexes have been widely employed in catalytic asymmetric synthesis of enantioenriched 3-hydroxyoxindoles, achieving good to excellent enantioselectivities and high yields. Pd-catalyzed allylation of isatins The palladium catalyst is widely used in
PDF
Album
Review
Published 18 May 2016

A convenient route to symmetrically and unsymmetrically substituted 3,5-diaryl-2,4,6-trimethylpyridines via Suzuki–Miyaura cross-coupling reaction

  • Dariusz Błachut,
  • Joanna Szawkało and
  • Zbigniew Czarnocki

Beilstein J. Org. Chem. 2016, 12, 835–845, doi:10.3762/bjoc.12.82

Graphical Abstract
  • unsymmetrically substituted diarylpyridines, difficult to access by other methods. Keywords: arylpyridines; cross coupling reaction; heteroaromatics; one-pot reaction; palladium catalyst; Introduction Nitrogen heterocycles are an important class of compounds widely present in agrochemical products [1][2] and
PDF
Album
Supp Info
Full Research Paper
Published 28 Apr 2016

Regioselective palladium-catalyzed ring-opening reactions of C1-substituted oxabicyclo[2,2,1]hepta-2,5-diene-2,3-dicarboxylates

  • Michael Edmunds,
  • Mohammed Abdul Raheem,
  • Rebecca Boutin,
  • Katrina Tait and
  • William Tam

Beilstein J. Org. Chem. 2016, 12, 239–244, doi:10.3762/bjoc.12.25

Graphical Abstract
  • 1 undergoing a ring-opening reaction, which used p-iodotoluene and a palladium catalyst (Scheme 2) [14]. The result was the addition of the aryl group to the unsubstituted double bond followed by dehydration to give an unsymmetrical biphenyl derivative. However, there have not been any
  • oxabenzonorbornadiene with the aryl group being added to the olefin carbon furthest from the C1 substituent followed by dehydration, producing a novel, highly-substituted, biphenyl derivative. Results and Discussion The optimization study focused on the palladium catalyst, Lewis acid additive and solvent, as the
  • yield of the reaction (48%, 17 h, Table 1, entry 2). The palladium catalyst, Pd(PPh3)4, gave a moderate yield (58%, 19 h, Table 1, entry 3), while the palladium(II) catalyst, PdCl2(PPh3)2, gave the best yield of the ring opened product and also reacted the fastest (88%, 16 h, Table 1, entry 4). A
PDF
Album
Supp Info
Full Research Paper
Published 09 Feb 2016

Hydroquinone–pyrrole dyads with varied linkers

  • Hao Huang,
  • Christoffer Karlsson,
  • Maria Strømme,
  • Martin Sjödin and
  • Adolf Gogoll

Beilstein J. Org. Chem. 2016, 12, 89–96, doi:10.3762/bjoc.12.10

Graphical Abstract
  • summary, after a four step synthesis route, the total yield of 3d from 8 was 36%. Attempts were made to convert alkyne 4d into the corresponding alkene 4c. However, partial hydrogenations with a Lindlar catalyst [23] or with trimethylsilane using a palladium catalyst [24] were either unsuccessful or
PDF
Album
Supp Info
Full Research Paper
Published 18 Jan 2016

Recent developments in copper-catalyzed radical alkylations of electron-rich π-systems

  • Kirk W. Shimkin and
  • Donald A. Watson

Beilstein J. Org. Chem. 2015, 11, 2278–2288, doi:10.3762/bjoc.11.248

Graphical Abstract
  • -pot procedure, enynes could be synthesized by introducing a second alkyne and a palladium catalyst to perform a tandem carbohalogenation/Sonagashira coupling. Conclusion Copper catalysis has recently emerged as a new means of harnessing the potential of alkyl radicals in catalytic alkylation chemistry
PDF
Album
Review
Published 23 Nov 2015

Pyridinoacridine alkaloids of marine origin: NMR and MS spectral data, synthesis, biosynthesis and biological activity

  • Louis P. Sandjo,
  • Victor Kuete and
  • Maique W. Biavatti

Beilstein J. Org. Chem. 2015, 11, 1667–1699, doi:10.3762/bjoc.11.183

Graphical Abstract
  • phosporus pentachloride to prepare 4-chloro-8-methoxy-5-nitroquinoline and secondly with triflate anhydride, dimethylaminopyridine, and 2,6-lutidine to prepare 8-methoxy-5-nitro-4-triflylquinoline. The Suzuki cross-coupling was carried out on both quinoline derivatives in the presence of a palladium
  • catalyst and the appropriate organoborane to give the expected product with 47 and 78% yield for the chloride and triflate substrates, respectively. The nitro function present in the product was then reduced by palladium-catalyzed hydrogenation to yield the amine product converted into a diazonium salt. A
PDF
Album
Review
Published 18 Sep 2015

Reactions of nitroxides 15. Cinnamates bearing a nitroxyl moiety synthesized using a Mizoroki–Heck cross-coupling reaction

  • Jerzy Zakrzewski and
  • Bogumiła Huras

Beilstein J. Org. Chem. 2015, 11, 1155–1162, doi:10.3762/bjoc.11.130

Graphical Abstract
  • -tetramethylpiperidine-1-oxyl (3) with iodobenzene (4a) and 4-methyliodobenzene (4b) were used as the test reactions to check the effectiveness of various palladium catalyst systems. The use of Pd(OAc)2/Ph3P/Bu3N [37] resulted in a low yield of the target products. No products were obtained when other catalyst systems
PDF
Album
Supp Info
Full Research Paper
Published 13 Jul 2015

Radical-mediated dehydrative preparation of cyclic imides using (NH4)2S2O8–DMSO: application to the synthesis of vernakalant

  • Dnyaneshwar N. Garad,
  • Subhash D. Tanpure and
  • Santosh B. Mhaske

Beilstein J. Org. Chem. 2015, 11, 1008–1016, doi:10.3762/bjoc.11.113

Graphical Abstract
  • absence of a palladium catalyst and without the need of inert atmosphere or Schlenk tube, however, the presence of both APS and DMSO was necessary. Optimization of the protocol using various permutations and combinations provided the ideal reaction conditions for imide synthesis (Table 1, entry 9). To the
PDF
Album
Supp Info
Full Research Paper
Published 12 Jun 2015
Other Beilstein-Institut Open Science Activities