Search results

Search for "palladium" in Full Text gives 630 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

Suzuki–Miyaura cross coupling is not an informative reaction to demonstrate the performance of new solvents

  • James Sherwood

Beilstein J. Org. Chem. 2020, 16, 1001–1005, doi:10.3762/bjoc.16.89

Graphical Abstract
  • coupling is often unaffected by the choice of solvent, and therefore the Suzuki–Miyaura reaction provides limited information regarding the usefulness of any particular solvent for organic synthesis. Keywords: cross coupling; green solvents; palladium; solvent selection; Suzuki reaction; Findings The
  • , equilibria, solubility, and ultimately product yield. If there is an observable change in reaction performance correlating to one or more solvent properties (often polarity), then it is possible to identify and implement an optimum solvent. Suzuki–Miyaura cross coupling is the premier method of palladium
  •  1b) [13]. Here, palladium acetate without an auxiliary ligand was used for a pre-catalyst and the base changed to potassium carbonate. Reaction conditions of 20 hours at 65 °C were decided after observing 11% conversion after 2 hours and 33% after 20 hours in NMP at room temperature. The majority of
PDF
Album
Supp Info
Letter
Published 13 May 2020

Accelerating fragment-based library generation by coupling high-performance photoreactors with benchtop analysis

  • Quentin Lefebvre,
  • Christophe Salomé and
  • Thomas C. Fessard

Beilstein J. Org. Chem. 2020, 16, 982–988, doi:10.3762/bjoc.16.87

Graphical Abstract
  • often exhibit different reactivity from unstrained substrates [8]. N-Arylation of spirocyclic amines would be a very efficient strategy for the modular synthesis of heterocyclic sp2–sp3 fragments, but their lack of stability to strongly basic conditions and heating might be an issue in palladium
  • findings on C–N cross-coupling conditions. In these works, nickel-photoredox-catalyzed cross-couplings were the most successful with success rates up to 50% [10]. This was an improvement from non-photocatalyzed conditions, where success rates of 11–33% were observed for palladium-catalyzed cross-coupling
PDF
Album
Supp Info
Full Research Paper
Published 12 May 2020

Synthesis and properties of tetrathiafulvalenes bearing 6-aryl-1,4-dithiafulvenes

  • Aya Yoshimura,
  • Hitoshi Kimura,
  • Kohei Kagawa,
  • Mayuka Yoshioka,
  • Toshiki Itou,
  • Dhananjayan Vasu,
  • Takashi Shirahata,
  • Hideki Yorimitsu and
  • Yohji Misaki

Beilstein J. Org. Chem. 2020, 16, 974–981, doi:10.3762/bjoc.16.86

Graphical Abstract
  • Abstract Novel multistage redox tetrathiafulvalenes (TTFs) bearing 6-aryl-1,4-dithiafulvene moieties were synthesized by palladium-catalyzed direct C–H arylation. In the presence of a catalytic amount of Pd(OAc)2, P(t-Bu3)·HBF4, and an excess of Cs2CO3, the C–H arylation of TTF with several aryl bromides
  • of 1,3-dithiole rings to aromatic rings appears very appealing since these allow to produce novel multistage redox systems. However, such molecules could formerly not be synthesized by conventional approaches. In 2011, a breakthrough synthesis of arylated TTF derivatives by a palladium-catalyzed
  • -dithiole rings, which are of interest as novel multistage redox systems as well as donor components for organic conductors [1][31][32][33][34][35][36][37][38][39][40][41]. The palladium-catalyzed C–H arylation might offer access to new cross-conjugated molecules bearing vinyl-extended TTF moieties (EBDTs
PDF
Album
Supp Info
Full Research Paper
Published 12 May 2020

Pd-catalyzed asymmetric Suzuki–Miyaura coupling reactions for the synthesis of chiral biaryl compounds with a large steric substituent at the 2-position

  • Yongsu Li,
  • Bendu Pan,
  • Xuefeng He,
  • Wang Xia,
  • Yaqi Zhang,
  • Hao Liang,
  • Chitreddy V. Subba Reddy,
  • Rihui Cao and
  • Liqin Qiu

Beilstein J. Org. Chem. 2020, 16, 966–973, doi:10.3762/bjoc.16.85

Graphical Abstract
  • bromobenzene substrates and the Pd···O interaction between carbonyl and palladium seem essential to achieve high enantioselectivity. Keywords: asymmetric catalysis; biaryls; monophosphine ligand; palladium catalyst; Suzuki–Miyaura couplings; Introduction Axially chiral molecules have received much attention
  • efficient synthesis of this scaffold [7][8][9][10][11][12][13][14][15][16][17][18][19][20][21][22], like Hiyama [23][24], Negishi [25][26] or Suzuki–Miyaura couplings [27][28][29][30][31][32][33][34][35][36]. In these synthetic strategies, the reaction system of palladium with chiral phosphine ligands was
  • compounds can be obtained in excellent yields and good enantioselectivities under mild conditions, by using brominated amides and arylboronic acids as substrates, as well as palladium and chiral-bridged biphenyl monophosphine ligands as catalysts. Results and Discussion 2-Bromo-3-methyl-N-phenylbenzamide
PDF
Album
Supp Info
Full Research Paper
Published 11 May 2020

Bipyrrole boomerangs via Pd-mediated tandem cyclization–oxygenation. Controlling reaction selectivity and electronic properties

  • Liliia Moshniaha,
  • Marika Żyła-Karwowska,
  • Joanna Cybińska,
  • Piotr J. Chmielewski,
  • Ludovic Favereau and
  • Marcin Stępień

Beilstein J. Org. Chem. 2020, 16, 895–903, doi:10.3762/bjoc.16.81

Graphical Abstract
  • Rennes, CNRS, ISCR (Institut des Sciences Chimiques de Rennes) UMR 6226, F-35000 Rennes, France 10.3762/bjoc.16.81 Abstract Boomerang-shaped bipyrroles containing donor–acceptor units were obtained through a tandem palladium-mediated reaction consisting of a cyclization step, involving double C–H bond
  • (NDA) and naphthalenemonoimide (NMI) moieties. The double C–H bond activation initially used palladium(II) acetate in acetic acid as the coupling system. The subsequent screening revealed, however, that a catalytic coupling could be also achieved in the presence of silver(I) carbonate as the
  • -dipyrrolylpropanes (NDA3H, NMI3H) were reacted with palladium(II) acetate in acetic acid to furnish the expected bipyrrole dilactams in 43–53% yields (Table 1, entries 1, 6, 11 and 17). Remarkably, it was found that cNMI2O and cNMI3O were obtained along with the intermediate α-unsubstituted boomerangs, cNMI2H and
PDF
Album
Supp Info
Full Research Paper
Published 04 May 2020

One-pot synthesis of dicyclopenta-fused peropyrene via a fourfold alkyne annulation

  • Ji Ma,
  • Yubin Fu,
  • Junzhi Liu and
  • Xinliang Feng

Beilstein J. Org. Chem. 2020, 16, 791–797, doi:10.3762/bjoc.16.72

Graphical Abstract
  • Road, Hong Kong, China 10.3762/bjoc.16.72 Abstract A novel dicyclopenta-fused peropyrene derivative 1 was synthesized via a palladium-catalyzed four-fold alkyne annulation of 1,3,6,8-tetrabromo-2,7-diphenylpyrene (5) with diphenylacetylene. The annulative π-extension reaction toward 1 involved a
  • methods towards the (di-)cyclopenta-fused pyrene congeners (i–iii, Scheme 1) have mainly been reliant on the flash vacuum pyrolysis of suitable precursors under harsh conditions (T ≥ 900 °C), which resulted in relatively low yields [21][22][23][24]. Palladium-catalyzed annulation has been recently proven
  • as an efficient route to get access to aromatic hydrocarbons with peri-fused five-membered rings [25][26][27]. For instance, the dicyclopenta-fused pyrene derivatives ii and iii (Scheme 1) were successfully synthesized through palladium-catalyzed carbannulation of brominated pyrene with
PDF
Album
Supp Info
Letter
Published 20 Apr 2020

Efficient synthesis of dipeptide analogues of α-fluorinated β-aminophosphonates

  • Marcin Kaźmierczak and
  • Henryk Koroniak

Beilstein J. Org. Chem. 2020, 16, 756–762, doi:10.3762/bjoc.16.69

Graphical Abstract
  • into the salts 14 was applied. The salts 14 could be then crystallized and subjected to X-ray studies. A standard N−debenzylation protocol was employed to remove the benzyl protecting group. The hydrogenolysis reaction was catalyzed by palladium on carbon (Pd/C), and was carried out in trifluoroethanol
PDF
Album
Supp Info
Full Research Paper
Published 16 Apr 2020

Combining enyne metathesis with long-established organic transformations: a powerful strategy for the sustainable synthesis of bioactive molecules

  • Valerian Dragutan,
  • Ileana Dragutan,
  • Albert Demonceau and
  • Lionel Delaude

Beilstein J. Org. Chem. 2020, 16, 738–755, doi:10.3762/bjoc.16.68

Graphical Abstract
  • -metal-catalyzed transformations were needed to achieve the total synthesis of des-epoxy-amphidinolide N, including a palladium asymmetric allylic alkylation (Pd-AAA), a Mukaiyama aldol reaction (with Sn), and a Krische allylation (with Ir) [72]. As special feature of this procedure, an Evans aldol
  • catalyst (3 mol %) to give the corresponding 1,3-diene intermediate in 85% yield (Scheme 16). The subsequent hydroboration and oxidation to homoallylic alcohol, followed by a palladium-catalyzed Heck cross-coupling, an allylic oxidation with SeO2, mesylation, and deprotection, afforded (−)-galanthamine (13
PDF
Album
Review
Published 16 Apr 2020

Recent advances in Cu-catalyzed C(sp3)–Si and C(sp3)–B bond formation

  • Balaram S. Takale,
  • Ruchita R. Thakore,
  • Elham Etemadi-Davan and
  • Bruce H. Lipshutz

Beilstein J. Org. Chem. 2020, 16, 691–737, doi:10.3762/bjoc.16.67

Graphical Abstract
  • alcohols 98–100, respectively. Palladium was no longer needed for these transformations. The scope of the reaction with ketones was not limited to diaryl species and aryl ketones participated in the reaction, including those with more hindered alkyl groups (Scheme 19) [44]. This discovery was followed by a
PDF
Album
Review
Published 15 Apr 2020

Synthesis of C70-fragment buckybowls bearing alkoxy substituents

  • Yumi Yakiyama,
  • Shota Hishikawa and
  • Hidehiro Sakurai

Beilstein J. Org. Chem. 2020, 16, 681–690, doi:10.3762/bjoc.16.66

Graphical Abstract
  • above results strongly suggested the existence of an equilibrium between the intermediates corresponding to products 5b and 5c. A possible mechanism is shown in Scheme 2 [21][22][23][24][25]. After the oxidative addition of 4b to Pd0 to generate intermediate A, the neutral palladium(II) intermediate B
  • is formed. Two competitive processes, the reductive elimination from B to give the product 5b, and the 1,5-palladium migration from A to C through B, might exist, and from C, after the bond rotation, the intermediate D would form to afford the isomer 5c. The selectivity of these two processes are
PDF
Album
Supp Info
Full Research Paper
Published 15 Apr 2020

Synthesis of organic liquid crystals containing selectively fluorinated cyclopropanes

  • Zeguo Fang,
  • Nawaf Al-Maharik,
  • Peer Kirsch,
  • Matthias Bremer,
  • Alexandra M. Z. Slawin and
  • David O’Hagan

Beilstein J. Org. Chem. 2020, 16, 674–680, doi:10.3762/bjoc.16.65

Graphical Abstract
  • developed by Bosch [16]. Accordingly, treatment of 16a with butyl vinyl ether, bathophenanthroline (BPhen) and Et3N, using palladium(II) trifluoroacetate as a catalyst, gave vinyl ether 17 in 62% yield in a single step. Finally, difluorocarbene-mediated difluorocyclopropanation generated 10 in 50% yield
PDF
Album
Supp Info
Full Research Paper
Published 14 Apr 2020

Towards the total synthesis of chondrochloren A: synthesis of the (Z)-enamide fragment

  • Jan Geldsetzer and
  • Markus Kalesse

Beilstein J. Org. Chem. 2020, 16, 670–673, doi:10.3762/bjoc.16.64

Graphical Abstract
  • provided alcohol 11 in an excellent yield which was subjected to aminolysis to provide amide 3 in seven steps and an overall yield of 16% [16][17][18][19][20]. Synthesis of (Z)-bromide 4 For the synthesis of (Z)-bromide 4 we chose a palladium-catalyzed, stereoselective dehalogenation as the key step
  • double bonds into the corresponding (Z)-monohalogenated derivatives using palladium(II) acetate, triphenylphosphine and tributyltin hydride. Following this procedure, we were able to obtain (Z)-bromide 4 in four steps and an overall yield of 39% [21][22][23][24]. Cross coupling of fragments 3 and 4 The
  • ][19][20], whereas the (Z)-bromide 4 can be generated in a four-step sequence with a 39% overall yield, including a palladium-catalyzed, stereoselective dehalogenation [21][22][23][24][25]. Retrosynthetic analysis of chondrochlorene A (1). Synthesis of amide 3 [16][17][18][19][20]. TIPDSCl2 = 1,3
PDF
Album
Supp Info
Full Research Paper
Published 14 Apr 2020

Rhodium-catalyzed reductive carbonylation of aryl iodides to arylaldehydes with syngas

  • Zhenghui Liu,
  • Peng Wang,
  • Zhenzhong Yan,
  • Suqing Chen,
  • Dongkun Yu,
  • Xinhui Zhao and
  • Tiancheng Mu

Beilstein J. Org. Chem. 2020, 16, 645–656, doi:10.3762/bjoc.16.61

Graphical Abstract
  • industrial waste and other side products. Particularly, the reductive carbonylation of aryl iodides to produce arylaldehydes with CO and H2 was seldom reported. Some homogeneous and heterogeneous catalytic systems based on palladium species using CO and H2 to complete the reductive carbonylation of aryl
PDF
Album
Supp Info
Full Research Paper
Published 08 Apr 2020

Regioselectively α- and β-alkynylated BODIPY dyes via gold(I)-catalyzed direct C–H functionalization and their photophysical properties

  • Takahide Shimada,
  • Shigeki Mori,
  • Masatoshi Ishida and
  • Hiroyuki Furuta

Beilstein J. Org. Chem. 2020, 16, 587–595, doi:10.3762/bjoc.16.53

Graphical Abstract
  • Glaser-coupling reactions [37]. Conventionally, an alkynylation of the BODIPY core has been achieved by palladium-catalyzed Sonogashira cross-coupling with halogenated BODIPYs (Figure 1b) [35][37]. However, due to the coexistence of multiple C–H bonds, a regioselective direct C–H alkynylation of the
PDF
Album
Supp Info
Full Research Paper
Published 01 Apr 2020

A systematic review on silica-, carbon-, and magnetic materials-supported copper species as efficient heterogeneous nanocatalysts in “click” reactions

  • Pezhman Shiri and
  • Jasem Aboonajmi

Beilstein J. Org. Chem. 2020, 16, 551–586, doi:10.3762/bjoc.16.52

Graphical Abstract
  • graphene oxide (rGO) with copper and palladium species (Scheme 15) [77]. In this study, graphite oxide (GO) was generated according to the modified Hummer’s method. Copper(II) was anchored on GO via ultrasonication. In the next step, copper ions were reduced by adding NaBH4. The mixture was then heated at
  • stirred using a catalytic amount of rGO/Cu2O (102) in water at 55 °C for 2 h to obtain 1,2,3-triazole derivatives substituted at the 1- and 4 position with high yields (Scheme 23). The catalyst 102 could be readily reused in six cycles [86]. In another study, charcoal-supported palladium and copper as a
  • multitask nanocatalyst was used for one-pot Sonogashira-“click”/“click”-Heck sequences [87]. Palladium(II) acetate, copper(II) acetate, and charcoal were dispersed in methanol. In order to remove oxygen from the reaction mixture, hydrogen gas was passed through the medium. Then, the solution was stirred at
PDF
Album
Review
Published 01 Apr 2020

Copper-catalyzed remote C–H arylation of polycyclic aromatic hydrocarbons (PAHs)

  • Anping Luo,
  • Min Zhang,
  • Zhangyi Fu,
  • Jingbo Lan,
  • Di Wu and
  • Jingsong You

Beilstein J. Org. Chem. 2020, 16, 530–536, doi:10.3762/bjoc.16.49

Graphical Abstract
  • ][24][25]. Our group has recently reported F+ reagent-promoted Pd-catalyzed C7–H arylation of 1‑naphthamides, but this method still suffers from a few disadvantages (Scheme 1) [26]. First, the precious metal palladium is employed as a catalyst. Moreover, stoichiometric F+ reagent is needed to oxidize
PDF
Album
Supp Info
Full Research Paper
Published 30 Mar 2020

Synthesis of six-membered silacycles by borane-catalyzed double sila-Friedel–Crafts reaction

  • Yafang Dong,
  • Masahiko Sakai,
  • Kazuto Fuji,
  • Kohei Sekine and
  • Yoichiro Kuninobu

Beilstein J. Org. Chem. 2020, 16, 409–414, doi:10.3762/bjoc.16.39

Graphical Abstract
  • , the transformation of the amino groups in phenoxasilin 3a into phenyl groups was carried out (Scheme 5). First, the ammonium salt 4 was prepared by treating 3a with MeOTf followed by a palladium-catalyzed cross-coupling reaction with the Grignard reagent (PhMgBr) that afforded the desired diphenylated
PDF
Album
Supp Info
Letter
Published 17 Mar 2020

Room-temperature Pd/Ag direct arylation enabled by a radical pathway

  • Amy L. Mayhugh and
  • Christine K. Luscombe

Beilstein J. Org. Chem. 2020, 16, 384–390, doi:10.3762/bjoc.16.36

Graphical Abstract
  • -temperature, Pd/Ag-catalyzed direct arylation systems are radical-mediated. This is in contrast to the commonly proposed two-electron mechanisms for direct arylation and appears to extend to other substrates such as benzo[b]thiophene and pentafluorobenzene. Keywords: direct arylation; indole; palladium
  • cycle. This type of mechanism has been previously proposed for aryl and alkene alkylations [27][28], but not for direct arylation systems. A possible mechanism is outlined in Scheme 3, informed by the previous reports [27][28][29][30]. The aryl iodide 4 undergoes SET with an excited palladium(0) species
  • to form hybrid palladium-radical intermediate 5. This carbon-centered radical can then add to the indole. From here, three different pathways to rearomatize 7 are possible, eventually affording the arylated product 10. Pd(0) can be regenerated by a base; in this case, the silver carboxylate. This
PDF
Album
Supp Info
Full Research Paper
Published 13 Mar 2020

Architecture and synthesis of P,N-heterocyclic phosphine ligands

  • Wisdom A. Munzeiwa,
  • Bernard Omondi and
  • Vincent O. Nyamori

Beilstein J. Org. Chem. 2020, 16, 362–383, doi:10.3762/bjoc.16.35

Graphical Abstract
  • (R = Me, iPr and Ph) in benzene produced chiral acetal 22. Subsequent palladium-catalyzed C–C coupling of the acetal with 4-fluorophenylboronic acid (FPBA) in the presence of caesium carbonate and tri-tert-butylphosphine afforded aryl fluorides 23. Pure ligands 24 (63–72%) were obtained by
  • treated with triflic anhydride to afford the corresponding triflate 55. Microwave-assisted reduction of compound 55 with pyridinium chloride afforded the α-chloropyridine derivative 56, which was further catalytically dehalogenated with palladium on carbon and formic acid to generate the pyridine scaffold
  • 57. Coupling of 57 with Ph2PCl·BH3 resulted in the boron-protected ligand 58, which was deprotected with Et3N. Alternatively, 1,1’-bis(diphenylphosphino)ferrocene (dppf) with palladium(II) acetate was used to catalyze the reduction of 55 generating the pyridine scaffold 57. Subsequent lithiation and
PDF
Album
Review
Published 12 Mar 2020

Synthesis and circularly polarized luminescence properties of BINOL-derived bisbenzofuro[2,3-b:3’,2’-e]pyridines (BBZFPys)

  • Ryo Takishima,
  • Yuji Nishii,
  • Tomoaki Hinoue,
  • Yoshitane Imai and
  • Masahiro Miura

Beilstein J. Org. Chem. 2020, 16, 325–336, doi:10.3762/bjoc.16.32

Graphical Abstract
  • available (S)- and (R)-1,1’-bi-2-naphthols through a palladium-catalyzed multiple intramolecular C–H/C–H coupling as the key ring-closure step. The effect of terminal tert-butyl substituents on the BBZFPy skeleton was systematically investigated to uncover a unique aggregation-induced enhancement of CPL
  • ; palladium; Introduction Densely-fused (hetero)aromatic compounds have been a key motif in a wide range of manufactured functional molecules, as they exhibit fundamentally useful electrochemical and photophysical properties. Considerable effort has therefore taken into the development of efficient methods
  • for the construction of such polycyclic scaffolds, and the last decade has witnessed a remarkable improvement in the palladium-catalyzed C–H/C–H oxidative coupling as one of the potential synthetic strategies [1]. This method is straightforward and highly step-economical, enabling us to produce
PDF
Album
Supp Info
Full Research Paper
Published 06 Mar 2020

Formal preparation of regioregular and alternating thiophene–thiophene copolymers bearing different substituents

  • Atsunori Mori,
  • Keisuke Fujita,
  • Chihiro Kubota,
  • Toyoko Suzuki,
  • Kentaro Okano,
  • Takuya Matsumoto,
  • Takashi Nishino and
  • Masaki Horie

Beilstein J. Org. Chem. 2020, 16, 317–324, doi:10.3762/bjoc.16.31

Graphical Abstract
  • -substituted thiophene 3 efficiently gave chlorobithiophene 4 in a facile manner (Scheme 2) [25]. The use of a palladium catalyst efficiently suppressed the undesired polymerization to afford the HT halobithiophene with different substituents [26]. Compared to an alternative pathway to 4, in which initial
  • RINT-2000(CuKα) device. Concerning the solvent of the nickel- and palladium-catalyzed reactions, THF (anhydrous grade) was purchased from Kanto Chemical Co. Ltd. and passed through alumina and copper columns (Nikko Hansen & Co. Ltd.) or distilled from a sodium dispersion in mineral oil/benzophenone
PDF
Album
Full Research Paper
Published 05 Mar 2020

Allylic cross-coupling using aromatic aldehydes as α-alkoxyalkyl anions

  • Akihiro Yuasa,
  • Kazunori Nagao and
  • Hirohisa Ohmiya

Beilstein J. Org. Chem. 2020, 16, 185–189, doi:10.3762/bjoc.16.21

Graphical Abstract
  • . The synergistic palladium/copper-catalyzed reaction of aromatic aldehydes, allylic carbonates, and a silylboronate produces the corresponding homoallylic alcohol derivatives. This process involves the catalytic formation of a nucleophilic α-silyloxybenzylcopper(I) species and the subsequent palladium
  • -catalyzed allylic substitution. Keywords: aldehyde; copper; copper catalysis; cross-coupling; palladium; synthetic method; Introduction α-Alkoxy-substituted carbanions (α-alkoxyalkyl anions) are useful C(sp3) nucleophiles for the construction of alcohol units found in a majority of pharmaceuticals
  • rearrangement and then successfully trapped with aryl bromides under palladium catalysis (Scheme 1). This system was extended to an asymmetric version using the chiral α-silyloxybenzylcopper(I) species having a chiral NHC ligand. In the asymmetric system, one example of allylic carbonate was used as the carbon
PDF
Album
Supp Info
Letter
Published 07 Feb 2020

Functionalization of the imidazo[1,2-a]pyridine ring in α-phosphonoacrylates and α-phosphonopropionates via microwave-assisted Mizoroki–Heck reaction

  • Damian Kusy,
  • Agata Wojciechowska,
  • Joanna Małolepsza and
  • Katarzyna M. Błażewska

Beilstein J. Org. Chem. 2020, 16, 15–21, doi:10.3762/bjoc.16.3

Graphical Abstract
  • compounds 1 and 2, was successfully achieved. We used tri(o-tolyl)phosphine (P(o-tol)3, palladium(II) acetate (Pd(OAc)2) and obtained the desired product (E)-benzyl 3-(6-aminopyridin-3-yl)acrylate in 86% yield (see Supporting Informaiton File 1). However, when the same conditions were applied to alkyl 3-(6
  • , replacing palladium acetate by tetrakis(triphenylphosphine)palladium(0), Pd(PPh3)4, or bis(dibenzylideneacetone)palladium(0), Pd(dba)2, resulted in a reduction in the yield and a partial recovery of substrate 1a (Table 1, entries 13 and 14). In the next series of experiments, different amines were tested
  • analyses. Compound 4 is the product of dehalogenation reaction, which is known to be catalyzed by palladium [1][18]. The second side product, compound 5, is probably the result of the cleavage of the C–CN bond in the solvent (propionitrile or acetonitrile), followed by a palladium-catalyzed cyanation of
PDF
Album
Supp Info
Full Research Paper
Published 03 Jan 2020

Synthesis of C-glycosyl phosphonate derivatives of 4-amino-4-deoxy-α-ʟ-arabinose

  • Lukáš Kerner and
  • Paul Kosma

Beilstein J. Org. Chem. 2020, 16, 9–14, doi:10.3762/bjoc.16.2

Graphical Abstract
  • ppm in compound 9). Full deprotection, including the conversion of the 4-azido group into an amino function, was accomplished by hydrogenation of 9 in the presence of palladium hydroxide in a 1:1 mixture of methanol/acetic acid to deliver the target derivative 11 in 30% yield after final HILIC
PDF
Album
Supp Info
Full Research Paper
Published 02 Jan 2020

Synthesis and characterization of bis(4-amino-2-bromo-6-methoxy)azobenzene derivatives

  • David Martínez-López,
  • Amirhossein Babalhavaeji,
  • Diego Sampedro and
  • G. Andrew Woolley

Beilstein J. Org. Chem. 2019, 15, 3000–3008, doi:10.3762/bjoc.15.296

Graphical Abstract
  • substituent in an attempt to enhance water solubility. Late-stage functionalization of the ortho-position was carried out through a palladium(II)-catalyzed C–H activation, resulting in ortho-brominated azobenzenes [21]. Photochemical characterization Despite the morpholino substituent, compound 4 was found to
  • -morpholinophenyl)diazene (4): (E)-1,2-Bis(2-methoxy-4-morpholinophenyl)diazene (9, 30 mg, 0.07 mmol) was dissolved in DCM. To this solution, palladium acetate (1.6 mg, 0.007 mmol) was added, and the resulting mixture was stirred for 15 minutes. Then, N-bromosuccinimide (28.6 mg, 0.16 mmol) was added to the
  • (3-bromo-5-methoxy-4,1-phenylene)-(E)-bis(1λ4-piperazine-1-carboxylate) (11): Di-tert-butyl 4,4'-(diazene-1,2-diyl)bis(3-methoxy-4,1-phenylene)-(E)-bis(1λ4-piperazine-1-carboxylate) (10, 30 mg, 0.04 mmol) was dissolved in DCM, palladium acetate (1 mg, 0.004 mmol) was added, and the resulting mixture
PDF
Album
Supp Info
Full Research Paper
Published 30 Dec 2019
Other Beilstein-Institut Open Science Activities