Search results

Search for "catalyst" in Full Text gives 1789 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

Hypervalent iodine-mediated intramolecular alkene halocyclisation

  • Charu Bansal,
  • Oliver Ruggles,
  • Albert C. Rowett and
  • Alastair J. J. Lennox

Beilstein J. Org. Chem. 2024, 20, 3113–3133, doi:10.3762/bjoc.20.258

Graphical Abstract
  • aza-heterocycles were synthesised in good yields. The authors proposed a mechanism for the fluorocyclisation reactions (Scheme 6), which relies on the activation of the fluoro-iodane reagent 12 with the zinc catalyst. The activation enables better orbital overlap to occur with the π bond of the alkene
  • from the styrenyl starting materials is stereoselective, giving the syn-diasteroisomer in high yields. A chiral iodoarene catalyst 16 was employed, along with a stoichiometric sacrificial oxidant, to give good to excellent levels of enantioselectivity. This elegant strategy led to a variety of β
  • the oxyfluorination of alkenes in 2015 [31]. Under identical conditions to the aminofluorination using 1-fluoro-3,3-dimethylbenziodoxole (12) with Zn(BF4)2 catalyst, unsaturated alcohols were cyclised to fluorinated tetrahydropyrans 26 and oxepanes 28 (Scheme 12) in 1–2 hours in good yields. Gulder
PDF
Album
Review
Published 28 Nov 2024

Advances in the use of metal-free tetrapyrrolic macrocycles as catalysts

  • Mandeep K. Chahal

Beilstein J. Org. Chem. 2024, 20, 3085–3112, doi:10.3762/bjoc.20.257

Graphical Abstract
  • interactions and accelerating aldol reactions. In the absence of a catalyst, no reaction between 2-(trimethylsilyloxy)furan (TMSOF, 13) and benzaldehyde (14) was observed, whereas all the tested macrocyclic compounds were found catalytically active, with 11 being the most efficient providing erythro/threo (15
  • , with TBAI as a co-catalyst, up to 74% yields (Table 1). The inactivity of porphyrin 18 was attributed to the inaccessibility of the inner core imine due to its planar structure. The mechanism of the epoxide ring-opening reaction was elucidated by DFT calculations, which suggested that the macrocycle
  • activates the Cu–Cl bond via chloride···calixpyrrole (N–H···Cl) hydrogen-bonding interactions toward the formation of the nitrene intermediate from chloramine-T (NaCl=NTs). Additionally, calix[4]pyrrole served as a phase-transfer catalyst in this reaction. Since chloramine-T had low solubility in
PDF
Album
Review
Published 27 Nov 2024

Enantioselective regiospecific addition of propargyltrichlorosilane to aldehydes catalyzed by biisoquinoline N,N’-dioxide

  • Noble Brako,
  • Sreerag Moorkkannur Narayanan,
  • Amber Burns,
  • Layla Auter,
  • Valentino Cesiliano,
  • Rajeev Prabhakar and
  • Norito Takenaka

Beilstein J. Org. Chem. 2024, 20, 3069–3076, doi:10.3762/bjoc.20.255

Graphical Abstract
  • systematic catalyst structure–reactivity and selectivity relationship study. The observed catalyst structure–enantioselectivity relationship of the present allenylation reaction was found exactly opposite to that of the analogous allylation reaction. The method provided eleven α-allenic alcohols in 22–99
  • enantioenriched form [11][12]. However, such metal/metalloid reagents and the corresponding metal catalyst-bound intermediates often equilibrate between possible regioisomeric forms and can undergo both, SE2 and SE2’ addition reactions, resulting in a mixture of homopropargylic alcohols and α-allenic alcohols [14
  • stable allenyltrichlorosilane that affords undesired homopropargylic alcohols [35][36] (Scheme 2b). Furthermore, Iseki [35] and Nakajima [36] evaluated only one chiral catalyst in their independent studies (i.e., no catalyst structure–reactivity and selectivity relationship study). In this context, we
PDF
Album
Supp Info
Letter
Published 25 Nov 2024
Graphical Abstract
  • , [3]rotaxane diol 10 was used as the initiator of the controlled ring-opening polymerization (ROP) of ε-caprolactone in the presence of a diphenyl phosphate catalyst to introduce the polyester main chain into the rotaxane framework; the successive end-capping reactions yielded macromolecular [3
PDF
Album
Review
Published 19 Nov 2024

Advances in radical peroxidation with hydroperoxides

  • Oleg V. Bityukov,
  • Pavel Yu. Serdyuchenko,
  • Andrey S. Kirillov,
  • Gennady I. Nikishin,
  • Vera A. Vil’ and
  • Alexander O. Terent’ev

Beilstein J. Org. Chem. 2024, 20, 2959–3006, doi:10.3762/bjoc.20.249

Graphical Abstract
  • of allylic peroxidation 2, 4 and 5 were observed (Scheme 4) [24]. Similar transformations were reported later using CuCl as the catalyst [39]. Later, Gade with coauthors demonstrated the allylic peroxidation of cyclohexane with TBHP using the alkylperoxocobalt(III) complexes [Co(BPI)(OAc)(OO-t-Bu
  • -t-Bu)2. Allylic peroxidation of 3-substituted prop-1-ene-1,3-diyldibenzenes 8 was performed with TBHP as the oxidant/peroxidation agent and with Cu2O as the catalyst [42] (Scheme 6). The proposed mechanism of peroxides 9 formation does not include peroxo–copper complexes and begins with the
  • ) [51]. The corresponding peroxides 30 are enough stable under the reaction conditions and were isolated in high yields (Scheme 12). Flow-modification of the 2-oxoindole peroxidation method using nanoparticles of iron oxide as the catalyst was proposed [52]. The summarized proposed reaction pathway is
PDF
Album
Review
Published 18 Nov 2024

Recent advances in transition-metal-free arylation reactions involving hypervalent iodine salts

  • Ritu Mamgain,
  • Kokila Sakthivel and
  • Fateh V. Singh

Beilstein J. Org. Chem. 2024, 20, 2891–2920, doi:10.3762/bjoc.20.243

Graphical Abstract
  • concentrate on various arylation reactions involving carbon and other heteroatoms, encompassing rearrangement reactions in the absence of any metal catalyst, and summarize advancements made in the last five years. Keywords: arylation reaction; diaryliodonium salts; electrophilic arylation reagent; metal-free
  • formation of the Nu–Ar product and aryl iodide [21]. Second, the arylation can take place in the presence of a metal catalyst via oxidative addition, followed by reduction elimination [48][49]. Thirdly, it proceeds through a ligand-coupled arylation which involves a five-membered transition state to yield
  • , C–C bond formation was reported by Chen and colleagues in 2020 via the arylation of vinyl pinacol boronates 23 by using diaryliodonium salts 16 to yield trans-arylvinylboronates 24 in the absence of a metal catalyst [62]. The optimized reaction conditions involve the reaction of substituted
PDF
Album
Review
Published 13 Nov 2024

Synthesis of pyrrole-fused dibenzoxazepine/dibenzothiazepine/triazolobenzodiazepine derivatives via isocyanide-based multicomponent reactions

  • Marzieh Norouzi,
  • Mohammad Taghi Nazeri,
  • Ahmad Shaabani and
  • Behrouz Notash

Beilstein J. Org. Chem. 2024, 20, 2870–2882, doi:10.3762/bjoc.20.241

Graphical Abstract
  • , and triazolobenzodiazepine under solvent- and catalyst-free conditions. Purposefully, this approach produced various bioactive scaffolds using environmentally friendly, mild, and simple conditions. Due to their bioactive moieties, these compounds with exclusive fluorescence properties may attract
  • solvent- and catalyst-free conditions (Scheme 1d). Results and Discussion Synthesis Dibenzoxazepine as imine component, cyclohexyl isocyanide, and the gem-diactivated olefin (2-benzylidenemalononitrile) were selected as the starting materials to screen the reaction conditions (Scheme 2, Table 1). First
  • isocyanides, gem-diactivated olefins, and cyclic imines (dibenzoxazepines, dibenzothiazepine, and triazolobenzodiazepine) under catalyst- and solvent-free conditions. Furthermore, the other advantages of this reaction include the manufacturing premium pharmaceutical scaffolds, a wide range of substrates
PDF
Album
Supp Info
Full Research Paper
Published 11 Nov 2024

Mechanochemical difluoromethylations of ketones

  • Jinbo Ke,
  • Pit van Bonn and
  • Carsten Bolm

Beilstein J. Org. Chem. 2024, 20, 2799–2805, doi:10.3762/bjoc.20.235

Graphical Abstract
  • sodium fluoride catalyst, with simple ketones, which resulted in the formation of difluoromethyl 2,2-difluorocyclopropyl ethers (Scheme 1B). Although the reactions worked well, it is also noteworthy that the use of TFDA as reagent, liberated fluoro(trimethyl)silane (TMSF), carbon dioxide, and ozone
PDF
Album
Supp Info
Letter
Published 04 Nov 2024

C–C Coupling in sterically demanding porphyrin environments

  • Liam Cribbin,
  • Brendan Twamley,
  • Nicolae Buga,
  • John E. O’ Brien,
  • Raphael Bühler,
  • Roland A. Fischer and
  • Mathias O. Senge

Beilstein J. Org. Chem. 2024, 20, 2784–2798, doi:10.3762/bjoc.20.234

Graphical Abstract
  • further substitution directly on the meso- or a meso-phenyl ortho/meta/para positions of a porphyrin, is the introduction of C–C bond forming chemistry. This is typically achieved using palladium and/or another transition-metal catalyst [20]. Sonagashira [21], Suzuki–Miyaura [22], Heck [23], Stille [24
  • the Suzuki coupling began with investigating first the Suzuki reaction compatibility of boronic acid 14 with porphyrin 13. Porphyrin 13 and phenylboronic acid (14) were subjected to coupling at 85 °C for 48 hours using Pd2dba3/SPhos as a catalyst/ligand giving porphyrin 26 in a 32% yield, based on a
  • [48], switching catalyst to Pd(PPh)3, and base to Na2CO3 (Table 1, entry 16) gave no product. Ultimately, an increased catalyst loading of 25 mol % per C–Br bond gave the desired porphyrin in a 16% yield when using Cs2CO3 as base. The synthesis of other heterocycle-appended dodecasubstituted
PDF
Album
Supp Info
Full Research Paper
Published 04 Nov 2024

Access to optically active tetrafluoroethylenated amines based on [1,3]-proton shift reaction

  • Yuta Kabumoto,
  • Eiichiro Yoshimoto,
  • Bing Xiaohuan,
  • Masato Morita,
  • Motohiro Yasui,
  • Shigeyuki Yamada and
  • Tsutomu Konno

Beilstein J. Org. Chem. 2024, 20, 2776–2783, doi:10.3762/bjoc.20.233

Graphical Abstract
  • published that the asymmetric conjugate addition of 4-methylphenylboronic acid towards (E)-5-bromo-4,4,5,5-tetrafluoro-1-phenyl-2-penten-1-one (8) in the presence of a rhodium catalyst coordinated with (S)-BINAP gave the corresponding Michael adduct 9 in 94% enantiomeric excess (reaction 2, Scheme 1) [22
  • chlorides in the presence of a copper catalyst to afford the corresponding tetrafluoroethylenated ketones 19. The ketones were then condensed with (R)-1-phenylethylamine under the influence of TiCl4 [34][35] to prepare various optically active imines (R)-16 in high yields (Scheme 3). Based on the result of
PDF
Album
Supp Info
Full Research Paper
Published 01 Nov 2024

Copper-catalyzed yne-allylic substitutions: concept and recent developments

  • Shuang Yang and
  • Xinqiang Fang

Beilstein J. Org. Chem. 2024, 20, 2739–2775, doi:10.3762/bjoc.20.232

Graphical Abstract
  • alkoxylation and alkylation products with the assistance of Lewis acid as co-catalyst (Scheme 9). Starting from four different racemic substrates, the same product 6g with 96% ee was obtained under standard conditions. This indicates that the reactions proceed through the same transition state and the
  • stereocenter of the product is controlled by the catalyst. A single crystal of Cu(I) was investigated by X-ray and proved to be the dicopper complex, while the Cu(II) catalyst was revealed as mononuclear copper coordinated with two ligands. Further kinetic isotope experiments and nonlinear relationship studies
  • demonstrated that the terminal alkyne unit is crucial for the process and the reactions using different isomers all proceed via the same intermediate. Nonlinear relationship experiments proved that the active catalyst is a mono-copper complex containing one ligand. A catalytic cycle is proposed in which copper
PDF
Album
Review
Published 31 Oct 2024

Synthesis of spiroindolenines through a one-pot multistep process mediated by visible light

  • Francesco Gambuti,
  • Jacopo Pizzorno,
  • Chiara Lambruschini,
  • Renata Riva and
  • Lisa Moni

Beilstein J. Org. Chem. 2024, 20, 2722–2731, doi:10.3762/bjoc.20.230

Graphical Abstract
  • -pot multistep synthesis of unprecedent 2,3-diaminoindolenines using graphene oxide (GO) as heterogeneous catalyst [21]. The protocol involves the three-component Ugi (3C-Ugi) reaction between aldehydes, isocyanides and 2 equivalents of electron-rich anilines to give α-aminoamidines, which undergo a C
  • ] 3. Based on our experience on the use of graphene oxide (GO) as heterogeneous catalyst to promote MCRs and subsequent C–N bond oxidation [16][21], we first investigated the GO-promoted oxidation of N-Ph-THIQ and the subsequent 3C Ugi reaction to give α-aminoamidine 2a. Applying the previously
  • isocyanide without the presence of a catalyst. In order to establish the role of GO we carried out the 3C Ugi-type reaction starting from iminium ion 1a, freshly prepared by visible light irradiation in the presence of bromochloroform [28]. This protocol resulted quite convenient as can be conducted under
PDF
Album
Supp Info
Full Research Paper
Published 29 Oct 2024

5th International Symposium on Synthesis and Catalysis (ISySyCat2023)

  • Anthony J. Burke and
  • Elisabete P. Carreiro

Beilstein J. Org. Chem. 2024, 20, 2704–2707, doi:10.3762/bjoc.20.227

Graphical Abstract
  • novel lipophilic cinchona squaramide organocatalyst. This organocatalyst was evaluated in a benchmark Michael addition of acetylacetone to trans-β-nitrostyrene, yielding the Michael adduct with high yield and enantioselectivity. The hydrophobic chain of the catalyst allowed the organocatalyst to be
  • easily recovered by precipitation using polar solvents. This catalyst proved to be excellent for the preparation of (S)-baclofen on a gram scale, furnishing the main chiral intermediate in high yield and enantioselectivity. Furthermore, the catalyst was recycled over five cycles while maintaining its
PDF
Album
Editorial
Published 28 Oct 2024

Synthesis of fluoroalkenes and fluoroenynes via cross-coupling reactions using novel multihalogenated vinyl ethers

  • Yukiko Karuo,
  • Keita Hirata,
  • Atsushi Tarui,
  • Kazuyuki Sato,
  • Kentaro Kawai and
  • Masaaki Omote

Beilstein J. Org. Chem. 2024, 20, 2691–2703, doi:10.3762/bjoc.20.226

Graphical Abstract
  • First, we optimized the conditions of the Suzuki–Miyaura cross-coupling in reference to the report by Yang et al. (Table 1) [43]. Upon the treatment of multihalogenated vinyl ether 1a with phenylboronic acid 4a (1.3 equiv) and palladium diacetate (10 mol %) as a catalyst at 40 °C, Suzuki–Miyaura cross
  • synthesized in 84% yield under reflux conditions (Table 1, entries 3 and 4). Next, we examined an effective catalyst for the cross-coupling. Reactions using palladium dichloride or bis(2,4-pentanedionato)palladium significantly reduced the yields of 2a (Table 1, entries 5 and 6, respectively). When an
  • allylpalladium chloride dimer or bis(triphenylphosphine)palladium dichloride were used as catalyst, the reaction proceeded with the same yield as that in Table 1, entry 4 (entries 7 and 8). Utilizing palladium catalyst such as bis(triphenylphosphine)palladium dichloride, all these reactions could convert 1a into
PDF
Album
Supp Info
Full Research Paper
Published 24 Oct 2024

Computational design for enantioselective CO2 capture: asymmetric frustrated Lewis pairs in epoxide transformations

  • Maxime Ferrer,
  • Iñigo Iribarren,
  • Tim Renningholtz,
  • Ibon Alkorta and
  • Cristina Trujillo

Beilstein J. Org. Chem. 2024, 20, 2668–2681, doi:10.3762/bjoc.20.224

Graphical Abstract
  • catalyst efficiency and selectivity in sustainable chemistry applications. Keywords: asymmetric catalysis; carbon dioxide; CO2; epoxide; frustrated Lewis pairs (FLPs); volcano plot; Introduction The field of frustrated Lewis pairs (FLPs) has flourished since their seminal discovery in 2006 by Stephan and
  • H2 over CO2 becomes crucial for effective CO2 reduction [7]. Additionally, the strength of the interaction between the catalyst and the resulting system after hydride transfer presents a limitation. The formation of a robust LA–oxygen interaction may impede proton transfer to the basic oxygen atom
  • catalyst [23][24][25]. Therefore, the stereochemical aspects of CO2 insertion into PO enabled by FLP catalysts should be investigated. To the best of our knowledge, only one paper has proposed an asymmetric approach to this reaction using a metal-based catalyst [23]. However, our approach differs
PDF
Album
Supp Info
Full Research Paper
Published 22 Oct 2024

Transition-metal-free decarbonylation–oxidation of 3-arylbenzofuran-2(3H)-ones: access to 2-hydroxybenzophenones

  • Bhaskar B. Dhotare,
  • Seema V. Kanojia,
  • Chahna K. Sakhiya,
  • Amey Wadawale and
  • Dibakar Goswami

Beilstein J. Org. Chem. 2024, 20, 2655–2667, doi:10.3762/bjoc.20.223

Graphical Abstract
  • (3H)-ones to 2-hydroxybenzophenones via decarbonylation–oxidation quickly and without the need of a transition-metal catalyst. Herein, a novel decarbonylation–oxidation method for 3-arylbenzofuran-2(3H)-ones has been developed for the synthesis of 2-hydroxybenzophenones via a transition-metal-free
  • catalyst was essential for this reaction to happen at a higher temperature, and the products were obtained in negligible yields without the catalyst. Our protocol established that the reaction proceeds without the need for a transition-metal catalyst, as well as at a lower temperature. Additionally, the
PDF
Album
Supp Info
Full Research Paper
Published 21 Oct 2024

The scent gland composition of the Mangshan pit viper, Protobothrops mangshanensis

  • Jonas Holste,
  • Paul Weldon,
  • Donald Boyer and
  • Stefan Schulz

Beilstein J. Org. Chem. 2024, 20, 2644–2654, doi:10.3762/bjoc.20.222

Graphical Abstract
  • concentrated under a stream of N2. Hydrogenation: The solvent of the natural extract (100 µL) was removed with a stream of N2 and taken up in pentane (100 µL) and a catalytic amount of Pd/C was added. The reaction was then stirred for 1 h under a H2 atmosphere. The catalyst was filtered and rinsed with pentane
PDF
Album
Supp Info
Full Research Paper
Published 18 Oct 2024

Efficient modification of peroxydisulfate oxidation reactions of nitrogen-containing heterocycles 6-methyluracil and pyridine

  • Alfiya R. Gimadieva,
  • Yuliya Z. Khazimullina,
  • Aigiza A. Gilimkhanova and
  • Akhat G. Mustafin

Beilstein J. Org. Chem. 2024, 20, 2599–2607, doi:10.3762/bjoc.20.219

Graphical Abstract
  • use of catalysts reduced the duration of the oxidation reaction and significantly increased the yield of sulfate derivatives. In the presence of РсМ, the optimal duration for the oxidation reaction of MU (1) was found to be 4 hours. When the catalyst was not applied, the yield of MU-5-ammonium sulfate
  • catalyst increasing by a factor of 10 in each successive experiment. As described in [13] PcFe(II), PcСo, and PcFe(III) exhibited the highest activity in oxidizing reactions of MU (1). Addition of these catalysts in the amount of 0.01–0.05 wt % increased the yield of MU-5-ammonium sulfate 2 to 82–95%. The
  • maximum yield of compound 2, equal to 95%, was obtained when 0.05 wt % PcFe(II) was introduced into the reaction. However, on enhancing the catalyst's quantity to 0.1 wt %, the product yield decreased to 33–45%. Further increase in the quantity of catalyst led to a greater decline in the yield of MU-5
PDF
Album
Supp Info
Full Research Paper
Published 16 Oct 2024

Transition-metal-free synthesis of arylboronates via thermal generation of aryl radicals from triarylbismuthines in air

  • Yuki Yamamoto,
  • Yuki Konakazawa,
  • Kohsuke Fujiwara and
  • Akiya Ogawa

Beilstein J. Org. Chem. 2024, 20, 2577–2584, doi:10.3762/bjoc.20.216

Graphical Abstract
  • with halogen or triflate groups. Recently, transition-metal-catalyzed direct borylation of arenes via C–H bond activation has been reported, although the design of the substrate and ligands is somewhat complicated [16][17][18][19][20][21][22]. Since the complete removal of catalyst-derived metal
PDF
Album
Supp Info
Full Research Paper
Published 11 Oct 2024

A review of recent advances in electrochemical and photoelectrochemical late-stage functionalization classified by anodic oxidation, cathodic reduction, and paired electrolysis

  • Nian Li,
  • Ruzal Sitdikov,
  • Ajit Prabhakar Kale,
  • Joost Steverlynck,
  • Bo Li and
  • Magnus Rueping

Beilstein J. Org. Chem. 2024, 20, 2500–2566, doi:10.3762/bjoc.20.214

Graphical Abstract
  • complex, C(sp3)–H bonds underwent azidation with high chemoselectivity, even in the absence of a directing group. The proposed mechanism involves the formation of the active catalyst Mn(III)(N3) via ligand exchange, followed by anodic oxidation to a Mn(IV)(N3)2 complex. This high–valent Mn(IV) species
  • benzylic position (Scheme 31). 1.3.2 Co-assisted anodic oxidation. In 2021, Xu and colleagues developed an electrocatalytic approach for the intramolecular oxidative allylic amination and C–H alkylation using cobalt–salen complexes as catalysts [43]. In this reaction, the cobalt catalyst [Co(II)] is first
  • back to [Co(II)] at the anode (Scheme 32). Recently, two additional studies on cobalt–salen complex-induced (cyclo)additions were reported by the Kim [44] and Findlater groups [45]. By employing cobalt–salen as a catalyst, along with PhMeSiH2 and dimethoxypyridine as additives, n-Bu4NPF6 as the
PDF
Album
Review
Published 09 Oct 2024

Visible-light-mediated flow protocol for Achmatowicz rearrangement

  • Joachyutharayalu Oja,
  • Sanjeev Kumar and
  • Srihari Pabbaraja

Beilstein J. Org. Chem. 2024, 20, 2493–2499, doi:10.3762/bjoc.20.213

Graphical Abstract
  • platforms, we herein present a photo-flow platform for Achmatowicz reactions. A novel photo-flow solar panel reactor was fabricated to test and validate the Achmatowicz rearrangement reaction (Figure S1, Supporting Information File 1), and the reaction conditions were optimized with a ruthenium catalyst. As
  • -off experiment was performed to examine the Achmatowicz rearrangement's dependence on light, and it was observed that continuous light irradiation was required (Table 1, entry 2). Next, we considered running the process without utilizing the Ru(bpy)3Cl2·6H2O catalyst (Table 1, entry 3). There was no
  • evidence of product formation, indicating that the Ru catalyst was required to pursue the photoinduced Achmatowicz rearrangement. Furthermore, it was observed that the product yield depended on resident time and it dropped over time as the residence time was reduced (see details in Supporting Information
PDF
Album
Supp Info
Letter
Published 08 Oct 2024

HFIP as a versatile solvent in resorcin[n]arene synthesis

  • Hormoz Khosravi,
  • Valeria Stevens and
  • Raúl Hernández Sánchez

Beilstein J. Org. Chem. 2024, 20, 2469–2475, doi:10.3762/bjoc.20.211

Graphical Abstract
  • as the catalyst (Table 1, entries 1–4). The removal of HCl from the reaction conditions unveiled the crucial role of the catalyst in the process (Table 1, entry 5), which was expected; however, note that here we use the acid in catalytic amounts and not in excess as reported in the literature [73][74
  • ][75]. Variations in catalyst nature between a Brønsted and Lewis acid, and the acid’s pKa (Table 1, entries 6–8) did not improve the yield compared to HCl (Table 1, entry 4). Last, further exploration of the conditions using HFIP/HCl revealed that the reaction progress achieves its maximum conversion
PDF
Album
Supp Info
Letter
Published 02 Oct 2024

Photoredox-catalyzed intramolecular nucleophilic amidation of alkenes with β-lactams

  • Valentina Giraldi,
  • Giandomenico Magagnano,
  • Daria Giacomini,
  • Pier Giorgio Cozzi and
  • Andrea Gualandi

Beilstein J. Org. Chem. 2024, 20, 2461–2468, doi:10.3762/bjoc.20.210

Graphical Abstract
  • the linked alkene moiety, followed by hydrogen transfer from the hydrogen atom transfer (HAT) catalyst. This process was used to successfully prepare 2-alkylated clavam derivatives. Keywords: β-lactam; acridinium photocatalyst; alkenes; amides; intramolecular radical reaction; photoredox catalysis
  • intramolecular nucleophilic attack induced by photocatalytic oxidation was reported by Yoon et al. with tosylamide derivatives [29]. Specifically, amides were employed in a photoredox cyclization process using a strong photooxidative acridinium catalyst such as the Fukuzumi catalyst (I, Figure 1B) [30][31
  • reaction was carried out in DCM with acridinium PC IV (5 mol %), 50 mol % of PhSSPh as HAT catalyst, and lutidine (50 mol %) as the base. Upon 72 hours of irradiation with a blue light at 456 nm, the product 11c was obtained in a satisfactory yield as a mixture of diastereoisomers in a 1.4:1 ratio (Table 1
PDF
Album
Supp Info
Full Research Paper
Published 01 Oct 2024

Evaluating the halogen bonding strength of a iodoloisoxazolium(III) salt

  • Dominik L. Reinhard,
  • Anna Schmidt,
  • Marc Sons,
  • Julian Wolf,
  • Elric Engelage and
  • Stefan M. Huber

Beilstein J. Org. Chem. 2024, 20, 2401–2407, doi:10.3762/bjoc.20.204

Graphical Abstract
  • abstractions, e.g. to activate gold chloride complexes [18][19]. Therefore, besides the development of new bidentate catalyst motifs, we were still interested in the optimization of these “simpler” derivatives. Thus, we designed a new catalyst motif [20] featuring an isoxazole ring, XB donor 7Z, and compared
  • 7Br, which hints that also in solution stronger binding to Lewis bases and therefore higher activity as catalyst may be expected (compared to prototypic iodolium 1Z). As a benchmark for the halogen-bonding strength in solution, the activation of (PPh3)AuCl was chosen. The activated gold(I) complex was
  • applied as catalyst for the cyclization of propargylic amide 11, a typical benchmark reaction in gold catalysis (Scheme 2) [24][25][26][27], which had previously already been activated by iodine(I) and iodine(III)-based XB donors [15][18]. To evaluate the activity of the new iodoloisoxazolium 7BArF, it
PDF
Album
Supp Info
Letter
Published 23 Sep 2024

Asymmetric organocatalytic synthesis of chiral homoallylic amines

  • Nikolay S. Kondratyev and
  • Andrei V. Malkov

Beilstein J. Org. Chem. 2024, 20, 2349–2377, doi:10.3762/bjoc.20.201

Graphical Abstract
  • enantioselectivities (90–99% ee) and good yields (75–94%) have been achieved on a wide range of aromatic and aliphatic N-acylimines 2 using chiral 3,3’-diaryl-BINOL 3 as catalyst (Scheme 2). The reaction proved to be highly tolerant to the nature of the R1 substituent in imine 2, demonstrating high yields and
  • 11. It was proposed, that the internal hydrogen bond between the catalyst 11 and the P=O fragment of the protecting group of imine 9 is responsible for the observed high enantioselectivities (76–98% ee). The scope included a wide range of substrates, such as aromatic, heteroaromatic, aliphatic, and α
  • -pyridyl)phenol (23) as an activator, 3 equivalents of HFIP and a slightly different catalyst, 3,3’-bis(3,5-bis(trifluoromethyl)phenyl)-BINOL 21 at 20 mol % loading (Scheme 5). Interestingly, the reaction showed an opposite trend and worked better with Z-geranylboronic acid (14). The scope was tested over
PDF
Album
Review
Published 16 Sep 2024
Other Beilstein-Institut Open Science Activities