Search results

Search for "cation" in Full Text gives 656 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

Confirmation of the stereochemistry of spiroviolene

  • Yao Kong,
  • Yuanning Liu,
  • Kaibiao Wang,
  • Tao Wang,
  • Chen Wang,
  • Ben Ai,
  • Hongli Jia,
  • Guohui Pan,
  • Min Yin and
  • Zhengren Xu

Beilstein J. Org. Chem. 2024, 20, 852–858, doi:10.3762/bjoc.20.77

Graphical Abstract
  • isotope labeling experiments [6], followed by a 2,7-cyclization, afforded C6 cationic intermediate IM-3 with cyclopiane skeleton. Quench of the cation IM-3 with water would give 4, while upon two 1,2-alkyl shifts of IM-3, followed by deprotonation of cation IM-4, would give spiroviolene (1). On the other
  • hand, the originally proposed cyclization mechanism (Scheme 1B) involves a 3,6-cyclization of cation IM-1 through a conformation shown as IM-1A to generate cation IM-5, which was proposed to undergo a dyotropic rearrangement, followed by a 1,2-alkyl shift of cation IM-6 to yield the spirocyclic cation
  • IM-7. A key 1,3-hydride shift of IM-7 from the β-face, followed by deprotonation of the formed C2-cation IM-8, would deliver the originally proposed structure 1' [6]. However, no related natural products that would be derived from the intermediates of this pathway have been found so far. A third
PDF
Album
Supp Info
Letter
Published 18 Apr 2024

Skeletal rearrangement of 6,8-dioxabicyclo[3.2.1]octan-4-ols promoted by thionyl chloride or Appel conditions

  • Martyn Jevric,
  • Julian Klepp,
  • Johannes Puschnig,
  • Oscar Lamb,
  • Christopher J. Sumby and
  • Ben W. Greatrex

Beilstein J. Org. Chem. 2024, 20, 823–829, doi:10.3762/bjoc.20.74

Graphical Abstract
  • vicinal proton. These results suggested that the formation of the allylic cation occurred readily from alcohols 15 and 18; however, the transition states leading to the rearrangement products were inaccessible and so only chloride addition occurred. The generation of the rearrangement products from the
PDF
Album
Supp Info
Full Research Paper
Published 16 Apr 2024

Advancements in hydrochlorination of alkenes

  • Daniel S. Müller

Beilstein J. Org. Chem. 2024, 20, 787–814, doi:10.3762/bjoc.20.72

Graphical Abstract
  • alkene reactivity is essential. Two reactivity scales for alkenes are available in the literature, one considering the reactivity of the alkene itself (Mayr scale) [25] and the other the stability of the corresponding cation after protonation (hydride affinities) [26]. In the polar hydrochlorination
  • ). Conveniently, these parameters are freely available on Mayr's database of reactivity parameters [28]. On the other hand, one can assess the stability of the in situ-generated cation. The greater its stability, the easier the protonation of the alkene will be, making it more reactive towards hydrochlorination
  • . Thermodynamic and theoretical data provide hydride affinities, which correspond to the negative heat of formation for the combination of a hydride anion with a given cation in the gas phase (Figure 2) [26][29]. It is noteworthy that, in contrast to the hydride affinity scale, the Mayr scale considers energetic
PDF
Album
Review
Published 15 Apr 2024

Synthesis and characterization of water-soluble C60–peptide conjugates

  • Yue Ma,
  • Lorenzo Persi and
  • Yoko Yamakoshi

Beilstein J. Org. Chem. 2024, 20, 777–786, doi:10.3762/bjoc.20.71

Graphical Abstract
  • ) showed much smaller aggregation (dotted green line, ≈12 nm), providing a transparent solution, while C60–oligo-Arg (5c) remained insoluble over the tested pH value range (4.0–9.2). This was presumably due to the strong cation–π interactions between the cationic Arg moieties and the aromatic C60, which is
PDF
Album
Supp Info
Full Research Paper
Published 12 Apr 2024

Recent developments in the engineered biosynthesis of fungal meroterpenoids

  • Zhiyang Quan and
  • Takayoshi Awakawa

Beilstein J. Org. Chem. 2024, 20, 578–588, doi:10.3762/bjoc.20.50

Graphical Abstract
  • resulting cation intermediate at C-4' to induce an acyl shift, forming the steroid-like structure of 7 with a 6-6-6-5 ring (Figure 2). Swapping terpenoid cyclases in heterologous expression systems A search of the genome database for Trt1-homolog CYCs revealed the enzyme AusL (41% identity with Trt1) in
  • -6-6-5-membered andrastin E (9) (Figure 2) [8]. Like Trt1, AusL and AdrI create the common cation intermediate from 6, but they deprotonate the cationic intermediate from H-1' and H-11, respectively [12]. The differences in the structural bases of Trt1, AusL, and AdrI are quite intriguing, in that
  • cyclization process is initiated by the protonation of the epoxide, leading to the formation of a cation at C-7. A cascade of migrations then occurs to form the final backbone: H-6 migrates to the cation C-7, the methyl group at C-13 shifts to C-6, H-10 migrates to C-11, and finally, deprotonation of the
PDF
Album
Review
Published 13 Mar 2024

Synthesis of photo- and ionochromic N-acylated 2-(aminomethylene)benzo[b]thiophene-3(2Н)-ones with a terminal phenanthroline group

  • Vladimir P. Rybalkin,
  • Sofiya Yu. Zmeeva,
  • Lidiya L. Popova,
  • Irina V. Dubonosova,
  • Olga Yu. Karlutova,
  • Oleg P. Demidov,
  • Alexander D. Dubonosov and
  • Vladimir A. Bren

Beilstein J. Org. Chem. 2024, 20, 552–560, doi:10.3762/bjoc.20.47

Graphical Abstract
  • the phenanthroline unit was also involved in a π–π-stacking interaction (blue plane–green plane in Figure 4), with the plane centroid–plane centroid distance being 3.6998(8) Å (plane shift 1.4919(17) Å, twist and fold angles 1.54° and 1.92°, respectively). Cation-induced transformations of the
PDF
Album
Supp Info
Full Research Paper
Published 11 Mar 2024

Switchable molecular tweezers: design and applications

  • Pablo Msellem,
  • Maksym Dekthiarenko,
  • Nihal Hadj Seyd and
  • Guillaume Vives

Beilstein J. Org. Chem. 2024, 20, 504–539, doi:10.3762/bjoc.20.45

Graphical Abstract
  • flourished since the early 2000s. It is worth mentioning, as stated by Leigh in a comprehensive review [15], that a pioneering example of a molecular machine was the photoswitchable molecular tweezers developed by Shinkai [16] in 1981 for photocontrolled cation binding. This novel class of tweezers
  • achieved by chemical (hydroxide/proton) or electrochemical (reduction/oxidation) stimuli allowing the enrichment of perylene from a mixture of polycyclic aromatic hydrocarbons (PAHs) in phase-transfer experiments into a perfluorocarbon phase. Metal cation responsive tweezers In the early 2000s, Lehn and co
  • -workers introduced a switchable system based on a terpyridine (terpy) ligand (Figure 8), which is structurally similar to the diphenylpyridine units used by Zimmerman in rigid clips, but can change their conformation upon complexation by a metal cation [32][33]. When substituted at the 6 and 6" positions
PDF
Album
Review
Published 01 Mar 2024

Ligand effects, solvent cooperation, and large kinetic solvent deuterium isotope effects in gold(I)-catalyzed intramolecular alkene hydroamination

  • Ruichen Lan,
  • Brock Yager,
  • Yoonsun Jee,
  • Cynthia S. Day and
  • Amanda C. Jones

Beilstein J. Org. Chem. 2024, 20, 479–496, doi:10.3762/bjoc.20.43

Graphical Abstract
  • data plots; where not visible they are smaller than the icon for the data point. (a) Schematic for synthesis of [L–Au–L]SbF6 where L = JPhos. (b) Perspective drawing of the cation in crystalline [Au(P(C4H9)2(C12H9))2](SbF6)CH2Cl2 where P are represented by dotted spheres, Au atoms are represented by
PDF
Album
Supp Info
Full Research Paper
Published 29 Feb 2024

Pseudallenes A and B, new sulfur-containing ovalicin sesquiterpenoid derivatives with antimicrobial activity from the deep-sea cold seep sediment-derived fungus Pseudallescheria boydii CS-793

  • Zhen Ying,
  • Xiao-Ming Li,
  • Sui-Qun Yang,
  • Hong-Lei Li,
  • Xin Li,
  • Bin-Gui Wang and
  • Ling-Hong Meng

Beilstein J. Org. Chem. 2024, 20, 470–478, doi:10.3762/bjoc.20.42

Graphical Abstract
  • pathway, the bergamotene sesquiterpenoid (I) is presumed to be a key intermediate cyclized from farnesyl diphosphate (FPP) via nerolidyl diphosphate (NPP) followed by a bisabolyl cation [14]. Subsequent oxidation (bishydroxylation) catalyzed by some oxygenase such as P450 would afford the key intermediate
PDF
Album
Supp Info
Full Research Paper
Published 28 Feb 2024

Green and sustainable approaches for the Friedel–Crafts reaction between aldehydes and indoles

  • Periklis X. Kolagkis,
  • Eirini M. Galathri and
  • Christoforos G. Kokotos

Beilstein J. Org. Chem. 2024, 20, 379–426, doi:10.3762/bjoc.20.36

Graphical Abstract
  • /SiO2 [75], ZnCl2/SiO2 [76], heteropoly-11-tungsto-1-vanadophosphoric acid, H4[PVV W11O40] (HPV) (20%) supported on natural clay (HPVAC-20) [77], V2O5/SiO2 [78], strongly acidic cation exchange resin (Seralite SRC-120) [79] or HCl/SiO2 [80]. Among all the protocols mentioned above, it is worth
PDF
Album
Review
Published 22 Feb 2024

Mechanisms for radical reactions initiating from N-hydroxyphthalimide esters

  • Carlos R. Azpilcueta-Nicolas and
  • Jean-Philip Lumb

Beilstein J. Org. Chem. 2024, 20, 346–378, doi:10.3762/bjoc.20.35

Graphical Abstract
  • . Radical intermediate 9 formed upon fragmentation of 25, adds to the styrene acceptor forming radical 26. Finally, a radical-polar crossover event between 26 and the IrIV complex regenerates the IrIII ground state while delivering cation 27 that is then trapped by the oxygen-nucleophile to form the
  • radical-polar crossover affords cation 27 that delivers functionalized product 31 upon nucleophilic addition. The Doyle and Knowles groups reported the use of NHPI esters as radical precursors in the context of a radical redox annulation method [48] (Scheme 8A). This transformation occurs through an
  • aromatic ring, forming intermediate 41, which was then oxidized to cation 42, thereby completing the photocatalytic cycle. The reaction proceeded by regioselective nucleophilic addition of H2O, accompanied by the loss of MeOH to deliver spirocycle 43. Notably, the dearomative spirocyclization of biaryl
PDF
Album
Perspective
Published 21 Feb 2024

Elucidating the glycan-binding specificity and structure of Cucumis melo agglutinin, a new R-type lectin

  • Jon Lundstrøm,
  • Emilie Gillon,
  • Valérie Chazalet,
  • Nicole Kerekes,
  • Antonio Di Maio,
  • Ten Feizi,
  • Yan Liu,
  • Annabelle Varrot and
  • Daniel Bojar

Beilstein J. Org. Chem. 2024, 20, 306–320, doi:10.3762/bjoc.20.31

Graphical Abstract
  • and subsequent reverse Ni-NTA affinity chromatography resulted in significant co-purification of E. coli contaminants, necessitating an extra purification step, where cation exchange chromatography allowed us to obtain pure fractions of CMA16–291. Of note, this additional purification step was not
  • (Val6 to Asp132) could be modelled, and unambiguous electron density permitted us to locate and model four cation binding sites (three in each structure) and one sugar binding site (Figure 4a,b and Figure S4, Supporting Information File 2). The complexed structures allowed us to shed light on the
  • from the expression of CMA1 protein in CHO-S cells. Note the smeared band indicating the presence of glycosylation. (d) Recombinant expression of CMA1 in bacteria. SDS-PAGE gels of the His-tag affinity chromatography and cation exchange chromatography from the expression of CMA1 protein in E. coli BL21
PDF
Album
Supp Info
Full Research Paper
Published 19 Feb 2024

Additive-controlled chemoselective inter-/intramolecular hydroamination via electrochemical PCET process

  • Kazuhiro Okamoto,
  • Naoki Shida and
  • Mahito Atobe

Beilstein J. Org. Chem. 2024, 20, 264–271, doi:10.3762/bjoc.20.27

Graphical Abstract
  • phenoxide in PhOH, the PhOH molecule is included in the hydrogen bond network along with the tetrabutylammonium cation (Bu4N+) to form a large aggregate. The hydrogen bonding between the amide and phosphate base in the small aggregates was stronger than in the large aggregates, which significantly enhanced
PDF
Album
Supp Info
Full Research Paper
Published 12 Feb 2024

Substitution reactions in the acenaphthene analog of quino[7,8-h]quinoline and an unusual synthesis of the corresponding acenaphthylenes by tele-elimination

  • Ekaterina V. Kolupaeva,
  • Narek A. Dzhangiryan,
  • Alexander F. Pozharskii,
  • Oleg P. Demidov and
  • Valery A. Ozeryanskii

Beilstein J. Org. Chem. 2024, 20, 243–253, doi:10.3762/bjoc.20.24

Graphical Abstract
  • is easily protonated, and its protic salt has been fully characterized as tetrafluoroborate 14H+BF4−. The 1H NMR spectrum of this salt confirmed the symmetrical structure of the heterocyclic cation with a chelated intramolecular [NHN]+ bond, whose proton in CD3CN solution resonates at 17.22 ppm
  • comparison, in the protonated cation of the starting diazine 5 in CD3CN, the chemical shift of the “acidic” proton is observed at 18.02 ppm [15]. Next, we evaluated the pKa value of base 14 by a competitive method in acetonitrile (NMR transprotonation involving an equivalent amount of "proton sponge" 1 as a
  • dynamically symmetric in solution), however, due to the noticeably larger internitrogen distance in cation 8H+, the H bond in it is significantly weakened, as evidenced by a lower degree of deshielding of the chelated NH proton (cf. δNH values; see also data for protonated 14) and a shorter counterion–NH
PDF
Album
Supp Info
Full Research Paper
Published 08 Feb 2024

Comparison of glycosyl donors: a supramer approach

  • Anna V. Orlova,
  • Nelly N. Malysheva,
  • Maria V. Panova,
  • Nikita M. Podvalnyy,
  • Michael G. Medvedev and
  • Leonid O. Kononov

Beilstein J. Org. Chem. 2024, 20, 181–192, doi:10.3762/bjoc.20.18

Graphical Abstract
  • ][33][34][37][39][43][44][46]. During a study of a possible influence of remote acyl protective groups [23] on the sialylation outcome (which could become possible through participation in stabilization of glycosyl cation [24][47][48] in a conformation with all-axial substituents in the pyranose ring
  • versus α:β = 13:1 for 2). One could speculate that a more nucleophilic carbonyl oxygen of the chloroacetyl group at O-9 in sialyl donor 2 might participate in a stabilization of the intermediate glycosyl cation from the α-side (as we discussed earlier [52][53]) diminishing the α/β ratio. Conversely, at
  • putative participation in stabilization [52][53] of the glycosyl cation. More importantly, in our opinion, this result indicates the existence of unexpected difficulties in the determination of relative reactivities of glycosyl donors (vide infra). Discussion It is generally believed that the molecular
PDF
Album
Supp Info
Full Research Paper
Published 31 Jan 2024

Tandem Hock and Friedel–Crafts reactions allowing an expedient synthesis of a cyclolignan-type scaffold

  • Viktoria A. Ikonnikova,
  • Cristina Cheibas,
  • Oscar Gayraud,
  • Alexandra E. Bosnidou,
  • Nicolas Casaretto,
  • Gilles Frison and
  • Bastien Nay

Beilstein J. Org. Chem. 2024, 20, 162–169, doi:10.3762/bjoc.20.15

Graphical Abstract
  • alternative mechanism, we envisionned a possible interconversion of 7 and 7’ through a [1,5]-sigmatropic rearrangement resulting in a hydrogen and cation shift towards 7’ (Scheme 4). To test this hypothesis, this rearrangement was computed at the DFT level. A cyclic transition state (TS) was found between
  • reactions. Highly nucleophilic arenes like 1,3,5-trimethoxybenzene react easily under mild conditions and result in a stabilized benzylic cation in acidic conditions, allowing a second intramolecular Friedel–Crafts reaction involving the aryl substituent of the substrate. These reactions are favored by π
PDF
Album
Supp Info
Full Research Paper
Published 25 Jan 2024
Graphical Abstract
  • lifetime of 71 ps in toluene. Meanwhile, the absorption at 640 nm, corresponding to the ZnP radical cation, was not observed. The results obtained from Rehm–Weller’s equation [146] suggested that photoinduced electron transfer was thermodynamically permitted in 86. However, ultrafast energy transfer from
  • , the emergence of the weak absorption band of the ZnP radical cation at 640 nm was hindered by the overwhelming absorption intensity of the residual porphyrin. The transient absorption spectra of 87 were obtained in benzonitrile solvent, which was expected to stabilize the CS state. Absorption
  • corresponding to the ZnP radical cation was clearly observed with a lifetime of 2.3 μs. The occurrence of such a long-living CS state can be rationally associated with the Marcus-inverted-region [143] behavior of the charge-recombination process. For 88, which has no spacer between ZnP and TCBD, as opposed to
PDF
Album
Review
Published 22 Jan 2024

Visible-light-induced radical cascade cyclization: a catalyst-free synthetic approach to trifluoromethylated heterocycles

  • Chuan Yang,
  • Wei Shi,
  • Jian Tian,
  • Lin Guo,
  • Yating Zhao and
  • Wujiong Xia

Beilstein J. Org. Chem. 2024, 20, 118–124, doi:10.3762/bjoc.20.12

Graphical Abstract
  • formed radical 7 can be oxidized by 2a or 4 giving a cation 8, which undergoes a deprotonation process and formation of the desired product. Conclusion In conclusion, we have developed a visible-light-promoted protocol for the synthesis of dihydropyrido[1,2-a]indolones bearing a trifluoromethyl group at
PDF
Album
Supp Info
Full Research Paper
Published 19 Jan 2024

Electron-beam-promoted fullerene dimerization in nanotubes: insights from DFT computations

  • Laura Abella,
  • Gerard Novell-Leruth,
  • Josep M. Ricart,
  • Josep M. Poblet and
  • Antonio Rodríguez-Fortea

Beilstein J. Org. Chem. 2024, 20, 92–100, doi:10.3762/bjoc.20.10

Graphical Abstract
  • and reversible process named phase 1. We find that the barriers for the radical cation mechanism are significantly lower than those found for the neutral pathway. The peapod is mainly providing one-dimensional confinement for the reaction to take place in a more efficient way. Car–Parrinello
  • the reaction either via singlet excitation or via radical cation formation (Scheme 1). Estimation of the activation barrier for the [2 + 2] cycloaddition when the nanotube acts as a sensitizer is 33.5 ± 6.8 kJ mol−1. This value agrees with computational predictions for the reaction via an excited
  • can also be activated through the formation of C60+• radical cation [3][9]. This mechanistic proposal for phase 1, which to our knowledge has not yet been explored in detail inside a carbon nanotube, is analyzed here and compared to the non-activated C60 dimerization. Finally, some intermediates for
PDF
Album
Supp Info
Full Research Paper
Published 17 Jan 2024

Multi-redox indenofluorene chromophores incorporating dithiafulvene donor and ene/enediyne acceptor units

  • Christina Schøttler,
  • Kasper Lund-Rasmussen,
  • Line Broløs,
  • Philip Vinterberg,
  • Ema Bazikova,
  • Viktor B. R. Pedersen and
  • Mogens Brøndsted Nielsen

Beilstein J. Org. Chem. 2024, 20, 59–73, doi:10.3762/bjoc.20.8

Graphical Abstract
  • Tetrathiafulvalene (TTF, Figure 1) is a redox-active molecule that has been widely explored in materials chemistry and supramolecular chemistry [1][2][3][4][5][6][7][8]. TTF reversibly undergoes two sequential one-electron oxidations, generating first a radical cation (TTF+•) and subsequently a dication (TTF2
  • rate: 0.1 V/s. All potentials are depicted against the Fc/Fc+ redox couple. Radical anion (left), dianion (middle), and radical cation (right) of compound 23; the radical anion has a 14πz-aromatic ring (highlighted in blue; only counting 2π-electrons of each triple bond, here defined as those in πz
  • orbitals), the dianion has an additional 6π-aromatic cyclopentadienyl anion (highlighted in green), while the cation has a 6π-aromatic 1,3-dithiolium ring (highlighted in red). ORTEP plots (50% probability) and crystal packing of compounds a) 25, b) 26, and c) 29. The respective crystal packing of each
PDF
Album
Supp Info
Full Research Paper
Published 15 Jan 2024

Using the phospha-Michael reaction for making phosphonium phenolate zwitterions

  • Matthias R. Steiner,
  • Max Schmallegger,
  • Larissa Donner,
  • Johann A. Hlina,
  • Christoph Marschner,
  • Judith Baumgartner and
  • Christian Slugovc

Beilstein J. Org. Chem. 2024, 20, 41–51, doi:10.3762/bjoc.20.6

Graphical Abstract
  • between P1 and C15 is slightly longer (1.824(2) Å in 2a; 1.828(3) in 2f) when compared to the P–CH2 distance of a tetra-n-butylphosphonium cation [45]. UV–vis spectroscopy All phosphonium phenolate compounds exhibit a bright yellow color in solution (see inset in Figure 3). Investigating the absorption
PDF
Album
Supp Info
Full Research Paper
Published 10 Jan 2024

1-Butyl-3-methylimidazolium tetrafluoroborate as suitable solvent for BF3: the case of alkyne hydration. Chemistry vs electrochemistry

  • Marta David,
  • Elisa Galli,
  • Richard C. D. Brown,
  • Marta Feroci,
  • Fabrizio Vetica and
  • Martina Bortolami

Beilstein J. Org. Chem. 2023, 19, 1966–1981, doi:10.3762/bjoc.19.147

Graphical Abstract
  • alkynes was studied in the ionic liquid 1-butyl-3-methylimidazolium tetrafluoroborate (BMIm-BF4) adding boron trifluoride diethyl etherate (BF3·Et2O) as catalyst. Different ionic liquids were used, varying the cation or the anion, in order to identify the best one, in terms of both efficiency and reduced
  • electrogeneration of BF3, the aim of this work was to explore the hydration of alkynes using ILs as reaction medium and BF3 as catalyst. First of all, we investigated the behaviour of diphenylacetylene in BMIm-BF4 containing BF3·Et2O. Then we evaluated the same reaction in different ILs, modifying the cation or the
  • ‒ hydrolysis in a 1:1 in volume IL/water solution kept at 45 °C for 24 h) [105]. It should be noted that the same treatment carried out on 1-methyl-3-octylimidazolium tetrafluoroborate (OMIm-BF4) evidenced a much higher extent of BF4‒ hydrolysis. This is probably due to the weaker interaction between cation
PDF
Album
Supp Info
Full Research Paper
Published 28 Dec 2023

Beyond n-dopants for organic semiconductors: use of bibenzo[d]imidazoles in UV-promoted dehalogenation reactions of organic halides

  • Kan Tang,
  • Megan R. Brown,
  • Chad Risko,
  • Melissa K. Gish,
  • Garry Rumbles,
  • Phuc H. Pham,
  • Oana R. Luca,
  • Stephen Barlow and
  • Seth R. Marder

Beilstein J. Org. Chem. 2023, 19, 1912–1922, doi:10.3762/bjoc.19.142

Graphical Abstract
  • the monomeric cation N-DMBI+ (Supporting Information File 1, Figure S15). Thus, the overall reaction is consistent with: Scope of reaction (Y-DMBI)2 with other benzyl, alkyl, and aryl halides Table 2 summarizes the conversions and product yields for the reactions of (N-DMBI)2 or (Cyc-DMBI)2 with
PDF
Album
Supp Info
Full Research Paper
Published 14 Dec 2023

Anion–π catalysis on carbon allotropes

  • M. Ángeles Gutiérrez López,
  • Mei-Ling Tan,
  • Giacomo Renno,
  • Augustina Jozeliūnaitė,
  • J. Jonathan Nué-Martinez,
  • Javier Lopez-Andarias,
  • Naomi Sakai and
  • Stefan Matile

Beilstein J. Org. Chem. 2023, 19, 1881–1894, doi:10.3762/bjoc.19.140

Graphical Abstract
  • ], followed by SWCNTs and MWCNTs two years later [13]. Particularly MWCNTs have the potential to couple anion–π and cation–π catalysis with electric-field-assisted catalysis [14]. While anion–π (and cation–π [15][16]) catalysis, compared to other unorthodox interactions, has been less impactful than expected
  • available so far on anion–π catalysis on carbon allotropes. Review Anion–π catalysis on fullerenes The use of fullerenes in catalysis is surprisingly underdeveloped [45][46][47][48][49][50][51]. Anion–π and cation–π interactions on fullerenes attract similarly little attention until today [52][53][54][55
  • –16 was as in the amide series from 8 and thus supported entropic contributions to anion–π catalysis. Steric increase of the secondary amide in 17 impeded anion–π catalysis, presumably because the catalytic π surface next to the ammonium cation became inaccessible for anions paired with the tethered
PDF
Album
Review
Published 12 Dec 2023

Recent advancements in iodide/phosphine-mediated photoredox radical reactions

  • Tinglan Liu,
  • Yu Zhou,
  • Junhong Tang and
  • Chengming Wang

Beilstein J. Org. Chem. 2023, 19, 1785–1803, doi:10.3762/bjoc.19.131

Graphical Abstract
  • a wide range of transformations [7][8]. Moreover, they proposed a plausible mechanism for the aforementioned conversions (Scheme 2). Initially, an NaI/PPh3 complex I was formed through a cation–π interaction. Subsequently, the combination of complex I with N-(cyclohexanecarbonyloxy)phthalimide
PDF
Album
Review
Published 22 Nov 2023
Other Beilstein-Institut Open Science Activities