Search results

Search for "decarboxylation" in Full Text gives 206 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

(Bio)isosteres of ortho- and meta-substituted benzenes

  • H. Erik Diepers and
  • Johannes C. L. Walker

Beilstein J. Org. Chem. 2024, 20, 859–890, doi:10.3762/bjoc.20.78

Graphical Abstract
  • decarboxylation then yields 1,2-cubane 88. This synthesis reduced the number of synthetic steps from eight, in the previously known patented synthesis from 2007 [54], to four. MacMillan and co-workers also developed a number of decarboxylative cross-coupling reactions to allow access to an even wider range of 1,2
  • ), deprotection, decarboxylation, oxidation, nitrogen extrusion (to cyclobutadiene) and Diels–Alder reaction yielded annulated tricycle 164. Further intramolecular [2 + 2] cycloaddition formed cubane precursor 165. From diketone 165, 1,3-cubane 166 was obtained by Favorskii ring contraction followed by
PDF
Album
Review
Published 19 Apr 2024

Chemical and biosynthetic potential of Penicillium shentong XL-F41

  • Ran Zou,
  • Xin Li,
  • Xiaochen Chen,
  • Yue-Wei Guo and
  • Baofu Xu

Beilstein J. Org. Chem. 2024, 20, 597–606, doi:10.3762/bjoc.20.52

Graphical Abstract
  • pathway (Figure 5). Briefly, the prenyl group is attached to tryptophan through a prenylation reaction catalyzed by ShnA, followed by the decarboxylation of the carboxy group by ShnC. Subsequently, compound 2 is formed by the addition of succinimide to N15 in 1 via a reaction catalyzed by ShnB. The
PDF
Album
Supp Info
Full Research Paper
Published 15 Mar 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
  • the αKG-dependent dioxygenase have been analyzed in detail due to its relatively small molecular weight and the low costs of its cofactors: αKG, ascorbic acid, and iron ions. The αKG reacts with iron and molecular oxygen to form the highly reactive Fe(IV)=O via oxidative decarboxylation. This active
PDF
Album
Review
Published 13 Mar 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
  • [25][26] (Scheme 1) and have found applications in a number of functional group interconversions mediated by radical chain decarboxylation [27]. However, their widespread use in synthesis, especially in complex molecular settings, suffers from significant disadvantages. These include thermal and
  • (PC•–) (Scheme 4A). This strong reducing agent mediates the one-electron reduction of the NHPI ester 10, forming radical anion intermediate 11. Fragmentation of 11 via N–O bond homolysis and decarboxylation forms the key tertiary radical 12 with concomitant formation of phthalimidyl anion (–Nphth) and
  • irradiation, EDA complex 74 triggered a radical chain process initiated by B–B bond cleavage, forming boryl-NHPI ester radical 75 and boryl radical 76. Subsequent decarboxylation of 75 yields carbon-centered radical 9 and boryl-phthalimide byproduct 77. Meanwhile DMA-ligated B2cat2 78 is formed upon
PDF
Album
Perspective
Published 21 Feb 2024

Synthesis of π-conjugated polycyclic compounds by late-stage extrusion of chalcogen fragments

  • Aissam Okba,
  • Pablo Simón Marqués,
  • Kyohei Matsuo,
  • Naoki Aratani,
  • Hiroko Yamada,
  • Gwénaël Rapenne and
  • Claire Kammerer

Beilstein J. Org. Chem. 2024, 20, 287–305, doi:10.3762/bjoc.20.30

Graphical Abstract
  • bronze in refluxing quinoline, thus triggering a S-extrusion along with a decarboxylation reaction (Scheme 3) [56]. From a retrosynthetic point of view, this example nicely illustrates the fact that polyannelated thiepines, and more generally S-, Se- and Te-based heteropines, are straightforward
  • chemically-induced S-extrusion (and concomitant decarboxylation) from a dibenzothiepine precursor [56]. Top: Conversion of dinaphthothiepine bisimides 3a,b and their sulfoxide analogues 4a,b into PBIs 6a,b by S-extrusion triggered by electron injection, photo- and thermal activation. Bottom
PDF
Album
Review
Published 15 Feb 2024

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
  • concert with sacrificial weak reductants (Figure 1b) [8][9][10][11]. Another approach is to add ambient-stable precursors to reaction mixtures: for example, reducing Wanzlick dimers and related species (Figure 1a, ii) have been formed from precursors through in situ decarboxylation [12] or deprotonation
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
  • forward. Decarboxylation of the resulting intermediate IV then affords the chiral addition product 6. This enolate addition is in kinetic competition with simple decarboxylation, yielding thioacetate 7. Under most conditions, this decarboxylation is favored. Anion–π catalysis selectively accelerates the
  • intrinsically disfavored but significant enolate addition by stabilizing the planar sp2 intermediate II that has to add before decarboxylation can occur [2][61]. The non-planar sp3 keto intermediate I that can decarboxylate through intermediate V without preceding enolate addition is less stabilized on the
  • planar π surfaces of anion–π catalysts. The A/D product ratio is thus a convenient measure for anion–π catalysis, the larger the better, with A/D > 1, the intrinsic selectivity for decarboxylation has been inverted [2][3][18]. Applying lessons from simpler anion–π catalysts, trialkylamines were 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
  • smoothly delivered an electron donor–acceptor (EDA) complex II via coulombic interactions. Upon 456 nm blue LED light irradiation, the EDA complex II underwent a single electron transfer (SET) process, followed by subsequent decarboxylation to produce the alkyl radical intermediate A, accompanied by
  • facilitated by the process of photoexcited radical decarboxylation. On the other hand, the copper catalytic cycle involved the capture of alkyl radicals by the copper complex B, the activation of heteroatom-containing substrates 30 by a base-mediated proton transfer, and the subsequent reductive elimination
  •  15) [27]. This method involved a series of steps, including the formation of an EDA complex, decarboxylation, radical addition, C–H functionalization, and annulation. Various primary, secondary, and tertiary alkyl N-hydroxyphthalimide esters 33 showed potential as viable substrates for the synthesis
PDF
Album
Review
Published 22 Nov 2023

Decarboxylative 1,3-dipolar cycloaddition of amino acids for the synthesis of heterocyclic compounds

  • Xiaofeng Zhang,
  • Xiaoming Ma and
  • Wei Zhang

Beilstein J. Org. Chem. 2023, 19, 1677–1693, doi:10.3762/bjoc.19.123

Graphical Abstract
  • -type AMYs in multicomponent, one-pot, and stepwise reactions for the synthesis of diverse heterocycles related to some bioactive compounds and natural products. Keywords: [3 + 2] cycloaddition; decarboxylation; 1,3-dipolar; double cycloaddition; one-pot synthesis; multicomponent reaction; semi
  • aldehydes and α-amino esters (via dehydration) or α-amino acids (via decarboxylation) could be classified based on the substitution groups on the N atom to: 1) N-substituted (N–R type), 2) hydrogen containing (N–H type), and 3) metal complexes (N–M type) (Figure 1) [16][17]. These AMYs could also be
PDF
Album
Perspective
Published 06 Nov 2023

Synthesis of ether lipids: natural compounds and analogues

  • Marco Antônio G. B. Gomes,
  • Alicia Bauduin,
  • Chloé Le Roux,
  • Romain Fouinneteau,
  • Wilfried Berthe,
  • Mathieu Berchel,
  • Hélène Couthon and
  • Paul-Alain Jaffrès

Beilstein J. Org. Chem. 2023, 19, 1299–1369, doi:10.3762/bjoc.19.96

Graphical Abstract
  • hypotensive effects. First, Wissner et al. reported the synthesis of racemic sn-1-deoxy-PAF 11.8 (Figure 11) [91]. First, n-octadecanoic acid chloride (11.1) reacted with tris[(trimethylsilyl)oxy]ethylene (11.2) [92] to produce, after acidic hydrolysis and subsequent decarboxylation, compound 11.3. Then, the
PDF
Album
Review
Published 08 Sep 2023

Radical ligand transfer: a general strategy for radical functionalization

  • David T. Nemoto Jr,
  • Kang-Jie Bian,
  • Shih-Chieh Kao and
  • Julian G. West

Beilstein J. Org. Chem. 2023, 19, 1225–1233, doi:10.3762/bjoc.19.90

Graphical Abstract
  • , such as hydrogen atom transfer (HAT), alkene addition, and decarboxylation. At least as important has been innovation in radical functionalization methods, including radical–polar crossover (RPC), enabling these intermediates to be engaged in productive and efficient bond-forming steps. However, direct
  • decarboxylation. Here, we provide an overview of the evolution of RLT catalysis from initial studies to recent advances and provide a conceptual framework we hope will inspire and enable future work using this versatile elementary step. Keywords: catalysis; cooperative catalysis; earth abundant elements
  • ], and decarboxylation [7], enabling these intermediates to be easily accessed from diverse starting materials. Functionalization methods have also seen significant development, with elementary steps such as radical–polar crossover (RPC) finding significant purchase [8]; however, these steps are not
PDF
Album
Perspective
Published 15 Aug 2023

Photoredox catalysis harvesting multiple photon or electrochemical energies

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

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

Graphical Abstract
PDF
Album
Review
Published 28 Jul 2023

Photoredox catalysis enabling decarboxylative radical cyclization of γ,γ-dimethylallyltryptophan (DMAT) derivatives: formal synthesis of 6,7-secoagroclavine

  • Alessio Regni,
  • Francesca Bartoccini and
  • Giovanni Piersanti

Beilstein J. Org. Chem. 2023, 19, 918–927, doi:10.3762/bjoc.19.70

Graphical Abstract
  • proton transfer from the oxidized indole radical cation [75], generated by SET from the activated photocatalyst. The α-amino radical generated by reductive decarboxylation of a DMAT derivative with a redox-active ester (−1.26 V to −1.37 V vs a saturated calomel electrode) would enable turnover of the
  • photoredox cycle (Figure 1b). Alternatively, we envisioned a more established approach expecting the direct oxidative photoredox decarboxylation of the carboxylic acid/carboxylate (by SET from the activated photocatalyst) of DMAT to generate the α-aminoalkyl radical that might readily be captured/trapped
  • radical is stabilized by both the indole ring and the Δ2-olefin. Next, the resonance-stabilized radical intermediate III was trapped by the active α-aminoalkyl radical, generated by reductive decarboxylation by Ir(II) produced in the photocatalytic cycle (which undergoes oxidation to afford the Ir(III
PDF
Album
Supp Info
Full Research Paper
Published 26 Jun 2023

Pyridine C(sp2)–H bond functionalization under transition-metal and rare earth metal catalysis

  • Haritha Sindhe,
  • Malladi Mounika Reddy,
  • Karthikeyan Rajkumar,
  • Akshay Kamble,
  • Amardeep Singh,
  • Anand Kumar and
  • Satyasheel Sharma

Beilstein J. Org. Chem. 2023, 19, 820–863, doi:10.3762/bjoc.19.62

Graphical Abstract
  • (Scheme 30). The reaction showed good compatibility with various functional groups. The proposed mechanism (Scheme 30b) involves the silver-catalyzed decarboxylation of heteroaryl acid 156 followed by transmetalation providing palladium intermediate 160. Further, C–H activation of pyridine N-oxide 9
PDF
Album
Review
Published 12 Jun 2023

Strategies in the synthesis of dibenzo[b,f]heteropines

  • David I. H. Maier,
  • Barend C. B. Bezuidenhoudt and
  • Charlene Marais

Beilstein J. Org. Chem. 2023, 19, 700–718, doi:10.3762/bjoc.19.51

Graphical Abstract
  • the enol carboxylate and subsequent 1,2 radical rearrangement and decarboxylation. Moderate to good yields of dibenzo[b,f]oxepine carboxylates 25 were achieved (63–85%). Stopka et al. [46] reported on tandem C–H functionalisation and ring expansion as an alternative to the Wagner–Meerwin rearrangement
PDF
Album
Review
Published 22 May 2023

Strategies to access the [5-8] bicyclic core encountered in the sesquiterpene, diterpene and sesterterpene series

  • Cécile Alleman,
  • Charlène Gadais,
  • Laurent Legentil and
  • François-Hugues Porée

Beilstein J. Org. Chem. 2023, 19, 245–281, doi:10.3762/bjoc.19.23

Graphical Abstract
  • formation of the fused cyclobutane acid 162 as the desired precursor for the cyclooctane formation. The ring expansion was achieved in the presence of an iridium catalyst and under blue LED irradiation, via the trapping by TEMPO or O2 of the cyclobutyl radical resulting from decarboxylation, which allowed a
PDF
Album
Review
Published 03 Mar 2023

Combining the best of both worlds: radical-based divergent total synthesis

  • Kyriaki Gennaiou,
  • Antonios Kelesidis,
  • Maria Kourgiantaki and
  • Alexandros L. Zografos

Beilstein J. Org. Chem. 2023, 19, 1–26, doi:10.3762/bjoc.19.1

Graphical Abstract
  • –Giese coupling, followed by reductive cleavage of the lactone moiety with LiI. Enzymatic hydroxylation by the BM3 MERO1 variant worked equally well to provide the 3-hydroxylated product 46. Photochemical radical decarboxylation of the formed mercaptopyridine derivative and radical capture by iodoform
PDF
Album
Review
Published 02 Jan 2023

One-pot double annulations to confer diastereoselective spirooxindolepyrrolothiazoles

  • Juan Lu,
  • Bin Yao,
  • Desheng Zhan,
  • Zhuo Sun,
  • Yun Ji and
  • Xiaofeng Zhang

Beilstein J. Org. Chem. 2022, 18, 1607–1616, doi:10.3762/bjoc.18.171

Graphical Abstract
  • . Subsequent decarboxylation of thiazolooxazol-1-one I affords non-stabilized azomethine ylide (AY) for 1,3-dipolar cycloaddition with olefinic oxindole 4a to give spirooxindolepyrrolothiazoles 5 and 7. The endo-TS is more favorable than exo-TS for the 1,3-dipolar cycloaddition to afford the major and minor
PDF
Album
Supp Info
Full Research Paper
Published 28 Nov 2022

Isolation and biosynthesis of daturamycins from Streptomyces sp. KIB-H1544

  • Yin Chen,
  • Jinqiu Ren,
  • Ruimin Yang,
  • Jie Li,
  • Sheng-Xiong Huang and
  • Yijun Yan

Beilstein J. Org. Chem. 2022, 18, 1009–1016, doi:10.3762/bjoc.18.101

Graphical Abstract
  • ) probably undergoes rearrangement, decarboxylation, methylation reaction, and further modification to generate compounds 1 and 2. Conclusion S. sp. KIB-H1544 was isolated from the rhizosphere soil of Datura stramonium L. Two novel diarylcyclopentenones daturamycins A and B and one new p-terphenyl
PDF
Album
Supp Info
Full Research Paper
Published 09 Aug 2022

Automated grindstone chemistry: a simple and facile way for PEG-assisted stoichiometry-controlled halogenation of phenols and anilines using N-halosuccinimides

  • Dharmendra Das,
  • Akhil A. Bhosle,
  • Amrita Chatterjee and
  • Mainak Banerjee

Beilstein J. Org. Chem. 2022, 18, 999–1008, doi:10.3762/bjoc.18.100

Graphical Abstract
  • dihalogenations. As known, the decarboxylation (or desulfonation) was observed in the case of salicylic acids and anthranilic acids (or sulfanilic acids) leading to 2,4,6-trihalogenated products when 3 equiv of NXS was used. Simple instrumentation, metal-free approach, cost-effectiveness, atom economy, short
  • aniline (entry 5, Table 2) afforded the 2,4,6-tribromo derivatives in the presence of NBS in excellent yields just by grinding for 5 min. Similarly, 3 equiv of NIS ensured the formation of 2,4,6-triiodo derivatives in over 90% yields (entries 8 and 11, Table 2). As known, easy decarboxylation (or
  • halogenation without much substituent effect and by just controlling the stoichiometry of NXS a series of mono-, di-, and trihalogenated phenols and anilines were obtained in a chemoselective manner in good to excellent yields within 2–15 min of grinding. Spontaneous decarboxylation (or desulfonation) was
PDF
Album
Supp Info
Full Research Paper
Published 09 Aug 2022

Morita–Baylis–Hillman reaction of 3-formyl-9H-pyrido[3,4-b]indoles and fluorescence studies of the products

  • Nisha Devi and
  • Virender Singh

Beilstein J. Org. Chem. 2022, 18, 926–934, doi:10.3762/bjoc.18.92

Graphical Abstract
  • ʟ-tryptophan (1) was much faster than with the tryptophan ester, taking only 45 minutes to complete. Interestingly, KMnO4 oxidation was selective, with no decarboxylation seen. Within 15 minutes, further treatment of 3a–e with methyl iodide in the presence of K2CO3 provided the corresponding methyl
PDF
Album
Supp Info
Full Research Paper
Published 26 Jul 2022

Synthesis of odorants in flow and their applications in perfumery

  • Merlin Kleoff,
  • Paul Kiler and
  • Philipp Heretsch

Beilstein J. Org. Chem. 2022, 18, 754–768, doi:10.3762/bjoc.18.76

Graphical Abstract
  • a Z/E ratio of 95:5. Additionally, the authors developed an alternative synthesis of civetone (69) by metathesis of ethyl 9-decenoate and subsequent Dieckmann cyclization in flow, followed by a saponification and decarboxylation process in batch providing (Z)-civetone in 48% yield over three steps
PDF
Album
Review
Published 27 Jun 2022

Structural basis for endoperoxide-forming oxygenases

  • Takahiro Mori and
  • Ikuro Abe

Beilstein J. Org. Chem. 2022, 18, 707–721, doi:10.3762/bjoc.18.71

Graphical Abstract
  • ) [64][65][66][67]. The Fe(II) is coordinated by the conserved 2-His-1-Asp residues and a 2OG in the active site. The binding of a substrate in the active site, where it is close to the Fe(II) center, provides a coordination site for O2. Subsequently, the oxidative decarboxylation of 2OG generates a
PDF
Album
Review
Published 21 Jun 2022

Inductive heating and flow chemistry – a perfect synergy of emerging enabling technologies

  • Conrad Kuhwald,
  • Sibel Türkhan and
  • Andreas Kirschning

Beilstein J. Org. Chem. 2022, 18, 688–706, doi:10.3762/bjoc.18.70

Graphical Abstract
  • 87%). Pericyclic reactions such as the Diels–Alder and hetero-Diels–Alder cycloadditions, the Alder-En reaction, as well as the decarboxylation of α-alkylated malonic acids, are also suitable for flow protocols in combination with inductive heating (Scheme 10, case A) [53]. The yields but especially
  • decarboxylation of the malonic acid derivative 42 to give pent-4-enecarboxylic acid (43). In many cases, the flow protocol provided improved yields compared to the corresponding batch syntheses. Palladium-catalyzed cross-coupling reactions require higher temperatures and thus can be realized in an inductively
PDF
Album
Review
Published 20 Jun 2022

Heteroleptic metallosupramolecular aggregates/complexation for supramolecular catalysis

  • Prodip Howlader and
  • Michael Schmittel

Beilstein J. Org. Chem. 2022, 18, 597–630, doi:10.3762/bjoc.18.62

Graphical Abstract
  • )] (k298 = 32.2 kHz) move across the deck 82 and prevent binding of the protonated DABCO at the ZnPor binding sites. Use of the fuel acid 112 or 113 instead of applying the TFA/DBU acid/base combination leads initially to protonation of DABCO but due to the decarboxylation of 114 the resulting strong base
PDF
Album
Review
Published 27 May 2022
Other Beilstein-Institut Open Science Activities