Search results

Search for "C" in Full Text gives 3991 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

Azide–alkyne cycloaddition (click) reaction in biomass-derived solvent CyreneTM under one-pot conditions

  • Zoltán Medgyesi and
  • László T. Mika

Beilstein J. Org. Chem. 2025, 21, 1544–1551, doi:10.3762/bjoc.21.117

Graphical Abstract
  • it is negative in the Ames test [25]. Recently, we determined key physicochemical properties of CyreneTM and showed that it has a negligible vapor pressure (<9.6 kPa) and low viscosity (<6.8 mPa·s) at typical transition-metal-catalyzed reaction temperatures (30–140 °C) [26]. CyreneTM has been
  • , a wide range of organic reactions, e.g., urea and amide formation [32][33], amide coupling [34], aldol condensation [35], C–H difluoromethylation [36], aromatic substitution [37], and MOFs synthesis [38] were demonstrated in CyreneTM. Very recently, Fasano and Citarella first reported a CuAAC
  • ) were tested in the conversion of 1.15 mmol benzyl azide (1a) and 1 mmol phenylacetylene (2a) in 2.5 mL CyreneTM at 30 °C. All the selected Cu salts catalyzed the cycloaddition; however, Cu chlorides and oxides resulted in unexpectedly low product yields after 0.5 h. The solubility of Cu oxides was
PDF
Album
Supp Info
Full Research Paper
Published 30 Jul 2025

General method for the synthesis of enaminones via photocatalysis

  • Paula Pérez-Ramos,
  • Raquel G. Soengas and
  • Humberto Rodríguez-Solla

Beilstein J. Org. Chem. 2025, 21, 1535–1543, doi:10.3762/bjoc.21.116

Graphical Abstract
  • -dimethylformamide (DMF) at 20 °C afforded the best results, giving the desired product 9a in 70% isolated yield (Table 1, entry 1). Changing NiBr2·diglyme to other nickel salts, such as Ni(OTf)2 and NiCl2·diglyme led to lower yields (Table 1, entries 2 and 3). Similarly, changing the ligand for dtbbpy or
  • . Simultaneously, acridinium photocatalyst PC1 absorbed energy and transitioned from the ground state to excited state under visible-light irradiation. This excited state PC1* is quenched by the amine, generating the amine radical cation and PC1 radical via a single-electron transfer (SET) process. Then, the C−Br
  • bond of I is cleaved by PC1•, generating a C-centered radical II, which can be further reduced to give an enolate III, that ultimately evolves to the more stable anion IV and undergoes protonation to afford the final enaminone product 9a. Conclusion In summary, a simple and effective nickel-assisted
PDF
Album
Supp Info
Letter
Published 29 Jul 2025

Azobenzene protonation as a tool for temperature sensing

  • Antti Siiskonen,
  • Sami Vesamäki and
  • Arri Priimagi

Beilstein J. Org. Chem. 2025, 21, 1528–1534, doi:10.3762/bjoc.21.115

Graphical Abstract
  • is assumed. Stronger electron donation gives rise to larger red-shift, and hence more pronounced color changes, upon protonation, as illustrated in Figure 1C. Figure 1D illustrates the shift in protonation equilibrium with increasing acidity when MSA is added to DCE solution of 3 at 25 °C. Due to the
  • with 1.44 mM MSA over a 10–80 °C temperature range. Comparing the maximum absorbances at the neutral and protonated peaks provides a measure of sensitivity and optimal working range. Figure 2B presents this comparison for 1–3 in MSA/DCE solutions, calculated as the ratio between neutral and protonated
  • , the rate of change should be both high and systematic. For example, in Figure 2B, the optimal range for 3 appears to be between 30 and 80 °C, where the rate of change is fastest and follows a clean, exponential trend. All three compounds show similar sensitivity in DCE, while the sensitivity of 3 is
PDF
Album
Supp Info
Full Research Paper
Published 28 Jul 2025

Calcium waste as a catalyst in the transesterification for demanding esters: scalability perspective

  • Anton N. Potorochenko and
  • Konstantin S. Rodygin

Beilstein J. Org. Chem. 2025, 21, 1520–1527, doi:10.3762/bjoc.21.114

Graphical Abstract
  • acetylene production, was investigated. The catalyst was obtained by calcination of calcium carbide slag at 600 °C (CS600) and characterized by XRD and FTIR analysis. The transesterification reactions were carried out with primary alcohols, producing fatty acid alkyl esters in 51–99% yields, depending on
  • was calcined at 600 °C before use and the catalyst CS600 was characterized using XRD and FTIR analysis, confirming the presence of CaO as the main phase. The primary alcohols successfully reacted in the transesterification reaction to give the corresponding fatty acid alkyl ester mixtures in yields
  • fine organic synthesis. Results and Discussion The carbide slag was prepared by hydrolysis of commercially available calcium carbide (Sigma-Aldrich) and oven dried at 80 °C for 3 hours. The active catalyst CS600 was prepared by calcining the carbide slag at 600 °C for 2 hours. The composition of the
PDF
Album
Supp Info
Full Research Paper
Published 28 Jul 2025

Ambident reactivity of enolizable 5-mercapto-1H-tetrazoles in trapping reactions with in situ-generated thiocarbonyl S-methanides derived from sterically crowded cycloaliphatic thioketones

  • Grzegorz Mlostoń,
  • Małgorzata Celeda,
  • Marcin Palusiak,
  • Heinz Heimgartner,
  • Marta Denel-Bobrowska and
  • Agnieszka B. Olejniczak

Beilstein J. Org. Chem. 2025, 21, 1508–1519, doi:10.3762/bjoc.21.113

Graphical Abstract
  • five-membered heterocycles via [3 + 2]-cycloaddition reactions. Depending on the type of the dipolarophile used in these reactions, the target heterocycles may contain only one sulfur atom (cycloadditions with C=C or CC dipolarophiles) or more heteroatoms (cycloadditions with C=S, C=O, C=N, N=N, S=O
  • sulfur (χ = 2.55 (for C) and χ = 2.58 (for S) according to the Pauling scale), thiocarbonyl S-methanides 1 are considered as electron-rich 1,3-dipoles with basic and nucleophilic reactivity displayed by the =S+–CH2‒ unit [6]. Therefore, acidic compounds of type R–XH (X = NR’, O, S), which are able to
  • ], were analogously treated with CH2N2 at ca. 0 °C and the expected cycloadducts 2c and 2d were formed with complete regioselectivity, and subsequently could be isolated in good yields without a remarkable decomposition (Scheme 3). Thermal decomposition of 2a and 2b, and the behavior of the corresponding
PDF
Album
Supp Info
Full Research Paper
Published 23 Jul 2025

Highly distinguishable isomeric states of a tripodal arylazopyrazole derivative on graphite through electron/hole-induced switching at ambient conditions

  • Himani Malik,
  • Sudha Devi,
  • Debapriya Gupta,
  • Ankit Kumar Gaur,
  • Sugumar Venkataramani and
  • Thiruvancheril G. Gopakumar

Beilstein J. Org. Chem. 2025, 21, 1496–1507, doi:10.3762/bjoc.21.112

Graphical Abstract
  • intermediates. (c) All 8 states are accessible reversibly even at lower sample voltages, indicating that switching can be performed between any given state at a low threshold voltage. That is, all states are accessible and are remarkably stable at all sample voltages. It is concluded that the high energy
  • pyrolytic graphite (HOPG, ZYB grade from μMasch). Approximately 4 μL of solutions of FNAAP were drop-casted on the HOPG surface, dried (a few minutes) in ambient conditions, and AFM/STM studies were performed. Relative humidity (≈50%) and temperature (22–25 °C) of the room were controlled by an air
  • conditioner and a dehumidifier. While drop casting, the sample was kept at ≈20–30 °C to ensure a smooth flow of the solvent over the substrate. After drop casting, all films were subjected to vacuum pumping (typically 0.01 mbar) for about 1 h. Before depositing molecules, freshly cleaved graphite surfaces
PDF
Album
Supp Info
Full Research Paper
Published 22 Jul 2025

Heterologous biosynthesis of cotylenol and concise synthesis of fusicoccane diterpenoids

  • Ye Yuan,
  • Zhenhua Guan,
  • Xue-Jie Zhang,
  • Nanyu Yao,
  • Wenling Yuan,
  • Yonghui Zhang,
  • Ying Ye and
  • Zheng Xiang

Beilstein J. Org. Chem. 2025, 21, 1489–1495, doi:10.3762/bjoc.21.111

Graphical Abstract
  • ][23][24][25][26]. Most of these synthetic approaches rely on similar strategies, i.e., coupling of the A ring and the C ring followed by the formation of the B ring. Additionally, the semisynthesis of analogues of 1 has been reported and led to the discovery of ISIR-050 (4), which shows higher
  • activity than cotylenin A in cell growth inhibition assays and less toxicity in single-agent treatments [27][28]. Recently, Jiang and Renata described a chemoenzymatic approach that combines the skeletal construction by chemical methods and enzymatic C–H oxidations [29]. The synthesis employs a catalytic
  • Nozaki–Hiyama–Kishi reaction and a one-pot Prins cyclization/transannular hydride transfer to construct the 5-8-5 tricyclic scaffold. Enzymatic oxidations were used to install the hydroxy group at the C-3 position. Ten fusicoccanes were synthesized in 8–13 steps each. Despite these efforts, a strategy
PDF
Album
Supp Info
Letter
Published 21 Jul 2025

Copper catalysis: a constantly evolving field

  • Elena Fernández and
  • Jaesook Yun

Beilstein J. Org. Chem. 2025, 21, 1477–1479, doi:10.3762/bjoc.21.109

Graphical Abstract
  • straightforward reactions. Complementarily, the Review article by Jang and Kim provides a deep understanding of recent advances in the combination of electrochemistry and copper catalysis for various organic transformations [3]. Their contribution elaborates various C–H functionalizations, olefin additions
  • , decarboxylative functionalizations, and Chan–Lam coupling reactions. In doing so, the authors point out the combination of transition-metal catalysis and electrochemistry as an efficient, sustainable method for the oftentimes challenging formation of CC and C–heteroatom bonds in complex molecules. Another Review
  • -workers demonstrates that the efficient and green direct C–H amination of benzoxazoles can be catalyzed by copper chloride salts in acetonitrile in the absence of any acidic, basic, or oxidizing additives [8]. Both CuCl and CuCl2 were found to be extremely efficient in promoting the reactions, which
PDF
Editorial
Published 17 Jul 2025

Microwave-enhanced additive-free C–H amination of benzoxazoles catalysed by supported copper

  • Andrei Paraschiv,
  • Valentina Maruzzo,
  • Filippo Pettazzi,
  • Stefano Magliocco,
  • Paolo Inaudi,
  • Daria Brambilla,
  • Gloria Berlier,
  • Giancarlo Cravotto and
  • Katia Martina

Beilstein J. Org. Chem. 2025, 21, 1462–1476, doi:10.3762/bjoc.21.108

Graphical Abstract
  • offers wide-ranging potential for substrate expansion and the functionalisation of bioactive compounds. This study presents a green and efficient C–H amination, catalysed by CuCl and CuCl2, in acetonitrile without acidic, basic or oxidant additives that is accelerated by microwave (MW) irradiation and is
  • , and the huge overall economic impact of these processes. Inspired by the logic behind cross-dehydrogenative CC-coupling methods [15], the direct C–H amination has been developed as a more straightforward, economical and environmentally friendly reaction, compared to its counterparts (such as the
  • classical Buchwald–Hartwig amination reaction [16] or the Ullman reaction [17][18]) which require pre-functionalisation steps and harsh conditions [19]. Many protocols for C–H aminations have been applied to heteroaromatic compounds, mainly 5-membered heteroarenes, and a great deal of attention has been
PDF
Album
Supp Info
Full Research Paper
Published 15 Jul 2025

Wittig reaction of cyclobisbiphenylenecarbonyl

  • Taito Moribe,
  • Junichiro Hirano,
  • Hideaki Takano,
  • Hiroshi Shinokubo and
  • Norihito Fukui

Beilstein J. Org. Chem. 2025, 21, 1454–1461, doi:10.3762/bjoc.21.107

Graphical Abstract
  • ruled out. The structures of compounds 3, 4, and 5 were determined by X-ray diffraction analysis (Figure 2). Mono-olefin 3 and bis-olefin 5 adopt a bathtub-like chiral macrocyclic structure rather than figure-eight conformation. Both compounds crystallize as a racemic pair of enantiomers with P21/c and
  • low racemization barrier as with structurally similar methylenedioxy-substituted DBC derivative [22]. Racemization dynamics The racemization barriers of CBBC 1, mono-olefin 3, and bis-olefin 5 were evaluated by monitoring the decrease of circular dichroism (CD) signals in 1,2-dichlorobenzene at 170 °C
  • conformation (Sa,Sa)-A is feasible with a small activation barrier of 9.9 kcal mol−1. The figure-eight conformer (M,M)-B untwists to adopt an achiral conformation C with the exo-alkene units rotated inwards in opposite directions. These conformational changes are almost identical to those of CBBC 1. However
PDF
Album
Supp Info
Letter
Published 14 Jul 2025

Advances in nitrogen-containing helicenes: synthesis, chiroptical properties, and optoelectronic applications

  • Meng Qiu,
  • Jing Du,
  • Nai-Te Yao,
  • Xin-Yue Wang and
  • Han-Yuan Gong

Beilstein J. Org. Chem. 2025, 21, 1422–1453, doi:10.3762/bjoc.21.106

Graphical Abstract
  • 2a–c featuring a nitrogen-embedded cyclopenta[ef]heptalene core [15]. These compounds exhibit λabs at 363, 452, and 580 nm, and PLQYs of 0.05, 0.33, and 0.32, respectively. While compounds 2a and 2b display broad emission near 505 nm, 2c shows dual-emission peaks at 588 and 634 nm with an ultranarrow
  • extending into the near-infrared region, suggesting potential for redox-responsive chiral photonic systems. In 2023, Chen’s group reported three nitrogen–nitrogen (NN)-embedded azahelicenes 21a–c, among which compound 21c, a structurally defined antiaromatic double aza[7]helicene – exhibited distinctive
  • , thereby advancing their application in multifunctional chiral photonic and sensing platforms. In 2025, Gryko’s group synthesized a series of heterohelicenes 34a–c, featuring a 1,4-dihydropyrrolo[3,2-b]pyrrole (DHPP) core [48] (Table 10). The compounds exhibit similar absorption and emission profiles
PDF
Album
Review
Published 11 Jul 2025

Tautomerism and switching in 7-hydroxy-8-(azophenyl)quinoline and similar compounds

  • Lidia Zaharieva,
  • Vera Deneva,
  • Fadhil S. Kamounah,
  • Nikolay Vassilev,
  • Ivan Angelov,
  • Michael Pittelkow and
  • Liudmil Antonov

Beilstein J. Org. Chem. 2025, 21, 1404–1421, doi:10.3762/bjoc.21.105

Graphical Abstract
  • E/Z isomerization around the C=N bond leads to reduced efficiency of the tautomeric based switching [52]. In this respect, based on the excellent stability of the azo compounds, it is interesting to understand the effect of the competitive proton transfer and E/Z switching by replacement of the
  • reasons are different. While in acetonitrile this is the effect of the increased polarity of the solvent, leading to stabilization of the more polar keto tautomers, in the non-polar chloroform it is caused by its proton-donative nature (formation of a stabilizing intermolecular hydrogen bonding with the C
  • equilibrium towards C=O-group-containing structures. In addition, the spectra in the remaining solvents show a single band around 500 nm independent on the nature of the solvent. Obviously, these data cannot be interpreted in the frame of the concept for a tautomeric equilibrium. The rotor in 2 contains a OH
PDF
Album
Supp Info
Full Research Paper
Published 10 Jul 2025

Reactions of acryl thioamides with iminoiodinanes as a one-step synthesis of N-sulfonyl-2,3-dihydro-1,2-thiazoles

  • Vladimir G. Ilkin,
  • Pavel S. Silaichev,
  • Valeriy O. Filimonov,
  • Tetyana V. Beryozkina,
  • Margarita D. Likhacheva,
  • Pavel A. Slepukhin,
  • Wim Dehaen and
  • Vasiliy A. Bakulev

Beilstein J. Org. Chem. 2025, 21, 1397–1403, doi:10.3762/bjoc.21.104

Graphical Abstract
  • as a model for searching the optimal synthesis conditions (Table 1). We found that when processing thioamide 1a with PhINTs (2a, 1.5 equiv) in chloroform at 50 °C in the presence of Rh2(Piv)4 (0.5 equiv), 2,3-dihydro-N-sulfonyl-1,2-thiazole 3aa is formed with a 48% yield (Table 1, entry 1). When
  • larger or even commercial scale. The structures of compounds 3 were confirmed by 1H and 13C NMR spectroscopy and high-resolution mass spectrometry. The 1H NMR spectra are characterized by signals from protons of the aromatic rings in the range of 8.08–7.32 ppm, singlets for the proton at the C-3 position
  • of the 2,3-dihydro-1,2-thiazole ring in the range of 6.44–6.06 ppm, and multiplets corresponding to the morpholine fragment in the range of 3.80–2.89 ppm or protons of residues of other amino groups in the range of 3.51–1.31 ppm. In the 13C NMR spectra, characteristic signals of the C-4 atom of the
PDF
Album
Supp Info
Full Research Paper
Published 10 Jul 2025

N-Salicyl-amino acid derivatives with antiparasitic activity from Pseudomonas sp. UIAU-6B

  • Joy E. Rajakulendran,
  • Emmanuel Tope Oluwabusola,
  • Michela Cerone,
  • Terry K. Smith,
  • Olusoji O. Adebisi,
  • Adefolalu Adedotun,
  • Gagan Preet,
  • Sylvia Soldatou,
  • Hai Deng,
  • Rainer Ebel and
  • Marcel Jaspars

Beilstein J. Org. Chem. 2025, 21, 1388–1396, doi:10.3762/bjoc.21.103

Graphical Abstract
  • in the soil and thereby inhibiting the growth and sporulation of several pathogenic microorganisms [14]. In our effort to discover new bioactive secondary metabolites, our group previously cultured the Pseudomonas UIAU-6B strain in a closed system at 28 °C, and the methanolic extract led to the
  • isolation of five phenolic siderophores with some of them exhibiting antimicrobial properties, including pseudomonins A–C and pseudomobactins A and B [15]. In this study we report the isolation and structural characterization of two previously undescribed as natural products (1 and 2) and two new compounds
  • File 1, Figures S1–S4) of 1 measured in deuterated DMSO revealed well-defined proton signals identified as four aromatic sp2 methines, two aliphatic sp3 methines, a methyl group and four quaternary carbon signals at δC 157.7 (C-1), δC117.8 (C-6), δC 167.2 (C-7), and δC 172.9 (C-10). The presence of a
PDF
Album
Supp Info
Full Research Paper
Published 04 Jul 2025

High-pressure activation for the solvent- and catalyst-free syntheses of heterocycles, pharmaceuticals and esters

  • Kelsey Plasse,
  • Valerie Wright,
  • Guoshu Xie,
  • R. Bernadett Vlocskó,
  • Alexander Lazarev and
  • Béla Török

Beilstein J. Org. Chem. 2025, 21, 1374–1387, doi:10.3762/bjoc.21.102

Graphical Abstract
  • of catalyst- and solvent-free processes is desirable. Chalcones are a privileged scaffold in medicinal chemistry and are used for the synthesis of a multitude of products [43]. The cyclization between chalcones and hydrazines usually occur via a C=O/NH2 condensation and a subsequent NH addition to
  • the CC double bond, that most commonly require some form of catalysis. Thus, developing catalyst-free processes presents significant challenges, although there are few successful examples in the literature [44]. Similar to the dihydrobenzimidazoles above, the investigations here also started with an
  • parentheses. As shown in Table 4 and Scheme 4, the HHP reactions occurred at 50 °C with low to moderate yields, and with nearly quantitative yields at 80 °C. HHP reactions afforded the products in higher yields than those of the control experiments under the same conditions (time, temp. etc.) but only at 1
PDF
Album
Supp Info
Full Research Paper
Published 02 Jul 2025

Oxetanes: formation, reactivity and total syntheses of natural products

  • Peter Gabko,
  • Martin Kalník and
  • Maroš Bella

Beilstein J. Org. Chem. 2025, 21, 1324–1373, doi:10.3762/bjoc.21.101

Graphical Abstract
  • tetrahedral value which results in a large ring strain of 25.5 kcal/mol, comparable to oxirane (27.3 kcal/mol) and much greater than tetrahydrofuran (5.6 kcal/mol) [4]. Moreover, the strained C–O–C bond angle effectively exposes the oxygen lone pairs, making oxetane a strong hydrogen-bond acceptor and Lewis
  • E [30] and daphnepapytone C [31] from the 2020s (Figure 4). Most of these compounds possess intriguing biological activities and selected examples from this list are discussed in more detail in chapter 4 with regards to their isolation, bioactivity and a recent total synthesis. The aim of this work
  • categorised into 6 synthetic strategies as depicted in Scheme 1: a) C–O bond-forming cyclisations, b) CC bond-forming cyclisations, c) [2 + 2] cycloadditions between carbonyls and alkenes, d) ring expansions, e) ring contractions and f) O–H insertions. In the following subchapters, these strategies will be
PDF
Album
Review
Published 27 Jun 2025

Recent advances in amidyl radical-mediated photocatalytic direct intermolecular hydrogen atom transfer

  • Hao-Sen Wang,
  • Lin Li,
  • Xin Chen,
  • Jian-Li Wu,
  • Kai Sun,
  • Xiao-Lan Chen,
  • Ling-Bo Qu and
  • Bing Yu

Beilstein J. Org. Chem. 2025, 21, 1306–1323, doi:10.3762/bjoc.21.100

Graphical Abstract
  • transfer (HAT) in photocatalytic reactions. These radicals display exceptional selectivity and efficiency in abstracting hydrogen atoms from C–H, Si–H, B–H, and Ge–H, positioning them as invaluable tools in synthetic chemistry. This review summarizes the latest advancements in the photocatalyzed generation
  • substrates. Keywords: amidyl radicals; C–H; HAT reagents; hydrogen-atom-transfer; late-stage functionalization; Introduction C–H bonds are the predominant chemical bonds in organic compounds, and their direct conversion can rapidly and efficiently increase the complexity and functionality of organic
  • molecules. On the other hand, C–H bonds exhibit low reactivity due to their relatively high bond dissociation energy (BDE) (Figure 1a). Therefore, the direct functionalization of C–H bonds is extremely challenging [1][2][3][4][5]. In recent decades, transition-metal-catalyzed C–H bond functionalization
PDF
Album
Review
Published 27 Jun 2025

Recent advances and future challenges in the bottom-up synthesis of azulene-embedded nanographenes

  • Bartłomiej Pigulski

Beilstein J. Org. Chem. 2025, 21, 1272–1305, doi:10.3762/bjoc.21.99

Graphical Abstract
  • synthesis of π-extended azulene was the non-benzenoid isomer of pyrene published by Ward and co-workers (Scheme 1) [31]. Cyclohept[bc]acenaphthylene (2) was obtained from a partially saturated precursor 1 via dehydrogenation using palladium on carbon. However, the reaction carried out at 300 °C gave 2 as a
  • dehydrogenated using sulfur in trichlorobenzene at 220 °C to yield the azulene-containing isomer of benzo[a]pyrene 9 in 18% isolated yield. Bestmann and Ruppert reported the synthesis of a dinaphthoazulene 14, a non-alternant isomer of benzo[a]perylene (Scheme 2) [34]. In their method, bisylide 10 was reacted
  • by Jutz and Kirchlechner in 1966 (Scheme 3) [36]. Condensation between phenalene 15 and pentafulvene 16 gave pentafulvene 17. Pentafulvene 17 was finally subjected to Ziegler–Hafner reaction in quinoline at 180 °C, resulting in the π-extended azulene 18 in 60% yield. A similar synthetic strategy was
PDF
Album
Review
Published 26 Jun 2025

Recent advances in oxidative radical difunctionalization of N-arylacrylamides enabled by carbon radical reagents

  • Jiangfei Chen,
  • Yi-Lin Qu,
  • Ming Yuan,
  • Xiang-Mei Wu,
  • Heng-Pei Jiang,
  • Ying Fu and
  • Shengrong Guo

Beilstein J. Org. Chem. 2025, 21, 1207–1271, doi:10.3762/bjoc.21.98

Graphical Abstract
  • transformations, (4) electrochemical approaches, and (5) metal-free or electron donor–acceptor (EDA)-driven systems. The substrate scope, limitations, and mechanistic aspects of these radical cascade cyclization strategies are critically examined. Review N-Arylalkenes: alkyl C(sp3)–H radicals Early investigations
  • primarily focused on substrates containing activated alkenes tethered within the molecular framework. N-Arylacrylamides were employed as model substrates, and a diverse range of functionalization reagents, including those with benzylic C–H bonds, C(sp3)–H bonds adjacent to heteroatoms, di
  • activated alkenes involving benzylic C(sp3)–H bonds through a cascade cyclization process (Scheme 1) [2]. This organomediated approach can be facilitated by a catalytic amount of Lewis acid. Using DTBP as the oxidant and IrCl3 as the promoter, a range of benzylic C–H bonds in arylmethanes
PDF
Album
Review
Published 24 Jun 2025

Optimized synthesis of aroyl-S,N-ketene acetals by omission of solubilizing alcohol cosolvents

  • Julius Krenzer and
  • Thomas J. J. Müller

Beilstein J. Org. Chem. 2025, 21, 1201–1206, doi:10.3762/bjoc.21.97

Graphical Abstract
  • . This can be obtained from renewable biomass and is therefore considered to be a "green" solvent. In 2-MeTHF (conditions (C)), product 1f was isolated with a yield of 95% and derivative 1h with a yield of 91%. This shows that the sustainable solvent is a potent alternative to 1,4-dioxane. With the new
  • ® and purified by flash chromatography on silica gel (n-hexane/acetone 3:1). Then, the crude product was suspended in n-hexane, the supernatant separated by filtration and the precipitate was dried under vacuum to afford aroyl-S,N-ketene acetal 1i (1.78 g, 4.93 mmol, 99%) as a yellow solid. Mp 159 °C
  • (lit. 155 °C [5]). 1H NMR (300 MHz, CDCl3) δ 5.24 (s, 2H), 6.43 (s, 1H), 6.92–7.04 (m, 3H), 7.09–7.16 (m, 3H), 7.19–7.29 (m, 4H), 7.57 (dd, 3J = 7.7 Hz, 4J = 1.3 Hz, 1H), 7.73–7.83 (m, 2H); MALDI–TOF–MS (m/z): 362.2, [C22H16FNOS + H]+. Retrosynthetic analysis of aroyl-S,N-ketene acetals 1 and tentative
PDF
Album
Supp Info
Full Research Paper
Published 20 Jun 2025

Synthesis of β-ketophosphonates through aerobic copper(II)-mediated phosphorylation of enol acetates

  • Alexander S. Budnikov,
  • Igor B. Krylov,
  • Fedor K. Monin,
  • Valentina M. Merkulova,
  • Alexey I. Ilovaisky,
  • Liu Yan,
  • Bing Yu and
  • Alexander O. Terent’ev

Beilstein J. Org. Chem. 2025, 21, 1192–1200, doi:10.3762/bjoc.21.96

Graphical Abstract
  • ) salts, including halides, nitrate, tetrafluoroborate, or perchlorate, were much less effective or completely inert. Keywords: C–P coupling; copper; enol acetates; β-ketophosphonates; phosphorylation; Introduction The construction of C–P bonds is a highly important task in key areas of modern chemistry
  • efficiency, strong basic or acidic conditions, and excess of organohalides as starting materials. Recent years have witnessed the upsurge of free-radical oxidative phosphorylation transformations that became a reliable strategy for the construction of C–P bonds in organophosphorus chemistry [2][32][33][34
  • leading to β-ketophosphonates have been reported [41][42][43][44][45][46][47][48][49][50][61], challenges in this area still exist primely in the search for new available synthetic equivalents of alkynes and alkenes for effective radical C–P bond formation. Enol acetates are potentially versatile
PDF
Album
Supp Info
Full Research Paper
Published 20 Jun 2025

Selective monoformylation of naphthalene-fused propellanes for methylene-alternating copolymers

  • Kenichi Kato,
  • Tatsuki Hiroi,
  • Seina Okada,
  • Shunsuke Ohtani and
  • Tomoki Ogoshi

Beilstein J. Org. Chem. 2025, 21, 1183–1191, doi:10.3762/bjoc.21.95

Graphical Abstract
  • thermogravimetric analysis (TGA) (Figure S703 in Supporting Information File 1). Temperatures at which weight loss reached 10% (T90) were 468–491 °C and carbonization yields at 900 °C (CY) were ca. 46 wt % for [4.3.3]_oligo and [4.3.3]_linear. T90 and CY of [4.3.3]_branch showed higher values of 543 °C and 75 wt
  • % probably owing to the network structure. By contrast, soluble [3.3.3]_oligo and [3.3.3]_linear had relatively high T90 of 528–532 °C and CY of 68–76 wt %. The high values were ascribed to two unsubstituted naphthalene rings in precursor [3.3.3]_CH2OH, which caused facile branching in the reaction or
  • heating process. T90 and CY of [3.3.3]_branch (415 °C and 64 wt %) were lower than those of [3.3.3]_linear because of two-step decay profile (Figure S703a in Supporting Information File 1). All the samples showed broad powder X-ray diffraction (PXRD) patterns with unclear peaks at around 2θ = 11° and 20
PDF
Album
Supp Info
Full Research Paper
Published 18 Jun 2025

Enhancing chemical synthesis planning: automated quantum mechanics-based regioselectivity prediction for C–H activation with directing groups

  • Julius Seumer,
  • Nicolai Ree and
  • Jan H. Jensen

Beilstein J. Org. Chem. 2025, 21, 1171–1182, doi:10.3762/bjoc.21.94

Graphical Abstract
  • Julius Seumer Nicolai Ree Jan H. Jensen Department of Chemistry, University of Copenhagen, Copenhagen, Denmark 10.3762/bjoc.21.94 Abstract The mild and selective functionalization of carbon–hydrogen (C–H) bonds remains a pivotal challenge in organic synthesis, crucial for developing complex
  • molecular architectures in pharmaceuticals, polymers, and agrochemicals. Despite advancements in directing group (DG) methodologies and computational approaches, predicting accurate regioselectivity in C–H activation poses significant difficulties due to the diversity and complexity of organic compounds
  • . This study introduces a novel quantum mechanics-based computational workflow tailored for the regioselective prediction of C–H activation in the presence of DGs. Utilizing (semi-empirical) quantum calculations hierarchically, the workflow efficiently predicts outcomes by considering concerted
PDF
Album
Supp Info
Full Research Paper
Published 16 Jun 2025

A multicomponent reaction-initiated synthesis of imidazopyridine-fused isoquinolinones

  • Ashutosh Nath,
  • John Mark Awad and
  • Wei Zhang

Beilstein J. Org. Chem. 2025, 21, 1161–1169, doi:10.3762/bjoc.21.92

Graphical Abstract
  • compounds [18]. Presented in this paper is a new synthetic route involving GBB, N-acylation and IMDA reactions for making intermediate III followed by dehydrative re-aromatization to give imidazopyridine-fused isoquinolinones C (Scheme 1C). Results and Discussion Following the reported procedures [10], the
  • initial GBB reaction of aminopyridines 1 (0.5 mmol), isocyanides 3 (1.2 equiv), and furfuraldehydes 2 (1.2 equiv) was conducted in 3:1 CH2Cl2/MeOH (4 mL) using Yb(OTf)3 (0.08 equiv) as a Lewis acid catalyst under microwave irradiation at 100 °C for 1 h (Scheme 2). Nineteen distinct adducts 4 were obtained
  • in 1,2-dichlorobenzene at 180 °C for 4 h, which gave 8a in 85% conversion and 82% isolated yield (Table 1, entry 3). Other solvents like toluene and xylene gave minimal or no product. Different combinations of temperature and reaction time couldn’t improve the yield. Among the various Lewis acids
PDF
Album
Supp Info
Full Research Paper
Published 13 Jun 2025

Synthetic approach to borrelidin fragments: focus on key intermediates

  • Yudhi Dwi Kurniawan,
  • Zetryana Puteri Tachrim,
  • Teni Ernawati,
  • Faris Hermawan,
  • Ima Nurasiyah and
  • Muhammad Alfin Sulmantara

Beilstein J. Org. Chem. 2025, 21, 1135–1160, doi:10.3762/bjoc.21.91

Graphical Abstract
  • macrolides with hydroxy groups at C20 or C7, identified as borrelidins C–E (Table 1, entries 4, 7, and 8) [19]. Borrelidins CR1 and CR2 (Table 1, entries 5 and 6), amide-containing congeners, were also isolated through bioassay-guided fractionation and purification of marine microorganisms from Costa Rica
  • these sulfones with epoxides 23b and 23a, respectively. Following the literature procedure for a similar reaction, using n-butyllithium in the presence of BF3·Et2O at −78 °C, the coupling reaction unfortunately resulted in the decomposition of the reactants (Scheme 4). The authors hypothesized that the
  • epoxides 23b and 23a with the non-protected variant 23c, and reacting it with sulfone 27 after pre-complexation with Ti(OiPr)4, again led only to decomposition. Given the unsatisfying results, Uguen and co-workers replaced the epoxides 23a–c to monoethers 46a and 46b, derived from trans-2,3-epoxy-1,4
PDF
Album
Review
Published 12 Jun 2025
Other Beilstein-Institut Open Science Activities