Search results

Search for "neat" in Full Text gives 306 result(s) in Beilstein Journal of Organic Chemistry. Showing first 200.

Palladium-catalyzed solid-state borylation of aryl halides using mechanochemistry

  • Koji Kubota,
  • Emiru Baba,
  • Tamae Seo,
  • Tatsuo Ishiyama and
  • Hajime Ito

Beilstein J. Org. Chem. 2022, 18, 855–862, doi:10.3762/bjoc.18.86

Graphical Abstract
  • serve as reactants and solvents (neat liquid conditions). In addition, a long reaction time (12 h) is required to complete the reaction [15]. A general and efficient solvent-free borylation protocol that can be applied to liquid as well as solid aryl halides remains undeveloped. Recently
  • 3). Reactions with DPEphos and XPhos, which are the optimal ligands for neat liquid conditions reported by Nechaev [15], also yielded only poor results (Table 1, entries 4 and 5). Other monophosphine ligands, such as SPhos, PCy3, and PAd3, did not improve the yield of 3a (Table 1, entries 6–8). The
PDF
Album
Supp Info
Letter
Published 18 Jul 2022

First series of N-alkylamino peptoid homooligomers: solution phase synthesis and conformational investigation

  • Maxime Pypec,
  • Laurent Jouffret,
  • Claude Taillefumier and
  • Olivier Roy

Beilstein J. Org. Chem. 2022, 18, 845–854, doi:10.3762/bjoc.18.85

Graphical Abstract
  • proton exchange. Overall, the NMR study suggests that the N-methylamino glycine monomer and oligomers have a strong propensity to form intermolecular hydrogen bonds, an interesting and sought-after property for self-assembling and interaction with biological targets. Neat peptoids 1–6 were also
PDF
Album
Supp Info
Full Research Paper
Published 14 Jul 2022

Post-synthesis from Lewis acid–base interaction: an alternative way to generate light and harvest triplet excitons

  • Hengjia Liu and
  • Guohua Xie

Beilstein J. Org. Chem. 2022, 18, 825–836, doi:10.3762/bjoc.18.83

Graphical Abstract
  • /TFA 5000:1 (v/v); device C, neat film of compound 7. Figure 6a and 6b were reproduced from [37], P. Zalar et al., “Color Tuning in Polymer Light-Emitting Diodes with Lewis Acids”, Angew. Chem., Int. Ed., with permission from John Wiley and Sons. Copyright © 2012 WILEY-VCH Verlag GmbH & Co. KGaA
PDF
Album
Review
Published 12 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
  • ) with molecular oxygen at elevated temperatures providing (+)-nootkatone (8) in 10% yield (Scheme 2). In this setup, neat (+)-valencene (7) is mixed with a stream of oxygen resulting in the formation of a segmented gas–liquid flow. In segmented flow a higher surface-to-volume ratio is achieved and
PDF
Album
Review
Published 27 Jun 2022

Complementarity of solution and solid state mechanochemical reaction conditions demonstrated by 1,2-debromination of tricyclic imides

  • Petar Štrbac and
  • Davor Margetić

Beilstein J. Org. Chem. 2022, 18, 746–753, doi:10.3762/bjoc.18.75

Graphical Abstract
  • ), even with the agitation by ultrasound (Table 1, entry 22). The ratios of side-products vary depending on the reaction conditions (Table S1, Supporting Information File 1). The endo-product 15 was dominant in the neat grinding experiment (Table 1, entry 1), whereas phthalimide 16 dominates when NaCl was
  • could be also obtained by cycloaddition parity reversal principle [16][36], employing norbornene dibromide 11 and furan (18). Ball milling reaction without THF (neat grinding) provided 22 as the major product, together with a minor amount of dehalogenated product 40 (Scheme 2). When THF was added for
PDF
Album
Supp Info
Full Research Paper
Published 24 Jun 2022

Mechanochemical halogenation of unsymmetrically substituted azobenzenes

  • Dajana Barišić,
  • Mario Pajić,
  • Ivan Halasz,
  • Darko Babić and
  • Manda Ćurić

Beilstein J. Org. Chem. 2022, 18, 680–687, doi:10.3762/bjoc.18.69

Graphical Abstract
  • using N-halosuccinimides as the halogen source under neat grinding or liquid-assisted grinding conditions in a ball mill has been described. Depending on the azobenzene substrate used, halogenation of the C–H bonds occurs in the absence or only in the presence of PdII catalysts. Insight into the
  • -bromosuccinimide (NBS) under neat grinding (NG) and liquid-assisted grinding (LAG) conditions in a ball mill [51]. Insight into the dynamics of the formation of reaction intermediates and products was obtained by in situ Raman monitoring that provided information on the nature of the catalytically active PdII
  • product as a single isomer in both the catalyzed and uncatalyzed reactions. Neat grinding of L3 and L4 with NCS produced the monochlorinated products L3Cl-I and L4Cl-I as single isomers within one hour (Table 1, entries 2 and 3). In contrast, the reaction of L5 with NCS resulted in a mixture of mono- and
PDF
Album
Supp Info
Full Research Paper
Published 15 Jun 2022

New synthesis of a late-stage tetracyclic key intermediate of lumateperone

  • Mátyás Milen,
  • Bálint Nyulasi,
  • Tamás Nagy,
  • Gyula Simig and
  • Balázs Volk

Beilstein J. Org. Chem. 2022, 18, 653–659, doi:10.3762/bjoc.18.66

Graphical Abstract
  • 24 was accomplished in neat conditions within 3 days of reaction time using a literature procedure [5]. Attempts in acetonitrile as solvent were unsuccessful: after 1 day at room temperature, product 28 was isolated in only 20% yield, while at reflux temperature, a decomposition has been observed
PDF
Album
Supp Info
Full Research Paper
Published 10 Jun 2022

Sesquiterpenes from the soil-derived fungus Trichoderma citrinoviride PSU-SPSF346

  • Wiriya Yaosanit,
  • Vatcharin Rukachaisirikul,
  • Souwalak Phongpaichit,
  • Sita Preedanon and
  • Jariya Sakayaroj

Beilstein J. Org. Chem. 2022, 18, 479–485, doi:10.3762/bjoc.18.50

Graphical Abstract
  • mg) was then washed with acetone to afford compound 5 (3.7 mg). Trichocitrinovirene A (1): Colorless gum; +46.1 (c 0.67, MeOH); UV (MeOH) λmax, nm (log ε): 210 (3.32); ECD (MeOH, c 0.0008) λmax, nm (Δε): 227 (+4.3); IR (neat) νmax: 3336, 1684, 1649 cm−1; 1H and 13C NMR (CD3OD) see Table 1; HRMS–ESI
  • (m/z): [M + Na]+ calcd for C15H22O5Na, 305.1356; found, 305.1359. Trichocitrinovirene B (2): Colorless gum; +44.6 (c 0.67, MeOH); UV (MeOH) λmax, nm (log ε): 210 (3.67); IR (neat) νmax: 3386, 1683, 1645 cm−1; 1H and 13C NMR (CD3OD) see Table 1; HRMS–ESI (m/z): [M + Na]+ calcd for C15H22O6Na
PDF
Album
Supp Info
Full Research Paper
Published 29 Apr 2022

Trichloroacetic acid fueled practical amine purifications

  • Aleena Thomas,
  • Baptiste Gasch,
  • Enzo Olivieri and
  • Adrien Quintard

Beilstein J. Org. Chem. 2022, 18, 225–231, doi:10.3762/bjoc.18.26

Graphical Abstract
  • 1). Presence of TCA in the 1H NMR spectra is easily monitored through a broad peak around 9 ppm and also through the upfield migration of the other peaks of the amine ammonium salt. Heating the amine salt precipitate neat at 100 °C also provided partial decarboxylation (Table 1, entry 2). In sharp
PDF
Album
Supp Info
Full Research Paper
Published 24 Feb 2022

Iron-catalyzed domino coupling reactions of π-systems

  • Austin Pounder and
  • William Tam

Beilstein J. Org. Chem. 2021, 17, 2848–2893, doi:10.3762/bjoc.17.196

Graphical Abstract
  • solvent-tuned [129]. In neat CH2Cl2, the reaction produced the expected β-trichloromethyl alkyl azide; however, the reaction was chemoselective for diazidation when tert-butanol was used as co-solvent. The authors hypothesized the presence of the alcohol suppresses the polar-unmatched HAT process from
PDF
Album
Review
Published 07 Dec 2021

The ethoxycarbonyl group as both activating and protective group in N-acyl-Pictet–Spengler reactions using methoxystyrenes. A short approach to racemic 1-benzyltetrahydroisoquinoline alkaloids

  • Marco Keller,
  • Karl Sauvageot-Witzku,
  • Franz Geisslinger,
  • Nicole Urban,
  • Michael Schaefer,
  • Karin Bartel and
  • Franz Bracher

Beilstein J. Org. Chem. 2021, 17, 2716–2725, doi:10.3762/bjoc.17.183

Graphical Abstract
  • Finnigan LTQ FT Ultra Fourier Transform Ion Cyclotron Resonance device at 250 °C for ESI. IR spectra were recorded on a Perkin Elmer FT-IR Paragon 1000 instrument as neat materials. Absorption bands were reported in wave numbers (cm−1), obtained on a ATR PRO450-S accessory (Jasco). Melting points were
PDF
Album
Supp Info
Full Research Paper
Published 05 Nov 2021

N-Sulfinylpyrrolidine-containing ureas and thioureas as bifunctional organocatalysts

  • Viera Poláčková,
  • Dominika Krištofíková,
  • Boglárka Némethová,
  • Renata Górová,
  • Mária Mečiarová and
  • Radovan Šebesta

Beilstein J. Org. Chem. 2021, 17, 2629–2641, doi:10.3762/bjoc.17.176

Graphical Abstract
  • conditions. Again a small amount of catalyst (S,R)-C2 (2.5 mol %) gave the best chemical yield, 70% using solvent-free, neat stirring at 30 °C. In comparison, the ball-mill reaction afforded 66% of the product (Table 4, entries 5–7). The Michael addition of aldehyde 6c gave under dry stirring products 8e and
PDF
Album
Supp Info
Full Research Paper
Published 25 Oct 2021

Silica gel and microwave-promoted synthesis of dihydropyrrolizines and tetrahydroindolizines from enaminones

  • Robin Klintworth,
  • Garreth L. Morgans,
  • Stefania M. Scalzullo,
  • Charles B. de Koning,
  • Willem A. L. van Otterlo and
  • Joseph P. Michael

Beilstein J. Org. Chem. 2021, 17, 2543–2552, doi:10.3762/bjoc.17.170

Graphical Abstract
  • 15a was heated in capped tubes under microwave conditions [35][36][37] for 10 minutes in solvents such as ethanol (100 °C), xylene (150 °C) or N,N-dimethylformamide (150 °C) (Table 1, entries 3–5). Dissolution in neat protic or Lewis acids (e.g., hydrochloric acid, trifluoroacetic acid
PDF
Album
Supp Info
Full Research Paper
Published 13 Oct 2021

Copper-catalyzed monoselective C–H amination of ferrocenes with alkylamines

  • Zhen-Sheng Jia,
  • Qiang Yue,
  • Ya Li,
  • Xue-Tao Xu,
  • Kun Zhang and
  • Bing-Feng Shi

Beilstein J. Org. Chem. 2021, 17, 2488–2495, doi:10.3762/bjoc.17.165

Graphical Abstract
  • in MeCN, the yield could be improved to 32% (Table 1, entry 4). Further screening of other oxidants revealed that NMO was the optimal (Table 1, entries 9–11). When the reaction was conducted in neat in the presence of 2-pyridone, 3a was obtained in 36% yield (Table 1, entry 12). However, significant
PDF
Album
Supp Info
Letter
Published 28 Sep 2021

Synthesis of 5-arylacetylenyl-1,2,4-oxadiazoles and their transformations under superelectrophilic activation conditions

  • Andrey I. Puzanov,
  • Dmitry S. Ryabukhin,
  • Anna S. Zalivatskaya,
  • Dmitriy N. Zakusilo,
  • Darya S. Mikson,
  • Irina A. Boyarskaya and
  • Aleksander V. Vasilyev

Beilstein J. Org. Chem. 2021, 17, 2417–2424, doi:10.3762/bjoc.17.158

Graphical Abstract
  • reaction of the acetylenyl-1,2,4-oxadiazoles with arenes in neat triflic acid TfOH (CF3SO3H) at room temperature for 1 h resulted in the formation of E/Z-5-(2,2-diarylethenyl)-3-aryl-1,2,4-oxadiazoles as products of regioselective hydroarylation of the acetylene bond. The addition of TfOH to the acetylene
PDF
Album
Supp Info
Full Research Paper
Published 15 Sep 2021

Development of N-F fluorinating agents and their fluorinations: Historical perspective

  • Teruo Umemoto,
  • Yuhao Yang and
  • Gerald B. Hammond

Beilstein J. Org. Chem. 2021, 17, 1752–1813, doi:10.3762/bjoc.17.123

Graphical Abstract
  • . disclosed the quaternary ammonium N-F reagent, N-fluoroquinuclidinium fluoride (6-1) [43]. They subsequently followed with more detailed results in 1988 [44]. Quinuclidine was fluorinated by neat fluorine in trichlorofluoromethane at −72 °C, affording the product 6-1 in 86% yield (Scheme 15). Fluorination
  • reactivity than the sulfonamide reagents such as Barnette’s N-fluoro-N-alkylarenesulfonamides, since the electronic density on the nitrogen was greatly decreased by two strong electron-withdrawing CF3SO2 groups. Reagent 7-1a reacted slowly with benzene and toluene under neat conditions, whereas activated
  • pentafluoropyridine, the precursor 11-1 was prepared as illustrated in Scheme 25. Treatment of 11-1 with neat F2 in acetonitrile at −10 °C under reduced pressure gave N-fluoro-sulfonamide 11-2 in 89% yield. This product was however a 9:1 mixture of the N-F reagent 11-2 and the protonated compound of 11-1. The
PDF
Album
Review
Published 27 Jul 2021

A straightforward conversion of 1,4-quinones into polycyclic pyrazoles via [3 + 2]-cycloaddition with fluorinated nitrile imines

  • Greta Utecht-Jarzyńska,
  • Karolina Nagła,
  • Grzegorz Mlostoń,
  • Heinz Heimgartner,
  • Marcin Palusiak and
  • Marcin Jasiński

Beilstein J. Org. Chem. 2021, 17, 1509–1517, doi:10.3762/bjoc.17.108

Graphical Abstract
  • , HMQC, and HMBC). The UV–vis spectra were measured on a PerkinElmer Lambda 45 spectrophotometer in spectroscopic grade CH2Cl2. MS (ESI) were performed with a Varian 500-MS LC Ion Trap. The IR spectra were measured neat with an Agilent Cary 630 FTIR spectrometer. Elemental analyses were obtained with a
  • , 133.4 (2i-C), 134.2, 135.0 (CH each), 138.1, 139.0 (2i-C), 140.8 (q, 2JC,F = 40.8 Hz, C(3)), 174.5, 177.3 (2C=O); 19F NMR (CDCl3, 565 MHz) δ −62.82 ppm; UV–vis (CH2Cl2) λmax (log ε) 247 (4.45), 266 (4.24), 275 (4.25), 340 (3.72), 409 (2.61), 496 nm (1.70); IR (neat) νmax: 3082, 1677 (C=O), 1588, 1521
  • ) δ 13.6, 13.9, 14.2, 14.6, 20.8 (5Me), 64.8, 80.1 (2i-C), 120.7 (q,1JC,F = 271.1 Hz, CF3), 120.7, 129.6 (2CH each), 134.8 (i-C), 140.0 (q, 2JC,F = 36.3 Hz, C(3)), 139.5, 146.9, 147.1 (3i-C), 192.6, 195.1 (2C=O) ppm; 19F NMR (CDCl3, 565 MHz) δ −61.45 ppm; IR (neat) νmax: 2930, 1677 (C=O), 1513, 1506
PDF
Album
Supp Info
Full Research Paper
Published 28 Jun 2021

Synthesis of 1-indolyl-3,5,8-substituted γ-carbolines: one-pot solvent-free protocol and biological evaluation

  • Premansh Dudhe,
  • Mena Asha Krishnan,
  • Kratika Yadav,
  • Diptendu Roy,
  • Krishnan Venkatasubbaiah,
  • Biswarup Pathak and
  • Venkatesh Chelvam

Beilstein J. Org. Chem. 2021, 17, 1453–1463, doi:10.3762/bjoc.17.101

Graphical Abstract
  • screening of several reaction conditions, we found that heating at 120 °C of a neat mixture consisting of 1-methyl-1H-indole-2-carbaldehyde (1a, 2.0 equiv), glycine methyl ester HCl salt (2a, 1.0 equiv), and DIPEA (3.5 equiv) in a sealed tube for 6 h, led to the isolation of γ-carboline 3aa in 70% yield
PDF
Album
Supp Info
Letter
Published 17 Jun 2021

A comprehensive review of flow chemistry techniques tailored to the flavours and fragrances industries

  • Guido Gambacorta,
  • James S. Sharley and
  • Ian R. Baxendale

Beilstein J. Org. Chem. 2021, 17, 1181–1312, doi:10.3762/bjoc.17.90

Graphical Abstract
PDF
Album
Review
Published 18 May 2021

Synthetic accesses to biguanide compounds

  • Oleksandr Grytsai,
  • Cyril Ronco and
  • Rachid Benhida

Beilstein J. Org. Chem. 2021, 17, 1001–1040, doi:10.3762/bjoc.17.82

Graphical Abstract
  • production of biodiesel [5][80], intermediates for triazine synthesis [48], etc. Some derivatives were grafted onto polystyrene resins for catalytic uses [5]. Various solvents and temperatures were used such as hot DMF [5], neat [79], toluene or hexane at 25–100 °C [79]. Generally, higher temperatures seem
  • curing processes. [79]. For example, Sakai et al. initially used quite harsh conditions for the reactions with dialkylcarbodiimides (neat, 100 °C, 2 h) that led to relatively low yields. Later, they applied smoother room temperature conditions for arylalkylcarbodiimides and diaryl derivatives: either
  • neat for 30 min, or 24 h diluted in an apolar solvent. Both of these conditions proved very efficient even for hindered reagents and displayed good to excellent yields (Scheme 39B). In the search for new biguanide-based organocatalysts for the transesterification of vegetable oils, Glasovac et al
PDF
Album
Review
Published 05 May 2021

Microwave-assisted multicomponent reactions in heterocyclic chemistry and mechanistic aspects

  • Shivani Gulati,
  • Stephy Elza John and
  • Nagula Shankaraiah

Beilstein J. Org. Chem. 2021, 17, 819–865, doi:10.3762/bjoc.17.71

Graphical Abstract
  • -assisted four-component reaction involving chromene-aldehyde (61), amines 32, acyclic 1,3-diketones 54 and nitromethane using silica-gel-supported polyphoshoric acid as catalyst under neat conditions for the synthesis of tetra-substituted pyrroles 62. A comparative study of the protocol employing the
PDF
Album
Review
Published 19 Apr 2021

Synthesis of bis(aryloxy)fluoromethanes using a heterodihalocarbene strategy

  • Carl Recsei and
  • Yaniv Barda

Beilstein J. Org. Chem. 2021, 17, 813–818, doi:10.3762/bjoc.17.70

Graphical Abstract
  • ); (b) AcCl, HCl, Et2O, rt, 24 h, 70%; (c) KHF2, neat, 105–110 °C, 60%. Attempted synthesis of 4. Reagents and conditions: (a) Ca(OH)2, 1,4-dioxane/water, reflux, 72 h, 5% (by HPLC integration); (b) ClF2CO2Na, K2CO3, DMF, 95 °C, 4 h, 7 (4%), 8 (19%). Synthesis of 10. Reagents and conditions: (a
PDF
Album
Supp Info
Letter
Published 12 Apr 2021

Synthesis of (Z)-3-[amino(phenyl)methylidene]-1,3-dihydro-2H-indol-2-ones using an Eschenmoser coupling reaction

  • Lukáš Marek,
  • Lukáš Kolman,
  • Jiří Váňa,
  • Jan Svoboda and
  • Jiří Hanusek

Beilstein J. Org. Chem. 2021, 17, 527–539, doi:10.3762/bjoc.17.47

Graphical Abstract
  • ATR diamond crystal (neat solid samples). Flash chromatography was performed using a Büchi Reveleris® X2 flash chromatography system equipped with a UV–vis/ELSD detector and Reveleris Flash pure cartridges (12–40 g, 35–45 μm, 53–80 Å) or puriFlash® Alumine N 32/63 µm cartridges (12 g). General
PDF
Album
Supp Info
Full Research Paper
Published 23 Feb 2021

Helicene synthesis by Brønsted acid-catalyzed cycloaromatization in HFIP [(CF3)2CHOH]

  • Takeshi Fujita,
  • Noriaki Shoji,
  • Nao Yoshikawa and
  • Junji Ichikawa

Beilstein J. Org. Chem. 2021, 17, 396–403, doi:10.3762/bjoc.17.35

Graphical Abstract
  • , 129.5, 129.6, 130.0, 130.7, 131.2, 131.6, 133.8, 134.0, 140.2, 140.9, 141.4; IR (neat) ν: 3055, 3020, 2976, 2883, 1471, 1442, 1433, 1398, 1194, 1130, 1038, 1009, 943, 872, 837, 822, 752, 706, 621, 573, 538 cm−1; HRMS (EI) m/z: [M]+ calcd. for C26H26O4, 402.1826; found, 402.1815. Gram-scale synthesis
  • , 124.8, 125.2, 126.8, 129.0, 129.05, 129.13, 129.9, 131.0, 133.8, 134.5, 138.8, 142.6; IR (neat) ν: 3053, 2966, 2883, 1489, 1396, 1132, 1043, 984, 831, 756 cm–1; HRMS (APCI+) m/z: [M + H]+ calcd. for C30H29O4, 453.2060; found, 453.2081. Gram-scale synthesis: Compound 4b was also prepared by the method
PDF
Album
Supp Info
Full Research Paper
Published 09 Feb 2021

1,2,3-Triazoles as leaving groups in SNAr–Arbuzov reactions: synthesis of C6-phosphonated purine derivatives

  • Kārlis-Ēriks Kriķis,
  • Irina Novosjolova,
  • Anatoly Mishnev and
  • Māris Turks

Beilstein J. Org. Chem. 2021, 17, 193–202, doi:10.3762/bjoc.17.19

Graphical Abstract
  • toluene, MeCN, and DCM, and in the presence of 1–20 equiv of P(OEt)3, the formation of the desired phosphonates 4 was not observed (Scheme 6). We started an optimization of the reaction conditions using substrate 6d, and reactions in neat phosphite at various temperatures were tried (Table 2). The
  • hours in neat P(OEt)3 at 160 °C. With the experimental conditions in hand, the SNAr–Arbuzov reaction between 2,6-bistriazolylpurines 6a–i and P(OEt)3 provided a library of novel purine phosphonates 4a–i in 27–82% yield (Table 3). The products 4a, 4d, 4e, and 4i were easily precipitated from hexane left
PDF
Album
Supp Info
Full Research Paper
Published 20 Jan 2021
Other Beilstein-Institut Open Science Activities