Search results

Search for "tosylate" in Full Text gives 117 result(s) in Beilstein Journal of Organic Chemistry.

Benzoimidazolium-derived dimeric and hydride n-dopants for organic electron-transport materials: impact of substitution on structures, electrochemistry, and reactivity

  • Swagat K. Mohapatra,
  • Khaled Al Kurdi,
  • Samik Jhulki,
  • Georgii Bogdanov,
  • John Bacsa,
  • Maxwell Conte,
  • Tatiana V. Timofeeva,
  • Seth R. Marder and
  • Stephen Barlow

Beilstein J. Org. Chem. 2023, 19, 1651–1663, doi:10.3762/bjoc.19.121

Graphical Abstract
  • iodomethane (or methyl tosylate) to afford the benzimidazolium iodides (or tosylates), 1+I− (or 1+OTs−), which were metathesized to the corresponding hexafluorophosphates, 1+PF6−. Either I− or PF6− salt can then be converted to the corresponding 1H derivative using NaBH4 in MeOH. The PF6− salts are somewhat
PDF
Album
Supp Info
Full Research Paper
Published 01 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
  • reacted immediately (one-pot procedure) with sodium borohydride to give 4.4. It must be noted that this sequence does not induce racemization and that 4.4 can be stored for months with 0.5% of solid KOH acting as a stabilizer [72]. Then, 4.4 was alkylated with stearyl tosylate to produce 4.5. The two
  • benzylation produced 4.13. Then, the deprotection of the primary alcohol in acidic conditions allows introducing the phosphocholine polar head group by using POCl3 and the choline tosylate salt as reagents to yield 4.14. Finally, the debenzylation of the secondary alcohol and its acylation produce PAF 4.15. b
  • from allyl alcohol (Figure 7) [82]. The Sharpless asymmetric epoxidation of allyl alcohol followed by tosylation produced glycidyl tosylate 7.1a (Figure 7). The reaction of palmityl alcohol (C16H33-OH) in the presence of a catalytic amount of BF3 open regio- and stereoselectively the epoxide to produce
PDF
Album
Review
Published 08 Sep 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

Synthesis, α-mannosidase inhibition studies and molecular modeling of 1,4-imino-ᴅ-lyxitols and their C-5-altered N-arylalkyl derivatives

  • Martin Kalník,
  • Sergej Šesták,
  • Juraj Kóňa,
  • Maroš Bella and
  • Monika Poláková

Beilstein J. Org. Chem. 2023, 19, 282–293, doi:10.3762/bjoc.19.24

Graphical Abstract
  • in two steps from known ʟ-ribitol 1 [34] in good overall yield. Next, it was converted to the C-5 deoxygenated N-benzylpyrrolidine 6 via trityl ether cleavage, tosylation of the deprotected OH group, and reduction of the tosylate 5. Hydrogenolysis of the N-benzyl group in 6 followed by a removal of
PDF
Album
Supp Info
Full Research Paper
Published 06 Mar 2023

Investigation of cationic ring-opening polymerization of 2-oxazolines in the “green” solvent dihydrolevoglucosenone

  • Solomiia Borova and
  • Robert Luxenhofer

Beilstein J. Org. Chem. 2023, 19, 217–230, doi:10.3762/bjoc.19.21

Graphical Abstract
  • polymerization initiated with methyl triflate (black), methyl tosylate (red), 2-ethyl-3-methyl-2-oxazolinium triflate (blue), benzyl bromide (green) after 3 h of incubation and acyl chloride (dark red), propionyl chloride (purple) after 24 h of incubation in DLG. Peaks marked with asterisks originate from
PDF
Album
Supp Info
Full Research Paper
Published 28 Feb 2023

Preparation of β-cyclodextrin-based dimers with selectively methylated rims and their use for solubilization of tetracene

  • Konstantin Lebedinskiy,
  • Volodymyr Lobaz and
  • Jindřich Jindřich

Beilstein J. Org. Chem. 2022, 18, 1596–1606, doi:10.3762/bjoc.18.170

Graphical Abstract
  • as the price of some reagents and relatively low yields. In this work, we describe a method using low-cost reactants, such as methyl tosylate and DIPEA, which provides moderate to high yields and is a reasonable alternative to the existing methods. The acetylation of compound 1 using standard
  • conditions gave the secondary side peracetylated 6 in a high yield. The selective deprotection of silyl groups, yielding compound 7, was also performed standardly using BF3O·Et2O. Our new procedure is based on the methylation of this acetyl-protected β-CD 7 using a DIPEA/methyl tosylate mixture without
  • mixture decreases during the first several hours due to the quaternization of DIPEA. We found that adding small portions of the base over the reaction course allows us to avoid this problem and finish the process successfully. Purifying the reaction mixture requires the removal of excess methyl tosylate
PDF
Album
Supp Info
Full Research Paper
Published 25 Nov 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
  • of L6–8 with N-chlorosuccinimide (NCS) gave no chlorinated product, which was also observed in the reaction of L1 with NCS (Table 2, entries 1–4) [51]. Since the monomeric monopalladated tosylate complex of azobenzene I1-I was identified as an intermediate in the solid-state bromination of L1 (Figure
  •  1) [51], the analogous complexes of L6 and L7 were prepared to investigate whether the halogenation of the para-halogenated azobenzene derivatives follows the reaction pathway of the bromination of L1. The molecular structures of the isolated monopalladated tosylate complexes I6-I and I7-I solved
  • from laboratory powder X-ray diffraction (PXRD) data (Figure 4 and Figures S31 and S32 in Supporting Information File 1), are similar to that of complex I1-I [51] in which the palladium center is bound to the MeCN and tosylate (OTs) via nitrogen and oxygen, respectively, and to the azobenzene via the
PDF
Album
Supp Info
Full Research Paper
Published 15 Jun 2022

Mechanistic studies of the solvolysis of alkanesulfonyl and arenesulfonyl halides

  • Malcolm J. D’Souza and
  • Dennis N. Kevill

Beilstein J. Org. Chem. 2022, 18, 120–132, doi:10.3762/bjoc.18.13

Graphical Abstract
  • expanded to give a two-term Equation 2 [34][36][37], where l is the sensitivity to changes in N, which is a measure of solvent nucleophilicity. The original N scale was based on the specific rates of solvolysis of methyl tosylate [38] but, with this as the standard, there is a significant problem in
  • fluorosulfates [76]. In discussing the k4OE/k97TFE ratios, a good starting point is with the bridgehead-substituted 1-adamantyl chlorides, where a rear side SN2 process is impossible, with a value for the ratio of 0.83. The ratio raises to 1.38 for the tosylate but such variations are to be expected because
PDF
Album
Review
Published 17 Jan 2022

Stepwise PEG synthesis featuring deprotection and coupling in one pot

  • Logan Mikesell,
  • Dhananjani N. A. M. Eriyagama,
  • Yipeng Yin,
  • Bao-Yuan Lu and
  • Shiyue Fang

Beilstein J. Org. Chem. 2021, 17, 2976–2982, doi:10.3762/bjoc.17.207

Graphical Abstract
  • heating was to prevent the removal of the base-labile protecting group in 3 by the excess strong base by allowing the excess base to be consumed selectively via β-elimination of the tosylate in 1. The product of premature removal of the base-labile protecting group – an alkoxide – would complicate the
  • reaction, while the product of β-elimination of the tosylate – a vinyl ether – is inert under the reaction conditions. Compounds 3a–l were subjected to the study. All the compounds except 3h were found to be stable under the coupling conditions while product 5 was formed as indicated by TLC analysis
  • mixture was cooled to −78 °C, and then the monomer solution was added and the reaction mixture was warmed to room temperature slowly before heating to 60 °C. The careful manipulation of the temperature allowed the excess base to be selectively consumed via β-elimination of the tosylate of the monomer
PDF
Album
Supp Info
Full Research Paper
Published 28 Dec 2021

Synthesis and antimicrobial activity of 1H-1,2,3-triazole and carboxylate analogues of metronidazole

  • Satya Kumar Avula,
  • Syed Raza Shah,
  • Khdija Al-Hosni,
  • Muhammad U. Anwar,
  • Rene Csuk,
  • Biswanath Das and
  • Ahmed Al-Harrasi

Beilstein J. Org. Chem. 2021, 17, 2377–2384, doi:10.3762/bjoc.17.154

Graphical Abstract
  • group. The first step was initiated by the protection of the primary hydroxy group of metronidazole (1) with p-toluenesulfonyl chloride in dry DCM in the presence of triethylamine at 0 °C to room temperature. The reaction afforded the desired metronidazole tosylate 2 in high yield (96%) [17]. In the
  • next step, the metronidazole tosylate 2 under treatment with NaN3 in DMF at 70 °C afforded the corresponding metronidazide 3 in 88% yield [18]. The 1H NMR spectrum of metronidazide 3 showed a singlet at δ 7.93 for the 1H-imidazole proton. Two triplet signals at δ 4.40 and δ 3.74 were assigned to four
PDF
Album
Supp Info
Full Research Paper
Published 09 Sep 2021

Regioselective N-alkylation of the 1H-indazole scaffold; ring substituent and N-alkylating reagent effects on regioisomeric distribution

  • Ryan M. Alam and
  • John J. Keating

Beilstein J. Org. Chem. 2021, 17, 1939–1951, doi:10.3762/bjoc.17.127

Graphical Abstract
  • distribution. For example, employing C-7 NO2 or CO2Me substituted indazoles conferred excellent N-2 regioselectivity (≥ 96%). Importantly, we show that this optimized N-alkylation procedure tolerates a wide structural variety of alkylating reagents, including primary alkyl halide and secondary alkyl tosylate
  • conditions A and B (Table 6). The high selectivity observed for N-1 alkylation using NaH in THF (conditions A) was mainly effective using primary halide and tosylate compounds as electrophiles. Similar to the regiospecificity observed when employing n-pentyl bromide (ratio N-1 (10):N-2 (11) > 99:1, Table 1
  • , entry 22), its tosylate counterpart gave the corresponding N-1 regioisomer 10 with a high degree of N-1 regioselectivity (ratio N-1 (10):N-2 (11) = 76:1) under conditions A (Table 6, entry 1). Furthermore, conditions A could be successfully applied to the synthesis of benzyl and alicyclic indazole
PDF
Album
Supp Info
Full Research Paper
Published 02 Aug 2021

Cascade intramolecular Prins/Friedel–Crafts cyclization for the synthesis of 4-aryltetralin-2-ols and 5-aryltetrahydro-5H-benzo[7]annulen-7-ols

  • Jie Zheng,
  • Shuyu Meng and
  • Quanrui Wang

Beilstein J. Org. Chem. 2021, 17, 1481–1489, doi:10.3762/bjoc.17.104

Graphical Abstract
  • -hydroxy-4-furyl-tetralin 14af into the PAT analogue 22 (see Figure 1) [26]. The reaction of 14af with p-toluenesulfonyl chloride in pyridine afforded the tosylate 21 in 90% yield, which was then treated with 40% aqueous dimethylamine to produce the tertiary amine containing PAT analogue 22 (cis/trans
  • basicity, the long reaction time of 20 hours may lead to an ion-pair species with 21 and hence erode the stereochemistry. To prove this idea, we performed the reaction with CH2Cl2 as the solvent in the presence of 5.0 equivalents of pyridine and 2.0 equivalents of TsCl. Under these conditions, the tosylate
  • 21 was obtained with full retention of the expected stereochemistry (cis/trans =1:99) (see Supporting Information File 1 for details). The conversion of tosylate 21 to product 22 proceeded in a typical SN2 manner resulting in the expected inversion of the configuration. To unequivocally support the
PDF
Album
Supp Info
Full Research Paper
Published 22 Jun 2021

Total synthesis of ent-pavettamine

  • Memory Zimuwandeyi,
  • Manuel A. Fernandes,
  • Amanda L. Rousseau and
  • Moira L. Bode

Beilstein J. Org. Chem. 2021, 17, 1440–1446, doi:10.3762/bjoc.17.99

Graphical Abstract
  • tosylate under basic conditions affording 24 in a yield of 83%. The displacement of the tosyl group with an azide whilst heating the reaction at 80 °C allowed for the isolation of azide 25 in a good yield of 75%. Heating at higher temperatures resulted in product decomposition. Hydrogenation of the azide
PDF
Album
Supp Info
Full Research Paper
Published 10 Jun 2021

N-tert-Butanesulfinyl imines in the asymmetric synthesis of nitrogen-containing heterocycles

  • Joseane A. Mendes,
  • Paulo R. R. Costa,
  • Miguel Yus,
  • Francisco Foubelo and
  • Camilla D. Buarque

Beilstein J. Org. Chem. 2021, 17, 1096–1140, doi:10.3762/bjoc.17.86

Graphical Abstract
  • an intramolecular nucleophilic substitution of the tosylate of the resulting primary alcohol (Scheme 15). This methodology was applicable to the synthesis of 1-phenyl-2-azaspiro[3.4]octane (45, n = 2, R = Ph) and 1-phenyl-2-azaspiro[3.5]nonane (45, n = 3, R = Ph). The structure and absolute
PDF
Album
Review
Published 12 May 2021

CF3-substituted carbocations: underexploited intermediates with great potential in modern synthetic chemistry

  • Anthony J. Fernandes,
  • Armen Panossian,
  • Bastien Michelet,
  • Agnès Martin-Mingot,
  • Frédéric R. Leroux and
  • Sébastien Thibaudeau

Beilstein J. Org. Chem. 2021, 17, 343–378, doi:10.3762/bjoc.17.32

Graphical Abstract
  •  5) [38]. Tidwell et al. explored the influence of a CF3 group on the solvolysis reaction of various benzylic sulfonate derivatives [39][40]. They found a linear free-energy relationship between the solvolysis rate of sulfonate 13f in different solvents compared to the one of 2-adamantyl tosylate
  • /kCD3 ratio of 1.54, highlighting an isotopic effect consistent with a solvolysis mechanism involving a carbenium ion (kCH3/kCD3 = 1.48 for 2-methyl-2-adamantyl tosylate). Also, kH/kCF3 = 2⋅105 was established, illustrating the retarding α-CF3 effect in the production of a carbenium ion [41]. In the
  • group. Remarkably, they found that CF3 deactivates to such an extent that benzylic tosylate 13f was approximately 10 times less reactive than benzylic tosylate 17 (Figure 2, top). Similarly to the previous study, the Grunwald–Winstein plot [44] gave a linear free-energy relationship between the
PDF
Album
Review
Published 03 Feb 2021

The preparation and properties of 1,1-difluorocyclopropane derivatives

  • Kymbat S. Adekenova,
  • Peter B. Wyatt and
  • Sergazy M. Adekenov

Beilstein J. Org. Chem. 2021, 17, 245–272, doi:10.3762/bjoc.17.25

Graphical Abstract
  • in 105, which is associated with a difference in the regioselectivity of the C–C bond cleavage. The dissociation of the tosylate 104 to generate a cyclopropylmethyl carbocation A was accompanied by the cleavage of the proximal bond to form homoallylic products. The regioselectivity of the ring
PDF
Album
Review
Published 26 Jan 2021

Synthesis of imidazo[1,5-a]pyridines via cyclocondensation of 2-(aminomethyl)pyridines with electrophilically activated nitroalkanes

  • Dmitrii A. Aksenov,
  • Nikolai A. Arutiunov,
  • Vladimir V. Maliuga,
  • Alexander V. Aksenov and
  • Michael Rubin

Beilstein J. Org. Chem. 2020, 16, 2903–2910, doi:10.3762/bjoc.16.239

Graphical Abstract
  • , 1466, 1436, 1423, 1273, 1208, 1129, 1069 cm−1; HRESIMS (TOF) m/z: [M + H]+ calcd for C12H10BrN2O, 276.9971; found, 276.9974. Biologically active imidazo[1,5-a]pyridines. Activation of nitroalkanes towards nucleophilic attack by amines. Mechanistic rationale. Reaction of the N-tosylate 17 with
PDF
Album
Supp Info
Full Research Paper
Published 26 Nov 2020

Syntheses of spliceostatins and thailanstatins: a review

  • William A. Donaldson

Beilstein J. Org. Chem. 2020, 16, 1991–2006, doi:10.3762/bjoc.16.166

Graphical Abstract
  • -methyl-3-buten-1-yl tosylate in the presence of Grubbs’ 2nd generation catalyst yielded 59, which, upon elimination with potassium tert-butoxide led to the diene 50. The reductive amination of 50 afforded an inseparable mixture of the C-14 amines (6:1 ratio). However, the amidation of this mixture with
PDF
Album
Review
Published 13 Aug 2020

Facile synthesis of 7-alkyl-1,2,3,4-tetrahydro-1,8-naphthyridines as arginine mimetics using a Horner–Wadsworth–Emmons-based approach

  • Rhys A. Lippa,
  • John A. Murphy and
  • Tim N. Barrett

Beilstein J. Org. Chem. 2020, 16, 1617–1626, doi:10.3762/bjoc.16.134

Graphical Abstract
  • available iodide 29. The formation of compound 30 proceeded in 21% yield, with alcohol 31 and dimer 32 also formed in 20% and 5% yield, respectively (Scheme 9). Indeed, when iodide 29 was replaced with bromide 33 and tosylate 34 no formation of compound 30 was observed, with alcohol 31 and dimer 32
  • formation of dimer 28. Conditions: KOt-Bu, THF, 1 h, 68% yield. Alkylation of phosphoramidate 13 by iodide 29 to afford compound 30 and byproducts alcohol 31 and dimer 32. Use of bromide 33 or tosylate 34 afforded only compounds 31 and 32. Conditions: (i) s-BuLi (1.3 equiv), iodide 29, THF, −78 °C, 30 min
PDF
Album
Supp Info
Full Research Paper
Published 08 Jul 2020

Recent synthesis of thietanes

  • Jiaxi Xu

Beilstein J. Org. Chem. 2020, 16, 1357–1410, doi:10.3762/bjoc.16.116

Graphical Abstract
  • ring-opening followed by an intramolecular substitution to afford chiral thietane 175 [61]. 2.2.6 Synthesis via the nucleophilic ring-opening of three-membered heterocycles and subsequent displacement from aziridine-2-methyl tosylate: (1R,2S,6R)-6-Methyl-7-tosyl-7-azabicyclo[4.1.0]heptan-2-yl tosylate
  • (179) is a derivative of aziridine-2-methyl tosylate. After the ring-opening with ammonium tetrathiomolybdate and subsequent intramolecular cyclization, the compound was converted into a bridged thietane 183 in 75% yield. The results indicated that, in the ring-opening step, tetrathiomolybdate
PDF
Album
Review
Published 22 Jun 2020

Recent advances in Cu-catalyzed C(sp3)–Si and C(sp3)–B bond formation

  • Balaram S. Takale,
  • Ruchita R. Thakore,
  • Elham Etemadi-Davan and
  • Bruce H. Lipshutz

Beilstein J. Org. Chem. 2020, 16, 691–737, doi:10.3762/bjoc.16.67

Graphical Abstract
  • an attempt to determine the existence of radical behavior of PhMe2Si-MgMe (2), they studied the reaction of this Grignard reagent with dodecyl tosylate (1, X = OTs), which led to the formation of dodecyl silane 3 (20%) along with tridecane 4 (3%) and dodecane 5 (36%). Similarly, dodecyl bromide (1, X
  • the Oestreich group in 2016 [27]. The reaction could be performed using CuCN as catalyst in the absence of a ligand. A wide variety of triflates 9, including some containing a remote tosylate, bromide, alkene, or alkyne functionality, afforded the desired alkylsilanes 10–16 in fair to good yields
PDF
Album
Review
Published 15 Apr 2020

Photophysics and photochemistry of NIR absorbers derived from cyanines: key to new technologies based on chemistry 4.0

  • Bernd Strehmel,
  • Christian Schmitz,
  • Ceren Kütahya,
  • Yulian Pang,
  • Anke Drewitz and
  • Heinz Mustroph

Beilstein J. Org. Chem. 2020, 16, 415–444, doi:10.3762/bjoc.16.40

Graphical Abstract
  • anions promoting the solubility in organic surroundings. Representative anions in the case of cationic absorbers relate to tosylate [5], [n-C12H25-Ph-SO3−] [5], FAP ([(C2F5)3PF3]−)[71], NTf2 ([(CF3SO2)2N]−) [5] or aluminates ([Al(t-C4F9O)4]−) [6] – just to count a few possible examples. Thus, replacement
PDF
Album
Supp Info
Review
Published 18 Mar 2020

Architecture and synthesis of P,N-heterocyclic phosphine ligands

  • Wisdom A. Munzeiwa,
  • Bernard Omondi and
  • Vincent O. Nyamori

Beilstein J. Org. Chem. 2020, 16, 362–383, doi:10.3762/bjoc.16.35

Graphical Abstract
  • chiral acetal ligands have been reported by Lyle et al. where the fluorine–metal exchange was achieved by treatment with potassium tert-butoxide for a relatively long period (24 h) (Scheme 4) [65]. Acid-catalyzed condensation of compound 20 with enantiomerically pure C2-symmetric 1,2-tosylate analogs 21
PDF
Album
Review
Published 12 Mar 2020

Oligomeric ricinoleic acid preparation promoted by an efficient and recoverable Brønsted acidic ionic liquid

  • Fei You,
  • Xing He,
  • Song Gao,
  • Hong-Ru Li and
  • Liang-Nian He

Beilstein J. Org. Chem. 2020, 16, 351–361, doi:10.3762/bjoc.16.34

Graphical Abstract
  • designable structure, tunable properties as well as superior solubility [26][27]. Furthermore, the thermal stability and negligible vapor pressure of ILs can facilitate the product separation after reaction. To the best of our knowledge, the IL 1-butylsulfonic-3-methylimidazolium tosylate ([HSO3-BMim]TS) can
PDF
Album
Supp Info
Full Research Paper
Published 10 Mar 2020

A chiral self-sorting photoresponsive coordination cage based on overcrowded alkenes

  • Constantin Stuckhardt,
  • Diederik Roke,
  • Wojciech Danowski,
  • Edwin Otten,
  • Sander J. Wezenberg and
  • Ben L. Feringa

Beilstein J. Org. Chem. 2019, 15, 2767–2773, doi:10.3762/bjoc.15.268

Graphical Abstract
  • formed revealing that a chiral self-sorting process takes place. In addition, two of the cage isomers can bind a tosylate anion in solution by formation of a host–guest complex. Results and Discussion Ligands Z-1 and E-1 (Scheme 1) were synthesized by a Suzuki cross-coupling reaction of 3
  • -sorting. Next, we were interested in the guest binding abilities of cages Pd2(stable Z-1)4 and Pd2(stable E-1)4. The tosylate anion was chosen as it has the appropriate size to fit inside the cages. A Job plot analysis revealed a 1:1 binding stoichiometry between both cage isomers and OTs− (Figures S3–S5
  • , Supporting Information File 1), which corresponds to the model in which OTs− serves as a guest molecule which is encapsulated inside the cages [55][56]. 1H NMR titrations with tetrabutylammonium tosylate revealed that both cages are able to bind OTs−, showing similar binding strengths (KB = 1604 ± 39 M−1 for
PDF
Album
Supp Info
Full Research Paper
Published 15 Nov 2019
Other Beilstein-Institut Open Science Activities