共查询到20条相似文献,搜索用时 15 毫秒
1.
2.
Dr. Jacek Cieślak Dr. Cristina Ausín Dr. Andrzej Grajkowski Dr. Serge L. Beaucage 《Chemistry (Weinheim an der Bergstrasse, Germany)》2013,19(14):4623-4632
The reaction of 2‐cyano‐2‐methyl propanal with 2′‐O‐aminooxymethylribonucleosides leads to stable and yet reversible 2′‐O‐(2‐cyano‐2,2‐dimethylethanimine‐N‐oxymethyl)ribonucleosides. Following N‐protection of the nucleobases, 5′‐dimethoxytritylation and 3′‐phosphitylation, the resulting 2′‐protected ribonucleoside phosphoramidite monomers are employed in the solid‐phase synthesis of three chimeric RNA sequences, each differing in their ratios of purine/pyrimidine. When the activation of phosphoramidite monomers is performed in the presence of 5‐benzylthio‐1H‐tetrazole, coupling efficiencies averaging 99 % are obtained within 180 s. Upon completion of the RNA‐chain assemblies, removal of the nucleobase and phosphate protecting groups and release of the sequences from the solid support are carried out under standard basic conditions, whereas the cleavage of 2′‐O‐(2‐cyano‐2,2‐dimethylethanimine‐N‐oxymethyl) protective groups is effected (without releasing RNA alkylating side‐products) by treatment with tetra‐n‐butylammonium fluoride (0.5 m) in dry DMSO over a period of 24–48 h at 55 °C. Characterization of the fully deprotected RNA sequences by polyacrylamide gel electrophoresis (PAGE), enzymatic hydrolysis, and matrix‐assisted laser desorption/ionization (MALDI) mass spectrometry confirmed the identity and quality of these sequences. Thus, the use of 2′‐O‐aminooxymethylribonucleosides in the design of new 2′‐hydroxyl protecting groups is a powerful approach to the development of a straightforward, efficient, and cost‐effective method for the chemical synthesis of high‐quality RNA sequences in the framework of RNA interference applications. 相似文献
3.
Dr. Renhua Qiu Prof. Dr. Xinhua Xu Dr. Lifeng Peng Dr. Yalei Zhao Dr. Ningbo Li Prof. Dr. Shuangfeng Yin 《Chemistry (Weinheim an der Bergstrasse, Germany)》2012,18(20):6172-6182
Strong Lewis acids of air‐stable metallocene bis(perfluorooctanesulfonate)s [M(Cp)2][OSO2C8F17]2?nH2O?THF (M=Zr ( 2 a ?3 H2O?THF), M=Ti ( 2 b ?2 H2O?THF)) were synthesized by the reaction of [M(Cp)2]Cl2 (M=Zr ( 1 a ), M=Ti ( 1 b )) with nBuLi and C8F17SO3H (2 equiv) or with C8F17SO3Ag (2 equiv). The hydrate numbers (n) of these complexes were variable, changing from 0 to 4 depending on conditions. In contrast to well‐known metallocene triflates, these complexes suffered no change in open air for a year. thermogravimetry–differential scanning calorimetry (TG‐DSC) analysis showed that 2 a and 2 b were thermally stable at 300 and 180 °C, respectively. These complexes exhibited unusually high solubility in polar organic solvents. Conductivity measurement showed that the complexes ( 2 a and 2 b ) were ionic dissociation in CH3CN solution. X‐ray analysis result confirmed 2 a ?3 H2O?THF was a cationic organometallic Lewis acid. UV/Vis spectra showed a significant red shift due to the strong complex formation between 10‐methylacridone and 2 a . Fluorescence spectra showed that the Lewis acidity of 2 a fell between those of Sc3+ (λem=474 nm) and Fe3+ (λem=478 nm). ESR spectra showed the Lewis acidity of 2 a (0.91 eV) was at the same level as that of Sc3+ (1.00 eV) and Y3+ (0.85 eV), while the Lewis acidity of 2 b (1.06 eV) was larger than that of Sc3+ (1.00 eV) and Y3+ (0.85 eV). They showed high catalytic ability in carbonyl‐compound transformation reactions, such as the Mannich reaction, the Mukaiyama aldol reaction, allylation of aldehydes, the Friedel–Crafts acylation of alkyl aromatic ethers, and cyclotrimerization of ketones. Moreover, the complexes possessed good reusability. On account of their excellent catalytic efficiency, stability, and reusability, the complexes will find broad catalytic applications in organic synthesis. 相似文献
4.
2‐Nitroveratryl as a Photocleavable Thiol‐Protecting Group for Directed Disulfide Bond Formation in the Chemical Synthesis of Insulin 下载免费PDF全文
John A. Karas Dr. Denis B. Scanlon Prof. Briony E. Forbes Dr. Irina Vetter Prof. Richard J. Lewis Dr. James Gardiner Prof. Frances Separovic Prof. John D. Wade Dr. Mohammed A. Hossain 《Chemistry (Weinheim an der Bergstrasse, Germany)》2014,20(31):9549-9552
Chemical synthesis of peptides can allow the option of sequential formation of multiple cysteines through exploitation of judiciously chosen regioselective thiol‐protecting groups. We report the use of 2‐nitroveratryl (oNv) as a new orthogonal group that can be cleaved by photolysis under ambient conditions. In combination with complementary S‐pyridinesulfenyl activation, disulfide bonds are formed rapidly in situ. The preparation of Fmoc‐Cys(oNv)‐OH is described together with its use for the solid‐phase synthesis of complex cystine‐rich peptides, such as insulin. 相似文献
5.
6.
A 3,4‐trans‐Fused Cyclic Protecting Group Facilitates α‐Selective Catalytic Synthesis of 2‐Deoxyglycosides 下载免费PDF全文
Dr. Edward I. Balmond Dr. David Benito‐Alifonso Dr. Diane M. Coe Prof. Roger W. Alder Dr. Eoghan M. McGarrigle Dr. M. Carmen Galan 《Angewandte Chemie (International ed. in English)》2014,53(31):8190-8194
A practical approach has been developed to convert glucals and rhamnals into disaccharides or glycoconjugates with high α‐selectivity and yields (77–97 %) using a trans‐fused cyclic 3,4‐O‐disiloxane protecting group and TsOH?H2O (1 mol %) as a catalyst. Control of the anomeric selectivity arises from conformational locking of the intermediate oxacarbenium cation. Glucals outperform rhamnals because the C6 side‐chain conformation augments the selectivity. 相似文献
7.
The phosphodiester linkage of 3′‐O‐levulinoylthymidine 5′‐methylphosphate ( 5 ) has been protected with 2‐[(acetyloxy)methyl]‐4‐(acetylsulfanyl)‐2‐(ethoxycarbonyl)‐3‐oxobutyl group (to give 1 ) to study the potential of this group as an esterase‐ and thermolabile protecting group. The group turned out to be unexpectedly thermolabile, being removed as ethyl 3‐(acetyloxy)‐4‐(acetylsulfanyl)‐2‐methylidenebut‐3‐enoate ( 10 ) without accumulation of any intermediates. The half‐life of this reaction at pH 7.5 and 37° is 14 min. Hog liver esterase (HLE), in turn, removes the protecting group as ethyl 4‐(acetylsulfanyl)‐2‐methylidene‐3‐oxobutanoate ( 12 ). On using 2.6 units of HLE in 1 ml, the rate of the enzymatic deprotection was still only one third of that of the nonenzymatic reaction. The mechanisms of both reactions have been studied and discussed. The crucial step seems to be removal of the O‐bound Ac group, either by esterase or by migration to the neighboring 3‐oxo group (nonenzymatic removal). This triggers the removal by retro‐aldol condensation/elimination mechanism. No alkylation of glutathione (GSH) upon the deprotection of 1 could be detected. 相似文献
8.
2‐Methoxy‐4‐methylsulfinylbenzyl: A Backbone Amide Safety‐Catch Protecting Group for the Synthesis and Purification of Difficult Peptide Sequences 下载免费PDF全文
Marta Paradís‐Bas Dr. Judit Tulla‐Puche Prof. Fernando Albericio 《Chemistry (Weinheim an der Bergstrasse, Germany)》2014,20(46):15031-15039
The use of 2‐methoxy‐4‐methylsulfinylbenzyl (Mmsb) as a new backbone amide‐protecting group that acts as a safety‐catch structure is proposed. Mmsb, which is stable during the elongation of the sequence and trifluoroacetic acid‐mediated cleavage from the resin, improves the synthetic process as well as the properties of the quasi‐unprotected peptide. Mmsb offers the possibility of purifying and characterizing complex peptide sequences, and renders the target peptide after NH4I/TFA treatment and subsequent ether precipitation to remove the cleaved Mmsb moiety. First, the “difficult peptide” sequence H‐(Ala)10‐NH2 was selected as a model to optimize the new protecting group strategy. Second, the complex, bioactive Ac‐(RADA)4‐NH2 sequence was chosen to validate this methodology. The improvements in solid‐phase peptide synthesis combined with the enhanced solubility of the quasi‐unprotected peptides, as compared with standard sequences, made it possible to obtain purified Ac‐(RADA)4‐NH2. To extend the scope of the approach, the challenging Aβ(1‐42) peptide was synthesized and purified in a similar manner. The proposed Mmsb strategy opens up the possibility of synthesizing other challenging small proteins. 相似文献
9.
10.
Back Cover: 2‐Methoxy‐4‐methylsulfinylbenzyl: A Backbone Amide Safety‐Catch Protecting Group for the Synthesis and Purification of Difficult Peptide Sequences (Chem. Eur. J. 46/2014) 下载免费PDF全文
Marta Paradís‐Bas Dr. Judit Tulla‐Puche Prof. Fernando Albericio 《Chemistry (Weinheim an der Bergstrasse, Germany)》2014,20(46):15256-15256
11.
A Convergent Strategy for the Synthesis of Type‐1 Elongated Mucin Cores 1–3 and the Corresponding Glycopeptides 下载免费PDF全文
Dipl. Chem. Christian Pett Dr. Ulrika Westerlind 《Chemistry (Weinheim an der Bergstrasse, Germany)》2014,20(24):7287-7299
Mucins are a class of highly O‐glycosylated proteins found on the surface of cells in epithelial tissues. O‐Glycosylation is crucial for the functionality of mucins and changes therein can have severe consequences for an organism. With that in mind, the elucidation of interactions of carbohydrate binding proteins with mucins, whether in morbidly altered or unaltered conditions, continue to shed light on mechanisms involved in diseases like chronic inflammations and cancer. Despite the known importance of type‐1 and type‐2 elongated mucin cores 1–4 in glycobiology, the corresponding type‐1 structures are much less well studied. Here, the first chemical synthesis of extended mucin type‐1 O‐glycan core 1–3 amino acid structures based on a convergent approach is presented. By utilizing differentiation in acceptor reactivity, shared early stage Tn‐ and T‐acceptor intermediates were elongated with a common type‐1 [β‐D ‐Gal‐1,3‐β‐D ‐GlcNAc] disaccharide, which allows for straightforward preparation of diverse glycosylated amino acids carrying the type‐1 mucin core 1–3 saccharides. The obtained glycosylated 9‐fluorenylmethoxycarbonyl (Fmoc)‐protected amino acid building blocks were employed in synthesis of type‐1 mucin glycopeptides, which are useful in biological applications. 相似文献
12.
Elisabeth Kapatsina Markus Mateescu Angelika Baro Wolfgang Frey Sabine Laschat 《Helvetica chimica acta》2009,92(10):2024-2037
The [1,1′‐biisoquinoline]‐4,4′‐diol ( 4a ), which was obtained as hydrochloride 4a ?2 HCl in two steps starting from the methoxymethyl (MOM)‐protected 1‐chloroisoquinoline 8 (Scheme 3), opens access to further O‐functionalized biisoquinoline derivatives. Compound 4a ?2 HCl was esterified with 4‐(hexadecyloxy)benzoyl chloride ( 5b ) to give the corresponding diester 3b (Scheme 4), which could not be obtained by Ni‐mediated homocoupling of 6b (Scheme 2). The ether derivative 2b was accessible in good yield by reaction of 4a ?2 HCl with the respective alkyl bromide 9 under the conditions of Williamson etherification (Scheme 4). Slightly modified conditions were applied to the esterification of 4a ?2 HCl with galloyl chlorides 10a – h as well as etherification of 4a ?2 HCl with 6‐bromohexyl tris(alkyloxy)benzoates 11b , d – h and [(6‐bromohexyl)oxy]‐substituted pentakis(alkyloxy)triphenylenes 14a – c (Scheme 5). Despite the bulky substituents, the respective target 1,1′‐biisoquinolines 12, 13 , and 15 were isolated in 14–86% yield (Table). 相似文献
13.
The electronic coupling between two amine redox sites bridged through the 5,5′‐positions of the [Re(CO)3Cl]‐chelated 2,2′‐bipyridine was studied by the electrochemical, spectroscopic, and EPR analysis. Interestingly, multiple near‐infrared bands were observed in this new organic mixed‐valent system. The results are interpreted with the aid of DFT and TDDFT calculations. 相似文献
14.
Chin‐Ping Yang Kuei‐Shi Hung Ruei‐Shin Chen 《Journal of polymer science. Part A, Polymer chemistry》2000,38(21):3954-3961
Organosoluble homopolyimides (PIs) and copolyimides (CoPIs) were synthesized from 2,2‐bis[4‐(4‐aminophenoxy)phenyl]propane (BAPP) or 2,2‐bis[4‐(4‐aminophenoxy)phenyl]hexafluoropropane (6FBAPP) and six kinds of commercial aromatic dianhydrides (PMDA, II a ; BTDA, II b ; BPDA, II c ; ODPA, II d ; DSDA, II e ; 6FDA, II f ). Although BAPP and II d∼f could prepare three kinds of soluble PIs ( III‐A d∼f ), likewise 6FBAPP and II c∼f could prepare four PIs ( III‐B c∼f ), the insoluble PIs were synthesized from these two diamines and other dianhydrides. However, soluble CoPIs could be prepared by alternative copolycondensation from a pair of dianhydrides of soluble PIs and insoluble PIs in certain molar ratios (m1/m2). The ratios of m1/m2 of BAPP/PMDA series CoPIs ( IV m1(d–f)/m2a ) ranged from 3–5, but ratios of 6FBAPP/PMDA series CoPIs ( V m1(c∼f)/m2a ) decreased to 2–3. The m1/m2 of the BAPP/BTDA and 6FBAPP/BTDA series CoPIs were 2, while the BAPP/BPDA series were between 1–2. Composition, solubility, tensile properties and thermal properties of these CoPIs synthesized via a two‐stage thermal cyclodehydration were determined and were compared with their corresponding PIs. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 3954–3961, 2000 相似文献
15.
Valerie V. Sheares Lifen Wu Yunxiao Li Tatania K. Emmick 《Journal of polymer science. Part A, Polymer chemistry》2000,38(22):4070-4080
The bulk free‐radical polymerization of 2‐[(N,N‐dialkylamino)methyl]‐1,3‐butadiene with methyl, ethyl, and n‐propyl substituents was studied. The monomers were synthesized via substitution reactions of 2‐bromomethyl‐1,3‐butadiene with the corresponding dialkylamines. For each monomer the effects of the polymerization initiator, initiator concentration, and reaction temperature on the final polymer structure, molecular weight, and glass‐transition temperature (Tg) were examined. Using 2,2′‐azobisisobutyronitrile as the initiator at 75 °C, the resulting polymers displayed a majority of 1,4 microstructures. As the temperature was increased to 100 and 125 °C using t‐butylperacetate and t‐butylhydroperoxide, the percentage of the 3,4 microstructure increased. Differential scanning calorimetry indicated that all of the Tg values were lower than room temperature. The Tg values were higher when the majority of the polymer structure was 1,4 and decreased as the percentage of the 3,4 microstructure increased. The Diels–Alder side products found in the polymer samples were characterized using NMR and gas chromatography‐mass spectrometry methods. The polymerization temperature and initiator concentration were identified as the key factors that influenced the Diels–Alder dimer yield. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 4070–4080, 2000 相似文献
16.
17.
18.
Markus W. Weishaupt Stefan Matthies Prof. Dr. Peter H. Seeberger 《Chemistry (Weinheim an der Bergstrasse, Germany)》2013,19(37):12497-12503
β‐Glucans are a group of structurally heterogeneous polysaccharides found in bacteria, fungi, algae and plants. β‐(1,3)‐D ‐Glucans have been studied in most detail due to their impact on the immune system of vertebrates. The studies into the immunomodulatory properties of these glucans are typically carried out with isolates that contain a heterogeneous mixture of polysaccharides of different chain lengths and varying degrees of branching. In order to determine the structure–activity relationship of β‐(1,3)‐glucans, access to homogeneous, structurally‐defined samples of these oligosaccharides that are only available through chemical synthesis is required. The syntheses of β‐glucans reported to date rely on the classical solution‐phase approach. We describe the first automated solid‐phase synthesis of a β‐glucan oligosaccharide that was made possible by innovating and optimizing the linker and glycosylating agent combination. A β‐(1,3)‐glucan dodecasaccharide was assembled in 56 h in a stereoselective fashion with an average yield of 88 % per step. This automated approach provides means for the fast and efficient assembly of linker‐functionalized mono‐ to dodecasaccharide β‐(1,3)‐glucans required for biological studies. 相似文献
19.
Norbert Szimhardt Michael S. Gruhne Marcus Lommel Andreas Hess Maximilian H. H. Wurzenberger Thomas M. Klaptke Jrg Stierstorfer 《无机化学与普通化学杂志》2019,645(3):354-361
This contribution covers the preparation and characterization of 2,2‐bis(5‐tetrazolyl)propane (5‐DTP) ( 1 ). The bridged bitetrazole is used as a neutral nitrogen‐rich ligand in 3d transition metal(II) based complexes for the first time and can be synthesized via [2+3] cycloaddition from sodium azide and dimethylmalononitrile. The combination with different anions (e.g., perchlorate, nitrate, sulfate, and chloride) yields materials with widely varying physicochemical properties. The obtained coordination compounds were characterized using low‐temperature single‐crystal X‐ray diffraction (except 14 ), IR spectroscopy, elemental analysis, and DTA (except 16 ). The sensitivities toward external stimuli (impact and friction) were determined according to the Bundesamt für Materialforschung und ‐prüfung (BAM) standard methods together with its sensitivities against electrostatic discharge (except 16 ). Complexes 10 and 14 were characterized in laser ignition experiments. For determination of the compounds' deflagration to detonation transition (DDT) capability, hot plate and hot needle tests were performed for the zinc(II) and copper(II) perchlorate complexes. 相似文献