首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Nucleotides, XVII. Synthesis of Homogeneous Adenosyl-3′,5′-oligomers by the Phosphotriester Method The chemical synthesis of the fully protected trimer 12 , the tetramers 11 and 23 as well as the pentamer 14 was achieved in preparative scales starting from the fully blocked adenosine-3′-phosphotriesters 1, 2 , and N6-benzoyl-2′,3′-bis-O-(tert-butyldimethylsilyl)adenosine (6) . All intermediates and end products have been isolated, purified and characterized by elemental analyses, UV, and CD spectra. Deprotection of the various blocking groups proceeds without difficulties to afford the free trimeric, tetrameric, and pentameric oligoadenylates 15, 16 , and 24 in high yields.  相似文献   

2.
A DNA duplex containing an N3-thymidine–butylene–N3-thymidine interstrand cross-link (ICL) was prepared using an on-column orthogonal deprotection strategy to permit different nucleotide sequence composition around the cross-linked site. The conditions used to remove 5′-O-allyloxycarbonyl and 3′-O-tert-butyldimethylsilyl protective groups for various on-column oligonucleotide intermediates did not affect the cross-linked lesion. Efficient removal of these groups enabled successful coupling of 2′-deoxyphosphoramidites to produce the desired duplex with a 31% yield after deprotection and purification.  相似文献   

3.
The electron impact mass spectra of monosilyl and mixed acyl-silyl derivatives of 2′-deoxynucleosides are described in detail. (Silyl = tert-butyldimethylsilyl, cyclo-tetramethylene-isopropylsilyl, or cyclo-tetramethylene-tert-butylsilyl; acyl = acetyl or trifluoroacetyl.) The interpretation of the fragmentation pathways was aided by metastable ion decomposition studies, precise mass and deuterium labelling measurements. Mass spectrally, the acyl substituents are mostly ‘passive’ and have (with possibly one exception) little fragmentation directing capability. In contrast, the silyl groups have powerful fragmentation directing properties. Elimination of the bulky alkyl radical R˙ (tert-butyl or isopropyl) from the molecular ion produces the siliconium ion, [M–R]+, which is the precursor for most of the other prominent ions in the spectra. These arise from ‘siliconium ion rearrangements’ resulting from the interaction of the positively charged siliconium ion centre with the electron dense regions (i.e. oxygens) in the molecule, to form cyclic silyloxonium ions which subsequently decompose. Since the interacting oxygen and silicon must be sterically accessible, the fragment ion types and their abundances are very dependent upon structure. Consequently, [M–R]+ ions formed from 3′- or 5′-O-silyl groups give rise to different sets of daughter ions which, for the most part, are not found, or have very low abundances, in the mass spectra of underivatized or trimethylsilylated nucleosides. Detailed information on sugar and base moieties and isomeric substitution is readily obtained.  相似文献   

4.
The synthesis of a series of heterocyclic analogues of (±)-4-((αR*)-α-((2S*,5R*)-4-allyl-2,5-dimethyl-1-piperazmyl)-3-hydroxybenzyl)-N,N-diethylbenzarrude (BW373U86) for screening against opioid receptors is described. The intermediate α-heterocyclic benzyl alcohols 24 were synthesized either by low temperature reaction of lithioheterocycles with 3-((tert-butyldimethylsilyl)oxy)benzaldehyde ( 10 ) or by reaction of 3-((tert-butyldimethylsilyl)oxy)phenylmagnesium bromide ( 19 ) with heterocyclic carbaldehydes. The α-heterocyclic benzyl alcohols 24 were converted to chloromethines ( 25 ) with thionyl chloride and used to alkylate with trans-1-allyl-2,5-dimethylpiperazine ( 5 ) to give diastereomeric pairs of the target compounds. The bromoheterocycles were then derivatized to produce amides. Compounds that are potent and selective for the 5 or μ opioid receptors and some mixed δ/μ analogues are reported.  相似文献   

5.
To measure the hydrophobic character of the ribose moiety of doridosine on the adenosine receptors, 2′,3′-didehydro-2′,3′-dideoxydoridosine (2) and 2′,3′-dideoxydoridosine (3) were prepared. Initial treatment of doridosine with N,N-dimethylformamide diethylacetal, and subsequently with tert-butyldimethylsilyl chloride gave 5. Compound 5 was then reacted with 1,1′-thiocarbonyldiimidazole and the resulting thionocarbonate 6 was heated with triethyl phosphite at 135°C to afford 7. Treatment of compound 7 with tetrabutylammonium fluoride and methanolic ammonia furnished compound 2 in good yield. Compound 2 was subjected to catalytic hydrogenation affording compound 3 in 85% yield.  相似文献   

6.
7.
The 1H and 13C NMR spectra of several isomeric N-substituted tetrazoles have been investigated. 13C NMR is shown to be more useful for distinguishing between structural isomers of N-substituted tetrazoles except for those carrying electropositive substituents like SnBu3. Correlations of δC-5 (inverse) and 1J(C-5,H) with s?1 found for 1-substituted tetrazole allowed the identification of the N SnBu3 derivative as 1-(tri-n-butylstannyl)tetrazole. The phenyl carbon chemical shift difference ΔC′ = δC-3′-δC-2′ is insignificant for structure elucidation and conformational studies of N-substituted 5-phenyltetrazoles; ΔH′ from 1H NMR spectra seems to be more useful.  相似文献   

8.
9.
The diamine 1,4-bis(4-aminophenoxy)-2,5-di-tert-butylbenzene, containing symmetric, bulky di-tert-butyl substituents and a flexible ether unit, was synthesized and used to prepare a series of polyamides by the direct polycondensation with various aromatic dicarboxylic acids in N-methyl-2-pyrrolidinone (NMP) using triphenyl phosphite and pyridine as condensing agents. All the polymers were obtained in quantitative yields with inherent viscosities of 0.32–1.27 dL g−1. Most of these polyamides, except II a , II d , and II e , showed an amorphous nature and dissolved in polar solvents and less polar solvents. Polyamides derived from 4,4′-sulfonyldibenzoic acid, 4,4′-(hexafluoro-isopropylidene)dibenzoic acid, and 5-nitroisophthalic acid were even soluble in a common organic solvent such as THF. Most polyamide films could be obtained by casting from their N,N-dimethylacetamide (DMAc) solutions. The polyamide films had a tensile strength range of 49–78 MPa, an elongation range at break of 3–5%, and a tensile modulus range of 1.57–2.01 GPa. These polyamides had glass transition temperatures ranging between 253 and 276°C, and 10% mass loss temperatures were recorded in the range 402–466°C in nitrogen atmosphere. © 1998 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 36: 1069–1074, 1998  相似文献   

10.
The direct enantioselective analysis of 3-, 4-, and 5-hydroxy fatty acids from biological material has been achieved by enantioselective multidimensional gas chromatography (enantio-MDGC) with heptakis(2,3-di-O-methyl-6-O-tert-butyldimethylsilyl)- or (2,3-di-O-acetyl-6-O-tert-butyldimethylsilyl)-β-cyclodextrin as chiral stationary phase. All the bacteria investigated produced polyesters of enatiomerically pure (R) configured compounds.  相似文献   

11.
Isolation and Structure of Long-Chain Alkylphenols and -catechols from Plectranthus albidus (Labiatae) From the title plant, a series of even-numbered long-chain, phenol- or pyrocatechol-derived 1-arylalkan-5-ones was isolated by classical chromatography and preparative reversed phase HPLC. By chemical and spectroscopic methods, including coupled chromatographic techniques (GC/MS/FT-IR, HPLC/MS), their structures were established to be 1-(4′-hydroxyphenyl)tetradecan-5-one ( 2a ), 1-(4′-hydroxyphenyl)hexadecan-5-one ( 2b ), 1-(4′-hydroxyphenyl)octadecan-5-one ( 2c ), and (Z)-1-(4′-hydroxyphenyl)octadec-13-en-5-one ( 2d ); (E,E)-1-(3′,4′-dihydroxyphenyl)deca-1,3-dien-5-one ( 1a ), 1-(3′,4′-dihydroxyphenyl)dodecan-5-one ( 3a ), 1-(3′,4′-dihydroxyphenyl)-tetradecan-5-one ( 3b ), 1-(3′,4′-dihydroxyphenyl)hexadecan-5-one ( 3c ), 1-(3′,4′-dihydroxyphenyl)octadecan-5-one ( 3d ), 1-(3′,4′-dihydroxyphenyl)icosan-5-one ( 3e ), and (Z)-1-(3′,4′-dihydroxyphenyl)octadec-13-en-5-one ( 3f ). In vitro, the compounds show significant antioxidant activity, the inhibitory concentration of the most potent one, 1a , being slightly lower than for 2-(tert-butyl)-4-methoxyphenol (BHA) and 2,6-di(tert-butyl)-4-methylphenol (BHT) in the Fe2+-catalysed autooxidation of linoleic acid, whereas the acitivities of phenols 2a–d are in the same order of magnitude as α-tocopherol.  相似文献   

12.
The antivirally active 3′-deoxyadenylyl-(2′–5′)-3′-deoxyadenylyl-(2′–5′)-3′-deoxyadenosine (cordycepin trimer core) was modified at the 2′- or 5′-terminus, by attachment of cholesterol via a carbonate bond (→ 15 ) or a succinate linker (→ 16 and 27 ) to improve cell permeability. The corresponding monomeric conjugates 4 , 7 , and 21 of cordycepin were prepared as model substances to study the applicability of the anticipated protecting groups – the monomethoxytrityl (MeOTr), the (tert-butyl)dimethylsilyl (tbds), and the β -eliminating 2-(4-nitrophenyl)ethyl (npe) and 2-(4-nitrophenyl)ethoxycarbonyl (npeoc) groups – for the final deblocking steps without harming the ester bonds of the conjugate trimers. The syntheses were performed in solution using phosphoramidite chemistry. The fully protected trimer conjugates 13 , 14 , and 26 as well as all intermediates were characterized by elemental analyses, UV and 1H-NMR spectra. The deblocked conjugates 15 , 16 , and 27 were pure according to HPLC and showed the correct compositions by mass spectra. Comparative biological studies indicated that cordycepincholesterol conjugate trimers 16 and 27 were 333- and 1000-fold, respectively, more potent inhibitors of HIV-1-induced syncytia formation than cordycepin trimer core.  相似文献   

13.
Alternative syntheses of 9-(3-deoxy-β-D-threo-pentofuranosyl)adenine ( 4 ), the core nucleoside of agrocin 84 [and its 2′-deoxy threo isomer 5 ] were devised: (1) direct conversion of 9-(β-D-arabinofuranosyl)adenine into 9-(2,3-anhydro-β-D-lyxofuranosyl)adenine and regioselective opening of its oxirane ring with sodium borohy-dride to give 4 and 5 (?7.5:1); (2) treatment of adenosine with sodium hydride and 2,4,6-triisopropylbenzene-sulfonyl chloride, and subjection of the resulting 2′(3′)-sulfonates to the reductive [1,2]-hydride shift rearrangement with lithium triethylborohydride to give 4 and 5 (? 2:1); and (3) subjection of the phenoxythiocar-bonyl esters of 9-[2(3),5-bis-O-(tert-butyldimethylsilyl)-β-D-arabinofuranosyl]adenine to Barton deoxygenation, and deprotection to give 4 and 2′-deoxyadenosine (?5:1). Methods (2) and (3) gave lower yields. Syntheses of simplified 6-N- and 5′-O-adenosine phosphoramidate model compounds were explored to examine potential access to such features in the structure proposed for agrocin 84.  相似文献   

14.
Blue fluorescent materials with a 2‐(diphenylamino)fluoren‐7‐ylvinylarene emitting unit and tert‐butyl‐based blocking units were synthesized. The photophysical properties of these materials, including UV/Vis absorption, photoluminescent properties, and HOMO–LUMO energy levels, were characterized and rationalized with quantum‐mechanical DFT calculations. The electroluminescent properties of these molecules were examined through the fabrication of multilayer devices with a structure of indium–tin oxide, 4,4′‐bis{N‐[4‐(N,N‐di‐m‐tolylamino)phenyl]‐N‐phenylamino}biphenyl, 4′‐bis[N‐(1‐naphthyl)‐N‐phenylamino]biphenyl, and blue materials doped in 2‐methyl‐9,10‐di(2‐naphthyl)anthracene/tris(8‐quinolinolato)aluminum/LiF/Al. All devices exhibit highly efficient blue electroluminescence with high external quantum efficiency (3.20–7.72 % at 20 mA cm?2). A deep‐blue device with Commission Internationale de l’Eclairage (CIE) coordinates of (0.15, 0.11) that uses 7‐[2‐(3′,5′‐di‐tert‐butylbiphenyl‐4‐yl)vinyl]‐9,9‐diethyl‐2‐N‐(3,5‐di‐tert‐butylphenyl)‐2,4‐difluorobenzenamino‐9H‐fluorene as a dopant in the emitting layer showed a luminous efficiency and external quantum efficiency of 3.95 cd A?1 and 4.23 % at 20 mA cm?2, respectively. Furthermore, a highly efficient sky‐blue device that uses the dopant 7‐{2‐[2‐(3,5‐di‐tert‐butylphenyl)‐9,9′‐spirobifluorene‐7‐yl]vinyl}‐9,9‐diethyl‐2‐N,N‐diphenylamino‐9H‐fluorene exhibited a luminous efficiency and high quantum efficiency of 10.3 cd A?1 and 7.7 % at 20 mA cm?2, respectively, with CIE coordinates of (0.15, 0.20).  相似文献   

15.
Silylhydrazines and Dimeric N,N′‐Dilithium‐N,N′‐bis(silyl)hydrazides – Syntheses, Reactions, Isomerisations Di‐tert.‐butylchlorosilane reacts with dilithiated hydrazine in a molar ratio to give the N,N′‐bis(silyl)hydrazine, [(Me3C)2SiHNH]2, ( 5 ). Isomeric tris(silyl)hydrazines, N‐difluorophenylsilyl‐N′,N′‐bis(dimethylphenylsilyl)hydrazine ( 7 ) and N‐difluorophenylsilyl‐N,N′‐bis(dimethylphenylsilyl)hydrazine ( 8 ) are formed in the reaction of N‐lithium‐N′‐N′‐bis(dimethylphenylsilyl)hydrazide and F3SiPh. Isomeric bis(silyl)hydrazines, (Me3C)2SiFNHNHSiMe2Ph ( 9 ) and (Me3C)2‐ SiF(PhMe2Si)N–NH2 ( 10 ) are the result of the reaction of di‐tert.‐butylfluorosilylhydrazine and ClSiMe2Ph in the presence of Et3N. Quantum chemical calculations for model compounds demonstrate the dyotropic course of the rearrangement. The monolithium derivative of 5 forms a N‐lithium‐N′,N′‐bis(silyl)hydrazide ( 11 ). The dilithium salts of 5 ( 13 ) and of the bis(tert.‐butyldiphenylsilyl)hydrazine ( 12 ) crystallize as dimers with formation of a central Li4N4 unit. The formation of 12 from 11 occurs via a N′ → N‐silyl group migration. Results of crystal structure analyses are reported.  相似文献   

16.
Lithiation of 1-(dimethylsulfamoyl)imidazole by n-butyllithium, followed by substitution with dimethylformamide provided 1-(dimethylsulfamoyl)-2-imidazolecarboxaldehyde in 19% yield. When 1-(dimethylsulfamoyl)-2-(tert-butyldimethylsilyl)imidazole was lithiated by sec-butyllithium, followed by methyl formate, there was obtained 1-(dimethylsulfamoyl)-2-(tert-butyldimethylsilyl)-5-imidazolecarbox-aldehyde (57%). Removal of the silyl group by acetic acid yielded 1-(dimethylsulfamoyl)-5-imidazolecarbxaldehyde ( 11 , 96%) as a gum. Isomerization of 11 took place slowly at room temperature (10 days), or faster in tetrahydrofuran solution containing triethylamine (2 hours) to form crystalline 1-(dimethylsul-famoyl)-4-imidazolecarboxaldehyde (12) in 68% yield. Proton and carbon-13 nmr spectra were analyzed to determine the structure of the isomers. However, only X-ray crystallography established the structure of 1-(dimethylsulfamoyl)-4-imidazolecarboxaldehyde, unequivocally. A mechanism for the isomerization of 11 to 12 is proposed.  相似文献   

17.
The preferred conformation of aminophosphanes with bulky amino groups ( 1–20 ) was determined by NMR spectroscopy in solution, in two cases in the solid state ( 11,17 ) and in one case ( 11 ) by X‐ray crystallography. Trimethylsilylaminodiphenylphosphanes Ph2PN(R)SiMe3 (R = Bu ( 1 ), Ph ( 2 ), 2‐pyridyl ( 3 ), 2‐pyrimidyl ( 4 ), Me3Si ( 5 )), amino(chloro)phenylphosphanes Ph(Cl)PNRR′ (R = Bz, R′ = Me ( 6 ), R = Bz, R′ = tBu ( 7 ), R = Et, R′ = Ph ( 8 )), amino(chloro)tert‐butylphosphanes tBu(Cl)PNRR′ (R = R′ = iPr ( 9 ), R = Me, R′ = tBu ( 10 ), R = Bz, R′ = tBu ( 11 ), R = H, R′ = tBu ( 12 ), R = Et, R′ = Ph ( 13 ), R = iPr, R′ = Ph ( 14 ), R = Bu, R′ = Ph ( 15 ), R = Bz, R′ = Ph ( 16 ), R = R′ = Ph ( 17 ), R = R′ = Me3Si ( 18 )), 3‐tert‐butyl‐2‐chloro‐1,3,2‐oxazaphospholane ( 19 ), and benzyl(tert‐butyl)aminodichlorophosphane ( 20 ) were studied by 1H, 13C, 15N, 29Si, and 31P NMR spectroscopy. In all cases, the more bulky substituent at the nitrogen atom prefers the syn‐position with respect to the assumed orientation of the phosphorus lone pair of electrons. Many of the derivatives studied adopt this preferred conformation even at room temperature. Numerous signs of coupling constants 1J(31P, 15N), 2J(31P, 13C), and 2J(31P, 29Si) were determined. Low temperature NMR spectra were measured for derivatives for which rotation about the P N bond at room temperature is fast, showing the presence of two rotamers at low temperature. The respective conformation of these rotamers could be assigned by 13C, 15N, and 31P NMR spectroscopy. Isotope‐induced chemical shifts 1Δ15/14N(31P) were determined for all compounds at natural abundance of 15N by using Hahn‐echo extended polarization transfer experiments. The molecular structure of 11 in the solid state reveals pyramidal surroundings of the nitrogen atom and mutual trans‐positions of the tert‐butyl groups at phosphorus and nitrogen. © 2002 Wiley Periodicals, Inc. Heteroatom Chem 13:667–676, 2002; Published online in Wiley InterScience (www.interscience.wiley.com). DOI 10.1002/hc.10084  相似文献   

18.
15N isotopic enrichment was necessary for the unequivocal assignment of the 1H NMR lines to the protons in the NH–OH fragment of benzohydroxamic acid, BHXA, C6H5CONHOH, in dry dimethyl sulfoxide solutions. The assignment [δ(NH) = 11.21, δ(OH) = 9.01, 1J(15N,1H) = 102.2 Hz, 2J(15N,1H) <1.5 Hz], which is opposite to that used by other authors, confirms the assignment extended to BHXA by Brown and co‐workers from the spectra of acetohydroxamic acid. The enrichment allowed also assignment of the 29Si lines in the spectra of disilylated benzohydroxamic acid, (Z)‐tert‐butyldimethylsilyl Ntert‐butyldimethylsilyloxybenzoimidate (2) and (Z)‐tert‐butyldiphenylsilyl Ntert‐butyldiphenylsilyloxybenzoimidate (3), and confirmed structure of the monosilylated products, Ntert‐butyldiphenylsilyloxybenzamide (4) and Ntert‐butyldiphenylsilyloxy benzoimidic acid (5). Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

19.
Nguyen  Vien  Bonds  Darius V.  Prokai  Laszlo 《Chromatographia》2008,68(1):57-62
A GC-MS method was developed for measuring hydroxyl-radical capture products of salicylic acid, a common trapping agent for this reactive oxygen species, in samples obtained by in vivo cerebral microdialysis experiments. The assay employed liquid–liquid extraction followed by derivatization of 2,3- and 2,5-dihydroxybenzoic acid, along with 3,5-dihydroxybenzoic acid added as an internal standard. Due to their simple electron ionization mass spectra featuring [M–57]+ ions through the loss of tertiary alkyl group from the corresponding molecular ions, tert-butyldimethylsilyl (TBDMS) derivatives afforded straightforward method development based on selected-ion monitoring. In addition, tandem mass spectrometry probing collision-induced dissociation of [M–57]+ ions obtained from the isomeric tert-butyldimethylsilyl derivatives revealed characteristic differences in the resultant product-ion spectra. Our work has demonstrated the applicability of GC-MS for the assay of microdialysates for 2,3- and 2,5-dihydroxybenzoic acid by confirming that local administration of the excitotoxic glutamate into the rat striatum significantly increased in vivo hydroxyl-radical production in this brain region and that subsequent systemic administration of α-phenyl-tert-butylnitrone reversed glutamate-induced oxidative stress.  相似文献   

20.
Procedures are described for the preparation of various bidentate and potentially tridentate chelating agents. These incorporate pyridyl, benzimidazole, imidazole or phenolic moieties. Phillips condensations of carboxylic acids with o-phenylenediamines were carried out in 4 M hydrochloric acid. Syntheses are reported for 2, 6-bis(N′-methylimidazol-2′-ylthiomethyl)pyridine, 2, 6-bis(benzimidazol-2′-ylthiomethyl)pyridine, 2-(4′-piperidyl)benzimidazole, 2-(3′-piperidyl)benzimidazole, 2-(3-N′-methylpiperidyl)benziinidazole, 2-(3-N′-methylpiperidyl)-N-methylbenzimidazole, 2-(2′-hydroxybenzyl)benzimidazole and 2-(2′-hydroxyben-zyl)N-methylbenzimidazole. The compounds were characterized where appropriate by their mass, uv, and 1H-nmr spectra. 2-(2′-Hydroxybenzyl)benzimidazole hydrochloride acts as a gelling agent in aqueous solution.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号