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 15 毫秒
1.
Cui JG  Lin CW  Zeng LM  Su JY 《Steroids》2002,67(13-14):1015-1019
Using stigmasterol as the starting material, 24-methylenecholest-4-en-3beta,6 alpha-diol (2) was synthesized in eight steps in 13% overall yield. The introduction of the sterol side-chain was carried out using (3-methyl-2-oxobutyl)-triphenylarsonium bromide (11) and K(2)CO(3) in a solid-liquid phase-transfer Wittig reaction. Construction of the steroidal nucleus was finished by oxidation of 24-methylenecholest-5-en-3beta-ol (9) with pyridinium chlorochromate (PCC) in dichloromethane at ambient temperature and by reduction of 24-methylenecholest-4-en-3,6-dione (10) with NaBH(4) in the presence of CeCl(3).7H(2)O.  相似文献   

2.
Demethoxylation reactions in the cultures of the brown-rot fungi Gloeophyllum trabeum and Poria placenta were studied by determining the evolution of (14)CO(2) from a non-phenolic lignin model, beta-O-4 dimer, [O(14)CH(3)]-labelled at position 4 in the A ring (model I), and from [O(14)CH(3)]-labelled vanillic acid (model II). The fungi were grown under oxygen or air atmosphere on an agar medium with or without spruce sapwood blocks. The dimeric model (I) was impregnated onto agar or wood block in cultures to clarify the possible effect of wood as growth substrate. In the case of vanillic acid (model II), birch wood was used. The effect of supplemented nutrient nitrogen (2 mM N) and glucose (0.1 or 1.0% w/v) on demethoxylation was also studied. G. trabeum enhanced the production of (14)CO(2) from the dimer in the presence of spruce wood blocks. It released (14)CO(2) from the methoxyl groups giving 30-60% of the applied activity in 8 weeks. P. placenta produced almost 30% (14)CO(2 )from vanillic acid (model II) in 9 weeks under oxygen, but from the methoxyl group of the dimer only 3% of (14)CO(2) was evolved in 4 weeks. The biomasses determined as ergosterol assay showed variation from 14 to 226 microg g(-1) dry weight of agar, and 2 to 9 microg g(-1 )of wood, but they did not correlate with the production of (14)CO(2). The results indicate that these brown-rot fungi possess different mechanisms for demethoxylation.  相似文献   

3.
The organometallic precursor (NEt(4))(2)[ReBr(3)(CO)(3)] was reacted with bidendate dithioethers (L) of the general formula H(3)C-S-CH(2)CH(2)-S-R (R = -CH(2)CH(2)COOH, CH(2)-C&tbd1;CH) and R'-S-CH(2)CH(2)-S-R' (R' = CH(3)CH(2)-, CH(3)CH(2)-OH, and CH(2)COOH) in methanol to form stable rhenium(I) tricarbonyl complexes of the general composition [ReBr(CO)(3)L]. Under these conditions, the functional groups do not participate in the coordination. As a prototypic representative of this type of Re compounds, the propargylic group bearing complex [ReBr(CO(3))(H(3)C-S-CH(2)CH(2)-S-CH(2)C&tbd1;CH)] Re2 was studied by X-ray diffraction analysis. Its molecular structure exhibits a slightly distorted octahedron with facial coordination of the carbonyl ligands. The potentially tetradentate ligand HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH was reacted with the trinitrato precursor [Re(NO(3))(3)(CO)(3)](2-) to yield a cationic complex [Re(CO)(3)(HO-CH(2)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(2)-OH)]NO(3) Re8 which shows the coordination of one hydroxy group. Re8 has been characterized by correct elemental analysis, infrared spectroscopy, capillary electrophoresis, and X-ray diffraction analysis. Ligand exchange reaction of the carboxylic group bearing ligands H(3)C-S-CH(2)CH(2)-S-CH(2)CH(2)-COOH and HOOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH with (NEt(4))(2)[ReBr(3)(CO)(3)] in water and with equimolar amounts of NaOH led to complexes in which the bromide is replaced by the carboxylic group. The X-ray structure analysis of the complex [Re(CO)(3)(OOC-CH(2)-S-CH(2)CH(2)-S-CH(2)-COOH)] Re6 shows the second carboxylic group noncoordinated offering an ideal site for functionalization or coupling a biomolecule. The no-carrier-added preparation of the analogous (99m)Tc(I) carbonyl thioether complexes could be performed using the precursor fac-[(99m)Tc(H(2)O)(3)(CO)(3)](+), with yields up to 90%. The behavior of the chlorine containing (99m)Tc complex [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 in aqueous solution at physiological pH value was investigated. In saline, the chromatographically separated compound was stable for at least 120 min. However, in chloride-free aqueous solution, a water-coordinated cationic species Tc1a of the proposed composition [(99m)Tc(H(2)O)(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))](+) occurred. The cationic charge of the conversion product was confirmed by capillary electrophoresis. By the introduction of a carboxylic group into the thioether ligand as a third donor group, the conversion could be suppressed and thus the neutrality of the complex preserved. Biodistribution studies in the rat demonstrated for the neutral complexes [(99m)TcCl(CO)(3)(CH(3)CH(2)-S-CH(2)CH(2)-S-CH(2)CH(3))] Tc1 and [(99m)TcCl(CO)(3)(CH(2)-S-CH(2)CH(2)-S-CH(2)-C&tbd1;CH)] Tc2 a significant initial brain uptake (1.03 +/- 0.25% and 0.78 +/- 0.08% ID/organ at 5 min. p.i.). Challenge experiments with glutathione clearly indicated that no transchelation reaction occurs in vivo.  相似文献   

4.
Poly(lactic acid) (PLA) and poly(acrylic acid) (PAA) biomaterials with luminescent ruthenium tris(bipyridine) centers couple drug delivery and imaging functions. Hydrophobic [Ru(bpyPLA2)3](PF6)2 (1) was generated from [Ru[bpy(CH2OH)2]3](PF6)2 in bulk monomer using 4-(dimethylamino)pyridine as the catalyst. The bromoesters, [Ru[bpy(CH2OR)2]3](PF6)2, [Ru[bpy(C13H27)2][bpy(CH2OR]2](PF6)2 (4), and [Ru[bpy(PLAOR)2]3]2+ (9) (R=COCBr(CH3)2), served as initiators for tert-butyl acrylate (tBA) polymerization. Conversion of PtBA to PAA via hydrolysis affords water soluble materials, [Ru(bpyPAA2)3]2+ (7) and [Ru[bpy(C13H27)2](bpyPAA2)2]2+ (8) and the amphiphilic star polymer [Ru[bpy(PLA-PAA)2]3)](PF6)2 (11), which is soluble in a H2O/CH3CN (1:1) mixture. Luminescence excitation and emission spectra of the Ru polymers were in agreement with the parent [Ru(bpy)3]2+ chromophore (lambdaex=468, lambdaem=621 nm). Lifetimes of tau approximately 700 ns in both air and nitrogen atmospheres are typical for most materials; however, the amphiphilic star block copolymer 11 is quenched by oxygen to some degree. Thermal analysis shows the expected glass transitions for the polymeric ruthenium complex materials.  相似文献   

5.
Two solution syntheses of cyclo(11----5 epsilon)-[Lys5]substance P-(5-11) (CLP) were carried out. The first synthesis involved the stepwise elongation of the peptide chain starting from glycine tert-butyl ester. At the stage of hexapeptide deprotection, the cleavage of Boc and But groups was accompanied by tert-butylation of the Met residue. Cyclization was carried out via a pentafluorophenyl ester intermediate. The benzyloxycarbonyl-cyclopeptide (Z-CLP) formed was deprotected by catalytic transfer hydrogenation. A (3 + 4) block coupling strategy was used in course of the repeated preparation of the linear precursor of CLP. Optimization of the cyclization and subsequent deprotecting stages lead to increased yields and facilitated the synthetic procedure. Z-CLP was found to possess myotropic activity on isolated guinea pig ileum (alpha = 0.55 +/- 0.18; pD2 = 7.97 +/- 0.20), whereas CLP was inactive in these experiments. Z-CLP causes a slight two-phase effect on arterial pressure in rats, CLP being inactive. Similar to substance P, CLP displays an antidepressant-like effect in mice as indicated by the swimming test.  相似文献   

6.
The structures of substituted (aminomethyl)lithium and (thiomethyl)lithium compounds have been examined. Geometric parameters, charge densities, bond orders, dipole moments and heats of formation for all the members of the two series of monomers and dimers of the units LiCN(R)2 and LiCSR where R=H, CH3(Me), C6H5(Ph) have been calculated. The structures of the three complex compounds containing the same units; [[Li(CH2SMe)(THF)]X], [Li2(CH2SPh)2(THF)4] and [Li2(CH2NPh2)2(THF)3] have also been modeled. Geometry optimizations have been performed with the semiempirical PM3 method. The molecular orbital calculations have been carried out by a self-consistent field method using the restricted Hartree-Fock formalism. Comparisons have been made with the corresponding properties of methyl lithium monomer and dimer. The results show that in all of the nitrogen-containing monomers, the C-Li bonds weaken and the Li-C-H(N) angles decrease due to the coordination of lithium with nitrogen. Substitution of hydrogen atoms by methyl or phenyl groups decreases the Li-N coordination. In the sulfur-containing compounds, sulfur behaves similarly to nitrogen but the changes are smaller because the 3p lone-pair orbital of sulfur is higher in energy than the 2p lone-pair of nitrogen. All the dimers of nitrogen/sulfur-containing methyl lithium derivatives form six-membered rings in which the Li-N(S) coordination is greater than the one in the corresponding monomers. Dimerization reactions have been found to be exothermic and the formation of all the dimers is favored. The results obtained for the three complex structures are comparable to the experimental results reported in the literature.Keywords:  相似文献   

7.
For the purpose of developing highly sensitive and convenient determination of plasmalogens, the high-performance liquid chromatography (HPLC) method using radioactive iodine ((125)I) was investigated. Radioactive triiodide (1-) ion ((125)I(3)(-)), which is an actual iodine form capable of reacting with vinyl ether bond ([bond]CH(2)[bond]O[bond]CH[double bond]CH[bond]) of plasmalogens, could be safely and efficiently produced by oxidizing a commercial radioactive sodium iodine (Na(125)I) with hydrogen peroxide (H(2)O(2)) under acid condition (pH 5.5-6.0), which is called iodine-125 reagent. I(3)(-) specifically reacted with plasmalogens at the molar ratio of 1:1 in methanol, and 1 or 2 mol of plasmalogens was involved in the binding with iodine per iodine atom, resulting in the formation of stable iodine-binding phospholipids. The HPLC system with Diol column and acetonitrile/water as a mobile phase was available for separating iodine-binding phospholipids from nonbinding free iodine and for separately eluting iodine-binding phospholipids derived from choline and ethanolamine plasmalogens. Using iodine-125 reagent (1.85 MBq/ml), plasmalogens were detectable at high sensitivity of 10,000-15,000 cpm/nmol, which is more than 1000-fold higher sensitivity than the classical determination with nonradioactive iodine. Plasmalogen concentrations in human plasma were measured with the HPLC system and determined as, on average, 129.1+/-31.3 microM (n=8) in a 1.2 content ratio of choline to ethanolamine plasmalogens, a concentration that nearly agrees with the value reported previously.  相似文献   

8.
A series of bis(11-oxo-11H-indeno[1,2-b]quinoline-6-carboxamides) linked through the 6-carboxamides were prepared by coupling the requisite acid imidazolides with various diamines. Compounds with mono-cationic linker chains were more potent cytotoxins than the corresponding monomer in a panel of rodent and human cell lines, while those with the dicationic linker chains (CH2)2NR(CH2)2NR(CH2)2 and (CH2)2NR(CH2)3NR(CH2)2 showed extraordinarily high potencies (for example, IC50s of 0.18-1.4 nM against human Jurkat leukemia; up to 1000-fold more potent than the parent monomer). As seen previously in the monomeric series, small, lipophilic 4-substituents significantly increased potency in cell culture. The dimeric compounds were all slightly to significantly more potent in the mutant JL(A) and JL(D) cell lines that under-express topo II, suggesting that this enzyme is not their primary target. An 11-imino-linked dimer was much less active, and an asymmetric indeno[1,2-b]quinoline-6-carboxamide/naphthalimide dimer was less active than the comparable symmetric bis(indeno[1,2-b]quinoline-6-carboxamide). Selected analogues were active against sub-cutaneously implanted colon 38 tumors in mice, giving growth delays comparable to that of the clinical topo I inhibitor irinotecan at up to 10-fold lower doses. These compounds form an interesting new class of putative topo I inhibitors.  相似文献   

9.
A reinvestigation of the reactions leading to arsonolipids (2,3-diacyloxypropylarsonic acids) has been carried out in order to understand why the yields of their preparation were only moderate, although they are better than those reported for 2,3-diacyloxypropylphosphonic acid (phosphotidic acid). Thus, the reaction of glycidol and of 3-chloro-1,2-propanediol with alkaline sodium arsenite, "Na3AsO3", gives the desired product, 2,3-dihydroxypropylarsonic acid, and approximately 10% of an arsenic-containing glycerol dimer which is removed during the preparation of these arsonolipids. The step which is mainly responsible for the diminished yields is due to the reaction of the -As(SPh)2 or -AsO3H- precursor with the activated acid chlorides or carboxylic acid anhydrides to give an intermediate which cyclizes with the primary hydroxy group of the 2,3-dihydroxypropyl moiety. This cyclization does not allow the primary hydroxy group to be acylated. Such cyclization could not be avoided with RCOCl/py, (RCO)2O/DMAP, or RCOOH/DCC/DMAP acylating systems.  相似文献   

10.
The crucial step in the preparation of the title arsonolipids starting from the dichloromethane-soluble dithioarsonite CH2(OH)CH(OH)CH2–As(SPh)2 is to avoid an internal cyclization during the acylation which protects the primary –OH group from being acylated. This was to a large extent accomplished by using fatty acyl chloride in the presence of the weak base pyridine and controlling the temperature and rate of the acyl chloride addition, giving ~70% yields of arsonolipids. The presence of catalytic amounts of 4-dimethylaminopyridine boosted the yields to 82–85%. This yield is a great improvement over the yields (20–55%) previously achieved. The acylating systems (RCO)2O or RCOCl and BF3·Et2O gave only moderate yields (25–60%) of arsonolipids.  相似文献   

11.
Geometry optimizations and energy calculations have been carried out via molecular orbital methods at the density functional B3LYP/LANL2DZ level on the molecules PO3-, OPO3(3-), HOPO3(2-), CH3OPO3(2-), H(CH3OPO3)-, O(PO3)2(4-), HO(PO3)2(3-), CH2(PO3)2(4-), (CH3OPO2)O(PO3)3-, O(PO3)3(5-), HO(PO3)3(4-), (PO3)3(3-), (CH3OPO2)O(PO3)2(4-), [Mg[O(PO3)2)]]2-, [Ca[O(PO3)2]]2-, [Ca[CH2(PO3)2]]2-, [Ca[CH3OPO2)O(PO3)]]-, [Ca(PO3)3]-, [Ca[O(PO3)3]]3-, and [Ca[CH3OPO2)O(PO3)2]]2- with the aim to find reliable and easily accessible computational methods to simulate some phosphate-containing molecules of importance for the living cells and to study the energetics for protonation and metal-complex formation reactions. The analysis is part of a general investigation on phosphate-containing molecules as potential dissolving agents for calcium pyrophosphate dihydrate (CPPD) crystals which deposit in certain articular diseases. The basis set was expanded to 6-31G** for the P atoms for all the molecules investigated and to 6-31G* for the O atoms for OPO3(3-). Calculations at the semiempirical MNDO/d level were also carried out for comparison purposes on the free ligand molecules and on [Mg[O(PO3)2]]2-. The density functional analysis reproduced well the geometry found at the solid state via X-ray diffraction. The analyses of the geometrical parameters and the total electronic energy of the molecules shows that O(PO3)2(4-) and other di- and tri-phosphates are versatile ligands for divalent metal ions like Ca2+. The computed P-O-P bond angle for free O(PO3)2(4-) is 180 degrees and the conformation of the two PO3- groupings is staggered along the P...P vector. The linear arrangement for P-O-P is assisted by P-O pi interactions. The bending of the P-O-P angle when accompanied by a slight P-O(b) elongation requires a very small amount of energy; 4.65 kcal/mol to pass from 180 to 140 degrees , as calculated at the DFT level. The computed Ca-O and Mg-O bond distances for [M[O(PO3)2]]2- are 2.378 and 2.079A, when the metal ions link two oxygen atoms from each PO3 group. The computed Ca-O bond lengths for [Ca[CH3OPO2)O(PO3)]]- are 2.482 (PalphaO2) and 2.358A (PbetaO2), showing a significant lengthening for Ca-OPalpha, when compared to the pyrophosphate derivative. The Ca-O bond lengths for [Ca[O(PO3)3]]3- and [Ca[CH3OPO2)O(PO3)2]]2- are 2.251A and 2.525 (PalphaO2), 2.407 and 2.338 (PbetaO2), and 2.251 and 2.228A (PgammaO2), showing a shortening for the Ca-OPgamma bond upon methylation. The (Pbeta)O-Pgamma bond length increases significantly (0.09 A) upon Ca(II) coordination to (CH3OPO2)O(PO3)2(4-) via all the three PO3 groups. This latter result suggests that metal complexes of linear organic-triphosphates have a larger tendency to release the PgammaO3 group when compared to the free ligand molecules. The electronic contribution to the energy of the complex formation reaction for [Ca[CH2(PO3)2]]2- is only slightly higher (some 1.8 kcal) than that for [Ca[O(PO3)2]]2-; but is much higher (some 63 kcal) than that relevant to the formation of [Ca[CH3OPO2)O(PO3)2]]2-. (ABSTRACT TRUNCATED)  相似文献   

12.
The preparation and spectroscopic characterization of duplex decamers containing site-specific cis-syn and trans-syn thymine dimers are described. Three duplex decamers, d(CGTATTATGC).d(GCATAATACG), d(CGTAT[c,s]TATGC).d(GCATAATACG), and d(CGTAT[t,s]TATGC).d(GCATAATACG), were prepared by solid-phase phosphoramidite synthesis utilizing cis-syn and trans-syn cyclobutane thymine dimer building blocks (Taylor et al., 1987; Taylor & Brockie, 1988). NMR spectra (500 MHz 2D 1H and 202 MHz 1D 31P) were obtained in "100%" D2O at 10 degrees C, and 1D exchangeable 1H spectra were obtained in 10% D2O at 10 degrees C. 1H NMR assignments for H5, H6, H8, CH3, H1', H2', and H2" were made on the basis of standard sequential NOE assignment strategies and verified in part by DQF COSY data. Comparison of the chemical shift data suggests that the helix structure is perturbed more to the 3'-side of the cis-syn dimer and more to the 5'-side of the trans-syn dimer. Thermodynamic parameters for the helix in equilibrium coil equilibrium were obtained by two-state, all or none, analysis of the melting behavior of the duplexes. Analysis of the temperature dependence of the T5CH3 1H NMR signal gave delta H = 44 +/- 4 kcal and delta S = 132 +/- 13 eu for the trans-syn duplex. Analysis of the concentration and temperature dependence of UV spectra gave delta H = 64 +/- 6 kcal and delta S = 178 +/- 18 eu for the parent duplex and delta H = 66 +/- 7 kcal and delta S = 189 +/- 19 eu for cis-syn duplex. It was concluded that photodimerization of the dTpdT unit to give the cis-syn product causes little perturbation of the DNA whereas dimerization to give the trans-syn product causes much greater perturbation, possibly in the form of a kink or dislocation at the 5'-side of the dimer.  相似文献   

13.
To evaluate the NH(3), N(2)O, and CH(4) emissions from composting of livestock waste without forced aeration in turned piles, and to investigate the possible relationship between the scale of the compost pile and gas emission rates, we conducted swine manure composting experiments in parallel on small- and large-scale compost piles. Continuous measurements of gas emissions during composting were carried out using a chamber system, and detailed gas emission patterns were obtained. The total amount of each gas emission was computed from the amount of ventilation and gas concentration. NH(3) emission was observed in the early period of composting when the material was at a high temperature. Sharp peaks in CH(4) emission occurred immediately after swine manure was piled up, although a high emissions level continued after the first turning only in the large-scale pile. N(2)O emissions started around the middle stage of the composting period when NH(3) emissions and the temperature of the compost material began to decline. The emission rates of each gas in the small and large piles were 112.8 and 127.4 g NH(3)-N/kg T-N, 37.2 and 46.5 g N(2)O-N/kg T-N, and 1.0 and 1.9 g CH(4)/kg OM, respectively. It was found that changing the piling scale of the compost material was a major factor in gas emission rates.  相似文献   

14.
The ability of two platinum(IV) antitumor agents, cis,cis,trans-PtIV[(CH3)2CHNH2]2Cl2(OH)2 (2) and cis,cis,trans-PtIV(NH3)2Cl2(OH)2 (4), to interact with PM2 DNA was examined. Analysis using gel electrophoresis showed that neither compound is able to alter the electrophoretic mobilities of the three forms of PM2 DNA in the gel. However, incubation of 2 and 4 with 2 equiv of Fe(ClO4)2 X 6H2O or 1 equiv of ascorbic acid results in reduction to yield the divalent complexes cis-PtII(NH3)2Cl2 (1) and cis-PtII-[(CH3)2CHNH2]2Cl2 (3). The structures of the reduction products were characterized by using elemental analysis as well as infrared and 195Pt NMR spectroscopies. Both 1 and 3 were found to bind to and unwind supercoiled form I PM2 DNA. The aforementioned observations support the suggestion that reduction is a means of activating the antitumor properties of 2 and 4.  相似文献   

15.
Thyroid peroxidase (TPO), the major enzyme in the thyroid hormone synthesis, multifunctionally catalyzes (1) iodide oxidation, (2) iodination of the precursor protein, and (3) a coupling reaction of iodotyrosyl residues. The present study was carried out to examine the mercurial effects on the iodination, the second step of TPO. Purified porcine thyroglobulin or bovine serum albumin as acceptor protein was iodinated with [125I]NaI and H2O2 by purified porcine TPO. Iodinated protein was separated by acid precipitation on membrane filter or paper chromatography. Both CH3HgCl and HgCl2 dose-dependently inhibited the iodination, but HgCl2 was more potent to inhibit the iodination than CH3HgCl. These mercurial effects on the second step resemble the effects on the third step which were already reported; but are in marked contrast to the effects on the first step, where TPO was inhibited by HgCl2 but never by CH3HgCl.  相似文献   

16.
The complex trans-bis(dimethylsulfoxide)chloromethylplatinum(II) (1) is fairly soluble in water, where it undergoes multiple equilibria involving the formation of geometrically distinct [Pt(H(2)O)(DMSO)Cl(CH(3))] aqua-species. On reacting an aqueous solution of 1 with monodentate nitrogen donor ligands L, such as pyridines or amines, two well distinct patterns of behavior can be recognized: (i) a single stage fast substitution of one DMSO by the entering ligand, yielding a complex of the type trans(C,N)-[Pt(DMSO)(L)Cl(CH(3))] which contains four different groups coordinated to the metal and which undergoes a slow conversion into its cis-isomer, (ii) a double substitution affording cationic complex ions of the type cis-[Pt(L)(2)(DMSO)(CH(3))](+). When this latter reaction is carried out using sterically hindered ligands, slow rotation of the bulk ligand around the Pt[bond]N bond allows for the identification of head-to-head and head-to-tail rotamers in solution, through (1)H NMR spectrometry. The addition of chloride anion to 1 leads to the anionic species cis-[Pt(DMSO)Cl(2)(CH(3))](-), where a molecule of DMSO still remains coordinated to the metal center, despite its quite fast rate of ligand exchange (k(exch) with free DMSO=12+/-1 s(-1)). The reaction of complex 1 with bidentate ligands, such as ethylenediamine (en) or simple amino acids, leads to the cationic species [Pt(en)(DMSO)(CH(3))](+) or to the neutral [Pt(DMSO)(N[bond]O)(CH(3))], (where N[bond]-O[double bond]GlyO(-), AlaO(-)).  相似文献   

17.
The aerobic oxidation of Co(II)bleomycin bound to calf thymus DNA has been investigated in relation to the mechanism of reaction in solution in the absence of DNA. Kinetics of dioxygenation of the Co(II) complex were followed by spectrophotometric and electron spin resonance spectroscopy as well as dioxygen analysis. The reaction is slower than when carried out in solution; its rate is inversely related to the ratio of DNA base pairs to Co(II)bleomycin. The subsequent oxidation reaction, observed spectrophotometrically and by dioxygen analysis, is second order in cobalt complex. The calculated second order rate constant is also inversely related to the base pair to metal complex ratio. Once this ratio exceeds three, the reaction rate slows significantly with each additional increment of DNA added to the starting reaction mixture. Taking advantage of the high stability of O(2)-Co(II)bleomycin bound to greater than a 3-fold excess of DNA base pairs, it could be demonstrated that the rate constant for oxidation of two O(2)-Co(II)bleomycin molecules is much slower than that for O(2)-Co(II)bleomycin plus Co(II)bleomycin. With the same technique it was observed that the metal centers of O(2)-Co(II)bleomycin and Fe(II)bleomycin also undergo oxidation. The binding to DNA of both solution products of the oxidation of Co(II)bleomycin by O2 was examined by 1H NMR spectroscopy. Peroxy-Co(III)bleomycin, Form I, binds with higher affinity than Co(III)bleomycin, Form II. At lower ionic strength, the size of the DNA binding site for each form is about 2 base pairs/molecule of drug.  相似文献   

18.
Purified 12-hydroperoxyeicosa - 5, 8, 10, 14 - tetrae noic acid (12-HPETE) containing deuterium atoms at 5, 6, 8, 9, 11, 12, 14, and 15 was prepared by incubating octadeuterated arachidonic acid with a platelet preparation in air or [18O]oxygen gas. A mixture of 12-HPETE containing 16O16O:18O18O (5:4) was subsequently prepared and incubated with hematin in phosphate buffer (pH 7.4); in another experiment the same mixture of 12-HPETE was incubated with a rat lung preparation (0-30% ammonium sulfate) which lacked epoxide hydratase activity. Gas chromatography-mass spectrometry negative ion chemical ionization analysis of the extracted products after conversion into pentafluorobenzyl-trimethylsilyl derivatives indicated that the products from both incubations, i.e. 12-hydroxyeicosa - 5,8,10,14 - tetraenoic acid (12 - HETE), 8 - hydroxy - 11,12-epoxyeicosa - 5,9,14 - trienoic acid (8H-11,12-EPETE), 10H - 11,12-epoxyeicosa - 5,8,14 - trienoic acid (10H-11,12-EPETE), and 8,11,12-trihydroxyeicosa-5,9,14-trienoic acid, all retained either two [16O]oxygen or two [18O]oxygen atoms of starting 12-HPETE and that no mixed species existed which contained one [16O]oxygen and one [18O]oxygen atom. These results demonstrate for the first time in the arachidonic acid series an intramolecular transfer of the terminal hydroxyl group of the hydroperoxide of 12-HPETE to the C-8 or C-10 alkyl positions to form the hydroxyepoxides 8H-11,12-EPETE and 10H-11,12-EPETE. This reaction carried out by both hematin in the absence of protein and the rat lung preparation is suggestive of a metal-hydroperoxide-olefin "cage" complex facilitating a concerted reaction in which the terminal hydroxyl group of the hydroperoxide is trapped by alkyl free radical centers.  相似文献   

19.
Various steps involved in the preparation of partially methylated alditol acetates (PMAAs) from glycoprotein-derived carbohydrates were improved to obtain the derivatives in a rapid manner with excellent yields. Carbohydrates were permethylated in dimethyl sulfoxide (DMSO), using a fine suspension of sodium hydroxide and methyl iodide (CH3I). The fine suspension of NaOH was prepared conveniently from commercially available 50% aqueous NaOH in DMSO by sonication and washing the precipitate with DMSO. Methylation of ovalbumin and fetuin glycopeptides using the fine suspension of NaOH and CH3I was complete within 5 min, and the methylation reaction did not generate any nonsugar artifacts. Methylated carbohydrates without any purification were hydrolyzed in a mixture of volatile organic acids, which permitted rapid removal of the acids from samples by evaporation. Acetylation of partially methylated alditols with acetic anhydride for 2-4 h at ambient temperature using 4-N,N'-dimethylaminopyridine as a catalyst and the reaction was free from generating nonsugar reaction artifacts. The reaction time course for methylation, hydrolysis, and acetylation was determined to obtain optimum reaction conditions for preparation of the PMAAs. The procedure facilitated rapid identification and quantitation of PMAAs due to diminished reaction artifacts and the quality of the chromatogram depended only on the purity of starting material and the reagents used for the methylation analysis. Utility of these simple methods for rapid methylation analysis was demonstrated in the characterization of oligosaccharides isolated in small amounts using a carbohydrate analyzer.  相似文献   

20.
The acceptor specificities of ST3Gal III, ST3Gal IV, ST6Gal I and ST6Gal II were investigated using a panel of beta-D-Galp-(1-->4)-beta-D-GlcpNAc-(1-->2)-alpha-D-Manp-(1-->O)(CH(2))(7)CH(3) analogues. Modifications introduced at either C2, C3, C4, C5, or C6 of terminal D-Gal, as well as N-propionylation instead of N-acetylation of subterminal D-GlcN were tested for their influence on the alpha-2,3- and alpha-2,6-sialyltransferase acceptor activities. Both ST3Gal enzymes displayed the same narrow acceptor specificity, and only accept reduction of the Gal C2 hydroxyl function. The ST6Gal enzymes, however, do not have the same acceptor specificity. ST6Gal II seems less tolerant towards modifications at Gal C3 and C4 than ST6Gal I, and prefers beta-D-GalpNAc-(1-->4)-beta-D-GlcpNAc (LacdiNAc) as an acceptor substrate, as shown by replacing the Gal C2 hydroxyl group with an N-acetyl function. Finally, a particularly striking feature of all tested sialyltransferases is the activating effect of replacing the N-acetyl function of subterminal GlcNAc by an N-propionyl function.  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