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.
2.
The two-domain (βα) mammalian metallothionein binds seven divalent metals, however, the binding mechanism is not well characterized and recent reports require the presence of the partially metallated protein. In this paper, step-wise metallation of the metal-free, two-domain βα-rhMT and the isolated β-rhMT using Cd(II) is shown to proceed in a noncooperative manner by analysis of electrospray ionization mass spectrometric data. Under limiting amounts of Cd(II), all intermediate metallation states up to the fully metallated Cd3-β-rhMT and Cd7-βα-rhMT were observed. Addition of excess Cd(II), resulted in formation of the supermetallated (metallation in excess of normal levels) Cd4-β- and Cd8-βα-metallothionein species. These data establish that noncooperative cadmium metallation is a property of each isolated domain and the complete two-domain protein. Our data now also establish that supermetallation is a property that may provide information about the mechanism of metal transfer to other proteins.  相似文献   

3.
4.
Two new steroid glycosides: distolasteroside D6, (24S)-24-O-(β-D-xylopyranosyl)-5α-cholestane-3β,6α,8,15β,16β,24-hexaol, and distolasteroside D7, (22E,24R)-24-O-(β-D-xylopyranosyl)-5α-cholest-22-ene-3β,6α,8,15β,24-pentaol were isolated along with the previously known distolasterosides D1, D2, and D3, echinasteroside C, and (25S)-5α-cholestane-3β4β,6α,7α,8,15α,16β,26-octaol from the Far Eastern starfish Distolasterias nipon. The structures of new compounds were elucidated by NMR spectroscopy and MALDI TOF mass spectrometry. Like neurotrophins, distolasterosides D1, D2, and D3 were shown to induce neuroblast differentiation in a mouse neuroblastoma C1300 cell culture.  相似文献   

5.
The expression level of electrophoretically separated S- and F-allozymes of β-specific esterase (EC 3.1.1.2) in genotypes of wild-type Drosophila melanogaster (males and females) that are monozygous or heterozygous with respect to the locus β-Est is determined by means of computerized densitometry; α-naphthylacetate, β-naphthylacetate, and α-naphthylpropionate are used as the substrates. The intensity of the expression of the esterase is judged from the quantity of reaction product created as a result of simultaneous azo coupling between naphthol and diazonium in 4, 24, 44, and 64 min incubation times. Reliable differences in the expressions of the S- and F-allozymes as a function of the structure of the β-Est locus of genotypically distinct individuals are established. In all the variant experiments, a higher level of summary activity of the S- and F-allozymes of the β-esterase of the heterozygotes by comparison with the individual activity of the F-and S-allozymes of the corresponding homozygotes was demonstrated, independently of the sex of the Drosophila individual. A comparative estimate of the temporal dynamics of the expression of in vitro allozymes of the dominant homozygotes (β-Est S /β-Est S ), heterozygotes (β-Est S /β-Est F ), and recessive homozygotes (β-Est F /β-Est F ) is performed. Possible mechanisms for the occurrence of heterosis according to the character of expression of S- and F-allozymes of β-esterase on the basis of the theory of biochemical enrichment of heterozygote genotypes are considered.  相似文献   

6.
The tritium recovery assay of 9-hydroxyprostaglandin dehydrogenase [Pace-Asciak, C. (1975) J. Biol. Chem.250, 2789] has been modified to ensure its applicability to both crude and purified enzyme preparations. The stereospecificity of NAD+-dependent 9-hydroxyprostaglandin dehydrogenase with respect to NAD+ was determined first and found to be A-side specific. Based on the stereospecificity of the enzyme, a simple and sensitive assay method for 9-hydroxyprostaglandin dehydrogenase has been developed. The assay is able to detect picomole quantities of substrate conversion. When 15-keto-13,14-dihydro-[9β-3H]PGF is employed as substrate, the tritium label of the tritiated prostaglandin is effected to transfer to lactate stereospecifically by coupling 9-hydroxyprostaglandin dehydrogenase with a saturating level of lactate dehydrogenase. The amount of prostaglandin oxidized is quantitated by the radioactivity of the labeled lactate produced, which is separated from labeled prostaglandin by charcoal precipitation. Simultaneous assays with the current tritium-release and thin-layer chromatography methods indicated excellent correlation. Using this method we have found that rat kidney possesses the highest enzyme activity among those tissues examined. Rat kidney enzyme activity is linear for the first 10 min it is studied and is nonlinear with increasing amounts of crude enzyme extract, indicating the possible presence of endogenous inhibitor(s). The apparent Km for 15-keto-13,14-dihydro-PGF is 0.66 μm. The enzyme is activated by imipramine, inhibited by indomethacin, but not affected by furosemide and ethacrynic acid. These results confirm previous findings reported in the literature.  相似文献   

7.
The reaction of 3β-hydroxy-21-hydroxymethylidenepregn-5-en-3β-ol-20-one (1) with phenylhydrazine (2a) affords two regioisomers, 17β-(1-phenyl-3-pyrazolyl)androst-3-en-3β-ol (5a) and 17β-(1-phenyl-5-pyrazolyl)androst-5-en-3β-ol (6a). The direction of the ring-closure reactions of 1 with p-substituted phenylhydrazines (2b-e) depends strongly on the electronic features of the substituents. Oppenauer oxidation of 3β-hydroxy-17β-exo-heterocyclic steroids 5a-e and 6a-e yielded the corresponding Δ4-3-ketosteroids 9a-e and 10a-e. The inhibitory effects (IC50) of these compounds on rat testicular C17,20-lyase were investigated by means of an in vitro radioligand incubation technique.  相似文献   

8.
Three new polar steroids identified as trofoside A, 20R,24S)-24-O-(3-O-methyl-β-D-xylopyranosyl)-3β,6α,8,15β,24-pentahydroxy-5α-cholestane, its 22(23)-dehydro derivative (trofoside B), and 15-sulfooxy-(20R,24S)-5α-cholestane-3β,6β,8,15α,24-pentaol sodium salt, were isolated fromTrofodiscus über starfish extracts collected in the Sea of Ohotsk. Two known compounds, trofoside A aglycone, (20R,24S)-3β,6α,8,15β,24-pentahydroxy-5α-cholestane, and triseramide, (20R,24R,25S,22E)-24-methyl-3β6α,8,15β-tetrahydroxy-5α-cholest-22-en-27-oic acid (2-sulfoethyl)amide sodium salt, were also found. The structures of the isolated polyoxysteroids were established from their spectra. Minimal concentrations causing degradation of unfertilized egg-cells of the sea-urchin Strongylocentrotus intermedius(C min) and terminating the cell division at the stage of the first division (C min embr.), as well as the concentrations causing 50% immobilization of sperm cells (OC50) and inhibiting their ability to fertilize egg-cells by 50% (IC50) were determined for the isolated compounds. Of three compounds highly toxic in embryos and sea-urchin sperm cells, the polyol with a sulfo group in the steroid core was the most active; two glycosides with monosaccharide chains located at C3 and C24 atoms were less toxic. Note that all the compounds with the spermiotoxic activities differently affected the embryo development. The positions of monosaccharide residues in the core considerably influence the compound activity. For example, both mono-and double chained glycosides with the monosaccharide fragment at C3 and fragments at C3 and C4 atoms are active against sea-urchin sperm cells and embryos, whereas the C24 glycosylated trofoside A does not affect embryos and displays a poor spermiotoxicity.  相似文献   

9.
10.
Ten flavone compounds, including three new flavonoid glycosides, were isolated from defatted rapeseed, and their protective antioxidant effect on H2O2-induced oxidative damage in human umbilical vein endothelial cells (ECV-304) was investigated. Three new flavonoid glycosides were identified as kaempferol-3-O-[(6-O-sinapoyl)-β-d-glucopyranosyl-(1  2)-β-d-glucopyranoside]-7-O-β-d-glucopyranoside (8), kaempferol-3,7-di-O-β-d-glucopyranoside-4'-O-(6-O-sinapoyl)-β-d-glucopyranoside (9), and kaempferol-3-O-[(3-O-sinapoyl)-β-d-glucopyranosyl-(1  2)-β-d-glucopyranoside]-7-O-β-d-glucopyranoside (10). The protective effects of all of the isolated compounds on H2O2-induced oxidative damage were assessed, and the activities of superoxide dismutase (SOD) and lactate dehydrogenase (LDH) were measured. All of compounds had a protective effect on H2O2-induced oxidative damage in ECV-304 cells and the presence of a substituted sinapoyl group and its position in the structures were used to elucidate the activity differences.  相似文献   

11.
Surfaces of cultured human lymphoid cells RPMI 1788, RPMI 4098, RPMI 8866, Raji, and WI-L2 were found to contain bothβ 2-microglobulin (β 2-μ) and HL-A determinants when tested by direct complement-dependent cytotoxicity andquantitative absorption with different cytotoxic antiβ 2-μ antisera and specific HL-A alloantisera. The same antigenic specificities were found in 3M KCl extracts of these cultured cells with a sensitiveβ 2-μ radioimmunoassay and an HL-A antigen blocking assay. Daudi cells provided a contrast, since noβ 2-μ or HL-A determinants were found on their surfaces or in 3 M KCl extracts prepared from them. Results from specific antibody blocking tests suggest a close association betweenβ 2-μ and HL-A determinants on plasma membranes of cultured human lymphoid cells. A solid state immunoadsorbent containing antiβ 2-μ antibodies effectively removed all detectable HL-A antigenic activity from some 3M KCl extracts of cultured human lymphoid cells as well as from some sera. Adsorption of HL-A antigens to these immunoadsorbents was specific since it was blocked only by prior addition ofβ 2-μ. Once on the antiβ 2-μ immunoadsorbents, HL-A antigens still reacted specifically with HL-A alloantibodies in quantitative absorption experiments. HL-A antigens andβ 2-μ could be eluted from antiβ 2-μ immunoadsorbents with a variety of chaotropic reagents and detergents, but thus far potassium bromide and sodium dodecyl sulfate (SDS) appear to be the most effective. SDS-PAGE of these eluates indicated that HL-A antigens were considerably purified by adsorption to antiβ 2-μ immunoadsorbents and that two major molecular size fragments were distinguishable, i.e., ∼33,000 for HL-A and ∼ 12,000 forβ 2-μ.  相似文献   

12.
The deduced amino acid sequence from a gene of the hyperthermophilic archaeon Pyrococcus sp. ST04 (Py04_0872) contained a conserved glycoside hydrolase family 57 (GH57) motif, but showed <13 % sequence identity with other known Pyrococcus GH57 enzymes, such as 4-α-glucanotransferase (EC 2.4.1.25), amylopullulanase (EC 3.2.1.41), and branching enzyme (EC 2.4.1.18). This gene was cloned and expressed in Escherichia coli, and the recombinant product (P yrococcus sp. ST04 maltose-forming α-amylase, PSMA) was a novel 70-kDa maltose-forming α-amylase. PSMA only recognized maltose (G2) units with α-1,4 and α-1,6 linkages in polysaccharides (e.g., starch, amylopectin, and glycogen) and hydrolyzed pullulan very poorly. G2 was the primary end product of hydrolysis. Branched cyclodextrin (CD) was only hydrolyzed along its branched maltooligosaccharides. 6-O-glucosyl-β-cyclodextrin (G1-β-CD) and β-cyclodextrin (β-CD) were resistant to PSMA suggesting that PSMA is an exo-type glucan hydrolase with α-1,4- and α-1,6-glucan hydrolytic activities. The half-saturation value (K m) for the α-1,4 linkage of maltotriose (G3) was 8.4 mM while that of the α-1,6 linkage of 6-O-maltosyl-β-cyclodextrin (G2-β-CD) was 0.3 mM. The k cat values were 381.0 min?1 for G3 and 1,545.0 min?1 for G2-β-CD. The enzyme was inhibited competitively by the reaction product G2, and the K i constant was 0.7 mM. PSMA bridges the gap between amylases that hydrolyze larger maltodextrins and α-glucosidase that feeds G2 into glycolysis by hydrolyzing smaller glucans into G2 units.  相似文献   

13.
α-D-Mannosyl-maltotriose (Man-G3) were synthesized from methyl α-mannoside and maltotriose by the transfer action of α-mannosidase. (Man-G3)-βCD and (Man-G3)2-βCD were produced in about 20% and 4% yield, respectively when Aerobacter aerogenes pullulanase (160 units per 1 g of Man-G3) was incubated with the mixture of 1.6 M Man-G3 and 0.16 M βCD at 50°C for 4 days. The reaction products, (Man-G3)-βCD were separated to three peaks by HPLC analysis on a YMC-PACK A-323-3 column and (Man-G3)2-βCD were separated to several peaks by HPLC analysis on a Daisopak ODS column. The major product of (Man-G3)-βCDs was identified as 6-O-α-(63-O-α-D-mannosyl-maltotriosyl)-βCD by FAB-MS and NMR spectroscopies. The structures of (Man-G3)2-βCDs were analyzed by TOF-MS and NMR spectroscopies, and confirmed by comparison of elution profiles of their hydrolyzates by α-mannosidase and glucoamylase on a graphitized carbon column with those of the authentic di-glucosyl-βCDs. The structures of three main components of (Man-G3)2-βCDs were identified as 61,62-, 61,63- and 61,64-di-O-(63-O-α-D-mannosyl-maltotriosyl)-βCD.  相似文献   

14.
Two hexacoordinated dinuclear compounds [Mn(L1)(dca)]2(ClO4/PF6)2·CH3OH (1/2) and two heptacoordinated coordination polymers [Mn(L2)(dca)]n(ClO4/PF6)n (3/4) [L1 = N,N′-(bis-(pyridin-2-yl)benzylidene)-1,3-propanediamine; L2 = N,N′-(bis-(pyridin-2-yl)benzylidene)diethylenetriamine; dca = dicyanamide] are synthesized and characterized. Structures of 1-3 have been solved by X-ray diffraction measurements. Each manganese(II) center in 1/2 is located in a distorted octahedral environment with an MnN6 chromophore coordinated by the four N atoms of L1 and two nitrile N atoms of bibridged μ1,5 dca. Interestingly, the coordination polymer 3 forms a 1D chain through single Mn-(NCNCN)-Mn units in which each manganese(II) center adopts a pentagonal bipyramidal geometry with an MnN7 chromophore occupied with five N atoms of L2 and two nitrile N atoms of monobridged μ1,5 dca. Magnetic susceptibility measurements of 1-3 in the 2-300 K temperature range reveal weak antiferromagnetic interactions.  相似文献   

15.
Bicyclic furano[2,3-d]pyrimidine ribonucleosides were synthesized by Pd(0)-and CuI-catalyzed coupling of 5-iodouridine with terminal alkynes. The treatment of the resulting nucleosides with ammonia or methylamine solution in aqueous alcohol resulted in pyrrolo-and N 7-methylpyrrolo[2,3-d]pyrimidine nucleosides. 5′-O-Triphosphates of bicyclic nucleosides were obtained by the treatment of the nucleosides with POCl3 in the presence of a “proton sponge.” The 5′-O-triphosphates are not substrates for HCV RNA-dependent RNA polymerase, but are effective substrates for HCV RNA helicase/NTPase and did not inhibit ATP hydrolysis. Only 3-(β-D-ribofuranosyl)-6-decyl-2,3-dihydrofuro-[2,3-d]pyrimidin-2-one showed a moderate anti-HCV activity in the HCV replicon system and efficiently inhibited replication of bovine viral diarrhea virus (BVDV) in KCT-cells, other compounds being inactive. None of the compounds were cytotoxic within the tested range of concentrations.  相似文献   

16.
The reaction of [Ag2(κ2-P,P′-DPEphos)2(μ-OTf)2] (1) (DPEphos = bis(2-(diphenylphosphino)phenyl]ether) with 1,10-phenanthroline (phen) and 4,4′-bipyridine in equimolar ratios afford, respectively, the mononuclear complex [Ag(κ2-P,P′-DPEphos)(phen)][OTf] (2) and the coordination polymer [Ag(κ2-P,P′-DPEphos)(μ-4,4′-bpy)]n[OTf]n (3). In complex 3, the silver atoms are bridged by 4,4′-bipyridine units to form a zigzag metallopolymer.  相似文献   

17.
The reaction of Pd(β-NaiR)Cl2 (2) [β-NaiR (1) = 1-alkyl-2-(naphthyl-β-azo)imidazoles] with ArNH2 in MeCN has yielded a C-N coupled product chloro[1-alkyl-2-{(7-imidoaryl)naphthyl-β-azo}imidazole-N,N′,N′′]palladium(II), Pd(β-NaiR-N-Ar)Cl (3-5) and coupling takes place at ortho-C-H position of pendant naphthyl group. The structural confirmation has been achieved by single crystal X-ray structure determination of the representative complexes, Pd(β-NaiEt)Cl2 (2b) and Pd(β-NaiEt-N-C6H4-Cl-p)Cl (5b). The electronic spectra of the products, 3-5, exhibit characteristic transition within 600-900 nm those are absent in Pd(β-NaiR)Cl2 (2). Cyclic voltammogram shows one oxidative response and two ligand reductions. The products are emissive. The excited state decays via radiative and non-radiative biexponential routes. The electronic structure, spectra and redox properties are explained by DFT computation.  相似文献   

18.
Klaus Denzel  Georg G. Gross 《Planta》1991,184(2):285-289
Cell-free extracts from leaves of Rhus typhina L. (sumach) were found to transfer the 1-O-galloyl moiety of l,6-di-O-galloyl-β-d-glucose to the 2-position of the same compound, yielding 1,2,6-tri-O-galloyl-β-d-glucose and leaving 6-O-galloylglucose as the deacylated by-product. The enzyme catalyzing this ‘disproportionation’ was purified almost 1700-fold. It had a molecular weight of approx. 56 000, a K m value of 11.5 mM, was stable between pH 4.5 and 6.5, and most active at pH 5.9 and 40° C. The systematic name “1,6-di-O-galloyl-glucose: 1,6-di-O-galloylglucose 2-O-galloyltransferase” (EC 2.3.1.) was proposed for this new enzyme whose detection provided evidence that, in addition to β-glucogallin (1-O-galloyl-β-d-glucose), higher substituted glucose esters also have the potential to serve as acyl donors in the biosynthesis of gallotannins.  相似文献   

19.
Li JB  Hashimoto F  Shimizu K  Sakata Y 《Phytochemistry》2008,69(18):3166-3171
Five anthocyanins, cyanidin 3-O-(2-O-β-xylopyranosyl-6-O-(Z)-p-coumaroyl)-β-galactopyranoside (2), cyanidin 3-O-(2-O-β-xylopyranosyl-6-O-(E)-p-coumaroyl)-β-galactopyranoside (3), cyanidin 3-O-(2-O-β-xylopyranosyl-6-O-(E)-caffeoyl)-β-galactopyranoside (4), cyanidin 3-O-(2-O-β-xylopyranosyl-6-O-acetyl)-β-galactopyranoside (5), and cyanidin 3-O-(2-O-β-xylopyranosyl-6-O-acetyl)-β-glucopyranoside (6), together with the known cyanidin 3-O-(2-O-β-xylopyranosyl)-β-galactopyranoside (1), were isolated from red flowers of Camellia cultivar ‘Dalicha’ (Camellia reticulata) by chromatography using open columns. Their structures were subsequently determined on the basis of spectroscopic analyses, i.e., 1H NMR, 13C NMR, HMQC, HMBC, HR ESI-MS and UV-vis.  相似文献   

20.
The reaction of Pt(COD)Cl2, where COD is 1,5-cyclooctadiene, with one equivalent of a diamidato-bis(phosphino) Trost ligand ((R,R)-2 = N,N′-bis(2-diphenylphosphino-1-benzoyl)-(1R,2R)-1,2-diaminocyclohexane, (R,R)-N,N′-bis(2-diphenylphosphino-1-naphthoyl)-(1R,2R)-1,2-diaminocyclohexane, or (±)-N,N′-bis(2-diphenylphosphino-1-benzoyl)-1,2-bis(aminobenzene)) in the presence of base afforded square planar diamidato-bis(phosphino) platinum(II) complexes (R,R)-2-Pt, (R,R)-3-Pt, (±)-4-Pt. Characterization of all complexes included the solution and solid state structure determination of each complex based on multinuclear NMR and X-ray analyses, respectively. Stability of the complexes in acid was examined on addition of HCl to (R,R)-2-Pt in chloroform and compared to the unreactive nature of the similar diamidato-bis(phosphino) complex 1-Pt (= 1,2-bis-N-[2′-(diphenylphosphino)benzoyl]diamino-benzene) in the presence of acid. Protonation of the bound amidato nitrogen atoms of (R,R)-2-Pt was observed along with decoordination of the nitrogen atoms from the platinum(II) center producing (R,R)-2-PtCl2 in quantitative yield by NMR analysis. Confirmation of the product was made on comparison of the NMR spectra to that of authentic (R,R)-2-PtCl2 prepared on reaction of Pt(COD)Cl2 with (R,R)-2 in CH2Cl2 and characterized by single-crystal X-ray diffraction analysis and NMR spectroscopy. Results add to the knowledge of rich coordination chemistry of bis(phosphino) ligands with late transition metals, metal-amidato chemistry, and has implications in catalysis.  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