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.
We have investigated the effects of a 3-thia fatty acid (TTA) and of temperature on the fatty acid (FA) metabolism of Atlantic salmon (Salmo salar). One experiment investigated the activity of the peroxisomal beta-oxidation enzyme, acyl-CoA oxidase (ACO), and the incorporation of TTA into phospholipid (PL) molecular species. Salmon hepatocytes in culture were incubated either without TTA (control(spades)) or with 0.8 mM TTA (TTA(spades)) in a short term (48 h) temperature study at 5 degrees C and at 12 degrees C. TTA was incorporated into the four PL classes studied: phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylinositol (PI) and phosphatidylserine (PS). TTA was preferentially esterified with 18:1, 16:1, 20:4 and 22:6 in the PLs. Hepatocytes incubated with TTA had higher ACO activity at 5 degrees C than at 12 degrees C. In a second experiment salmon were fed a diet based on fish meal-fish oil without any TTA added (control) or a fish meal-fish oil diet supplemented with 0.6% TTA for 8 weeks at 12 degrees C and 20 weeks at 5 degrees C. At the end of the feeding trial, hepatocytes from fish acclimated to high or low temperatures were isolated from both dietary groups and incubated with either [1-(14)C]18:1 n-9 or [1-(14)C]20:4 n-3 at 5 degrees C or 12 degrees C. Radiolabelled 18:1 n-9 was mainly esterified into neutral lipids (NL), whereas [1-(14)C]20:4 n-3 was mainly esterified into PL at both temperatures. The rate of elongation of [1-(14)C]18:1 n-9 to 20:1 n-9 was twice as high in hepatocytes from fish fed the control diet than it was in hepatocytes from fish fed the TTA diet, at both temperatures. The amount of [1-(14)C]20:4 n-3 converted to 22:6 n-3 was approximately the same in hepatocytes from the two dietary groups, but there was a tendency to higher production of 22:6 n-3 at the lower temperature. Oxidation of [1-(14)C]18:1 n-9 to acid soluble products (ASP) and CO(2) was approximately 10-fold greater in hepatocytes kept at 5 degrees C than in those kept at 12 degrees C and the main oxidation products formed were acetate, oxaloacetate and malate.  相似文献   

2.
【目的】通过构建的人工耐酸系统,筛选耐受低pH值、乳酸及琥珀酸的菌株。【方法】构建人工耐酸系统长期驯化菌株,利用不同p H的酸性平板进行筛选,从环境中筛选出一株对低p H值、高浓度乳酸以及琥珀酸有很好耐受性的菌株。通过形态学特征、生理生化特征研究,并结合18S rDNA基因序列分析及分子系统发育树的构建结果,确定菌株的种类。【结果】经过酸性人工系统的长期驯化,筛选分离出一株耐受低pH值、高浓度乳酸以及琥珀酸的菌株WJ-2,经鉴定该菌株为酿酒酵母(Saccharomyces cerevisiae),其最适生长温度为30°C。酸性平板实验显示该菌株能够耐受pH2.5的酸性环境,同时对9%的乳酸及8%的琥珀酸也有很好的耐受性。另外,耐酸菌株WJ-2在pH 2.5、9%乳酸和8%琥珀酸的培养环境中仍能保持相对中性的细胞内pH值。【结论】通过构建人工酸性系统,成功筛选出一株对低pH值、高浓度乳酸以及琥珀酸具有耐受性的菌株——酿酒酵母菌WJ-2,该方法可为筛选具有特定耐受能力菌株提供一个新思路。  相似文献   

3.
The [3H]triamcinolone acetonide ([3H]TA)-binding ability of the rat liver glucocorticoid receptor (GR) was investigated under acidic conditions, ranging from pH 2 to 7.3. Both in the presence and absence of 10 mM molybdate, the [3H]TA-binding ability decreased below pH 6.5 and was almost completely lost below pH 5, pH 5.9 +/- 0.1 giving 50% [3H]TA-binding. The binding ability was recovered when the pH of the cytosol was reversed to 7.3 or the precipitate obtained on acidification was dissolved in a buffer of pH 7.3. Moreover, in the absence of molybdate, the [3H]TA-GR complexes formed at pH 7.3 remained unchanged until pH 5. Then they decreased, pH 3.9 +/- 0.1 giving 50% binding, and completely disappeared at pH 3. [3H]TA-binding activity recovered from the precipitate also decreased in a similar pH region (a 50% decrease in binding being observed at pH 4.2 +/- 0.04). These results suggest that rat liver GR is rather resistant under acidic conditions and that it exists in a peculiar state below pH 5.9 to approximately 4 as to its ligand binding property: unoccupied GR has no [3H]TA-binding ability but [3H]TA-GR complexes once formed at neutral pH do not dissociate. [3H]TA-GR complexes recovered from the precipitate at pH 5 had a Stokes radius of 7.5 nm, little DNA-cellulose-binding ability and sedimented at 8.6S on glycerol gradient centrifugation, indicating that the receptor existed in a nontransformed state. In addition, both occupied and unoccupied GR were transformed at about pH 4, their being 50% transformation. This transformation was accompanied by irreversible denaturation of the receptor.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
1. Propylene oxide reacts with DNA in aqueous buffer solution at about neutral pH to yield two principal products, identified as 7-(2-hydroxypropyl)guanine and 3-(2-hydroxypropyl)adenine, which hydrolyse out of the alkylated DNA at neutral pH values at 37 degrees C. 2. These products were obtained in quantity by reactions between propylene oxide and guanosine or adenine respectively. 3. The reactions between propylene oxide and adenine in acetic acid were parallel to those between dimethyl sulphate and adenine in neutral aqueous solution; the alkylated positions in adenine in order of decreasing reactivity were N-3, N-1 and N-9. A method for separating these alkyladenines is described. 4. Deoxyguanylic acid sodium salt was alkylated at N-7 by propylene oxide in neutral aqueous solution. 5. The nature of the side chain in the principal alkylation products was established by mass spectrometry, and the nature of the products is consistent with their formation by the bimolecular reaction mechanism.  相似文献   

5.
To investigate properties of hydrophilic bundled peptides and their interactions with phospholipid membranes, bundled peptides named [Trp2]- and [Trp12]-4alpha-46S9, which are composed of four fragments of amphiphilic 24-mer peptide, were designed and synthesized. Tryptophan (Trp) was introduced at the 2nd position from the N-terminal or at the centre (12th) of the helix to monitor the peptide-lipid interaction. Circular dichroism measurements indicated that the peptides had low alpha-helicities in a buffer solution (pH 7.4) and also in the presence of dipalmitoyl-DL-3-phosphatidylcholine (DPPC) vesicles. In the presence of DPPC/dipalmitoyl-DL-3-phosphatidylglycerol (DPPG) (3:1) vesicles, the measurement could not be taken because of turbidity induced by vesicle aggregation. Both peptides had moderate perturbation activity for both the neutral and acidic vesicles at 25 degrees C. The perturbation patterns at 50 degrees C were much different from those at 25 degrees C and the maximum activity reached 100% at a low peptide concentration. The results of the measurement of membrane fusion activity of peptides showed a similar tendency to that found in the perturbation experiment. A quenching experiment indicated that the Trp2 and Trp12 residues in [Trp2]- and [Trp12]-4alpha-46S9 were scarcely embedded in neutral lipid membranes.  相似文献   

6.
Abstract Reactions between the anticancer drug titanocene dichloride (Cp2TiCl2) and various nucleotides and their constituents in aqueous solution or N,N-dimethylformamide (DMF) have been investigated by 1H and 31P NMR spectroscopy and in the solid state by IR spectroscopy. In aqueous solution over the pH* (pH meter reading in D2O) range 2.3-6.5, CMP forms one new species with Ti(IV) bound only to the phosphate group. In acidic media at pH*<4.6, three species containing titanocene bound to the phosphate group of dGMP, AMP, dTMP and UMP are formed rapidly. The bases also appear to influence titanocene binding. Only one of these Ti(IV)-bound species can be detected in the pH* range of 4.6-6.5 in each case. The order of reactivity towards Cp2TiCl2(aq) at pH* ca. 3 is GMP>TMP approximately AMP > CMP. At pH* > 7.0, hydrolysis of Cp2TiCl2 predominated and little reaction with the nucleotides was observed. Binding of deoxyribose 5'-phosphate and 4-nitrophenyl phosphate to Cp2TiCl2(aq) via their phosphate groups was detected by 31P NMR spectroscopy, but no reaction between Cp2TiCl2(aq) and deoxyguanosine, 9-ethylguanine or deoxy-D-ribose was observed in aqueous solution. The nucleoside phosphodiesters 3',5'-cyclic GMP and 2',3'-cyclic CMP did not react with Cp2TiCl2(aq) in aqueous solution; however, in the less polar solvent DMF, 3',5'-cyclic GMP coordination to [Cp2Ti]2+ via its phosphodiester group was readily observed. Binding of titanocene to the phosphodiester group of the dinucleotide GpC was also observed in DMF by 31P NMR. The nucleoside triphosphates ATP and GTP reacted more extensively with Cp2TiCl2(aq) than their monophosphates; complexes with bound phosphate groups were formed in acidic media and to a lesser extent at neutral pH. Cleavage of phosphate bonds in ATP (and GTP) by Cp2TiCl2(aq) to form inorganic phosphate, AMP (or GMP) and ADP (or GDP) was observed in aqueous solutions. In addition, titanocene binding to ATP was not inhibited by Mg(II), but the ternary complex titanocene-ATP-Mg appeared to form. These reactions contrast markedly with those of the drug cisplatin, which binds predominantly to the base nitrogen atoms of nucleotides and only weakly to the phosphate groups. The high affinity of Ti(IV) for phosphate groups may be important for its biological activity.  相似文献   

7.
Irradiation of trans-[RuCl(cyclam)(NO)](2+), cyclam is 1,4,8,11-tetraazacyclotetradecane, at pHs 1-7.4, with near UV light results in the release of NO and formation of trans-[Ru(III)Cl(OH)(cyclam)](+) with pH dependent quantum yields (from approximately 0.01 to 0.16 mol Einstein(-1)) lower than that for trans-[RuCl([15]aneN(4))(NO)](2+), [15]aneN(4) is 1,4,8,12-tetaazacyclopentadecane, (0.61 mol Einstein(-1)). After irradiation with 355 nm light, the trans-[RuCl([15]aneN(4))(NO)](2+) induces relaxation of the aortic ring, whereas the trans-[RuCl(cyclam)(NO)](2+) complex does not. The relaxation observed with trans-[RuCl([15]aneN(4))(NO)](2+) is consistent with a larger quantum yield of release of NO from this complex.  相似文献   

8.
A series of bifunctional molecules containing macrocyclic polyamines [12]aneN(3) and naphthyl moieties 1-3(a, b) have been designed and synthesized through efficient N-alkylation and copper-mediated alkyne-azide click reactions. Experiments on gel electrophoresis, dynamic light scattering and atomic force microscopy confirmed that 2b and 3b with two [12]aneN(3) units efficiently induced the DNA condensation at the concentration of 120 μM in less than 5 min. The condensation mechanism was studied by EB displacement fluorescence spectra, viscosity titration, and ionic strength effects. The condensation process was found to be reversible, and the presence of both naphthyl and [12]aneN(3) units in the molecules was proved to be necessary for the effective DNA condensation inductions. Cytotoxicity assay showed that the presence of triazole moieties can result in lower toxicity.  相似文献   

9.
This paper reports on a combined two-dimensional NMR and energy minimization computational characterization of the conformation of the N-(deoxyguanosyl-8-yl)aminofluorene adduct [(AF)G] positioned across adenosine in a DNA oligomer duplex as a function of pH in aqueous solution. This study was undertaken on the d[C1-C2-A3-T4-C5-(AF)G6-C7-T8-A9-C10-C11].[G12-G13-T14 -A15-G16-A17-G18- A19-T20-G21-G22] complementary undecamer [(AF)G 11-mer duplex]. The modification of the single G6 on the pyrimidine-rich strand was accomplished by reaction of the oligonucleotide with N-acetoxy-2-(acetylamino)fluorene and subsequent deacetylation under alkaline conditions. The HPLC-purified modified strand was annealed with the unmodified purine-rich strand to generate the (AF)G 11-mer duplex. The exchangeable and nonexchangeable protons are well resolved and narrow in the NMR spectra of the (AF)G 11-mer duplex so that the base and the majority of sugar nucleic acid protons, as well as several aminofluorene ring protons, have been assigned following analysis of two-dimensional NOESY and COSY data sets at pH 6.9, 30 degrees C in H2O and D2O solution. The NOE distance constraints establish that the glycosidic torsion angle is syn at (AF)G6 and anti at A17, which results in the aminofluorene ring being positioned in the minor groove. A very large downfield shift is detected at the H2' sugar proton of (AF)G6 associated with the (AF)G6[syn].A17[anti] alignment in the (AF)G 11-mer duplex. The NMR parameters demonstrate formation of Watson-Crick C5.G18 and C7.G16 base pairs on either side of the (AF)G6[syn].A17[anti] modification site with the imino proton of G18 more stable to exchange than the imino proton of G16. Several nonexchangeable aminofluorene protons undergo large downfield shifts as do the imino and H8 protons of G16 on lowering of the pH from neutrality to acidic values for the (AF)G 11-mer duplex. Both the neutral and acidic pH conformations have been defined by assigning the NOE constraints in the [C5-(AF)G6-C7].[G16-A17-G18] segment centered about the modification site and incorporating them in distance constrained minimized potential energy calculations in torsion angle space with the DUPLEX program. A series of NOEs between the aminofluorene protons and the DNA sugar protons in the neutral pH conformation establish that the aminofluorene ring spans the minor groove and is directed toward the G16-A17-G18 sugar-phosphate backbone on the partner strand.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

10.
The potency of indole-3-carbinol (I3C) to form condensation products under acidic aqueous conditions was studied. After identifying a known dimer, 3,3′-diindolylmethane (DIM), we elucidated the structures of two trimers also found in acid reaction mixtures: 5,6,11,12,17,18-hexahydrocyclonona[1,2-b:4,5-b′:7,8-b″]tri-indole (CTI), and 2,3-bis[3-indolylmethyl] indole (BII). The formation of these indole oligomers was shown to be pH dependent. The highest amounts of DIM and BII were formed in aqueous solutions having a pH value ranging from 4 to 5. No CTI could be detected at pH values above 4.5. In rats that received an oral dose of I3C we could detect DIM and BII in gastric contents, stomach tissue, small intestine and liver. No CTI could be detected in vivo after oral exposure to I3C. In in vitro experiments, using rat hepatocytes, the cytochrome P-450IA1 apoprotein level, 7-ethoxyresorufin O-deethylation activity (EROD) and DT-diaphorase activity (DTD) were markedly enhanced by DIM and CTI as well as BII.  相似文献   

11.
The novel tetrameric structure of human beta-tryptase faces each active site into the central pore, thereby restricting access of most biologic protease inhibitors. The mechanism by which the anti-tryptase mAb B12 inhibits human beta-tryptase peptidase and proteolytic activities at neutral pH, but augments proteolytic activity at acidic pH, was examined. At neutral pH, B12-beta-tryptase complexes are inactive. At acidic pH, B12 (intact and Fab) minimally affects peptidase activity when added to beta-tryptase tetramers, but does induce susceptibility to inhibition by soybean trypsin inhibitor and antithrombin III. Surprisingly, B12 Fab-beta-tryptase complexes formed at both neutral and acidic pH exhibit the apparent molecular mass of a complex with 1 beta-tryptase monomer and 1 Fab by gel filtration. B12 does not compete with heparin for binding to tryptase at either neutral or acidic pH. Thus, B12 directly disrupts beta-tryptase tetramers to monomers that are inactive at neutral pH, whereas at acidic pH, are active and more accessible to protein inhibitors and substrates.  相似文献   

12.
Mimics of carboxypeptidase A, a prototypical metalloprotease, were synthesized by linking macrocyclicpolyamines to the primary side of beta-cyclodextrin followed by complexing with Zn(II). These enzyme mimics exhibit saturation kinetics in hydrolysis of p-nitrophenyl acetate (PNPA) and enhance the rate of hydrolysis reaction by almost 300-fold. The effective molarities (EM) of the mimics range from 0.2 to 1.9 M. Origin of the rate acceleration was examined: the reactivity of Zn(II) complexes of [12]aneN3 [12]aneN4, and [14]aneN4 for hydrolyzing PNPA increases with increase in basicity of the zinc bound hydroxides [Zn(II)-OH], yielding a linear Br?nsted plot. Free hydroxide fits well on this plot. A similar plot was obtained with the enzyme mimics. The Br?nsted relationships indicate that the Zn(II)-OH in the catalytic systems hydrolyzes the ester by direct nucleophilic attack on the ester carbonyl of cyclodextrin-bound but not Zn(II)-coordinated PNPA.  相似文献   

13.
Amination of guanosine (Guo) with 2,4-dinitrophenoxyamine in aqueous DMF gave 7-amino-Guo, which was readily converted to 8,5'-O-cyclo-Guo, and 8-hydroxy-Guo. Deoxyguanosine (dG) gave only deglycosylated 7-amino-G under the same reaction condition. Aminations of Guo and dG with hydroxylamine-O-sulfonic acid above pH 9 gave the corresponding 1-amino derivatives, whereas those in acidic media at pH 2-4 gave 8-amino-Guo and 7-amino-G as the main products, respectively. Amination of Guo in neutral media gave 1-amino-Guo and decomposed products of 7-amino-Guo. The mechanisms of these amination reactions are described.  相似文献   

14.
In this paper, partitioning behaviors of typical neutral (Alanine), acidic (Glutamic acid) and basic (Lysine) amino acids into imidazolium-based ionic liquids [C4mim][PF6], [C6mim][PF6], [C8mim][PF6], [C6mim][BF4] and [C8mim][BF4] as extracting solvents were examined. [C6mim][BF4] showed the best efficiency for partitioning of amino acids. The partition coefficients of amino acids in ionic liquids were found to depend strongly on pH of the aqueous solution, amino acid and ionic liquid chemical structures. Different chemical forms of amino acids in aqueous solutions were pH dependent, so the pH value of the aqueous phase was a determining factor for extraction of amino acids into ionic liquid phase. Both water content of ionic liquids and charge densities of their anionic and cationic parts were important factors for partitioning of cationic and anionic forms of amino acids into ionic liquid phase. Extracted amino acids were back extracted into phosphate buffer solutions adjusted on appropriate pH values. The results showed that ionic liquids could be used as suitable modifiers on the stationary phase of an HPLC column for efficient separation of acidic, basic, and neutral amino acids.  相似文献   

15.
Three unprecedented purine-containing compounds, named [6]-, [8]-, and [10]-zingerines as they are 5-(6-amino-9H-purin-9-yl) analogs of [6]-, [8]-, and [10]-gingerols, respectively, were isolated from a methanolic extract of ginger rhizomes using a phase trafficking-based method that utilizes solid phase reagents allowing for fast and selective simultaneous separation of basic, acidic, and neutral components of natural products extracts.  相似文献   

16.
Dihydroflavins are facile reducing agents and potent nucleophiles. The dihydroflavin nucleophilic reactivity, as measured by the rate of covalent flavin adduct formation with tetrahydronaphthalene epoxides, is comparable to that of the thiolate anion (Y. T. Lee and J. F. Fisher (1993) J. Org. Chem. 58, 3712). In these reactions there appears subsequent to the nucleophilic cleavage of the epoxide by the dihydroflavin the product corresponding to formal hydride reduction product (at the benzylic carbon) of these epoxides. Thus the reaction of (+/-)-1a,2,3, 7b-tetrahydro-(1aalpha,2alpha,3beta,7balpha)-naphth[1,2-b]oxirene-2,3-diol (1), (+/-)-1a,2,3,7b-tetrahydro-(1aalpha,2beta,3alpha,7balpha)-naphth[1,2-b]oxirene-2,3-diol (2), and (+/-)-1a,2,3,7b-tetrahydro-(1aalpha,7balpha)-naphth[1,2-b]oxirene (3) in 9:1 (v/v) aqueous Tris buffer-dioxane, at both acidic and neutral pH, with FMNH(2) and 1,5-dihydrolumiflavin (LFH(2)) gave (following covalent flavin-epoxide adduct formation) the products having a methylene group at the benzylic position. The reduction product yield was proportional to the yield of the N(5) flavin-epoxide adduct intermediate, and the rate of the reaction was proportional to the dihydroflavin concentration. These observations are consistent with these reduction products resulting from bimolecular reaction between the dihydroflavin-epoxide adduct and a second molecule of dihydroflavin. Copyright 2000 Academic Press.  相似文献   

17.
The DNA binding orientation and dynamic behavior of Cu(II) complexes of 1,4,7-triazacyclononane ([9]aneN(3)), 1, and an acridine conjugate, 2, were investigated by DNA fiber EPR (EPR=electron paramagnetic resonance) spectroscopy. Crystal and molecular structure of 2 were determined by X-ray diffraction. It has been shown that 1 binds to DNA in two different modes at room temperature; one species is rapidly rotating and the other is immobilized randomly on the DNA. The introduction of acridine to [9]aneN(3) fixed the [Cu([9]aneN(3))](2+) moiety of 2 in two different environments on the DNA: the g(mid R:mid R:) axis of one species (g( parallel)=2.26) is aligned perpendicularly to the DNA fiber axis whereas that of the other (g( parallel)=2.24) aligns<90 degrees with the DNA fiber axis. The different DNA binding structures of 1 and 2 are reflected also in their different efficiencies of DNA cleavage; 2 was found to be more effective both in oxidative and hydrolytic cleavage reactions.  相似文献   

18.
pH-dependent pore formation properties of pardaxin analogues   总被引:4,自引:0,他引:4  
The interaction of pardaxin, a shark-repellent neurotoxin, and its charge-modified analogues with vesicles and human erythrocytes is described. The following six analogues and derivatives were synthesized by a solid phase method: [Glu8, Glu16]pardaxin, [N1-succinamido,Glu8,Glu16]pardaxin, [N1,Lys8,Lys16-triacetyl]pardaxin, des-[1----9]pardaxin (Shai, Y., Bach, D., and Yanovsky, A. (1990) J. Biol. Chem. 265, 20202-20209), and des-[1----9] [Glu16]pardaxin. The relative hydrophobic characteristics of the analogues were examined using reverse-phase high performance liquid chromatography. The pH-dependent spectroscopic and functional characteristics of the analogues were also investigated at either neutral or acidic pH. Spectroscopic characterization was achieved by measuring circular dichroism both before and after binding to vesicles, at either neutral or acidic pH. The ability of the peptides to dissipate a diffusion potential, to cause calcein release or the pH-dependent release of 8-aminonaphthalene-1,3,6-trisulfonic acid disodium salt/p-xylene-bis[pyridinium bromide] from sonicated unilamellar liposomes, as well as measurements of cytolytic activity on human erythrocytes, served to functionally characterize the peptides. We show a direct correlation between alpha-helical content, the analogues' hydrophobicity, and their pore-forming properties at the different pH values tested. We also demonstrate that the charge of the N terminus and of the peptide backbone, but not of the C terminus, affects the secondary structure as well as the activities of the analogues. Finally, we show that the cytolytic activity of pardaxin at neutral pH is not retained by any of the analogues.  相似文献   

19.
Calf pancreas microsomes incorporated radioactive D-mannose from GDP-D-[14C]mannose into lipid-bound oligosaccharides extracted with chloroform/methanol/water (10/10/2.5, v/v). Several products, which probably differed in the size of the oligosaccharide moiety, were labeled. These could be partially resolved by thin layer chromatography and DEAE-cellulose chromatography. The labeled lipid-bound oligosaccharides were retained on DEAE-cellulose more strongly than synthetic dolichyl alpha-D-[14C]mannopyranosyl phosphate. They were stable to mild alkali, but labile to acid and hot alkali. Acid treatment yielded a neutral 14C-labeled oligosaccharide fraction which was estimated by gel filtration to have a minimum of 8 monosaccharide residues. Hot alkali treatment yielded a mixture of neutral and acidic 14C-labeled oligosaccharides which could be transformed into neutral products by alkaline phosphatase. The D-[14C]mannose residues were alpha-linked at the nonreducing terminus of the oligosaccharides since they could be removed completely with alpha-mannosidase. Most of the D-[14C]mannose-labeled oligosaccharides were retained on concanavalin A Sepharose and eluted with methyl alpha-D-mannopyranoside. Pancreatic dolichyl beta-D-[14C]mannopyranosyl phosphate incubated with calf pancreas microsomes in the presence of sodium taurocholate was efficiently utilized as donor of alpha-D-mannosyl residues in lipid-bound oligosaccharides. The products formed from dolichyl beta-D-[14C]mannopyranosyl phosphate were identical with those formed from GDP-D-[14C]mannose, and evidence was obtained to show that the dolichyl beta-D-[14C]mannopyranosyl phosphate was serving as donor without prior conversion to GDP-D-[14C]mannose. Transfer of mannose from dolichyl beta-D-[14C]mannopyranosyl phosphate to lipid-bound oligosaccharides took place at a pH optimum of 7.3, whereas transfer to the precipitate containing glycoproteins was greatest at pH 6.0 in Tris/maleate buffer. The addition of divalent cation was not required, but low concentrations of EDTA were extremely inhibitory. The carbohydrate composition of the lipid-bound oligosaccharides of microsomal membranes was investigated by gas-liquid chromatography and by reduction with sodium borotritide. A heterogeneous mixture of oligosaccharides containing N-acetyl-D-glucosamine, D-mannose, and D-glucose varying in proportions from approximately 1/2.5/0.5 to 1/5/1.5 was obtained with glucosamine at the reducing end. Acid treatment of the lipid-bound oligosaccharide fraction yielded dolichyl pyrophosphate, suggesting that at least some of the oligosaccharides were linked to dolichol through a pyrophosphate group.  相似文献   

20.
Disulfide bond reduction by the CO2.- radical was investigated in aponeocarzinostatin, aporiboflavin-binding protein, and bovine immunoglobulin. Protein-bound cysteine free thiols were formed under gamma-ray irradiation in the course of a pH-dependent and protein concentration dependent chain reaction. The chain efficiency increased upon acidification of the medium, with an apparent pKa around 5, and decreased abruptly below pH 3.6. It decreased also at neutral pH as cysteine accumulated. From pulse radiolysis analysis, CO2.- proved able to induce rapid one-electron oxidation of thiols and of tyrosine phenolic groups in addition to one-electron donation to exposed disulfide bonds. The bulk rate constant of CO2.- uptake by the native proteins was 5- to 10-fold faster at pH 3 than at pH 8, and the protonated form of the disulfide radical anion, [symbol: see text], appeared to be the major protein radical species formed under acidic conditions. The main decay path of [symbol: see text] consisted of the rapid formation of a thiyl radical intermediate [symbol: see text] in equilibrium with the closed, cyclic form. The thiyl radical was subsequently reduced to the sulfhydryl level [symbol: see text] on reaction with formate, generating 1 mol of the CO2.- radical, thus propagating the chain reaction. The disulfide radical anion [symbol: see text] at pH 8 decayed through competing intramolecular and/or intermolecular routes including disproportionation, protein-protein cross-linking, electron transfer with tyrosine residues, and reaction with sulfhydryl groups in prereduced systems. Disproportionation and cross-linking were observed with the riboflavin-binding protein solely. Formation of the disulfide radical cation [symbol: see text], phenoxyl radical Tyr-O. disproportionation, and phenoxyl radical induced oxidation of preformed thiol groups should also be taken into consideration to explain the fate of the oxygen-centered phenoxyl radical.  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