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.
The rapid reaction of diisopropylfluorophosphate with a tyrosine residue of human serum albumin at 0.02 m ionic strength involves prior rapid reversible binding characterized by a dissociation constant of 3.6 × 10?3m and an apparent pKa of 8.3. The rapid reaction of p-nitrophenyl acetate with human serum albumin (G. E. Means and M. L. Bender, 1975, Biochemistry14, 4989–4994) appears to involve the same tyrosine residue and is thus stoichiometrically inhibited by prior reaction with diisopropylfluorophosphate. Both reactions are strongly inhibited by decanoate anion, strongly retarded at higher ionic strength, and reflect strong rapidly reversible binding and abnormally low tyrosine pKa values. This reactive tyrosine residue thus appears to be located in a primary binding site for small apolar anions and to be closely associated with several cationic groups.  相似文献   

2.
The association rate constant for the binding of bilirubin to bovine serum albumin has been determined in a continuous-flow experiment. The value obtained is 0.9 x 106m?1S?1. Furthermore the dissociation rate constant is determined from the rate of the peroxidase-catalyzed oxidation of bilirubin in a bilirubin-albumin solution. This figure is 3.1 × 10?2s 1. Calculation of the apparent binding equilibrium constant from the two rate constants gives 2.9 x 107m?1. The above mentioned peroxidase oxidation has also been used for a direct estimation of the binding equilibrium constant giving 2.7 × 107m?1. All experiments are carried out at 36 °C and pH 7.4.  相似文献   

3.
The ATPase of avian myeloblastosis virus (AMV) is not a recognizable cellular enzyme. It hydrolyzes ATP, GTP, ITP, UTP, and dCTP at equal rates, is inhibited by high concentrations of dithiothreitol, and is partially inhibited by 1 × 10?5mp-chloromercuribenzoic acid (PCMB) and p-chloromercuribenzene sulfonate acid (PCMBS). The inhibition by the mercurials is reversed by increasing the concentration of PCMB or PCMBS to 1 × 10?3m. The enzyme requires phospholipid for activity. Incubation with phospholipase C inhibits activity and subsequent addition of lecithin-containing saturated fatty acids partially restores activity, whereas lecithin-containing unsaturated fatty acids further inhibit activity.  相似文献   

4.
The kinetics of the binding of cyanide to ferric chloroperoxidase have been studied at 25°C and ionic strength 0.11 M using a stopped-flow apparatus. The dissociation constant (KCN) of the peroxidase-cyanide complex and both forward (k+) and reverse (k?) rate constants are independent of the H+ concentration over the pH range 2.7 to 7.1. The values obtained are kcn = (9.5 ± 1.0) × 10-5 M, k+. = (5.2 ± 0.5) × 104 M?1 sec?1 and k- = (5.0± 1.4) sec-1. In the presence of 0 06 M potassium nitrate the affinity of cyanide for chloroperoxidase decreases due to the inhibition of the forward reaction. The dissociation rate is not affected. The nitrate anion exerts its influence by binding to a protonated form of the enzyme, whereas the cyanide binds to the unprotonated form. Binding of nitrate results in an apparent shift towards higher pKa values of the ionization of a crucial heme-linked acid group. Hence the influence of this group can be detected in the accessible pH range. Extrapolation to zero nitrate concentration yields a value of 3.1±0.3 for the pKa of the heme-linked acid group.  相似文献   

5.
The objective of the current study was to observe the impact of two seasons viz. summer (February–May) and monsoon (June–August) on the blood biochemical and hormonal responses in different indigenous goats of tropical island agro-ecological environment maintained under extensive management system. Sixty animals of three different indigenous goat breeds were included in the study: Andaman local goat (AL, n = 20), Andaman local?×?Malabari (AL?×?M, n = 20), and Teressa goat (n = 20). Sixty serum samples (n = 10/season/breed) from the three groups of animals were analyzed. Study revealed that there was a significant increase (p ≤ 0.05) in serum total protein in Teressa goats during summer than monsoon. Serum albumin showed significant variation (p ≤ 0.05) between AL and AL?×?M during summer whereas significant variation of albumin (p ≤ 0.05) was observed between AL and Teressa, AL?×?M and Teressa during monsoon season. Significant differences in serum albumin (p ≤ 0.05) were also observed in AL and AL?×?M during summer and monsoon seasons, respectively. The serum cortisol levels were significantly higher (p ≤ 0.05) in AL goats during summer than in monsoon season. Thus, the study could able to establish the seasonal variation in biochemical and hormonal values of indigenous goat breeds in hot and humid tropical island environment.  相似文献   

6.
The kinetics of the reaction of Helix pomatia haemocyanin with oxygen have been studied under conditions where ligand binding is co-operative (n = 4.5). The dissociation of oxygen from oxyhaemocyanin in the presence of sodium dithionite and the combination of deoxyhaemocyanin with oxygen were studied by the stopped-flow technique. The combination with oxygen, as well as the dissociation of oxyhaemocyanin, are clearly autocatalytic. The initial rate constant for oxygen combination to the fully deoxygenated state is 0.2 to 0.3 × 106m?1 s?1; during the course of the reaction the rate constant increases to a value higher than 106m?1s?1.The initial rate of oxygen dissociation from fully saturated haemocyanin is 10 s?1, increasing to about 30 s?1 as the reaction proceeds. Thus, both the combination and the dissociation rate constants contribute to the co-operativity of oxygen binding.Temperature-jump relaxation experiments were carried out at fractional oxygen saturations larger than 0.7. The dependence of the relaxation rate upon the concentration of the reactants indicates the presence of one principal bimolecular process. The calculated combination and dissociation rate constants for this process are: 3.8 × 106m?1 s?1 and 10 s?1, respectively. Evidence is presented which shows that the transition from the T-state to the R-state of the protein is relatively slow. Both the T and R-state seem to be largely stabilized at the expense of intermediate states.Under other conditions, where oxygen binding is non-co-operative, temperature-jump and stopped-flow experiments reveal considerable kinetic heterogeneity.  相似文献   

7.
The partial purification of shikimate dehydrogenase (SDH) from tomato fruit was achieved by precipitation with ammonium sulphate, and chromatography on DEAE-cellulose and hydroxyapatite. The enzyme has a MW of 73000, shows an optimum at pH 9.1 and Km values of 3.8 × 10?5 M and 1.0 × 10?5 M with shikimic acid and NADP as substrates. NADP could not be replaced by NAD. The tomato enzyme is competitively inhibited by protocatechuic acid with a Ki value of 7.7 × 10?5 M. On the other hand, cinnamic acid derivatives and 2-hydroxybenzoic acid were ineffective. At 50° for 5 min the SDH is inactivated by 85%. The activity was inhibited by pCMB and N-ethylmaleimide, suggesting a requirement for SH groups. The inactivation plot of oxidation by pCMB was biphasic, and NADP decreased the reactivity of sulphydryl groups to the reagent. The activation energy was found to be 14.2kcal/mol. The properties of the SDH are discussed in relation to the enzymes from other sources.  相似文献   

8.
Dissociation and alkali complex formation equilibria of nitrilotris(methylenephosphonic acid) (NTMP, H6L) have been studied by dilatometric, potentiometric and 31P NMR-controlled titrations. Dilatometry indicated the formation of alkali complexes ML (M=Li, Na, K, Rb, Cs) at high pH with a stability decreasing from Li to Cs. An efficient combination of potentiometric and NMR methods confirmed two types of alkali metal complexes MHL and ML. Stability constants for the equilibria following M+ + HL5− ? MHL4− and M+ + L6− ? ML5−, respectively, were determined: logKNaHL=1.08(0.07), logKKHL=0.86(0.08), logKNaL=2.24(0.03). Systematic errors are introduced by using alkali metal hydroxides as titrants for routine potentiometric determinations of dissociation constants pKa5app and pKa6app. Correction formulae were derived to convert actual dissociation constants pKa into apparent dissociation constants pKaapp (or vice versa). The actual dissociation constants were found: pKa5(H2L4− ? H+ + HL5−)=7.47(0.03) and pKa6(HL5− ? H+ + L6−)=14.1(0.1). The anisotropy of 31P chemical shifts of salts MnH6 − nL (M=Li, Na, n=0-5) is more sensitive towards titration (n) than isotropic solution state chemical shifts.  相似文献   

9.
The kinetics of uptake and retention of β-ecdysone by imaginal discs from late third instar larvae of Drosophila melanogaster correspond well with those of the first synthetic response of discs to hormone, an increase in RNA synthesis.Competition studies indicate the presence of two types of hormone binding sites, specific and non-specific. The specific sites are saturated at hormone concentrations which fully induce morphogenesis. Results are consistent with the hypothesis that analogs which induce morphogenesis at differing concentrations bind to the same sites. Experiments with the inhibitors N-ethylmaleimide, actinomycin d, and cycloheximide suggest that the binding sites are pre-existing in the cell and require functional sulfhydryl groups for binding.Specific binding, binding that is competed by excess unlabeled β-ecdysone, is saturable (70–80 nM). Kinetic rate constants for this specific binding were estimated to be ka = 1.5 × 105M?1 min?1, kd = 3 × 10?2 min?1. The equilibrium dissociation constant calculated from the kinetic rate constants was Keq = 2 × 10?7M compared to 1.7 × 10?7M β-ecdysone required to induce morphogenesis in vitro and 2.5 × 10?7M determined to be the in vivo concentration at the time of induction of morphogenesis.  相似文献   

10.
We developed a novel human leukocyte antigen HLA–ABC locus-specific quantitative real-time polymerase chain reaction (PCR) to determine the locus-specific gene expression of HLA–ABC in peripheral blood leukocytes (PBLs, n?=?53), colon mucosa (n?=?15), and larynx mucosa (n?=?15). Laser-assisted tissue microdissection allowed us to study the selected cells without interference from surrounding stroma. We report evidence on the specificity of the technique, describing the HLA–ABC locus-specific gene expression patterns found in the PBLs and two solid tissues studied. PBLs showed a higher gene expression of HLA-B than of HLA-A or HLA-C (p?=?4.7?×?10?10 and p?=?1.6?×?10?6, respectively). In solid tissue, HLA-A and HLA-B gene expressions were similar and HLA-C expression lower. In particular, in larynx mucosa, significant differences were found between HLA-A and HLA-C expressions and between HLA-B and HLA-C expressions (p?=?6.5?×?10?4 and p?=?8.1?×?10?4, respectively). The same differences were observed in colon mucosa, but significance was not reached (p?=?0.08 and p?=?0.06, respectively). Differences in locus-specific regulation may be related to the control of cytotoxic responses of NK and CD8 positive T cells. Gene expression of HLA–ABC specific locus showed no intra-individual variability, but there was a high inter-individual variability. This may result from differences in the expression of common regulatory factors that control HLA–ABC constitutive expression.  相似文献   

11.
The interactions of fatty acids with porcine and bovine β-lactoglobulins were measured using tryptophan fluorescence enhancement. In the case of bovine β-lactoglobulin, the apparent binding constants for most of the saturated and unsaturated fatty acids were in the range of 10?7 M at neutralpH. Bovine β-lactoglobulin displays only one high affinity binding site for palmitate with an apparent dissociation constant of 1·10?7 M. The strength of the binding was decreasing in the following way: palmitate > stearate > myristate > arachidate > laurate. Caprylic and capric acids are not bound at all. The affinity of β-lactoglobulin for palmitate decreased as thepH of the incubation medium was lowered and BLG/palmitate complex was not observed atpH's lower than 4.5. Surprisingly, chemically modified bovine β-lactoglobulin and porcine β-lactoglobulin did not bind fatty acids in the applied conditions.  相似文献   

12.
Docosahexaenoic acid is found to be bound to three equivalent sites on albumin with the same affinities as palmitic acid at 0–38°C, which demonstrates that ethene-1,2-diyl- and methylene-groups contribute equally to the affinity. The equilibrium dissociation constants (K dm s) for red cell membrane binding sites of linoleic- and docosahexaenoic acid at pH 7.3 are determined at temperatures between 0 and 37°C. The temperature-independent capacities for binding are 12 ± 1 and 25.4 ± 3.0 nmoles g−1 ghosts respectively. Double isotope binding experiments reveal that the unsaturated fatty acids: arachidonic-, linoleic-, docosahexaenoic-, and oleic acid have partially shared capacities in ratios approximately 1:2:4:5, in contrast to the noncompetitive binding of palmitic acid. The observations suggest a two-tier binding limitation. One is the number of protein sites binding fatty acid anions electrostatically and the other is the number of suitable annular lipids adaptively selected among membrane lipids by the hydrocarbon chain. These competition conditions are confirmed by measurements of the tracer exchange efflux at near 0°C from albumin-free and albumin-filled ghosts of linoleic- and docosahexaenoic acid, either alone or in the presence of arachidonic- and palmitic acid. Under equilibrium conditions, the calculated ratios of inside to outside membrane binding is below 0.5 for four unsaturated fatty acids. The unidirectional rate constants of translocation between the inside and the outside correlate with the number of double bonds in these fatty acids, which are also correlated with the dissociation rate constants of the complexes with albumin. The membrane permeation occurs presumably by binding of the anionic unsaturated fatty acids to an integral protein followed by channeling of the neutral form between opposite binding sites of the protein through annular lipids encircling the protein. Received: 30 June 1997/Revised: 23 February 1998  相似文献   

13.
Genome-wide association (GWA) studies have identified many candidate genes that are associated with blood lipid and lipoprotein concentrations. In this study, we want to know whether the results from European for lipid-related single-nucleotide polymorphisms (SNPs) are generalizable to Chinese children. We genotyped seven SNPs in Chinese school-age children (n = 3,503) and assessed the associations of these SNPs with lipids profiles and dyslipidemia. After false discovery rate correction, of the seven SNPs, six (rs2144300, p ~ 9.30 × 10?3; rs1260333, p ~ 6.20 × 10?11; rs1260326, p ~ 8.73 × 10?11; rs10105606, p ~ 0.010; rs1748195, p ~ 0.016 and rs964184, p ~ 2.33 × 10?13) showed strong association with triglycerides. Three SNPs (rs1260333, p ~ 3.30 × 10?3; rs1260326, p ~ 4.39 × 10?3 and rs2954029, p ~ 6.36 × 10?4) showed strong association with total cholesterol. Two SNPs (rs10105606, p ~ 6.66 × 10?4 and rs1748195, p ~ 2.55 × 10?3) showed strong association with high density lipoprotein cholesterol. Four SNPs (rs1260333, p ~ 0.017; rs1260326, p ~ 0.013; rs2954029, p ~ 1.09 × 10?3 and rs964184, p ~ 5.51 × 10?3) showed strong association with low density lipoprotein cholesterol. There were significant associations between rs1260333 (OR is 0.82, 95 % CI 0.74–0.92, p ~ 3.96 × 10?4), rs1260326 (OR is 0.82, 95 % CI 0.74–0.92, p ~ 5.31 × 10?4), and rs964184 (OR is 1.36, 95 % CI 1.20–1.55, p ~ 1.89 × 10?6) and dyslipidemia. These SNPs generated strong combined effects on lipid profiles and dyslipidemia. Our study demonstrates that SNPs associated with lipids from European GWA studies also play roles in Chinese children, which broadened the understanding of lipids metabolism.  相似文献   

14.
Recently, microalgae have gained a lot of attention because of their ability to produce fatty acids in their surrounding environments. The present paper describes the influence of organic carbon on the different fatty acid pools including esterified fatty acids, intracellular free fatty acids and extracellular free fatty acids in Ochromonas danica. It also throws light on the ability of O. danica to secrete free fatty acids in the growth medium under photoautotrophic and mixotrophic conditions. Biomass production of photoautotrophically grown O. danica was higher than that of mixotrophically grown, where a cellular biomass formation of 1.8 g L?1 was observed under photoautotrophic condition which was about five folds higher than that under mixotrophic conditions. Contrary, the esterified fatty acid content reached up to 99 mg g?1 CDW under photoautotrophic conditions at the late exponential phase, while during mixotrophic conditions a maximum of 212 mg g?1 CDW was observed at the stationary phase. Furthermore, O. danica cells grown under mixotrophic conditions showed higher intracellular free fatty acid and extracellular free fatty acid contents (up to 51 and 20 mg g?1 CDW, respectively) than cells grown under photoautotrophic conditions (up to 26 and 4 mg g?1 CDW, respectively). The intra- and extracellular free fatty acids consisted of a high proportion of polyunsaturated fatty acids, mainly C18:2n?6, C18:3n?3 and C20:4n?6.  相似文献   

15.
Human serum heme–albumin (HSA–heme–Fe) displays reactivity and spectroscopic properties similar to those of heme proteins. Here, the nitrite reductase activity of ferrous HSA–heme–Fe [HSA–heme–Fe(II)] is reported. The value of the second-order rate constant for the reduction of $ {\text{NO}}_{2}^{ - } $ to NO and the concomitant formation of nitrosylated HSA–heme–Fe(II) (i.e., k on) is 1.3 M?1 s?1 at pH 7.4 and 20 °C. Values of k on increase by about one order of magnitude for each pH unit decrease between pH 6.5 to 8.2, indicating that the reaction requires one proton. Warfarin inhibits the HSA–heme–Fe(II) reductase activity, highlighting the allosteric linkage between the heme binding site [also named the fatty acid (FA) binding site 1; FA1] and the drug-binding cleft FA2. The dissociation equilibrium constant for warfarin binding to HSA–heme–Fe(II) is (3.1 ± 0.4) × 10?4 M at pH 7.4 and 20 °C. These results: (1) represent the first evidence for the $ {\text{NO}}_{2}^{ - } $ reductase activity of HSA–heme–Fe(II), (2) highlight the role of drugs (e.g., warfarin) in modulating HSA(–heme–Fe) functions, and (3) strongly support the view that HSA acts not only as a heme carrier but also displays transient heme-based reactivity.  相似文献   

16.
The binding of alizarin yellow G—an azo derivative of salicylic acid—by bovine serum albumin has been investigated using the method of equilibrium dialysis. Six strong and a number of additional, weak binding sites have been found to be present. The system is characterized by strong positive cooperativity between the first and second sites. Six binding constants have been determined on the basis of a simplified mathematical model. The results are ~2 × 104m?1 for the first binding site, 6 × 105m?1 for the second, and between 4 × 104 and 105m?1 for the rest. The phenomenon is discussed in terms of the existence of various conformers or of the conformational adaptability of albumin. Cobinding by salicylic acid does not displace alizarin yellow G but induces a conformational change in the protein which affects the absorption spectrum of the bound dye. As expected for this kind of heterotropic interaction, the spectrum of the system albumin-salicylic acid is similarly affected by the cobinding of alizarin yellow G.  相似文献   

17.
Arthrobacter sialophilus neuraminidase catalyzes the hydration of 5-acetamido-2,6-anhydro-3, 5-dideoxy-d-glycero-d-galacto-non-2-enonic acid (2,3-dehydro-AcNeu) with Km and kcat values of 8.9 × 10?4m and 6.40 × 10?4 s?1, respectively. The methyl ester of 2,3-dehydro-AcNeu as well as 2,3-dehydro-4-epi-AcNeu are also hydrated by the enzyme. The product resulting from the enzymatic hydration of 2,3-dehydro-AcNeu is N-acetylneuraminic acid. A series of derivatives of 2,3-dehydro-AcNeu (KI 1.60 × 10?6m) including 2,3-dehydro-4-epi-AcNeu (2.10 × 10?4m) and 2,3-dehydro-4-keto-AcNeu (KI = 6.10 × 10?5 m) were each competitive inhibitors of the enzyme. The methyl esters of these ketal derivatives were also competitive enzyme inhibitors. Dissociation constants for these ketals were determined independently by fluorescence enzyme titrations which gave values similar to those found kinetically. These six relatives of 2,3-dehydro-AcNeu were also competitive inhibitors for the influenza viral neuraminidases. For the viral neuraminidases, the dissociation constant for 2,3-dehydro-AcNeu and its methyl ester were 2.40 × 10?6 and 1.17 × 10?3m, respectively. The interpretation placed upon the KI values determined for these ketals against the Arthrobacter versus influenza neuraminidases is that the bacterial enzyme has a more flexible glycone binding site.  相似文献   

18.
Out of the three cadmium-binding proteins (CD-BPs) in rat liver parenchyma (40K, 29K, and 24K CdBPs), the 40K Cd-BP showed the highest affinity for cadmium (Cd), with a dissociation constant (KD) of 1.2 × 10?8M. This is in between the affinity of human serum albumin KD = 3.8 × 10?5M) and metallothionein (KD = < 10?11). These Cd-BPs may be responsible for hepatic sequestration of cadmium.  相似文献   

19.
The analysis of the effect of pH upon the rate of polymerization indicates that the activity of yeast RNA polymerase I is optimal between pH 7.5 and 9 and depends on the ionization state of two groups with apparent pKa values of 6.5 and 10. Yeast RNA polymerase I is extremely labile at acid pH. Below pH 5 the enzyme is irreversibly inactivated by [H+], with a second-order rate constant of 1.6 × 10?4m?1 min?1. Sucrose gradient sedimentation and gel electrophoresis analysis of the enzyme inactivated at acid pH indicates the sequential dissociation of several enzyme subunits. The polypeptides of 44,000 and 24,000 daltons dissociate first from the enzyme core followed by the dissociation of the polypeptides of 48,000 and 36,000 daltons.  相似文献   

20.
The binding characteristics of the interaction between 3-(2-cyanoethyl) cytosine (CECT) and human serum albumin (HSA) were investigated using fluorescence, UV absorption spectroscopic and molecular modeling techniques under simulative physiological conditions. The intrinsic fluorescence intensity of HSA was decreased with the addition of CECT. The fluorescence data handled by Stern–Volmer equation proved that the quenching mechanism of the interaction between CECT and HSA was a static quenching procedure. The binding constants evaluated utilizing the Lineweaver–Burk equation at 17, 27 and 37?°C, were 2.340?×?104, 2.093?×?104 and 1.899?×?104?L?mol?1, respectively. The thermodynamic parameters were calculated according to van’t Hoff equations. Negative enthalpy (ΔH) and positive entropy (ΔS) values indicated that both hydrogen bond and hydrophobic force played a major role in the binding process of CECT to HSA, which was consistent with the results of the molecular modeling study. In addition, the effect of other ions on the binding constant of CECT-HSA was examined.  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