首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
M Le Hir 《Enzyme》1991,45(4):194-199
An impurity, probably an anion, present in some batches of the buffer substances 4-(2-hydroxyethyl) piperazine-1-ethanesulfonic acid (HEPES), 2-morpholinoethane sulfonic acid (Mes) and piperazine-1,4-bis(2-ethane sulfonic acid (Pipes), activates the soluble 5'-nucleotidase from rat kidney. The affinity of the enzyme for 5'-IMP and the Vmax were both increased by the unidentified activator. ATP and 2,3-diphosphoglycerate, known activators of the soluble 5'-nucleotidase, had no effect if the incubation media were buffered with batches containing high concentrations of the activating impurity. These results suggest that the impurity interacts with the soluble 5'-nucleotidase at the same site as ATP and 2,3-diphosphoglycerate, however with a much higher affinity than these two compounds. It is possible that the same impurity might interfere with other proteins for which ATP is a substrate or a ligand.  相似文献   

2.
Interactions of both purified tubulin and microtubule protein (tubulin plus associated proteins) with two commonly used sulfonate buffers were examined. 1,4-Piperazineethanesulfonate (Pipes) and 4-morpholineethanesulfonate (Mes) at high concentrations induce the polymerization of purified tubulin in reactions requiring only buffer, tubulin and GTP. While both reactions were temperature-dependent, cold-reversible and inhibited by GDP, colchicine or Ca2+, there were significant differences between them. Substantially lower tubulin and buffer concentrations were required for Pipes-induced polymerization; and turbidity was much more intense in the Pipes-induced than in the Mes-induced reaction at the same protein concentration. Electron microscopy demonstrated that for the most part typical smooth-walled microtubules were formed in Mes, while aberrant forms were the predominant structures formed in Pipes. When the polymerization of microtubule protein was examined as a function of buffer concentration, biphasic patterns were observed with both Pipes and Mes: polymerization occurred at both low and high, but not intermediate, buffer concentrations. The turbidity observed at high concentrations of Pipes greatly exceeded that at low concentrations. With Mes, equivalent turbidity developed at both high and low buffer concentrations. Although associated proteins copolymerized with tubulin at low buffer concentrations, they were excluded from the polymerized material at high buffer concentrations. Pipes and Mes were compared to sodium phosphate, Tris/HCl and imidazole/HCl buffers at 0.1 M in several polymerization systems using both purified tubulin and microtubule protein. The sulfonate buffers were invariably associated with more vigorous reactions than the other buffers.  相似文献   

3.
The interaction of DNA with various neutral pH, amine-based buffers has been analyzed by free solution capillary electrophoresis, using a mixture of a plasmid-sized DNA molecule and a small DNA oligonucleotide as the reporter system. The two DNAs migrate as separate, nearly Gaussian-shaped peaks in 20-80 mM TAE (TAE, Tris-acetate-EDTA; Tris, tris[hydroxymethyl]aminomethane) buffer. The separation between the peaks gradually increases with increasing TAE buffer concentration because of differences in solvent friction between large and small DNA molecules. The two DNAs form complexes with the borate ions in TBE (Tris-borate-EDTA) buffer, with mobilities that depend on the DNA/borate ratio. In 45 mM TBE buffer, the two DNAs comigrate as a single sharp peak, with a mobility that is faster than either of the constituent DNAs in the same buffer. Hence, the mixed DNA-borate complex is stabilized by the binding of additional borate ions, possibly forming bridges between the different DNAs. The mixed DNA-borate complex is gradually dissociated into its component DNAs by increasing the TBE concentration, possibly because the borate binding sites become saturated at high buffer concentrations. Other neutral pH, amine-based buffers, such as Mops (3-[N-morpholino]propanesulfonic acid), Hepes (N-[2-hydroxyethyl]piperazine-N'-[2-ethanesulfonic acid]), Bes (N,N-bis[2-hydroxyethyl]-2-aminoethanesulfonic acid), Tes (N-tris[hydroxymethyl]methyl-2-aminoethanesulfonic acid), and tricine (N-tris[hydroxymethyl]methylglycine) also form complexes with DNA, giving distorted peaks in the electropherograms. The combined results indicate that borate buffers and most neutral pH, amine-based buffers interact with DNA.  相似文献   

4.
The functional characteristics of hemoglobin (Hb) depend on oxygenation-linked proton and anion binding and thus on solvent buffer groups and ionic composition. This study compares the oxygenation properties of human Hb in ionic [tris(hydroxymethyl)aminomethane (Tris) and BisTris] buffers with those in zwitterionic N-2-hydroxy-ethylpiperazine-N'-2-ethanesulfonic acid (HEPES) buffer under strictly controlled chloride concentrations at different pH values, two temperatures, and in the absence and presence of the erythrocytic cofactor, 2,3-diphosphoglycerate (DPG). In contrast to earlier studies (carried out at the same or different chloride concentrations) it shows only small buffer effects that are manifested at low chloride concentration and high pH. These observations suggest chloride binding to the Tris buffers, which reduces the interaction with specific chloride binding sites in the Hb. The findings indicate that HEPES allows for more accurate assessment of Hb-oxygen affinity and its anion and temperature sensitivities than ionic buffers and advocates standard use of HEPES in studies on Hb function. Precise oxygen affinities of Hb dissolved in both buffers are defined under standard conditions.  相似文献   

5.
Homogenization of rat liver in Hepes (N-2-hydroxyethylpiperazine-N′-2-ethane-sulfonic acid), MOPS (2-[N-morpholino]ethanesulfonic acid), Na phosphate, Pipes (piperazine-N,N′-bis[2-ethanesulfonic acid]), TEA (triethanolamine), TES (N-tris[hydroxymethyl]-methyl-2-aminoethanesulfonic acid), Tricine (N-tris-[hydroxymethyl]methylglycine), or Tris (tris[hydroxymethyl]aminomethane), and subsequent assay for supernatant total and holo tyrosine aminotransferase activity using these buffers yields apparent enzyme concentrations which vary depending upon the buffer composition, the ionic strength, and the fold-dilution of the supernatant. A precipitous decrease in the apparent holoenzyme concentration results from a slight dilution of the supernatant with most of the buffers. Some of the dilution effects may be due to dissociation of pyridoxal phosphate from the apoenzyme or to competition between the buffer and pyridoxal phosphate for association with the enzyme. The percentage of the apparent total enzyme which exists as holoenzyme varies from 3% for supernatant prepared in Na phosphate buffer up to 94% for that prepared in Hepes. Inactivation of total enzyme activity occurs to a similar extent resulting from incubation of liver homogenates prepared with Na phosphate, Hepes, or Pipes. The residual apparent holoenzyme activity observed when assayed in the presence of Na phosphate may be due to reaction of an enzyme other than tyrosine aminotransferase. The data provide a basis for explaining the large variation in reported percentage holoenzyme and should also serve as a warning for other holoenzyme assays which use pyridoxal phosphate as a cofactor.  相似文献   

6.
Good's zwitterionic buffers are widely used in biological and biochemical research in which hydrogen peroxide is a solution component. This study was undertaken to determine whether Good's buffers exhibit reactivity toward H(2)O(2). It is found that H(2)O(2) oxidizes both morpholine ring-containing buffers (e.g., Mops, Mes) and piperazine ring-containing zwitterionic buffers (e.g., Pipes, Hepes, and Epps) to produce their corresponding N-oxide forms. The percentage of oxidized buffer increases as the concentration of H(2)O(2) increases. However, the rate of oxidation is relatively slow. For example, no oxidized Mops was detected 2h after adding 0.1M H(2)O(2) to 0.1M Mops (pH 7.0), and only 5.7% was oxidized after 24h exposure to H(2)O(2). Thus, although all of these buffers can be oxidized by H(2)O(2), their slow reaction does not significantly perturb levels of H(2)O(2) in the time frame and at the concentrations of most biochemical studies. Therefore, the previously reported rapid loss of H(2)O(2) produced from the ferroxidase reaction of ferritin is unlikely due to reaction of H(2)O(2) with buffer, a conclusion supported by the fact that H(2)O(2) is also lost rapidly when the solution pH of the ferroxidase reaction is controlled by a pH stat apparatus in the absence of buffer.  相似文献   

7.
The steady-state kinetics of CO2 hydration catalyzed by human carbonic anhydrase I (carbonate hydro-lyase, EC 4.2.1.1) has been investigated at three pH values corresponding to different parts of the pH-rate profile. Two buffer systems with similar pKa values were used at each pH. The results show that the catalyzed rates depend on the buffer concentration but also on the chemical nature of the buffer. For example, at pH 8.8 the buffer 1,2-dimethylimidazole behaves formally as a second substrate in a 'ping-pong' mechanism yielding a maximal kcat value of 2.2 x 10(5) s-1, whereas much lower rates were obtained with Taps buffers. Similarly, at pH 7.3 1-methylimidazole yields higher rates than Mops and at pH 6.3 3,5-lutidine is more efficient than Mes. Non-Michaelis-Menten kinetics were observed with all buffers except 1,2-dimethylimidazole. In addition, while the apparent buffer activation by 1,2-dimethylimidazole can be described by a single Km value of 26 mM, the Mes concentration dependence is consistent with the presence of two components of similar magnitudes with Km values of 45 mM and 0.15 mM. These results are interpreted within the framework of the 'zinc-hydroxide' mechanism in terms of multiple pathways for the rate-contributing transfer of a proton from the zinc-bound water molecule, formed during CO2/HCO3- interconversion, to the reaction medium, thus, regenerating zinc-bound OH-.  相似文献   

8.
The kinetics of microperoxidase-11 (MP-11) in the oxidation reaction of guaiacol (AH) by hydrogen peroxide was studied, taking into account the inactivation of enzyme during reaction by its suicide substrate, H2O2. Concentrations of substrates were so selected that: 1) the reaction was first-order in relation to benign substrate, AH and 2) high ratio of suicide substrate to the benign substrate, [H2O2] > [AH]. Validation and reliability of the obtained kinetic equations were evaluated in various nonlinear and linear forms. Fitting of experimental data into the obtained integrated equation showed a close match between the kinetic model and the experimental results. Indeed, a similar mechanism to horseradish peroxidase was found for the suicide-peroxide inactivation of MP-11. Kinetic parameters of inactivation including the intact activity of MP-11, alphai, and the apparent inactivation rate constant, ki, were obtained as 0.282 +/- 0.006 min(-1) and 0.497 +/- 0.013(-1) min at [H2O2] = 1.0 mM, 27 degrees C, phosphate buffer 5.0 mM, pH = 7.0. Results showed that inactivation of microperoxidase as a peroxidase model enzyme can occur even at low concentrations of hydrogen peroxide (0.4 mM).  相似文献   

9.
The isothermal titration calorimetry (ITC) technique supported by potentiometric titration data was used to study the interaction of zinc ions with pH buffer substances, namely 2‐(N‐morpholino)ethanesulfonic acid (Mes), piperazine‐N,N′‐bis(2‐ethanesulfonic acid) (Pipes), and dimethylarsenic acid (Caco). The displacement ITC titration method with nitrilotriacetic acid as a strong, competitive ligand was applied to determine conditional–independent thermodynamic parameters for the binding of Zn(II) to Mes, Pipes, and Caco. Furthermore, the relationship between the proposed coordination mode of the buffers and the binding enthalpy has been discussed. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

10.
At pH 7.4, CO2, rather than HCO3-, markedly enhances the oxidation of diverse substrates by SOD1 plus H2O2. Since the concentration of CO2 would fall with rising pH in HCO3- buffers, it was of interest to explore the effects of pH on the peroxidase activity of SOD1 in the presence and in the absence of HCO3-. The rate of NADPH peroxidation in the HCO3- buffer was minimally affected by pH in the range of 8-10.5; in a pyrophosphate buffer, the rate increased markedly, such that at pH 10.5 the rates in the two buffers were nearly identical. Similar results were obtained when urate was used as the peroxidizeable substrate. These results are explicable on the basis of an increase in the rate with pH due to the ionization of H2O2 to the effective HO2- coupled with a decrease in [CO2] due to the ionizations of H2CO3, which displaces the hydration equilibrium to the right. These two opposing effects counteract in the HCO3(-)-buffered reaction mixtures; in the pyrophosphate buffer, only the effect of increasing [H02-] was seen.  相似文献   

11.
Arabidopsis thaliana P1 protein was crystallized using the hanging drop vapor-diffusion method in 0.1 M piperazine-1, 4-bis(2-ethanesulfonic acid) buffer, containing 14% polyethylene glycol 6000 and 0.2 M magnesium acetate at pH 6.5 and 20 degrees C. The crystals are orthorhombic and belong to the space group P2(1)2(1)2(1) with unit cell dimensions of a=49.8, b=122.4 and c=149. 9 A. The diffraction data up to 2.9 A were collected by a multiwire area detector.  相似文献   

12.
Zimmerman DC 《Plant physiology》1968,43(10):1656-1660
The delay in, or loss of, flaxseed lipoxidase activity in N-tris (hydroxymethyl) methylglycine and N-2-hydroxyethylpiperazine-N′-2-ethanesulfonic acid buffers with linolenic acid as a substrate appears due to an alteration of the lipid micelle. Flaxseed lipoxidase activity is dependent on the ionic strength of the assay solution. These effects are not observed with linoleic acid as substrate. The influence of these 2 buffers on linolenic acid micelles may have a direct bearing on recent reports of chloroplast structure and activity in these buffers.  相似文献   

13.
Euglena were cultured under 3 W m-2 constant white light. In culture medium, cells show immediate and long lasting step-down photophobic responses and photoaccumulation behavior to blue light if dim red light-adapted for 30 min. However, if cells are suspended in buffered, saltcontaining solutions (adaptation buffers), strong step-down photobehavior and photoaccumulation responses are not observed for several hours. These behaviors gradually increase in strength to reach a maximum after 6–12 h; after which a stable response is maintained. The relative rates of appearance and the relative strengths of the responses are influenced by the concentrations of Ca2+ and K+, but not H+ or Na+ ions, in the adaptation buffers. Expression of the stepdown photobehavior thus requires that the cells adapt to the chemical environment in which they are suspended.Abbreviations Hepes N-2-hydroxypiperazine-N-2-ethanesulfonic acid - Mes 2(N-morpholino)-ethanesulfonic acid - Pipes piperazine-N,N-bis (2-ethanesulfonic acid) - Taps tris(hydroxymethyl) methylaminopropanesulfonic acid This work was supported, in part, by grant No. PCM-79-05320 from the U.S. National Science Foundation to B.D.  相似文献   

14.
Ascorbate and several polyphenolic compounds have been reported to undergo oxidation in cell culture media to generate hydrogen peroxide (H?0?), but the mechanism underlying this has not been established. We therefore investigated the parameters affecting H?0? production. H?0? gene ration from ascorbate, gallic acid and other phenolic compounds in Dulbecco's Modified Eagles' Medium (DMEM) at 37°C under 95% air - 5% C0? was not significantly inhibited by high (5-10 mM) concentration of EGTA, o-phenanthroline or desferrioxamine, but partial inhibition by EDTA and diethylenetriaminepentaacetic acid (DTPA) was observed. Incubation of DMEM alone at 37°C led to an upward drift of pH, even under an atmosphere of 95% air - 5% C0?. Prevention of this pH rise by increasing the concentration of N-[2-hydroxyethyl]piperazine-N'-[2-ethanesulfonic acid] (Hepes) buffer lowered the levels of H?0? generated by ascorbate and phenolic compounds, but there was still substantial H?0? generated at pH 7.4. Mixtures of ascorbate and phenolic compounds led to less H?0? generation than would be expected from the rates observed with ascorbate or phenolic compounds alone. Ascorbate prevented the loss of gallic acid incubated in DMEM. The role of metal ions and other constituents of the culture medium in promoting H?0? generation is discussed.  相似文献   

15.
S L Johnson  P T Tuazon 《Biochemistry》1977,16(6):1175-1183
The rate of the primary acid modification reaction of 1,4-dihydronicotinamide adenine dinucleotide (NADH) and 1,4-dihydro-3-acetylpyridine adenine dinucleotide (APADH) and their analogues has been studied over a wide pH range (pH 1-7) with a variety of general acid catalysts. The rate depends on [H+] at moderate pH and becomes independent of [H+] at low pH. This behavior is attributed to substrate protonation at the carbonyl group (pK of NADH = 0.6). The reaction is general acid catalyzed; large solvent deuterium isotope effects are observed for the general acid and lyonium ion terms. Most buffers cause a linear rate increase with increasing buffer concentration, but certain buffers cause a hyperbolic rate increase. The nonlinear buffer effects are due to complexation of the buffer with the substrate, rather than to a change in rate-limiting step. The rate-limiting step is a proton transfer from the general acid species to the C5 position of the substrate. Anomerization is not a necessary first step in the case of the primary acid modification reaction of beta-NADH, in which beta to alpha anomerization takes place.  相似文献   

16.
Ascorbate and several polyphenolic compounds have been reported to undergo oxidation in cell culture media to generate hydrogen peroxide (H
2
O
2
), but the mechanism underlying this has not been established. We therefore investigated the parameters affecting H
2
O
2
production. H
2
O
2
generation from ascorbate, gallic acid and other phenolic compounds in Dulbecco's Modified Eagles' Medium (DMEM) at 37°C under 95% air - 5% CO
2
was not significantly inhibited by high (5-10 mM) concentration of EGTA, o-phenanthroline or desferrioxamine, but partial inhibition by EDTA and diethylenetriaminepentaacetic acid (DTPA) was observed. Incubation of DMEM alone at 37°C led to an upward drift of pH, even under an atmosphere of 95% air - 5% CO
2
. Prevention of this pH rise by increasing the concentration of N-[2-hydroxyethyl]piperazine-N'-[2-ethanesulfonic acid] (Hepes) buffer lowered the levels of H
2
O
2
generated by ascorbate and phenolic compounds, but there was still substantial H
2
O
2
generated at pH 7.4. Mixtures of ascorbate and phenolic compounds led to less H
2
O
2
generation than would be expected from the rates observed with ascorbate or phenolic compounds alone. Ascorbate prevented the loss of gallic acid incubated in DMEM. The role of metal ions and other constituents of the culture medium in promoting H
2
O
2
generation is discussed.  相似文献   

17.
Increased visualization of microtubules by an improved fixation procedure.   总被引:10,自引:0,他引:10  
We have found that when a buffer utilized for in vitro polymerization of microtubules, i.e., 1 mM guanosine triphosphate, 1 mM MgSO4, 2 mM ethylene glycol bis(beta-aminoethyl ether)-N, N'-tetraacetic acid 100 mM piperazine-N,N'-bis(2-ethanesulfonic acid), pH 6.9 polymerization mix, was used in the glutaraldehyde prefixation regimen instead of classical fixative buffers, i.e., isotonic cacodylate or phosphate buffer, the following features were observed in thin-sections of the cytoplasm of interphase HeLa cells: (a) a greater than 2-fold increase in total microtubule contour length, (b) a 2-fold increase in a number of microtubules greater than or equal to 1 mu long, (c) an enhanced association of microtubules with cytoplasmic organelles, and (d) an increased clustering of 100 A filaments located in a perinuclear region of the cell. Furthermore, we found that after we incubated purified chick brain microtubules on a Sephadex G-25 column pre-equilibrated with polymerization mix, cacodylate or phosphate buffer at 37 degrees C, and then eluted the microtubules at 37 degrees C, the exposure to cacodylate or phosphate buffer caused extensive depolymerization, but exposure to polymerization mix buffer allowed reisolation of highly polymerized microtubules. Our results imply that prefixation with cacodylate or phosphate buffered glutaraldenyde destabilizes microtubules leading to the decreased visualization of microtubules.  相似文献   

18.
The effects of lead on the uptake and release of gamma-[3H]aminobutyric acid [( 3H]GABA) from rat brain slices were examined in solutions buffered with Tris-HCl, sodium phosphate, and sodium bicarbonate. Lead acetate (10-250 microM) inhibited uptake and potassium-stimulated release and facilitated spontaneous efflux only in solutions buffered with Tris-HCl. Calcium-independent binding of [3H]GABA was unaffected by lead acetate (1-100 microM) in Tris-citrate buffer but was significantly inhibited by 3 microM lead acetate in Tris-HCl solution. At the rat soleus neuromuscular junction, lead caused a dose-dependent reduction of end-plate potential amplitude at concentrations of 10-100 microM lead acetate in HEPES-buffered solution but had no effect at these concentrations in phosphate-buffered solution. Stability constants of lead complexes indicate that buffers containing carbonate and phosphate are unlikely to contain a significant concentration of Pb2+, as complexing by these anions would reduce the availability of free Pb2+. This study indicates that the choice of buffer is important when investigating the effects of lead on biological systems and that negative findings may result from the use of inappropriate buffers. It also has important clinical implications suggesting that some effects of lead poisoning may result from its ability to affect neurotransmitter systems directly and that local changes in pH and complexing anion concentrations in the CNS may influence its biological availability and, hence, variable biological responses.  相似文献   

19.
A rapid assay for activity of phospholipase A2 using radioactive substrate   总被引:1,自引:0,他引:1  
A rapid method for the assay of phospholipase A2 has been developed using a radioactive substrate, L-alpha-dipalmitoyl-(2-[9,10(N)-3H]palmitoyl)-phosphatidylcholine. The substrate diluted with cold carrier (1 mM) is dissolved in 80% ethanol containing 25 mM sodium deoxycholate. The enzymatic reaction is performed in 1.0 ml 0.1 M glycine-NaOH buffer, pH 9.0, containing 2 mumol CaCl2, 10 micrograms bovine serum albumin, 2.5 mumol sodium deoxycholate, 0.01 unit (or less) phospholipase A2, and 40-100 nmol substrate. The enzymatic reaction is terminated by adding 0.2 ml 5% Triton X-100 solution containing 40 mumol EDTA. The product of the enzymatic reaction, radioactive palmitic acid, is extracted by 10 ml hexane containing 0.1% acetic acid in the presence of anhydrous sodium sulfate (0.5 g/ml). Activity of phospholipase A2 is directly determined from the radioactivity in the hexane extract. The present method achieves a quick separation of the radioactive product, [3H]palmitic acid, from the radioactive substrate, L-alpha-dipalmitoyl-(2-[3H]palmitoyl)-phosphatidylcholine, without the need of separation by TLC.  相似文献   

20.
【背景】酚酸脱羧酶催化分解酚酸产生的4-乙烯基酚类物质可用于食品添加剂及香精香料行业,而酚酸脱羧酶的表达水平相对较低,因此,高水平的酚酸脱羧酶是工业规模生产4-乙烯基酚类物质的先决条件。【目的】克隆解淀粉芽胞杆菌的酚酸脱羧酶基因,实现在大肠杆菌中的高效异源表达,分析酚酸脱羧酶的底物特异性,并对其表达条件进行优化。【方法】通过PCR技术获得酚酸脱羧酶的基因,构建重组基因工程菌,将测序结果与其他酚酸脱羧酶序列进行比对,利用IPTG诱导方法高效表达蛋白。将重组酚酸脱羧酶与4种不同的底物进行反应,设计响应面试验对诱导条件进行优化。【结果】酚酸脱羧酶对对香豆酸、阿魏酸、咖啡酸、芥子酸的比酶活比率为:100:23.33:15.39:10.51。结合与其他酚酸脱羧酶比对结果发现酚酸脱羧酶家族的C末端区域氨基酸序列的变异率最高,这与酚酸脱羧酶的底物特异性和催化机制有关。通过单因素和响应面试验得到酚酸脱羧酶诱导表达的最佳条件为:2×YT培养基,诱导温度30°C,接种量1.78%,诱导时机3.8 h,IPTG1.25mmol/L,诱导时间18h,此时预测酶活和实际酶活分别为47.61IU/mL和47.55IU/mL。【结论】应用响应面法优化酚酸脱羧酶的诱导表达是可行的,本试验为以后生产稳定、高产的酚酸脱羧酶以及了解其催化机理提供了重要的理论基础。  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