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.
The combined potentiometric and spectroscopic studies of interactions of N-2-hydroxyethylpiperazine-N'-2-ethanesulfonic acid (HEPES) with Cu(II) demonstrated that this popular buffer, commonly labelled as "non-coordinating" forms a CuL+ complex, with the logbeta(CuL) value of 3.22. This complex undergoes alkaline hydrolysis above pH 6, resulting in Cu(OH)2 precipitation. However, the presence of HEPES at a typical concentration of 100 mM at pH 7.4 elevates the apparent binding constant, being determined for a complex of another ligand, by a factor of 80. HEPES does not form ternary complexes with aminoacids Ala, Trp, and His, but may do so with other bioligands, such as nucleotides. Therefore, HEPES can still be recommended for Cu(II) studies in place of other common buffers, such as Tris and phosphate, but appropriate corrections and precautions should be applied in quantitative experiments.  相似文献   

2.
J R Schullek  I B Wilson 《Peptides》1989,10(2):431-434
Phosphate, borate, and Tris inhibit angiotensin converting enzyme (ACE), but HEPES buffer is inert. Measurements of substrate inhibition were made in HEPES buffer at pH 7.0 and 25 degrees C and 37 degrees C. Substrate inhibition was marked and goes to completion. A new equation for substrate inhibitions enables one, under favorable circumstances, to determine whether there is cooperativity in the binding of substrate to the inhibitory and active sites. Cooperativity does occur with ACE using Hipp-His-Leu as substrate. The kinetic parameters were measured (Km = 0.21 mM, K* = 0.65 mM at 37 degrees C). The enzyme concentration (1.94 X 10(-8) M) was determined by titration with lisinopril so that kcat (5 X 10(3) at 37 degrees C) could be determined. Using this value and the molecular weight the specific activity of ACE was calculated for different common buffers. The specific activity in HEPES calculated from Vmax was 33.7 units/mg at 37 degrees C.  相似文献   

3.
The amine buffer Tris slowly reduces tetranitromethane (TNM) to the nitroform anion in a non-accelerating reaction. The amine buffers HEPES and MOPS also (slowly) react with TNM but the dialkylaminoalkyl radicals formed from these two buffers undergo further reactions resulting in a rapid, accelerating, free radical chain process whereby the amine is oxidized and TNM reduced. The chemical functionality in any reaction component, not necessarily the buffer, required for this radical chain mechanism is >N-CH<. In the presence of such groups, the quantification of superoxide by TNM is impossible.  相似文献   

4.
The amine buffer Tris slowly reduces tetranitromethane (TNM) to the nitroform anion in a non-accelerating reaction. The amine buffers HEPES and MOPS also (slowly) react with TNM but the dialkylaminoalkyl radicals formed from these two buffers undergo further reactions resulting in a rapid, accelerating, free radical chain process whereby the amine is oxidized and TNM reduced. The chemical functionality in any reaction component, not necessarily the buffer, required for this radical chain mechanism is >N-CH<. In the presence of such groups, the quantification of superoxide by TNM is impossible.  相似文献   

5.
We investigated the production of hydrogen peroxide (HOOH) in illuminated seawater media containing a variety of zwitterionic buffers. Production rates varied extensively among buffers, with 4‐(2‐hydroxyethyl)1‐piperazineethanesulfonic acid (HEPES) highest and N‐Tris(hydroxymethyl)methyl‐3‐aminopropanesulfonic acid (TAPS) among the lowest. The rate of HOOH accumulation was remarkably consistent over many days, and increased linearly with buffer concentration, natural seawater concentration, and light level. Concentrations of HEPES commonly used in culture media (1–10 mM) generated enough HOOH to kill the axenic Prochlorococcus strain VOL1 during growth in enriched seawater media at lower, environmentally realistic cell concentrations and/or under high light exposure. We also demonstrated that HEPES can be used experimentally to study the biological effects of chronic exposure to sublethal levels of HOOH such as may be experienced by light‐exposed microorganisms.  相似文献   

6.
Fatty acid binding and oxidation kinetics for wild type P450BM3 (CYP102A1) from Bacillus megaterium have been found to display chain length-dependent homotropic behavior. Laurate and 13-methyl-myristate display Michaelis-Menten behavior while there are slight deviations with myristate at low ionic strengths. Palmitate shows Michaelis-Menten kinetics and hyperbolic binding behavior in 100 mmol/L phosphate, pH 7.4, but sigmoidal kinetics (with an apparent intercept) in low ionic strength buffers and at physiological phosphate concentrations. In low ionic strength buffers both the heme domain and the full-length enzyme show complex palmitate binding behavior that indicates a minimum of four fatty acid binding sites, with high cooperativity for the binding of the fourth palmitate molecule, and the full-length enzyme showing tighter palmitate binding than the heme domain. The first flavin-to-heme electron transfer is faster for laurate, myristate and palmitate in 100 mmol/L phosphate than in 50 mmol/L Tris (pH 7.4), yet each substrate induces similar high-spin heme content. For palmitate in low phosphate buffer concentrations, the rate constant of the first electron transfer is much larger than kcat. The results suggest that phosphate has a specific effect in promoting the first electron transfer step, and that P450BM3 could modulate Bacillus membrane morphology and fluidity via palmitate oxidation in response to the external phosphate concentration.  相似文献   

7.
Survival rates of Escherichia coli and Staphylococcus aureus after high-pressure treatment in buffers that had large or small reaction volumes (ΔV°), and which therefore underwent large or small changes in pH under pressure, were compared. At a low buffer concentration of 0.005 M, survival was, as expected, better in MOPS (morpholinepropanesulfonic acid), HEPES, and Tris, whose ΔV° values are approximately 5.0 to 7.0 cm3 mol−1, than in phosphate or dimethyl glutarate (DMG), whose ΔV° values are about −25 cm3 mol−1. However, at a concentration of 0.1 M, survival was unexpectedly better in phosphate and DMG than in MOPS, HEPES, or Tris. This was because the baroprotective effect of phosphate and DMG increased much more rapidly with increasing concentration than it did with MOPS, HEPES, or Tris. Further comparisons of survival in solutions of salts expected to cause large electrostriction effects (Na2SO4 and CaCl2) and those causing lower electrostriction (NaCl and KCl) were made. The salts with divalent ions were protective at much lower concentrations than salts with monovalent ions. Buffers and salts both protected against transient membrane disruption in E. coli, but the molar concentrations necessary for membrane protection were much lower for phosphate and Na2SO4 than for HEPES and NaCl. Possible protective mechanisms discussed include effects of electrolytes on water compressibility and kosmotropic and specific ion effects. The results of this systematic study will be of considerable practical significance in studies of pressure inactivation of microbes under defined conditions but also raise important fundamental questions regarding the mechanisms of baroprotection by ionic solutes.  相似文献   

8.
Electrophoresis through agarose and polyacrylamide-type gels is the standard method to separate, identify, and purify nucleic acids. Properties of electrophoresis buffers such as pH, ionic strength, and composition affect performance. The buffers in use contain a weak acid or weak base buffered by a compound with a dissimilar pK. Herein, three pK-matched buffers were developed, each containing two effective buffering components: one weak base and one weak acid which have similar pKa at 25 degrees C (within 0.3 pK units): (i) Ethanolamine/Capso, pH 9.6; (ii) triethanolamine/Tricine, pH 7.9; and (iii) Bis-Tris/Aces, pH 6.7. On agarose gels, the buffers in various concentrations were tested for separation of double-stranded DNA fragments with various DNA markers, agarose gel concentrations, and field strengths. Mobility was inversely proportional to the logarithm of molecular weight. The buffers provided high resolution without smearing at more dilute concentration than is possible with standard TAE (Tris/acetate, pH 8.0) or TBE (Tris/borate, pH 8.3) buffers. The buffers were also tested in 7 M urea denaturing LongRanger sequencing gels and in nondenaturing polyacrylamide SSCP gels. The pK-matched buffers provide good separation and high resolution, at a broad range of potential pH values. In comparison to TAE and TBE, pK-matched buffers provide higher voltage and current stability, lower working concentration, more concentrated stock solutions (up to 200x), and lower current per unit voltage, resulting in less heat generation.  相似文献   

9.
A new type of hyaluronidase was isolated from squid cranial cartilage. The enzyme seems to be localised extracellularly, since it is extracted from the tissue by 0.5 M sodium acetate, pH 7.0, in the presence of proteinase inhibitors. Degradation studies suggest that the enzyme belongs to the family of endoglycosidases generating oligosaccharides of rather large size. The best activity of the enzyme was observed at pH 7.0 and 37 degrees C and the optimum buffer for digestion was 0.15 M Tris acetate. It is inactive in sodium phosphate, morpholine acetate and HEPES buffers. The enzyme degrades aggrecan, hyaluronan, chondroitin sulphate and oversulphated chondroitin sulphate.  相似文献   

10.
The introduction of antigen retrieval (AR) techniques has dramatically improved the sensitivity of immunohistochemical detection of various antigens in formalin-fixed, paraffin-embedded tissues. The microwave-heating and pressure-cooking procedures are the most effective AR methods reported to date. Although extensive efforts have been made to optimize AR procedures using these two methods, previous studies have not led to a standard protocol applicable to all antibodies derived from different clones. In this study we have investigated the optimal AR buffer conditions for 29 antibodies that are in common use for diagnostic purposes in hospitals worldwide. Borate (pH 8.0) and Tris buffer (pH 9.5) yielded the highest retrieved antigen immunoreactivity against most antibodies as compared to other buffers tested. In addition, the microwave pressure-cooking gave better results than microwave-heating alone. Therefore, borate (pH 8.0) or Tris (pH 9.5) buffer used in conjunction with the pressure-cooking procedure is strongly recommended for standard routine use.  相似文献   

11.
The surface fouling of food processing equipment by proteins was studied by investigating the adsorption of egg white proteins to the surface of stainless steel (SS) at pH 7.4 and 30 °C, and particularly the effects of different types of ionic substances. Ovalbumin and ovomucoid, acidic egg white proteins, were less adsorbed in the presence of phosphate (P(i)), a multivalent anion, than in the presence of HEPES, an amphoteric ion. On the other hand, lysozyme, a basic egg white protein, was more adsorbed in the presence of P(i) than in the presence of HEPES. Citrate as another multivalent anion and taurine as another amphoteric ion affected the respective adsorption of those egg white proteins similarly to P(i) and HEPES. The adsorption of an egg white protein to an SS surface therefore depended on the combination of the type of protein and the effective charge of the coexisting ionic substance. This behaviour can be well explained by assuming that a small ionic substance precedes a protein in attaching to an SS surface, resulting in an alteration to the effective surface charge. Pretreating SS with a P(i) buffer lowered the amount of ovalbumin adsorbed with the HEPES buffer, demonstrating that P(i) can attach to and remain on the SS surface to affect the subsequent protein adsorption.  相似文献   

12.
The effects of various chemical agents on longevity of the cercariae of Schistosoma mansoni Sambon, 1907, were investigated. The median lifetime of cercariae maintained in dechlorinated tap water (DTW) was 10.5 hr. Increasing concentrations of sodium chloride added to distilled water increased the median lifetime to an optimum of 10 hr at 0.01 M; higher salt levels decreased longevity. At this optimum sodium chloride level, concentrations of glucose between 0.003 M and 0.03 M enhanced median survival to 13–14 hr. Using Tris-HCl buffers the effects of pH and ionic strength were examined. Cercarial longevity increased from 3.5 to 25 hr as pH increased from 6.4 to 9.0, and 0.01 M Tris was superior to 0.001 M Tris at a given pH. The greatest median lifetime (26 hr) was obtained in 0.01 M Tris, pH 9.0. Infectivity of cercariae in Tris at this optimal pH was compared to that in DTW. Maintenance in DTW resulted in greater worm burdens in mice than did treatment with Tris. This suggested that factors which affect cercarial longevity may not influence infectivity in the same manner. The effects of rotenone, Antimycin A, dinitrophenol, and potassium cyanide on cercarial viability suggested the existence of a functioning terminal electron transport chain similar to that of mammalian systems.  相似文献   

13.
Bakers yeast enolase A binds 2 moles of magnesium with a total enthalpy of +11,000 ± 1,100 cal/mol of protein at 25 °C in 0.05 ionic strength Tris buffer, pH 7.8. Measurements of the pH of unbuffered solutions of enolase indicate that at 0.05 ionic strength, 2 moles of protons are released per 2 moles of metal bound. The binding of magnesium to yeast enolase is consequently produced by a favorable entropy change. The enthalpies of binding observed in Tris buffer appear to be different at 0–1 and 1–2 moles of metal per mole of protein, suggesting a difference in binding sites.  相似文献   

14.
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   

15.
To obtain the label enzyme for enzyme-linked-immunoabsorbent-assay of two components each time in one well with conventional microplate readers, molecular engineering of Pseudomonas aeruginosa arylsulfatase (PAAS) is needed. To compare thermostability of PAAS/mutants of limited purity, effects of buffers on the half-activity time (t 0.5) at 37 °C were tested. At pH 7.4, PAAS showed non-exponential decreases of activity, with the apparent t 0.5 of ~6.0 days in 50 mM HEPES, but ~42 days in 10 mM sodium borate with >85 % activity after 15 days; protein concentrations in both buffers decreased at slower rates after there were significant decreases of activities. Additionally, the apparent t 0.5 of PAAS was ~14 days in 50 mM Tris–HCl, and ~21 days in 10 mM sodium phosphate. By sodium dodecyl-polyacrylamide gel electrophoresis, the purified PAAS gave single polypeptide; after storage for 14 days at 37 °C, there were many soluble and insoluble fragmented polypeptides in the HEPES buffer, but just one principal insoluble while negligible soluble fragmented polypeptides in the borate buffer. Of tested mutants in the neutral borate buffer, rates for activity decreases and polypeptide degradation were slower than in the HEPES buffer. Hence, dilute neutral borate buffers were favorable for examining thermostability of PAAS/mutants.  相似文献   

16.
Effect of Mg(2+), Ca(2+), Ni(2+) and Cd(2+) ions on parameters of DNA helix-coil transition in sodium cacodylate (pH 6.5), Tris (pH 8.5) and sodium tetraborate (pH 9.0) buffers have been studied by differential UV-visible spectroscopy and by thermal denaturation. Anomalous behavior of the melting temperature T(m) and the melting interval ΔT in the presence of MgCl(2) was observed in Tris, but not in cacodylate or tetraborate buffers. Changes in the buffer type and pH did not influence T(m) and ΔT dependence on Ca(2+) and Cd(2+) concentrations. Decrease of the T(m) and ΔT of DNA in the presence of Ni(2+) and Cd(2+) was caused by preferential ion interaction with N7 of guanine. This type of interaction was also found for Mg(2+) in Tris buffer. The anomalous decrease in the T(m) and ΔT values was connected to formation of complexes between metal ions and Tris molecules. Transition of DNA single-stranded regions into a compact form with the effective radius of the particles of 300±100 ? was induced by Mg(2+) ions in Tris buffer.  相似文献   

17.
Gap junction channels regulate cell-cell communication by passing metabolites, ions, and signaling molecules. Gap junction channel closure in cells by acidification is well documented; however, it is unknown whether acidification affects connexins or modulating proteins or compounds that in turn act on connexins. Protonated aminosulfonates directly inhibit connexin channel activity in an isoform-specific manner as shown in previously published studies. High-resolution atomic force microscopy of force-dissected connexin26 gap junctions revealed that in HEPES buffer, the pore was closed at pH < 6.5 and opened reversibly by increasing the pH to 7.6. This pH effect was not observed in non-aminosulfonate buffers. Increasing the protonated HEPES concentration did not close the pore, indicating that a saturation of the binding sites occurs at 10 mM HEPES. Analysis of the extracellular surface topographs reveals that the pore diameter increases gradually with pH. The outer connexon diameter remains unchanged, and there is a approximately 6.5 degrees rotation in connexon lobes. These observations suggest that the underlying mechanism closing the pore is different from an observed Ca2+-induced closure.  相似文献   

18.
The separation of proteins on stationary phases consisting of a bound organic chelator and a chelated divalent transition metal has been studied as a function of (A) metal ion species; (B) mobile phase composition and pH; and (C) anion and cation concentration. Optimum separation was observed at alkaline pH on chelated nickel stationary phases. Ammonium and Tris salts reduced the affinity of the metal chelate packing for serum proteins. Halide ions caused the proteins to be more strongly bound to the stationary phase. High salt concentrations had only a small effect on the binding of serum proteins in the absence of amine containing buffers or salts. It was also observed that the ease of elution and the recovery of protein were dependent on pH and upon the presence of halides. The general order of elution of serum proteins, based on isoelectric focusing, was independent of metal ion species and elution conditions, suggesting that a single mechanism or a unique sequence of mechanisms was operative. The results suggest that ligand exchange is the major mechanism of separation under basic conditions and that hydrophobic effects are the result of the competition of nonnitrogen ions with ammonium ions or amines for ligand binding sites modifying or participating in protein binding. Protein binding studies under weak acidic conditions are also presented although the mechanism responsible for protein binding is unclear.  相似文献   

19.
The purpose of the present study was to quantify the binding affinity between Tamm-Horsfall protein (THP) and complement 1q (C1q) using ELISA and a resonant mirror biosensor. In ELISA, immobilized THP was incubated with soluble C1q under both low and physiological ionic-strength conditions. Tamm-Horsfall protein bound C1q with an equilibrium dissociation constant (KD) of 1.9 +/- 0.6 nmol/L in low ionic-strength Tris buffers (20 mmol/L NaCl, pH 7.5) and with a lower affinity (KD of 13.4 +/- 4.7 nmol/L) in physiological-strength Tris buffers (154 mmol/L NaCl, pH 7.5). A resonant mirror biosensor, which monitors binding events in real-time, was used to quantify the KD of this reaction, as well as to estimate the kinetic parameters. In these studies, THP and C1q bound with an association rate constant, kass, of 1.25 x 105 L/mol per s and a dissociation rate constant, kdiss, of 0.002-0.005/s. The calculated KD for the THP/C1q binding in low ionic-strength buffers was higher (averages of 10-15 nmol/L) than that obtained by the ELISA, while physiological ionic-strength buffers still reduced the affinity of this binding by an order of magnitude. In conclusion, THP consistently bound C1q with high affinity using several techniques. At least a portion of this interaction involved electrostatic events, as demonstrated by the influence of ionic strength on the binding affinity.  相似文献   

20.
Shewanella oneidensis MR-1 possesses a periplasmic [NiFe]-hydrogenase (MR-1 [NiFe]-H(2)ase) that has been implicated in H(2) production and oxidation as well as technetium [Tc(VII)] reduction. To characterize the roles of MR-1 [NiFe]-H(2)ase in these proposed reactions, the genes encoding both subunits of MR-1 [NiFe]-H(2)ase were cloned and then expressed in an MR-1 mutant without hyaB and hydA genes. Expression of recombinant MR-1 [NiFe]-H(2)ase in trans restored the mutant's ability to produce H(2) at 37% of that for the wild type. Following purification, MR-1 [NiFe]-H(2)ase coupled H(2) oxidation to reduction of Tc(VII)O(4)(-) and methyl viologen. Change of the buffers used affected MR-1 [NiFe]-H(2)ase-mediated reduction of Tc(VII)O(4)(-) but not methyl viologen. Under the conditions tested, all Tc(VII)O(4)(-) used was reduced in Tris buffer, while in HEPES buffer, only 20% of Tc(VII)O(4)(-) was reduced. The reduced products were soluble in Tris buffer but insoluble in HEPES buffer. Transmission electron microscopy analysis revealed that Tc precipitates reduced in HEPES buffer were aggregates of crystallites with diameters of ~5 nm. Measurements with X-ray absorption near-edge spectroscopy revealed that the reduction products were a mixture of Tc(IV) and Tc(V) in Tris buffer but only Tc(IV) in HEPES buffer. Measurements with extended X-ray adsorption fine structure showed that while the Tc bonding environment in Tris buffer could not be determined, the Tc(IV) product in HEPES buffer was very similar to Tc(IV)O(2)·nH(2)O, which was also the product of Tc(VII)O(4)(-) reduction by MR-1 cells. These results shows for the first time that MR-1 [NiFe]-H(2)ase catalyzes Tc(VII)O(4)(-) reduction directly by coupling to H(2) oxidation.  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