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.
l-Histidine as pseudobiospecific ligand was immobilized onto poly(ethylene vinyl alcohol) hollow-fiber membranes to obtain an affinity support for immunoglobulin G (IgG) purification. The interaction of human IgG with the affinity membranes was studied by chromatography and equilibrium binding analysis. Adsorption was possible over a broad pH range and was found to depend strongly on the nature of the buffer ions rather than on ionic strength. With zwitterionic buffers like morpholinopropanesulfonic acid (Mops) and hydroxyethylpiperazineethanesulfonic acid (Hepes), much higher adsorption capacities were obtained than with other buffers like Tris-HCl and phosphate buffers. An inhibition analysis revealed that non-zwitterionic buffers competitively inhibit IgG binding, whereas Mops and Hepes in their zwitterionic form do not. By choosing the appropriate buffer system, it was possible to adsorb specifically different IgG subsets. The IgG molecules were found to adsorb on membrane immobilized histidine via their Fab part. Determination of dissociation constants at different temperatures allowed calculation of thermodynamic adsorption parameters. Decrease in KD with increasing temperature and a positive entropy value between 20 and 35°C (in Mops buffer) indicated that adsorption is partially governed by hydrophobic forces in that temperature range, whereas at lower temperatures, electrostatic forces are more important for adsorption.  相似文献   

2.
Hydrostatic pressure is a useful tool in the study of varied fields such as protein aggregation, association, folding, ligand binding, and allostery. Application of pressure can have a significant effect on the pK(a) values of buffers commonly used for biochemical analysis. Consequently, cationic buffers, rather than neutral ones, are generally used to minimize pH effects; however, even with these buffers, the change in pH over 3 kbar may be consequential in highly pH-sensitive biochemical systems. Using fluorescence-based assays, we have systematically examined the effects of pressure on various buffers in the neutral pH range. We show that many commonly used cationic and Good's buffers increase in pH with pressure on the order of 0.1 to 0.3 pH units/kbar, in agreement with other published values. Carboxylates and phosphate decrease in pH to a similar extent. Buffer mixtures, composed of both cationic and carboxylate or phosphate components, are shown to be an order of magnitude less pressure sensitive than the individual component buffers. Using various relative concentrations of Tris and either phosphate, tricarballylate (1,2,3-propanetricarboxylate), or CDA (1,1-cyclohexane diacetate) at pH values between 7 and 8 yields baroresistant buffer mixtures. Buffer mixtures can be optimized for a specific pH, and a list of mixtures is presented for general laboratory use.  相似文献   

3.
Direct microcalorimetric measurements were made of the reaction between acetylcholine chloride and acetylcholinesterase (EC 3.1.1.7) that was extracted from electric eel (Electrophorus electricus) and purified by affinity chromatography. Tris-HCl, sodium phosphate and potassium phosphate were used as buffers and sources of ions for the reaction. At pH 7.2 and in 0.1-0.2 M phosphate buffer, the delta H for acetylcholine hydrolysis was found to be -0.107 kcal/mol (under buffered conditions) and -0.931 kcal/mol under unbuffered conditions (water). At pH 8.0 in 0.1 M Tris-HCl buffer, values greater than -2.5 kcal/mol were obtained, with the highest value of -9.2 kcal/mol being seen with bovine erythrocyte acetylcholinesterase. Tris-HCl buffer at 4 X 10(-2) M enhanced the reaction velocity by 51.2% over that of 4 X 10(-3) M buffer. Enzyme purity, pH and ionic milieu of reaction mixture, and substrate concentration affected the measured delta H value.  相似文献   

4.
《Free radical research》2013,47(3):149-160
Iron autoxidation in Mops and Hepes buffers is characterized by a lag phase that becomes shorter with increasing FeCl2 concentration and pH. During iron oxidation in these buffers a yellow colour develops in the solution. When the reaction is conducted in the presence of nitro blue tetrazolium (NBT), blue formazan is formed. Of the many OH' scavengers tested, mannitol and sorbitol are most effective in inhibiting Fe2+ oxidation, yellow colour development and NBT reduction. Some inhibition was also noted with catalase. The iron product of the oxidative reaction differs from Fe3+ in its absorption spectrum and its low reactivity with thiocyanate. Similar results are obtained when iron autoxidation is studied in unbuffered solutions brought to alkaline pH with NaOH. In phosphate buffer, no lag phase is evident and the absorption spectrum of the final solution is identical to that of Fe3+ in this buffer. The iron product reacts immediately with thiocyanate. When iron oxidation is conducted in the presence of NBT the formation of formazan is almost undetectable. Of the many compounds tested only catalase inhibits iron autoxidation in this buffer. The sequence of reactions leading to iron autoxidation in Good-type buffers1 thus resembles that occurring in unbuffered solutions brought to alkaline pH with NaOH and greatly differs from that occurring in phosphate buffer. These results are in agreement with the observation that these buffers have very low affinity for iron.1 The data presented define experimental conditions where Fe2+ is substantially stable for a considerable length of time in Mops buffer.  相似文献   

5.
Data are presented regarding the establishment of the pH (designated pH*) of a standard buffer solution suitable as a pH reference in 50 mass% glycerol/water mixtures at temperatures ranging from -20 to 25 degrees C. The buffer material selected was the ampholyte Mops [(3-N-morpholino)-propane sulfonic acid], and the reference standard consists of equal molal amounts of Mops and its sodium salt. The assignment of pH* values is based on measurements of the electromotive force (emf) of cells without liquid junction of the type: Pt;H2(g, 1 atm) / Mops, Na Mopsate, NaCl / AgCl;Ag and the pH* was derived from a determination of K2, the equilibrium constant for the dissociation process (Mops) +/- in equilibrium with (Mopsate)- + H+. The standard emf of the silver-silver chloride electrode in 30, 40, and 50 mass% glycerol/water mixtures was determined from emf measurements of the cell at subzero temperatures with HCl solutions replacing the buffer-chloride mixtures.  相似文献   

6.
The catalytic activity of phosvitin in Fe(II) oxidation and the addition of iron to transferrin were studied under various conditions. It was concluded that the Fe(II) oxidized by phosvitin would bind to apotransferrin, although an appreciable fraction of Fe(III) remained bound to phosvitin. Fe(III) also migrated from phosvitin to apotransferrin. This reaction was first-order with respect to Fe(III)-phosvitin concentration with a half-time (t1/2) of 10 min, and a first-order rate constant, k=0.069min-1, in 700 muM-phosphate buffer, pH 7.2, at 30 degrees C. The catalysis of the oxidation of Fe(III) by phosvitin was proportional to O2 concentration, and is quite different from the relative O2 independence of Fe(II) oxidation as catalysed by ferroxidase. A scheme for the mobilization and transfer of iron in the chicken, including the role of ferroxidase, phosyitin and transferrin, is presented.  相似文献   

7.
The substrate specificity and the intraperoxisomal localization of alpha-hydroxyacid oxidase in rat liver has been investigated cytochemically by the cerium technique and biochemically with a luminometric assay. Rat liver is fixed by perfusion with a low concentration (0.25%) of glutaraldehyde and vibratome sections are incubated for 60 min at 37 degrees C in a medium containing 3 mM CeCl3, 100 mM NaN3 and 5 mM of an alpha-hydroxyacid in 0.1 M of one of the following buffers: Pipes, Mops, Na-cacodylate, Tris-maleate, all adjusted to pH 7.8. Ten different alpha-hydroxyacids with a chain length between 2 and 8 carbon atoms were tested. The best results were obtained with glycolic, argininic and L-alpha-isocaproic acids. These cytochemical findings were confirmed also biochemically using purified peroxisomal fractions isolated by gradient centrifugation in metrizamide. The pattern of the intraperoxisomal localization of the enzyme was influenced markedly by the type of buffer used for the cytochemical incubation. Whereas in the Tris-maleate medium both the cores and the matrix stained with the same intensity, with all other buffers the reaction in cores was more prominent. The staining of cores was abolished by pretreating sections in Tris-maleate (pH 7.8) or alkaline pyrophosphate buffers. These observations establish the substrate specificity of alpha-hydroxyacid oxidase in rat liver and demonstrate the delicate association of this enzyme with the crystalline cores and the matrix of peroxisomes in rat liver.  相似文献   

8.
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   

9.
The discrepancy between predicted and measured H(2)O(2) formation during iron deposition with recombinant heavy human liver ferritin (rHF) was attributed to reaction with the iron protein complex [Biochemistry 40 (2001) 10832-10838]. This proposal was examined by stopped-flow kinetic studies and analysis for H(2)O(2) production using (1) rHF, and Azotobacter vinelandii bacterial ferritin (AvBF), each containing 24 identical subunits with ferroxidase centers; (2) site-altered rHF mutants with functional and dysfunctional ferroxidase centers; and (3) recombinant human liver light ferritin (rLF), containing no ferroxidase center. For rHF, nearly identical pseudo-first-order rate constants of 0.18 s(-1) at pH 7.5 were measured for Fe(2+) oxidation by both O(2) and H(2)O(2), but for rLF, the rate with O(2) was 200-fold slower than that for H(2)O(2) (k = 0.22 s(-1)). A Fe(2+)/O(2) stoichiometry near 2.4 was measured for rHF and its site altered forms, suggesting formation of H(2)O(2). Direct measurements revealed no H(2)O(2) free in solution 0.5-10 min after all Fe(2+) was oxidized at pH 6.5 or 7.5. These results are consistent with initial H(2)O(2) formation, which rapidly reacts in a secondary reaction with unidentified solution components. Using measured rate constants for rHF, simulations showed that steady-state H(2)O(2) concentrations peaked at 14 muM at approximately 600 ms and decreased to zero at 10-30 s. rLF did not produce measurable H(2)O(2) but apparently conducted the secondary reaction with H(2)O(2). Fe(2+)/O(2) values of 4.0 were measured for AvBF. Stopped-flow measurements with AvBF showed that both H(2)O(2) and O(2) react at the same rate (k = 0.34 s(-1)), that is faster than the reactions with rHF. Simulations suggest that AvBF reduces O(2) directly to H(2)O without intermediate H(2)O(2) formation.  相似文献   

10.
Bacterioferritin (EcBFR) of Escherichia coli is an iron-mineralizing hemoprotein composed of 24 identical subunits, each containing a dinuclear metal-binding site known as the "ferroxidase center." The chemistry of Fe(II) binding and oxidation and Fe(III) hydrolysis using H(2)O(2) as oxidant was studied by electrode oximetry, pH-stat, UV-visible spectrophotometry, and electron paramagnetic resonance spin trapping experiments. Absorption spectroscopy data demonstrate the oxidation of two Fe(II) per H(2)O(2) at the ferroxidase center, thus avoiding hydroxyl radical production via Fenton chemistry. The oxidation reaction with H(2)O(2) corresponds to [Fe(II)(2)-P](Z) + H(2)O(2) --> [Fe(III)(2)O-P](Z) + H(2)O, where [Fe(II)(2)-P](Z) represents a diferrous ferroxidase center complex of the protein P with net charge Z and [Fe(III)(2)O-P](Z) a micro-oxo-bridged diferric ferroxidase complex. The mineralization reaction is given by 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2FeOOH((core)) + 4H(+), where two Fe(II) are again oxidized by one H(2)O(2). Hydrogen peroxide is shown to be an intermediate product of dioxygen reduction when O(2) is used as the oxidant in both the ferroxidation and mineralization reactions. Most of the H(2)O(2) produced from O(2) is rapidly consumed in a subsequent ferroxidase reaction with Fe(II) to produce H(2)O. EPR spin trapping experiments show that the presence of EcBFR greatly attenuates the production of hydroxyl radical during Fe(II) oxidation by H(2)O(2), consistent with the ability of the bacterioferritin to facilitate the pairwise oxidation of Fe(II) by H(2)O(2), thus avoiding odd electron reduction products of oxygen and therefore oxidative damage to the protein and cellular components through oxygen radical chemistry.  相似文献   

11.
Jameson GN  Jin W  Krebs C  Perreira AS  Tavares P  Liu X  Theil EC  Huynh BH 《Biochemistry》2002,41(45):13435-13443
The catalytic step that initiates formation of the ferric oxy-hydroxide mineral core in the central cavity of H-type ferritin involves rapid oxidation of ferrous ion by molecular oxygen (ferroxidase reaction) at a binuclear site (ferroxidase site) found in each of the 24 subunits. Previous investigators have shown that the first detectable reaction intermediate of the ferroxidase reaction is a diferric-peroxo intermediate, F(peroxo), formed within 25 ms, which then leads to the release of H(2)O(2) and formation of ferric mineral precursors. The stoichiometric relationship between F(peroxo), H(2)O(2), and ferric mineral precursors, crucial to defining the reaction pathway and mechanism, has now been determined. To this end, a horseradish peroxidase-catalyzed spectrophotometric method was used as an assay for H(2)O(2). By rapidly mixing apo M ferritin from frog, Fe(2+), and O(2) and allowing the reaction to proceed for 70 ms when F(peroxo) has reached its maximum accumulation, followed by spraying the reaction mixture into the H(2)O(2) assay solution, we were able to quantitatively determine the amount of H(2)O(2) produced during the decay of F(peroxo). The correlation between the amount of H(2)O(2) released with the amount of F(peroxo) accumulated at 70 ms determined by M?ssbauer spectroscopy showed that F(peroxo) decays into H(2)O(2) with a stoichiometry of 1 F(peroxo):H(2)O(2). When the decay of F(peroxo) was monitored by rapid freeze-quench M?ssbauer spectroscopy, multiple diferric mu-oxo/mu-hydroxo complexes and small polynuclear ferric clusters were found to form at rate constants identical to the decay rate of F(peroxo). This observed parallel formation of multiple products (H(2)O(2), diferric complexes, and small polynuclear clusters) from the decay of a single precursor (F(peroxo)) provides useful mechanistic insights into ferritin mineralization and demonstrates a flexible ferroxidase site.  相似文献   

12.
Steady-state kinetics of carboxypeptidase Y, a proteinase from yeast, were studied by using the reaction of 4-nitrophenyl trimethylacetate as a probe. The pH profile of kcat. is sigmoidal in H2O-based buffers for the carboxypeptidase Y-catalysed hydrolysis of this ester (kcat. referring to the rate of deacylation of trimethylacetyl-carboxypeptidase Y). The corresponding pD profile in 2H2O is doubly sigmoidal, with inflexions at pD approximately 3.8 and approximately 6.8. The ionization of pKDapp. approximately 3.8 is caused by a rapid inactivation in 2H2O media by a process that is only slowly reversed on transfer to pH 7.00 phosphate buffer in H2O. The corresponding inactivation in H2O-based buffers of low pH is considerably slower (approximately 30-fold), follows a first-order rate-dependence and is very strongly pH-dependent, indicating some form of co-operative change in enzyme tertiary structure.  相似文献   

13.
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   

14.
Bacterioferritins are members of a class of spherical shell-like iron storage proteins that catalyze the oxidation and hydrolysis of iron at specific sites inside the protein shell, resulting in formation of a mineral core of hydrated ferric oxide within the protein cavity. Electrode oximetry/pH stat was used to study iron oxidation and hydrolysis chemistry in E. coli bacterioferritin. Consistent with previous UV-visible absorbance measurements, three distinct kinetic phases were detected, and the stoichiometric equations corresponding to each have been determined. The rapid phase 1 reaction corresponds to pairwise binding of 2 Fe(2+) ions at a dinuclear site, called the ferroxidase site, located within each of the 24 subunits, viz., 2Fe(2+) + P(Z) --> [Fe(2)-P](Z) + 4H(+), where P(Z) is the apoprotein of net charge Z and [Fe(2)-P](Z) represents a diferrous ferroxidase complex. The slower phase 2 reaction corresponds to the oxidation of this complex by molecular oxygen according to the net equation: [Fe(2)-P](Z) + (1)/(2)O(2) --> [Fe(2)O-P](Z) where [Fe(2)O-P](Z) represents an oxidized diferric ferroxidase complex, probably a mu-oxo-bridged species as suggested by UV-visible and EPR spectrometric titration data. The third phase corresponds to mineral core formation according to the net reaction: 4Fe(2+) + O(2) + 6H(2)O --> 4FeO(OH)((core)) + 8H(+). Iron oxidation is inhibited by the presence of Zn(2+) ions. The patterns of phase 2 and phase 3 inhibition are different, though inhibition of both phases is complete at 48 Zn(2+)per 24mer, i.e., 2 Zn(2+) per ferroxidase center.  相似文献   

15.
The rate of the 4 to 5 S estrogen-binding protein (EBP) in vitro transformation was measured by sucrose gradient centrifugation analysis. The temperature-activated 4 to 5 S EBP transformation is found to be highly reproducible without loss of [3H]estradiol-binding activity in a buffer containing an excess of [3H]estradiol, 40 mM Tris, 1 mM dithiothreitol, and 1 M urea at pH 7.4. The presence of [3H]estradiol is necessary for the 4 to 5 EBP transformation. A kinetic analysis of the 4 to 5 EBP transformation shows that it is a bimolecular reaction, the dimerization of the 4 S EBP with a second (similar or dissimilar) monomer or subunit. In buffers containing 0.4 M KCl the apparent second order rate constant is 2.3 plus or minus 0-2 times 10-7 M minus 1 min minus 1 at 28 degrees. The reaction is independent of the initial receptor concentration, suggesting that the 4 S EBP is dissociated into monomeric units in buffers of high ionic strength. In buffers without KCl or with 0.1 M KCl the apparent second order rate constant of receptor transformation increases with decreasing receptor concentration. This suggests that the 4 S EBP is associated weakly with another macromolecule (or macromolecules) in buffers of low ionic strength. The rate of 4 to 5 S EBP transformation shows a 200-fold increase between 0 and 35 degrees. The Arrhenius energy of activation is 21.3 kcal mol minus 1 in buffer without KCl and 19.1 kcal mol minus 1 in buffer with 0.4 M KCl. Following the temperature-activated dimerization, the avidity of binding between the 4 S EBP and its complementary subunit is increased, 0.4 M KCl can no longer cause dissociation, and the 5 S EBP dimer appears. This kinetic analysis indicates that the avidity of binding between the subunits of the estrogen receptor is modulated by estradiol, temperature, and ionic strength. We propose that these interactions of the estrogen receptor's subunits reflect conformational changes involved in receptor activation.  相似文献   

16.
The production of formaldehyde from tris(hydroxymethyl) aminomethane(Tris) by interaction with hydroxyl radicals(.OH) was studied, since the reaction mixture from the Fenton reaction performed in Tris/HCl buffer was found to be color-developed by colorimetric determination of formaldehyde. The absorption spectrum of chromogens was identical to that of authentic formaldehyde. Color development, which required the presence of Tris, hydrogen peroxide and cupric ions in the Fenton reaction mixture, was inhibited by the addition of hydroxyl radical scavengers such as glucose or hyaluronic acid. These results indicated that formaldehyde was produced when Tris interacted with ·OH. With structures similar to Tris, Good's buffers were also found to produce formaldehyde by interaction with ·OH. Analysis of formaldehyde derived from these buffers may provide a simple and convenient assay for detecting ·OH generation. In evaluating effects of ·OH on the biological system in Tris/HCl buffer or certain Good's Buffers, ·OH loss may be due to interactions of ·OH with these buffers. The formaldehyde produced as a result of such interactions may affect biological systems.  相似文献   

17.
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   

18.
The rate of ferritin formation in the buffers 4-morpholinepropanesulphonic acid (Mops), 4-morpholineethanesulphonic acid (Mes) and imidazole at pH values from 5.0 to 6.5 is quite similar. However, the rate of iron deposition is much greater in Mops and Mes at pH values above 6.5 than in imidazole. Increasing the concentration of imidazole inhibits ferritin formation and also leads to a transformation in the shape of the kinetic curves observed. This inhibiton is also observed at constant ionic strength but is not found for non-complexing buffers such as Mops. An inhibition of ferritin formation in imidazole and in Mops buffers is also observed with increasing ionic strength. We conclude that the unprotonated form of imidazole inhibits iron deposition, possibly by binding to the active site of the apoferritin molecule. The temperature dependence of iron deposition was examined. An optimum temperature of 50 degrees C was found but the Arrhenius plots were non-linear. On the basis of these and previous results, a kinetic model is developed which accounts well for ferritin formation at pH values below 6.5 and above 7.0 in non-complexing buffers. The model does not account for the kinetics observed at pH values close to neutrality.  相似文献   

19.
Hydroxyl radicals, generated by reaction of an iron-EDTA complex with H2O2 in the presence of ascorbic acid, attack deoxyribose to form products that, upon heating with thiobarbituric acid at low pH, yield a pink chromogen. Added hydroxyl radical "scavengers" compete with deoxyribose for the hydroxyl radicals produced and diminish chromogen formation. A rate constant for reaction of the scavenger with hydroxyl radical can be deduced from the inhibition of color formation. For a wide range of compounds, rate constants obtained in this way are similar to those determined by pulse radiolysis. It is suggested that the deoxyribose assay is a simple and cheap alternative to pulse radiolysis for determination of rate constants for reaction of most biological molecules with hydroxyl radicals. Rate constants for reactions of ATP, ADP, and Good's buffers with hydroxyl radicals have been determined by this method.  相似文献   

20.
Yamazaki S  Morioka C  Itoh S 《Biochemistry》2004,43(36):11546-11553
Tyrosinase is a copper monooxygenase containing a coupled dinuclear copper active site (type-3 copper), which catalyzes oxygenation of phenols (phenolase activity) as well as dehydrogenation of catechols (catecholase activity) using O(2) as the oxidant. In this study, catalase activity (conversion of H(2)O(2) to (1/2)O(2) and H(2)O) and peroxygenase activity (H(2)O(2)-dependent oxygenation of substrates) of mushroom tyrosinase have been examined kinetically by using amperometric O(2) and H(2)O(2) sensors. The catalase activity has been examined by monitoring the initial rate of O(2) production from H(2)O(2) in the presence of a catalytic amount of tyrosinase in 0.1 M phosphate buffer (pH 7.0) at 25 degrees C under initially anaerobic conditions. It has been found that the catalase activity of mushroom tyrosinase is three-order of magnitude greater than that of mollusk hemocyanin. The higher catalase activity of tyrosinase could be attributed to easier accessibility of H(2)O(2) to the dinuclear copper site of tyrosinase. Mushroom tyrosinase has also been demonstrated for the first time to catalyze oxygenation reaction of phenols with H(2)O(2) (peroxygenase activity). The reaction has been investigated kinetically by monitoring the H(2)O(2) consumption rate in 0.5 M borate buffer (pH 7.0) under aerobic conditions. Similarity of the substituent effects of a series of p-substituted phenols in the peroxygenase reaction with H(2)O(2) to those in the phenolase reaction with O(2) as well as the absence of kinetic deuterium isotope effect with a perdeuterated substrate (p-Cl-C(6)D(4)OH vs p-Cl-C(6)H(4)OH) clearly demonstrated that the oxygenation mechanisms of phenols in both systems are the same, that is, the electrophilic aromatic substitution reaction by a (micro-eta(2):eta(2)-peroxo)dicopper(II) intermediate of oxy-tyrosinase.  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