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 interaction of the cupric ion with phytic acid as a function of pH has been studied by potentiometric and thermal titration and by the determination of ligand binding. As has been found for the reaction of zinc and calcium cations with phytate, the presence of the Cu(II) ion results in a displacement of the titration curves to more acid values. Evaluation of the parameters that describe such changes in ionization behavior by curve-fit analysis showed that as the Cu(II):phytate mol ratio was increased from one to eight, the pK' values of the ionizable group sets of phytic acid (ranging from 1.59 to 9.79) were consolidated into just two sets with curve-fit (CP) values ca. 1.5 and 3.7. Marked pH hysteresis effects are seen in such systems because of the pronounced acid strength of the Cu(II):aqua ion and the Cu(II) ligand aqua ion complex. The combined heat of binding and precipitation (plus solvation changes, etc.) of Cu(II) to phytate is endothermic (21.8–22.2 kcal mol−1). This is similar in magnitude to that reported for the binding of either Zn(II) or Ca(II) to phytate. In the titration of Cu(NO3)2 with KOH, presumably to form Cu(OH)2, ΔH° was exothermic (−12.5 kcal mol−1). From measurements of free Cu(II) cation concentration in the presence of phytate the binding reaction was found to be stoichiometric with 6 mols Cu(II) bound at pH 6. Binding occurs within the pH range 2–6. An apparent necessary requirement for binding is the availability of the oxo dianion structure formed from the second dissociation step of a phosphoryl group. Curve-fit analysis of the binding data as a function of pH showed that a group or group set with CP value ca. 4 governs the binding reaction(s) at all mol ratios of Cu(II) to phytate examined. It is suggested that the binding of cupric ions to phytate may occur to the equatorial rather than the axial configuration as suggested for Ca(II) binding. A space-filling molecular model to illustrate this has been constructed. Soluble Cu(II):phytate complexes are formed within the pH range from 2 to ca. 3.4. This is supported by the results of difference absorption spectrometry.  相似文献   

2.
Using an isoperibolic titration microcalorimeter, the ionization characteristics and associated heat changes of phytic acid (myo-inositol hexaphosphate) and phytic acid in the presence of varying Zn(II) concentrations have been examined over the pH range 2.5–11 at 25°C in 0.2 M KCl. In the absence of Zn(II), ca. 7 of the 12 ionizable protons in phytic acid are titrated in this pH range with ionization heats varying from ca. 2 to −3 kcal-mol-1. At Zn(II): phytate mol ratios of 4:1 and greater, the dissociation of all protons and complex formation of phytate with Zn(II) occurs below pH 6. From the difference titration curves of phytic acid plus Zn(II) versus Zn(II) alone, ca. 3.5 mol Zn(II) bind per mol phytate. Since Zn(II):phytate complexes are insoluble, the observed heat changes contain contributions not only from heats of precipitation but also from binding, ionization, neutralization, and hydration effects. From the heat change for the titration of (a) phytic acid, pH 2.6–10.4; (b) phytic acid + Zn(II), pH 2.6–6.1; and (c) Zn(II), pH 2.6–6.1 at Zn(II): phytate ratios of 4 to 10, the value of 24.7 ± 0.5 kcal mol−1 phytate has been obtained for the binding of 3.5 mols Zn(II). This figure also includes the heat of precipitation of the complex. In pH-drop experiments, with the initial pH at 8.65, the value of 23.9 kcal mol-1 was obtained for ΔH°. Hysteresis effects are prevalent in these reaction solutions. Time-dependent changes in pH occur with a change in pH. For the phytate-Zn(II) reactions, the time-course curves are biphasic and fit a rate equation for two simultaneous first order reactions. Hysteresis effects seen in the titration of Zn(II) fit simple first-order kinetics. These effects most probably arise from the ejection of a proton from the aqua ion or aqua ion ligand complex(es).  相似文献   

3.
In vitro interactions among phytic acid (PA), Cu(II) ions, and Ca(II) ions were examined as functions of PA:Cu(II):Ca(II) molar ratios and pH. Ca(II) ions competed with Cu(II) ions for binding by the soluble phytate species for PA:Cu(II) molar ratios ranging from 10:1 to 1:6 and pH values in the 2.4-5.9 range. At pH values where precipitation occurred, Ca(II) ions potentiated Cu(II) ion binding by the precipitated phytate species for PA:Cu(II) molar ratios of 10:1 to 1:3. At lower PA:Cu(II) molar ratios, Ca(II) ions competed with Cu(II) ions for binding by the precipitated phytate species. Compositions of the precipitated copper-calcium phytates are reported.  相似文献   

4.
The interactions of phytic acid with Cu(II) and Zn(II) ions were examined as functions of metal ion concentrations and pH. Cu(II) ion-selective potentiometric and electron spin resonance (ESR) experiments provide strong evidence for the binding of Cu(II) ions to the phytic acid molecule at low pH (2.4–3.4) values. The relative stabilities of the copper and zinc phytates at low pH values were found to be very similar. For systems with metal ion:phytic acid molar ratios of 1:1–4:1 and 5:1–6:1 and pH values in the 3.4–5.9 and 3.4–5.0 ranges, respectively, Zn(II) ions were found to form complexes with phytic acid that were more stable than those of Cu(II) ions with phytic acid. The phytic acid molecule, however, was found to accommodate Cu(II) ions more readily than Zn(II) ions. For example, in systems containing equal amounts of Cu(II) and Zn(II) ions, 2 Zn(II) ions and 2, 3, 4, or 4.5 Cu(II) ions were found per phytic acid molecule depending upon metal ion:phytic acid molar ratios in the systems and pH. Total metal ion:phytic acid molar ratios and pH affected resultant metal ion solubilities and were factors influencing the effects of Zn(II) and Cu(II) ions on the binding of each other by phytic acid. Zn(II) and Cu(II) ions were observed to potentiate the binding of each other by phytic acid in some systems and compete with each other for phytate binding sites in others.  相似文献   

5.
6.
On the basis of the known interaction of phytic acid to form soluble or insoluble complexes with cations, the effect of this naturally occurring polydentate ligand on carboxypeptidase A, a zinc-containing metalloenzyme, and its Co(II)-substituted derivative, has been studied. Under conditions of rigorous exclusion of adventitious metal ions, phytate showed no inhibitory effect. However, the addition of Cu(II) ions to form soluble phytate-Cu(II) complexes at pH 7.2 and 25 degrees C caused more than a 95% decrease in activity. The Cd(II) ion was nearly as effective but other ions showed only a small or no effect. In the absence of phytate, incubation of the enzyme with Cu(II) or Cd(II) at the same concentration produced only about a 25% reduction in activity. The decrease in activity followed first-order kinetics, and the rate constant was the same (1.2 x 10(-4) sec-1) as seen upon incubation with EDTA. However, in contrast to that observed upon incubation of the enzyme with phytate and Cu(II), exposure to EDTA produced a complete loss in activity which could be regained by addition of Zn(II) to the assay solution. In the former case, not only was there residual activity left after incubation at pH 7.2 for 24 hrs at 25 degrees C, but the initial activity could not be regained under similar assay treatment. An increase in either the Cu(II) or phytate concentration while the other was kept constant, yielded saturation curves with maximal effect at 3 x 10(-5) M for Cu(II) and at 5 x 10(-5) M for phytate (enzyme at ca. 10(-6) M). At these ratios, all of the cupric ions are completely bound to phytate as determined by ion-selective potentiometry. A preparative scale reaction of phytate and Cu(II) with carboxypeptidase A (kcat 8460 min-1; K'M 0.23 mM with CBZ-glycyl-glycyl-L-phenylalanine as substrate at pH 7.5, 25 degrees C) gave a product isolated in 95% yield but with lower activity (kcat 198 min-1; K'M 0.25 mM). A Cu(II)-carboxypeptidase preparation had similar kinetic parameters (kcat 207 min-1; K'M 0.34 mM). This near identity of constants suggested that a metal exchange reaction had occurred, i.e., incubation of Zn(II)-carboxypeptidase with a phytate-Cu(II) complex resulted in not only the removal of the zinc ion from the active site but also the sequential and rapid incorporation of a cupric ion into the apoenzyme so formed.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
CopC is a periplasmic copper carrier that, in contrast to cytoplasmic copper chaperones, has a beta-barrel fold and two metal-binding sites distinct for Cu(II) and Cu(I). The copper sites are located in each end of the molecule: the Cu(I) site involves His and Met coordination whereas the Cu(II) site consists of charged residues. To reveal biophysical properties of this protein, we have explored the effects of the cofactors on CopC unfolding in vitro. We demonstrate that Cu(II) coordination affects both protein stability and unfolding pathway, whereas Cu(I) has only a small effect on stability. Apo-CopC unfolds in a two-state reaction between pH 4 and 7.5 with maximal stability at pH 6. In contrast, Cu(II)-CopC unfolds in a three-state reaction at pH6 that involves a partly folded intermediate that retains Cu(II). This intermediate exhibits high thermal and chemical stability. Unique energetic and structural properties of different metalated CopC forms may help facilitate metal transport to many partners in vivo.  相似文献   

8.
The interaction of phytic acid with Ca(II) has been studied by potentiometric titration and by measurement of free Ca(II) concentrations using an ion Selective electrode. With increasing Ca(II) concentration, the titration curve of phytic acid is displaced to regions of lower pH. In the binding of calcium ions to phytic acid, there is no evidence that significant binding occurs below approximately pH 5. Above this pH, the extent of binding is dependent upon both pH and the calcium to phytic acid ratios. Maximum binding obtains at a Ca(II):phytate ratio of 6 with 4.8 mol of Ca(II) bound per mol of phytate above pH ca. 8. Binding constants are apparently very large since binding isotherms at any Ca(II):phytate ratio are a linear function of the total calcium ion concentration. In all cases, binding occurs only when one or more phosphate groups have been converted to the oxo dianion form. The apparent pK' values (curve-fit parameters) that describe the potentiometric titration data are in good agreement with the constants evaluated from the binding of Ca(II) to phytate as a function of pH. Using CPK space-filling models, structures containing six metal ions in coordinate linkage to pairs of oxo dianions have been constructed and discussed within the framework of the axial conformation of phytic acid and the order of proton removal with an increase in pH based upon NMR studies.  相似文献   

9.
The acid base behavior of phytate has been studied (at t=25 degrees C by potentiometry, ISE-H+ glass electrode) in NaNO3aq at different ionic strengths (0.1 < or = I/mol L(-1) < or = 1.0). The interactions with copper(II) were investigated in the same experimental conditions in different metal to ligand (Phy) ratios (1:1 < or = Cu2+ :Phy < or = 4:1), by using both ISE-H+ and ISE-Cu2+ electrodes. Phytate acid base behavior in sodium nitrate is very similar to that in sodium chloride, previously investigated. In the experimental conditions adopted, the formation of three CuiHjPhy(12-2i-j)- species is observed: the mononuclear CuH4Phy6- and CuH5Phy5-, and the dinuclear Cu2H5Phy3-. Analysis of complex formation constants at different ionic strengths reveals that both ISE-H+ and ISE-Cu2+ electrodes gave, within the experimental error, analogous values. Dependence of complex formation constants on ionic strength was modeled by EDH (Extended Debye-Hückel) and SIT (Specific ion Interaction Theory) equations. The sequestering ability of phytate toward copper(II) has been evaluated by the calculation of pL50 (the total ligand concentration, as -log CL, able to bind 50% of metal cation), an empirical parameter already proposed for an objective "quantification" of this ability. A thorough analysis of literature data on phytate-copper(II) complexes has been performed.  相似文献   

10.
DNA was found to be cleaved in neutral solutions containing arenes and copper (II) salts. The reaction is comparable in efficiency with the DNA cleavage by such systems as Cu(II)-phenanthroline and Cu(II)-ascorbic acid, but, in contrast to the latter, the system Cu(2+)-arene does not require the presence of an exogenous reducing agent or hydrogen peroxide. The system Cu(2+)-arene does not cleave DNA under anaerobic conditions. Catalase, sodium azide, and bathocuproine, which is a specific chelator of Cu(I), completely inhibit the reaction. The data obtained allow one to suppose that Cu(I) ions, superoxide radical, and singlet oxygen participate in the reaction. It has been shown by the EPR method using spin traps that the reaction proceeds with formation of alkoxyl radicals, which can insert breaks in the DNA molecule. For effective cleavage of DNA in the Cu(II)-o-bromobenzoic acid system, the radicals have to be generated by a specific copper-DNA-o-bromobenzoic acid complex, in which copper ions are most probably coordinated with oxygen atoms of the DNA phosphate groups. The English version of the paper: Russian Journal of Bioorganic Chemistry, 2003, vol. 29, no. 6; see also http://www.maik.ru.  相似文献   

11.
The kinetics of reconstitution of Pseudomonas aeruginosa azurin from its apoprotein and copper(II) salts have been studied using absorbance at 625 nm and fluorescence emission at 308 nm as monitors of the process. At low Cu(II) concentrations the rates of both absorbance and fluorescence changes are linearly dependent on Cu(II) concentration. At higher Cu(II) concentrations the rate of absorbance change is independent of Cu(II) concentration. The rates of both absorbance and fluorescence changes as a function of pH suggest that the titration of a single ionizable group is important for the Cu(II)-dependent reaction. Overall analysis of the kinetics suggests that the fluorescence change and the absorbance change are associated with at least two steps in the overall pathway of the formation of the metal-protein complex, and that the copper(II) and tryptophan environments in this protein, though perhaps spatially close, may be distinct.  相似文献   

12.
Novel homodinuclear Cu(II) (K1), heterodinuclear Cu(II)-Mn(II) (K2) and homotrinuclear Cu(II) (K3) complexes with a novel oxime-type ligand have been prepared and their nucleolytic activities on pCYTEXP were established by neutral agarose gel electrophoresis. The analyses of the cleavage products obtained electrophoretically indicate that although the examined complexes induces very similar conformational changes on supercoiled DNA by converting supercoiled form to nicked form than linear form in a sequential manner as the complex concentration or reaction period is increased, K3 is less effective than the two others. The oxime complexes were nucleolytically active at physiological pH values but the activities of K1 or K2 were diminished by increasing the pH of the reaction mixture. In contrast, K3 makes dominantly single strand nicking by producing nicked circles on DNA at almost all the applied pH values. Metal complex induced DNA cleavage was also tested for inhibition by various radical scavengers as superoxide dismutase (SOD), azide, thiourea and potassium iodide. The antioxidants inhibited the nucleolytic acitivities of the oxime complexes but SOD afforded no protection indicating that the nucleolytic mechanism involves of copper and/or manganese complex-mediated reactive oxygen species such as hydroxyl radicals being responsible for the oxidative DNA cleavage.  相似文献   

13.
Preparation and characterisation of copper(II) hyaluronate   总被引:1,自引:0,他引:1  
Amorphous copper complexes of the general composition Cu(C14H20O11N)2 x xH2O have been prepared with high- and low-molecular-weight hyaluronic acid (HA). Optimal conditions for preparation are obtained at pH values from 5.0 to 5.5, with a molar ratio of HA versus Cu2+ of 1:1, and at a mass concentration of 5 and 10 mg/mL for high- (Mw = 1.8 x 10(6) Da) and low-molecular-weight sodium hyaluronate (Mw = 2 x 10(5) Da), respectively. The coordination polyhedron of the copper ion has been elucidated by EXAFS and XANES spectroscopy. Copper atoms are octahedrally coordinated in both cases with four equatorial Cu-O bond lengths of 1.95 A, and two axial Cu-O bonds of 2.46 A. Visible spectra of acidic aqueous solution suggest that substitution of axial oxygens by NH groups occurs at pH 6.5 or higher. If the pH value of the copper(II) hyaluronate solution increases above 6.5, the coordination of copper(II) changes. It is very likely that the N atom coming from the acetamido group enters into the coordination sphere of the copper(II) ion.  相似文献   

14.
Interactions of inosine derivatives with copper(II) were studied in the pH range 1.4–13 in 50% H2O-50% DMSO solution. The distinct pH dependence of the optical spectra observed in copper(II)-inosine complexes are correlated to their respective EPR changes as a function of pH. It was concluded that a simple 1:1 complex of copper(II)-inosine is formed in the pH range 1.4–5.0 and bis complexes are present in the pH 5.0–6.2 region solutions of inosine and Cu(II). From pH 6.2 to 7.8 a diamagnetic, hydroxybridged complex dominates. At pH 7.8–9.2 an insoluble, oxybridged species is formed in addition to the soluble paramagnetic Cu(NI)4 complex. Starting from pH 9.1 the N-polymeric complex is formed which is stable up to pH 12.5, and above pH 12.5 the only species is the Cu(ribose)2 complex.  相似文献   

15.
This study examined the effect of diet-induced, marginal zinc deficiency for 7 wks in 15 men (aged 25.3 +/- 3.3 yrs; mean +/- SD) on selected indices of iron and copper status. The regimen involved low-zinc diets based on egg albumin and soy protein with added phytate and calcium such that mean [phytate]/[Zn] and [phytate] X [Ca]/[Zn] molar ratios were 209 and 4116, respectively, for 1 wk, followed by 70 and 2000, respectively, for 6 wks. Subjects were then repleted with 30 mg Zn/d for 2 wks. Plasma copper, Cu,Zn-superoxide dismutase (Cu,Zn-SOD) activity in plasma and red blood cells (RBC), hemoglobin, hematocrit, and serum ferritin were determined weekly on fasting blood samples. Significant reductions (p less than 0.05) after 7 wks in RBC Cu,Zn-superoxide dismutase (49.5 +/- 7.2 vs 33.6 +/- 6.3 U/mg Hb) and serum ferritin (69.2 +/- 38.7 vs 53.8 +/- 33.7 micrograms/L) occurred; no comparable decline was noted for plasma Cu, hemoglobin, or hematocrit. Significant (p less than 0.05) but less consistent changes were also observed in plasma superoxide dismutase activity. None of the changes were associated with the decreases in plasma, urinary and hair zinc concentrations, and alkaline phosphatase activity in RBC membranes. Results indicate that the biochemical iron and copper status of the subjects was marginally impaired, probably from the dietary regimen that induced marginal zinc deficiency.  相似文献   

16.
L-cysteine, D-penicillamine, and L-glutathione were oxidized to symmetrical disulfides in the presence of Cu(II)(3,5-DIPS)2 and air-oxygen at physiologic pH, 7.3. Air-oxygen caused the oxidation of thiol reduced copper, Cu(I), to Cu(II), as evidenced by expected spectrophotometric changes in these reaction mixtures. L-cysteine, D-penicillamine, and L-glutathione formed mixed disulfides and TNB with the addition of DTNB to solutions of these thiols. The observed order of reactivity for these thiols with DTNB was: L-cysteine greater than D-penicillamine greater than L-glutathione. Surprisingly, Cu(II)(3,5-DIPS)2 converted these mixed disulfides to their symmetrical disulfides and DTNB, and although the initial conversion rate was rapid, complete conversion required more than two hours. These observations suggest caution with regard to the spectrophotometric determination of thiols immediately after the addition of Ellman's reagent. These results also clarify an earlier report concerning the oxidation of thiols by Cu(II)(o-phenanthroline)2 and offer caution with regard to the determination of thiols using DTNB in the presence of copper complexes. Spectrophotometric data are provided in support of the suggestion that analysis of plasma or cellular samples for thiols be done in the absence of copper(II) complexes to avoid false negative results.  相似文献   

17.
Several features of the catalytic oxidation of cysteine by ceruloplasmin and nonenzymic Cu(II) at pH 7 have been compared. The oxidation of cysteine by ceruloplasmin has several properties in common with the Cu(II) catalyzed oxidation of cysteine: pH maxima, thiol specificity, lack of inhibition by anions, and high sensitivity to inhibition by copper complexing reagents. These two catalysts differed in their molecular activity, in their ability to oxidize penicillamine and thioglycolate, and in that H2O2 was produced as a primary product only during Cu(II) oxidation. The oxidation of cysteine by ceruloplasmin was compared also with the ceruloplasmin catalyzed oxidation of o-dianisidine, a classical pH 5.5 substrate. The mechanism of the oxidation of cysteine by ceruloplasmin at pH 7 differed from that of o-dianisidine oxidation because the latter substrate was inhibited by anions but not by copper complexing agents. Spectral and other data suggest that during the ceruloplasmin reaction with cysteine there is a one electron transfer from cysteine to ceruloplasmin resulting in the specific reduction of type lb Cu(II).  相似文献   

18.
Miura T  Sasaki S  Toyama A  Takeuchi H 《Biochemistry》2005,44(24):8712-8720
The physiological function of the prion protein (PrP) remains enigmatic despite its established involvement in the pathogenesis of spongiform encephalopathies. PrP is a glycolipid-anchored membrane protein, which constitutively recycles between the cell surface and an endosomal compartment. The N-terminal region of PrP contains a four tandem repeat (OP4) of the octapeptide PHGGGWGQ (OP) that binds and reduces Cu(II) ions. We have examined the kinetic properties of the OP4-mediated Cu(II) reduction and found that OP4 exhibits the highest reduction activity around pH 6.5, close to the pH in early endosomes. All four OP units and at least one tryptophan side chain are essential for Cu(II) reduction. The reaction is described by an uncompetitive substrate inhibition mechanism involving a 1:1 Cu(II)-OP4 active intermediate. Structural analysis by Raman spectroscopy has revealed that the Cu(II) ion is coordinated by four histidine Ntau atoms in the active intermediate and the feasibility of formation of this intermediate correlates with the Cu(II) reduction over a pH range from 5.0 to 8.2. Molecular mechanics calculations suggest that two tryptophan residues of OP4 are located near the Cu(II) site, being consistent with the importance of redox-active tryptophan in the Cu(II) reduction. PrP has been proposed to capture Cu(II) ions in the extracellular space and release them in the endosome. The results of this study strongly suggest that PrP also plays a role in the reduction of captured Cu(II) ions prior to their transfer to Cu(I)-specific intracellular copper trafficking proteins.  相似文献   

19.
Wu YW  Tsai YH 《Bioconjugate chemistry》2008,19(9):1822-1830
Thioureas (Tus) are widely used in chemical and pharmaceutical industries. This study demonstrated that copper induced the disulfide-linkage between Tus, such as alpha-naphthylthiourea (ANTU) and fluorescein-5-isothiocyanate cadaverine (FTC), with albumin (Alb), a major carrier protein in plasma with multiple functions. This reaction was absolutely copper-dependent, whereas cobalt, nickel, calcium, magnesium, zinc, iron, and manganese ions could not induce the same reaction. The reaction was substrate dose-dependent, and occurred optimally at pH 6.5. The resulting conjugated product was heat-labile, but stable in pH 6.0-8.0 buffer at 25 degrees C. The linkage could be reduced by Cu(I) (in acidic pH) and thiol-reducing agents. The mechanism of albumin thioureation was concluded: (i) the binding of Cu(II) with albumin is not necessary for the reaction, while the formation of Tus-Cu(II) complex is essential; (ii) thioureation resulted from the attack of Tus-Cu(II) at Alb-Cys(34)-SH to form the Alb-Cys(34)-S-S-Tus complex accompanied by the release of Cu(I); (iii) the released Cu(I) would back inhibit the reaction because of its competition with Cu(II) for Tus binding. These phenomenons may have important implications for the pharmacokinetics of thiourea-based drugs in plasma.  相似文献   

20.
Galactose oxidase is a copper metalloenzyme containing a novel protein-derived redox cofactor in its active site, formed by cross-linking two residues, Cys228 and Tyr272. Previous studies have shown that formation of the tyrosyl-cysteine (Tyr-Cys) cofactor is a self-processing step requiring only copper and dioxygen. We have investigated the biogenesis of cofactor-containing galactose oxidase from pregalactose oxidase lacking the Tyr-Cys cross-link but having a fully processed N-terminal sequence, using both Cu(I) and Cu(II). Mature galactose oxidase forms rapidly following exposure of a pregalactose oxidase-Cu(I) complex to dioxygen (t(1/2) = 3.9s at pH7). In contrast, when Cu(II) is used in place of Cu(I) the maturation process requires several hours (t(1/2) = 5.1 h). EDTA prevents reaction of pregalactose oxidase with Cu(II) but does not interfere with the Cu(I)-dependent biogenesis reaction. The yield of cross-link corresponds to the amount of copper added, although a fraction of the pregalactose oxidase protein is unable to undergo this cross-linking reaction. The latter component, which may have an altered conformation, does not interfere with analysis of cofactor biogenesis at low copper loading. The biogenesis product has been quantitatively characterized, and mechanistic studies have been developed for the Cu(I)-dependent reaction, which forms oxidized, mature galactose oxidase and requires two molecules of O2. Transient kinetics studies of the biogenesis reaction have revealed a pH sensitivity that appears to reflect ionization of a protein group (pKa = 7.3) at intermediate pH resulting in a rate acceleration and protonation of an early oxygenated intermediate at lower pH competing with commitment to cofactor formation. These spectroscopic, kinetic, and biochemical results lead to new insights into the biogenesis mechanism.  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