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 788 毫秒
1.
The association process of FAD and apo-electron-transferring flavoprotein (apoETF) from hog kidney was investigated. The reaction schemes which involve the association-dissociation of the protein species could be excluded by the light scattering data, which indicated that the molecular weights of apoETF and holoETF are identical. The binding reaction between FAD and a large excess of apoETF was monophasic and obeyed pseudo-first order kinetics. On the other hand, the reaction between apoETF and a large excess of FAD was biphasic: the fast phase obeyed a pseudo-first order reaction, and the rate of the slow phase was almost independent of FAD concentration. These results suggest the existence of two different forms of apoETF, as represented in the following reaction scheme: [formula: see text] where "F" is FAD, "H" is holoETF, and "A" and "A" are the different forms of apoETF. The kinetic parameters were determined as k-1 = 3.9 x 10(4) M-1.s-1, k-1 approximately 10(-5) s-1, k+2 = 1.0 x 10(-3) s-1, and k-2 = 3.1 x 10(-3) s-1, in 50 mM potassium phosphate buffer, pH 7.6, containing 0.3 mM EDTA, and 5% v/v glycerol, at 7 degrees C. The elution patterns of apoETF on molecular sieve chromatography were very different from that of holoETF although the true molecular weights were identical. This result suggests that the structure of apoETF differs greatly from that of holoETF.  相似文献   

2.
Apoprotein of electron-transferring flavoprotein (ETF) reacts with FAD as follows: A*<-->A, A+FAD<-->holoETF. Two different forms of apoETF (A* and A) convert into each other and only one of them, A, can associate with FAD [Sato, K. et al. (1991) J. Biochem. 109, 734-740]. In the present study, the reactions between apoETF and ATP, ADP, AMP, riboflavin, or FMN were investigated. It was revealed that all three adenine nucleotides bind with apoETF with the same kinetic reaction scheme as FAD, and compete with FAD. These results suggest that the nucleotides bind to A with the same location as the ADP part of FAD in holoETF and that the ADP-binding site of apoETF is generated upon conversion from A* to A. Neither riboflavin nor FMN bound to apoETF regardless of the presence or absence of the nucleotides, indicating that the ADP part of the FAD molecule is essential to the incorporation of the isoalloxazine ring into ETF. The binding rate constant of FAD to A was 1/20 of that of ADP while the dissociation rate constant was 1/1,000. This indicates that the riboflavin part of FAD inhibits the binding of FAD by steric hindrance, while after the binding, it stabilizes the complex.  相似文献   

3.
Three very unstable mutant forms of staphylococcal nuclease were used to quantitate the change in the apparent equilibrium constant for reversible denaturation (Kapp) as a function of denaturant concentration for a variety of different denaturing solutes. The value of this equilibrium constant in the absence of denaturant (Kapp,0) was determined by renaturation of the mutant proteins with a combination of glycerol and calcium ion, the latter of which binds at the active site in the native conformation. Because Kapp,0 fell in the easily measurable range between 0.1 and 1, the change in Kapp, and thus the change in free energy (delta Gapp), at very low concentrations of denaturants could be accurately measured. With guanidine hydrochloride (GuHCl), the rate of change of the apparent free energy of denaturation with respect to denaturant concentration (d(delta Gapp)/dCGuHCl or mGuHCl) was found to be remarkably constant down to zero denaturant concentration, even though this value was different for each of the three proteins. Unlike GuHCl, urea exhibited a slightly reduced value of d delta Gapp/dCurea at low concentrations. Results with a number of thiocyanate, perchlorate, and iodide salts confirmed that the Hofmeister series holds for concentrations below 0.1 M; that is, with regard to efficacy as a denaturant SCN- greater than ClO4- greater than I- and Li+,NH4+ greater than Na+,K+. However, all of the chaotropic salts analyzed exhibited markedly increased values of d(delta Gapp)/dCsalt at concentrations below 0.2 M. One possible explanation for these large deviations from a linear relationship between delta Gapp and salt concentration is that weak binding or adsorption of chaotropic anions is occurring at a saturable number of sites in hydrophobic regions of the denatured state.  相似文献   

4.
The energetics of the Na+-dependent transport of D-glucose into osmotically active membrane vesicles, derived from the brush borders of the rabbit renal proximal tubule, was studied by determining how alterations in the electrochemical potential of the membrane induced by anions, ionophores, and a proton conductor affect the uptake of the sugar. The imposition of a large NaCl gradient (medium is greater than vesicle) resulted in the transient uptake of D-glucose into brush border membranes against its concentration gradient. In the presence of Na+ salts of isethionate or sulfate, both relatively impermeable anions, there was no accumulation of D-glucose above the equilibrium value. With Na+ salts of two highly permeable lipophilic anions, NO3- and SCN-, the transient overshoot was enhanced relative to that with Cl-. With Na+ salts whose mode of membrane translocation is electroneutral, i.e. acetate, bicarbonate, and phosphate, no overshoot was found. These findings suggest that only anions which penetrate the brush border membrane and generate an electrochemical potential, negative on the inside, permit the uphill Na+-dependent transport of D-glucose.  相似文献   

5.
An in-depth analysis of the kinetics of 5 alpha-reductase in human prostatic tissue gave findings inconsistent with the claim that the enzyme is michaelian. In both hyperplastic and malignant tissue, the time-course of the conversion of testosterone (T) into dihydrotestosterone (DHT) was non-linear under conditions ensuring less than 15% conversion of substrate and cofactor. An initial rapid phase of conversion was followed by a long steady-state phase. This time-dependent change in conversion rate was not due to enzyme denaturation, fast inhibition by substrate or product effects. It resulted from a true slow transient kinetic process induced in the reactive enzyme by the substrates. Under our experimental conditions at pH 5.5, 5 alpha-reductase appeared to undergo a conformational change from an initially highly reactive form to a less reactive form. Since this "hysteretic" behavior was correlated with apparently negative cooperativity in enzyme kinetics, we postulate that, as previously described for other key metabolic enzymes, regulation of 5 alpha-reductase activity in the prostate depends on the molecular flexibility of the enzyme and on changes in the cooperativity of different enzyme forms over time. This original non-michaelian behavior may explain the conflicting kinetics reported so far in the literature for this enzyme. The clinical implications of 5 alpha-reductase hysteresis and its involvement in the damping of DHT production within the prostate are discussed.  相似文献   

6.
Xanthine oxidase employs four electron transport sites (flavin adenine dinucleotide (FAD), molybdenum, and two FeS centers) in catalyzing a variety of redox reactions. To determine whether the redox sites reside in independent domains of the enzyme, the temperature of heat inactivation of each site's catalytic activity was determined, except that no attempt was made to distinguish between the two FeS sites. In the oxidase form of xanthine oxidase, the order of thermal stabilities was Mo greater than FAD greater than FeS, while after conversion to its dehydrogenase form the above ranking was Mo greater than FeS greater than FAD. The small but reproducible difference in heat inactivation temperatures among the redox sites demonstrated that the sites are located in separate domains of the enzyme. To confirm the above segregation of redox centers, the temperature of heat-induced release of each redox cofactor from its site on the enzyme was examined. These temperatures were found to be different for each redox cofactor and agreed closely with the heat inactivation temperatures measured above. The data thus demonstrate that both heat inactivation and cofactor release derive from thermal unfolding of independent domains. Using a technique termed "thermal digestion analysis," the FAD domain was located in a approximately equal to 42,000-Da tryptic fragment, while the FeS and Mo domains were isolated in a trypsin-resistant 92,000-Da fragment. We conclude that xanthine oxidase is constructed in modular fashion with the redox sites located in independent structural domains.  相似文献   

7.
With 45 % or more oil content that contains more than 55 % alpha linolenic (LIN) acid, linseed (Linum usitatissimum L.) is one of the richest plant sources of this essential fatty acid. Fatty acid desaturases 2 (FAD2) and 3 (FAD3) are the main enzymes responsible for the Δ12 and Δ15 desaturation in planta. In linseed, the oilseed morphotype of flax, two paralogous copies, and several alleles exist for each gene. Here, we cloned three alleles of FAD2A, four of FAD2B, six of FAD3A, and seven of FAD3B into a pYES vector and transformed all 20 constructs and an empty construct in yeast. The transformants were induced in the presence of oleic (OLE) acid substrate for FAD2 constructs and linoleic (LIO) acid for FAD3. Conversion rates of OLE acid into LIO acid and LIO acid into LIN acid were measured by gas chromatography. Conversion rate of FAD2 exceeded that of FAD3 enzymes with FAD2B having a conversion rate approximately 10 % higher than FAD2A. All FAD2 isoforms were active, but significant differences existed between isoforms of both FAD2 enzymes. Two FAD3A and three FAD3B isoforms were not functional. Some nonfunctional enzymes resulted from the presence of nonsense mutations causing premature stop codons, but FAD3B-C and FAD3B-F seem to be associated with single amino acid changes. The activity of FAD3A-C was more than fivefold greater than the most common isoform FAD3A-A, while FAD3A-F was fourfold greater. Such isoforms could be incorporated into breeding lines to possibly further increase the proportion of LIN acid in linseed.  相似文献   

8.
NAD(P)H quinone oxidoreductase 1 is involved in antioxidant defence and protection from cancer, stabilizing the apoptosis regulator p53 towards degradation. Here, we studied the enzymological, biochemical and biophysical properties of two cancer-associated variants (p.R139W and p.P187S). Both variants (especially p.187S) have lower thermal stability and greater susceptibility to proteolysis compared to the wild-type. p.P187S also has reduced activity due to a lower binding affinity for the FAD cofactor as assessed by activity measurements and direct titrations. Native gel electrophoresis and dynamic light scattering also suggest that p.P187S has a higher tendency to populate unfolded states under native conditions. Detailed thermal stability studies showed that all variants irreversibly denature causing dimer dissociation, while addition of FAD restores the stability of the polymorphic forms to wild-type levels. The kinetic destabilization induced by polymorphisms as well as the kinetic protection exerted by FAD was confirmed by measuring denaturation kinetics at temperatures close to physiological. Our data suggest that the main molecular mechanisms associated with these cancer-related variants are their low binding affinity for FAD and/or kinetic instability. Thus, pharmacological chaperones may be useful in the treatment of patients bearing these polymorphisms.  相似文献   

9.
Stearoyl-acyl carrier protein desaturase-mediated conversion of stearic acid to oleic acid (18:1) is the key step that regulates the levels of unsaturated fatty acids (FAs) in cells. Our previous work with the Arabidopsis (Arabidopsis thaliana) ssi2/fab2 mutant and its suppressors demonstrated that a balance between glycerol-3-phosphate (G3P) and 18:1 levels is critical for the regulation of salicylic acid (SA)- and jasmonic acid-mediated defense signaling in the plant. In this study, we have evaluated the role of various genes that have an impact on SA, resistance gene-mediated, or FA desaturation (FAD) pathways on ssi2-mediated signaling. We show that ssi2-triggered resistance is dependent on EDS1, PAD4, EDS5, SID2, and FAD7 FAD8 genes. However, ssi2-triggered defects in the jasmonic acid pathway, morphology, and cell death phenotypes are independent of the EDS1, EDS5, PAD4, NDR1, SID2, FAD3, FAD4, FAD5, DGD1, FAD7, and FAD7 FAD8 genes. Furthermore, the act1-mediated rescue of ssi2 phenotypes is also independent of the FAD2, FAD3, FAD4, FAD5, FAD7, and DGD1 genes. Since exogenous application of glycerol converts wild-type plants into ssi2 mimics, we also studied the effect of exogenous application of glycerol on mutants impaired in resistance-gene signaling, SA, or fad pathways. Glycerol increased SA levels and induced pathogenesis-related gene expression in all but sid2, nahG, fad7, and fad7 fad8 plants. Furthermore, glycerol-induced phenotypes in various mutant lines correlate with a concomitant reduction in 18:1 levels. Inability to convert glycerol into G3P due to a mutation in the nho1-encoded glycerol kinase renders plants tolerant to glycerol and unable to induce the SA-dependent pathway. A reduction in the NHO1-derived G3P pool also results in a partial age-dependent rescue of the ssi2 morphological and cell death phenotypes in the ssi2 nho1 plants. The glycerol-mediated induction of defense was not associated with any major changes in the lipid profile and/or levels of phosphatidic acid. Taken together, our results suggest that glycerol application and the ssi2 mutation in various mutant backgrounds produce similar effects and that restoration of ssi2 phenotypes is not associated with the further desaturation of 18:1 to linoleic or linolenic acids in plastidal or extraplastidal lipids.  相似文献   

10.
The FAD of p-hydroxybenzoate hydroxylase (PHBH) is known to exist in two conformations. The FAD must be in the in-position for hydroxylation of p-hydroxybenzoate (pOHB), whereas the out-position is essential for reduction of the flavin by NADPH. In these investigations, we have used 8-mercapto-FAD and 8-hydroxy-FAD to probe the movement of the flavin in catalysis. Under the conditions employed, 8-mercapto-FAD (pK(a) = 3.8) and 8-hydroxy-FAD (pK(a) = 4.8) are mainly anionic. The spectral characteristics of the anionic forms of these flavins are very sensitive to their environment, making them sensitive probes for detecting movement of the flavin during catalysis. With these flavin analogues, the enzyme hydroxylates pOHB efficiently, but at a rate much slower than that of enzyme with FAD. Reaction of oxygen with reduced forms of these modified enzymes in the absence of substrate appears to proceed through the formation of the flavin-C4a-hydroperoxide intermediate, as with normal enzyme, but the decay of this intermediate is so fast compared to its formation that very little accumulates during the reaction. However, after elimination of H2O2 from the flavin-C4a-hydroperoxide, a perturbed oxidized enzyme spectrum is observed (Eox*), and this converts slowly to the spectrum of the resting oxidized form of the enzyme (Eox). In the presence of pOHB, PHBH reconstituted with 8-mercapto-FAD also shows the additional oxidized intermediate (Eox*) after the usual oxygenated C4a-intermediates have formed and decayed in the course of the hydroxylation reaction. This Eox* to Eox step is postulated to be due to flavin movement. Furthermore, binding of pOHB to resting (Eox) follows a three-step equilibrium mechanism that is also consistent with flavin movement being the rate-limiting step. The rate for the slowest step during pOHB binding is similar to that observed for the conversion of Eox* to Eox during the oxygen reaction in the absence or presence of substrate. Steady-state kinetic analysis of PHBH substituted with 8-mercapto-FAD demonstrated that the apparent k(cat) is also similar to the rate of Eox* conversion to Eox. Presumably, the protein environment surrounding the flavin in Eox* differs slightly from that of the final resting form of the enzyme (Eox).  相似文献   

11.
Autoflavinylation of apo6-hydroxy-D-nicotine oxidase   总被引:2,自引:0,他引:2  
6-Hydroxy-D-nicotine oxidase (6-HDNO) was expressed in Escherichia coli JM109 cells from the recombinant plasmid pAX-6-HDNO as a beta-galactosidase-6-HDNO fusion protein. Affinity chromatography of the fusion protein on p-aminobenzyl-1-thio-beta-galactopyranoside-agarose and subsequent digestion with protease Xa yielded highly purified apo6-HDNO. Incubation of the purified protein with [14C]FAD demonstrated that flavinylation of apo6-HDNO proceeds autocatalytically. Phosphorylated three-carbon compounds such as glycerol-3-P, which are known to stimulate the formation of the histidyl (N3)-(8 alpha) FAD between apo6-HDNO and FAD (Brandsch, R., and Bichler, V. (1989) Eur. J. Biochem. 182, 125-128), could be replaced in their action by high concentrations of glycerol (45%) or sucrose (20%). These substances apparently induced and stabilized a conformational state of the apoenzyme compatible with covalent attachment of FAD. In the absence of glycerol the apoenzyme readily lost the ability to form holoenzyme at temperatures above 30 degrees C. Holoenzyme formation protected the 6-HDNO polypeptide from this thermal denaturation. Autoflavinylation of 6-HDNO was inhibited by the sulfhydryl reagents dithionitrobenzoate or N-ethylmaleimide. Inhibition was prevented by mercaptoethanol or FAD, but not 6-hydroxy-D-nicotine, the substrate of the holoenzyme. A cysteine-thiol group may therefore be involved in reactions leading to the covalent attachment of FAD to apo6-HDNO. When flavinylation of apo6-HDNO proceeded under anaerobic conditions, the amount of incorporation of [14C]FAD into the polypeptide was indistinguishable from reactions performed in the presence of O2. None of the FAD-derivatives (8-demethyl-FAD, 8-chloro-FAD, and 5-deaza-FAD) could replace FAD in holoenzyme formation. The failure of covalent attachment of 5-deaza-FAD to apo6-HDNO is in agreement with the assumption that the quinone methide form of the isolloxazine ring is an intermediate in the flavinylation reaction.  相似文献   

12.
The conformation of the polypeptide antibiotic gramicidin is greatly influenced by its environment. In methanol, it exists as an equilibrium mixture of four interwound double-helical conformers that differ in their handedness, chain orientation, and alignment. Upon the addition of multivalent cationic salts, there is a shift in the equilibrium to a single conformer, which was monitored in this study by circular dichroism spectroscopy. With increasing concentrations of multivalent cations, both the magnitude of the entire spectrum and the ratio of the 229-nm to the 210-nm peak were increased. The spectral change is not related to the charge on the cation, but appears to be related to the cationic radius, with the maximum change in ellipticity occurring for cations with a radius of approximately 1 A. The effect requires the presence of an anion whose radius is greater than that of a fluoride ion, but is otherwise not a function of anion type. It is postulated that multivalent cations interact with a binding site in one of the conformers, known as species 1 (a left-handed, parallel, no stagger double helix), stabilizing a modified form of this type of structure.  相似文献   

13.
Physically cross-linked novel block copolymer hydrogels with tunable hydrophilic properties for biomedical applications were synthesized by controlled radical polymerization of polyurethane macroiniferter and (2,2-dimethyl-1,3-dioxolane) methyl methacrylate. The block copolymers were converted to hydrogels by the selective hydrolysis of poly[(2,2-dimethyl-1,3-dioxolane) methyl methacrylate] block to poly(glycerol methacrylate). The block copolymerization has been monitored by monomer conversion and molecular weight increase as a function of time. It was observed that the polymerization proceeded with a characteristic "living" behavior where both monomer conversion and molecular weight increased linearly, with increasing reaction time. The resulting hydrogels were investigated for their equilibrium water content (EWC), dynamic water contact angles, swelling kinetics, thermodynamic interaction parameters, plasma protein adsorption, and platelet adhesion. Similar to our previous mechanically responsive hydrogels (Mequanint, K.; Sheardown, H. J. Biomater. Sci. Polym. Ed. 2005, 10, 1303-1318), the present results indicated that block copolymer hydrogels have excellent hydrophilicity and swelling behavior with improved modulus of elasticity. The equilibrium swelling was affected by the hydrolysis time, block length of poly(glycerol methacrylate), temperature, and the presence of soluble salts. Fibrinogen adsorption and platelet adhesion were significantly lower for the hydrogels than for the control polyurethane, whereas albumin adsorption increased for the hydrogels in proportion to the contents of poly(glycerol methacrylate). These hydrogels have potential in a number of biomedical applications such as drug delivery and scaffolds for tissue engineering.  相似文献   

14.
Ferric yeast cytochrome c peroxidase in the presence of different anions may assume a number of forms which differ in optical spectra and chemical properties. In solutions whose only anion is acetate, two spectral forms are present together in an equilibrium. Each of these spectral species is believed to bear bound acetate anion. A form characterized by an intense absorption maximum at 620 nm is unreactive enzymatically and does not react with hydrogen peroxide or with dithionite. A form characterized by a less intense absorption near 645 nm is enzymatically and chemically reactive. Increasing temperature and increasing pH displace the equilibrium toward the 645 nm form. Increasing cytochrome c peroxidase concentration favors the 620 nm form. In kinetic experiments in which the 645 nm form is removed by rapid reaction with H2O2 or dithionite, the 620 nm form is converted in a first order reaction (k = 0.36 s-1, 15 degrees C) to the 645 nm form. In solutions whose sole anion is phosphate a 645 nm form is the only demonstrable spectral species. The enzymatic activity and rates of chemical reaction of 645 nm spectral forms occurring in acetate and in phosphate buffers are the same.  相似文献   

15.
The kinetic properties of NADP-dependent sorbitol dehydrogenase from G. oxydans cell extract were studied at pH 8.8 and 9.3 in the direction of D-sorbitol oxydation. It was shown that the shape of the kinetic curves of NADPH accumulation in time is characterised by initial burst whose magnitude depends on the concentration of the enzyme extract used. Preincubation of the enzyme with NADP or D-sorbitol eliminated the initial burst on these curves and transformed them into straight lines coming from the start of co-ordinates. The dependence of the stationary reaction rate on the enzyme extract concentration is not a linear one. The kinetic dependences of stationary rate of the reaction catalysed by the enzyme on the concentration of D-sorbitol and NADP at pH 8.8 and 9.3 were examined under all conditions studied; the shape of these kinetic curves altered to considerable extent with the alteration of the enzyme extract concentration in the reaction mixture and pH. At pH 9.3 several intermiediate plateaux were found on the curves of the D-sorbitol concentration dependent stationary rate of the reaction. The preincubation of the enzyme extract with NADP during 1.5 h removed the intermediate plateau on these curves and made them hyperbolic. Disk-electrophoresis of the enzyme extract in PAAG concentration gradient showed that at pH 8.8 the enzyme exists in one active form, while at pH 9.3 it exists in three major and three minor active forms of the enzyme differing in their molecular weights are found. It is assumed that the enzyme from G. oxydans cell extract can exist in a great number of molecular equilibrium forms, the rate of quilibrium being comparable or significantly less than that of the enzymatic reaction. NADP significantly influences on the equilibrium of the molecular forms of the enzyme.  相似文献   

16.
"Biosynthetic" L-threonine dehydratase (EC 4.2.1.16) was purified to a homogeneous state with 29% yield of total activity from Escherichia coli K-12. The homogeneity of the enzyme was shown by polyacrylamide gel disc electrophoresis in the presence of dodecyl sulphate. The enzyme consisted of equal subunits having a molecular weight of about 57 000. The polyacrylamide gel disc electrophoresis has shown that the native enzyme consisted of a set of oligomeric forms. The multiplicity of molecular organization of the enzyme was reflected in complicated kinetic behaviour: at pH greater than 9 on the plots of initial reaction rate (v) versus initial substrate concentration ([S]o) there were four inflexion points (two intermediate plateaux), the position and deepness of which depended on enzyme concentration. At pH 8.3 on the v versus [S]o plots appeared two inflexion points (one intermediate plateu), the position of which practically did not depend on enzyme concentration in the reaction mixture, but strongly depended on the enzyme concentration in the stock solution. Repeated polyacrylamide gel disc electrophoresis of several oligomeric forms, isolated by the first electrophoresis, has shown that the oligomeric forms underwent a slow polymerization. It was suggested that "biosynthetic" L-threonine dehydratase from E. coli K-12 is a set of multiple oligomeric forms, having different kinetic parameters. Probably, each form of the enzyme has a "simple" kinetics characterized by hyperbolic or sigmoidal shape of v versus [S]o plots. The rate of equilibrium installation between the oligomeric forms was small in comparison with the enzyme reaction velocity, that lead to the complex kinetic curves, appearing as a result of summing up of the kinetics inherent to theindividual forms.  相似文献   

17.
S N Ali  H D Zeller  M K Calisto  M S Jorns 《Biochemistry》1991,30(45):10980-10986
Sarcosine oxidase contains 1 mol of covalently bound plus 1 mol of noncovalently bound FAD per active site. The first phase of the anaerobic reduction of the enzyme with sarcosine converts oxidized enzyme to an equilibrium mixture of two-electron-reduced forms (EH2) and occurs at a rate (2700 min-1, pH 8.0) similar to that determined for the maximum rate of aerobic turnover in steady-state kinetic studies (2600 min-1). The second phase of the anaerobic half-reaction converts EH2 to the four-electron-reduced enzyme (EH4) and occurs at a rate (k = 350 min-1) which is 7-fold slower than aerobic turnover. Reaction of EH2 with oxygen is 1.7-fold faster (k = 4480 min-1) than aerobic turnover and 13-fold faster than the anaerobic conversion of EH2 to EH4. The results suggest that the enzyme cycles between fully oxidized and two-electron-reduced forms during turnover with sarcosine. The long wavelength absorbance observed for EH2 is attributable to a flavin biradical (FADH.FAD.-) which is generated in about 50% yield at pH 8.0 and in nearly quantitative yield at pH 7.0. The rate of biradical formation is determined by the rate of electron transfer from sarcosine to the noncovalent flavin since electron equilibration between the two flavins (k = 750 s-1 or 45,000 min-1, pH 8.0) is nearly 20-fold faster, as determined in pH-jump experiments. Only two of the three possible isoelectronic forms of EH2 are likely to transfer electrons to oxygen since the reaction is known to occur at the covalent flavin. However, equilibration among EH2 forms is probably maintained during reoxidation, consistent with the observed monophasic kinetics, since interflavin electron transfer is 10-fold faster than electron transfer to oxygen.  相似文献   

18.
The structures of two forms of a recombinant flavoprotein have been determined at high resolution and compared. These proteins are (1) the flavocytochrome c p-cresol methylhydroxylase (rPCMH, 1.85 A resolution) and (2) the cytochrome-free flavoprotein subunit of rPCMH (PchF, 1.30 A resolution). A significant conformational difference is observed in a protein segment that is in contact with the re face of the isoalloxazine ring of FAD when the structure of PchF is compared to the subunit in the intact flavocytochrome. This structural change is important for optimum catalytic function of the flavoprotein, which has been shown to be dependent on the presence of the cytochrome subunit. This change results in different protein-flavin and apparently different protein-substrate interactions that have a "tuning effect" on the electronic and redox properties of bound p-cresol and the covalently bound FAD. The conformational change in the segment in the cofactor-binding site is induced by a small rearrangement in the flavoprotein-cytochrome interface region of the flavoprotein.  相似文献   

19.
G S Greaney  G N Somero 《Biochemistry》1979,18(24):5322-5332
The effect of anions on the thermodynamic activation functions for a model enzyme, calf intestinal alkaline phosphatase (EC 3.1.3.1), have been studied in order to examine the role of protein hydration changes in establishing the energetics of enzyme catalysis. The influences of these anions on the activation volume (delta V) and activation free energy (delta G) reflected clear Hofmeister (lyotropic) series effects, in the order F- greater than Cl- greater than Br- greater than I- (order of increasing salting-out potential). A pronounced covariation was observed between the influences of these anions on Vmax, which is proportional to delta G, and on the negative activation volume of the reaction. Fluoride was able to counteract the influences of Br- and I- on both Vmax and delta V when combinations of these anions were employed. The effects of Br- and I- on Vmax and delta V were more pronounced at lower temperatures. The control delta V was increasingly negative at reduced temperatures. The effects of the neutral salts and propanol on delta V and delta G, as well as the effects of salting-in anions on the activation enthalpy and the negative activation entropy of the reaction, are consistent with a model which proposes that peptide groups or polar side chains on the native enzyme exergonically increase their exposure to solvent during the catalytic activation event. These conclusions are in accord with the known free energy, enthalpy, entropy, and volume changes which occur when model peptide groups are transferred between water and concentrated salt solutions. Consistent with the kinetic results, the fluorescence emission wavelength maximum of alkaline phosphatase increased in the presence of anions in the order F- greater than Cl- greater than Br- greater than I-. The salting-out ion (F-) and the salting-in ions (Br- and I-) shifted lambda max in different directions, and these lambda max shifts could be counterbalanced by using equimolar combinations of salting-in and salting-out anions. Control experiments with a model compound, N-acetyltryptophanamide, showed that the spectra shifts caused by the salts did not result solely from differential quenching by the anions of the solvent-exposed tryptophan(s) on the enzyme. Hofmeister additivity phenomena indicated that the solvent is at the basis of these salt-induced enzyme structural changes. It is concluded that changes in protein solvation during enzymic reactions contribute significantly to the thermodynamic activation parameters in both the native and the salt-perturbed enzyme.  相似文献   

20.
Human adenosine deaminase. Distribution and properties.   总被引:20,自引:0,他引:20  
Adenosine deaminase exists in multiple molecular forms in human tissue. One form of the enzyme appears to be "particulate". Three forms of the enzyme are soluble and interconvertible with apparent molecular weights of approximately 36,000, 114,000, and 298,000 (designated small, intermediate, and large, respectively). The small form of adenosine deaminase is convertible to the large form only in the presence of a protein, which has an apparent molecular weight of 200,000 and has no adenosine deaminase activity. This conversion of the small form of the enzyme to the large form occurs at 4 degrees, exhibits a pH optimum of 5.0 to 8.0, and is associated with a loss of conversion activity. The small form of the enzyme predominates in tissue preparations exhibiting the higher enzyme-specific activities and no detectable conversion activity. The large form of adenosine deaminase predominates in tissue extracts exhibiting the lower enzyme specific activities and abundant conversion activity. The small form of adenosine deaminase shows several electrophoretic variants by isoelectric focusing. The electrophoretic heterogeneity observed with the large form of the enzyme is similar to that observed with the small form, with the exception that several additional electrophoretic variants are uniformly identified. No organ specificity is demonstrable for the different electrophoretic forms. The kinetic characteristics of the three soluble molecular species of adenosine deaminase are identical except for pH optimum, which is 5.5 for the intermediate species and 7.0 to 7.4 for the large and small forms.  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