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.
Carbon monoxide binding to Chromatium vinosum ferrocytochrome c' has been studied by high-precision equilibrium methods. In contrast to the CO binding properties of Rhodospirillum molischianum cytochrome c' [Doyle, M. L., Weber, P. C., & Gill, S. J. (1985) Biochemistry 24, 1987-1991], CO binding to C. vinosum cytochrome c' is found to be unusual in the following ways. The binding curve is found to be cooperative with typical Hill coefficients equal to 1.25. The shape of the binding curve is asymmetrical. The heat of CO ligation is measured by two independent methods, both of which yield large endothermic values of approximately 10 kcal [mol of CO(aq)]-1. The overall affinity for CO increases as the concentration of cytochrome c' decreases. These observations suggest the CO binding properties of C. vinosum cytochrome c' are complicated by CO-linked association-dissociation processes. Further investigation by gel filtration chromatography shows that at micromolar concentrations the dimeric state is tightly associated in both the reduced and oxidized forms of the cytochrome but addition of saturating concentrations of CO causes the reduced ligated dimer to dissociate largely into monomers. A model is presented that quantitatively fits the data, involving a ligand-linked dimer-monomer dissociation reaction. In this model, CO binds to the dimer form noncooperatively with an intrinsic affinity constant equal to 5600 +/- 1200 M-1 at 25 degrees C. The unligated dimer form is tightly associated, but addition of CO causes dissociation of the dimer into the monomer with a monomer-dimer association constant equal to 450 +/- 200 M-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Molecular weight of native apotransketolase from baker's yeast is found to be 159000 +/- 6000 by means of sedimentation equilibrium and sedimentation-diffusion rate. The enzyme in a relatively low concentration reversibly dissociates into two subunits with molecular weight of about 80 000 at pH 7.6 and 20 degrees C. The equilibrium constant of the reaction monomer-dimer is 4.4 . 10(3) M-1. A decrease of the temperature stimulates the association of monomers into dimer, while the shift of pH 7.6 into acid or alkaline region stimulates the dissociation process. Dissociation becames irreversible at pH less than 5 and greater than 10.5.  相似文献   

3.
Receptor dimerization is important for many signaling pathways. However, the monomer-dimer equilibrium has never been fully characterized for any receptor with a 2D equilibrium constant as well as association/dissociation rate constants (termed super-quantification). Here, we determined the dynamic equilibrium for the N-formyl peptide receptor (FPR), a chemoattractant G protein-coupled receptor (GPCR), in live cells at 37°C by developing a single fluorescent-molecule imaging method. Both before and after liganding, the dimer-monomer 2D equilibrium is unchanged, giving an equilibrium constant of 3.6 copies/μm(2), with a dissociation and 2D association rate constant of 11.0 s(-1) and 3.1 copies/μm(2)s(-1), respectively. At physiological expression levels of ~2.1 receptor copies/μm(2) (~6,000 copies/cell), monomers continually convert into dimers every 150 ms, dimers dissociate into monomers in 91 ms, and at any moment, 2,500 and 3,500 receptor molecules participate in transient dimers and monomers, respectively. Not only do FPR dimers fall apart rapidly, but FPR monomers also convert into dimers very quickly.  相似文献   

4.
The dimeric Crotalus atrox venom PLA2 is part of the secreted phospholipase A2 (PLA2) enzyme family that interacts at the lipid-solution interface to hydrolyze the sn-2 acyl ester bond of phospholipids. We have employed fluorescence correlation spectroscopy (FCS) to study the monomer-dimer equilibrium of the C. atrox venom PLA2 in solution, in the presence of urea, and in the presence of monomeric and micellar n-dodecylphosphocholine (C12-PN), a phosphatidylcholine analogue. Dilution experiments show that PLA2 is an extremely tight dimer, Kd < or = 0.01 nM, in solution. Urea was introduced to weaken the subunit's association, and an estimate for the PLA(2) dimer dissociation constant in buffer was obtained by linear extrapolation. The derived dissociation constant was at least several orders of magnitude greater than that suggested from the dilution experiments, indicating a complex interaction between urea and the PLA2 dimer. FCS data indicate that the PLA2 dimer begins to dissociate at 10 mM C12-PN in 10 mM Ca2+ and at 5 mM C12-PN in 1 mM EDTA. The PLA2 tryptophan fluorescence displayed spectral shifts and intensity changes upon interacting with C12-PN. On the basis of the FCS and tryptophan fluorescence results, we postulate an intermediate state where the two monomers are in loose interaction within a protein-lipid comicelle. As the concentration of C12-PN was increased, complete dissociation of the dimer was observed, inferred from the doubling of the particle number, and the average diffusion constant decreased to approximately 60 microm2/s, consistent with PLA2 associated with a C12-PN micelle. The presence of Ca2+ makes the comicelle intermediate more stable, retarding the separation of the monomers in the micellar suspension. Our data clearly indicate that PLA2, though a strong dimer in the absence of lipids, is dissociated by micellar C12-PN and supports the monomer hypothesis for PLA2 action.  相似文献   

5.
The bacterial phosphoenolpyruvate:glycose phosphotransferase system plays an essential role in diverse physiological phenomena. To perform these functions, the system is stringently regulated, although the underlying molecular regulatory mechanisms have not been established. A potential target for this type of regulation is the first protein in the phosphotransfer sequence, Enzyme I, which catalyzes the following reaction: P-enolpyruvate + Enzyme I Mg2+ in equilibrium phospho-I + pyruvate. We reported previously that Enzyme I from Salmonella typhimurium consists of identical subunits which associate in a temperature-dependent manner; the mode of association was found to be either monomer-dimer or isodesmic. The association reaction has now been investigated by analytical gel chromatography at 8, 11, and 23 degrees C. At each temperature, the mode of association was strictly monomer-dimer. The apparent association equilibrium constant, K'a, increased dramatically with temperature, with an enthalpy of 54.8 +/- 6.3 kcal/mol. At 23 degrees C, K'a decreased slightly when the enzyme solution contained either Mg2+ or phosphoenolpyruvate. However, when both ligands were present, i.e. under conditions where Enzyme I is phosphorylated, K'a decreased significantly (25-fold at 11 degrees C and 50-fold at 23 degrees C). These results are in accord with a model for the action of Enzyme I which involves a cycle of association and dissociation. This model has potentially important implications for regulating Enzyme I and the bacterial phosphoenolpyruvate:glycose phosphotransferase system.  相似文献   

6.
The kinetics of association and dissociation for the ouabain-Na+,K+- dependent ATPase complex have been studied in intact turkey erythrocytes as a function of external Na+ concentration, K+ concentration, and temperature. At free ligand concentrations substantially exceeding the concentration of available binding sites, the association reaction exhibits pseudo-first-order kinetics with an association rate constant (k1) that is conveniently determined over a wide range of temperatures (5-37 degrees C). The dissociation reaction exhibits strict first-order kinetics with a dissociation rate constant (k-1) that has the unusual property, in the turkey cell, of being sufficiently great to permit its direct determination even at temperatures as low as 5 degrees C. Values for the equilibrium binding constant for the ouabain-ATPase complex (KA) predicted from the ratio of the association and dissociation rate constants agree closely with independently measured values of KA determined directly under conditions of equilibrium binding. KA is a sensitive function of the composition of the external ionic environment, rising with increasing Na+ concentration and falling with increasing K+ concentration. These changes in KA are shown to be quantitatively attributable to changes in the rate constant k1, k-1 in contrast being unaffected at any given temperature by even very large changes in Na+ or K+ concentration. Arrhenius plots of k1 and k-1 both yield straight lines over the entire temperature range corresponding to activation energies for association and dissociation of 29.5 and 24.2 kcal/mol, respectively. These observations have made it possible to calculate the following standard values for the ouabain binding reaction in the presence of 150 mM Na+: delta G degree = -9.8 kcal/mol; delta H degree = +5.3 kcal/mol; delta S degree = +48.7 cal/degree/mol. The large positive value of delta S degree presumably reflects a highly ordered configuration of the ouabain-free ATPase molecule that is lost upon ouabain binding and that "drives" the reaction despite the positive value of delta H degree.  相似文献   

7.
Barbar E  Kleinman B  Imhoff D  Li M  Hays TS  Hare M 《Biochemistry》2001,40(6):1596-1605
Cytoplasmic dynein is a multisubunit ATPase that transforms chemical energy into motion along microtubules. LC8, a 10 kDa light chain subunit of the dynein complex, is highly conserved with 94% sequence identity between Drosophila and human. The precise function of this protein is unknown, but its ubiquitous expression and conservation suggest a critical role in the function of the dynein motor complex. We have overexpressed LC8 from Drosophila melanogaster and characterized its dimerization and folding using analytical ultracentrifugation, size-exclusion chromatography, circular dichroism, and fluorescence spectroscopy. Sedimentation equilibrium measurements of LC8 at pH 7 reveal a reversible monomer-dimer equilibrium with a dissociation constant of 12 microM at 4 degrees C. At lower pH, LC8 dissociates to a monomer, with a transition midpoint at pH 4.8. Far-UV CD and fluorescence spectra demonstrate that pH-dissociated LC8 retains native secondary and tertiary structures, while the diminished near-UV CD signal shows loss of quaternary structure. The observation that dimeric LC8 dissociates at low pH can be explained by titration of a histidine pair in the dimer interface. Equilibrium denaturation experiments with a protein concentration range spanning almost 2 orders of magnitude indicate that unfolding of LC8 dimer is a two-stage process, in which global unfolding is preceded by dissociation to a folded monomer. The nativelike tertiary structure of the monomer suggests a role for the monomer-dimer equilibrium of LC8 in dynein function.  相似文献   

8.
Interaction of flavin mononucleotide (FMN) with dimeric and tetrameric forms of rabbit muscle glycogen phosphorylase beta has been studied under the conditions when allosteric activator binding sites are saturated by AMP (1 mM AMP; pH 6.8; 17 degrees C). Simultaneous use of schlieren optical system and photoelectric scanning absorption optical system of analytical ultracentrifuge Spinco, model E, makes it possible to register the oligomeric state of the enzyme and calculate the degree of saturation of individual oligomeric enzyme forms by FMN. The apparent association constant for the equilibrium dimer in equilibrium with tetramer decreased with increasing FMN concentration. The microscopic dissociation constants for the complexes of dimeric and tetrameric forms of glycogen phosphorylase beta with FMN have been found to be equal to 10 and 79 microM, respectively.  相似文献   

9.
Self-association of bovine serum albumin (BSA) was explored using fluorescence resonance energy transfer (FRET) between two populations of the protein labeled separately with either fluorescein-5'-isothiocyanate (FITC) or eosin-5'-isothiocyanate (EITC). The energy transfer reached the steady state after 5 s at 25 degrees C, indicating a fast exchange between oligomer subunits. The dependence of the energy transfer efficiency on the protein concentration and its reversion by unlabeled BSA demonstrate that association between BSA monomers occurs through a reversible path that involves specific interactions between the protein molecules. Because energy transfer took place even after blocking Cys 34 with iodoacetamide, this residue might not be involved in the reversible self-association process. The number of subunits forming the oligomer and its dissociation constant were determined from measurements of energy transfer as a function of the donor-acceptor ratio and of the total protein concentration. Analysis of these data indicated that BSA is in a monomer-dimer equilibrium with a dissociation constant of 10 +/- 2 microM at 25 degrees C in 10 mM MOPS-K (pH 5.8).  相似文献   

10.
The influence of the constitutive metal ions on the equilibrium properties of dimeric Photobacterium leiognathi Cu,Zn superoxide dismutase has been studied for the wild-type and for two mutant protein forms bearing a negative charge in the amino acid clusters at the dimer association interface. Depletion of copper and zinc dissociates the two mutant proteins into monomers, which reassemble toward the dimeric state upon addition of stoichiometric amounts of zinc. Pressure-dependent dissociation is observed for the copper-depleted wild-type and mutated enzymes, as monitored by the fluorescence shift of a unique tryptophan residue located at the subunit association interface. The spectral shift occurs slowly, reaching a plateau after 15-20 minutes, and is fully reversible. The recovery of the original fluorescence properties, after decompression, is fast (less than four minutes), suggesting that the isolated subunit has a relatively stable structure, and excluding the presence of stable intermediates during the dimer-monomer transition. The dimer dissociation process is still incomplete at 6.5 kbar for the copper-depleted wild-type and mutated enzymes, at variance with what is generally observed for oligomeric proteins that dissociate below 3 kbar. Measurement of the degree of dissociation, at two different protein concentrations, allows us to calculate the standard volume variation upon association, Delta V, and the dissociation constant K(d0), at atmospheric pressure, (25 ml/mol and 3 x 10(-7)M, respectively). The holoprotein is fully dimeric even at 6.5 kbar, which allows us to evaluate a lower Delta G degrees limit of 11.5 kcal/mol, corresponding to a dissociation constant K(d0)<10(-9)M.  相似文献   

11.
It has been previously reported that the structure of a human pancreatic ribonuclease variant, namely PM8, constitutes a dimer by the exchange of an N-terminal domain, although in an aqueous solution it is found mainly as a monomer. First, we investigated the solution conditions that favour the dimerization of this variant. At 29 degrees C in a 20% (v/v) ethanol buffer, a significant fraction of the protein is found in dimeric form without the appearance of higher oligomers. This dimer was isolated by size-exclusion chromatography and the dimerization process was studied. The dissociation constant of this dimeric form is 5 mm at 29 degrees C. Analysis of the dependence of the dimerization process on the temperature shows that unlike bovine pancreatic ribonuclease, a decrease in the temperature shifts the monomer-dimer equilibrium to the latter form. We also show that a previous dissociation of the exchangeable domain from the main protein body does not take place before the dimerization process. Our results suggest a model for the dimerization of PM8 that is different to that postulated for the dimerization of the homologous bovine pancreatic ribonuclease. In this model, an open interface is formed first and then intersubunit interactions stabilize the hinge loop in a conformation that completely displaces the equilibrium between nonswapped and swapped dimers to the latter one.  相似文献   

12.
R Koren  G G Hammes 《Biochemistry》1976,15(5):1165-1171
Kinetic studies have been carried out of the monomer-dimer interaction of insulin, beta-lactoglobulin, and alpha-chymotrypsin using stopped-flow and temperature-jump techniques. The pH indicators bromothymol blue, bromophenol blue, and phenol red were used to monitor pH changes associated with the monomer-dimer interaction. In all three cases a kinetic process was observed which could be attributed to a simple monomer-dimer equilibrium, and association (k1) and dissociation (k-1) rate constants were determined. The results obtained are as follows: for insulin at 23 degrees C, pH 6.8, 0.125 M KNO3, k1 = 1.14 X 10(8) M-1 s-1, k-1 - 1.48 X 10(4)s(-1); for beta-lactoglobulin AB at 35 degrees C, pH 3.7, 0.025 M KNO3, d1 = 4.7 X 10(4) M-1 s-1, k-1 = 2.1 s-1; for alpha-chymotrypsin at 25 degreesC, pH 4.3, 0.05 M KNO3 k1 - 3.7 X 10(3) M-1 s-1, k-1 - 0.68 s-1. The kinetic behavior of the separated beta-lactoglobulin A and B was similar to that of the mixture. In the case of chymotrypsin, bromophenol blue was found to activate the enzyme catalyzed hydrolysis of p-nitrophenyl acetate, and a rate process was observed with the temperature jump which could be attributed to a conformational change of the indicator-protein complex. The association rate constant for dimer formation of insulin approaches the value expected for a diffusion-controlled process, while the values obtained for the other two proteins are below those expected for a diffusion-controlled reaction unless unusally large steric and electrostatic effects are present.  相似文献   

13.
The Apple 4 (A4) domain of human plasma factor XI (FXI) was used to investigate the process of FXI noncovalent dimer formation. Recombinant 6-histidine-tagged A4 domain proteins were prepared utilizing a bacterial expression system. Purification was accomplished under denaturing conditions, followed by a refolding protocol to facilitate correct disulfide bond formation. Analysis of the A4 domain (C321S mutant) by size exclusion chromatography indicated the presence of a slowly equilibrating reversible monomer-dimer equilibrium. The elution profiles reveal highly symmetrical peaks for both dimeric and monomeric species with elution times that were highly reproducible for varying amounts of both the dimeric and monomeric species. The monomer-dimer equilibrium was found to be dependent upon changes in both pH and salt concentration. Under conditions approximating physiologic salt concentration and pH (20 mm HEPES, 100 mm NaCl, and 1 mm EDTA, pH 7.4), it was determined that the monomer-dimer equilibrium was characterized by a dissociation constant (K(D)) value of 229 +/- 26 nm with a calculated Delta G value of 9.1 kcal/mol. This report identifies electrostatic contributions and the presence of a hydrophobic component that mediate interactions at the A4 domain interface. The rate of dissociation for the recombinant A4 domain C321S mutant was examined by monitoring the increase in 4,4'-dianilino-1,1'-binaphthyl-5,5'-disulfonic acid dipotassium salt fluorescence under dissociating conditions, giving a value for a dissociation rate constant (k(off)) of 4.3 x 10(-3) s(-1).  相似文献   

14.
Intact AraC protein is poorly soluble and difficult to purify, whereas its dimerization domain is the opposite. Unexpectedly, the DNA binding domain of AraC proved also to be soluble in cells when overproduced and is easily purified to homogeneity. The DNA binding affinity of the DNA binding domain for its binding site could not be measured by electrophoretic mobility shift because of its rapid association and dissociation rates, but its affinity could be measured with a fluorescence assay and was found to have a dissociation constant of 1 x 10(-8)M in 100 mM KCl. The binding of monomers of the DNA binding domain to adjacent half-sites occurs without substantial positive or negative cooperativity. A simple analysis relates the DNA binding affinities of monomers of DNA binding domain and normal dimeric AraC protein.  相似文献   

15.
Although bovine beta-lactoglobulin assumes a monomeric native structure at pH 3 in the absence of salt, the addition of salts stabilizes the dimer. Thermodynamics of the monomer-dimer equilibrium dependent on the salt concentration were studied by sedimentation equilibrium. The addition of NaCl, KCl, or guanidine hydrochloride below 1 M stabilized the dimer in a similar manner. On the other hand, NaClO(4) was more effective than other salts by about 20-fold, suggesting that anion binding is responsible for the salt-induced dimer formation, as observed for acid-unfolded proteins. The addition of guanidine hydrochloride at 5 M dissociated the dimer into monomers because of the denaturation of protein structure. In the presence of either NaCl or NaClO(4), the dimerization constant decreased with an increase in temperature, indicating that the enthalpy change (DeltaH(D)) of dimer formation is negative. The heat effect of the dimer formation was directly measured with an isothermal titration calorimeter by titrating the monomeric beta-lactoglobulin at pH 3.0 with NaClO(4). The net heat effects after subtraction of the heat of salt dilution, corresponding to DeltaH(D), were negative, and were consistent with those obtained by the sedimentation equilibrium. From the dependence of dimerization constant on temperature measured by sedimentation equilibrium, we estimated the DeltaH(D) value at 20 degrees C and the heat capacity change (DeltaC(p)) of dimer formation. In both NaCl and NaClO(4), the obtained DeltaC(p) value was negative, indicating the dominant role of burial of the hydrophobic surfaces upon dimer formation. The observed DeltaC(p) values were consistent with the calculated value from the X-ray dimeric structure using a method of accessible surface area. These results indicated that monomer-dimer equilibrium of beta-lactoglobulin at pH 3 is determined by a subtle balance of hydrophobic and electrostatic effects, which are modulated by the addition of salts or by changes in temperature.  相似文献   

16.
Dimeric proteins can arise by the swapping of structural domains between monomers. The prevalence of this occurrence is unknown. Ribonuclease A (RNase A) is assumed to be a monomer near physiological conditions. Here, this hypothesis is tested and found to be imprecise. The two histidine residues (His12 and His119) in the active site of RNase A arise from two domains (S-peptide and S-protein) of the protein. The H12A and H119A variants have 10(5)-fold less ribonucleolytic activity than does the wild-type enzyme. Incubating a 1:1 mixture of the H12A and H119A variants at pH 6.5 and 65 degrees C results in a 10(3)-fold increase in ribonucleolytic activity. A large quantity of active dimer can be produced by lyophilizing a 1:1 mixture of the H12A and H119A variants from acetic acid. At pH 6.5 and 65 degrees C, the ribonucleolytic activity of this dimer converges to that of the dimer formed by simply incubating the monomers, as expected for a monomer-dimer equilibrium. The equilibrium dissociation constant for the dimer is near 2 mM at both 65 and 37 degrees C. This value of Kd is only 20-fold greater than the concentration of RNase A in the cow pancreas, suggesting that RNase A dimers exist in vivo. The intrinsic ability of RNase A to form dimers under physiological conditions is consistent with a detailed model for the evolution of homodimeric proteins. Dimers of "monomeric" proteins could be more prevalent than is usually appreciated.  相似文献   

17.
Dissociation of purified erythrocyte Ca(2+)-ATPase by hydrostatic pressure.   总被引:2,自引:0,他引:2  
Subunit interactions in the Ca(2+)-ATPase from erythrocyte plasma membranes were investigated through a combination of fluorescence spectroscopy and high-pressure techniques. Application of hydrostatic pressure in the range of 1 bar to 2.4 kbar promoted full dissociation of the ATPase, as revealed by spectral shifts of the intrinsic fluorescence emission and by changes in the fluorescence polarization of dansyl-conjugated ATPase. Pressure dissociation of the ATPase displayed a dependence on protein concentration compatible with dissociation of a dimer. Calculated from pressure-dissociation curves, the standard volume change dV0 for the association of subunits was 43-50 ml/mol and K0, the dissociation constant at atmospheric pressure, was 6-9 x 10(-8) M. Addition of Ca2+ stabilized the dimeric ATPase structure against pressure dissociation, whereas addition of vanadate facilitated dissociation by pressure. These results suggest that intersubunit interactions depend on the equilibrium between the two major conformational states E1 and E2 of the ATPase. Addition of calmodulin in the presence of Ca2+ had no additional effect when compared to that observed in the presence of Ca2+ alone. This finding is interpreted in terms of the mechanism of calmodulin activation of ATPase catalysis.  相似文献   

18.
Thermodynamics of reversible monomer-dimer association of tubulin   总被引:1,自引:0,他引:1  
D L Sackett  R E Lippoldt 《Biochemistry》1991,30(14):3511-3517
The equilibrium between the rat brain tubulin alpha beta dimer and the dissociated alpha and beta monomers has been studied by analytical ultracentrifugation with use of a new method employing short solution columns, allowing rapid equilibration and hence short runs, minimizing tubulin decay. Simultaneous analysis of the equilibrium concentration distributions of three different initial concentrations of tubulin provides clear evidence of a single equilibrium characterized by an association constant, Ka, of 4.9 X 10(6) M-1 (Kd = 2 X 10(-7) M) at 5 degrees, corresponding to a standard free energy change on association delta G degrees = -8.5 kcal mol-1. Colchicine and GDP both stabilize the dimer against dissociation, increasing the Ka values (at 4.5 degrees C) to 20 X 10(6) and 16 X 10(6) M-1, respectively. Temperature dependence of association was examined with multiple three-concentration runs at temperatures from 2 to 30 degrees C. The van't Hoff plot was linear, yielding positive values for the enthalpy and entropy changes on association, delta S degrees = 38.1 +/- 2.4 cal deg-1 mol-1 and delta H degrees = 2.1 +/- 0.7 kcal mol-1, and a small or zero value for the heat capacity change on association, delta C p degrees. The entropically driven association of tubulin monomers is discussed in terms of the suggested importance of hydrophobic interactions to the stability of the monomer association and is compared to the thermodynamics of dimer polymerization.  相似文献   

19.
1. The fluorescence polarization, P, of FAD increased on complex formation with the apoenzyme of D-amino acid oxidase [D-amino acid: O2 ocidoreductase (deaminating), EC 1.4.3.3]. The time course of the increase was monophasic. The values of P were extimated to be 0.04, 0.4, and 0.4 for FAD, the enzyme and the enzyme-benzoate complex, respectively. 2. The value of P of the enzyme is dependent on its concentration, indicating that the degrees of dissociation of FAD in the monomer and dimer are different. The dissociation constant was calculated to be 7 times 10-minus 7 M for the monomeric form of the enzyme. This value is far larger than the value for the dimeric form of the enzyme, 1 times 10-minus 8 M, calculated from equilibrium dialysis data. 3. Changes in fluorescence polarization of the enzyme due to changes in solution pH or temperature can be explained in terms of the monomer-dimer equilibrium.  相似文献   

20.
The enzyme concentration dependence of spectrophotometric titrations of hog kidney D-amino acid oxidase [EC 1.4.3.3] with p-aminobenzoate was studied. The monomer-dimer equilibrium constant of the oxidized holoenzyme at 25 degrees C was estimated to be 7 X 10(5)M-1 at pH 7.5 and 4X 10(6)M-1 at pH 8.3. The energetics of subunit association are discussed.  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