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.
2.
The fructose-1,6-bis(phosphate) aldolase isologous tetramer tightly associates through two different subunit interfaces defined by its 222 symmetry. Both single- and double-interfacial mutant aldolases have a destabilized quaternary structure, but there is little effect on the catalytic activity. These enzymes are however thermolabile. This study demonstrates the temperature-dependent dissociation of the mutant enzymes and determines the dissociation free energies of both mutant and native aldolase. Subunit dissociation is measured by sedimentation equilibrium in the analytical ultracentrifuge. At 25 degrees C the tetramer-dimer dissociation constants for each single-mutant enzyme are similar, about 10(-6) M. For the double-mutant enzyme, sedimentation velocity experiments on sucrose density gradients support a tetramer-monomer equilibrium. Furthermore, sedimentation equilibrium experiments determined a dissociation constant of 10(-15) M3 for the double-mutant enzyme. By the same methods the upper limit for the dissociation constant of wild-type aldolase A is approximately 10(-28) M3, which indicates an extremely stable tetramer. The thermodynamic values describing monomer-tetramer and dimer-tetramer equilibria are analyzed with regard to possible cooperative interaction between the two subunit interfaces.  相似文献   

3.
1. The transient-state kinetics of ligand-displacement reactions have been analyzed. Methods based on this analysis have been used to obtain reliable estimates of on-velocity and off-velocity constants for coenzyme binding to liver alcohol dehydrogenase at different pH values between 6 and 10. 2. The rate of NADH dissociation from the enzyme shows no pronounced dependence on pH. The rate of NAD+ dissociation is controlled by a group with a pKa of 7.6, agreeing with the pKa reported to regulate the binding of certain inhibitory substrate analogues to the enzyme . NAD+ complex. 3. Critical experiments have been performed to test a recent proposal that on-velocity constants for the binding of NADH and NAD+ are controlled by proton equilibria exhibiting different pKa values. The results show that association rates for NADH and NAD+ exhibit the same pH dependence corresponding to a pKa of 9.2. Titrimetric evidence is presented indicating that the latter effect of pH derives from ionization of a group which affects the anion-binding capacity of the coenzyme-binding site.  相似文献   

4.
A pyrrolo[3,2-d]pyrimidine-based type-II vascular endothelial growth factor receptor 2 (VEGFR2) kinase inhibitor, compound 20d, displayed time-dependent inhibition of the non-phosphorylated catalytic domain of VEGFR2. In contrast, 20d did not show time-dependent inhibition of the phosphorylated enzyme. Dissociation of 20d from non-phosphorylated VEGFR2 was slow and the half-life of the complex was longer than 4h. In contrast, dissociation of 20d from the phosphorylated enzyme was very fast (half-life <5min). A fluorescent tracer based displacement assay and surface plasmon resonance (SPR) analysis confirmed the slow dissociation of 20d from only non-phosphorylated VEGFR2. Thus, activity based and binding kinetic analyses both supported slow dissociation of 20d from only non-phosphorylated VEGFR2. Additionally SPR analysis revealed that association rates were rapid and nearly identical for these two phosphorylation forms of VEGFR2. From these results, the preferential effect of 20d on non-phosphorylated VEGFR2 is dominated by its slow dissociation from the enzyme and this characteristically long residence time may increase its potency in vivo. The present findings may assist in the design of novel type-II kinase inhibitors.  相似文献   

5.
W T Jenkins 《Biochimie》1989,71(4):405-410
A reinvestigation of the effects of pH and salt concentration on the proton and dicarboxylic acid dissociation constants of pig heart aspartate aminotransferase shows that both anions and cations were involved concomitantly, both as stoichiometric reactants and bioenergetically. Equations are presented which can be used experimentally, to determine the numbers of salt ions (their thermodynamic stoichiometries) involved in biochemical equilibria such as proton and ligand dissociations from macromolecules. These equations were used to reinvestigate the effects of salts on the chromophoric pKa of the enzyme prosthetic group, the interaction of the enzyme with dicarboxylic acids, and the overall equilibrium for the transamination half-reaction.  相似文献   

6.
The pH-induced reversible dissociation of pigeon liver malic enzyme (EC 1.1.1.40) was studied by combined use of chemical cross-linking and SDS/polyacrylamide-gel electrophoresis. The tetrameric enzyme showed a pH-dependent dissociation in an acidic environment. At pH values above 8.0 most molecules existed as tetramers. The enzyme was gradually dissociated at lower pH. When the pH was below 5.0 most of the enzyme was present as the monomeric forms. Reassociation of the subunits was accomplished by adjusting the pH to neutrality. The dissociation and reassociation were almost instantaneous. No trimer was detected. The pigeon liver malic enzyme was thus shown to have a double-dimer quaternary structure with D2 symmetry. In the presence of substrates, the monomer-dimer-tetramer equilibrium favours the direction of dissociation. Tartronate, an L-malate analogue, was found to be more effective than L-malate in this process. When the monomeric forms were immobilized, the enzyme subunits were found to be fully active in catalysis. A possible arrangement of the four identical subunits of the enzyme molecule is proposed to account for the results obtained in this investigation. The origin of the half-of-the-sites reactivity of pigeon liver malic enzyme is also discussed.  相似文献   

7.
A two‐way soft resolution method will fail when applied to a simultaneous equilibria system due to rank deficiency in its concentration profiles. Increasing the dimensionality of measurements from two‐way to three‐way data can be used to overcome this problem. Simultaneous dissociation of two weak acids is considered as a model for simultaneous equilibria. Three‐way data obtained from excitation–emission spectrofluorimetric monitoring of a pH‐metric titration is analyzed using a proper combination of well‐known soft‐modeling methods. Multivariate curve resolution‐alternating least squares is used for calculating the excitation and emission spectral profiles of involved species and rank annihilation factor analysis for obtaining the contribution of each species in measured excitation–emission matrices at different pHs. The results of simulated and real simultaneous acids dissociation equilibria showed that the proposed combined method performs well even in situation when the equilibrium constants are close to each other. The applicability of method for study of an acidic dissociation is also shown. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

8.
The state of oligomerization of the tumor suppressor p53 is an important factor in its various biological functions. It has a well-defined tetramerization domain, and the protein exists as monomers, dimers and tetramers in equilibrium. The dissociation constants between oligomeric forms are so low that they are at the limits of measurement by conventional methods in vitro. Here, we have used the high sensitivity of single-molecule methods to measure the equilibria and kinetics of oligomerization of full-length p53 and its isolated tetramerization domain, p53tet, at physiological temperature, pH and ionic strength using fluorescence correlation spectroscopy (FCS) in vitro. The dissociation constant at 37 °C for tetramers dissociating into dimers for full-length p53 was 50 ± 7 nM, and the corresponding value for dimers into monomers was 0.55 ± 0.08 nM. The half-lives for the two processes were 20 and 50 min, respectively. The equivalent quantities for p53tet were 150 ± 10 nM, 1.0 ± 0.14 nM, 2.5 ± 0.4 min and 13 ± 2 min. The data suggest that unligated p53 in unstressed cells should be predominantly dimeric. Single-molecule FCS is a useful procedure for measuring dissociation equilibria, kinetics and aggregation at extreme sensitivity.  相似文献   

9.
The dimeric formaldehyde dehydrogenase from bovine liver has been resolved into three nearly homogeneous enzyme forms by the successive use of ion-exchange, affinity, and ampholine (chromatofocusing) chromatography. The different enzyme species were isolated in the approximate proportions 3:2:1, having pI values of 6.5, 6.2, and 6.0, respectively. The subunit molecular weights of the three forms are all similar (Mr congruent to 41,000), on the basis of sodium dodecyl sulfategel electrophoresis. The enzyme species appear to arise from covalent differences unrelated either to partial proteolysis during isolation or to differential sialization of homodimeric protein. Human liver contains a single major form and two minor forms of formaldehyde dehydrogenase having pI values very similar to those found for the bovine liver enzyme. The macroscopic kinetic constants (V, V/K) for the three forms of the dehydrogenase from bovine liver are all similar in magnitude, using NADH and S-hydroxymethylglutathione as substrates. The isotope-sensitive hydride transfer step is not significantly rate-limiting during catalysis by any of the forms, as evidenced by the near-unity primary deuterium isotope effects on both V and V/KS (for S-hydroxymethylglutathione); catalysis may be limited by the rate of dissociation of at least one (and possibly both) of the product molecules. In support of rate-limiting dissociation of NAD+ in the normal reaction, V increases by approximately 22-fold and isotope effects of approximately 1.4 are observed on both V and V/KS, using the coenzyme analog 3-acetylpyridine adenine dinucleotide. Product dissociation from the active site appears to be accelerated by the presence of dilute denaturing agents, perhaps indicative of a rate-limiting conformational transition associated with product release.  相似文献   

10.
The specific functional roles of various parts of the third transmembrane segment (M3) of the sarcoplasmic reticulum Ca(2+)-ATPase were examined by functionally characterizing a series of mutants with multiple or single substitutions of M3 residues. Steady-state and transient kinetic measurements, assisted by computer simulation of the time and Ca(2+) dependences of the phosphorylation level, were used to study the partial reaction steps of the enzyme cycle, including the binding and dissociation of Ca(2+) at the high affinity cytoplasmically facing sites. The mutation Lys-Leu-Asp-Glu(255) --> Glu-Ile-Glu-His resulted in a conspicuous increase in the rate of Ca(2+) dissociation as well as a displacement of the major conformational equilibria of the phosphoenzyme and dephosphoenzyme forms. The point mutant Phe(256) --> Ala also showed an increased rate of Ca(2+) dissociation, whereas a conspicuous decrease both in the rate of Ca(2+) dissociation and in the rate of Ca(2+) binding was found for the mutant Gly-Glu-Gln-Leu(260) --> Ile-His-Leu-Ile. These findings suggest that the NH(2)-terminal half of M3 is involved in control of the gateway to the Ca(2+) sites. The main effect of two mutations to the COOH-terminal half of M3, Ser-Lys-Val-Ile-Ser(265) --> Thr-Gly-Val-Ala-Val and Leu-Ile-Cys-Val-Ala-Val-Trp-Leu-Ile(274) --> Phe-Leu-Gly-Val-Ser-Phe-Phe-Ile-Leu, was a block of the dephosphorylation.  相似文献   

11.
In this work, we have postulated a comprehensive and unified chemical mechanism of action for yeast alcohol dehydrogenase (EC 1.1.1.1, constitutive, cytoplasmic), isolated from Saccharomyces cerevisiae. The chemical mechanism of yeast enzyme is based on the integrity of the proton relay system: His-51....NAD+....Thr-48....R.CH2OH(H2>O)....Zn++, stretching from His-51 on the surface of enzyme to the active site zinc atom in the substrate-binding site of enzyme. Further, it is based on extensive studies of steady-state kinetic properties of enzyme which were published recently. In this study, we have reported the pH-dependence of dissociation constants for several competitive dead-end inhibitors of yeast enzyme from their binary complexes with enzyme, or their ternary complexes with enzyme and NAD+ or NADH; inhibitors include: pyrazole, acetamide, sodium azide, 2-fluoroethanol, and 2,2,2-trifluorethanol. The unified mechanism describes the structures of four dissociation forms of apoenzyme, two forms of the binary complex E.NAD+, three forms of the ternary complex E.NAD+.alcohol, two forms of the ternary complex E.NADH.aldehyde and three binary complexes E.NADH. Appropriate pKa values have been ascribed to protonation forms of most of the above mentioned complexes of yeast enzyme with coenzymes and substrates.  相似文献   

12.
When estimating binding affinities of a ligand, which can exists in multiple forms, for a target molecule, one must consider all possible competing equilibria. Here, a method is presented that estimates the contribution of the protonation equilibria of a ligand in solution to the measured or calculated binding affinity. The method yields a correction to binding constants that are based on the total concentration of inhibitor (the sum of all ionized forms of the inhibitor in solution) to account for the complexed form of the inhibitor only. The method is applied to the calculation of the difference in binding affinity of two inhibitors, 2‐phosphoglycolate (PGA) and its phoshonate analog 3‐phosphonopropionate (3PP), for the glycolytic enzyme triosephosphate isomerase. Both inhibitors have three titrating sites and exist in solution as a mixture of different forms. In this case the form that actually binds to the enzyme is present at relative low concentrations. The contributions of the alternative forms to the difference in binding energies is estimated by means of molecular dynamics simulations and corrections. The inhibitors undergo a pKa shift upon binding that is estimated by ab initio calculations. An interesting finding is that the affinity difference of the two inhibitors is not due to different interactions in the active site of the enzyme, but rather due to the difference in the solvation properties of the inhibitors. Protein 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

13.
Quantitative analysis of the time courses of fluorescence anisotropy changes due to the binding of fructose-1,6-bisphosphate aldolase to the dissociable cytoplasmic glycerol-3-phosphate dehydrogenase covalently labelled with fluorescent dye was carried out. The behaviour of the aldolase-dehydrogenase system seems to be consistent with a cyclic reversible model characterized by the formation and dissociation of complexes of both the monomeric and the dimeric forms of dehydrogenase with aldolase, and rapid equilibrium between the free monomeric and dimeric forms of dehydrogenase. The half-life time of the formation of dimeric dehydrogenase-aldolase complex at the concentration of the enzymes expected to exist in the cell (i.e. in the micromolar range) is some minutes, and the time needed for equilibration between the aldolase-bound dimeric and monomeric forms of dehydrogenase is a few minutes as well. Consequently, one may expect that both the formation and the dissociation of this heterologous enzyme complex have physiological relevance.  相似文献   

14.
The soluble form of dopamine beta-hydroxylase from bovine adrenal medulla has previously been shown to exist as a tetrameric species of Mr = 290,000 composed of two disulfide-linked dimers. Here we report that this enzyme can also undergo a reversible tetramerdimer dissociation which is dependent on pH. Gel permeation chromatography of dopamine beta-hydroxylase at pH 5.0 demonstrates a Stokes radius of 5.8 nm. When the pH is shifted to 5.7, the Stokes radius changes to 6.9 nm. Sedimentation equilibrium analysis of the purified enzyme demonstrates that this change in molecular size is due to a change in molecular weight. At low protein concentration, the estimated Mr of the enzyme is 145,000 at pH 5.0 and at high protein concentration approaches 290,000 at pH 5.7. This change in Mr is consistent with the existence of a tetramer-dimer dissociation and a change in the equilibrium constant from 1.8 X 10(-6) M to 1.16 X 10(-9) M when the pH is increased from 5.0 to 5.7. This pH-dependent subunit dissociation is correlated with pH-dependent changes in enzyme activity. Purified bovine-soluble dopamine beta-hydroxylase activity is a hyperbolic function of tyramine concentration at pH 5.0. However, the hydroxylase activity displays non-hyperbolic kinetics at pH 6.0. The kinetic data obtained at pH 6.0 can be accounted for by fitting to a model containing two nonidentical catalytic forms of enzyme generated by the pH-dependent partial dissociation of tetrameric enzyme to dimeric subunits. The two catalytic forms have apparently identical maximal velocities; however, they differ in their Michaelis constants for the substrate; the dimeric form having a low Km and the tetrameric form having a high Km. Since the pH inside bovine adrenal medullary chromaffin granules is approximately 5.5, we conclude that the subunits of dopamine beta-hydroxylase are in dynamic dissociation in a physiologically important pH range.  相似文献   

15.
Arylsulfatase B, purified to homogeneity from human eosinophils, is a tetrameric enzyme whose activity varied in accordance with the state of association of its monomeric subunits. The rate of dissociation of oligomeric forms was slow relative to the rate of the enzymatic reaction so that the kinetic properties of the enzyme depended on the concentration of the enzyme before assay. For concentrated enzyme solutions (14 micrograms/ml), Lineweaver-Burk analysis demonstrated substrate inhibition at greater than or equal to 20 mM substrate and revealed two distinct regions of activity at low and intermediate substrate concentrations. The addition of bovine serum albumin (60 micrograms/ml) or sucrose (0.25 M), which prevent subunit dissociation, yielded a linear relationship on Lineweaver-Burk analysis at non-inhibitory substrate concentrations. For dilute enzyme concentrations (4.7 micrograms/ml), inhibition occurred at greater than or equal to 2 mM substrate. Nanomolar amounts of leukotriene C4 (LTC4), relative to millimolar concentrations of substrate, inhibited eosinophil arylsulfatase B. On Lineweaver-Burk analysis, the pattern of inhibition of LTC4 with concentrated enzyme was compatible with competitive inhibition of only one oligomeric form of the enzyme, whereas at low enzyme concentrations the pattern of inhibition was apparently competitive. These findings suggest that LTC4 is a potent competitive inhibitor of a dissociated, possibly dimeric, form of the enzyme. Nanomolar concentrations of LTC4, leukotriene D4, and leukotriene E4 were equally inhibitory, whereas leukotriene B4 and isomeric 5,12-dihydroxyeicosatetraenoic acids had no inhibitory activity, indicating a requirement for a thiopeptide at C-6. Thiopeptide leukotriene analogs without an intact triene structure also lacked inhibitory activity. Sulfoxide analogs of LTC4 and leukotriene D4 were potent inhibitors, although two sulfone analogs of leukotriene D4 were not inhibitory. Arylsulfatase B did not inactivate the spasmogenic activity of sulfidopeptide leukotrienes. These findings indicate that sulfidopeptide leukotrienes and their sulfoxide derivatives may regulate by competitive inhibition the activity of oligomeric forms of the eosinophil lysosomal hydrolase, arylsulfatase B.  相似文献   

16.
The molecular weights of different aggregational states of phosphoenolpyruvate carboxylase purified from the leaves of Zea mays have been determined by measurement of the molecular diameter using a Malvern dynamic light scattering spectrometer. Using these data to identify the monomer, dimer, tetramer, and larger aggregate(s) the effect of pH and various ligands on the aggregational equilibria of this enzyme have been determined. At neutral pH the enzyme favored the tetrameric form. At both low and high pH the tetramer dissociated, followed by aggregation to a "large" inactive form. The order of dissociation at least at low pH appeared to be two-step: from tetramer to dimers followed by dimer to monomers. The monomers then aggregate to a large aggregate, which is inactive. The presence of EDTA at pH 8 protected the enzyme against both inactivation and large aggregate formation. Dilution of the enzyme at pH 7 at room temperature results in driving the equilibrium from tetramer to dimer. The presence of malate with EDTA stabilizes the dimer as the predominant form at low protein concentrations. The presence of the substrate phosphoenolpyruvate alone and with magnesium and bicarbonate induced formation of the tetramer, and decreased the dissociation constant (Kd) of the tetrameric form. The inhibitor malate, however, induced dissociation of the tetramer as evidenced by an increase in the Kd of the tetramer.  相似文献   

17.
We have determined the low temperature EPR spectra and room temperature ligand dissociation rate constants of human NO-hemoglobins F and A as a function of pH and inositol hexaphosphate levels in order to assess the contribution of a quaternary structural equilibrium in the two proteins to their spectral and functional properties. Our results are consistent with an increased stability of a ligated low affinity structure in the fetal protein; the functional properties of this structure appear to be essentially the same in the two hemoglobins, even though its stability relative to a high affinity conformation is different. The pH dependence of the NO dissociation constant in both adult and fetal hemoglobin can be assigned primarily to the pH-dependent equilibria of high and low affinity forms as monitored by EPR.  相似文献   

18.
A novel inactivation mechanism of the NAD-dependent hydrogenase from Alcaligenes eutrophus Z1 comprising redox-dependent steps is described. The model of the hydrogenase inactivation process is proposed which implies that the enzyme may exist in several forms which differ in their stability and spectral properties. One of these forms, existing within a limited (approximately -200 +/- 30 mV) potential range, undergoes a rapid and irreversible inactivation. The dissociation of the FMN prosthetic group from the apohydrogenase appears to be the main reason for the enzyme inactivation. The rationale for the enzyme stabilization under real operational conditions based on the chemical modification of the hydrogenase molecule is suggested.  相似文献   

19.
The influence that fructose 2,6-bisphosphate (Fru-2,6-BP) has on the aggregation properties of rat liver phosphofructokinase has been studied by observing the fluorescence polarization of the enzyme covalently bound to the fluorescent probe pyrenebutyric acid. Fru-2,6-BP dramatically slows the dissociation of the high molecular weight aggregate forms of the enzyme when the enzyme is diluted to 3.2 micrograms/ml (4 X 10(-8) M subunits). Furthermore, Fru-2,6-BP is a strong promoter of reassociation to tetramer and larger forms if the enzyme has been previously allowed to dissociate to the dimer in its absence. Unlike many other positive effectors of liver phosphofructokinase, Fru-2,6-BP is also able to overcome the tendency of MgATP to promote tetramer formation and instead stabilize a very high degree of high molecular weight aggregate formation even in the presence of MgATP. The apparent affinity of liver phosphofructokinase for Fru-2,6-BP was measured by its ability to promote reassociation and compared to that for Fru-1,6-BP. The apparent dissociation constant for Fru-2,6-BP under these conditions is 36 microM, about 40-fold lower than the value of 1.4 mM measured for Fru-1,6-BP. Both ligands demonstrate synergism with the substrate Fru-6-P, which can lower the dissociation constant for Fru-2,6-BP 9-fold to 4 microM and that for Fru-1,6-BP 5-fold to 0.28 mM. These data are interpreted to suggest that influencing the aggregation state of rat liver phosphofructokinase may be one way in which Fru-2,6-BP achieves its effects on the enzyme in vivo.  相似文献   

20.
Both solution and crystallographic studies suggest that the 4'-aminopyrimidine ring of the thiamin diphosphate coenzyme participates in catalysis, likely as an intramolecular general acid-base catalyst via the unusual 1',4'-iminopyrimidine tautomer. It is indeed uncommon for a coenzyme to be identified in its rare tautomeric form on its reaction pathways, yet this has been possible with thiamin diphosphate, in some cases even in the absence of substrate [Nemeria, N., Chakraborty, S., Baykal, A., Korotchkina, L., Patel, M. S., and Jordan, F. (2007) Proc. Natl. Acad. Sci. U.S.A. 104, 78-82.]. The ability to detect both the aminopyrimidine and iminopyrimidine tautomeric forms of thiamin diphosphate on enzymes has enabled us to assign the predominant tautomeric form present in individual intermediates on the pathway. Herein, we report the pH dependence of these tautomeric forms providing the first data for the internal thermodynamic equilibria on thiamin diphosphate enzymes for the various ionization and tautomeric forms of this coenzyme on four enzymes: benzaldehyde lyase, benzoylformate decarboxylase, pyruvate oxidase, and the E1 component of the human pyruvate dehydrogenase multienzyme complex. Evidence is provided for an important function of the enzyme environment in altering both the ionization and tautomeric equilibria on the coenzyme even prior to addition of substrate. The pKa for the 4'-aminopyrimidinium moiety coincides with the pH for optimum activity thereby ensuring that all ionization states and tautomeric states are accessible during the catalytic cycle. The dramatic influence of the protein on the internal equilibria also points to conditions under which the long-elusive ylide intermediate could be stabilized.  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