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.
NADP-dependent malate dehydrogenase (decarboxylating) from sugar cane leaves was inhibited by increasing the ionic strength in the assay medium. The inhibitory effect was higher at pH 7.0 than 8.0, with median inhibitory concentrations (IC50) of 89 mM and 160 mM respectively, for inhibition by NaCl. Gel-filtration experiments indicated that the enzyme dissociated into dimers and monomers when exposed to high ionic strength (0.3 M NaCl). By using the enzyme-dilution approach in the absence and presence of 0.3 M NaCl, the kinetic properties of each oligomeric species of the protein was determined at pH 7.0 and 8.0. Tetrameric, dimeric and monomeric structures were shown to be active but with different V and Km values. The catalytic efficiency of the oligomers was tetramer greater than dimer greater than monomer, and each quaternary structure exhibited higher activity at pH 8.0 than 7.0. Dissociation constants for the equilibria between the different oligomeric forms of the enzyme were determined. It was established that Kd values were affected by pH and Mg2+ levels in the medium. Results suggest that the distinct catalytic properties of the different oligomeric forms of NADP-dependent malate dehydrogenase and changes in their equilibrium could be the molecular basis for an efficient physiological regulation of the decarboxylation step of C4 metabolism.  相似文献   

2.
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   

3.
The crystal structure of dimeric D-amino acid transaminase shows that the two Trp-139 sites are located in a hydrophobic pocket at the interface between the subunits and that the two indole side chains face one another and are within 10 A of coenzyme. This enzyme prefers an aromatic character at position 139, as previously demonstrated by the finding that Phe-139 but no other substitution tested provides the maximum degree of thermostability and catalytic efficiency. Here we show that an equilibrium between active dimers and inactive monomers can be demonstrated with the W139F mutant enzyme, whereas with the wild-type enzyme the subunit interface is so tight that a study of this equilibrium is precluded. We show how the processes of dimerization of monomers and dissociation of dimers to monomers are controlled. Lower pH (5.0) favors monomer formation from dimers. Gel filtration and activity analysis show that at higher pH (7.0) the monomers combine to form active dimers with a K(d) of 0.17 microM. This assembly process is relatively slow and takes several hours for completion, thereby permitting accurate measurement of kinetics and equilibrium parameters. Absorption and circular dichroism spectra of dimers and monomers are significantly different, indicating that the environment around the cofactor is very likely altered between them. The circular dichroism peak of the W139F dimer at 418 nm is less negative than that of the wild-type enzyme in accordance with its lower visible absorbance; the circular dichroism peak of the W139F monomer at 418 nm is more negative than that of the wild-type enzyme. The dissociation of dimers to monomers has also been studied by taking advantage of these spectral differences, thus permitting the rates of the dissociation and the reassociation to be calculated and compared. 2-Mercaptoethanol assists in the conversion of monomers to dimers. The results here describe dissociation/reassociation in the dimeric enzyme under native conditions without denaturants.  相似文献   

4.
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   

5.
In dilute solution phosphoenolpyruvate carboxylase of Escherichia coli undergoes a spontaneous inactivation that can be described mathematically by a two-component declining exponential equation. The rate constant for the decay of the first component is 3.05 ± 0.52 × 10?2 min?, whereas that for the second component is variable, smaller in magnitude, and dependent upon the dilution conditions. Analysis of the coefficients for the exponential equation suggests that the decline of enzymatic activity with time is a function of the initial concentrations of catalytically active dimer and tetramer. From the concentrations of these two species, as determined from their initial activities, an equilibrium constant of 3 × 10?7m for the tetramer-dimer dissociation was determined.The diluted enzyme exhibits properties similar to those ascribed to hysteretic enzymes. The appearance of hysteresis is a function of the time after dilution and the presence of modifiers of catalytic activity, i.e., it is not present immediately after dilution and can be prevented from occurring if aspartate is present in the dilution buffer. The data are consistent with a scheme in which dimeric and tetrameric forms of the enzyme undergo inactivation by dissociation to monomers. The tetramer can dissociate directly to monomers and become inactivated or it can dissociate first to dimers than to monomers before undergoing inactivation. Monomer-to-dimer reassociation occurs to form a catalytically active species, but monomer-to-tetramer reassociation to an active species is not apparent. Hysteresis is presumed to result from reversible isomerization of the monomeric species to a form that can also result in an irreversibly inactivated enzyme.  相似文献   

6.
Analytical ultracentrifugation studies performed on spinach chloroplast fructose bisphosphatase show that the tetrameric oxidized (inactive) or reduced (active) enzyme dissociates into inactive dimers and monomers at alkaline pH. The dissociation process is, at least, partially reversible if the enzyme is dimeric. Moreover, the oxidized inactive tetrameric enzyme is less prone to dissociation into dimers and monomers than the reduced active tetramer. The irreversibility of the dissociation process may be explained by a sulfhydryl-disulfide interchange. Together with the findings from previously published sulfhydryl group titration experiments (J. Pradel et al., Eur. J. Biochem., 113 (1981) 507), the above results suggest that the activation of the oxidized tetramer involves the reduction of two inter-protomeric disulfide bonds.  相似文献   

7.
Monomers and dimers of bovine heart cytochrome c oxidase (EC 1.9.3.1.) were separated by gel filtration chromatography on Ultrogel AcA 34 or by sucrose gradient centrifugation. Factors influencing the interconversion of the two aggregation states of this enzyme were analyzed. At very low ionic strength, in the presence of dodecyl maltoside, monomers were the main species. Salts appeared to stabilize the dimeric form, divalent cations being more efficient than monovalent. High enzyme concentrations favoured the formation of dimers, also at low ionic strength. The type of detergent had a strong influence on the monomer-dimer interconversion; in Triton X-100 and dodecyl maltoside (at high ionic strength) cytochrome c oxidase was homogenously dispersed in its dimeric form, while in Tween-80 gel filtration showed only large particles eluting in the void volume. In cholate monomers and aggregates were observed but no dimers. The aggregation state had an influence on the steady state kinetics of the ferrocytochrome c oxidase activity. Monomers showed linear Eadie-Hofstee plots, whilst the dimeric and aggregated enzyme gave nonlinear Eadie-Hofstee plots. Ionic strength, enzyme concentration and type of detergent were affecting the enzyme's kinetics in a way consistent with the molecular form obtained by the gel filtration or sedimentation analysis. The data support a negative cooperative mechanism for the interaction of cytochrome c with the dimeric enzyme, as proposed earlier (K.A. Na?ecz et al., (1983) Biochem. Biophys. Res. Commun., 114, 822-828).  相似文献   

8.
Binding of ligands to the catalytic center of mammalian triosephosphate isomerase (TPI) induces a conformational change(s) that enhances the specific deamidation of Asn71 at the subunit interface. Deamidation initiates dissociation and degradation of the enzyme in vivo and in vitro. We have utilized circular dichroism spectroscopy to examine the conformational changes in the enzyme upon ligand binding and subunit dissociation/reassociation. Native TPI from rabbit, chicken, and yeast exhibit similar spectra at pH 7.5, but are substantially different at pH 9.5. Covalent reaction of the active site Glu 165 with the substrate analogue 3-chloroacetol phosphate results in a conformational change (decrease in beta-sheet) which is similar in TPI from all three species. Reversible dissociation of the dimeric enzyme in guanidine followed by dialysis, although permitting full recovery of catalytic activity, results in refolded dimers with decreased alpha-helix. These conformational changes induced by ligand binding, pH, or reversible dissociation explain, in part, the differences in the chemical and physical properties of the enzyme from the three species at alkaline pH, the increased lability of the dissociated/reassociated enzyme, and corroborate 31P NMR data on substrate-induced conformational changes. These studies also support the concept of molecular wear and tear whereby ligand binding at the catalytic center induces conformational changes that increase the probability of covalent modification and ultimate degradation of the protein.  相似文献   

9.
The reversible inactivation and dissociation of the allosteric phosphofructokinase from Escherichia coli has been studied in relatively mild conditions, i.e., in the presence of the chaotropic agent KSCN. At moderate KSCN concentration, the loss of enzymatic activity involves two separated phases: first, a rapid dissociation of part of the tetramer into dimers, second, a slower displacement of the dimer-tetramer equilibrium upon further dissociation of the dimer into monomers. These two reactions can no longer be distinguished above 0.3 M KSCN since complete inactivation occurs in a single reaction. Different changes are observed for the fluorescence and the activity of the enzyme in KSCN: the fluorescence is not affected by the dissociation into dimers which is responsible for inactivation. The decrease in fluorescence reflects the change in environment of the unique tryptophan residue, Trp 311, during the dimer to monomer dissociation. This residue belongs to the interface containing the regulatory site, and its native fluorescence indicates that this interface is still present in the dimer. The substrate fructose 6-phosphate protects phosphofructokinase from inactivation by binding to the tetramer and prevents its dissociation into dimers. The presence of phosphoenolpyruvate prevents the slow dissociation of the dimer into monomers, which shows the ability of the dimer to bind the inhibitor. Two successive processes can be observed during reassociation of the protein upon KSCN dilution. First, a fast reaction (k1 = 2 x 10(5) M-1.s-1) is accompanied by a fluorescence increase and results in the formation of the dimeric species.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Among proteins utilized as sweeteners, neoculin and miraculin are taste-modifying proteins that exhibit pH-dependent sweetness. Several experiments on neoculin have shown that His11 of neoculin is responsible for pH dependence. We investigated the molecular mechanism of the pH dependence of neoculin by molecular dynamics (MD) calculations. The MD calculations for the dimeric structures of neoculin and His11 mutants showed no significant structural changes for each monomer at neutral and acidic pH levels. The dimeric structure of neoculin dissociated to form isolated monomers under acidic conditions but was maintained at neutral pH. The dimeric structure of the His11Ala mutant, which is sweet at both neutral and acidic pH, showed dissociation at both pH 3 and 7. The His11 residue is located at the interface of the dimer in close proximity to the Asp91 residue of the other monomer. The MD calculations for His11Phe and His11Tyr mutants demonstrated the stability of the dimeric structures at neutral pH and the dissociation of the dimers to isolated monomers. The dissociation of the dimer caused a flexible backbone at the surface that was different from the dimeric interface at the point where the other monomer interacts to form an oligomeric structure. Further MD calculations on the tetrameric structure of neoculin suggested that the flexible backbone contributed to further dissociation of other monomers under acidic conditions. These results suggest that His11 plays a role in the formation of oligomeric structures at pH 7 and that the isolated monomer of neoculin at acidic pH is responsible for sweetness.  相似文献   

11.
Data from sedimentation analysis suggest that modification of about 40% of free amino groups of inorganic pyrophosphatase by maleic anhydride, pH 10.5, results in a loss of the enzyme ability to form dimers at neutral values of pH. The specific activity of monomeric pyrophosphatase is 50-80% of that of the dimeric form. The monomer has a pH optimum of about 7, requires metal ions for activation of both enzyme and substrate and is capable of exergonic synthesis of PPi in the active center. The enzyme binding to PPi is strongly stabilized by fluoride. The experimental data indicate that the individual subunit of inorganic pyrophosphatase possesses all the main catalytic properties of native dimeric molecule.  相似文献   

12.
A method was developed for the purification of phosphoenolpyruvate carboxylase from darkened maize leaves so that the enzyme retained its sensitivity to inhibition by malate. The procedure depended on the prevention of proteolysis by the inclusion of chymostatin in the buffers used during the purification. The purified enzyme was indistinguishable from that in crude extracts as judged by native polyacrylamide-gel electrophoresis. SDS/polyacrylamide-gel electrophoresis followed by immunoblotting, and Superose 6 gel filtration. Gel-filtration studies showed that the purified enzyme and the enzyme in extracts of darkened or illuminated leaves showed a concentration-dependent dissociation of tetrameric into dimeric forms. Purified phosphoenolpyruvate carboxylase and enzyme in crude extracts from darkened leaves were equally sensitive to inhibition by malate (Ki approx. 0.30 mM) under conditions where it existed in the tetrameric or dimeric forms, but the enzyme in crude extracts from illuminated leaves was less sensitive to malate inhibition (Ki approx. 0.95 mM) whether it was present as a tetramer or as a dimer. It is concluded that changes in the oligomerization state of phosphoenolpyruvate carboxylase are not directly involved in its regulation by light.  相似文献   

13.
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   

14.
Incubation of soluble human liver arginase (L-arginine amidinohydrolase, EC 3.5.3.1) with p-hydroxymercuribenzoate resulted in the dissociation of the enzyme into active dimers. Addition of 2-mercaptoethanol resulted in the regeneration of the tetrameric enzyme. When arginase, bound covalently to nylon, was incubated with p-hydroxymercuribenzoate, matrix-bound dimers were obtained. Incubation of these species with 2-mercaptoethanol resulted in stable, unmodified dimers. Based on this dissociation of arginase, a model with D2-symmetry is suggested for this enzyme. The specific activity, the Km value for arginine, pH optimum and the inhibition constants for ornithine and lysine were determined for monomeric, dimeric and tetrameric forms. It is concluded that the behaviour of the active sites of the monomers is not substantially altered by the interaction of these species in the oligomeric molecule.  相似文献   

15.
pH,温度、离子强度及效应剂等对固定化烟草RuBP羧化酶在2.5mol/L尿素处理下的解离作用有各种不同的影响。在pH6.0时,仅小亚基从大亚基核(L_8)解离,当pH为中性偏碱时,大亚基核也解离。低温和低离子强度均促进酶的解离,而温度和离子强度对大亚基之间的解离的影响显著大于对大、小亚基之间的影响。这表明酶的亚基之间存在着不同的极性和疏水作用,而大亚基之间的疏水作用比大、小亚基之间的强。6-PG对大、小亚基之间解离的抑制作用表明大亚基上的催化位置与小亚基之间有一定的密切关系。  相似文献   

16.
D Panda  S Roy  B Bhattacharyya 《Biochemistry》1992,31(40):9709-9716
Concentration-dependent dissociation of dimers of goat brain tubulin S and tubulin was studied by fluorescence anisotropy. Upon dilution, assembly-competent fluorescein 5'-maleimide labeled dimers of tubulin S and tubulin show a progressive decrease in fluorescence anisotropy. That this lowering of anisotropy results from the dissociation of tubulin S dimers into monomers was shown by dilution experiments with unlabeled homologous and heterologous proteins. A nonlinear least-squares fit of the data gave a dissociation constant of 7.1 x 10(-8) M for tubulin S compared to 7.2 x 10(-7) M for tubulin at 25 degrees C in 0.1 M PEM buffer, pH 7.0. van't Hoff plots of dimer-monomer dissociation of tubulin S and tubulin also show considerable differences in delta H and delta S. Effects of ionic strength and colchicine on the equilibrium constants are also substantially different for tubulin and tubulin S. The implications of these observations on the influence of C-terminal tails on tubulin structure are discussed.  相似文献   

17.
Umbilicus rupestris (pennywort) switches from C3 photosynthesis to an incomplete form of crassulacean acid metabolism (referred to as 'CAM-idling') when exposed to water stress (drought). This switch is accompanied by an increase in the activity of phosphoenolpyruvate carboxylase. This enzyme also shows several changes in properties, including a marked decrease in sensitivity to acid pH, a lower Km for phosphoenolpyruvate, very much decreased sensitivity to the allosteric inhibitor malate, and increased responsiveness to the allosteric effector glucose 6-phosphate. The Mr of the enzyme remains unchanged, at approx. 185 000. These changes in properties of phosphoenolpyruvate carboxylase are discussed in relation to the roles of the enzyme in C3 and in CAM plants.  相似文献   

18.
Maize (Zea mays L.) leaf phosphoenopyruvate (PEP) carboxylase activity at subsaturating levels of PEP was increased by the inclusion of glycerol (20%, v/v) in the assay medium. The extent of activation was dependent on H+ concentration, being more marked at pH 7 (with activities 100% higher than in aqueous medium) than at pH 8 (20% activation). The determination of the substrate concentration necessary to achieve half-maximal enzyme activity (S0.5) (PEP) and maximal velocity (V) between pH 6.9 and 8.2 showed a uniform decrease in S0.5 in the presence of glycerol over the entire pH range tested, and only a slight decrease in V at pH values near 8. Including NaCl (100 millimolar) in the glycerol containing assay medium resulted in additional activation, mainly due to an increase in V over the entire range of pH. Glucose-6-phosphate (5 millimolar) activated both the native and the glycerol-treated enzyme almost to the same extent, at pH 7 and 1 millimolar PEP. Inhibition by 5 millimolar malate at pH 7 and subsaturating PEP was considerably lower in the presence of glycerol than in an aqueous medium (8% against 25%, respectively). Size-exclusion high performance liquid chromatography in aqueous buffer revealed the existence of an equilibrium between the tetrameric and dimeric enzyme forms, which is displaced to the tetramer as the pH was increased from 7 to 8. In the presence of glycerol, only the 400 kilodalton tetrameric form was observed at pH 7 or 8. However, dissociation into dimers by NaCl could not be prevented by the polyol. We conclude that the control of the aggregation state by the metabolic status of the cell could be one regulatory mechanism of PEP carboxylase.  相似文献   

19.
1. Incubation of prolyl 4-hydroxylase (prolyl-glycyl-peptide, 2-oxoglutarate : oxygen oxidoreductase (4-hydroxylating), EC 1.14.11.2) with H2O2 leads to a decrease of 50% in the specific activity of enzyme tetramers, followed by dissociation into inactive dimers in which the monomers are covalently cross-linked by S-S bridge formation. 2. Incubation of the enzyme with K3Fe(CN)6 leads to a comparable decrease in activity of enzyme tetramers. Addition of urea leads to dissociation into inactive dimers with similarly cross-linked monomers. 3. Removal of the dissociating agent leads to reassociation of cross-linked dimers to tetramers and to about 50% reactivation. The enzyme is further reactivated by preincubation with dithiothreitol. 4. Dissociation of the enzyme with dithiothreitol, urea or LiCl, or at low pH (4.15) produces inactive monomers, which could not be reassociated.  相似文献   

20.
Porcine heart citrate synthase, a dimeric protein of Mr = 100,000 composed of two identical subunits, is shown to undergo a monomer-dimer equilibrium. The extent of dimerization is found to be dependent on the concentration of citrate synthase, pH, ionic strength, and the specific buffer system employed. Oxaloacetate and citrate, substrates for the forward and reverse reaction catalyzed by citrate synthase, affect dimerization at concentrations of the protein which exists as monomer in their absence. The dissociation of citrate synthase dimers has been demonstrated utilizing the techniques of gel permeation chromatography, fluorescence polarization, fluorescence energy transfer, and heat denaturation. Earlier studies of citrate synthase quarternary structure found the protein to be nondissociable except under denaturing conditions or extensive modification; however, most former studies were performed at relatively high protein concentration, ionic strength, and pH, conditions which stabilize the dimer. In light of recent evidence derived from x-ray crystallographic studies showing amino acid residues from one subunit contributing to the citrate and CoA binding sites of the other, the dissociation into monomers would be expected to have profound effects on citrate synthase activity and regulation, as well as overall tricarboxylic acid cycle activity.  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