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.
Human liver arginase (L-arginine amidinohydrolase, EC 3.5.3.1) was immobilised by attachment to nylon with glutaraldehyde as a crosslinking agent. Incubation of the immobilised tetrameric enzyme with EDTA followed by dialysis resulted in the dissociation of the enzyme into inactive matrix-bound and solubilised subunits. Both species recovered enzymatic activity after incubation with Mn2+, and the activity of the reactivated matrix-bound subunits was nearly 25% of that shown by the enzyme initially attached to the support in the tetrameric form. When the reactivated bound subunits were incubated with soluble subunits in the presence of Mn2+, they 'picked-up' from the solution an amount of protein and enzymatic activity almost identical to that initially lost by the immobilised tetramer after the dissociating treatment with EDTA. This occurred only in the presence of Mn2+. It is suggested that the reactivation of the subunits of arginase involves the initial formation of an active monomer, which then acquires a conformation that favours a reassociation to the tetrameric state.  相似文献   

2.
The specific activity of chicken liver pyruvate carboxylase has been shown to decrease with decreasing enzyme concentration, even at 100 microM, which is close to the estimated physiological concentration. The kinetics of the loss of enzyme specific activity following dilution were biphasic. Incubation of dilution-inactivated enzyme with ATP, acetyl CoA, Mg2+ + ATP or, to a lesser degree, with Mg2+ alone resulted in a high degree of reactivation, while no reactivation occurred in the presence of pyruvate. The association state of the enzyme before, during, and after dilution inactivation has been assessed by gel filtration chromatography. These studies indicate that on dilution, there is dissociation of the catalytically active tetrameric enzyme species into inactive dimers. Reactivation of the enzyme resulted in reassociation of enzymic dimers into tetramers. The enzyme was shown to form high molecular weight aggregates at high enzyme concentrations.  相似文献   

3.
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   

4.
Experimental conditions favouring the dissociation of tetrameric rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase into active monomers were elaborated. The urea-induced dissociation of the tetramer was shown to be a stepwise process (in 2 M urea only dimers are formed; an increase in urea concentration up to 3 M causes the splitting of the dimers into monomers). The specific activity of immobilized monomers in the glyceraldehyde-3-phosphate oxidation reaction does not differ from that of the parent immobilized tetrameric form. The tetrameric enzyme molecule binds the coenzyme with a negative cooperativity (the first two NAD+ molecules bind with KD below 0.1 microM; for the third and fourth molecules the dissociation constant was determined to be equal to 5.5 +/- 1.5 microM (50 mM medinal buffer, 10 mM sodium phosphate, pH 8.2). The cooperativity of NAD+ binding is preserved in the immobilized preparation of tetrameric dehydrogenase. The immobilized monomers bind NAD+ with KD of 1.6 +/- 1.0 microM. The experimental results are consistent with the hypothesis according to which the association of catalytically active subunits into a tetramer changes their coenzyme-binding properties in such a way that the first two NAD+ molecules bind more firmly to a tetramer than to a monomer, whereas the third and the fourth NAD+ molecules bind less firmly.  相似文献   

5.
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   

6.
β-Galactosidase or lactase is a very important enzyme in the food industry, being that from the yeast Kluyveromyces lactis the most widely used. Here we report its three-dimensional structure both in the free state and complexed with the product galactose. The monomer folds into five domains in a pattern conserved with the prokaryote enzymes of the GH2 family, although two long insertions in domains 2 and 3 are unique and related to oligomerization and specificity. The tetrameric enzyme is a dimer of dimers, with higher dissociation energy for the dimers than for its assembly. Two active centers are located at the interface within each dimer in a narrow channel. The insertion at domain 3 protrudes into this channel and makes putative links with the aglycone moiety of docked lactose. In spite of common structural features related to function, the determinants of the reaction mechanism proposed for Escherichia coli β-galactosidase are not found in the active site of the K. lactis enzyme. This is the first X-ray crystal structure for a β-galactosidase used in food processing.  相似文献   

7.
Cytoplasmic serine hydroxymethyltransferase (cSHMT) is a tetrameric, pyridoxal phosphate (PLP)-dependent enzyme that catalyzes the reversible interconversion of serine and tetrahydrofolate to glycine and methylenetetrahydrofolate. The enzyme has four active sites and is best described as a dimer of obligate dimers. Each monomeric subunit within the obligate dimer contributes catalytically important amino acid residues to both active sites. To investigate the interchange of subunits among cSHMT tetramers, a dominant-negative human cSHMT enzyme (DNcSHMT) was engineered by making three amino acid substitutions: K257Q, Y82A, and Y83F. Purified recombinant DNcSHMT protein was catalytically inactive and did not bind 5-formyltetrahydrofolate. Coexpression of the cSHMT and DNcSHMT proteins in bacteria resulted in the formation of heterotetramers with a cSHMT/DNcSHMT subunit ratio of 1. Characterization of the cSHMT/DNcSHMT heterotetramers indicates that DNcSHMT and cSHMT monomers randomly associate to form tetramers and that cSHMT/DNcSHMT obligate dimers are catalytically inactive. Incubation of recombinant cSHMT protein with recombinant DNcSHMT protein did not result in the formation of hetero-oligomers, indicating that cSHMT subunits do not exchange once the tetramer is assembled. However, removal of the active site PLP cofactor does permit exchange of obligate dimers among preformed cSHMT and DNcSHMT tetramers, and the formation of heterotetramers from cSHMT and DNcSHMT homodimers does not affect the activity of the cSHMT homodimers. The results of these studies demonstrate that PLP inhibits dimer exchange among cSHMT tetramers and suggests that cellular PLP concentrations may influence the stability of cSHMT protein in vivo.  相似文献   

8.
We have shown here that the cytosolic bacterial chaperone SecB is a structural dimer of dimers that undergoes a dynamic equilibrium between dimer and tetramer in the native state. We demonstrated this equilibrium by mixing two tetrameric species of SecB that can be distinguished by size. We showed that the homotetrameric species exchanged dimers, because when the mixture was analyzed both by size exclusion chromatography and native polyacrylamide gel electrophoresis a third hybrid tetrameric species was detected. Furthermore, treatment of SecB with 5,5'-dithiobis-(2-nitrobenzoic acid), which modifies the sulfhydryl group on cysteines, caused irreversible dissociation to a dimer indicating that cysteine must be involved in the stabilizing interactions at the dimer interface. It is clear that the two dimer-dimer interfaces of the SecB tetramer are differentially stable. Dissociation at one interface allows for a dynamic dimer-tetramer equilibrium. Because only dimers were exchanged it is clear that the other interface between dimers is significantly more stable, otherwise oligomers should have formed with a random distribution of monomers.  相似文献   

9.
The influence of ATP and KCl on the quaternary structure and the enzymatic activity of D-glyceraldehyde-3-phosphate dehydrogenase from yeast(Y-GAPDH) has been studied by ultracentrifugation, gel chromatography and standard optical tests. In 0.1 M imidazole buffer pH 7.0, at low temperature (0°C) both complete deactivation and dissociation to dimers occur in the presence of 2 mM ATP and 0.1 M 2-mercaptoethanol. In 0.067 M phosphate buffer pH 7.0, containing 2 mM ATP and 1 mM dithiothreitol, only slight deactivation paralleled by minor changes of the native quaternary structure is observed. In this same buffer, increasing temperature leads to stabilization of both the tetrameric state and the catalytic activity of the enzyme. Deactivation and dissociation in the presence of 0.15 M KCl (in 0.2 M glycine buffer 9.1 ≥ pH ≥ 8.0) is a function of pH rather than electrolyte concentration; at neutral pH the enzyme is stabilized in its native state. Contrary to earlier assumptions in the literature, ATP and KCl under the above experimental conditions do not appear to play an important role in the in vivo regulation of Y-GAPDH.  相似文献   

10.
Sergienko EA  Jordan F 《Biochemistry》2002,41(19):6164-6169
The tetrameric enzyme yeast pyruvate decarboxylase (YPDC) has been known to dissociate into dimers at elevated pH values. However, the interface along which the dissociation occurs, as well as the fundamental kinetic properties of the resulting dimers, remains unknown. The active sites of YPDC are comprised of amino acid residues from two subunits, a property which we utilize to address the issue as to which dimer interface is cleaved under different conditions of dissociation. Hydroxide-induced dissociation of the active site D28A (or D28N) and E477Q variants, each at least 100 times less reactive than wild-type YPDC, followed by reassociation of D28A (or D28N) and E477Q variants led to a remarkable 35-50-fold increase in activity. This result is possible only if the hydroxide-induced dissociation results in a cleavage along the interface between two subunits so that residues D28 and E477 are now separated. Upon reassociation, one of the two active sites of the hybrid dimer will have both residues substituted, whereas the second one will be of the wild-type phenotype. In contrast to the hydroxide-induced dimers, the urea-induced dissociation recently proposed results in dissociation along dimer-dimer interfaces, without separating the active sites, and therefore, on reassociation, these dimers do not regain activity. The significance of the results is discussed in light of a recently proposed alternating sites mechanism for YPDC. A preparative ion-exchange method is reported for the separation and purification of hybrid enzymes.  相似文献   

11.
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   

12.
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   

13.
Inactivation of apo-glyceraldehyde-phosphate dehydrogenase (D-glyceraldehyde-3-phosphate: NAD+ oxidoreductase(phosphorylating) (EC 1.2.1.12) from rat skeletal muscle at 4 degrees C in 0.15 M NaC1, 5 mM EDTA, 4 mM 2-mercaptoethanol pH 7.2 is a first-order reaction. The rate constant of inactivation depends on protein concentration. With one molecule of NAD bound per tetrameric enzyme, a 50 per cent loss in activity is observed and the rate constant of inactivation becomes independent of the protein concentration over a 30-fold range. Two moles of NAD bound per mole of enzyme fully protect it against inactivation. NADH affords a cooperative effect on enzyme structure similar to that of NAD. Inactivation of 7.8 S apoenzyme is reflected in its dissociation into 4.8-S dimers. In the case of enzyme-NAD1 complex, no direct relationship between the extent of inactivation and dissociation is observed, suggesting that these two processes do not occur simultaneously; we may say that dissociation is slower than inactivation. A mechanism in which the rate-limiting step for inactivation is a conformational change in the tetramer occurring prior to dissociation and affecting only the structure of the non-liganded dimer, is consistent with the experimental observations. Inorganic phosphate protects apoenzyme against inactivation. Its effect is shown to be due to the anion binding at specific sites on the protein with a dissociation constant of 2.6 plus or minus 0.4 mM. The NaC1-induced cold inactivation of glyceraldehyde-phosphate dehydrogenase is fully reversible at 25 degrees C in the presence of 20 mM dithiothreitol and 50 mM inorganic phosphate. The rate of reactivation is independent of protein concentration. Inactivated enzyme retains the ability to bind specific antibodies produced in rabbits, but diminishes its precipitating capability.  相似文献   

14.
Unlike other beta-class eukaryotic DNA polymerases, the enzyme purified from the Novikoff hepatoma is inhibited by both sulfhydryl blocking agents N-ethylmaleimide (NEM) and p-hydroxymercuribenzoate (pHMB). The degree of sensitivity varies depending on the enzyme purity, pH of the reaction, and the presence of sulfhydryl reducing agents. Novikoff beta-polymerase activity is unaffected by the presence of 2-mercaptoethanol (2-Me) or dithiothreitol (DTT); however, the combination of 2-mercaptoethanol and NEM or pHMB acts to reverse the inhibition of the sulfhydryl blocking agent. The reversal of inhibition involves more than just a titration of NEM with 2-mercaptoethanol since a) the combination of these two reagents actually stimulates the DNA polymerase, and b) dithiothreitol did not reverse the inhibition. Binding of the polymerase to DNA did not affect the enzyme sensitivity to NEM.  相似文献   

15.
The unfolding and dissociation of the tetrameric enzyme fructose-1,6-bisphosphatase from pig kidney by guanidine hydrochloride have been investigated at equilibrium by monitoring enzyme activity, ANS binding, intrinsic (tyrosine) protein fluorescence, exposure of thiol groups, fluorescence of extrinsic probes (AEDANS, MIANS), and size-exclusion chromatography. The unfolding is a multistate process involving as the first intermediate a catalytically inactive tetramer. The evidence that indicates the existence of this intermediate is as follows: (1) the loss of enzymatic activity and the concomitant increase of ANS binding, at low concentrations of Gdn.HCl (midpoint at 0.75 M), are both protein concentration independent, and (2) the enzyme remains in a tetrameric state at 0.9 M Gdn.HCl as shown by size-exclusion chromatography. At slightly higher Gdn.HCl concentrations the inactive tetramer dissociates to a compact dimer which is prone to aggregate. Further evidence for dissociation of tetramers to dimers and of dimers to monomers comes from the concentration dependence of AEDANS-labeled enzyme anisotropy data. Above 2.3 M Gdn.HCl the change of AEDANS anisotropy is concentration independent, indicative of monomer unfolding, which also is detected by a red shift of MIANS-labeled enzyme emission. At Gdn.HCl concentrations higher than 3.0 M, the protein elutes from the size-exclusion column as a single peak, with a retention volume smaller than that of the native protein, corresponding to the completely unfolded monomer. In the presence of its cofactor Mg(2+), the denaturated enzyme could be successfully reconstituted into the active enzyme with a yield of approximately 70-90%. Refolding kinetic data indicate that rapid refolding and reassociation of the monomers into a nativelike tetramer and reactivation of the tetramer are sequential events, the latter involving slow and small conformational rearrangements in the refolded enzyme.  相似文献   

16.
《FEBS letters》1985,192(2):215-219
Pyruvate carboxylase purified from Rhizopus arrhizus exhibits projections when examined in the electron microscope which indicate that this enzyme is a tetrameric molecule in which the subunits are arranged at the corners of a tetrahedron. The tetrameric molecule is stabilised by addition of acetyl-CoA or of pyruvate but is labilised in the presence of 2-oxoadipate. Addition of EDTA causes a decrease in the stability of the tetrameric molecule with a time course similar to that observed for loss of acetyl-CoA-dependent catalytic activity [(1984) FEBS Lett. 127, 157-160]. The data suggest that the hysteretic responses induced by exposure to EDTA are associated with dissociation of the tetrameric molecule to dimers and monomers having a decreased sensitivity to allosteric activation.  相似文献   

17.
Tetrameric D-glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12) isolated from rabbit skeletal muscle was covalently bound to CNBr-activated Sepharose 4B via a single subunit. Catalytically active immobilized dimer and monomeric forms of the enzyme were prepared after urea-induced dissociation of the tetramer. A study of the coenzyme-binding properties of matrix-bound tetrameric, dimeric and monomeric species has shown that: (1) an immobilized tetramer binds NAD+ with negative cooperativity, the dissociation constants being 0.085 microM for the first two coenzyme molecules and 1.3 microM for the third and the fourth one; (2) coenzyme binding to the dimeric enzyme form also displays negative cooperativity with Kd values of 0.032 microM and 1.1 microM for the first and second sites, respectively; (3) the binding of NAD+ to a monomer can occur with a dissociation constant of 1.6 microM which is close to the Kd value for low-affinity coenzyme binding sites of the tetrameric or dimeric enzyme forms. In the presence of NAD+ an immobilized monomer acquires a stability which is not inferior to that of a holotetramer. The catalytic properties of monomeric and tetrameric enzyme forms were compared and found to be different under certain conditions. Thus, the monomers of rabbit muscle D-glyceraldehyde-3-phosphate dehydrogenase displayed a hyperbolic kinetic saturation curve for NAD+, whereas the tetramers exhibited an intermediary plateau region corresponding to half-saturating concentrations of NAD+. At coenzyme concentrations below half-saturating a monomer is more active than a tetramer. This difference disappears at saturating concentrations of NAD+. Immobilized monomeric and tetrameric forms of D-glyceraldehyde-3-phosphate dehydrogenase from baker's yeast were also used to investigate subunit interactions in catalysis. The rate constant of inactivation due to modification of essential arginine residues in the holoenzyme decreased in the presence of glyceraldehyde 3-phosphate, probably as a result of conformational changes accompanying catalysis. This effect was similar for monomeric and tetrameric enzyme forms at saturating substrate concentrations, but different for the two enzyme species under conditions in which about one-half of the active centers remained unsaturated. Taken together, the results indicate that association of D-glyceraldehyde-3-phosphate dehydrogenase monomers into a tetramer imposes some constraints on the functioning of the active centers.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
The effect of 2,3-diphospho-D-glycerate on the sedimentation coefficient of carbon monoxide hemoglobin was correlated with the fraction of rapidly reacting hemoglobin observed subsequent to flash photolysis at 23 degrees C at pH 7.30 in buffers of 0.1 M ionic strength. Concentrations of the organic phosphate up to about 5 mM resulted in an increase in S20,w, consistent with an increase in the fraction of tetrameric hemoglobin. A decrease in rapidly reacting hemoglobin parallelled the increase in the sedimentation coefficient. Between 5 and 20 mM 2,3-diphosphoglycerate, S20,w decreased, suggesting that dissociation to dimers was enhanced. An increase in rapidly reacting hemoglobin was also observed in this concentration range. Similar sedimentation results were obtained with oxyhemoglobin at pH 7.00 and carbon monoxide hemoglobin at pH 7.06. Assuming single binding sites on each species, the dissociation constants for 2,3-diphosphoglycerate binding to tetrameric and dimeric HbCO are 0.2-0.3 mM and 2-5 mM at pH 7.30. This biphasic effect of this physiologically important organic phosphate on the state of aggregation of R state hemoglobin has not been previously reported, but it is similar to that previously noted with inositol hexaphosphate, which enhanced tetramer formation at low concentrations, while at higher concentrations it promoted hemoglobin dissociation to dimers (White, S. L. (1976) J. Biol. Chem. 251, 4763-4769; Gray, R. D. (1980) J. Biol. Chem. 255, 1812-1818).  相似文献   

19.
The structure of the active center of glyceraldehyde-3-phosphate dehydrogenase and the arrangement of subunits in the tetrameric molecule is delineated. The mechanism of cooperative effects in the oligomer is considered, and the involvement of various regions of the active center and of different-subunit contact area in the realization of the cooperative phenomena is discussed. A special attention is paid to the effect of NAD+ bound to one of the subunits of the tetramer on the structure of an adjacent subunit and to the problem of the participation of the coenzyme in the creation of anion-binding sites of the enzyme. The conditions of reversible dissociation of the tetrameric apoenzyme molecule into dimers are depicted, and the role of NAD+ in the organization of the quaternary structure of the dehydrogenase is discussed. The problem of catalytic activity of the dimeric form of the enzyme is argued.  相似文献   

20.

Background  

Oligomeric enzymes can undergo a reversible loss of activity at low temperatures. One such enzyme is tryptophanase (Trpase) from Escherichia coli. Trpase is a pyridoxal phosphate (PLP)-dependent tetrameric enzyme with a Mw of 210 kD. PLP is covalently bound through an enamine bond to Lys270 at the active site. The incubation of holo E. coli Trpases at 2°C for 20 h results in breaking this enamine bond and PLP release, as well as a reversible loss of activity and dissociation into dimers. This sequence of events is termed cold lability and its understanding bears relevance to protein stability and shelf life.  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