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 46 毫秒
1.
Interaction of AMP deaminase with RNA   总被引:1,自引:0,他引:1  
tRNA, 18 S and 28 S ribosomal RNAs were found to activate muscle AMP deaminase (AMP aminohydrolase, EC 3.5.4.6) but inhibit liver and heart AMP deaminases. The macromolecular structures are essential for modulation of enzyme activity, since the effects of RNA disappeared after RNAase treatment. Sucrose density centrifugation experiments clearly demonstrated the binding of purified muscle AMP deaminase to tRNA, 18 S and 28 S RNAs. The binding is reversible and responsive to alterations of pH and KCl concentration. The binding was stable at pH 5.1-7.0 in 0.1 M KCl, but most of the enzyme dissociated at pH 7.5. KCl below 0.1 M concentration had no effect on dissociation of enzyme-RNA complex, but in 0.15 M KCl the complex was partially dissociated and in 0.2 M KCl most of the enzyme was released. Various nucleotides were also effective in dissociation of the enzyme from complex. The binding is saturable and the maximum number of muscle AMP deaminase molecules bound per mol 28 S RNA was calculated to be approx. 30. Liver and heart AMP deaminases were also found to interact with RNA.  相似文献   

2.
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   

3.
The pigeon liver fatty acid synthetase complex (14 S) is dissociated in low ionic strength buffer containing dithiothreitol to form a half-molecular weight subunits (9 S) which are completely inactive for the synthesis of saturated fatty acids. The dithiothreitol-protected (reduced) subunits are rapidly reassociated and reactivated to form the active enzyme complex, not only by an increase in salt concentration but also by micromolar concentrations of NADP+ or NADPH. Increases in KCl or NADPH concentration result in an increase in the extent of reactivation (equilibrium) with no change in the over-all rate of the reaction or the half-life ofreactivation of the enzyme. The extent (equilibrium) of reactivation of the enzyme is the same in 0.2 M potassium phosphate buffer, pH 7.0; 0.2 M KCl in 5 mM Tris-35 mM glycine buffer, PH 8.3; and 50 muM NADP+ or NADPH in the Tris-glycine buffer. The extent and rate of reactivation of the enzyme is dependent not only on ionic strength and NADPH concentration, but also on pH and temperature. Reactivation with 0.2 M KCl is optimal between pH 7.3 and 8.5. At higher and lower pH values the rate and extent of reactivation are lowered. The rate and extent of reactivation are also decreased as the temperature is lowered below 10 degrees. At 0 degrees there is little reactivation of enzyme activity. However, in the presence of 0.2 M KCl containing 15 to 40% glycerol at 0 degrees, reactivation of the enzyme is about 50% complete. The rate of reactivation of enzyme in the presence of KCl or NADPH conforms to first order kinetics. This result suggests that the subunits first combine to form an inactive complex which is subsequently transformed to an enzymatically active complex. Evidence for the presence of inactive complex was obtained in experiments carried out in 0.2 M KCl at pH 6.0, and in 0.2 M KCl at pH 8.3, at both 6 and 3 degrees. Under these conditions the amount of complex observed upon ultracentrifugation was greater than expected from determinations of enzyme activity. The above findings suggest that ionic and hydrophobic interactions, and possibly the water structure surrounding the interacting sites, are of prime importance in reassociation and reactivation of enzyme. In addition, NADP+ and NADPH have very specific effects in bringing about reassociation and in maintaining the structural integrity of the multienzyme complex.  相似文献   

4.
We have investigated the CO-recombination kinetics after flash photolysis of CO from the "half-reduced" cytochrome c oxidase as a function of pH. In addition, the reaction was investigated in mutant enzymes in which Lys(I-362) and Ser(I-299), located approximately in the middle of the K-pathway and near the enzyme surface, respectively, were modified. Laser-flash induced dissociation of CO is followed by rapid internal electron transfer from heme a(3) to a. At pH>7 this electron transfer is associated with proton release to the bulk solution (tau congruent with 1 ms at pH 8). Thus, the CO-recombination kinetics reflects protonation events at the catalytic site. In the wild-type enzyme, below pH approximately 7, the main component in the CO-recombination displayed a rate of approximately 20 s(-1). Above pH approximately 7, a slow CO-recombination component developed with a rate that decreased from approximately 8 s(-1) at pH 8 to approximately 1 s(-1) at pH 10. This slow component was not observed with KM(I-362), while with the SD(I-299)/SG(I-299) mutant enzymes at each pH it was slower than with the wild-type enzyme. The results are interpreted in terms of proton release from H(2)O in the catalytic site after CO dissociation, followed by OH(-) binding to the oxidized heme a(3). The CO-recombination kinetics is proposed to be determined by the protonation rate of OH(-) and not dissociation of OH(-), i.e. the K-pathway transfers protons and not OH(-). With the KM(I-362) mutant enzyme the proton is not released, i.e. OH(-) is not formed. With the SD(I-299)/SG(I-299) mutant enzymes the proton is released, but both the release and uptake are slowed by the mutations. During reaction of the reduced enzyme with O(2), the H(2)O at the binuclear center is most likely involved as a proton donor in the O-O cleavage reaction.  相似文献   

5.
A double mixing stopped-flow method is described for studying the dimer-tetramer equilibria of oxyhemoglobins and the kinetics of association of unliganded dimers. The three hemoglobins studied were: Hb Carp, Hb A, and Hb Rothschild (Trp beta 37 (C3)----Arg). The new method reproduces the data obtained for oxyHb A by other established methods. In agreement with previous studies, the new method indicates little, if any, dissociation of oxyHb carp into dimers even in 2 M urea solutions (0.1 M Bis-Tris pH 7.0). OxyHb Rothschild, on the other hand, is extensively dissociated into dimers (K(Hb4L4 in equilibrium with 2Hb2) = 37.3 x 10(-6) M) and the rate constant for the association of deoxy dimers of Hb Rothschild is about one-tenth of the value for Hb A indicating that the deoxy tetramer of Hb Rothschild is at least 10 times more dissociated into dimers than deoxyHb A.  相似文献   

6.
1. The hemocyanin of the bivalve, Yoldia limatula (Say) was found by light-scattering to have a mol. wt of 8.0 +/- 0.6 x 10(6). Mass measurements by scanning transmission electron microscopy (STEM) gave a particle mass of 8.25 +/- 0.42 x 10(6) for the native particle and 4.09 +/- 0.20 x 10(6) for the half-molecule. 2. The hemocyanin subunits fully dissociated in 8.0 M urea and 6.0 M GdmCl at pH 8.0, and at pH 11.0, 0.01 M EDTA have mol. wts of 4.38 x 10(5), 4.22 x 10(5) and 4.71 x 10(5), close to one-twentieth of the parent molecular weight of Y. limatula hemocyanin and most gastropod hemocyanins. 3. Analyses of the urea dissociation transitions studied at pH 8.0, 1 x 10(-2) M Mg2+, 1 x 10(-2) M Ca2+ and pH 8.0, 3 x 10(-3) M Ca2+ suggest few hydrophobic amino acid groups, of the order of 10 to 15 at the contact areas of each half-molecule or decamer. 4. The further dissociation of the decamers to dimers and the dimers to monomers indicates the presence of a larger number of amino acid groups of ca 35-40/dimer and 100-120/monomer. 5. This suggests hydrophobic stabilization of the dimer to dimer and monomer to monomer contacts within the decamers, as observed with other molluscan hemocyanins.  相似文献   

7.
Peroxidase-mediated toxicity to schistosomula of Schistosoma mansoni   总被引:16,自引:0,他引:16  
Guinea pig eosinophil peroxidase (EPO) was capable of killing schistosomula of Schistosoma mansoni in vitro when combined with hydrogen peroxide and a halide. Killing was measured by 51Cr release, by microscopic evaluation of viability, and by reinfection experiments in mice. Parasite killing was dependent on each component of the EPO-H2O2-halide system, was completely inhibited by catalase and azide, and was partially inhibited by cyanide. The EPO-mediated system required 10(-4) M H2O2 and 10(-4) M iodide at pH 7.0, and the schistosomula were killed with exposure to this system of less than 30 min at 37 degrees C. At pH 6.0, the EPO-mediated system showed significant cidal activity with 10(-6) M iodide. Canine neutrophil peroxidase (myeloperoxidase [MPO]) was also able to kill schistosomula in vitro in the presence of 10(-4) M H2O2 and 10(-4) iodide at pH 7.0 and pH 6.0. Physiologic concentrations of chloride (0.1 M) could substitute for iodide at pH 7.0 and pH 6.0 as the halide cofactor; however, at pH 7.0, a higher concentration of enzyme was required. These findings with isolated enzyme systems are compatible with a role for peroxidase in the host defense against schistosomula.  相似文献   

8.
The expression of the Paenibacillus sp. A11 cyclodextrinase (CDase) gene using the pUC 18 vector in Escherichia coli JM 109 resulted in the formation of an insoluble CDase protein in the cell debris in addition to a soluble CDase protein in the cytoplasm. Unlike the expression in Paenibacillus sp. A11, CDase was primarily observed in cytoplasm. However, by adding 0.5 M sorbitol as an osmolyte, the formation of insoluble CDase was prevented while a three-fold increase in cytoplasmic CDase activity was achieved after a 24 h-induction. The recombinant CDase protein was purified to approximately 14-fold with a 31% recovery to a specific activity of 141 units/mg protein by 40-60% ammonium sulfate precipitation, DEAE-Toyopearl 650 M, and Phenyl Sepharose CL-4B chromatography. It was homogeneous by non-denaturing and SDS-PAGE. The enzyme was a single polypeptide with a molecular weight of 80 kDa, as determined by gel filtration and SDS-PAGE. It showed the highest activity at pH 7.0 and 40 degrees C. The catalytic efficiency (k(cat)/K(m)) values for alpha-, beta-, and gamma- CD were 3.0 x 10(5), 8.8 x 10(5), and 5.5 x 10(5) M(-1) min(-1), respectively. The enzyme hydrolyzed CDs and linear maltooligosaccharides to yield maltose and glucose with less amounts of maltotriose and maltotetraose. The rates of hydrolysis for polysaccharides, soluble starch, and pullulan were very low. The cloned CDase was strongly inactivated by N-bromosuccinimide and diethylpyrocarbonate, but activated by dithiothreitol. A comparison of the biochemical properties of the CDases from Paenibacillus sp. A11 and E. coli transformant (pJK 555) indicates that they were almost identical.  相似文献   

9.
The cyclomaltodextrinase from alkalophilic Bacillus sp. I-5 (CDase I-5) was expressed in Escherichia coli and the purified enzyme was used for characterization of the enzyme action. The hydrolysis products were monitored by both HPLC and high-performance ion chromatography analysis that enable the kinetic analysis of the cyclomaltodextrin (CD)-degrading reaction. Analysis of the kinetics of cyclomaltodextrin hydrolysis by CDase I-5 indicated that ring-opening of the cyclomaltodextrin was the major limiting step and that CDase I-5 preferentially degraded the linear maltodextrin chain by removing the maltose unit. The substrate binding affinity of the enzyme was almost same for those of cyclomaltodextrins while the rate of ring-opening was the fastest for cyclomaltoheptaose. Acarbose and methyl 6-amino-6-deoxy-alpha-d-glucopyranoside were relatively strong competitive inhibitors with K(i) values of 1.24 x 10(-3) and 8.44 x 10(-1) mM, respectively. Both inhibitors are likely to inhibit the ring-opening step of the CD degradation reaction.  相似文献   

10.
The structure and function of Mycobacterium smegmatis Dps (DNA-binding proteins from starved cells) and of the protein studied by Gupta and Chatterji, in which the C terminus that is used for binding DNA contains a histidine tag, have been characterized in parallel. The native dodecamer dissociated reversibly into dimers above pH 7.5 and below pH 6.0, with apparent pK(a) values of approximately 7.65 and 4.75; at pH approximately 4.0, dimers formed monomers. Based on structural analysis, the two dissociation steps have been attributed to breakage of the salt bridges between Glu(157) and Arg(99) located at the 3-fold symmetry axes and to protonation of Asp(66) hydrogen-bonded to Lys(36) across the dimer interface, respectively. The C-terminal tag did not affect subunit dissociation, but altered DNA binding dramatically. At neutral pH, protonation of the histidine tag promoted DNA condensation, whereas in the native C terminus, compensation of negative and positive charges led to DNA binding without condensation. This different mode of interaction with DNA has important functional consequences as indicated by the failure of the native protein to protect DNA from DNase-mediated cleavage and by the efficiency of the tagged protein in doing so as a result of DNA sequestration in the condensates. Chemical protection of DNA from oxidative damage is realized by Dps proteins in a multistep iron oxidation/uptake/mineralization process. Dimers have a decreased protection efficiency due to disruption of the dodecamer internal cavity, where iron is deposited and mineralized after oxidation at the ferroxidase center.  相似文献   

11.
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   

12.
Apoferritins prepared from horse spleen and heart and rat heart and liver were dissociated by treatment with acetic acid (pH 1.3-3.0). Sedimentation velocity studies showed that apoferritins of spleen and liver (16-17 S) and heart (18-19 S) dissociated into material sedimenting near 3.2 S. Sedimentation equilibrium measurements determined that most of the material had a molecular weight of 38,000-43,000, corresponding to subunit dimers. Failure to dissociate into subunit monomers was confirmed by gel chromatography on Sephadex G-75 and G-150. With the exception of boiling in sodium dodecyl sulfate, further treatments with 0.1-0.4 M KCl, NaCl, 4-9 M urea, 0.01-0.5 M KSCN, 0.1-0.5% Triton X-100, 5-52% dimethylsulfoxide, 10% ethylene glycol, or 0.1% trifluoroacetic acid all failed to cause dissociation into individual subunits, as did exposure to 6 M guanidine-HCl or formic acid, or prior succinylation and/or nitration of the protein. Reassociation occurred between pH 4 and 7 but was not aided by the addition of Fe(II) or reducing agents. It is concluded that ferritins readily dissociate to subunit dimer units and that further dissociation does not occur without full denaturation of the protein.  相似文献   

13.
Fragments of 32, 26 and 17 kDa of the porcine estradiol receptor were prepared, all of which contain the ligand-binding site. While dimers of the 32 and 26 kDa fragments like those of intact receptor can be dissociated by protonation, the dimer of the 17 kDa fragment obtained by trypsination of the 26 kDa fragment is resistant to lowering the pH from 7.0 to 6.5 and below. Its dissociation can be achieved by 0.5 M MgCl2 at pH 7.0. All fragments are recognized by the MAB 13H2 in Western blots. The antibody also reacts with native receptor and the three fragments, both in their monomer and dimer states. The combining ratios of antibody with receptor, or its fragments, in the monomer and dimer states and the weakening of the estradiol-receptor bond by antibody attachment support the back to back and head to toe model of receptor dimers.  相似文献   

14.
Dissociation of alpha beta and alpha gamma dimers of human hemoglobins (Hb) A and F into monomers was studied by alpha chain exchange (Shaeffer, J. R., McDonald, M. J., Turci, S. M., Dinda, D. M., and Bunn, H. F. (1984) J. Biol. Chem. 259, 14544-14547). Unlabeled carbonmonoxy-Hb A was incubated with trace amounts of preparatively purified, native, 3H-alpha subunits in 10 mM sodium phosphate, pH 7.0, at 25 degrees C. At appropriate times, free alpha monomers were separated from Hb A tetramers by anion exchange high performance liquid chromatography. Transfer of radioactivity from the alpha chain pool into Hb A was measured, yielding a first order dimer dissociation rate constant, k2 = (3.2 +/- 0.3) X 10(-3) h-1. The Arrhenius plot of k2 was linear between 7 and 37 degrees C, yielding an enthalpy of activation of 23 kcal/alpha beta dimer. As the chloride concentration was raised from 0 to 0.2 M, the dissociation rate increased 3-fold; with higher salt concentrations, however, the rate gradually returned to baseline. This rate was not altered by raising the pH from 6.5 to 7.2, but as pH was further raised to 8.4, kappa 2 increased about 3-fold. Hb F, which has an increased stability at alkaline pH, dissociated into alpha and gamma monomers 3 times more slowly than Hb A. Moreover, the dimer-monomer dissociation of Hb F was characterized by a significantly reduced pH dependence. These results demonstrate that both alpha beta and alpha gamma dimers of Hb A and Hb F dissociate reversibly into monomers under physiologic conditions. The differential pH dependence for dimer dissociation between Hb A and Hb F suggests that specific amino acid replacement at the alpha 1 gamma 1 interface confers increased resistance to alkaline denaturation.  相似文献   

15.
The molecular weight of delta-5-3-ketosteroid isomerase from Pseudomonas testosteroni was determined by means of sedimentation equilibrium and exclusion chromatography over a wide range of enzyme concentrations in 0.2 M potassium phosphate buffer, pH 7.0. In addition, the sedimentation constant of the enzyme was determinded over an extended range of concentrations. The enzyme was found to have a molecular weight of 26,000 plus or equal to 1,000, suggesting that it is a dimer of identical or similar 13,400 molecular weight polypeptide chains. In the ultracentrifuge this dimeric species was found to undergo aggregation at enzyme concentrations above 2 mg per ml and dissociation at enzyme concentrations below 0.05 mg per ml. Exclusion chromatography studies indicate that under the conditions of chromatography the oligomeric enzyme is partially dissociated at enzyme concentrations in the range 0.2 to 0.002 mug per ml. These results suggest that under conditions of enzyme assay in 0.2 M potassium phosphate buffer, pH 7.0, isomerase is in a monomeric state of aggregation.  相似文献   

16.
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   

17.
H+ and Ca2+ concentration changes in the reaction medium following MgATP addition at pH 6.0 were determined with the partially purified Ca-ATPase from sarcoplasmic reticulum vesicles in the presence of 25-50 microM CaCl2 and 5 mM MgCl2 at 4 degrees C. Previously, we showed a sequential occurrence of H+ binding and H+ dissociation in the Ca-ATPase during ATP hydrolysis and further suggested that the H+ binding takes place inside the vesicles (Yamaguchi, M., and Kanazawa, T. (1984) J. Biol. Chem. 259, 9526-9531). The present results demonstrate that the H+ binding occurred coincidently with Ca2+ dissociation from the enzyme upon conversion of the phosphoenzyme (EP) intermediate from the ADP-sensitive form to the ADP-insensitive form in the catalytic cycle of ATP hydrolysis. As KCl decreased in the medium, the extent of the H+ binding increased almost proportionately with the extent of either the Ca2+ dissociation or the accumulation of ADP-insensitive EP. Both the H+ binding and the Ca2+ dissociation were prevented by a modification of the specific SH group of the enzyme essential for the conversion of ADP-sensitive EP to ADP-insensitive EP. In the late stage of the reaction, H+ dissociation from the enzyme occurred coincidently with Ca2+ binding to the dephosphoenzyme which was formed by EP decomposition. These results are consistent with the possibility that the H+ ejection during the Ca2+ uptake with the intact vesicles previously shown by several investigators takes place through a Ca2+/H+ exchange directly mediated by the membrane-bound Ca-ATPase.  相似文献   

18.
Thermal unfolding of dodecameric manganese glutamine synthetase (622,000 M(r)) at pH 7 and approximately 0.02 ionic strength occurs in two observable steps: a small reversible transition (Tm approximately 42 degrees C; delta H approximately equal to 0.9 J/g) followed by a large irreversible transition (Tm approximately 81 degrees C; delta H approximately equal to 23.4 J/g) in which secondary structure is lost and soluble aggregates form. Secondary structure, hydrophobicity, and oligomeric structure of the equilibrium intermediate are the same as for the native protein, whereas some aromatic residues are more exposed. Urea (3 M) destabilizes the dodecamer (with a tertiary structure similar to that without urea at 55 degrees C) and inhibits aggregation accompanying unfolding at < or = 0.2 mg protein/mL. With increasing temperature (30-70 degrees C) or incubation times at 25 degrees C (5-35 h) in 3 M urea, only dodecamer and unfolded monomer are detected. In addition, the loss in enzyme secondary structure is pseudo-first-order (t1/2 = 1,030 s at 20.0 degrees C in 4.5 M urea). Differential scanning calorimetry of the enzyme in 3 M urea shows one endotherm (Tmax approximately 64 degrees C; delta H = 17 +/- 2 J/g). The enthalpy change for dissociation and unfolding agrees with that determined by urea titrations by isothermal calorimetry (delta H = 57 +/- 15 J/g; Zolkiewski M, Nosworthy NJ, Ginsburg A, 1995, Protein Sci 4: 1544-1552), after correcting for the binding of urea to protein sites exposed during unfolding (-42 J/g). Refolding and assembly to active enzyme occurs upon dilution of urea after thermal unfolding.  相似文献   

19.
1. The effect of alkaline earth cations on the dissociation of the extracellular hemoglobin of Lumbricus terrestris and the effect of ionic strength on the dissociation of the hemoglobins of L. terrestris and Tubifex tubifex at concentrations of ca 2.5 mg/ml, over the pH range 9.0-10.5 was investigated using ultracentrifugation to separate the dissociated from the undissociated molecules. 2. Mg(II), Ca(II) and Sr(II) at concentrations of up to 0.2 M, decreased the dissociation of Lumbricus oxyhemoglobin from 70% at pH 9.0 and 100% at pH 9.5 and higher, to 20-30% at 0.05 M. The three cations were equally effective in decreasing the extent of dissociation of L. terrestris oxyhemoglobin over the pH range 9.0-10.5, with a K1/2 of ca 10 mM. 3. The dissociation of L. terrestris oxyhemoglobin over the pH range 9.0-10.5 was decreased only to 50-60% in the presence of up to 0.5 M NaCl or KCl; there was no further decrease in dissociation at concentrations of the two salts up to 1.5 M. 4. The dissociation of T. tubifex oxyhemoglobin over the pH range 9.0-10.0 was decreased from 100% to ca 40-50% in the presence of 0.5 M NaCl or KCl with little or no change at higher concentrations. At pH 10.5 and 11.0 the decrease in dissociation was more gradual, reaching ca 50% at 1.5 M NaCl.  相似文献   

20.
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   

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