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.
Ribonucleotide reductase activity is rate-limiting for DNA synthesis, and inhibition of this enzyme supports cytostatic antitumor effects of inducible NO synthase. The small R2 subunit of class I ribonucleotide reductases contains a stable free radical tyrosine residue required for activity. This radical is destroyed by peroxynitrite, which also inactivates the protein and induces nitration of tyrosine residues. In this report, nitrated residues in the E. coli R2 protein were identified by UV-visible spectroscopy, mass spectrometry (ESI-MS), and tryptic peptide sequencing. Mass analysis allowed the detection of protein R2 as a native dimer with two iron clusters per subunit. The measured mass was 87 032 Da, compared to a calculated value of 87 028 Da. Peroxynitrite treatment preserved the non-heme iron center and the dimeric form of the protein. A mean of two nitrotyrosines per E. coli protein R2 dimer were obtained at 400 microM peroxynitrite. Only 3 out of the 16 tyrosines were nitrated, including the free radical Tyr122. Despite its radical state, that should favor nitration, the buried Tyr122 was not nitrated with a high yield, probably owing to its restricted accessibility. Dose-response curves for Tyr122 nitration and loss of the free radical were superimposed. However, protein R2 inactivation was higher than nitration of Tyr122, suggesting that nitration of the nonconserved Tyr62 and Tyr289 might be also of importance for peroxynitrite-mediated inhibition of E. coli protein R2.  相似文献   

2.
The reaction of tetranitromethane with human chorionic gonadotropin and its subunits has been investigated. The hormone consists of two subunits, α and β, containing four and three tyrosyl residues, respectively. Introduction of 1 nitrated tyrosine residue into the native hormone was accompanied by a 20% loss in immunological reactivity and a 50% loss in biological activity. This initial reaction occurred at α Tyr-88 and/or α Tyr-89. Exhaustive nitration of the hormone modified α tyrosines 65, 88, and 89 and resulted in 75% inactivation biologically and 50% immunologically. Either nitrated α subunit obtained by dissociation of the nitrated hormone recombined with the native β subunit to give a hormone whose activity was in reasonable agreement with that of the corresponding nitrated monomer. These results indicate involvement of α Tyr-88 and/or α Tyr 89 in binding of the hormone to its receptor. These residues are not required for binding to the β subunit, however. Tyr-65 of the α subunit is probably not involved in binding to either the β subunit or the hormone receptor. The β subunit obtained from the exhaustively nitrated hormone was unmodified and recombined with native α to give fully active hormone. About 25% of the protein was recovered as polymeric material following nitration; lesser amounts of crosslinked monomer were formed. Both were biologically inactive. The polymer products retained about 30% of the native immunological competence.Nitration of the isolated α subunit fully converted the remaining tyrosine (Tyr-37) to 3-nitrotyrosine in a two-step reaction. The fully nitrated α subunit did not recombine well with the native β subunit and the recombinant hormone has 10% or less of the native activity. Involvement of α Tyr-37 in binding to the β subunit is suggested by these data. However, exposure of this residue by a conformational change in the α subunit after dissociation of the native hormone, while it seems unlikely in view of the high disulfide content, is also consistent with the data. Reaction of the free β subunit with tetranitromethane resulted in complete nitration of Tyr-37, 85% nitration of Tyr-59, and 25% nitration of Tyr-82. The nitrated β subunit did not recombine well with native α but the isolated recombinant had two-thirds of the native activity. From these data we conclude that β Tyr-37 and/or β Tyr-59 are possibly involved in binding to the α subunit but do not have a role in the biological activity. Tyr-82 of β is apparently not involved in either subunit interactions or hormone-receptor binding.  相似文献   

3.
The molecular weight of 7 S nerve growth factor has been studied in the analytical ultracentrifuge between pH values 6.8 and 7.8. At pH 6.8, where no dissociation is observed, the molecular weight was found to be 137,000 plus and minus 7,000. Between pH values 7.4 and 7.8 there is some dissociation. Using the data from this study and results in the literature, a model of 7 S nerve growth factor, (alpha beta gamma)2, in reversible equilibrium with a subunit complex, (alpha beta gamma), is proposed.  相似文献   

4.
Factor VIIIa is a heterotrimer of the factor VIII heavy chain-derived A1 and A2 subunits plus the factor VIII light chain-derived A3-C1-C2 subunit. While the A1 and A3-C1-C2 subunits can be isolated as a stable dimer, the A2 subunit is weakly associated with the dimer. In the human protein, the association of A2 with dimer is reversible and governed by a pH-dependent dissociation constant. Using the specific activity of factor VIIIa as an indicator of trimer concentration, the Kd (pH 6.0) was determined to be 28 nM whereas at the more physiologic pH (pH 7.4) this value was approximately 260 nM. Results from pH shift experiments confirmed the reversible binding of A2 to dimer as did the capacity for high levels of exogenous A2 subunit to inhibit the spontaneous decay of factor VIIIa activity. A2 subunit associated with the A1 subunit in the A1/A3-C1-C2 dimer based upon the capacity for free A1 subunit to inhibit the reconstitution of factor VIIIa from A2 subunit and dimer. These results indicate that the primary mechanism for the spontaneous decay of human factor VIIIa is the reversible dissociation of A2 subunit from the A1 subunit of the A1/A3-C1-C2 dimer.  相似文献   

5.
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   

6.
A method is described for the purification of native hexokinases P-I and P-II from yeast using preparative isoelectric focussing to separate the isozymes. The binding of glucose to hexokinase P-II, and the effect of this on the monomer--dimer association--dissociation reaction have been investigated quantitatively by a combination of titrations of intrinsic protein fluorescence and equilibrium ultracentrifugation. Association constants for the monomer-dimer reaction decreased with increasing pH, ionic strength and concentration of glucose. Saturating concentrations of glucose did not bring about complete dissociation of the enzyme showing that both sites were occupired in the dimer. At pH 8.0 and high ionic strength, where the enzyme existed as monomer, the dissociation constant of the enzyme-glucose complex was 3 X 10(-4) mol 1(-1) and was independent of the concentration of enzyme. Binding to the dimeric form at low pH and ionic strength (I=0.02 mol 1(-1), pH less than 7.5) was also independent of enzyme concentration (in the range 10-1000 mug ml-1) but was much weaker. The process could be described by a single dissociation constant, showing that the two available sites on the dimer were equivalent and non-cooperative; values of the intrinsic dissociation constant varied from 2.5 X 10(-3) mol 1(-1) at pH 7.0 to 6 X 10(-3) at pH 6.5. Under intermediate conditions (pH 7.0, ionic strength=0.15 mol 1(-1)), where monomer and dimer coexisted, the binding of glucose showed weak positive cooperatively (Hill coefficient 1.2); in addition, the binding was dependent upon the concentration of enzyme in the direction of stronger binding at lower concentrations. The results show that the phenomenon of half-sites reactivity observed in the binding of glucose to crystalline hexokinase P-II does not occur in solution; the simplest explanation of our finding the two sites to be equivalent is that the dimer results from the homologous association of two identical subunits.  相似文献   

7.
1. The pKa of the phenolic hydroxy group of the Tyr(3NO2)-237 residue in pig heart [Tyr(3NO2)237]lactate dehydrogenase is 7.2 in the apoenzyme, 7.4 in the enzyme-NADH complex and 7.8 in the enzyme-NADH-oxamate complex. The alkaline shift from apoenzyme to ternary complex is ascribed to the approach of the Glu-107 residue during the movement of the polypeptide loop residues 98-110. 2. The affinities of the nitrated enzyme for NADH and for oxamate (in the presence of NADH) are slightly less than those of the native enzyme. The turnover number for the nitrated enzyme in the pyruvate-to-lactate direction is about 0.75 of the value for the native enzyme. 3. Temperature-jump relaxation experiments of the enzyme saturated with NADH but fractionally saturated with oxamate are interpreted to show that the pKa of the nitrotyrosine residue responds to a protein rearrangement after oxamate binds to the binary enzyme-NADH complex. 4. Transient-kinetic experiments show the environment of the Tyr(3NO2)-237 residue in the enzyme-NADH-pyruvate complex of the steady state to be similar to that in the enzyme-NADH-oxamate inhibitor complex.  相似文献   

8.
9.
Nyarko A  Cochrun L  Norwood S  Pursifull N  Voth A  Barbar E 《Biochemistry》2005,44(43):14248-14255
LC8 is a highly conserved light-chain subunit of cytoplasmic dynein that interacts with a wide variety of cellular proteins and is presumed to play a fundamental role in dynein assembly and cargo recruitment and in the assembly of protein complexes unrelated to dynein. LC8 is a dimer at physiological pH but dissociates to a folded monomer at pH < 4.8. We have suggested that acid-induced dimer dissociation is due to protonation of His 55, which is stacked against His 55' and completely buried in the dimer interface. In this work, we show that the pH-induced dissociation is reversible and indeed governed by the ionization state of His 55. Mutagenesis of His 55 to Lys results in a monomer in the pH range of 3-8, while the mutation to Ala results in a dimer in the same pH range. Mutations that disrupt intermolecular hydrogen bonds between Tyr 65 and Lys 44' and His 55 and Thr 67' do not change the association state of the dimer. Titration curves for His 55 and the two other histidines, His 72 and 68, were determined by (13)C-(1)H NMR for H55K and for WT-LC8 in the monomeric and dimeric states. The pK(a) values of His 72 and His 68 are 6 in the WT dimer and 6.2-6.5 in monomeric H55K, while the pK(a) of His 55 is about 4.5 in the WT dimer. These results indicate that deprotonation of His 55 is linked to dimer formation and that mutation of His 55 to a small neutral residue or to a positively charged residue uncouples the protonation and dissociation processes.  相似文献   

10.
11.
Complete dissociation of dimeric plasma sex steroid-binding protein (SBP or SHBG) was obtained in 6 M urea at 10 degrees C. Removal of urea resulted in the refolding of monomers, followed by reformation of dimeric SBP, which migrates with the same mobility as the native protein. Dimerization does not require Ca+2 or steroid. Renatured monomers yield dimers with dissociation constants for 5 alpha-dihydrotesterone (DHT) and 17 beta-estradiol (E2) indistinguishable from those of native human SBP. This phenomenon was also demonstrated by mixing human and rabbit SBPs that, upon renaturation, form a hybrid dimer composed of one human subunit and one rabbit subunit. The hybrid binds both DHT and E2 in contrast to rSBP, which only binds the androgen. Therefore, we conclude that (1) docking of the two subunits creates an asymmetric steroid-binding site located at the interface between the subunits, and (2) only one face of the dimer defines the specificity for binding E2 by encompassing portion of a structural motif that recognizes the flat ring A of E2. The remaining portion, which recognizes the saturated ring A of DHT, is shared by both faces of the dimer. Because native monomers do not exist alone, the often-asked question of whether the SBP monomer binds steroid can be considered meaningless; steroid-binding activity is expressed only in the dimeric state. Finally, formation of the hybrid indicates that SBP dimerization represents a conserved event during the molecular evolution of SBP, suggesting that the structural elements responsible for dimerization will be homologous in SBPs from other species.  相似文献   

12.
Bovine heart phosphorylase kinase has been isolated by a procedure involving precipitation with polyethylene glycol, DEAE-Sephacel chromatography and calmodulin-Sepharose affinity chromatography. The isolated enzyme had a specific activity of 8.3 IU/mg of protein at pH 8.2 at 30 degrees C in the presence of 1% glycogen. The native enzyme had a sedimentation coefficient of 23 S and the Mr of the alpha', beta, gamma, and delta subunits, were 140,000, 130,000, 46,000, and 18,000, respectively. Activation of the phosphorylase kinase by the catalytic subunit of bovine heart cAMP-dependent protein kinase increases the pH 6.8/8.2 activity ratio from 0.01 to 0.32-0.38. Glycogen (1%) decreased the Km of the activated phosphorylase kinase at pH 6.8 for phosphorylase b from 5.5 to 1.25 mg/ml. Trypsin treatment increased the pH 6.8 activity but decreased the pH 8.2 activity. During this process the alpha' subunit was converted to a Mr 110,000 polypeptide and the enzyme activity was converted essentially to a 5.9 S species having an apparent Mr of 100,000 as determined by gel filtration. On extended trypsin treatment only one major polypeptide corresponding to the beta subunit remained. The same polypeptide was present in the active fractions following gel filtration of the trypsinized kinase.  相似文献   

13.
The pH dependence of the apparent tetramer to dimer dissociation constant has been determined at 20 degrees for both oxy- and deoxyhemoglobins A and Kansas. These measurements were made by three different procedures: gel chromatography, sedimentation velocity, and kinetic methods in either of three buffer systems: 0.05 M cacodylate, Tris, or glycine with 1 mM EDTA and 0.1 M NaCl between pH 6.5 and 11. The tetramer-dimer dissociation constant of human oxyhemoglobin A decreases from about 3.2 X 10(-6) M at pH 6.0 to about 3.2 X 10(-8) M at pH 8.5. The slope of this line indicates that the dissociation of tetramer to dimer is accompanied by the uptake of about 0.6 protons per mol of tetramer in this region. The corresponding dissociation constant for deoxyhemoglobin in the same pH region increases apparently almost linearly from 1.0 x 10(-12) M at pH 6.5 to about 1.0 x 10(-5) M at pH 11. To dimer is associated with the release of about 1.6 protons per mol of tetramer. Comparison of these data with the known proton release accompanying the oxygenation of tetramers confirms that the pH dependence of oxygen binding by dimers must be very small. The present data predict that the overall proton release or uptake per oxygen bound by dimer should be less than 0.1. The tetramer-dimer dissociation equilibria of oxy- and deoxyhemoglobins above pH 8.5 have identical pH dependences. In this range the dissociation constant of deoxy-Hb is about one-tenth that of oxyhemoglobin. Human oxyhemoglobin Kansas is known to have an enhanced tetramer-dimer dissociation compared with that of hemoglobin A. Below pH 8.5 the tetramer-dimer dissociation constant of Hb Kansas is about 400 times greater than that of HbA in the absence of phosphate buffers. In contrast, the tetramer-dimer dissociation constants of deoxyhemoglobins A and Kansas appear to be identical. These findings are consistent with previous structural observations on these hemoglobins. The data on the tetramer-dimer dissociation of human hemoglobin were used to calculate the total free energy of binding of oxygen to the tetramer and the median oxygen pressure on the basis of fundamental linkage relations and a pH-independent estimate of the total free energy of binding oxygen to dimer. Simulated oxygen binding curves were generated with the equations of Ackers and Halvorson (Ackers, G. K., and Halvorson, H. (1974) Proc. Natl. Acad. Sci. U.S.A. 71, 4312-4316) by making two assumptions: (a) that the dimers are noncooperative and pH-independent in O2 binding and (b) that the distribution of cooperative energy in the oxygenation of tetramers is independent of pH. We have compared these simulations with experimental data obtained at low protein concentrations (30 to 124 muM heme) to show that the variation in oxygen affinity with pH can be described in terms of the subunit equilibria. We conclude that an accurate analysis of the contributions of individual oxygen binding steps to the Bohr effect cannot be made without considering the contributions of the dimers to oxygen binding...  相似文献   

14.
The binding of oxygen by the haemocyanin of the gastropod Lymnaea stagnalis was studied by equilibrium and kinetic methods. The studies were performed under conditions in which the haemocyanin molecule was in the native state. Over the pH range 6.8-7.6, in the presence of 10mM-CaCl2 the haemocyanin bound O2 cooperatively. Over this pH range the haemocyanin molecule displayed a normal Bohr effect whereby the O2 affinity of the molecule decreased with a fall in the pH of the solution. The maximum slope of the Hill plot (hmax.) was 3.5, obtained at pH 7.5. An increase in the CaCl2 concentration from 5 to 20 mM at pH 6.8 resulted in a slight increase in the oxygen affinity, with hmax. remaining virtually unchanged. At constant pH and CaCl2 concentration, an increase in NaCl concentration from 0 to 50 mM resulted in a small decrease in O2 affinity, but a significant increase in the value of hmax. from 3.5 to 8.6. Temperature-jump relaxation experiments over a range of O2 concentrations produced single relaxation times. The dependence of the relaxation time on the reactant concentrations indicated a simple bimolecular binding process. The calculated association and dissociation rate constants for this process at pH 7.5 are 29.5 X 10(6) M-1 X S-1 and 49 S-1 respectively. The association rate constant kon was found to be essentially independent of pH and CaCl2 concentration. The dissociation rate constant, koff, however, increased with a decrease in the pH, but was also independent of CaCl2 concentration. These results indicate that the stability of the haemocyanin-O2 complex is determined by the dissociation rate constant.  相似文献   

15.
S Pin  C A Royer  E Gratton  B Alpert  G Weber 《Biochemistry》1990,29(39):9194-9202
The dissociation of the subunits of human adult oxyhemoglobin has been investigated by using steady-state fluorescence anisotropy, multifrequency phase fluorometry, and high hydrostatic pressure. Human hemoglobin obtained by using two purification procedures (bulk preparation by centrifugation or further fractionation using anion-exchange chromatography) was labeled with an extrinsic fluorescent probe, 5-(dimethylamino)naphthalene-1-sulfonyl chloride (DNS-Cl). The long fluorescence lifetime of this probe allows for the observation of the macromolecular tumbling, and thus provides a method for observing changes in the size of the complex upon subunit dissociation under differing solution conditions of proton and organic phosphate concentration. At pH 7, the dansylated preparations of bulk and fractionated hemoglobin showed a concentration-dependent decrease in the anisotropy which though not identical can only arise from the tetramer to dimer dissociation. We observed primarily the dimer at pH 9 and a small destabilization of the tetramer in the presence of saturating inositol hexaphosphate (IHP). High-pressure experiments allowed for the observation of the dissociation of the hemoglobin dimer into monomers. From these measurements, we estimate the dimer dissociation constant to be between 0.1 and 1 nM. We compare the present results on the subunit affinities in hemoglobin obtained from steady-state and time-resolved fluorescence data with those obtained previously by using gel filtration, sedimentation, and kinetic techniques. These comparisons are indicative of a certain degree of conformational heterogeneity in the hemoglobin preparations.  相似文献   

16.
1. The hemocyanin of the chiton, Stenoplax conspicua, has a molecular weight determined by light-scattering of 4.2 X 10(6) daltons, (dt) and a sedimentation coefficient of 60 S. 2. The fully dissociated subunits in 6.0 and 8.0 M urea, and at pH 8.9-10 in the absence of divalent ions, have molecular weights of 4.15-4.30 x 10(5) and 4.17-4.75 x 10(5) dt, which is close to one-tenth of the molecular weight of the parent hemocyanin assembly. 3. The pH dependence of the molecular weights from pH 4.5 to 11 exhibit bell-shaped transition profiles, best accounted for by a three-species, decamer to dimer to monomer scheme of subunit dissociation, with one acidic and one basic ionizing group per dimer and 5-8 acidic and basic groups per monomer. 4. In the absence of stabilizing divalent ions S. conspicua hemocyanin is relatively unstable. At pH 7.4 in the presence of 0.01 M EDTA, it is predominantly in the dimeric state, characterized by a sedimentation constant of 18 S. It is also more readily dissociated to monomers at high pHs (8-9 and above) than are the C. stelleri and A. granulata hemocyanins. 5. Urea and GdmCl are effective dissociating agents of S. conspicua hemocyanin. The urea dissociation profile obtained at pH 8.5, 0.01 M Mg2+, 0.01 M Ca2+, and analyzed by means of the decamer-dimer-monomer scheme of subunit dissociation gave estimates of about 30 amino acid groups (Napp) at the dimer contacts within the hemocyanin decamers and about 120 groups per monomer within each dimer, suggesting hydrophobic stabilization of hemocyanin assembly.  相似文献   

17.
The activity of human immunodeficiency virus 1 (HIV-1) protease has been examined as a function of solvent composition, incubation time, and enzyme concentration at 37 degrees C in the pH 4.5-5.5 range. Glycerol and dimethyl sulfoxide inhibit the enzyme, while polyethylene glycol and bovine serum albumin activate the enzyme. When incubated at a concentration of 50-200 nM, the activity of the protease decreases irreversibly with an apparent first-order rate constant of 4-9 x 10(-3) min-1. The presence of 0.1% (w/v) polyethylene glycol or bovine serum albumin in the reaction buffer dramatically stabilizes enzyme activity. In the absence of prolonged incubation of the enzyme at submicromolar concentration, the specific activity of HIV-1 protease in buffers of either high or low ionic strength is constant over the enzyme concentration range of 0.25-5 nM, indicating that dissociation of the dimeric protease, if occurring, can only be governed by a picomolar dissociation constant. Similarly, the variation of the specific activity of HIV-2 protease over the enzyme concentration of 4-85 nM is consistent only with a dimer dissociation constant of less than 10 nM. We conclude that: 1) the assumption of a nondissociating HIV-1 protease is a valid one for kinetic studies of tight-binding inhibitors where nanomolar concentrations of the enzymes are employed; 2) stock protease solutions of submicromolar concentration in the absence of activity-stabilizing compounds may lead to erroneous kinetic data and complicate mechanistic interpretations.  相似文献   

18.
The role of the hydroxyl group of tyrosine 6 in the catalytic mechanism of isoenzyme 3-3 of rat glutathione S-transferase has been examined by x-ray crystallography and site-specific replacement of the residue with phenylalanine and evaluation of the catalytic properties of the mutant enzyme. This particuar tyrosine residue is conserved in the sequences of all of the cytosolic enzymes and is found, in crystal structures of both isoenzyme 3-3 from the mu-gene class and an isoenzyme from the pi-gene class, to be proximal to the sulfur of glutathione (GSH) or glutathione sulfonate bound at the active site. The 2.2-A structure of the binary complex of isoenzyme 3-3 and GSH indicates that the hydroxyl group of Tyr6 is located 3.2-3.5 A from the sulfur of GSH, well within hydrogen bonding distance. Removal of the hydroxyl group of Tyr6 has essentially no effect on the dissociation constant (22 +/- 3 microM) for GSH. Nevertheless the Y6F mutant exhibits a turnover number which is only about 1% that of the native enzyme when assayed at pH 6.5 with either 1-chloro-2,4-dinitrobenzene (CDNB) or 4-phenyl-3-buten-2-one. UV difference spectra of the binary enzyme-GSH complexes suggest that the predominant ionization state of GSH in the active site of the Y6F mutant is the neutral thiol (e.g. EY6F.GSH) which is in contrast to the native enzyme in which the thiol is substantially deprotonated (e.g. E.GS-). Spectrophotometric titration suggests that the pKa of the thiol is 6.9 +/- 0.3 in the E.GSH complex and greater than or equal to 8 in the EY6F.GSH binary complex. In addition, the pH dependence of kcat/KmCDNB reveals that the reactions catalyzed by the native enzyme and the Y6F mutant are dependent on a single ionization in the E.GSH and EY6F.GSH complexes with pKa = 6.2 +/- 0.1 and 7.8 +/- 0.3, respectively. The results suggest that the hydrogen bond between Tyr6 and the enzyme-bound nucleophile helps to lower the pKa of GSH in the binary enzyme-substrate complex.  相似文献   

19.
A recombinant fragment of the human receptor for epidermal growth factor containing both its extracellular domain and its membrane-spanning segment, when dissolved with Triton X-100, was observed to dimerize in response to addition of epidermal growth factor (EGF) even at the lowest concentration of this fragment that could be assayed (4 nM). Consequently, the dissociation constant for the dimer of this fragment is at least 10,000-fold smaller than that for dimers of soluble, recombinant forms of the extracellular domain lacking the membrane-spanning segment. The second-order rate constant for dimerization of the fragment containing the extracellular domain and the membrane-spanning segment was estimated to be greater than 0.3 nM(-1) min(-1), more than 10-fold that of the native enzyme under the same conditions. This result suggests that the cytoplasmic domain of the intact enzyme sterically hinders its dimerization. When EGF is removed from the dimer of the fragment, the rate constant for its dissociation is greater than 0.2 min(-1), more than 40-fold that of the native enzyme. This result suggests that interfaces between cytoplasmic domains of intact EGF receptor impart significant stabilization to the dimer of the enzyme.  相似文献   

20.
 用荧光光谱法、截流荧光法和酶活力测定法研究了在盐酸胍溶液中米曲霉氨基酰化酶变性动力学。我们发现在4.8mol/L盐酸胍溶液作用下(0.05mol/L磷酸缓冲溶液,pH7.4,25℃),氨基酰化酶二聚体解离成单亚基过程是一个十分快速的过程,反应速率常数k为3361l/s,即约需3ms时间完成;而单亚基分子的构象变化需要约20min方能到达平衡态,这是一个逐渐变化的缓慢过程。酶分子在胍作用下的失活现象同酶分子的结构变化紧密相关,在胍浓度大于4mol/L时酶完全失活。在高浓度盐酸胍下酶失活主要是因为酶二聚体迅速解离成单亚基的过程和单亚基构象逐渐变化的缓慢过程。双亚基解离常数大小标志着酶分子亚基间作用力的强弱。  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