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 62 毫秒
1.
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   

2.
The influence of the constitutive metal ions on the equilibrium properties of dimeric Photobacterium leiognathi Cu,Zn superoxide dismutase has been studied for the wild-type and for two mutant protein forms bearing a negative charge in the amino acid clusters at the dimer association interface. Depletion of copper and zinc dissociates the two mutant proteins into monomers, which reassemble toward the dimeric state upon addition of stoichiometric amounts of zinc. Pressure-dependent dissociation is observed for the copper-depleted wild-type and mutated enzymes, as monitored by the fluorescence shift of a unique tryptophan residue located at the subunit association interface. The spectral shift occurs slowly, reaching a plateau after 15-20 minutes, and is fully reversible. The recovery of the original fluorescence properties, after decompression, is fast (less than four minutes), suggesting that the isolated subunit has a relatively stable structure, and excluding the presence of stable intermediates during the dimer-monomer transition. The dimer dissociation process is still incomplete at 6.5 kbar for the copper-depleted wild-type and mutated enzymes, at variance with what is generally observed for oligomeric proteins that dissociate below 3 kbar. Measurement of the degree of dissociation, at two different protein concentrations, allows us to calculate the standard volume variation upon association, Delta V, and the dissociation constant K(d0), at atmospheric pressure, (25 ml/mol and 3 x 10(-7)M, respectively). The holoprotein is fully dimeric even at 6.5 kbar, which allows us to evaluate a lower Delta G degrees limit of 11.5 kcal/mol, corresponding to a dissociation constant K(d0)<10(-9)M.  相似文献   

3.
Mitochondrial malate dehydrogenase (mMDH) shows a complex dependence upon ionic environment that includes kinetic and structural effects. We measured mMDH activity in several buffers (phosphate, MOPS, and MES) at pH 6.5 and 7.5, and in the presence of a number of anions, at highly diluted enzyme concentrations where mMDH showed significant loss of activity. Under these conditions, mMDH activity shows a non-linear dependence on enzyme concentration, in agreement with the existence of a dimer–monomer equilibrium, where only the dimeric form is active. According to this hypothesis, the dissociation constant of mMDH dimer has been determined to be 5.4 nM in the MES buffer at pH 6.5. Either the presence of a small anion like phosphate, or an increase of the pH from 6.5 to 7.5 shifts the equilibrium in favor of the dimeric form with the two effects appearing to be additive. To extend the study, we analysed the effect of a number of anions on the mMDH activity in 50 mM MOPS buffer at pH 7.5. All the anions had a dual effect: at low concentrations, they increased the activity of mMDH, while at high concentrations, they inhibited it. A more accurate analysis of the data revealed that the activation capacity of all the anions tested was similar, although they differed in their inhibitory influence. To show these differences more clearly, the experiment was repeated in 50 mM phosphate buffer at pH 7.5, under conditions where almost all activations were due to the buffer. The analysis of the results obtained under these conditions revealed the following sequence of inhibition potency: phosphate相似文献   

4.
The structure-function relationship of mitochondrial malate dehydrogenase was investigated at low enzyme concentration, as well as in the presence of polyethylene glycol (PEG 6000) and structure making ions. Previous reports claimed the dimeric enzyme to undergo dissociation in dilute solution, and PEG-induced pairing of dimers in the crystalline state. Sedimentation analysis and gel filtration in 0.1 M sodium phosphate pH 7.6 plus 1 mM EDTA and 1 mM dithioerythritol prove the enzyme to be a stable dimer at c greater than or equal to 0.2 microgram/ml (5 nM). In the presence of 8-20% (w/v) PEG 6000, association of the dimer to tetramers and higher aggregates is observed. At 20% (w/v) polyethylene glycol, ultracentrifugal analysis yields up to 50% tetramers; chemical cross-linking by glutaraldehyde confirms the association in a qualitative way. The enzymatic properties of mMDH (specific activity, Km for oxaloacetate and NADH) in the absence and in the presence of PEG 6000 are indistinguishable. At high polyethylene glycol concentrations (greater than or equal to 20%), the thermal stability of the enzyme is found to be increased. The fluorescence emission, as well as the far-UV and near-UV circular dichroism remain unaffected. Accumulated evidence from equilibrium experiments at low enzyme concentration and reconstitution kinetics (after dissociation at acid pH) proves the active species of mMDH to be the dimer.  相似文献   

5.
Gel exclusion chromatographic studies demonstrate that cytosolic and mitochondrial malate dehydrogenases (cMDH and mMDH) dissociate into subunits in the presence of 0.1% of the non-ionic detergent Triton X-100 (TX-100). The presence of cofactor and catalytically competent cofactor-substrate pairs does not protect mMDH against this dissociation. In contrast, cMDH dimers resist dissociation in the presence of either addition. Since steady state kinetic studies indicate both enzymes are fully active in the presence of 0.1% TX-100, we conclude that quaternary structure is not a kinetically important feature of mMDH structure and cooperativity does not account for mMDH kinetic anomalies. In contrast, cooperativity is a reasonable explanation for cMDH kinetic properties even in the presence of 0.1% TX-100, since this enzyme's subunits associate in the presence of active site ligands. The existence of fully active mMDH subunits raises the possibility that this species rather than the dimer may be a constituent of proposed multi-enzyme complexes of the mitochondrion. Preliminary chromatographic experiments involving gently disrupted mitochondria have found MDH activity in differently sized complexes, all with molecular weights larger than the mMDH dimer but smaller than complexes anticipated for multi-enzyme complexes involving other enzymes and the mMDH dimer.  相似文献   

6.
A kinetic and thermodynamic study has been carried out to characterize quantitatively the conformational equilibrium of gramicidin A (GA) in tetrahydrofuran at different peptide concentrations in the absence and presence of egg yolk phosphatidylcholine by using size-exclusion high-performance liquid chromatographic analysis. In the absence of lipid, the experimental data fit a simple dimer-monomer equilibrium, the rate and equilibrium constants for the dissociation process being (1.6 +/- 0.7) X 10(-7) s-1 and (8.5 +/- 0.3) X 10(-6) M, respectively. A higher extent of monomerization and a decrease in the time required for reaching equilibrium are detected in the presence of phospholipid, the kinetic and thermodynamic effects depending on both lipid and GA concentrations. In order to account for these observations a cyclic equilibrium mechanism is proposed which is analysed in terms of four conformational species, namely, free monomer, free dimer, lipid-bound monomer and lipid-bound dimer. The results obtained are discussed in relation to recent literature data on lipid-protein interactions.  相似文献   

7.
FtsZ is a major protein in bacterial cytokinesis that polymerizes into single filaments. A dimer has been proposed to be the nucleating species in FtsZ polymerization. To investigate the influence of the self-assembly of FtsZ on its unfolding pathway, we characterized its oligomerization and unfolding thermodynamics. We studied the assembly using size-exclusion chromatography and fluorescence spectroscopy, and the unfolding using circular dichroism and two-photon fluorescence correlation spectroscopy. The chromatographic analysis demonstrated the presence of monomers, dimers, and tetramers with populations dependent on protein concentration. Dilution experiments using fluorescent conjugates revealed dimer-to-monomer and tetramer-to-dimer dissociation constants in the micromolar range. Measurements of fluorescence lifetimes and rotational correlation times of the conjugates supported the presence of tetramers at high protein concentrations and monomers at low protein concentrations. The unfolding study demonstrated that the three-state unfolding of FtsZ was due to the mainly dimeric state of the protein, and that the monomer unfolds through a two-state mechanism. The monomer-to-dimer equilibrium characterized here (K(d) = 9 μM) indicates a significant fraction (~10%) of stable dimers at the critical concentration for polymerization, supporting a role of the dimeric species in the first steps of FtsZ polymerization.  相似文献   

8.
Tubulin dimer dissociation detected by fluorescence anisotropy   总被引:1,自引:0,他引:1  
M R Mejillano  R H Himes 《Biochemistry》1989,28(15):6518-6524
We have demonstrated a concentration-dependent dissociation of bovine brain tubulin dimer covalently labeled with 5-[(4,6-dichlorotriazin-2-yl)amino]fluorescein (DTAF) or with fluorescein isothiocyanate (FITC) by fluorescence anisotropy and size-exclusion HPLC. The fluorescence anisotropy values decreased to a limiting value upon dilution of tubulin from 10(-5) to 8 x 10(-8) M. A dissociation constant in 0.1 M Pipes, pH 6.9, 1 mM EGTA, and 1 mM MgSO4 at 20 degrees C was estimated to be (8.4 x 10(-7) +/- (0.4 x 10(-7) M. Control experiments using monomeric and other dimeric proteins, urea-denatured tubulin, and DTAF-tubulin diluted into solutions of bovine serum albumin or unlabeled tubulin were consistent with the finding that the changes in anisotropy upon dilution are due to protein dissociation. These results were supported by size-exclusion HPLC data where an increase in the elution volume of DTAF-tubulin and FITC-tubulin was observed with decreasing protein concentrations. Reversibility of the dissociation process and the lack of denaturation at high dilution were shown by the ability of reconcentrated protein to assemble into microtubules to about the same extent as undiluted protein. Fluorescent lifetimes and limiting anisotropy values were found to be approximately identical at different tubulin concentrations, indicating that the anisotropy changes reflect changes in size or rotational correlation time of the protein. Studies on the effects of tubulin ligands and promoters or inhibitors of assembly demonstrated that their effects on tubulin dimer-monomer equilibria are small but reproducible.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
High-pressure methods have become an interesting tool of investigation of structural stability of proteins. They are used to study protein unfolding, but dissociation of oligomeric proteins can be addressed this way, too. HIV-1 protease, although an interesting object of biophysical experiments, has not been studied at high pressure yet. In this study HIV-1 protease is investigated by high pressure (up to 600 MPa) fluorescence spectroscopy of either the inherent tryptophan residues or external 8-anilino-1-naphtalenesulfonic acid at 25°C. A fast concentration-dependent structural transition is detected that corresponds to the dimer-monomer equilibrium. This transition is followed by a slow concentration independent transition that can be assigned to the monomer unfolding. In the presence of a tight-binding inhibitor none of these transitions are observed, which confirms the stabilizing effect of inhibitor. High-pressure enzyme kinetics (up to 350 MPa) also reveals the stabilizing effect of substrate. Unfolding of the protease can thus proceed only from the monomeric state after dimer dissociation and is unfavourable at atmospheric pressure. Dimer-destabilizing effect of high pressure is caused by negative volume change of dimer dissociation of −32.5 mL/mol. It helps us to determine the atmospheric pressure dimerization constant of 0.92 μM. High-pressure methods thus enable the investigation of structural phenomena that are difficult or impossible to measure at atmospheric pressure.  相似文献   

10.
M Herold  K Kirschner 《Biochemistry》1990,29(7):1907-1913
The unfolding and dissociation of the dimeric enzyme aspartate aminotransferase (D) from Escherichia coli by guanidine hydrochloride have been investigated at equilibrium. The overall process was reversible, as judged from almost complete recovery of enzymic activity after dialysis of 0.7 mg of denatured protein/mL against buffer. Unfolding and dissociation were monitored by circular dichroism and fluorescence spectroscopy and occurred in three separate phases: D in equilibrium 2M in equilibrium 2M* in equilibrium 2U. The first transition at about 0.5 M guanidine hydrochloride coincided with loss of enzyme activity. It was displaced toward higher denaturant concentrations by the presence of either pyridoxal 5'-phosphate or pyridoxamine 5'-phosphate and toward lower denaturant concentrations by decreasing the protein concentration. Therefore, bound coenzyme stabilizes the dimeric state, and the monomer (M) is inactive because the shared active sites are destroyed by dissociation of the dimer. M was converted to M* and then to the fully unfolded monomer (U) in two subsequent transitions. M* was stable between 0.9 and 1.1 M guanidine hydrochloride and had the hydrodynamic radius, circular dichroism, and fluorescence of a monomeric, compact "molten globule" state.  相似文献   

11.
The beta subunit of DNA polymerase III holoenzyme is in a dimer-monomer equilibrium at physiological beta concentrations. Dissociation is accompanied by the fluorescence enhancement of a fluorophore attached to a unique sulfhydryl group of beta (Griep, M. A., and McHenry, C. S. (1988) Biochemistry 27, 5210-5215). Sequencing of the isolated tryptic peptides of beta revealed that the fluorescent maleimide group was attached to cysteine 333. The 2 residues, lysine 332 and glutamate 334, that flank this residue are hydrophilic and may place cysteine 333 on the surface of beta, explaining its high reactivity. Fluorescence energy transfer permitted us to locate the uniquely labeled cysteines 333 of beta at the distal ends of the beta dimer. When the beta dimer was dissociated to monomers, the accompanying alteration of the conformational state was reported by the fluorescein-5-maleimide (fluorescein)-labeled cysteines which were located far from the dimer interface. The carboxyl of fluorescein had a fluorescence pKa of 6.9 when beta was in its dimeric state. The pKa decreased by 0.3 pH unit upon dissociation to monomers and resulted in the fluorescence enhancement that was observed when the signal was monitored at constant pH. The adjacent glutamate 334 apparently increased the pKa of the attached fluorescein when beta was in its dimeric state. Movement of either the adjacent lysine 332 amino side chain to a closer position or glutamate 334 to a position further away could lower the pKa upon beta monomerization. Thus, beta undergoes a conformational change concomitant with dimer dissociation that was transmitted to the opposite ends of the beta dimer. The pKa of fluorescein attached to the distal cysteines was shifted, leading to greater ionization and enhanced fluorescence.  相似文献   

12.
G R Parr  G G Hammes 《Biochemistry》1976,15(4):857-862
The kinetics of dissociation and reassembly of rabbit skeletal muscle phosphofructokinase has been studied using fluorescence, stopped-flow fluorescence and enzyme activity measurements. The dissociation of the fully active tetramer in 0.8 M guanidine hydrochloride (0.1 M potassium phosphate, pH 8.0) occurs in three kinetic phases as measured by changes in the protein fluorescence emission intensity: dissociation of tetramer to dimer with a relaxation time of a few milliseconds; dissociation of dimer to monomer with a relaxation time of a few seconds; and a conformational change of the monomer with a relaxation time of a few minutes. All three phases exhibit first-order kinetics; ATP (0.05 mM) retards the second step but does not influence the rate of the other two processes. The rate of the second process increases with decreasing temperature; this may be due to the involvement of hydrophobic interactions in the stabilization of the dimeric enzyme. A further unfolding of the monomer polypeptide chain occurs at higher guanidine concentrations, and the relaxation time associated with this process was found to be 83 ms in 2.5 M guanidine, 0.1 M potassium phosphate (pH 8.0) at 23 degrees C. The phosphofructokinase monomers were reassembled from 0.8 M guanidine chloride by 1:10 dilution of the guanidine hydrochloride concentration and yielded a protein with 70-94% of the original activity, depending on the protein concentration. The reactivation process follows second-order kinetics; ATP (5 mM) increases the rate of reactivation without altering the reaction order, while fructose 6-phosphate does not influence the rate of reaction. The rate-determining step is probably the association of monomers to form the dimer.  相似文献   

13.
Carbon monoxide binding to Chromatium vinosum ferrocytochrome c' has been studied by high-precision equilibrium methods. In contrast to the CO binding properties of Rhodospirillum molischianum cytochrome c' [Doyle, M. L., Weber, P. C., & Gill, S. J. (1985) Biochemistry 24, 1987-1991], CO binding to C. vinosum cytochrome c' is found to be unusual in the following ways. The binding curve is found to be cooperative with typical Hill coefficients equal to 1.25. The shape of the binding curve is asymmetrical. The heat of CO ligation is measured by two independent methods, both of which yield large endothermic values of approximately 10 kcal [mol of CO(aq)]-1. The overall affinity for CO increases as the concentration of cytochrome c' decreases. These observations suggest the CO binding properties of C. vinosum cytochrome c' are complicated by CO-linked association-dissociation processes. Further investigation by gel filtration chromatography shows that at micromolar concentrations the dimeric state is tightly associated in both the reduced and oxidized forms of the cytochrome but addition of saturating concentrations of CO causes the reduced ligated dimer to dissociate largely into monomers. A model is presented that quantitatively fits the data, involving a ligand-linked dimer-monomer dissociation reaction. In this model, CO binds to the dimer form noncooperatively with an intrinsic affinity constant equal to 5600 +/- 1200 M-1 at 25 degrees C. The unligated dimer form is tightly associated, but addition of CO causes dissociation of the dimer into the monomer with a monomer-dimer association constant equal to 450 +/- 200 M-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Alkaline phosphatase (AP) from Atlantic cod (Gadus morhua) is a zinc and magnesium containing homodimer that requires the oligomeric state for activity. Its kinetic properties are indicative of cold-adaptation. Here, the effect of urea on the structural stability was studied in order to correlate the activity with metal content, the microenvironment around tryptophan residues, and events at the subunit interface. At the lowest concentrations of urea, the first detected alteration in properties was an increase in the activity of the enzyme. This was followed by inactivation, and the release of half of the zinc content when the amount of urea reached levels of 2 M. Intrinsic tryptophan fluorescence and circular dichroism ellipticity changed in the range 2.5 to 8 M urea, signaling dimer dissociation, followed by one major monomer unfolding transition at 6-8 M urea as indicated by ANS fluorescence and KI fluorescence quenching. Gibbs free energy was estimated by the linear extrapolation method using a three-state model as 8.6 kcal/mol for dimer stability and 11.6 kcal/mol for monomer unfolding giving a total of 31.8 kcal/mol. Dimer association had a very small ionic contribution. Dimers were stable in relatively high concentration of urea, whereas the immediate vicinity around the active site was vulnerable to low concentrations of urea. Thus, inactivation did not coincide with dimer dissociation, suggesting that the active site is the most dynamic part of the molecule and closest related to cold-adaptation of its enzymatic activity.  相似文献   

15.
M A Griep  C S McHenry 《Biochemistry》1988,27(14):5210-5215
The beta subunit of Escherichia coli DNA polymerase III holoenzyme binds Mg2+. Reacting beta with fluoresceinmaleimide (FM) resulted in one label per beta monomer with full retention of activity. Titration of FM-beta with Mg2+ resulted in a saturable 11% fluorescence enhancement. Analysis indicated that there was one noncooperative magnesium binding site per beta monomer with a dissociation constant of 1.7 mM. Saturable fluorescence enhancement was also observed when titration was with Ca2+ or spermidine(3+) but not with the monovalent cations Na+ and K+. The Mg2+-induced fluorescence enhancement was specific for FM-beta and was not observed with FM-glutathione, dimethoxystilbenemaleimide-beta, or pyrenylmaleimide-beta. Gel filtration studies indicated that the beta dimer-monomer dissociation occurred at physiologically significant beta concentrations and that the presence of 10 mM Mg2+ shifted the dimer-monomer equilibrium to favor monomers. Both the gel-filtered dimers and the gel-filtered monomers were active in the replication assay. These and other results suggested that the fluorescence increase which accompanies beta dissociation is due to a relief from homoquenching of FM when the beta dimer dissociates into monomers.  相似文献   

16.
Binding of Nile Red to tubulin enhances and blue-shifts fluorescence emission to about 623 nm with a "shoulder" around 665 nm. Binding is reversible and saturable with an apparent Kd of approximately 0.6 microM. Nile Red does not alter tubulin polymerization, and polymerization in 2-(N-morpholino)ethanesulfonic acid (Mes) buffer does not alter the spectrum of the Nile Red-tubulin complex. In contrast, polymerization in glutamate buffer results in a red shift, reduction of intensity, and a decrease in lifetime, suggesting an increase in "polarity" of the binding environment. Lifetimes of 4.5 and 0.6 ns fluorescence in Mes buffer are associated with the 623-nm peak and the 665-nm shoulder, respectively. Indirect excitation spectra for these components are distinct and the 4.5-ns component exhibits tryptophan to Nile Red energy transfer. Acrylamide quenching yields linear Stern-Volmer plots with unchanged lifetimes, indicating static quenching. Apparent quenching constants are wavelength-dependent; global analysis reveals a quenchable component corresponding to the 4.5 ns component and an "unquenchable" component superposing the 0.6-ns spectrum. Analysis of anisotropy decay required an "associative" model which yielded rotational correlation times of greater than 50 ns for the 4.5-ns lifetime and 0.3 ns for the 0.6-ns lifetime. Dilution of tubulin in Mes results in an apparent red shift of emission without lifetime changes, due only to loss of the 623-nm component. These data are reconciled in terms of a model with two binding sites on the tubulin dimer. The more "nonpolar" site is located in a region of subunit-subunit contact which accounts for the fluorescence changes upon dilution; this permits estimation of a subunit dissociation constant of 1 microM.  相似文献   

17.
The influence of substrates and cofactors on the oligomeric structure of the cytosolic form of NADP+-specific isocitrate dehydrogenase (IDH) from lactating bovine mammary gland was investigated using analytical ultracentrifugation and kinetic methods. In guanidine-HCl, the monomer molecular weight for reduced and carboxymethylated IDH was found to be 50,000 to 52,000. In nondenaturing solvents IDH behaves as a homogeneous solute with a molecular weight of 97,200. When added separately, manganous isocitrate, isocitrate, manganous citrate (substrate analog), and a mixture of the substrate analog and NADP+ do not significantly alter the sedimentation coefficient or the molecular weight of IDH as judged by direct observation of the enzyme at 0.1 to 3 microM using sedimentation velocity and equilibrium. Active enzyme sedimentation (AES) was used to assess the degree of dissociation of IDH at lower concentrations, and Kd for the dimer-monomer equilibrium was estimated to be 2 nM. In enzymatic studies, the specific activity at several levels of substrate does not vary as the subunit concentration of enzyme is reduced from 10 to 0.3 nM. Estimates for Kd by AES indicate the presence of a significant fraction of monomer at assay concentrations of 1 nM and below, where the weight fraction of monomer is predicted to be 0.6. If the monomer has a lower activity than the dimer, a drop in specific activity is expected below 1 nM. Significant decreases occur only when the IDH is not protected from denaturation. The concentration of cytoplasmic IDH in bovine mammary tissue is estimated to be 5.7 microM, at least 100-fold greater than our estimates of Kd. Since over 90% of the enzyme is present in the dimeric form, ligand-induced changes in aggregation state cannot play a significant role in the regulation of the cytosolic form of IDH in situ in this tissue.  相似文献   

18.
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   

19.
A site-directed mutation, F235C, was created at the penultimate residue of the lambda-repressor. Measurement of dimer-monomer dissociation constant suggested that dimer-monomer dissociation of the mutant repressor is similar to that of the wild-type. Affinity towards a single operator O(R)1 is also similar to that of the wild-type repressor. The mutant repressor gene in a multi-copy plasmid confers immunity towards infection by a cI(-) lambda phage, suggesting preservation of functional integrity. Far-UV circular dichroism spectra show no major change in the secondary structure. Fluorescence quenching experiments, however, suggest increased exposure of some tryptophan residues. The urea denaturation profile indicates decreased stability of a part of the C-terminal domain. Under non-denaturing conditions, cysteine-235 shows half-of-the-sites reactivity, i.e. on average only one out of two cysteine-235 residues in the dimer shows reactivity towards sulfhydryl reagents. Fluorescence energy transfer between randomly labeled donor and acceptor fluorescent probes indicates that only one sulfhydryl per dimer is reactive, suggesting true half-of-the-sites reactivity. The structural role of the C-terminal tail in the whole repressor dimer is discussed.  相似文献   

20.
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   

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