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 93 毫秒
1.
Protein interactions with urea and guanidinium chloride. A calorimetric study.   总被引:33,自引:0,他引:33  
The interaction of urea and guanidinium chloride with proteins has been studied calorimetrically by titrating protein solutions with denaturants at various fixed temperatures, and by scanning them with temperature at various fixed concentrations of denaturants. It has been shown that the observed heat effects can be described in terms of a simple binding model with independent and similar binding sites. Using the calorimetric data, the number of apparent binding sites for urea and guanidinium chloride have been estimated for three proteins in their unfolded and native states (ribonuclease A, hen egg white lysozyme and cytochrome c). The intrinsic and total thermodynamic characteristics of their binding (the binding constant, the Gibbs energy, enthalpy, entropy and heat capacity effect of binding) have also been determined. It is found that the binding of urea and guanidinium chloride by protein is accompanied by a significant decrease of enthalpy and entropy. At all concentrations of denaturants the enthalpy term slightly dominates the entropy term in the Gibbs energy function. Correlation analysis of the number of binding sites and structural characteristics of these proteins suggests that the binding sites for urea and guanidinium chloride are likely to be formed by several hydrogen bonding groups. This type of binding of the denaturant molecules should lead to a significant restriction of conformational freedom within the polypeptide chain. This raises a doubt as to whether a polypeptide chain in concentrated solutions of denaturants can be considered as a standard of a random coil conformation.  相似文献   

2.
The guanidinium chloride- and urea-induced unfolding of FprA, a mycobacterium NADPH-ferredoxin reductase, was examined in detail using multiple spectroscopic techniques, enzyme activity measurements and size exclusion chromatography. The equilibrium unfolding of FprA by urea is a cooperative process where no stabilization of any partially folded intermediate of protein is observed. In comparison, the unfolding of FprA by guanidinium chloride proceeds through intermediates that are stabilized by interaction of protein with guanidinium chloride. In the presence of low concentrations of guanidinium chloride the protein undergoes compaction of the native conformation; this is due to optimization of charge in the native protein caused by electrostatic shielding by the guanidinium cation of charges on the polar groups located on the protein side chains. At a guanidinium chloride concentration of about 0.8 m, stabilization of apo-protein was observed. The stabilization of apo-FprA by guanidinium chloride is probably the result of direct binding of the Gdm+ cation to protein. The results presented here suggest that the difference between the urea- and guanidinium chloride-induced unfolding of FprA could be due to electrostatic interactions stabilizating the native conformation of this protein.  相似文献   

3.
Human fibrinogen was clotted under conditions that promote latent factor XIII activity and in the presence of a radioactive substitute cross-linking donor ([14C]glycine ethyl ester). The labeled fibrin was reduced and alkylated in the presence of 6 M guanidinium chloride. After dialysis and freeze-drying, the preparation was separated into its constituent polypeptide subunits by chromatography on (carboxymethyl)cellulose in the presence of 8 M urea. Under the incorporation conditions used, the radioactivity was limited to gamma chains (one donor molecule/chain) and alpha chains (two donor molecules/chain). The labeled alpha chains were digested with cyanogen bromide and fractionated on Sephadex G-50. All the radioactivity was found in a fragment previously designated H alpha CNI, the largest of the cyanogen bromide fragments in the alpha chain. The fragment was further fragmented by digestion with plasmin, trypsin, chymotrypsin, and/or staphylococcal protease. The incorporated radioactivity was found to reside in equal amounts at two different sites located 38 residues apart. These were determined to be positions 88 and 126 in H alpha CNI, which correspond to glutamine-328 and glutamine-366 in the alpha chain.  相似文献   

4.
Endoglucanase D from Clostridium thermocellum was purified from inclusion bodies formed upon its overproduction in Escherichia coli, using 5 M urea as a solubilizing solution. We examined the effects of denaturing agents upon the stability of the pure soluble enzyme as a function of the temperature. At room temperature, guanidinium chloride induces an irreversible denaturation. By comparison, we observed no structural or functional effects at room temperature using high concentrations of urea as denaturing agent. The irreversible denaturation process observed with guanidinium chloride also occurs with urea but only at elevated temperature (greater than or equal to 60 degrees C); in 6 M urea, the activation energy of the denaturation reaction is decreased by a factor of only 1.8. We interpret the high resistance of this protein to urea as reflecting a reduced flexibility of its structure at normal temperatures which should be correlated to the thermophilic origin of this protein.  相似文献   

5.
The quaternary structure of streptavidin in urea   总被引:2,自引:0,他引:2  
We report on the interactions of urea and guanidinium salts with streptavidin. Gel filtration chromatography in 0, 4, 6, and 7 M urea indicates that the streptavidin tetramer remains intact in urea. Biotin alters the electrophoretic mobility of streptavidin whether or not 6 M urea is present. The intrinsic fluorescence of streptavidin is increased and blue-shifted in 6 M urea. The fluorescence changes indicate the absence of unfolding. A conformational response to urea is possible, but much of the fluorescence change is due to urea binding as a weak biotin analog (Ka approximately 1.3 M-1). The resistance to structural perturbation by urea reflects the structural stability of streptavidin's anti-parallel beta-barrel motif. Unfolding is sluggish in 6 M guanidinium hydrochloride (half-time, approximately 50 days). After guanidinium thiocyanate unfolding, streptavidin can be refolded, but the unfolding and refolding transitions are centered at different concentrations of perturbant. Slow unfolding, with a 15th power dependence on guanidinium thiocyanate concentration, may be partially responsible for the noncoincidence of the unfolding and refolding processes. Nonequilibrium behavior is also seen in 6 M urea, as native streptavidin does not unfold and guanidinium thiocyanate unfolded streptavidin does not refold. Refolding does occur at lower concentrations of urea. Guanidinium thiocyanate only slowly unfolds the biotin-streptavidin complex. In the presence of biotin, unfolded streptavidin does not refold in 6 M guanidinium thiocyanate or in 6 M urea.  相似文献   

6.
Two polypeptides of 74 kDa and 55 kDa have been isolated from human platelets by immunoaffinity and lectin affinity chromatography and their effects on thrombin reactivity have been examined. These proteins in combination enhanced the aggregation of platelets by thrombin while aggregation induced by trypsin, collagen and adenosine diphosphate was not significantly affected. An enhancement in the action of thrombin on fibrinogen, N-benzoylarginine ethyl ester and H-D-phenylalanyl-L-pipecolyl-L-arginine-p-nitroanilide dihydrochloride was also observed in the presence of the platelet proteins. Under similar conditions, the proteins did not influence the esterolytic activity of trypsin or plasmin. Studies at different thrombin and protein concentrations showed maximum enhancement of enzyme reactivity when the ratio between the peptides and thrombin was optimal. In the presence of these proteins, the affinity of thrombin for N-benzoylarginine ethyl ester was about twofold higher than in the control. Two polypeptides with properties similar to those described above have also been isolated from human plasma. Antibodies to the above proteins isolated from either platelets or plasma were raised in rabbits. Intact platelets solubilized in Triton X-100 or plasma showed two precipitin lines in immunoelectrophoresis against both of the above antisera and a similar pattern was observed with the isolated polypeptides. The polypeptides did not interact in immunoelectrophoresis with antisera to whole serum, antithrombin, C4 binding protein or protein S. These 74-kDa and 55-kDa polypeptides contained radioactivity when radioiodinated platelets were used suggesting that they are located on the cell surface. Fresh plasma was analyzed by gel electrophoresis under nondenaturing and denaturing conditions and the proteins were transferred to nitrocellulose sheets. Staining with antibody to these thrombin-reactive proteins and 125I-protein A showed several reactive plasma proteins under nondenaturing conditions with the major band migrating in the albumin area. In plasma treated with sodium dodecyl sulfate, the 74-kDa and 55-kDa components were observed. A prominent 74-kDa band and a fainter 55-kDa component were again observed when platelets solubilized in sodium dodecyl sulfate were analysed by the above procedure. It is proposed that human platelets and plasma contain polypeptides which may directly modulate thrombin reactivity.  相似文献   

7.
The copper- and zinc-containing superoxide dismutase of bovine erythrocytes retains its native molecular weight of 32 000 in 8.0 M urea for at least 72 h at 25 degrees C, as evidenced by sedimentation equilibrium analysis. Subsequent to prolonged exposure to urea, the dimeric enzyme could be dissociated by sodium dodecyl sulfate in the absence of reductants, indicating the absence of unnatural disulfide cross-links. The sulfhydryl group of cysteine-6 was unreactive toward 5,5'-dithiobis(2-nitrobenzoic acid) or bromoacetic acid in both neutral buffer and 8.0 M urea. The histidine residues of the enzyme were resistant to carboxymethylation in neutral buffer and 8.0 M urea. However, when the enzyme was exposed to bromoacetic acid in the presence of 6.0 M guanidinium chloride and 1 mM (ethylenedinitriol)tetraacetic acid (EDTA), both sulfhydryl and histidine alkylation were observed. Guanidinium chloride (6.0 M) increased the reactivity of the sulfhydryl group of cysteine-6 and allowed the oxidative formation of disulfide-bridged dimers. This was prevented by 1 mM EDTA. It follows that 8.0 M urea neither dissociates the native enzyme into subunits nor produces a conformation detectably different than that possessed under native conditions.  相似文献   

8.
Chemical denaturants are frequently used to unfold proteins and to characterize mechanisms and transition states of protein folding reactions. The molecular basis of the effect of urea and guanidinium chloride (GdmCl) on polypeptide chains is still not well understood. Models for denaturant--protein interaction include both direct binding and indirect changes in solvent properties. Here we report studies on the effect of urea and GdmCl on the rate constants (k(c)) of end-to-end diffusion in unstructured poly(glycine-serine) chains of different length. Urea and GdmCl both lead to a linear decrease of lnk(c) with denaturant concentration, as observed for the rate constants for protein folding. This suggests that the effect of denaturants on chain dynamics significantly contributes to the denaturant-dependence of folding rate constants for small proteins. We show that this linear dependency is the result of two additive non-linear effects, namely increased solvent viscosity and denaturant binding. The contribution from denaturant binding can be quantitatively described by Schellman's weak binding model with binding constants (K) of 0.62(+/-0.01)M(-1) for GdmCl and 0.26(+/-0.01)M(-1) for urea. In our model peptides the number of binding sites and the effect of a bound denaturant molecule on chain dynamics is identical for urea and GdmCl. The results further identify the polypeptide backbone as the major denaturant binding site and give an upper limit of a few nanoseconds for residence times of denaturant molecules on the polypeptide chain.  相似文献   

9.
Lipid content of purified fatty acid synthetase preparations from the Dipterous Ceratitis capitata correlated with the enzyme activity. Delipidation of the enzyme by extracting with a series of organic solvents rendered a protein without any residual activity and the treatment with phospholipase A2 for 30 min reduced the activity to 10%. Addition of lipid classes to either the native enzyme or the phospholipase-treated preparation enhanced the activity in a different manner, phosphatidylethanolamine being the most effective lipid. The role of the lipids in the lipoprotein structure of the complex was studied by circular dichroism spectra of the native enzyme and in the presence of a concentration range of urea, guanidinium chloride, sodium dodecyl sulfate, sodium cholate, and sodium chloride. 3 M urea and 1.5 M guanidinium chloride induced a conformational transition of the lipoprotein that lost its alpha-helical structure at higher concentrations of both reagents. Sodium dodecyl sulfate and sodium cholate had little effect on the alpha-helical structure, although both reagents induced the loss of enzyme activity. Cholate had essentially the same effect as phospholipids on the maintenance of the native structure but it was unable to support the enzyme activity.  相似文献   

10.
T A Alston  R H Abeles 《Biochemistry》1987,26(13):4082-4085
L-Histidine methyl ester inactivates histidine decarboxylase in a time-dependent manner. The possibility was considered that an irreversible reaction between enzyme and inhibitor occurs [Recsei, P. A., & Snell, E. E. (1970) Biochemistry 9, 1492-1497]. We have confirmed time-dependent inactivation by histidine methyl ester and have investigated the structure of the enzyme-inhibitor complex. Upon exposure to either 8 M guanidinium chloride or 6% trichloroacetic acid, unchanged histidine methyl ester is recovered. Formation of the complex involves Schiff base formation, most likely with the active site pyruvyl residue [Huynh, Q. K., & Snell, E. E. (1986) J. Biol. Chem. 261, 4389-4394], but does not involve additional irreversible covalent interaction between inhibitor and enzyme. Complex formation is a two-step process involving rapidly reversible formation of a loose complex and essentially irreversible formation of a tight complex. For the formation of the tight complex, Ki = 80 nM and koff = 2.5 X 10(-4) min-1. Time-dependent inhibition was also observed with L-histidine ethyl ester, L-histidinamide, and DL-3-amino-4-(4-imidazolyl)-2-butanone. No inactivation was observed with glycine methyl ester or histamine. We propose that in the catalytic reaction the carboxyl group of the substrate is in a hydrophobic region. The unfavorable interaction between the carboxylate group and the hydrophobic region facilitates decarboxylation [Crosby, J., Stone, R., & Liehard, G. E. (1970) J. Am. Chem. Soc. 92, 2891-2900]. With histidine methyl ester this unfavorable interaction is no longer present; hence, there is tight binding.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Felitsky DJ  Record MT 《Biochemistry》2004,43(28):9276-9288
Two thermodynamic models have been developed to interpret the preferential accumulation or exclusion of solutes in the vicinity of biopolymer surface and the effects of these solutes on protein processes. The local-bulk partitioning model treats solute (and water) as partitioning between the region at/or near the protein surface (the local domain) and the bulk solution. The solvent exchange model analyzes a 1:1 competition between water and solute molecules for independent surface sites. Here we apply each of these models to interpret thermodynamic data for the interactions of urea and the osmoprotectant glycine betaine (N,N,N-trimethylglycine; GB) with the surface exposed in unfolding the marginally stable lacI HTH DNA binding domain. The partition coefficient K(P) quantifying accumulation of urea at this protein surface (K(P) approximately equal 1.1) is only weakly dependent on urea concentration up to 6 M urea. However, K(P) quantifying exclusion of GB from the vicinity of this protein surface increases from 0.83 (extrapolated to 0 M GB) to 1.0 (indicating that local and bulk GB concentrations are equal) at 4 M GB (activity > 40 M). We interpret the significant concentration dependence of K(P) for GB, predicted to be general for excluded, nonideal solutes such as GB, as a modest (8%) attenuation of the GB concentration dependence of solute nonideality in the local domain relative to that in the bulk solution. Above 4 M, K(P) for the interaction of GB with the surface exposed in protein unfolding is predicted to exceed unity, which explains the maximum in thermal stability observed for RNase and lysozyme at 4 M GB (Santoro, M. M., Liu, Y. F., Khan, S. M. A., Hou, L. X., and Bolen, D. W. (1992) Biochemistry 31, 5278-5283). Both thermodynamic models provide good two-parameter fits to GB and urea data for lacI HTH unfolding over a wide concentration range. The solute partitioning model allows for a full spectrum of attenuation effects in the local domain, encompasses the cases treated by the competitive binding model, and provides a somewhat better two-parameter fit of effects of high GB concentration on lacI HTH stability. Parameters of this fit should be applicable to isothermal and thermal unfolding data for all proteins with similar compositions of surface exposed in unfolding.  相似文献   

12.
Heparin cofactor, a thrombin inhibitor, is purified from human plasma by affinity chromatography on heparin-agarose. The nature of the binding between thrombin and the inhibitor is studied by treatment of the complex with 6 M guanidinium chloride, hydroxylamine, and dilute alkali. The complex is not dissociated during gel chromatography in 6 M guanidinium chloride. This result supports an earlier proposal that formation of the complex includes the formation of a covalend bond. Treatment of dodecylsulfate-denatured complex with hydroxylamine results in dissociation of the complex to yield free thrombin and heparin cofactor. The complex is also dissociated in dilute NaOH (pH 12) solutions. These results indicate that the covalent bond between thrombin and the inhibitor is a carboxylic ester.  相似文献   

13.
To interpret effects of urea and guanidinium (GuH(+)) salts on processes that involve large changes in protein water-accessible surface area (ASA), and to predict these effects from structural information, a thermodynamic characterization of the interactions of these solutes with different types of protein surface is required. In the present work we quantify the interactions of urea, GuHCl, GuHSCN, and, for comparison, KCl with native bovine serum albumin (BSA) surface, using vapor pressure osmometry (VPO) to obtain preferential interaction coefficients (Gamma(mu3)) as functions of nondenaturing concentrations of these solutes (0-1 molal). From analysis of Gamma(mu3) using the local-bulk domain model, we obtain concentration-independent partition coefficients K(nat)(P) that characterize the accumulation of these solutes near native protein (BSA) surface: K(nat)(P,urea)= 1.10 +/- 0.04, K(nat)(P,SCN(-)) = 2.4 +/- 0.2, K(nat)(P,GuH(+)) = 1.60 +/- 0.08, relative to K(nat)(P,K(+)) identical with 1 and K(nat)(P,Cl(-)) = 1.0 +/- 0.08. The relative magnitudes of K(nat)(P) are consistent with the relative effectiveness of these solutes as perturbants of protein processes. From a comparison of partition coefficients for these solutes and native surface (K(nat)(P)) with those determined by us previously for unfolded protein and alanine-based peptide surface K(unf)(P), we dissect K(P) into contributions from polar peptide backbone and other types of protein surface. For globular protein-urea interactions, we find K(nat)(P,urea) = K(unf)(P,urea). We propose that this equality arises because polar peptide backbone is the same fraction (0.13) of total ASA for both classes of surface. The analysis presented here quantifies and provides a physical basis for understanding Hofmeister effects of salt ions and the effects of uncharged solutes on protein processes in terms of K(P) and the change in protein ASA.  相似文献   

14.
1. The interactions of histone fractions with 8-anilinonaphthalene-1-sulphonic acid were investigated by fluorimetry and spectrofluorimetry and the results were interpreted with the aid of equilibrium-dialysis techniques. 2. Characteristic differences were found between the various histone fractions, and with fractions F3 and F2a the binding was found to be salt-dependent. 3. Evidence was obtained indicating a slow change of the physical state of fractions F3 and F2a in the presence of salt, and the binding by these two fractions in the presence of salt was greater by an order of magnitude than by fractions F1 and F2b. 4. Conditions favouring binding were also those favouring histone aggregation; SO(4) (2-) ions activated binding at a lower concentration than Cl(-) ions; urea, guanidinium ions and high concentrations of I(-) ions were inhibitory to binding. 5. After histones had been kept in the presence of salt for a long time the reversal of interaction on decreasing the salt concentration was incomplete. 6. The inhibition of binding by fraction F2a in the presence of urea or fraction F2b depended on the time sequence of addition of the reagents. 7. Artificial nucleoproteins made by precipitating DNA with the histone fractions in neutral 0.14m-sodium chloride showed the same order of interaction as was found for the fractions in solution. 8. Comparison of the binding by fraction F2a with that by bovine plasma albumin showed that in both cases there were a large number of weakly binding sites but that fraction F2a lacked the small number of strongly binding sites found in albumin. No slow change of binding in the presence of salt was found for albumin. 9. Binding by fraction F2b increased the affinity of the protein for further molecules of the adsorbate. 10. The results are discussed in relation to the close relationship between binding and aggregation and the possible role of non-polar interactions as determined by the balance between polar and non-polar amino acids in the histone fractions.  相似文献   

15.
The properties of a simple model for solvation in mixed solvents are explored in this paper. The model is based on the supposition that solvent replacement is a simple one-for-one substitution reaction at macromolecular sites which are independent of one another. This leads to a new form for the binding polynomial in which all terms are associated with ligand interchange rather than ligand addition. The principal solvent acts as one of the ligands. Thermodynamic analysis then shows that thermodynamic binding (i.e., selective interaction) depends on the properties of K'-1, whereas stoichiometric binding (site occupation) depends on K'. K' is a 'practical' interchange equilibrium constant given by (f3/f1)K, where K is the true equilibrium constant for the interchange of components 3 and 1 on the site and f3 and f4 denote their respective activity coefficients on the mole fraction scale. Values of K' less than unity lead to negative selective interaction. It is selective interaction and not occupation number which determines the thermodynamic effects of solvation. When K' greater than 100 on the mole fraction scale or K' greater than 2 on the molality scale (in water), the differences between stoichiometric binding and selective interaction become less than 1%. The theory of this paper is therefore necessary only for very weak binding constants. When K'-1 is small, large concentrations of the added solvent component are required to produce a thermodynamic effect. Under these circumstances the isotherms for the selective interaction and for the excess (or transfer) free energy are strongly dependent on the behavior of the activity coefficients of both solvent components. Two classes of behavior are described depending on whether the components display positive or negative deviations from Raoult's law. Examples which are discussed are aqueous solutions of urea and guanidinium chloride for positive deviations and of sucrose and glucose for negative deviations. Examination of the few studies which have been reported in the literature shows that most of the qualitative features of the stabilization of proteins by sugars and their destabilization by urea and guanidinium chloride are faithfully represented with the model. This includes maxima in the free energy of stabilization and destabilization, decreased and zero selective interaction at high concentrations, etc. These phenomena had no prior explanation. Deficiencies in the model as a representation of solvation in aqueous solution are discussed in the appendix.  相似文献   

16.
Replacement of the Arg residue at position 82 in bacteriorhodopsin by Gln or Ala was previously shown to slow the rate of proton release and raise the pK of Asp 85, indicating that R82 is involved both in the proton release reaction and in stabilizing the purple form of the chromophore. We now find that guanidinium chloride lowers the pK of D85, as monitored by the shift of the 587-nm absorbance maximum to 570 nm (blue to purple transition) and increased yield of photointermediate M. The absorbance shift follows a simple binding curve, with an apparent dissociation constant of 20 mM. When membrane surface charge is taken into account, an intrinsic dissociation constant of 0.3 M fits the data over a range of 0.2-1.0 M cation concentration (Na+ plus guanidinium) and pH 5.4-6.7. A chloride counterion is not involved in the observed spectral changes, as chloride up to 0.2 M has little effect on the R82Q chromophore at pH 6, whereas guanidinium sulfate has a similar effect to guanidinium chloride. Furthermore, guanidinium does not affect the chromophore of the double mutant R82Q/D85N. Taken together, these observations suggest that guanidinium binds to a specific site near D85 and restores the purple chromophore. Surprisingly, guanidinium does not restore rapid proton release in the photocycle of R82Q. This result suggests either that guanidinium dissociates during the pump cycle or that it binds with a different hydrogen-bonding geometry than the Arg side chain of the wild type.  相似文献   

17.
Karmodiya K  Surolia N 《Proteins》2008,70(2):528-538
The urea and guanidinium chloride (GdmCl) induced unfolding of FabG, a beta-ketoacyl-ACP reductase of Plasmodium falciparum, was examined in detail using intrinsic fluorescence of FabG, UV-circular dichroism (CD), spectrophotometric enzyme activity measurements, glutaraldehyde cross-linking, and size exclusion chromatography. The equilibrium unfolding of FabG by urea is a multistep process as compared with a two-state process by GdmCl. FabG is fully unfolded at 6.0M urea and 4.0M GdmCl. Approximately 90% of the enzyme activity could be recovered on dialyzing the denaturants, showing that denaturation by both urea and GdmCl is reversible. We found two states in the reversible unfolding process of FabG in presence of NADPH; one is an activity-enhanced state and the other, an inactive state in case of equilibrium unfolding with urea. On the contrary, in presence of NADPH, there is no stabilization of FabG in case of equilibrium unfolding with GdmCl. We hypothesize that the hydrogen-bonding network may be reorganized by the denaturant in the activity-enhanced state formed in presence of 1.0M urea, by interrupting the association between dimer-dimer interface and help in accommodating the larger substrate in the substrate binding tunnel thus, increasing the activity. Furthermore, binding of the active site organizer, NADPH leads to compaction of the FabG in presence of urea, as evident by acrylamide quenching. We have shown here for the first time, the detailed inactivation kinetics of FabG, which have not been evaluated in the past from any of the FabG family of enzymes from any of the other sources. These findings provide impetus for exploring the influences of ligands on the structure-activity relationship of Plasmodium beta-ketoacyl-ACP reductase.  相似文献   

18.
Solution properties of beta recombinase were studied by circular dichroism and fluorescence spectroscopy, size exclusion chromatography, analytical ultracentrifugation, denaturant-induced unfolding and thermal unfolding experiments. In high ionic strength buffer (1 M NaCl) beta recombinase forms mainly dimers, and strongly tends to aggregate at ionic strength lower than 0.3 M NaCl. Urea and guanidinium chloride denaturants unfold beta recombinase in a two-step process. The unfolding curves have bends at approximately 5 M and 2.2 M in urea and guanidinium chloride-containing buffers. Assuming a three-state unfolding model (N2-->2I-->2U), the total free energy change from 1 mol of native dimers to 2 mol of unfolded monomers amounts to deltaG(tot) = 17.9 kcal/mol, with deltaG(N2-->2I) = 4.2 kcal/mol for the first transition and deltaG(I-->U) = 6.9 kcal/mol for the second transition. Using sedimentation-equilibrium analytical ultracentrifugation, the presence of beta recombinase monomers was indicated at 5 M urea, and the urea dependence of the circular dichroism at 222 nm strongly suggests that folded monomers represent the unfolding intermediate.  相似文献   

19.
Cathepsin C, a tetrameric lysosomal dipeptidyl-peptide hydrolase, is activated by chloride ion. The activation is shown here to be specific and pH-dependent, dissociation constants for chloride being lower at low pH. Bound chloride decreases the Km for the hydrolysis of chromophore labelled substrates without any significant change in Vmax, confirming its involvement in substrate binding. Determination of the kinetic parameters of chloride activation, using unlabelled substrates, has enabled its site of action to be located. The lower Km for the hydrolysis of simple amide substrates in the presence of Cl- shows that the S sites are involved. Possible involvement of the S' sites is excluded by the finding that the Km for the nucleophile in the transferase reaction is unaffected by chloride. The rates of inhibition by E-64 and iodoacetate are both chloride-dependent and, from the structure of the papain-E-64 complex, it is concluded that chloride binds close to the S2 site. The binding of guanidinium ion, a positively charged inhibitor, to the S site is dependent on chloride. Based on these results, a model is proposed to explain the chloride activation of cathepsin C. The possible physiological role of chloride in the regulation of proteolysis in the lysosome is discussed.  相似文献   

20.
The unfolding of shikimate kinase (SK) from Erwinia chrysanthemi by urea and its subsequent refolding on dilution of the denaturing agent has been studied in detail [Eur. J. Biochem. 269 (2002) 2124]. Comparison of the effects of urea on the enzyme with those of guanidinium chloride (GdmCl) and NaCl indicated that chloride ions significantly weakened the binding of shikimate. This finding prompted a more detailed examination of the effects of salts on the structure, function and stability of the enzyme; the effects of NaCl and Na(2)SO(4) were investigated in detail. These salts have very small effects on the secondary structure as judged by far UV CD circular dichroism (CD), although the near UV CD spectra suggest that some limited changes in the environment of aromatic amino acids may occur. Both salts inhibit SK activity and analysis of the kinetic and substrate binding parameters point to a complex mechanism for the inhibition. Inclusion of salts leads to a marked stabilisation against unfolding of the enzyme by urea. When the enzyme is unfolded by incubation in 4 M urea, addition of NaCl or Na(2)SO(4) leads to a relatively slow refolding of the enzyme as judged by the regain of native-like CD and fluorescence. In addition, the refolded enzyme can bind shikimate, though more weakly than the native enzyme. However, the refolded enzyme does not appear to be capable of binding nucleotides, nor does it possess detectable catalytic activity. The refolding process brought about by addition of salt in the presence of 4 M urea is not associated with any change in the fluorescence of the probe 8-anilino-1-naphthalenesulfonic acid (ANS), indicating that an intermediate formed by hydrophobic collapse is unlikely to be significantly populated. The results point to both specific and general effects of salts on SK. These are discussed in the light of the structural information available on the enzyme.  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