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 773 毫秒
1.
For the first time, the enzyme rhodanese (thiosulfate:cyanide sulfurtransferase; EC 2.8.1.1) has been renatured from 6 M guanidinium chloride (GdmCl) by direct dilution of the denaturant at relatively high protein concentrations. This has been made possible by using the nonionic detergent dodecyl-beta-D-maltoside (lauryl maltoside). Lauryl maltoside concentration dependence of the renaturation and reactivation time courses were studied using 50 micrograms/ml rhodanese. There was no renaturation at lauryl maltoside (less than 0.1 mg/ml), and the renaturability increased, apparently cooperatively, up to 5 mg/ml detergent. This may reflect weak binding of lauryl maltoside to intermediate rhodanese conformers. The renaturability began to decrease above 5 mg/ml lauryl maltoside and was significantly reduced at 20 mg/ml. Individual progress curves of product formation, for rhodanese diluted into lauryl maltoside 90 min before assay, showed induction phases as long as 7 min before an apparently linear steady state. The induction phase increased with lauryl maltoside concentration and could even be observed in native controls above 1 mg/ml detergent. These results are consistent with suggestions that refolding of GdmCl-denatured rhodanese involves an intermediate with exposed hydrophobic surfaces that can partition into active and inactive species. Further, lauryl maltoside can stabilize those surfaces and prevent aggregation and other hydrophobic interaction-dependent events that reduce the yield of active protein. The rhodanese-lauryl maltoside complex could also form with native enzyme, thus explaining the induction phase with this species. Finally, it is suggested that renaturation of many proteins might be assisted by lauryl maltoside or other "nondenaturing" detergents.  相似文献   

2.
F X Schmid 《FEBS letters》1986,198(2):217-220
The trans----cis isomerization of Pro 93 was measured during refolding of bovine ribonuclease A. This isomerization is slow (tau = 500 s) under marginally stable folding conditions of 2.0 M GdmCl, pH 6, at 10 degrees C. However, it is strongly accelerated (tau = 100 s) in samples which, prior to isomerization, had been converted to a folding intermediate by a 15 s refolding pulse under strongly native conditions (0.8 M ammonium sulfate, 0 degree C). The results demonstrate that extensive folding is possible before Pro 93 isomerizes to its native cis state and that the presence of structural folding intermediates leads to a marked increase in the rate of subsequent proline isomerization.  相似文献   

3.
The refolding and unfolding kinetics of the all-beta-sheet protein human basic fibroblast growth factor (hFGF-2) were studied by fluorescence spectroscopy. The kinetics of the unfolding transition are monophasic. The refolding reaction at high and low guanidinium chloride (GdmCl) concentrations is best described by mono- and biphasic folding, respectively. Refolding and unfolding of hFGF-2 (155 amino acids) is very slow compared with other non-disulfide-bonded monomeric proteins of similar size. For example, the rate constant for unfolding at 4.5 mol.liter(-1) GdmCl is 0.006 s(-1), and the refolding rate constants at 0.4 mol.liter(-1) GdmCl are 0.01 s(-1) and 0.0009 s(-1) (15 degrees C, pH 7.0). A characterization of the thermodynamic nature of the folding process using transition state theory revealed that the slow refolding is almost exclusively controlled by entropic factors, namely the strong loss of conformational freedom during refolding. The rate of the slow unfolding kinetics is mainly (and at low denaturant concentrations exclusively) controlled by the large positive change in enthalpy. hFGF-2 shows similar slow folding kinetics to that of its structural homolog interleukin-1beta. Since both proteins show very little sequence identity, it is suggested that their slow folding kinetics are determined by the complex beta-sheet arrangement of the native molecules.  相似文献   

4.
It has been recognized that the artificial chaperone system, cetyltrimethylammonium bromide and beta-cyclodextrin, is effective for enhancing protein renaturation. In this work, we studied the effect of the artificial chaperone system and guanidinium chloride (GdmCl) on the oxidative renaturation of lysozyme at 0.21-1.05 mg/mL, and a kinetic model based on the competition between protein folding and aggregation was employed to express the renaturation process. The refolding rate constant increased, while the aggregation rate constant decreased, with increasing concentration of the artificial chaperones. With increasing GdmCl concentration (0.28-2 M), both rate constants decreased, but there existed a specific GdmCl concentration that maximized the ratio of the two rate constants and thus the renaturation yield. The results obviously indicated the cooperative effect of GdmCl and the artificial chaperones on enhancing protein renaturation.  相似文献   

5.
We have previously described the complexity of the folding of the lipolytic enzyme cutinase from F. solani pisi in guanidinium chloride. Here we extend the refolding analysis by refolding from the pH-denatured state and analyze the folding behaviour in the presence of the weaker denaturant urea and the stronger denaturant guanidinium thiocyanate. In urea there is excellent consistency between equilibrium and kinetic data, and the intermediate accumulating at low denaturant concentrations is off-pathway. However, in GdmCl, refolding rates, and consequently the stability of the native state, vary significantly depending on whether refolding takes place from the pH- or GdmCl-denatured state, possibly due to transient formation of aggregates during folding from the GdmCl-denatured state. In GdmSCN, stability is reduced by several kcal/mol with significant aggregation in the unfolding transition region. The basis for the large variation in folding behaviour may be the denaturants' differential ability to support formation of exposed hydrophobic regions and consequent changes in aggregative properties during refolding.  相似文献   

6.
The reactivation efficiency in the refolding of denatured luciferase in the presence and the absence of monoclonal antibodies (mAbs) has been studied. Luciferase could be partially reactivated when the protein was denatured in high concentrations of guanidium chloride (GdmCl; >4.5 M) and the refolding was carried out in very low protein concentrations. The refolding yield was, however, significantly lower when it was performed on luciferase that had been denatured with lower concentrations of GdmCl. The efficiency of refolding decreases when the formation of aggregates increases. Three of the five luciferase mAbs tested (4G3, N2E3, S2G10) dramatically increased the yield of reactivation and simultaneously eliminated the formation of aggregates. It is proposed that these mAbs assisted the refolding of luciferase by binding to the exposed hydrophobic surface of the refolding intermediate, thus preventing it from aggregating. The epitopes interacting with these refolding-assisting mAbs are all located in the A-subdomain of the N-terminal region of luciferase. These results have also shed light on the structural features of the intermediate and its interface involved in protein aggregate formation, contributing to the understanding of the protein folding mechanism.  相似文献   

7.
Biosynthetic folding, beginning with the growing nascent chain and leading to the biologically active structure within its proper cellular context, is one function shared by all proteins. We show that the bacterial luciferase beta subunit reaches its final native form in the alphabeta heterodimer much more rapidly during biosynthetic folding than during refolding from urea. The rate of formation of active enzyme is determined by a short-lived folding intermediate, which is able to associate with the alpha subunit very rapidly following release from the ribosome. This intermediate appears to involve a transient interaction of the C-terminal region of the beta subunit, a region distant from the subunit interface, but intimately involved in heterodimerization. Refolding of the beta subunit under similar conditions proceeds much more slowly. We have characterized both pathways and show that the basic difference between biosynthetic folding and refolding from urea is that the newly synthesized beta subunit enters the folding pathway at a point beyond the slow, rate-determining step that limits the rate of the renaturation process and constitutes a kinetic trap. This mechanism embodies a major strategy, the avoidance of slow-folding intermediates and kinetic traps, that may be employed by many proteins to achieve fast and efficient biosynthetic folding.  相似文献   

8.
The chaperonin protein cpn60 from Escherichia coli protects the monomeric, mitochondrial enzyme rhodanese (thiosulfate:cyanide sulfurtransferase, EC 2.8.1.1) against heat inactivation. The thermal inactivation of rhodanese was studied for four different states of the enzyme: native, refolded, bound to cpn60 in the form of a binary complex formed from unfolded rhodanese, and a thermally perturbed state. Thermal stabilization is observed in a range of temperatures from 25 to 48 degrees C. Rhodanese that had been inactivated by incubation at 48 degrees C, in the presence of cpn60 can be reactivated at 25 degrees C, upon addition of cpn10, K+, and MgATP. A recovery of about 80% was achieved after 1 h of the addition of those components. Thus, the enzyme is protected against heat inactivation and kept in a reactivable form if inactivation is attempted using the binary complex formed between rhodanese folding intermediate(s) and cpn60. The chaperonin-assisted refolding of urea-denatured rhodanese is dependent on the temperature of the refolding reaction. However, optimal chaperonin assisted refolding of rhodanese observed at 25 degrees C, which is achieved upon addition of cpn10 and ATP to the cpn60-rhodanese complex, is independent of the temperature of preincubation of the complex, that was formed previously at low temperature. The results are in agreement with a model in which the chaperonin cpn60 interacts with partly folded intermediates by forming a binary complex which is stable to elevated temperatures. In addition, it appears that native rhodanese can be thermally perturbed to produce a state different from that achieved by denaturation that can interact with cpn60.  相似文献   

9.
Thein vitro refolding of the monomeric, mitochondrial enzyme rhodanese (thiosulfate: cyanide sulfurtransferase, EC 2.8.1.1), which is assisted by theE. coli chaperonins, is modulated by the 23 amino acid peptide (VHQVLYRALVSTKWLAESVRAGK) corresponding to the amino terminal sequence (1–23) of rhodanese. In the absence of the peptide, a maximum recovery of active enzyme of about 65% is achieved after 90 min of initiation of the chaperonin assisted folding reaction. In contrast, this process is substantially inhibited in the presence of the peptide. The maximum recovery of active enzyme is peptide concentration-dependent. The peptide, however, does not prevent the interaction of rhodanese with the chaperonin 60 (cpn60), which leads to the formation of the cpn60-rhodanese complex. In addition, the peptide does not affect the rate of recovery of active enzyme, although it does affect the extent of recovery. Further, the unassisted refolding of rhodanese is also inhibited by the peptide. Thus, the peptide interferes with the folding of rhodanese in either the chaperonin assisted or the unassisted refolding of the enzyme. A 13 amino acid peptide (STKWLAESVRAGK) corresponding to the amino terminal sequence (11–23) of rhodanese does not show any significant effect on the chaperonin assisted or unassisted refolding of the enzyme. The results suggest that other sequences of rhodanese, in addition to the N-terminus, may be required for the binding of cpn60, in accord with a model in which cpn60 interacts with polypeptides through multiple binding sites.  相似文献   

10.
Oxidative renaturation of lysozyme at high concentrations   总被引:18,自引:0,他引:18  
Newly synthesized cloned gene proteins expressed in bacteria frequently accumulate in insoluble aggregates or inclusion bodies. Active protein can be recovered by solubilization of inclusion bodies followed by renaturation of the solubilized (unfolded) protein. The recovery of active protein is highly dependent on the renaturation conditions chosen. The renaturation process is generally conducted at low protein concentrations (0.01-0.2 mg/mL) to avoid aggregation. We have investigated the potential of successfully refolding reduced and denatured hen egg white lysozyme at high concentrations (1 and 5 mg/mL). By varying the composition of the renaturation media, optimum conditions which kinetically favor proper folding over inactivation were found. Solubilizing agents such as guanidinium chloride (GdmCl) and folding aids such as L-arginine present in low concentrations during refolding effectively enhanced renaturation yields by suppressing aggregation resulting in reactivation yields as high as 95%. Quantitatively the kinetic competition between lysozyme folding and aggregation can be described using first-order kinetics for the renaturation reaction and third-order kinetics for the overall aggregation pathway. The rate constants for both reactions have been found to be strongly dependent on denaturant and thiol concentration. This strategy supercedes the necessity to reactivate proteins at low concentrations using large renaturation volumes. The marked increase in volumetric productivity makes this a viable option for recovering biologically active protein efficiently and in high yield in vitro from proteins produced as inclusion bodies within microbial cells. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 54: 221-230, 1997.  相似文献   

11.
The in vitro folding of rhodanese involves a competition between formation of properly folded enzyme and off-pathway inactive species. Co-solvents like glycerol or low temperature, e.g. refolding at 10 degrees C, successfully retard the off-pathway formation of large inactive aggregates, but the process does not yield 100% active enzyme. These data suggest that mis-folded species are formed from early folding intermediates. GroEL can capture early folding intermediates, and it loses the ability to capture and reactivate rhodanese if the enzyme is allowed first to spontaneously fold for longer times before it is presented to GroEL, a process that leads to the formation of unproductive intermediates. In addition, GroEL cannot reverse large aggregates once they are formed, but it could capture some folding intermediates and activate them, even though they are not capable of forming active enzyme if left to spontaneous refolding. The interaction between GroEL and rhodanese substantially but not completely inhibits intra-protein inactivation, which is responsible for incomplete activation during unassisted refolding. Thus, GroEL not only decreases aggregation, but it gives the highest reactivation of any method of assistance. The results are interpreted using a previously suggested model based on studies of the spontaneous folding of rhodanese (Gorovits, B. M., McGee, W. A., and Horowitz, P. M. (1998) Biochim. Biophys. Acta 1382, 120--128 and Panda, M., Gorovits, B. M., and Horowitz, P. M. (2000) J. Biol. Chem. 275, 63--70).  相似文献   

12.
In vitro refolding of the monomeric mitochondrial enzyme, rhodanese (thiosulfate sulfurtransferase; EC 2.8.1.1) is facilitated by molecular chaperonins. The four components: two proteins from Escherichia coli, chaperonin 60 (groEL) and chaperonin 10 (groES), MgATP, and K+, are necessary for the in vitro folding of rhodanese. These were previously shown to be necessary for the in vitro folding of ribulose-1,5-bisphosphate carboxylase at temperatures in excess of 25 degrees C (Viitanen, P. V., Lubben, T. H., Reed, J., Goloubinoff, P., O'Keefe, D. P., and Lorimer, G. H. (1990) Biochemistry 29, 5665-5671). The labile folding intermediate, rhodanese-I, which rapidly aggregates at 37 degrees C in the absence of the chaperonins, can be stabilized by forming a binary complex with chaperonin 60. The discharge of the binary chaperonin 60-rhodanese-I complex, results in the formation of active rhodanese, and requires the presence of chaperonin 10. Optimal refolding is associated with a K(+)-dependent hydrolysis of ATP. At lower protein concentrations and 25 degrees C, where aggregation is reduced, a fraction of the rhodanese refolds to an active form in the absence of the chaperonins. This spontaneous refolding can be arrested by chaperonin 60. There is some refolding (approximately equal to 20%) when ATP is replaced by nonhydrolyzable analogs, but there is no refolding in the presence of ADP or AMP. ATP analogs may interfere with the interaction of rhodanese-I with the chaperonins. Nondenaturing detergents facilitate rhodanese refolding by interacting with exposed hydrophobic surfaces of folding intermediates and thereby prevent aggregation (Tandon, S., and Horowitz, P. (1986) J. Biol. Chem. 261, 15615-15618). The chaperonin proteins appear to play a similar role in as much as they can replace the detergents. Consistent with this view, chaperonin 60, but not chaperonin 10, binds 2-3 molecules of the hydrophobic fluorescent reporter, 1,1'-bi(4-anilino)naphthalene-S,5'-disulfonic acid, indicating the presence of hydrophobic surfaces on chaperonin 60. The number of bound probe molecules is reduced to 1-2 molecules when chaperonin 10 and MgATP are added. The results support a model in which chaperonins facilitate folding, at least in part, by interacting with partly folded intermediates, thus preventing the interactions of hydrophobic surfaces that lead to aggregation.  相似文献   

13.
Creatine kinase (CK) is a dimeric enzyme important in ATP regeneration in cells where energy demands are high. The folding of CK under equilibrium and transient conditions has been studied in detail and is found to be complex. At equilibrium in 0.8 M GuHCl, 90% of CK molecules are in the form of a partially structured, monomeric intermediate. We exploit this property to measure kinetics of refolding and unfolding to and from this equilibrium intermediate (EI), using far-UV circular dichroism and intrinsic fluorescence as structural probes. We are thus able to compare the properties of EI and the kinetic intermediate formed during the burst phase in refolding. Native CK and EI unfold with rate constants in seconds and milliseconds, respectively. As is observed for refolding of fully-denatured CK, refolding from EI to the native state shows a burst phase followed by two exponential phases. The burst phase refolding intermediate is inferred to have more structure and greater stability than the equilibrium intermediate. When refolding from the fully-denatured state in 0.8 M GuHCl, the equilibrium intermediate is formed within the dead-time of mixing in the stopped-flow apparatus. The equilibrium intermediate may thus represent a kinetic intermediate formed early during folding.  相似文献   

14.
Proline isomerization is well known to cause additional slow phases during protein refolding. We address a new question: does the presence of prolines significantly affect the very fast kinetics that lead to the formation of folding intermediates? We examined both the very slow (10-100 min) and very fast (4 micro s-2.5 ms) folding kinetics of the two-domain enzyme yeast phosphoglycerate kinase by temperature-jump relaxation. Phosphoglycerate kinase contains a conserved cis-proline in position 204, in addition to several trans-prolines. Native cis-prolines have the largest effect on folding kinetics because the unfolded state favors trans isomerization, so we compared the kinetics of a P204H mutant with the wild-type as a proof of principle. The presence of Pro-204 causes an additional slow phase upon refolding from the cold denatured state, as reported in the literature. Contrary to this, the fast folding events are sped up in the presence of the cis-proline, probably by restriction of the conformational space accessible to the molecule. The wild-type and Pro204His mutant would be excellent models for off-lattice simulations probing the effects of conformational restriction on short timescales.  相似文献   

15.
Unfolded bovine rhodanese, a sulfurtransferase, does not regain full activity upon refolding due to the formation of aggregates and disulfide-linked misfolded states unless a large excess of reductant such as 200 mM -ME and 5 mg/ml detergent are present [Tandon and Horowitz (1990), J. Biol. Chem. 265, 5967]. Even then, refolding is incomplete. We have studied the unfolding and refolding of three rhodanese forms whose crystal structures are known: ES, containing the transferred sulfur as a persulfide; E, without the transferred sulfur, and carboxymethylated rhodanese (CMR), in which the active site was blocked by chemical modification. The X-ray structures of ES, E, and CMR are virtually the same, but their tertiary structures in solution differ somewhat as revealed by near-UV CD. Among these three, CMR is the only form of rhodanese that folds reversibly, requiring 1 mM DTT. A minimum three-state folding model of CMR (NIU) followed by fluorescence at 363 nm, (NI) by fluorescence at 318 nm, and CD (IU) is consistent with the presence of a thermodynamically stable molten globule intermediate in 5–6 M urea. We conclude that the active-site sulfhydryl group in the persulfide form is very reactive; therefore, its modification leads to the successful refolding of urea-denatured rhodanese even in the absence of a large excess of reductant and detergent. The requirement for DTT for complete reversibility of CMR suggests that oxidation among the three non-active-site SH groups can represent a minor trap for refolding through species that can be easily reduced.  相似文献   

16.
The refolding of mitochondrial aspartate aminotransferase (mAAT; EC 2.6.1.1) has been studied following unfolding in 6 m guanidine hydrochloride for different periods of time. Whereas reactivation of equilibrium-unfolded mAAT is sigmoidal, reactivation of the short term unfolded protein displays a double exponential behavior consistent with the presence of fast and slow refolding species. The amplitude of the fast phase decreases with increasing unfolding times (k approximately 0.75 min(-1) at 20 degrees C) and becomes undetectable at equilibrium unfolding. According to hydrogen exchange and stopped-flow intrinsic fluorescence data, unfolding of mAAT appears to be complete in less than 10 s, but hydrolysis of the Schiff base linking the coenzyme pyridoxal 5'-phosphate (PLP) to the polypeptide is much slower (k approximately 0.08 min(-1)). This implies the existence in short term unfolded samples of unfolded species with PLP still attached. However, since the disappearance of the fast refolding phase is about 10-fold faster than the release of PLP, the fast refolding phase does not correspond to folding of the coenzyme-containing molecules. The fast refolding phase disappears more rapidly in the pyridoxamine and apoenzyme forms of mAAT, both of which lack covalently attached cofactor. Thus, bound PLP increases the kinetic stability of the fast refolding unfolding intermediates. Conversion between fast and slow folding forms also takes place in an early folding intermediate. The presence of cyclophilin has no effect on the reactivation of either equilibrium or short term unfolded mAAT. These results suggest that proline isomerization may not be the only factor determining the slow refolding of this cofactor-dependent protein.  相似文献   

17.
Patra AK  Udgaonkar JB 《Biochemistry》2007,46(42):11727-11743
The mechanisms of folding and unfolding of the small plant protein monellin have been delineated in detail. For this study, a single-chain variant of the natively two-chain monellin, MNEI, was used, in which the C terminus of chain B was connected to the N terminus of chain A by a Gly-Phe linker. Equilibrium guanidine hydrochloride (GdnHCl)-induced unfolding experiments failed to detect any partially folded intermediate that is stable enough to be populated at equilibrium to a significant extent. Kinetic experiments in which the refolding of GdnHCl-unfolded protein was monitored by measurement of the change in the intrinsic tryptophan fluorescence of the protein indicated the accumulation of three transient partially structured folding intermediates. The fluorescence change occurred in three kinetic phases: very fast, fast, and slow. It appears that the fast and slow changes in fluorescence occur on competing folding pathways originating from one unfolded form and that the very fast change in fluorescence occurs on a third parallel pathway originating from a second unfolded form of the protein. Kinetic experiments in which the refolding of alkali-unfolded protein was monitored by the change in the fluorescence of the hydrophobic dye 8-anilino-1-naphthalenesulfonic acid (ANS), consequent to the dye binding to the refolding protein, as well as by the change in intrinsic tryptophan fluorescence, not only confirmed the presence of the three kinetic intermediates but also indicated the accumulation of one or more early intermediates at a few milliseconds of refolding. These experiments also exposed a very slow kinetic phase of refolding, which was silent to any change in the intrinsic tryptophan fluorescence of the protein. Hence, the spectroscopic studies indicated that refolding of single-chain monellin occurs in five distinct kinetic phases. Double-jump, interrupted-folding experiments, in which the accumulation of folding intermediates and native protein during the folding process could be determined quantitatively by an unfolding assay, indicated that the fast phase of fluorescence change corresponds to the accumulation of two intermediates of differing stabilities on competing folding pathways. They also indicated that the very slow kinetic phase of refolding, identified by ANS binding, corresponds to the formation of native protein. Kinetic experiments in which the unfolding of native protein in GdnHCl was monitored by the change in intrinsic tryptophan fluorescence indicated that this change occurs in two kinetic phases. Double-jump, interrupted-unfolding experiments, in which the accumulation of unfolding intermediates and native protein during the unfolding process could be determined quantitatively by a refolding assay, indicated that the fast unfolding phase corresponds to the formation of fully unfolded protein via one unfolding pathway and that the slow unfolding phase corresponds to a separate unfolding pathway populated by partially unfolded intermediates. It is shown that the unfolded form produced by the fast unfolding pathway is the one which gives rise to the very fast folding pathway and that the unfolded form produced by the slower unfolding pathway is the one which gives rise to the slow and fast folding pathways.  相似文献   

18.
For the first time, the enzyme rhodanese has been refolded after denaturation in guanidinium chloride (GdmHCl). Renaturation was by either (a) direct dilution into the assay, (b) intermediate dilution into buffer, or (c) dialysis followed by concentration and centrifugation. Method (c) preferentially retained active enzyme whose specific activity was 1140 IU/mg, which fell to 898 IU/mg after 6 days. The specific activity of native enzyme is 710 IU/mg. Progress curves were linear for the dialyzed enzyme, and kinetic analysis showed it had the same Km for thiosulfate as the native enzyme, but apparently displayed a higher turnover number. Progress curves for denatured enzyme directly diluted into assay mix showed as many as three phases: a lag during which no product formed; a first order reactivation; and an apparently linear steady state. An induction period was determined by extrapolating the steady-state line to the time axis. The percent reactivation fell to 7% (t1/2 = 10 min) as the time increased between GdmHCl dilution and the start of the assay, independent of the presence of thiosulfate. The induction period, which decreased to zero as the incubation time increased, was retained in the presence of thiosulfate. There were no observable differences between native and renatured protein by electrophoresis or fluorescence spectroscopy. Previous reports of some refolding of urea-denatured rhodanese (Stellwagen, E. (1979) J. Mol. Biol. 135, 217-229) were confirmed, extended, and compared with results using GdmHCl. A working hypothesis is that rhodanese refolding involves intermediates that partition into active and inactive products. These intermediates may result from nucleation of the two rhodanese domains, which exposes hydrophobic surfaces that become the interdomain interface in the correctly folded protein.  相似文献   

19.
The competition between protein aggregation and folding has been investigated using rhodanese (thiosulfate:cyanide sulfurtransferase, EC 2.8.1.1) as a model. During folding from a urea-denatured state, rhodanese rapidly forms associated species or intermediates, some of which are large and/or sticky. The early removal of such particles by filtration results in a decreased refolding yield. With time, a portion of the smaller aggregates can partition back first to intermediates and then to refolded protein, while a fraction of these irreversibly form unproductive higher aggregates. Dynamic light scattering measurements indicate that the average sizes of the aggregates formed during rhodanese folding increase from 225 to 325 nm over 45 min and they become increasingly heterogeneous. Glycerol addition or the application of high hydrostatic pressure improved the final refolding yields by stabilizing smaller particles. Although addition of glycerol into the refolding mixture blocks the formation of unproductive aggregates, it cannot dissociate them back to productive intermediates. The presence of 3.9 M urea keeps the aggregates small, and they can be dissociated to monomers by high hydrostatic pressure even after 1 h of incubation. These studies suggest that early associated intermediates formed during folding can be reversed to give active species.  相似文献   

20.
To gain insight into the molecular aspects of unfolding/refolding of enzymes from extremophilic organisms, we have used xylanase from an alkalophilic thermophilic Bacillus as the model system. Kinetics of denaturation/renaturation were monitored using intrinsic fluorescence studies. The protein fluorescence measurements suggested a putative intermediate state present in 0.08 M guanidine hydrochloride with an emission maximum of 345 nm; the far-UV circular dichroism spectra revealed content of secondary structure similar to the native enzyme. Studies with the fluorescent apolar probe 1-anilinonapthalene-8-sulfonate (1,8-ANS) were consistent with the presence of increased hydrophobic surfaces as compared with the native or fully unfolded protein. The refolding of Xyl II, was attempted by a relatively new strategy using an artificial chaperone assisted two-step method. The unfolded xylanase was found to bind to the detergent transiently and the subsequent addition of methyl-beta-cyclodextrin helped to strip the detergent and assist in the folding. Our findings suggested that the detergent stabilized a putative intermediate in the folding pathway seemingly equivalent to the folding state described as molten globule. The reactivation of Xyl II was affected by ionic as well as nonionic detergents. However, the cationic detergent cetyltrimethylammonium bromide (CTAB) provided a maximum reactivation (threefold) of the enzyme. The 'delayed detergent addition' experiments revealed that the detergent acts by suppressing the initial aggregate formation and not by dissolving aggregates. The relevance of our findings to the role of artificial chaperones in vivo is discussed.  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