首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Studies have been performed to quantitate the binding of the fluorescent probe 8-anilinonaphthalene-2-sulfonic acid (2,8-ANS) to catalytic intermediates of the enzyme rhodanese: the sulfur-substituted form (ES) and the sulfur-free form (E). The molecular 2,8-ANS has not been extensively used for protein studies, and some characterization is presented to demonstrate its usefulness as a probe for apolar binding sites. The molecule 2,8-ANS binds to at least two classes of sites on rhodanese. One class (class 1) is present in the ES form and has a Kd of 1.7 mM. The E form of rhodanese appears to have a second class of sites (class 2) in addition to the class 1 sites. Two independent fluorometric methods of analyzing the class 2 binding of 2,8-ANS to the E form gave an average value for Kd congruent to 179 microM. These fluorometric titrations, together with a Job plot, clearly indicate that 2,8-ANS binds to more than one site on the E form of rhodanese. The apparent apolarity is slightly higher for class 2 sites than for the class 1 sites, but both give Z factors of greater than 85. The substrate thiosulfate is able to displace the probe that is bound to the class 2 sites on the E form of the enzyme. Further, 2,8-ANS is found to be a competitive inhibitor of the catalyzed reaction with an apparent Kd of 170 microM. Circular dichroism measurements detect no significant changes in the average conformation of rhodanese that can be ascribed to the presence of 2,8-ANS.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The enzyme rhodanese in the form without transferred sulfur, (E), was inactivated by carboxymethylation with iodoacetic acid (E.IAA), and its conformation was compared with that of E inactivated by oxidative processes (Eox). Formation of E.IAA led to the exposure of binding sites for the fluorescent apolar probe 1,1'-bi(4-anilino)naphthalene-5,5'-disulfonic acid (BisANS). The dissociation constant for BisANS decreased as the concentration of E.IAA decreased and ranged from approximately 200 microM at 1 mg/ml protein to approximately 2 microM at protein concentrations below 0.1 mg/ml. Centrifugation confirmed that E.IAA, but not the underivatized enzyme, could associate. E.IAA was proteolyzable by chymotrypsin or endoproteinase Glu C (V8), while rhodanese containing bound sulfur, ES, was totally refractory, and E was only clipped to a small extent. This constellation of consequences was only previously observed with oxidatively inactivated rhodanese. Fluorescence depolarization measurements of bound BisANS were consistent with exposure of apolar surfaces and association of the protein. The fluorescence spectra of BisANS bound to E.IAA or Eox were identical and distinct from the spectrum of BisANS bound to phenylglyoxal-inactivated ES. Digestion with chymotrypsin was followed using protein and BisANS fluorescence and showed a similar response for E.IAA and Eox. These results indicate that the consequences of forming Eox and E.IAA are very similar. Thus, reaction of the active site sulfhydryl group apparently triggers a conformational change leading to increased protein flexibility and increased exposure of hydrophobic surfaces. In the case of oxidation, the trigger might involve initial formation of an active site sulfenic acid which ultimately gives higher oxidation states that could include disulfides.  相似文献   

3.
The intrinsic fluorescence of the enzyme rhodanese is quenched by as much as 30% when sulfur is transferred to the free enzyme form, E, giving the sulfur-substituted enzyme, ES. This fluorescence change (lambda ex = 295 nm and lambda em = 335 nm) has been used to quantitate the E and ES forms which are isolatable, obligatory intermediates in rhodanese catalysis. Fluorescence titration was performed using cyanide to irreversibly remove sulfur from ES. The results show a stoichiometry corresponding to 1 bound sulfur/molecule of the ES form of rhodanese (Mr = 33,000). The fluorescence changes were used to measure the concentrations of E and ES when these were in reversible equilibria induced by interactions with the substrates S2O3(2-) and SO3(2-). These results were compared with an equilibrium constant derived from published kinetic studies for the reaction (formula; see text) The very close agreement between the physical and kinetic methods indicate that there are no significant concentrations of intermediates other than E and ES. Overall, the results are compatible with the formation of a persulfide intermediate in rhodanese catalysis and are consistent with conclusions from x-ray crystallography and absorption spectroscopy. In addition, these procedures offer a facile method to measure equilibria between catalytic intermediates in the rhodanese reaction using functionally relevant concentrations.  相似文献   

4.
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   

5.
Rhodanese is oxidatively inactivated by several reagents, some of which are not normally considered oxidants. Rhodanese, in a form not containing persulfide sulfur (E), was inactivated by phenylglyoxal under conditions where disulfides are formed. There was the concomitant increase in the fluorescence of the apolar probe 1,1'-bi(4-anilino)naphthalene-5,5'-disulfonic acid (bisANS). At 0.2 mg/ml protein, there was no turbidity, while at 1 mg/ml, turbidity formed after an induction period of 23 min. Phenylglyoxal-inactivated E was extensively digested by endoproteinase glutamate C (V8 protease) to give two discrete high molecular weight fragments (Mr = 29,500 and 16,000). Enzymatically active E or ES, the form of rhodanese containing transferred sulfur (Mr = 33,000) was totally refractory to V8 protease and gave only small fluorescent enhancement of bisANS. Phenylglyoxal inactivated ES (reaction at arginine) gave very little fluorescence enhancement of bisANS and was not digested by V8. Hydrogen peroxide rapidly inactivated E (t1/2 less than 2 min) giving a slow increase in bisANS fluorescence (t1/2 greater than 10 min) identical to that observed with phenylglyoxal. The turbidity also increased after an induction period of approximately 30 min. Inactivation of E by hydrogen peroxide gave the same digestion pattern as that observed with phenylglyoxal inactivation. The turbidity was associated with the formation of disulfide-bonded structures that formed with the stoichiometry of E, 2E, 4E, 6E, 8E, etc. relative to the native enzyme, E. E was inactivated with several other reagents that lead to oxidatively inactivated rhodanese including NADH, dithiothreitol, mercaptoethanol, and m-dinitrobenzene. Enzyme inactivated with dithiothreitol or NADH gave an identical digestion pattern as above. In addition, with the exception of NADH which could not be used due to optical interference, each of the reagents gave rise to increased fluorescence of bisANS after inactivation. The results are consistent with a model in which the oxidized rhodanese resulting from diverse treatments is in a new conformation that has extensive exposed apolar surfaces and can form both noncovalent and disulfide-bonded aggregates.  相似文献   

6.
Summary The biosynthesis of rhodanese was studied in human hepatoma cell lines by immunoblotting and pulselabeling experiments using polyclonal antibodies raised against the bovine liver enzyme. Rhodanese, partially purified from human liver, showed an apparent molecular weight of 33,000 daltons, coincident with that of rhodanese from Hep 3B cells. After pulse labeling of Hep 3B cells both at 37°C and 25°C, rhodanese in the cytosol fraction exhibited the same molecular weight as the enzyme isolated from the particulate fraction containing mitochondria. Moreover, newly synthesized rhodanese from total Hep 3B RNA translation products showed the same electrophoretic mobility as rhodanese from Hep 3B cells. These results suggest that rhodanese, unlike most mitochondrial proteins, is not synthesized as a higher molecular weight precursor.  相似文献   

7.
The active-site sulfhydryl group in the enzyme thiosulfate sulfurtransferase (rhodanese; thiosulfate:cyanide sulfurtransferase; EC 2.8.1.1) is alkylated rapidly by iodoacetic acid in the free enzyme form, E, with complete loss of sulfurtransferase activity. Iodoacetic acid is completely ineffective with the sulfur-substituted form of the enzyme, ES. Iodoacetamide, on the other hand, has no effect on either enzyme form. The competitive enzyme inhibitor, toluenesulfonic acid, protects against inactivation in a strictly competitive way and analysis gives an apparent binding constant for toluenesulfonic acid of 12.5 mM, which is in agreement with studies of its effect on the catalyzed reaction. These results are taken to indicate that iodoacetic acid is an affinity analog for the substrate, thiosulfate, and inactivates because it can use the specific thiosulfate binding interactions, correctly orient its reactive center and displace intraprotein interactions which appear to protect the active-site sulfhydryl group in the E form.  相似文献   

8.
Sulfhydryl groups of bovine liver rhodanese (thiosulfate: cyanide sulfurtransferase, EC 2.8.1.1) were modified by treatment with tetrathionate. There was a linear relationship between loss of enzyme activity and the amount of tetrathionate used. At a ratio of one tetrathionate per mole of rhodanese, 100% of enzyme activity was lost in the sulfur-free E-form as compared with a 70% loss for the sulfur-containing ES-form of the enzyme. Addition of up to a 100-fold molar excess of tetrathionate to ES gave no further inactivation. Addition of cyanide to the maximally inactivated ES-tetrathionate complex gave complete loss of activity. Kinetic studies of maximally inactivated ES and partially inactivated E gave Km (Ks) values that were essentially the same as native enzyme, indicating that the active enzyme, in all cases, bound thiosulfate similarly. Reactivation was faster with the ES-form than with the E-form. The substrate, thiosulfate, could reactivate the enzyme up to 70% in 1 h with ES as compared to 24 h with E. Tetrathionate modification of rhodanese could be correlated with the changes in intrinsic fluorescence and with the binding of the active site reporter 2-anilinonaphthalene-8-sulfonic acid (2,8-ANS). Circular dichroism spectra of the protein suggested increased ordered secondary structure in the protein after reaction with tetrathionate. Cadmium chloride and phenylarsine oxide totally inactivated the enzyme at levels usually associated with their effect on enzymes containing vicinal sulfhydryl groups. Further, cadmium inhibition could be reversed by EDTA. Tetrathionate modification of rhodanese may proceed through the formation of sulfenylthiosulfate intermediates at sulfhydryl groups, close to but not identical with the active-site sulfhydryl group, which then can react further with the active-site sulfhydryl group to form disulfide bridges.  相似文献   

9.
Sulfurtransferase are enzymes involved in the formation, conversion and transport of compounds containing sulfane-sulfur atoms. Although the three-dimensional structure of the rhodanese from the nitrogen-fixing bacterium Azotobacter vinelandii is known, the role of its two domains in the protein conformational stability is still obscure. We have evaluated the susceptibility to proteolytic degradation of the two domains of the enzyme. The two domains show different resistance to the endoproteinases and, in particular, the N-terminal domain shows to be more stable to digestion during time than the C-terminal one. Cloning and overexpression of the N-terminal domain of the protein was performed to better understand its functional and structural role. The recombinant N-terminal domain of rhodanese A. vinelandii is soluble in water solution and the spectroscopic studies by circular dichroism and heteronuclear NMR spectroscopy indicate a stable fold of the protein with the expected alpha/beta topology. The results indicate that this N-terminal domain has already got all the elements necessary for an C-terminal domain independent folding. Its solution structure by NMR, actually under course, will be a valid contribution to understand the role of this domain in the folding process of the sulfurtransferase.  相似文献   

10.
Bovine liver rhodanese (thiosulfate:cyanide sulfurtransferase, EC 2.8.1.1) was prepared in dilute solutions and subjected to conditions that led to a time-dependent loss of enzyme activity. The rate of this activity loss was found to be dependent upon the sulfur substitution state of the enzyme, and the presence or absence of the substrates, thiosulfate and cyanide. In the absence of excess substrates, free enzyme (E), and the covalent intermediate form of the enzyme bearing a divalent sulfur atom in the active site (ES), are of approximately equal functional stability. In comparison, E, in the presence of excess cyanide, was markedly more labile, while ES, supported by 10-50 mM thiosulfate, showed no significant loss of activity under any of the conditions tested. All the enzyme solutions were shown to be losing assayable protein from solution. However, it was demonstrated that, for rhodanese in the E form, the amount of protein lost was insufficient to account for the activity lost, and a marked decline in specific activity was observed. Enzyme in the ES form, whether supported by additional thiosulfate or not, did not decline in the specific activity, though comparable protein loss did occur from these solutions. Intrinsic fluorescence measurements of rhodanese in the ES form, before and after removal of the persulfide sulfur through the addition of cyanide, indicated that loss of enzymic activity was not accompanied by loss of the bound sulfur atom. Therefore, the stabilizing effect observed with thiosulfate could not be explained simply by its ability to maintain enzyme in the sulfur-substituted state. Since the concentration of thiosulfate employed in these experiments was insufficient to maintain all the enzyme in ES.S2O3 form, thiosulfate was acting as a chemical reagent rather than a substrate in stabilizing enzyme activity.  相似文献   

11.
For the first time, the enzyme rhodanese has been proteolytically cleaved to give species that most likely correspond to individual domains. This indicates cleavage can occur in the interdomain tether. Further, the conditions for cleavage show that availability of the susceptible bond(s) depends on conformational changes triggered by oxidative inactivation. Rhodanese, without persulfide sulfur (E), was oxidized consequent to incubation with phenylglyoxal, NADH, or hydrogen peroxide. The oxidized enzyme (Eox) was probed using the proteolytic enzymes endoproteinase glutamate C (V8), trypsin, chymotrypsin, or subtilisin. The proteolytic susceptibility of Eox, formed using hydrogen peroxide, was compared with that of E and the form of the enzyme containing transferred sulfur, ES. ES was totally refractory to proteolysis, while E was only clipped to a small extent by trypsin or V8 and not at all by chymotrypsin or subtilisin. Eox was susceptible to proteolysis by all the proteases used, and, although there were some differences among the proteolytic patterns, there was always a band on sodium dodecyl sulfate-polyacrylamide gel electrophoresis corresponding to Mr = 16,500. This was the only band observed in addition to the parent species (Mr = 33,000) when Eox was digested with chymotrypsin, and conservation of total protein was observed after digestion up to 90 min. No additional species were observable on silver staining, although there was some indication that the band at 16,500 might be a doublet. The results are consistent with the occurrence of a conformational change after oxidation that results in increased exposure and/or flexibility of the interdomain tether which contains residues that meet the specificity requirements of the proteases used.  相似文献   

12.
The conformations of sulfur-free and sulfur-containing rhodanese were followed with and without the detergent lauryl maltoside after guanidinium chloride (GdmCl) addition to 5 M to study the apparent irreversibility of denaturation. Without lauryl maltoside, sulfur-containing rhodanese denatured in a transition giving, at approximately 2.3 M GdmCl, 50% of the total denaturation induced change observed by activity, CD, or intrinsic fluorescence. Sulfur-free rhodanese gave more complex behavior by intrinsic fluorescence and CD. CD showed loss of secondary structure in a broad, complex, and apparently biphasic transition extending from 0.5 to 3 M GdmCl. The interpretation of the transition was complicated by time-dependent aggregation due to noncovalent interactions. Results with the apolar fluorescence probe 2-anilinonaphthalene-8-sulfonic acid, implicated apolar exposure in aggregation. Sulfhydryl reactivity indicated that low GdmCl concentrations induced intermediates affecting the active site conformation. Lauryl maltoside prevented aggregation with no effect on activity or any conformational parameter of native enzyme. Transitions induced by GdmCl were still observed and consistent with several phases. Even in lauryl maltoside, an increase in apolar exposure was detected by 2-anilinonaphthalene-8-sulfonic acid, and by protein adsorption to octyl-Sepharose well below the major unfolding transitions. These results are interpreted with a model in which apolar interdomain interactions are disrupted, thereby increasing active site accessibility, before the intradomain interactions.  相似文献   

13.
The question of whether the solvent's water-immiscibility is relevant to enzymatic activity was addressed by assaying four different hydrolases (three lipases and one protease) in nine anhydrous solvents of similar hydrophobicities of which four were infinitely miscible with water and five were not. For no enzyme was a jump in activity observed upon a transition from water-miscible to water-immiscible solvent. The relevance of solvent apolarity to enzymatic efficiency was also examined. To this end, three groups of isomeric anhydrous solvents were selected where within each group of isomeric anhydrous solvents were selected where within each group one solvent was apolar (i.e., lacked a permanent dipole moment). For none of the four enzymes studied was activity significantly higher in apolar solvents than in their polar counterparts. Thus we conclude that often-cited solvent's immiscibility with water and apolarity by themselves are irrelevant to enzymatic activity. (c) 1993 John Wiley & Sons, Inc.  相似文献   

14.
GroEL C138W is a mutant form of Escherichia coli GroEL, which forms an arrested ternary complex composed of GroEL, the co-chaperonin GroES and the refolding protein molecule rhodanese at 25 degrees C. This state of arrest could be reversed with a simple increase in temperature. In this study, we found that GroEL C138W formed both stable trans- and cis-ternary complexes with a number of refolding proteins in addition to bovine rhodanese. These complexes could be reactivated by a temperature shift to obtain active refolded protein. The simultaneous binding of GroES and substrate to the cis ring suggested that an efficient transfer of substrate protein into the GroEL central cavity was assured by the binding of GroES prior to complete substrate release from the apical domain. Stopped-flow fluorescence spectroscopy of the mutant chaperonin revealed a temperature-dependent conformational change in GroEL C138W that acts as a trigger for complete protein release. The behavior of GroEL C138W was reflected closely in its in vivo characteristics, demonstrating the importance of this conformational change to the overall activity of GroEL.  相似文献   

15.
Rhodanese (thiosulfate cyanide sulfurtransferase; E.C. 2.8.1.1) is a mitochondrial enzyme that is unprocessed after import. We describein vitro experiments showing that partially folded rhodanese can interact with lipid bilayers. The interaction was monitored by measuring the ability of rhodanese to disrupt small unilamellar vesicles composed of phosphatidylserine and to release 6-carboxyfluorescein that was trapped in the liposomes. Partially folded rhodanese, derived by dilution of urea-unfolded enzyme, efficiently induced liposome leakage. Native rhodanese had no effect on liposome integrity. Liposome disruption progressively decreased as rhodanese was given the opportunity to refold or aggregate before introduction of the liposomes. A synthetic 23 amino acid peptide representing the N-terminal sequence of rhodanese was very efficient at disrupting the liposomes. Shorter peptides chosen from within this sequence (residues 11–23 or residues 1–17) had no effect on liposome disruption. A peptide representing the tether region that connects the domains of the enzyme was also without effect. These results are consistent with the hypothesis that the N-terminal sequence of rhodanese is an uncleaved leader sequence, and can interact with membrane components that are involved in the mitochondrial uptake of this protein.  相似文献   

16.
17.
In previous studies on the rhodanese activity of bovine liver mitochondria, we have shown that in addition to activity observed in the soluble protein fraction, there is rhodanese activity that is bound to the mitochondrial membrane. The latter activity accounts for as much as 40% of the total and, in situ, is associated in a multiprotein complex that forms iron-sulfur centers. In the present studies, we have investigated the rhodanese activity of bovine heart muscle. We have found that the major part of this enzyme activity is localized in the mitochondria and, further, that at least 25% of the total rhodanese activity of heart mitochondria is membrane-bound. As in liver tissue, the heart activity at least in part is associated in a multiprotein complex that forms iron-sulfur centers. Upon purification of the heart rhodanese in the soluble protein fraction, there is a 10- to 30-fold decrease inK m values for the standard assay substrates thiosulfate and cyanide ions. These observations are consistent with the interpretation that there are activated and deactivated (low activity) forms of the heart enzyme in crude extracts, but only the activated form survives purification. The present results, together with our recent finding that liver mitochondrial rhodanese is subject to phosphorylation, lend support to our proposal that the rhodaneses serve as converter enzymes which regulate the rate of electron transport through sulfuration of respiratory chain components. The rhodaneses, in turn, are controlled by protein kinases and the local ATP concentration.  相似文献   

18.
Differential chemical modification ofE. coli chaperonin 60 (cpn60) was achieved by using one of several sulfhydryl-directed reagents. For native cpn60, the three cysteines were accessible for reaction with N-ethylmaleimide (NEM), while only two of them are accessible to the larger reagent 4,4-dipyridyl disulfide (4-PDS). However, no sulfhydryl groups were modified when the even larger reagents 5,5-dithiobis-(2-nitrobenzoic acid) (DTNB) or 2-(4-(iodoacetamido)anilino) naphthalene-6-sulfonic acid (IAANS), were employed, unless the chaperonin was unfolded. The cpn60 that had been covalently modified with NEM or IAANS, was not able to support the chaperonin-assisted refolding of the mitochondrial enzyme rhodanese, which also requires cpn10 and ATP hydrolysis. However, both modified forms of cpn60 were able to form binary complexes with rhodanese, as demonstrated by their ability to arrest the spontaneous refolding of the enzyme. That is, chemical modification with these sulfhydryl-directed reagents produced a species that was not prevented from interaction with partially folded rhodanese, but that was prevented from supporting a subsequent step(s) during the chaperonin-assisted refolding process.  相似文献   

19.
Weikl TR  Boehr DD 《Proteins》2012,80(10):2369-2383
Protein function often involves changes between different conformations. Central questions are how these conformational changes are coupled to the binding or catalytic processes during which they occur, and how they affect the catalytic rates of enzymes. An important model system is the enzyme dihydrofolate reductase (DHFR) from Escherichia coli, which exhibits characteristic conformational changes of the active‐site loop during the catalytic step and during unbinding of the product. In this article, we present a general kinetic framework that can be used (1) to identify the ordering of events in the coupling of conformational changes, binding, and catalysis and (2) to determine the rates of the substeps of coupled processes from a combined analysis of nuclear magnetic resonance R2 relaxation dispersion experiments and traditional enzyme kinetics measurements. We apply this framework to E. coli DHFR and find that the conformational change during product unbinding follows a conformational‐selection mechanism, that is, the conformational change occurs predominantly prior to unbinding. The conformational change during the catalytic step, in contrast, is an induced change, that is, the change occurs after the chemical reaction. We propose that the reason for these conformational changes, which are absent in human and other vertebrate DHFRs, is robustness of the catalytic rate against large pH variations and changes to substrate/product concentrations in E. coli. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

20.
Ming-Yao Chew 《Phytochemistry》1973,12(10):2365-2367
Rhodanese activity was detected in crude leaf extracts of 12 randomly selected plant species consisting of 9 non-cyanophoric and 3 cyanophoric species. In each case, the enzyme exhibited high activity at pH 10·4 and 55°. There appeared to be no correlation between rhodanese activity and the cyanophoric nature of the plant.  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