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.
To study the distribution of rhodanese (E.C. 2.8.1.1) in plants, rhodanese activity was assayed on 13 cyanogenic and 12 non-cyanogenic species. All the species tested had the enzyme activity. This phenomenon leads to a hypothesis that the enzyme is generally distributed in plants.  相似文献   

2.
The three-dimensional structure of the rhodanese homology domain At4g01050(175-195) from Arabidopsis thaliana has been determined by solution nuclear magnetic resonance methods based on 3043 upper distance limits derived from NOE intensities measured in three-dimensional NOESY spectra. The structure shows a backbone root mean square deviation to the mean coordinates of 0.43 A for the structured residues 7-125. The fold consists of a central parallel beta-sheet with five strands in the order 1-5-4-2-3 and arranged in the conventional counterclockwise twist, and helices packing against each side of the beta-sheet. Comparison with the sequences of other proteins with a rhodanese homology domain in Arabidopsis thaliana indicated residues that could play an important role in the scaffold of the rhodanese homology domain. Finally, a three-dimensional structure comparison of the present noncatalytic rhodanese homology domain with the noncatalytic rhodanese domains of sulfurtransferases from other organisms discloses differences in the length and conformation of loops that could throw light on the role of the noncatalytic rhodanese domain in sulfurtransferases.  相似文献   

3.
The rate of rhodanese inactivation by 2,4,6-trinitrobenzenesulphonate is increased in the presence of diethylbarbiturate in the reaction medium. A "rate saturation effect" indicates the formation of a rhodanese-diethylbarbiturate complex, prior to modification-induced enzyme inactivation. The dissociation constant of this complex is 19.0 mM. Diethylbarbiturate has no effect on the trinitrophenylation rate of the free amino groups of rhodanese. When rhodanese modification, in the presence of diethylbarbiturate in the reaction medium, is carried out by the use of a 2,4,6-trinitrobenzenesulphonate concentration much lower than the concentration of rhodanese modifiable amino groups, reaction stoichiometry indicates that 3 to 5 moles of rhodanese are rendered inactive for each mole of 2,4,6-trinitrobenzenesulphonate utilized. This finding indicates the existence of a chain-reaction type mechanism of rhodanese inactivation.  相似文献   

4.
Summary The role of rhodanese in the detoxication of acute cyanide exposure is controversial. The debate involves questions of the availability of rhodanese to cyanide in the peripheral circulation. Blood-borne cyanide will distribute to the brain and may induce lesions or even death. The present study addresses the dispute by determining the distribution of rhodanese in tissues considered to have the highest rhodanese activity and thought to serve as major detoxication sites. The results indicate that rhodanese levels are highest in (1) hepatocytes that are in close proximity to the blood supply of the liver (2) epithelial cells surrounding the bronchioles (a major entry route for gaseous cyanide) and (3) proximal tubule cells of the kidney (serving to facilitate cyanide detoxication and elimination as thiocyanate). Rhodanese activity in the brain is low compared with liver and kidney (Mimoriet al., 1984; Drawbaugh & Marrs, 1987); the brain is not considered to be a major site of cyanide detoxication. The brain, however, is the target for cyanide toxicity. In this study our goal was also to differentiate the distribution of rhodanese in an area of the brain. We found that the enzyme level is highest in fibrous astrocytes of the white matter. Cyanide-induced brain lesions may thus occur in areas of the brain lacking sufficient sites for detoxication.  相似文献   

5.
The enzyme rhodanese is greatly stabilized in the range pH 4-6, and samples at pH 5 are fully active after several days at 23 degrees C. This is very different from results at pH greater than 7, where there is significant loss of activity within 1 h. A pH-dependent conformational change occurs below pH 4 in a transition centered around pH 3.25 that leads slowly to inactive rhodanese at pH 3 (t 1/2 = 22 min at pH3). The inactive rhodanese can be reactivated by incubation under conditions required for detergent-assisted refolding of denatured rhodanese. The inactive enzyme at pH 3 has the maximum of its intrinsic fluorescence spectrum shifted to 345 nm from 335 nm, which is characteristic of native rhodanese at pH greater than 4. At pH 3, rhodanese shows increased exposure of organized hydrophobic surfaces as measured by 1,1'-bis(4-anilino)naphthalene-5,5'-disulfonic acid binding. The secondary structure is maintained over the entire pH range studied (pH 2-7). Fluorescence anisotropy measurements of the intrinsic fluorescence provide evidence suggesting that the pH transition produces a state that does not display greatly increased average flexibility at tryptophan residues. Pepsin digestibility of rhodanese follows the pH dependence of conformational changes reported by activity and physical methods. Rhodanese is resistant to proteolysis above pH 4 but becomes increasingly susceptible as the pH is lowered. The form of the enzyme at pH 3 is cleaved at discrete sites to produce a few large fragments. It appears that pepsin initially cleaves close to one end of the protein and then clips at additional sites to produce species of a size expected for the individual domains into which rhodanese is folded. Overall, it appears that in the pH range between pH 3 and 4, titration of groups on rhodanese leads to opening of the structure to produce a conformation resembling, but more rigid than, the molten globule state that is observed as an intermediate during reversible unfolding of rhodanese.  相似文献   

6.
When the enzyme rhodanese was inactivated with hydrogen peroxide (H(2)O(2)), it underwent significant conformational changes, leading to an increased exposure of hydrophobic surfaces. Thus, this protein seemed to be an ideal substrate for GroEL, since GroEL uses hydrophobic interactions to bind to its substrate polypeptides. Here, we report on the facilitated reactivation (86%) of H(2)O(2)-inactivated rhodanese by GroEL alone. Reactivation by GroEL required a reductant and the enzyme substrate, but not GroES or ATP. Further, we found that GroEL interacted weakly and/or transiently with H(2)O(2)-inactivated rhodanese. A strong interaction with rhodanese was obtained when the enzyme was pre-incubated with urea, indicating that exposure of hydrophobic surfaces alone on oxidized rhodanese was not sufficient for the formation of a strong complex and that a more unfolded structure of rhodanese was required to interact strongly with GroEL. Unlike prior studies that involved denaturation of rhodanese through chemical or thermal means, we have clearly shown that GroEL can function as a molecular chaperone in the reactivation of an oxidatively inactivated protein. Additionally, the mechanism for the GroEL-facilitated reactivation of rhodanese shown here appears to be different than that for the chaperonin-assisted folding of chemically unfolded polypeptides in which a nucleotide and sometimes GroES is required.  相似文献   

7.
THiocystine (bis-[2-amino-2-carboxyethyl]trisulfide) is a natural substrate for rhodanese (thiosulfate:cyanide sulfurtransferase, EC 2.8.1.1). Analogs of thiocystine were prepared by eliminating the carboxyl or amino group or by lengthening the carbon chain. Of these only homothiocystine (bis-[2-amino-2-carboxypropyl]trisulfide) had appreciable activity as a substrate. At pH 8.6, the optimum for rhodanese, transfer of sulfane sulfur to cyanide in the presence of rhodanese was nonspecific. Only the sulfane sulfur of 35S-labeled thiocystine was transferred to rhodanese. Thus, thiocystine and thiosulfate both produce a rhodanese persulfide as a stable intermediate in sulfur transfer.  相似文献   

8.
Bovine mitochondrial rhodanese is a phosphoprotein   总被引:1,自引:0,他引:1  
The mitochondrial sulfurtransferase, rhodanese, has been analyzed for phosphate content. Significant amounts of protein-bound phosphate (30-40%) were measured in the six rhodanese preparations examined. Chromatographic experiments followed by phosphate analyses done on two of the preparations indicated that rhodanese A and rhodanese B, two enzyme forms that were previously resolved on DEAE-Sephadex by Blumenthal and Heinrikson (Blumenthal, K., and Heinrikson, R. L. (1971) J. Biol. Chem. 240, 2430-2437), correspond to dephospho- and phosphorhodanese, respectively. The phosphorylation of rhodanese by [gamma-32P]ATP is catalyzed by cAMP-dependent protein kinase. The stoichiometry of 32P incorporation based on the amount of dephosphorhodanese in the enzyme preparation approaches 1.0. The phosphorylation site is accessible in rhodanese that is free of substrate sulfur but not in the covalent enzyme-sulfur intermediate which is formed as an obligatory step during the course of catalysis. Because the cellular localization of cAMP-dependent protein kinase makes it unlikely as the physiologic modulator of rhodanese activity, liver extracts have been tested for a rhodanese kinase that does not require cAMP. Rhodanese kinase activity which is independent of cAMP is observed in extract fractions resolved by Affi-Gel Blue chromatography and freed from endogenous rhodanese by chromatography on Sephadex G-100. These results together with previous findings from this and other laboratories have led to a working model of a bicyclic cascade system that can modulate the rate of mitochondrial respiration. The essence of the model is a transduction and amplification of cellular signals into the altered covalent phosphorylation of rhodanese. Rhodanese, in turn, serves as a converter enzyme which directly alters the rate of the respiratory chain and, thus, ATP production by the reversible sulfuration of key iron-sulfur centers. The model, when expanded to include signal pathways initiated by hormones or neurotransmitters, represents a mechanism by which mitochondria can recognize and meet changing energy demands.  相似文献   

9.
For the first time, the enzyme rhodanese had been refolded after thermal denaturation. This was previously not possible because of the strong tendency for the soluble enzyme to aggregate at temperatures above 37 degrees C. The present work used rhodanese that was covalently coupled to a solid support under conditions that were found to preserve enzyme activity. Rhodanese was immobilized using an N-hydroxymalonimidyl derivative of Sepharose containing a 6-carbon spacer. The number of immobilized competent active sites was measured by using [35S]SO3(2-) to form an active site persulfide that is the obligatory catalytic intermediate. Soluble enzyme was irreversibly inactivated in 10 min at 52 degrees C. The immobilized enzyme regained at least 30% of its original activity even after boiling for 20 min. The immobilized enzyme had a Km and Vmax that were each approximately 3 times higher than the corresponding values for the native enzyme. After preincubation at high temperatures, progress curves for the immobilized enzyme showed induction periods of up to 5 min before attaining apparently linear steady states. The pH dependence of the activity was the same for both the soluble and the immobilized enzyme. These results indicate significant stabilization of rhodanese after immobilization, and instabilities caused by adventitious solution components are not the sole reasons for irreversibility of thermal denaturation seen with the soluble enzyme. The results are consistent with models for rhodanese that invoke protein association as a major cause of inactivation of the enzyme. Furthermore, the induction period in the progress curves is consistent with studies which show that rhodanese refolding proceeds through intermediate states.  相似文献   

10.
Reactivation of denatured rhodanese (thiosulfate:cyanide sulfurtransferase, EC 2.8.1.1) was found to be aided by the presence of serum albumin. Both the rate and the extent of reactivation of the urea-denatured enzyme were optimal at low rhodanese and moderate serum albumin concentrations. Similarly, stabilization of the sulfurtransferase activity of rhodanese that had been partially unfolded at 40°C was aided by the presence of serum albumin. All the observations are in accord with a model in which enzyme that has been partially refolded from the urea-denatured state or partially unfolded thermally interacts directly with serum albumin in a way that prevents rhodanese self-association. Serum albumin thus acts as a molecular chaperone in these systems.  相似文献   

11.
1. The activity of rhodanese in different tissues of some domestic animals was measured. 2. Rhodanese was present in all tissues studied. 3. The activity of rhodanese in most tissues of sheep was higher than other animals studied. 4. In sheep and cattle the epithelium of rumen, omasum and reticulum were the richest sources of rhodanese. Significant activity of rhodanese was also present in liver and kidney. 5. In camel the liver contained the highest level of rhodanese followed by lung and rumen epithelium. Camel liver contained a third of the activity of sheep liver. 6. Equine liver had a third of the activity of sheep liver. Other tissues showed low levels of rhodanese activity. 7. Dog liver contained only 4% of the activity of sheep liver. In this animal, brain was the richest source of rhodanese. 8. The results are discussed in terms of efficacy of different tissues of animals in cyanide detoxification.  相似文献   

12.
Rhodanese has been utilized as a model enzyme for the study of protein structure-function relationships. The enzyme has recently been cloned and the recombinant enzyme is now available for investigation. However, prior to use in structure-function studies, the recombinant enzyme must be shown to have the same structure and activity as the bovine liver enzyme used in the previous studies. An immunological study of the conformations of these enzyme conformers is described. Three antibodies (two monoclonal and one polyclonal, site-directed antibody) were shown to detect distinct and nonoverlapping epitopes. The epitopes of the monoclonal antirhodanese antibodies (R207 and MAB11) were mapped to the same CNBr digest fragment of the amino terminal domain of rhodanese, and the epitope of the site-directed antibody prepared against the interdomain tether sequence of rhodanese (PAT-T1) was mapped to that region of rhodanese (residues 142–156). The rhodanese conformers were studied by monitoring the accessibility of the epitopes recognized by each antibody in each conformer using an indirect ELISA. None of the antibodies could detect its epitope on the purified liver enzyme. Two of the antibodies (R207 and PAT-T1) could also not detect their epitopes on the recombinant enzyme. However, MAB11 did detect a conformational difference between the natural and recombinant rhodanese conformers, indicating the conformational difference is localized in the first 73 amino acids of rhodanese. This difference presumably reflects the difference in the histories of the two enzymes and may be due to differences in enzyme folding, differences in the purification procedures, and differences in storage conditions—all of which could influence the final conformation of the enzyme.  相似文献   

13.
Mitochondrial rhodanese: membrane-bound and complexed activity   总被引:3,自引:0,他引:3  
We have proposed that phosphorylated and dephosphorylated forms of the mitochondrial sulfurtransferase, rhodanese, function as converter enzymes that interact with membrane-bound iron-sulfur centers of the electron transport chain to modulate the rate of mitochondrial respiration (Ogata, K., Dai, X., and Volini, M. (1989) J. Biol. Chem. 204, 2718-2725). In the present studies, we have explored some structural aspects of the mitochondrial rhodanese system. By sequential extraction of lysed mitochondria with phosphate buffer and phosphate buffer containing 20 mM cholate, we have shown that 30% of the rhodanese activity of bovine liver is membrane-bound. Resolution of cholate extracts on Sephadex G-100 indicates that part of the bound rhodanese is complexed with other mitochondrial proteins. Tests with the complex show that it forms iron-sulfur centers when incubated with the rhodanese sulfur-donor substrate thiosulfate, iron ions, and a reducing agent. Experiments on the rhodanese activity of rat liver mitochondria give similar results. Taken together, the findings indicate that liver rhodanese is in part bound to the mitochondrial membrane as a component of a multiprotein complex that forms iron-sulfur centers. The findings are consistent with the role we propose for rhodanese in the modulation of mitochondrial respiratory activity.  相似文献   

14.
1. Rhodanese has been extracted from Thiobacillus denitrificans by ultrasonic disintegration of the cells. 2. Studies with Sephadex columns have shown that the enzyme aggregates, forming a tetramer. 3. The molecular weights of the monomer and of an enzymically active sub-unit one-quarter this size have been determined by gel filtration. 4. Higher-molecular-weight forms of rhodanese are broken down by mercaptoethanol to enzymically active fragments of mol.wt. 7000 and 2000 respectively. 5. It is suggested that these fragments are linked in vivo via disulphide bridges to form the monomer, which can then aggregate via further disulphide links. 6. The fragment of mol.wt. 7000 has been obtained in a substantially pure state. 7. Both disulphide and thiol groups are necessary for enzyme activity. 8. Similarities and differences existing between bacterial rhodanese, mammalian rhodanese and beta-mercaptopyruvate sulphurtransferase are discussed.  相似文献   

15.
We report a purification of bovine liver rhodanese (thiosulfate:cyanide sulfurtransferase, EC 2.8.1.1) using column chromatography under conditions that take advantage of recent information regarding the structure and stability of this enzyme. At low pH (e.g., pH 4-6), rhodanese is stabilized against inactivation processes. By maintaining rhodanese at low pH, column chromatography, and especially ion-exchange chromatography, becomes practical, without loss of enzymatic activity. A purification method involving the sequential use of cation-exchange, size-exclusion, and hydrophobic-interaction chromatography was developed, and rhodanese was purified with good yield to electrophoretic purity and high specific activity. Previous methods for purifying bovine liver rhodanese employ repeated ammonium sulfate fractionations and crystallization of the rhodanese. In these methods, it is difficult to separate rhodanese from yellow-brown contaminants in the final stages of the procedures. Here, yellow-brown contaminants, which copurify with rhodanese on the first two columns, are completely resolved by hydrophobic interaction chromatography. This method can be readily scaled up, requires no special equipment, eliminates the variability inherent in previous methods, and is less dependent upon experience.  相似文献   

16.
The enzyme rhodanese was investigated for the conformational transition associated with its urea unfolding. When rhodanese was treated with 0 or 3 M urea, the activity was not significantly affected. 4.25 M urea treatment led to a time-dependent loss of activity in 60 min. Rhodanese was completely inactivated within 2 min in 6 M urea. The 1,1'-bi(4-anilino)naphthalene-5,5'-disulfonic acid fluorescence intensity was not significantly increased during 0, 3, and 6 M urea equilibrations, and the fluorescence was dramatically increased with 4.25 M urea, indicating that hydrophobic surfaces are exposed. After 0 and 3 M urea equilibration, rhodanese was not significantly proteolyzed with trypsin. Treatment with 4.25 M urea led to simultaneous formation of major 12-, 15.9-, 17-, and 21.2-kDa fragments, followed by progressive emergence of smaller peptides. The N termini of the 17- and 21.2-kDa bands were those of intact rhodanese. The N terminus of the 15.9-kDa band starts at the end of the interdomain tether. The 12-kDa band begins with either residue 183 or residue 187. The size and sequence information suggest that the 17- and 15.9-kDa bands correspond to the two domains. The 21.2- and 12-kDa bands appear to be generated through one-site tryptic cleavage. It is concluded that urea disrupts interaction between the two domains, increasing the accessibility of the interdomain tether that can be digested by trypsin. The released domains have increased proteolytic susceptibility and produce smaller peptides, which may represent subdomains of rhodanese.  相似文献   

17.
Considerably larger quantities of cyanide are required to solubilize gold following the bio-oxidation of gold-bearing ores compared with oxidation by physical-chemical processes. A possible cause of this excessive cyanide consumption is the presence of the enzyme rhodanese. Rhodanese activities were determined for the bacteria most commonly encountered in bio-oxidation tanks. Activities of between 6.4 and 8.2 micromol SCN min(-1) mg protein(-1) were obtained for crude enzyme extracts of Thiobacillus ferrooxidans, Thiobacillus thiooxidans and Thiobacillus caldus, but no rhodanese activity was detected in Leptospirillum ferrooxidans. Rhodanese activities 2-2.5-fold higher were found in the total mixed cell mass from a bio-oxidation plant. T. ferrooxidans synthesized rhodanese irrespective of whether it was grown on iron or sulphur. With a PCR-based detection technique, only L. ferrooxidans and T. caldus cells were detected in the bio-oxidation tanks. As no rhodanese activity was associated with L. ferrooxidans, it was concluded that T. caldus was responsible for all of the rhodanese activity. Production of rhodanese by T. caldus in batch culture was growth phase-dependent and highest during early stationary phase. Although the sulphur-oxidizing bacteria were clearly able to convert cyanide to thiocyanate, it is unlikely that this rhodanese activity is responsible for the excessive cyanide wastage at the high pH values associated with the gold solubilization process.  相似文献   

18.
Rhodanese has been extensively utilized as a model protein for the study of enzyme structure-function relationships. An immunological study of conformational changes occurring in rhodanese as a result of oxidation or thermal inactivation was conducted. Three monoclonal antibodies (MABs) to rhodanese were produced. Each MAB recognized a unique epitope as demonstrated by binding of the MABs to different proteolytic fragments of rhodanese. While none of the MABs significantly bound native, soluble, sulfur-substituted bovine rhodanese, as indicated in indirect enzyme-linked immunosorbent assay experiments, each MAB was immunoadsorbed from solution by soluble rhodanese as a function of the time rhodanese was incubated at 37 degrees C. Thus, as rhodanese was thermally inactivated, conformational changes resulted in the expression of three new epitopes. Catalytic conformers demonstrated different rates of thermally induced antigen expression. Each MAB also recognized epitopes expressed when rhodanese was immobilized on microtiter plates at 37 degrees C. Two conformers resulting from oxidation of rhodanese by hydrogen peroxide were identified immunologically. All MABs recognized rhodanese that was oxidized with peroxide and allowed to undergo a secondary cyanide-dependent reaction which also resulted in a fluorescence shift and increased proteolytic susceptibility. Only one MAB was capable of recognizing an epitope expressed when rhodanese was oxidized with peroxide and then separated from the reactants to prevent induction of the secondary conformational changes.  相似文献   

19.
The ability of the cytoplasmically synthesized mitochondrial enzyme rhodanese and its putative import signal sequence to interact with model phospholipid membranes was characterized. Membrane perturbation assays were used to test a current hypothesis that the initial step in protein translocation may involve binding of signal sequences with membrane lipids. Here we show comparative studies on the effect of native and various forms of denatured rhodanese, as well as two peptides, rho(1-23) and rho(11-23), derived from its NH2-terminal sequence, on the perturbation of 6-carboxyfluorescein-containing large unilamellar vesicles composed of either cardiolipin, phosphatidylcholine, or phosphatidylserine. We monitored the degree of perturbation by measuring dye leakage and found differential perturbation by either peptide or protein. Unfolded rhodanese perturbed vesicles in the order phosphatidylserine > cardiolipin > phosphatidylcholine. Denatured rhodanese was approximately 25 times more effective (on a molar basis) than rho(1-23) in the disruption of anionic liposomes. Rho(11-23) was unable to perturb liposomes. We found an inverse correlation between degree of activity of rhodanese folding intermediates and their ability to perturb liposomes. On urea denaturation, enzymatic activity was completely lost before membrane perturbation ability reached significant levels. Analysis of the peptides by circular dichroism showed that anionic liposomes can induce alpha-helical structure only in rho(1-23) and denatured rhodanese. Intrinsic peptide fluorescence studies showed that only rho(1-23) and denatured rhodanese partitioned into these model membranes. Results obtained here imply that peptides from naturally occurring alpha-helical structures may need adjacent motifs for helical structure induction in lipid environments, and the subsequent secondary structure may, in turn, promote partitioning of these segments into the lipid phase and ultimately lead to membrane perturbation.  相似文献   

20.
The role of rhodanese as an intermediate catalyst in the oxidation of elemental S (S°) is not well understood. This study investigated the effect of 26 soil properties and steam sterilization in relation to S° oxidation and rhodanese activity in 33 soils (27 Oregon soils and six Chinese soils). S° oxidation potential was determined by incubating (7 d at 23 °C) soil amended with 500 mg S° kg-1 soil and measuring the SO4 released. Both total S° oxidation (TSO) and rhodanese activity varied widely among the 33 soils, ranging from 0 to 143 mg SO4-S kg-1 soil 7 d-1 and 22 to 2109 nmoles SCN- g-1 soil h-1 respectively. S° oxidation but not rhodanese activity had a significant positive correlation with soil pH. In sterile soils, chemical S° oxidation (CSO) averaged 3% of the total S° oxidation and apparent rhodanese activity averaged 11% of the total rhodanese activity. S° oxidation was not significantly correlated with rhodanese activity. However, development of stepwise regression models predicting S° oxidation revealed that rhodanese activity was an important explanatory variable in predicting biological S° oxidation (TSO minus CSO). Also, microbial biomass C was found to be an important parameter in models for both S° oxidation and rhodanese activity. Investigations of the effect of acidification during S° oxidation showed that biological S° oxidation was negatively correlated with S° oxidation-induced-pH-change for soils with pH > 6 but no such significant relationship was found on soils with pH> 6. This suggested that extreme acidity may inhibit S° oxidation but not rhodanese activity.  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