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.
Eukaryotic sulfite oxidase is a dimeric protein that contains the molybdenum cofactor and catalyzes the metabolically essential conversion of sulfite to sulfate as the terminal step in the metabolism of cysteine and methionine. Nitrate reductase is an evolutionarily related molybdoprotein in lower organisms that is essential for growth on nitrate. In this study, we describe human and chicken sulfite oxidase variants in which the active site has been modified to alter substrate specificity and activity from sulfite oxidation to nitrate reduction. On the basis of sequence alignments and the known crystal structure of chicken sulfite oxidase, two residues are conserved in nitrate reductases that align with residues in the active site of sulfite oxidase. On the basis of the crystal structure of yeast nitrate reductase, both positions were mutated in human sulfite oxidase and chicken sulfite oxidase. The resulting double-mutant variants demonstrated a marked decrease in sulfite oxidase activity but gained nitrate reductase activity. An additional methionine residue in the active site was proposed to be important in nitrate catalysis, and therefore, the triple variant was also produced. The nitrate reducing ability of the human sulfite oxidase triple mutant was nearly 3-fold greater than that of the double mutant. To obtain detailed structural data for the active site of these variants, we introduced the analogous mutations into chicken sulfite oxidase to perform crystallographic analysis. The crystal structures of the Mo domains of the double and triple mutants were determined to 2.4 and 2.1 ? resolution, respectively.  相似文献   

2.
Little is known about the homeostasis of sulfite levels, a cytotoxic by-product of plant sulfur turnover. By employing extended dark to induce catabolic pathways, we followed key elements of the sulfite network enzymes that include adenosine-5′-phosphosulfate reductase and the sulfite scavengers sulfite oxidase (SO), sulfite reductase, UDP-sulfoquinovose synthase, and β-mercaptopyruvate sulfurtransferases. During extended dark, SO was enhanced in tomato (Solanum lycopersicum) wild-type leaves, while the other sulfite network components were down-regulated. SO RNA interference plants lacking SO activity accumulated sulfite, resulting in leaf damage and mortality. Exogenous sulfite application induced up-regulation of the sulfite scavenger activities in dark-stressed or unstressed wild-type plants, while expression of the sulfite producer, adenosine-5′-phosphosulfate reductase, was down-regulated. Unstressed or dark-stressed wild-type plants were resistant to sulfite applications, but SO RNA interference plants showed sensitivity and overaccumulation of sulfite. Hence, under extended dark stress, SO activity is necessary to cope with rising endogenous sulfite levels. However, under nonstressed conditions, the sulfite network can control sulfite levels in the absence of SO activity. The novel evidence provided by the synchronous dark-induced turnover of sulfur-containing compounds, augmented by exogenous sulfite applications, underlines the role of SO and other sulfite network components in maintaining sulfite homeostasis, where sulfite appears to act as an orchestrating signal molecule.As a plant macronutrient, sulfur (S) is important for yield production and the quality of crops. In nature, S is mostly available in its fully oxidized anion sulfate form, which is taken up, reduced, and incorporated into Cys via sulfite and sulfide generation in the sulfate assimilation pathway (Saito, 2000). In addition to the presence of S in the amino acids Cys and Met, it is an important component of oligopeptides such as reduced glutathione (GSH), coenzymes, prosthetic groups, vitamins, secondary metabolites, and lipids (Saito, 2000; Leustek, 2002). In the latter case, the chloroplast membrane containing sulfoquinovosyldiacylglycerol (SQDG) is one of the primary S-containing components in higher plants (Shimojima, 2011).Sulfite, a less oxidized form of sulfate, is an intermediate in the assimilation of S and a potentially cytotoxic molecule (Leustek et al., 2000) that, if not rapidly metabolized, can wreak havoc at the cellular (Sanda et al., 2001; Davidian and Kopriva, 2010) and whole-plant (Murray, 1997; Brychkova et al., 2007) levels. Roots obtain sulfate from the soil, and sulfite is generated from sulfate in the leaves by the chloroplast-localized adenosine-5′-phosphosulfate (APS) reductase (APR; Enzyme Commission [EC] 1.8.4.9; Vauclare et al., 2002). Another source of sulfite is atmospheric, originating from microbial, volcanic, or anthropogenic activities and entering the plant via the stomata or through the root system. In mammalian tissue, endogenous sulfite is thought to be formed during the degradation of the S-containing amino acids (Amy, 1988; Heber and Huve, 1998; Hänsch and Mendel, 2005); however, such a pathway has yet to be explored in plants.Known avenues for sulfite usage include sulfite assimilation, incorporation into metabolites, and detoxification, which together make up a potential network for the control of sulfite turnover (Fig. 1). These include ferredoxin-dependent sulfite reduction catalyzed by the chloroplast-localized sulfite reductase (SiR; EC 1.8.7.1), which yields the reduced sulfide for primary sulfate assimilation (Khan et al., 2010). Another pathway for sulfite utilization is its incorporation into sulfolipids, catalyzed by the chloroplast-localized UDP-sulfoquinovose synthase (SQD1; EC 3.13.1.1; Sanda et al., 2001). Sulfite can be reoxidized back to sulfate by the molybdenum cofactor-containing enzyme, sulfite oxidase (SO; EC 1.8.3.1), localized in the peroxisomes (Eilers et al., 2001). Alternatively, the mitochondrion- and cytosol-localized β-mercaptopyruvate sulfurtransferases (STs; EC 2.8.1.2.), MST1 and MST2, respectively, have been shown to catalyze the synthesis of the less toxic compound thiosulfate in the presence of β-mercaptopyruvate and sulfite (Papenbrock and Schmidt, 2000a; Tsakraklides et al., 2002).Open in a separate windowFigure 1.Schematic representation of sulfite network enzymes in plants. Sulfite is generated from sulfate in two consecutive steps. In the first step, ATP sufurylase catalyzes the adenylation of sulfate to APS, and then sulfite is produced in the chloroplast by the glutathione-dependent APR. In the chloroplast, the generated sulfite can be further reduced to sulfide by the ferredoxin-dependent SiR. The sulfide together with O-acetyl-l-Ser are the substrates for Cys biosynthesis catalyzed by OAS-TL. Alternatively, the chloroplast-localized sulfite can enter the sulfolipid reductive pathway to generate SQDG in two consecutive steps. In the first step, UDP-sulfoquinovase is catalyzed by SQD1, employing sulfite and UDP-Glc as substrates, while in the second step, SQDG is catalyzed by SQDG synthase, employing UDP-sulfoquinovose and diacylglycerols as substrates. Sulfite localized to the cytosol and mitochondria may be detoxified to thiosulfate by the STs or be generated by these enzymes from thiosulfate and cyanide (Nakamura et al., 2000; Papenbrock and Schmidt, 2000a; Tsakraklides et al., 2002). Sulfite can be oxidized to sulfate by the molybdenum cofactor-containing enzyme, peroxisomal SO. Fdox, Oxidized ferredoxin; Fdred, reduced ferredoxin; GSSG, oxidized glutathione; H2O2, hydrogen peroxide; PPi, diphosphate.Since sulfite is cytotoxic (Leustek et al., 2000), it can be expected that its cellular levels are tightly regulated through an interplay between its production and conversion. However, virtually nothing is known about the factors regulating its homeostasis in plants and the interaction of enzymes involved in internal sulfite detoxification. In this work, extended dark stress was employed as a means to induce catabolic pathways (Keskitalo et al., 2005; Brychkova et al., 2007) that stimulate the turnover of S-containing metabolites and result in sulfite flux. We analyzed known components of the sulfite homeostasis network in tomato (Solanum lycopersicum) plants and demonstrated the essential role of SO. During dark stress, SO expression was enhanced while the expression of the other known sulfite network components, APRs, SiR, SQD1, and STs, was inhibited. In the absence of the SO activity, toxic sulfite levels accumulated in the dark and was accompanied by increased leaf damage and plant mortality. Direct sulfite application stimulated components of the sulfite network, indicating that sulfite might play an important role as a signal molecule in orchestrating sulfite homeostasis.  相似文献   

3.
Resting-cell suspensions of Desulfomonile tiedjei consumed H2 with 3-chloro-, 3-bromo-, and 3-iodobenzoate as electron acceptors with rates of 0.50, 0.44, and 0.04 mumol h-1 mg-1, respectively. However, benzoate and 3-fluorobenzoate were not metabolized by this bacterium. In addition, H2 uptake was at least fourfold faster when sulfate, sulfite, or thiosulfate was available as the electron acceptor instead of a haloaromatic substrate. When sulfite and 3-chlorobenzoate were both available for this purpose, the rate of H2 uptake by D. tiedjei was intermediate between that obtained with either electron acceptor alone. Hydrogen concentrations were reduced to comparably low levels when either 3-chlorobenzoate, sulfate, or sulfite was available as an electron acceptor, but significantly less H2 depletion was evident with benzoate or nitrate. Rates of 3-chlorobenzoate dechlorination increased from an endogenous rate of 14.5 to 17.1, 74.0, 81.1, and 82.3 nmol h-1 mg-1 with acetate, pyruvate, H2, and formate, respectively, as the electron donors. Sulfite and thiosulfate inhibited dehalogenation, but sulfate and NaCl had no effect. Dehalogenation and H2 metabolism were also inhibited by acetylene, molybdate, selenate, and metronidazole. Sulfite reduction and dehalogenation were inhibited by the same respiratory inhibitors. These results suggest that the reduction of sulfite and dehalogenation may share part of the same electron transport chain. The kinetics of H2 consumption and the direct inhibition of dehalogenation by sulfite and thiosulfate in D. tiedjei cells clearly indicate that the reduction of sulfur oxyanions is favored over aryl dehalogenation for the removal of reducing equivalents under anaerobic conditions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Resting-cell suspensions of Desulfomonile tiedjei consumed H2 with 3-chloro-, 3-bromo-, and 3-iodobenzoate as electron acceptors with rates of 0.50, 0.44, and 0.04 mumol h-1 mg-1, respectively. However, benzoate and 3-fluorobenzoate were not metabolized by this bacterium. In addition, H2 uptake was at least fourfold faster when sulfate, sulfite, or thiosulfate was available as the electron acceptor instead of a haloaromatic substrate. When sulfite and 3-chlorobenzoate were both available for this purpose, the rate of H2 uptake by D. tiedjei was intermediate between that obtained with either electron acceptor alone. Hydrogen concentrations were reduced to comparably low levels when either 3-chlorobenzoate, sulfate, or sulfite was available as an electron acceptor, but significantly less H2 depletion was evident with benzoate or nitrate. Rates of 3-chlorobenzoate dechlorination increased from an endogenous rate of 14.5 to 17.1, 74.0, 81.1, and 82.3 nmol h-1 mg-1 with acetate, pyruvate, H2, and formate, respectively, as the electron donors. Sulfite and thiosulfate inhibited dehalogenation, but sulfate and NaCl had no effect. Dehalogenation and H2 metabolism were also inhibited by acetylene, molybdate, selenate, and metronidazole. Sulfite reduction and dehalogenation were inhibited by the same respiratory inhibitors. These results suggest that the reduction of sulfite and dehalogenation may share part of the same electron transport chain. The kinetics of H2 consumption and the direct inhibition of dehalogenation by sulfite and thiosulfate in D. tiedjei cells clearly indicate that the reduction of sulfur oxyanions is favored over aryl dehalogenation for the removal of reducing equivalents under anaerobic conditions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
The Mo(V) center of plant sulfite oxidase from Arabidopsis thaliana (At-SO) has been studied by continuous wave and pulsed EPR methods. Three different Mo(V) EPR signals have been observed, depending on pH and the technique used to generate the Mo(V) oxidation state. At pH 6, reduction by sulfite followed by partial reoxidation with ferricyanide generates an EPR spectrum with g-values similar to the low-pH (lpH) form of vertebrate SOs, but no nearby exchangeable protons can be detected. On the other hand, reduction of At-SO with Ti(III) citrate at pH 6 generates a Mo(V) signal with large hyperfine splittings from a single exchangeable proton, as is typically observed for lpH SO from vertebrates. Reduction of At-SO with sulfite at high pH generates the well-known high-pH (hpH) signal common to all sulfite oxidizing enzymes. It is proposed that, depending on the conformation of Arg374, the active site of At-SO may be in "closed" or "open" forms that differ in the degree of accessibility of the Mo center to substrate and water molecules. It is suggested that at low pH the sulfite-reduced At-SO has coordinated sulfate and is in the "closed form". Reoxidation to Mo(V) by ferricyanide leaves bound sulfate trapped at the active site, and consequently, there are no ligands with exchangeable protons. Reduction with Ti(III) citrate injects an electron directly into the active site to generate the [Mo(V)[triple bond]O(OH)]2+ unit that is well-known from model chemistry and which has a single exchangeable proton with a large isotropic hyperfine interaction. At high pH, the active site is in the "open form", and water can readily exchange into the site to generate the hpH SO.  相似文献   

6.
7.
In the crystal structure of chicken sulfite oxidase, the residue Tyr(322) (Tyr(343) in human sulfite oxidase) was found to directly interact with a bound sulfate molecule and was proposed to have an important role in mediating the substrate specificity and catalytic activity of this molybdoprotein. In order to understand the role of this residue in the catalytic mechanism of sulfite oxidase, steady-state and stopped-flow analyses were performed on wild-type and Y343F human sulfite oxidase over the pH range 6-10. In steady-state assays of Y343F sulfite oxidase using cytochrome c as the electron acceptor, k(cat) was somewhat impaired ( approximately 34% wild-type activity at pH 8.5), whereas the K(m)(sulfite) showed a 5-fold increase over wild type. In rapid kinetic assays of the reductive half-reaction of wild-type human sulfite oxidase, k(red)(heme) changed very little over the entire pH range, with a significant increase in K(d)(sulfite) at high pH. The k(red)(heme) of the Y343F variant was significantly impaired across the entire pH range, and unlike the wild-type protein, both k(red)(heme) and K(d)(sulfite) were dependent on pH, with a significant increase in both kinetic parameters at high pH. Additionally, reduction of the molybdenum center by sulfite was directly measured for the first time in rapid reaction assays using sulfite oxidase lacking the N-terminal heme-containing domain. Reduction of the molybdenum center was quite fast (k(red)(Mo) = 972 s(-1) at pH 8.65 for wild-type protein), indicating that this is not the rate-limiting step in the catalytic cycle. Reduction of the molybdenum center of the Y343F variant by sulfite was more significantly impaired at high pH than at low pH. These results demonstrate that the Tyr(343) residue is important for both substrate binding and oxidation of sulfite by sulfite oxidase.  相似文献   

8.
Tyrosine 343 in human sulfite oxidase (SO) is conserved in all SOs sequenced to date. Intramolecular electron transfer (IET) rates between reduced heme (Fe(II)) and oxidized molybdenum (Mo(VI)) in the recombinant wild-type and Y343F human SO were measured for the first time by flash photolysis. The IET rate in wild-type human SO at pH 7.4 is about 37% of that in chicken SO with a similar decrease in k(cat). Steady-state kinetic analysis of the Y343F mutant showed an increase in K(m)(sulfite) and a decrease in k(cat) resulting in a 23-fold attenuation in the specificity constant k(cat)/K(m)(sulfite) at the optimum pH value of 8.25. This indicates that Tyr-343 is involved in the binding of the substrate and catalysis within the molybdenum active site. Furthermore, the IET rate constant in the mutant at pH 6.0 is only about one-tenth that of the wild-type enzyme, suggesting that the OH group of Tyr-343 is vital for efficient IET in SO. The pH dependences of IET rate constants in the wild-type and mutant SO are consistent with the previously proposed coupled electron-proton transfer mechanism.  相似文献   

9.
The molybdenum cofactor (Moco)-containing enzymes are divided into three classes that are named after prototypical members of each family, viz. sulfite oxidase, DMSO reductase and xanthine oxidase. Functional or structural models have been prepared for these three prototypical enzymes: (i) The complex [MoO2(mnt)2]2- (mnt2- = 1,2-dicyanoethylenedithiolate) has been found to be able to oxidize hydrogen sulfite to HSO4- and is thus a functional model of sulfite oxidase. Kinetic and computational studies indicate that the reaction proceeds via attack of the substrate at one of the oxo ligands of the complex, rather than at the metal. (ii) The coordination geometries of the mono-oxo [Mo(VI)(O-Ser)(S2)2] entity (S2 = dithiolene moiety of molybdopterin) found in the crystal structure of R. sphaeroides DMSO reductase and the corresponding des-oxo Mo(IV) unit have been reproduced in the complexes [M(VI)O(OSiR3)(bdt)2] and [M(VI)O(OSiR3)(bdt)2] (M = Mo,W; bdt = benzene dithiolate). (iii) A facile route has been developed for the preparation of complexes containing a cis-Mo(VI)OS molybdenum oxo, sulfido moiety similar to that detected in the oxidized form of xanthine oxidase.  相似文献   

10.
Biodesulfurization of organic-sulfur compounds   总被引:2,自引:0,他引:2  
A screening assay in which dibenzothiophene (DBT) or DBT-sulfone served as the only bioavailable source of sulfur was used to obtain two new bacterial isolates, strains UM9 and UM3, that desulfurized either substrate. Strain UM9 produced the desulfurized product, 2-hydroxybiphenyl (HBP); no other identifiable desulfurized products or released sulfate or sulfite were detected. Biodesulfurization activity occurred only for growing cultures and was depressed by free sulfate. Neither isolate grew on DBT, DBT-sulfone, or HBP as sole carbon sources. Under optimized conditions of pH and temperature, strain UM9 exhibited up to 35% greater biodesulfurization of DBT-sulfone than did UM3, and both isolates also desulfurized several other organic-sulfur compounds. The kinetics and characteristics of biodesulfurization by either UM3 or UM9, tentatively identified as species ofRhodococcus, indicated mechanisms different from those reported in the literature for other bacteria.  相似文献   

11.
The inactivation of sulfite oxidase, a molybdoenzyme containing the Mo cofactor, by arsenite and periodate was investigated. In contrast to ferricyanide (Gardlik, S., and Rajagopalan, K.V. (1991) J. Biol. Chem. 266, 4889-4895), neither of these reagents causes oxidation of the pterin ring of the Mo cofactor. Instead, inactivation by these reagents appears to involve attack on sulfhydryl groups at the active site of the enzyme. The inactivation of sulfite oxidase by arsenite was shown to be dependent on the presence of O2 and on the enzymatic oxidation of arsenite to arsenate. The inactivation was preventable by the presence of sulfite, or by the use of cytochrome c as the electron acceptor instead of O2. It is concluded that inactivation by arsenite is the result of arsenite displacement of Mo during enzymatic oxidation of arsenite to arsenate, when Mo transiently breaks its bond to protein or molybdopterin sulfhydryl(s) in order to provide a site for transfer of electrons to O2. Data indicate that arsenite is properly oriented to displace Mo only once every 20,800 turnovers, thus accounting for the slow rate of inactivation by this reagent. Inactivation of sulfite oxidase by periodate is believed to occur as the result of direct attack of periodate on the thiolate ligands of Mo, either those of the protein and/or molybdopterin, leading to Mo loss. Treatment of enzyme with even low levels of periodate resulted in loss of Mo and both sulfite:cytochrome c and sulfite:O2 activities. Molybdopterin of periodate-inactivated enzyme retained the ability to reconstitute nitrate reductase apoprotein in nit-1 extracts and the ability to reduce dichlorophenolindophenol, indicating that the pterin ring had not been oxidized.  相似文献   

12.
The inhibition of aryl reductive dehalogenation reactions by sulfur oxyanions has been demonstrated in environmental samples, dehalogenating enrichments, and the sulfate-reducing bacterium Desulfomonile tiedjei; however, this phenomenon is not well understood. We examined the effects of sulfate, sulfite, and thiosulfate on reductive dehalogenation in the model microorganism D. tiedjei and found separate mechanisms of inhibition due to these oxyanions under growth versus nongrowth conditions. Dehalogenation activity was greatly reduced in extracts of cells grown in the presence of both 3-chlorobenzoate, the substrate or inducer for the aryl dehalogenation activity, and either sulfate, sulfite, or thiosulfate, indicating that sulfur oxyanions repress the requisite enzymes. In extracts of fully induced cells, thiosulfate and sulfite, but not sulfate, were potent inhibitors of aryl dehalogenation activity even in membrane fractions lacking the cytoplasmically located sulfur oxyanion reductase. These results suggest that under growth conditions, sulfur oxyanions serve as preferred electron acceptors and negatively influence dehalogenation activity in D. tiedjei by regulating the amount of active aryl dehalogenase in cells. Additionally, in vitro inhibition by sulfur oxyanions is due to the interaction of the reactive species with enzymes involved in dehalogenation and need not involve competition between two respiratory processes for reducing equivalents. Sulfur oxyanions also inhibited tetrachloroethylene dehalogenation by the same mechanisms, further indicating that chloroethylenes are fortuitously dehalogenated by the aryl dehalogenase. The commonly observed inhibition of reductive dehalogenation reactions under sulfate-reducing conditions may be due to similar regulation mechanisms in other dehalogenating microorganisms that contain multiple respiratory activities.  相似文献   

13.
Sulfite oxidase (SOX) is a homodimeric molybdoheme enzyme that oxidizes sulfite to sulfate at the molybdenum center. Following substrate oxidation, molybdenum is reduced and subsequently regenerated by two sequential electron transfers (ETs) via heme to cytochrome c. SOX harbors both metals in spatially separated domains within each subunit, suggesting that domain movement is necessary to allow intramolecular ET. To address whether one subunit in a SOX dimer is sufficient for catalysis, we produced heterodimeric SOX variants with abolished sulfite oxidation by replacing the molybdenum-coordinating and essential cysteine in the active site. To further elucidate whether electrons can bifurcate between subunits, we truncated one or both subunits by deleting the heme domain. We generated three SOX heterodimers: (i) SOX/Mo with two active molybdenum centers but one deleted heme domain, (ii) SOX/Mo_C264S with one unmodified and one inactive subunit, and (iii) SOX_C264S/Mo harboring a functional molybdenum center on one subunit and a heme domain on the other subunit. Steady-state kinetics showed 50% SOX activity for the SOX/Mo and SOX/Mo_C264S heterodimers, whereas SOX_C264S/Mo activity was reduced by two orders of magnitude. Rapid reaction kinetics monitoring revealed comparable ET rates in SOX/Mo, SOX/Mo_C264S, and SOX/SOX, whereas in SOX_C264S/Mo, ET was strongly compromised. We also combined a functional SOX Mo domain with an inactive full-length SOX R217W variant and demonstrated interdimer ET that resembled SOX_C264S/Mo activity. Collectively, our results indicate that one functional subunit in SOX is sufficient for catalysis and that electrons derived from either Mo(IV) or Mo(V) follow this path.  相似文献   

14.
CULLIN4-RING E3 ubiquitin ligases (CRL4s) regulate key developmental and stress responses in eukaryotes. Studies in both animals and plants have led to the identification of many CRL4 targets as well as specific regulatory mechanisms that modulate their function. The latter involve COP10-DET1-DDB1 (CDD)–related complexes, which have been proposed to facilitate target recognition by CRL4, although the molecular basis for this activity remains largely unknown. Here, we provide evidence that Arabidopsis thaliana DET1-, DDB1-ASSOCIATED1 (DDA1), as part of the CDD complex, provides substrate specificity for CRL4 by interacting with ubiquitination targets. Thus, we show that DDA1 binds to the abscisic acid (ABA) receptor PYL8, as well as PYL4 and PYL9, in vivo and facilitates its proteasomal degradation. Accordingly, we found that DDA1 negatively regulates ABA-mediated developmental responses, including inhibition of seed germination, seedling establishment, and root growth. All other CDD components displayed a similar regulatory function, although they did not directly interact with PYL8. Interestingly, DDA1-mediated destabilization of PYL8 is counteracted by ABA, which protects PYL8 by limiting its polyubiquitination. Altogether, our data establish a function for DDA1 as a substrate receptor for CRL4-CDD complexes and uncover a mechanism for the desensitization of ABA signaling based on the regulation of ABA receptor stability.  相似文献   

15.
Direct oxidation of sulfite to sulfate occurs in various photo- and chemotrophic sulfur oxidizing microorganisms as the final step in the oxidation of reduced sulfur compounds and is catalyzed by sulfite:cytochrome c oxidoreductase (EC ). Here we show that the enzyme from Thiobacillus novellus is a periplasmically located alphabeta heterodimer, consisting of a 40.6-kDa subunit containing a molybdenum cofactor and an 8.8-kDa mono-heme cytochrome c(552) subunit (midpoint redox potential, E(m8.0) = +280 mV). The organic component of the molybdenum cofactor was identified as molybdopterin contained in a 1:1 ratio to the Mo content of the enzyme. Electron paramagnetic resonance spectroscopy revealed the presence of a sulfite-inducible Mo(V) signal characteristic of sulfite:acceptor oxidoreductases. However, pH-dependent changes in the electron paramagnetic resonance signal were not detected. Kinetic studies showed that the enzyme exhibits a ping-pong mechanism involving two reactive sites. K(m) values for sulfite and cytochrome c(550) were determined to be 27 and 4 micrometer, respectively; the enzyme was found to be reversibly inhibited by sulfate and various buffer ions. The sorAB genes, which encode the enzyme, appear to form an operon, which is preceded by a putative extracytoplasmic function-type promoter and contains a hairpin loop termination structure downstream of sorB. While SorA exhibits significant similarities to known sequences of eukaryotic and bacterial sulfite:acceptor oxidoreductases, SorB does not appear to be closely related to any known c-type cytochromes.  相似文献   

16.
Selected biochemical features of sulfonate assimilation in Escherichia coli K-12 were studied in detail. Competition between sulfonate-sulfur and sulfur sources with different oxidation states, such as cysteine, sulfite and sulfate, was examined. The ability of the enzyme sulfite reductase to attack the C-S linkage of sulfonates was directly examined. Intact cells formed sulfite from sulfonate-sulfur. In cysteine-grown cells, when cysteine was present with either cysteate or sulfate, assimilation of both of the more oxidized sulfur sources was substantially inhibited. In contrast, none of three sulfonates had a competitive effect on sulfate assimilation. In studies of competition between different sulfonates, the presence of taurine resulted in a decrease in cysteate uptake by one-half, while in the presence of isethionate, cysteate uptake was almost completely inhibited. In sulfite-grown cells, sulfonates had no competitive effect on sulfite utilization. An E. coli mutant lacking sulfite reductase and unable to utilize isethionate as the sole source of sulfur formed significant amounts of sulfite from isethionate. In cell extracts, sulfite reductase itself did not utilize sulfonate-sulfur as an electron acceptor. These findings indicate that sulfonate utilization may share some intermediates (e.g. sulfite) and regulatory features (repression by cysteine) of the assimilatory sulfate reductive pathway, but sulfonates do not exert regulatory effects on sulfate utilization. Other results suggest that unrecognized aspects of sulfonate metabolism, such as specific transport mechanisms for sulfonates and different regulatory features, may exist.  相似文献   

17.
18.
A new Desulfovibrio strain ThAc01 was isolated from freshwater mud; the strain conserved energy for growth under strictly anaerobic conditions by disproportionation of thiosulfate or sulfite to sulfate and sulfide according to the following reactions: $$\begin{gathered} S_2 O_3^{2 - } + H_2 O \to SO_4^{2 - } + HS^ - + H^ + \hfill \\ 4SO_3^{2 - } + H^ + {\text{ }} \to 3SO_4^{2 - } + HS^ - \hfill \\ \end{gathered}$$ Strain ThAc01 required acetate as a carbon source, but was unable to utilize acetate as an oxidizable energy source. In a defined medium with acetate and bicarbonate as carbon sources, the growth yields per mol of substrate disproportionated were 2.1 g or 3.2 g dry cell mass on thiosulfate or sulfite, respectively. Strain ThAc01 was also able to grow by dissimilatory sulfate reduction with lactate, ethanol, propanol, or butanol as electron donors and carbon sources which were incompletely oxidized to the corresponding fatty acids. However, growth by sulfate reduction was slower than by disproportionation. Elemental sulfur, nitrate, fumarate, or malate did not serve as electron acceptors. Strain ThAc01 contained desulfoviridin and cytochromes; it required panthothenate and biotin as growth factors and had a DNA base ratio of 64.1 mol% G+C. Disproportionating bacteria similar to strain ThAc01 were enriched with either thiosulfate or sulfite from various freshwater, brackish or marine mud samples. Most probable number enumeration indicated that 2×106 thiosulfate-disproportionating bacteria were present per ml freshwater mud. Of various other sulfate-reducing bacteria tested, only Desulfobacter curvatus (strain AcRM3) was able to disproportionate thiosulfate or sulfite. Desulfovibrio vulgaris (strain Marburg) slowly disproportionated sulfite, but effected only a slight increase in cell density. Strain ThAc01 is proposed as the type strain of a new species, Desulfovibrio sulfodismutans.  相似文献   

19.
Sulfate-reducing bacteria, Desulfovibrio vulgaris, strain Miyazaki, were grown on either sulfate, sulfite, or thiosulfate as the terminal electron acceptor. Better growth was observed on sulfite and less growth on thiosulfate than on sulfate. Enzyme levels of adenylylsulfate (APS) reductase [EC 1.8.99.2], reductant-activated inorganic pyrophosphatase [EC 3.6.1.1], sulfite reductase [EC 1.8.99.1] (desulfoviridin), hydrogenase [EC 1.12.2.1], and Mg2+-activated ATPase [EC 3.6.1.3] were compared in crude extracts of these cells at various stages of growth. 1) The specific activity of APS reductase in sulfite-grown cells was only one-fourth that in sulfate-grown cells throughout growth. Thiosulfate-grown cells had an activity intermediate between those of sulfate- and sulfite-grown cells. 2) Cells grown on sulfite had lower specific activity of reductant-activated inorganic pyrophosphatase than cells grown on sulfate or thiosulfate. 3) The specific activity of sulfite reductase (desulfoviridin) was highest in sulfite-grown cells. The sulfite medium gave the enzyme in high yield as well as with high specific activity. 4) The specific activities of hydrogenase and Mg2+-ATPase were not significantly altered by electron acceptors in the growth medium.  相似文献   

20.
Recent studies have shown that long-term in vivo exposure of dogs to neutral sulfur(IV)/sulfite aerosols induces mild inflammatory reactions, whereas the combination of neutral sulfite with acidic sulfur(VI)/sulfate aerosols evokes less pronounced effects. To understand underlying mechanisms, we studied in vitro the role of lipid mediators in the responses of alveolar macrophages (AMs) to sulfur-related compounds under neutral (pH 7) or moderate acidic (pH 6) conditions. Canine AMs incubated with sulfite at pH 7 released threefold higher amounts of platelet-activating factor than control (P < 0.005). Generation of arachidonic acid, leukotriene B4, 5-hydroxy-eicosatetraenoic acid, prostaglandin E2, thromboxane B2 and 12-hydroxyheptadecatrienoic acid increased twofold (P < 0.0005). However, these metabolites remained unchanged following incubation of AMs with sulfite at pH 6 or with sulfate at pH 7 or pH 6. Mediator release by sulfite-treated AMs at pH 7 stimulated respiratory burst activity of neutrophils. Inhibition of MAPK pathway by PD 98059, of cytosolic (cPLA2) and secretory phospholipases A2 by AACOCF3 and thioetheramide-PC, respectively, reduced sulfite-induced eicosanoid formation in AMs. Sulfite activated cPLA2 activity twofold at pH 7. This mechanism of sulfite-stimulated responses in phospholipid metabolism predicts that chronic exposure to sulfur(IV)/sulfite is associated with a considerable health risk.  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