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.
The first direct voltammetric response from a molybdenum enzyme under non-turnover conditions is reported. Cyclic voltammetry of dimethylsulfoxide reductase from Rhodobacter capsulatus reveals a reversible Mo(VI/V) response at +161 mV followed by a reversible Mo(V/IV) response at -102 mV versus NHE at pH 8. The higher potential couple exhibits a pH dependence consistent with protonation upon reduction to the Mo(V) state and we have determined the p K(a) for this semi-reduced species to be 9.0. The lower potential couple is pH independent within the range 5相似文献   

2.
Hoke KR  Cobb N  Armstrong FA  Hille R 《Biochemistry》2004,43(6):1667-1674
Arsenite oxidase from Alcaligenes faecalis, an unusual molybdoenzyme that does not exhibit a Mo(V) EPR signal during oxidative-reductive titrations, has been investigated by protein film voltammetry. A film of the enzyme on a pyrolytic graphite edge electrode produces a sharp two-electron signal associated with reversible reduction of the oxidized Mo(VI) molybdenum center to Mo(IV). That reduction or oxidation of the active site occurs without accumulation of Mo(V) is consistent with the failure to observe a Mo(V) EPR signal for the enzyme under a variety of conditions and is indicative of an obligate two-electron center. The reduction potential for the molybdenum center, 292 mV (vs SHE) at pH 5.9 and 0 degrees C, exhibits a linear pH dependence for pH 5-10, consistent with a two-electron reduction strongly coupled to the uptake of two protons without a pK in this range. This suggests that the oxidized enzyme is best characterized as having an L(2)MoO(2) rather than L(2)MoO(OH) center in the oxidized state and that arsenite oxidase uses a "spectator oxo" effect to facilitate the oxo transfer reaction. The onset of the catalytic wave observed in the presence of substrate correlates well with the Mo(VI/IV) potential, consistent with catalytic electron transport that is limited only by turnover at the active site. The one-electron peaks for the iron-sulfur centers are difficult to observe by protein film voltammetry, but spectrophotometric titrations have been carried out to measure their reduction potentials: at pH 6.0 and 20 degrees C, that of the [3Fe-4S] center is approximately 260 mV and that of the Rieske center is approximately 130 mV.  相似文献   

3.
Several aspects of the interaction of xanthine oxidase with arsenite are investigated. Room temperature potentiometric titrations using EPR to monitor Molybdenum reduction reveal midpoint potentials of -225 mV for the Mo(VI)-arsenite/Mo(V)-arsenite couple and -440 mV for the Mo(V)-arsenite/Mo(IV)-arsenite couple at pH 8.3. Under the same conditions, the values for native enzyme are -395 mV and -420 mV, respectively. The predicted effects of the altered Mo(VI)/Mo(V) potential on the distributions of reducing equivalents in partially reduced enzyme are compared with the experimentally observed effects in optical experiments. The bleaching that occurs on reduction of the chromophore that is generated when arsenite binds to oxidized enzyme is characterized and found to be associated with reduction of Mo(V)-arsenite to Mo(V)-arsenite. This probe enables determination of the midpoint potential for this conversion using optical data. From such data at a series of pH values ranging from 6.15 to 9.9, a pH dependence of -60 mV/pH unit increase is determined for this couple above pH 7. The ability of arsenite to bind to reduced xanthine oxidase and to desulfo enzyme are also investigated. Reduced active enzyme binds arsenite much more tightly (Kd less than 0.1 microM) and more rapidly than does oxidized active enzyme (Kd = 8 microM); oxidized desulfo enzyme binds arsenite almost as tightly (Kd = 20 microM) as does the oxidized active enzyme.  相似文献   

4.
Bray RC  Adams B  Smith AT  Richards RL  Lowe DJ  Bailey S 《Biochemistry》2001,40(33):9810-9820
The bis-molybdopterin enzyme dimethylsulfoxide reductase (DMSOR) from Rhodobacter capsulatus catalyzes the conversion of dimethyl sulfoxide (DMSO) to dimethyl sulfide (DMS), reversibly, in the presence of suitable e(-)-donors or e(-)-acceptors. The catalytically significant intermediate formed by reaction of DMSOR with DMS ('the DMS species') and a damaged enzyme form derived by reaction of the latter with O(2) (DMS-modified enzyme, DMSOR(mod)D) have been investigated. Evidence is presented that Mo in the DMS species is not, as widely assumed, Mo(IV). Formation of the DMS species is reversed on removing DMS or by addition of an excess of DMSO. Equilibrium constants for the competing reactions of DMS and DMSO with the oxidized enzyme (K(d) = 0.07 +/- 0.01 and 21 +/- 5 mM, respectively) that control these processes indicate formation of the DMS species occurs at a redox potential that is 80 mV higher than that required, according to the literature, for reduction of Mo(VI) to Mo(IV) in the free enzyme. Specificity studies show that with dimethyl selenide, DMSOR yields a species analogous to the DMS species but with the 550 nm peak blue-shifted by 27 nm. It is concluded from published redox potential data that this band is due to metal-to-ligand charge transfer from Mo(V) to the chalcogenide. Since the DMS species gives no EPR signal in the normal or parallel mode, a free radical is presumed to be in close proximity to the metal, most likely on the S. The species is thus formulated as Mo(V)-O-S(*)Me(2). Existing X-ray crystallographic and Raman data are consistent with this structure. Furthermore, 1e(-) oxidation of the DMS species with phenazine ethosulfate yields a Mo(V) form without an -OH ligand, since its EPR signal shows no proton splittings. This form presumably arises via dissociation of DMSO. The structure of DMSOR(mod)D has been determined by X-ray crystallography. All four thiolate ligands and Ogamma of serine-147 remain coordinated to Mo, but there are no terminal oxygen ligands and Mo is Mo(VI). Thus, it is a dead-end species, neither oxo group acceptance nor e(-)-donation being possible. O(2)-dependent formation of DMSOR(mod)D represents noncatalytic breakdown of the DMS species by a pathway alternative to that in turnover, with oxidation to Mo(VI) presumably preceding product release. Steps in the forward and backward catalytic cycles are discussed in relation to earlier stopped-flow data. The finding that in the back-assay the Mo(IV) state may at least in part be by-passed via two successive 1e(-) reactions of the DMS species with the e(-)-acceptor, may have implications in relation to the existence of separate molybdopterin enzymes catalyzing DMSO reduction and DMS oxidation, respectively.  相似文献   

5.
Bacterial cytoplasmic assimilatory nitrate reductases are the least well characterized of all of the subgroups of nitrate reductases. In the present study the ferredoxin-dependent nitrate reductase NarB of the cyanobacterium Synechococcus sp. PCC 7942 was analyzed by spectropotentiometry and protein film voltammetry. Metal and acid-labile sulfide analysis revealed nearest integer values of 4:4:1 (iron/sulfur/molybdenum)/molecule of NarB. Analysis of dithionite-reduced enzyme by low temperature EPR revealed at 10 K the presence of a signal that is characteristic of a [4Fe-4S](1+) cluster. EPR-monitored potentiometric titration of NarB revealed that this cluster titrated as an n = 1 Nernstian component with a midpoint redox potential (E(m)) of -190 mV. EPR spectra collected at 60 K revealed a Mo(V) signal termed "very high g" with g(av) = 2.0047 in air-oxidized enzyme that accounted for only 10-20% of the total molybdenum. This signal disappeared upon reduction with dithionite, and a new "high g" species (g(av) = 1.9897) was observed. In potentiometric titrations the high g Mo(V) signal developed over the potential range of -100 to -350 mV (E(m) Mo(6+/5+) = -150 mV), and when fully developed, it accounted for 1 mol of Mo(V)/mol of enzyme. Protein film voltammetry of NarB revealed that activity is turned on at potentials below -200 mV, where the cofactors are predominantly [4Fe-4S](1+) and Mo(5+). The data suggests that during the catalytic cycle nitrate will bind to the Mo(5+) state of NarB in which the enzyme is minimally two-electron-reduced. Comparison of the spectral properties of NarB with those of the membrane-bound and periplasmic respiratory nitrate reductases reveals that it is closely related to the periplasmic enzyme, but the potential of the molybdenum center of NarB is tuned to operate at lower potentials, consistent with the coupling of NarB to low potential ferredoxins in the cell cytoplasm.  相似文献   

6.
The respiratory molybdoenzyme nitrate reductase (NarGHI) from Escherichia coli has been studied by protein film voltammetry, with the enzyme adsorbed on a rotating disk pyrolytic graphite edge (PGE) electrode. Catalytic voltammograms for nitrate reduction show a complex wave consisting of two components that vary with pH, nitrate concentration, and the presence of inhibitors. At micromolar levels of nitrate, the activity reaches a maximum value at approximately -25 mV and then decreases as the potential becomes more negative. As the nitrate concentration is raised, the activity at more negative potentials increases and eventually becomes the dominant feature at millimolar concentrations. This leads to the hypothesis that nitrate binds more tightly to Mo(V) than Mo(IV), so that low levels of nitrate are more effectively reduced at a higher potential despite the lower driving force. However, an alternative interpretation, that nitrate binding is affected by a change in the redox state of the pterin, cannot be ruled out. This proposal, implicating a specific redox transition at the active site, is supported by experiments carried out using the inhibitors azide and thiocyanate. Azide is the stronger inhibitor of the two, and each inhibitor shows two inhibition constants, one at high potential and one at low potential, both of which are fully competitive with nitrate; closer analysis reveals that the inhibitors act preferentially upon the catalytic activity at high potential. The unusual potential dependence therefore derives from the weaker binding of nitrate or the inhibitors to a more reduced state of the active site. The possible manifestation of these characteristics in vivo has interesting implications for the bioenergetics of E. coli.  相似文献   

7.
Periplasmic nitrate reductase catalyzes the reduction of nitrate into nitrite using a mononuclear molybdenum cofactor that has nearly the same structure in all enzymes of the DMSO reductase family. In previous electrochemical investigations, we found that the enzyme exists in several inactive states, some of which may have been previously isolated and mistaken for catalytic intermediates. In particular, the enzyme slowly and reversibly inactivates when exposed to high concentrations of nitrate. Here, we study the kinetics of substrate inhibition and its dependence on electrode potential and substrate concentration to learn about the properties of the active and inactive forms of the enzyme. We conclude that the substrate-inhibited enzyme never significantly accumulates in the EPR-active Mo(+ V) state. This conclusion is relevant to spectroscopic investigations where attempts are made to trap a Mo(+ V) catalytic intermediate using high concentrations of nitrate.  相似文献   

8.
C J Kay  L P Solomonson  M J Barber 《Biochemistry》1990,29(48):10823-10828
Potentiometric titrations of assimilatory nitrate reductase from Chlorella vulgaris were performed within the pH range 6.0-9.0. Mo(V) was measured by room temperature EPR spectroscopy while the reduction state of FAD was monitored by CD spectroscopy. Between pH 6 and 8.5, the line shape of the Mo(V) EPR signal was constant, exhibiting superhyperfine coupling to a single, exchangeable proton. Potentiometric titrations indicated the Em values for the Mo(VI)/Mo(V) (+61 mV, pH 6) and Mo(V)/Mo(IV) (+35 mV, pH 6) couples decreased with increasing pH by approximately -59 mV/pH unit, consistent with the uptake of a single proton upon reduction of Mo(VI) to Mo(V) and Mo(V) to Mo(IV). The pKa values for the dissociation of these redox-coupled protons appeared to lie outside the pH range studied: pKo(MoVI), pKo(MoV) less than 5.5; pKr(MoV), pKr(MoIV) greater than 9. The Em (n = 2) for FAD (-250 mV, pH 7) varied by approximately -30 mV/pH unit within the pH range 6.0-9.0. Low-temperature EPR potentiometry at the extreme pH values indicated less than 0.5% conversion of FAD to the semiquinone form at the midpoint of the titrations. In contrast, NADH-reduced enzyme exhibited approximately 3-5% of the FAD in the semiquinone form, present as the anionic (FAD.-) species, the spectrum characterized by a line width of 1.3 mT at both pH 6.0 and 9.0.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Glutathione reductase from S. cerevisiae (EC 1.6.4.2) catalyzes the NADPH oxidation by glutathione in accordance with a "ping-pong" scheme. The catalytic constant kcat) is 240 s-1 (pH 7.0, 25 degrees C); kcat for the diaphorase reaction is 4-5 s-1. The enzyme activity does not change markedly at pH 5.5-8.0. At pH less than or equal to 7.0, NADP+ acts as a competitive inhibitor towards NADPH and as a noncompetitive inhibitor towards glutathione. NADP+ increases the diaphorase activity of the enzyme. The maximal activity is observed, when the NADP+/NADPH ratio exceeds 100. At pH 8.0, NADP+ acts as a mixed type inhibitor during the reduction of glutathione. High concentrations of NADP+ also inhibit the diaphorase activity due to the reoxidation of the reduced enzyme by NADP+ at pH 8.0. The redox potential of glutathione reductase calculated from the inhibition data is--306 mV (pH 8.0). Glutathione reductase reduces quinoidal compounds in an one-electron way. The hyperbolic dependence of the logarithm of the oxidation constant on the one electron reduction potential of quinone is observed. It is assumed that quinones oxidize the equilibtium fraction of the two-electron reduced enzyme containing reduced FAD.  相似文献   

10.
The reduction potentials (Em) of the purple acid phosphatase from porcine uterus, uteroferrin (Uf), and its phosphate, arsenate, and molybdate complexes were determined by coulometric methods at various pH values. The midpoint potential of Uf at the pH value for optimal enzyme activity (pH 5) was found to be +367 mV versus a normal hydrogen electrode (NHE), while at pH 6.01 Uf exhibits a reduction potential of +306 mV. At pH 6.01 molybdate was found to shift the potential of Uf more positive by 192 mV, while phosphate and arsenate shift the potential of Uf more negative by 193 and 89 mV, respectively. These shifts are consistent with the different susceptibilities of Uf to aerobic oxidation in the presence of these anions. Comparison of the reduction potential of Uf at pH 7.0 with those reported for other dinuclear non-heme iron enzymes and various (mu-oxo)diiron model complexes suggest that the potential of Uf is too positive to be consistent with a mu-oxo-bridge in Ufo. The pH dependence of the reduction potentials of Uf (60 mV/pH unit) and the fact that the electron transfer rate increases with decreasing pH indicate a concomitant participation of a proton during the oxidation-reduction process. This process was assigned to the protonation of a terminally bound hydroxide ligand at the Fe(II) center upon reduction of Ufo. Structural implications provided by the electrochemical data indicate that molybdate affects the dinuclear core in a manner that differs from that of phosphate and arsenate. This observation is consistent with previous spectroscopic and biochemical studies.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
The thermodynamic and catalytic properties of flavocytochrome c3 from Shewanella frigidimarina have been studied using a combination of protein film voltammetry and solution methods. As measured by solution kinetics, maximum catalytic efficiencies for fumarate reduction (kcat/Km = 2.1 x 10(7) M-1 s-1 at pH 7.2) and succinate oxidation (kcat/Km = 933 M-1 s-1 at pH 8.5) confirm that flavocytochrome c3 is a unidirectional fumarate reductase. Very similar catalytic properties are observed for the enzyme adsorbed to monolayer coverage at a pyrolytic graphite "edge" electrode, thus confirming the validity of the electrochemical method for providing complementary information. In the absence of fumarate, the adsorbed enzyme displays a complex envelope of reversible redox signals which can be deconvoluted to yield the contributions from each active site. Importantly, the envelope is dominated by the two-electron signal due to FAD [E degrees ' = -152 mV vs the standard hydrogen electrode (SHE) at pH 7.0 and 24 degrees C] which enables quantitative examination of this center, the visible spectrum of which is otherwise masked by the intense absorption bands due to the hemes. The FAD behaves as a cooperative two-electron center with a pH-dependent reduction potential that is modulated (pKox at 6.5) by ionization of a nearby residue. In conjunction with the kinetic pKa values determined for the forward and reverse reactions (7.4 and 8.6, respectively), a mechanism for fumarate reduction, incorporating His365 and an anionic form of reduced FAD, is proposed. The reduction potentials of the four heme groups, estimated by analysis of the underlying envelope, are -102, -146, -196, and -238 mV versus the SHE at pH 7.0 and 24 degrees C and are comparable to those determined by redox potentiometry.  相似文献   

12.
The dimethylsulfoxide reductase (DMSOR) from Rhodobacter capsulatus is known to retain its three-dimensional structure and enzymatic activity upon substitution of molybdenum, the metal that occurs naturally at the active site, by tungsten. The redox properties of tungsten-substituted DMSOR (W-DMSOR) have been investigated by a dye-mediated reductive titration with the concentration of the W(V) state monitored by EPR spectroscopy. At pH 7.0, E(m)(W(VI)/W(V)) is -194 mV and E(m)(W(V)/W(IV)) is -134 mV. Each E(m) value of W-DMSOR is significantly lower (220 and 334 mV, respectively) than that of the corresponding couple of Mo-DMSOR. These redox potentials are consistent with the ability of Mo-DMSOR to catalyze both the reduction of DMSO to DMS and the back reaction, whereas W-DMSOR is very effective in catalyzing the forward reaction, but shows no ability to catalyze the oxidation of DMS to DMSO.  相似文献   

13.
The heterogeneity of arginases in rat tissues.   总被引:11,自引:0,他引:11       下载免费PDF全文
1. The mid-point reduction potentials of the various groups in xanthine oxidase from bovine milk were determined by potentiometric titration with dithionite in the presence of dye mediators, removing samples for quantification of the reduced species by e.p.r. (electron-paramagnetic-resonance) spectroscopy. The values obtained for the functional enzyme in pyrophosphate buffer, pH8.2, are: Fe/S centre I, -343 +/- 15mV; Fe/S II, -303 +/- 15mV; FAD/FADH-; -351 +/- 20mV; FADH/FADH2, -236 +/-mV; Mo(VI)/Mo(V) (Rapid), -355 +/- 20mV; Mo(V) (Rapid)/Mo(IV), -355 +/- 20mV. 2. Behaviour of the functional enzyme is essentially ideal in Tris but less so in pyrophosphate. In Tris, the potential for Mo(VI)/Mo(V) (Rapid) is lowered relative to that in pyrophosphate, but the potential for Fe/S II is raised. The influence of buffer on the potentials was investigated by partial-reduction experiments with six other buffers. 3. Conversion of the enzyme with cyanide into the non-functional form, which gives the Slow molybdenum signal, or alkylation of FAD, has little effect on the mid-point potentials of the other centres. The potentials associated with the Slow signal are: Mo(VI)/Mo(V) (Slow), -440 +/- 25mV; Mo(V) (Slow)/Mo(IV), -480 +/- 25 mV. This signal exhibits very sluggish equilibration with the mediator system. 4. The deviations from ideal behaviour are discussed in terms of possible binding of buffer ions or anti-co-operative interactions amongst the redox centres.  相似文献   

14.
The succinate dehydrogenases (SDH: soluble, membrane-extrinsic subunits of succinate:quinone oxidoreductases) from Escherichia coli and beef heart mitochondria each adsorb at a pyrolytic graphite 'edge' electrode and catalyse the interconversion of succinate and fumarate according to the electrochemical potential that is applied. E. coli and beef heart mitochondrial SDH share only ca. 50% homology, yet the steady-state catalytic activities, when measured over a continuous potential range, display very similar catalytic operating potentials and energetic biases (the relative ability to catalyse succinate oxidation vs. fumarate reduction). Importantly, E. coli SDH also exhibits the interesting 'tunnel-diode' behaviour previously reported for the mitochondrial enzyme. Thus as the potential is lowered below ca. -60 mV (pH 7, 38 degrees C) the rate of catalytic fumarate reduction decreases abruptly despite an increase in driving force. Since the homology relates primarily to residues associated with active site regions, the marked similarity in the voltammetry reaffirms our previous conclusions that the tunnel-diode behaviour is a characteristic property of the enzyme active site. Thus, succinate dehydrogenase is an excellent fumarate reductase, but its activity in this direction is limited to a very specific range of potential.  相似文献   

15.
IsdG and IsdI are paralogous heme degrading enzymes from the bacterium Staphylococcus aureus. Heme bound by these enzymes is extensively ruffled such that the meso-carbons at the sites of oxidation are distorted toward bound oxygen. In contrast, the canonical heme oxygenase family degrades heme that is bound with minimal distortion. Trp-66 is a conserved heme pocket residue in IsdI implicated in heme ruffling. IsdI variants with Trp-66 replaced with residues having less bulky aromatic and alkyl side chains were characterized with respect to catalytic activity, heme ruffling, and electrochemical properties. The heme degradation activity of the W66Y and W66F variants was approximately half that of the wild-type enzyme, whereas the W66L and W66A variants were inactive. A crystal structure and NMR spectroscopic analysis of the W66Y variant reveals that heme binds to this enzyme with less heme ruffling than observed for wild-type IsdI. The reduction potential of this variant (−96 ± 7 mV versus standard hydrogen electrode) is similar to that of wild-type IsdI (−89 ± 7 mV), so we attribute the diminished activity of this variant to the diminished heme ruffling observed for heme bound to this enzyme and conclude that Trp-66 is required for optimal catalytic activity.  相似文献   

16.
Dimethyl sulfide dehydrogenase from the purple phototrophic bacterium Rhodovulum sulfidophilum catalyzes the oxidation of dimethyl sulfide to dimethyl sulfoxide. Recent DNA sequence analysis of the ddh operon, encoding dimethyl sulfide dehydrogenase (ddhABC), and biochemical analysis (1) have revealed that it is a member of the DMSO reductase family of molybdenum enzymes and is closely related to respiratory nitrate reductase (NarGHI). Variable temperature X-band EPR spectra (120-122 K) of purified heterotrimeric dimethyl sulfide dehydrogenase showed resonances arising from multiple redox centers, Mo(V), [3Fe-4S](+), [4Fe-4S](+), and a b-type heme. A pH-dependent EPR study of the Mo(V) center in (1)H(2)O and (2)H(2)O revealed the presence of three Mo(V) species in equilibrium, Mo(V)-OH(2), Mo(V)-anion, and Mo(V)-OH. Above pH 8.2 the dominant species was Mo(V)-OH. The maximum specific activity occurred at pH 9.27. Comparison of the rhombicity and anisotropy parameters for the Mo(V) species in DMS dehydrogenase with other molybdenum enzymes of the DMSO reductase family showed that it was most similar to the low-pH nitrite spectrum of Escherichia coli nitrate reductase (NarGHI), consistent with previous sequence analysis of DdhA and NarG. A sequence comparison of DdhB and NarH has predicted the presence of four [Fe-S] clusters in DdhB. A [3Fe-4S](+) cluster was identified in dimethyl sulfide dehydrogenase whose properties resembled those of center 2 of NarH. A [4Fe-4S](+) cluster was also identified with unusual spin Hamiltonian parameters, suggesting that one of the iron atoms may have a fifth non-sulfur ligand. The g matrix for this cluster is very similar to that found for the minor conformation of center 1 in NarH [Guigliarelli, B., Asso, M., More, C., Augher, V., Blasco, F., Pommier, J., Giodano, G., and Bertrand, P. (1992) Eur. J. Biochem. 307, 63-68]. Analysis of a ddhC mutant showed that this gene encodes the b-type cytochrome in dimethyl sulfide dehydrogenase. Magnetic circular dichroism studies revealed that the axial ligands to the iron in this cytochrome are a histidine and methionine, consistent with predictions from protein sequence analysis. Redox potentiometry showed that the b-type cytochrome has a high midpoint redox potential (E degrees = +315 mV, pH 8).  相似文献   

17.
B Adams  A T Smith  S Bailey  A G McEwan  R C Bray 《Biochemistry》1999,38(26):8501-8511
Improved assays for the molybdenum enzyme dimethylsulfoxide reductase (DMSOR) with dimethyl sulfoxide (DMSO) and with dimethyl sulfide (DMS) as substrates are described. Maximum activity was observed at pH 6.5 and below and at 8.3, respectively. Rapid-scan stopped-flow spectrophotometry has been used to investigate the reduction of the enzyme by DMS to a species previously characterized by its UV-visible spectrum [McAlpine, A. S., McEwan, A. G., and Bailey, S. (1998) J. Mol. Biol. 275, 613-623], and its subsequent reoxidation by DMSO. Both these two-electron reactions were faster than enzyme turnover under steady-state conditions, indicating that one-electron reactions with artificial dyes were rate-limiting. Second-order rate constants for the two-electron reduction and reoxidation reactions at pH 5.5 were (1.9 +/- 0.1) x 10(5) and (4.3 +/- 0.3) x 10(2) M-1 s-1, respectively, while at pH 8.0, the catalytic step was rate-limiting (62 s-1). Kinetically, for the two-electron reactions, the enzyme is more effective in DMS oxidation than in DMSO reduction. Reduction of DMSOR by DMS was incomplete below approximately 1 mM DMS but complete at higher concentrations, implying that the enzyme's redox potential is slightly higher than that of the DMS-DMSO couple. In contrast, reoxidation of the DMS-reduced state by DMSO was always incomplete, regardless of the DMSO concentration. Evidence for the existence of a spectroscopically indistinguishable reduced state, which could not be reoxidized by DMSO, was obtained. Brief reaction (less than approximately 15 min) of DMS with DMSOR was fully reversible on removal of the DMS. However, in the presence of excess DMS, a further slow reaction occurred aerobically, but not anaerobically, to yield a stable enzyme form having a lambdamax at 660 mn. This state (DMSORmod) retained full activity in steady-state assays with DMSO, but was inactive toward DMS. It could however be reconverted to the original resting state by reduction with methyl viologen radical and reoxidation with DMSO. We suggest that in this enzyme form two of the dithiolene ligands of the molybdenum have dissociated and formed a disulfide. The implications of this new species are discussed in relation both to conflicting published information for DMSOR from X-ray crystallography and to previous spectroscopic data for its reduced forms.  相似文献   

18.
Redox potentials for the various centres in the enzyme xanthine dehydrogenase (EC 1.2.1.37) from turkey liver determined by potentiometric titration in the presence of mediator dyes, with low-temperature electron-paramagnetic-resonance spectroscopy. Values at 25 degrees C in pyrophosphate buffer, pH 8.2, are: Mo(VI)/Mo(V)(Rapid),-350 +/- 20mV; Mo(V) (Rapid)/Mo(IV), -362 +/- 20mV; Fe-S Iox./Fe-S Ired., -295 +/- 15mV; Fe-S IIox./Fe-S IIred., -292 +/- 15mV; FAD/FADH,-359+-20mV; FADH/FADH2, -366 +/- 20mV. This value of the FADH/FADH2 potential, which is 130mV lower than the corresponding one for milk xanthine oxidase [Cammack, Barber & Bray (1976) Biochem. J. 157, 469-478], accounts for many of the differences between the two enzymes. When allowance is made for some interference by desulpho enzyme, then differences in the enzymes' behaviour in titration with xanthine [Barber, Bray, Lowe & Coughlan (1976) Biochem. J. 153, 297-307] are accounted for by the potentials. Increases in the molybdenum potentials of the enzymes caused by the binding of uric acid are discussed. Though the potential of uric acid/xanthine (-440mV) is favourable for full reduction of the dehydrogenase, nevertheless, during turnover, for kinetic reasons, only FADH and very little FADH2 is produced from it. Since only FADH2 is expected to react with O2, lack of oxidase activity by the dehydrogenase is explained. Reactivity of the two enzymes with NAD+ as electron acceptor is discussed in relation to the potentials.  相似文献   

19.
Trevor D. Rapson 《BBA》2008,1777(10):1319-1325
Under hydrodynamic electrochemical conditions with slow cyclic voltammetry sweep rates we have been able to probe catalytic events at the molybdenum active site of sulfite dehydrogenase (SDH) from Starkeya novella adsorbed on an edge plane graphite electrode within a polylysine film. The electrochemically driven catalytic behaviour of SDH mirrors that seen in solution assays suggesting that the adsorbed enzyme retains its native activity. However, at high sulfite concentrations, the voltammetric waveform transforms from the expected sigmoidal profile to a peak-shaped response, similar to that reported for the molybdenum enzymes DMSO reductase and nitrate reductase (NarGHI and NapAB) where a redox reaction at the active site has been associated with a switch to lower activity at high overpotentials. This is the first time a similar phenomenon has been observed in a Mo-containing oxidase/dehydrogenase, which raises a number of interesting mechanistic problems. The potential at which the activity of SDH becomes attenuated only emerges at saturating substrate conditions and occurs at a potential (ca. + 320mV vs NHE) well removed from any known redox couple in the enzyme. These results cannot be explained by the same mechanism adopted for DMSO reductase and nitrate reductase catalysis.  相似文献   

20.
Cyclic voltammetry was used for simultaneous formation and immobilization of nickel oxide nano-scale islands and catalase on glassy carbon electrode. Electrodeposited nickel oxide may be a promising material for enzyme immobilization owing to its high biocompatibility and large surface. The catalase films assembled on nickel oxide exhibited a pair of well defined, stable and nearly reversible CV peaks at about -0.05 V vs. SCE at pH 7, characteristic of the heme Fe (III)/Fe (II) redox couple. The formal potential of catalase in nickel oxide film were linearly varied in the range 1-12 with slope of 58.426 mV/pH, indicating that the electron transfer is accompanied by single proton transportation. The electron transfer between catalase and electrode surface, (k(s)) of 3.7(+/-0.1) s(-1) was greatly facilitated in the microenvironment of nickel oxide film. The electrocatalytic reduction of hydrogen peroxide at glassy carbon electrode modified with nickel oxide nano-scale islands and catalase enzyme has been studied. The embedded catalase in NiO nanoparticles showed excellent electrocatalytic activity toward hydrogen peroxide reduction. Also the modified rotating disk electrode shows good analytical performance for amperometric determination of hydrogen peroxide. The resultant catalase/nickel oxide modified glassy carbon electrodes exhibited fast amperometric response (within 2 s) to hydrogen peroxide reduction (with a linear range from 1 microM to 1 mM), excellent stability, long term life and good reproducibility. The apparent Michaelis-Menten constant is calculated to be 0.96(+/-0.05)mM, which shows a large catalytic activity of catalase in the nickel oxide film toward hydrogen peroxide. The excellent electrochemical reversibility of redox couple, high stability, technical simplicity, lake of need for mediators and short preparations times are advantages of this electrode. Finally the activity of biosensor for nitrite reduction was also investigated.  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