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 reaction of fully reduced and mixed-valence cytochrome oxidase with O2 has been followed in flow-flash experiments, starting from the CO complexes, at 428, 445, 605 and 830 nm between pH 5.8b and 9.0 in the temperature range of 2-40 degrees C. With the fully reduced enzyme, four kinetic phase with rate constants at pH 7.4 and 25 degrees C of 9 x 10(4), 2.5 x 10(4), 1.0 x 10(4) and 800 s(-1), respectively, are observed. The rates of the three last phases display a very small temperature dependence, corresponding to activation energies in the range 13-54 kJ x mol(-1). The rates of the third and fourth phases decrease at high pH due to the deprotonation of groups with pKa values of 8.3 and 8.8, respectively, but also the second phase appears to have a small pH dependence. In the reaction of the mixed-valence enzyme, three kinetic phases with rate constants at pH 7.4 and 25 degrees C of 9 x 10(4), 6000 and 150 s(-1), respectively, are observed. The third phase only has a small temperature dependence, corresponding to an activation energy of 20 kJ x mol(-1). No pH dependence could be detected for any phase. Reaction schemes consistent with the experimental observations are presented. The pH dependencies of the rates of the two final phase in the reaction of the fully reduced enzyme are proposed to be related to the involvement of protons in the reduction of a peroxide intermediate. The temperature dependence data suggest that the reorganization energies and driving forces are closely matched in all electron transfer steps with both enzyme forms. It is suggested that the slowest step in the reaction of the mixed-valence enzyme is a conformation change involved in the reaction cycle of cytochrome oxidase as a proton pump.  相似文献   

2.
A study is presented on proton transfer associated with the reaction of the fully reduced, purified bovine heart cytochrome c oxidase with molecular oxygen or ferricyanide. The proton consumption associated with aerobic oxidation of the four metal centers changed significantly with pH going from approximately 3.0 H(+)/COX at pH 6.2-6.3 to approximately 1.2 H(+)/COX at pH 8.0-8.5. Rereduction of the metal centers was associated with further proton uptake which increased with pH from approximately 1.0 H(+)/COX at pH 6.2-6.3 to approximately 2.8 H(+)/COX at pH 8.0-8.5. Anaerobic oxidation of the four metal centers by ferricyanide resulted in the net release of 1.3-1.6 H(+)/COX in the pH range 6.2-8.2, which were taken up by the enzyme on rereduction of the metal centers. The proton transfer elicited by ferricyanide represents the net result of deprotonation/protonation reactions linked to anaerobic oxidoreduction of the metal centers. Correction for the ferricyanide-induced pH changes of the proton uptake observed in the oxidation and rereduction phase of the reaction of the reduced oxidase with oxygen gave a measure of the proton consumption in the reduction of O(2) to 2H(2)O. The results show that the expected stoichiometric proton consumption of 4H(+) in the reduction of O(2) to 2H(2)O is differently associated, depending on the actual pH, with the oxidation and reduction phase of COX. Two H(+)/COX are initially taken up in the reduction of O(2) to two OH(-) groups bound to the binuclear Fe a(3)-Cu(B) center. At acidic pHs the third and fourth protons are also taken up in the oxidative phase with formation of 2H(2)O. At alkaline pHs the third and fourth protons are taken up with formation of 2H(2)O only upon rereduction of COX.  相似文献   

3.
The pH dependencies of the rate constants in the photocycles of recombinant D96N and D115N/D96N bacteriorhodopsins were determined from time-resolved difference spectra between 70 ns and 420 ms after photoexcitation. The results were consistent with the model suggested earlier for proteins containing D96N substitution: BR hv----K----L----M1----M2----BR. Only the M2----M1 back-reaction was pH-dependent: its rate increased with increasing [H+] between pH 5 and 8. We conclude from quantitative analysis of this pH dependency that its reverse, the M1----M2 reaction, is linked to the release of a proton from a group with a pKa = 5.8. This suggests a model for wild-type bacteriorhodopsin in which at pH greater than 5.8 the transported proton is released on the extracellular side from this as yet unknown group and on the 100-microseconds time scale, but at pH less than 5.8, the proton release occurs from another residue and later in the photocycle most likely directly from D85 during the O----BR reaction. We postulate, on the other hand, that proton uptake on the cytoplasmic side will be by D96 and during the N----O reaction regardless of pH. The proton kinetics as measured with indicator dyes confirmed the unique prediction of this model: at pH greater than 6, proton release preceded proton uptake, but at pH less than 6, the release was delayed until after the uptake. The results indicated further that the overall M1----M2 reaction includes a second kinetic step in addition to proton release; this is probably the earlier postulated extracellular-to-cytoplasmic reorientation switch in the proton pump.  相似文献   

4.
The kinetic mechanism of lactose transport across the cytoplasmic membrane has been investigated and the results related to standard models for the lactose-H+ symport reaction using computer simulation. It is shown that the biphasic kinetics reported for lactose uptake (Kaczorowski, G.J. and Kaback, H.R. (1979) Biochemistry 18, 3691-3697) are consistent with random binding of lactose and protons and rapid subsequent translocation of the ternary lactose-H+-permease complex. Such a model is also shown to explain the observed dependence of the kinetic parameters on the magnitude of the protonmotive force. Both sugar and protons are shown to cause product inhibition of lactose flux and the ability of standard models to account for the pattern of inhibition is discussed. Three apparent dissociation constants have been determined for the protonation reactions in the external medium: two (pKa 6.3 and 9.6) control the activity of the permease, whilst the third (pKa 8.3) controls the affinity of the permease for galactosides. A similar set of dissociation constants has been determined for the internal reactions. Again two (pKa 6 and 9.8) control activity and a third (pKa 8.8) controls the affinity for galactosides. The dissociation reactions characterised by pKa 8.3, 8.8, 9.6 and 9.8 are attributed to the dissociation of the substrate (symported) proton from the binary proton-permease complexes (pKa 8.3 and 8.8) and the ternary proton-galactoside-permease complexes (pKa 9.6 and 9.8). The third pair (pKa 6.3 and 6.0) must be interpreted as describing a separate protonation reaction which may have a regulatory or auxiliary role in transport.  相似文献   

5.
The pH-dependence of the kinetic parameters in H2O2 decomposition by beef liver catalase was investigated. At pH 7.0, the ternary complex (ESS) decomposition rate was about 100 times faster than ESS formation (42 microM H2O2), and the value of the Michaelis constant was 0.025 M. From ethanol competition experiments, two different proton dissociation constants of the enzyme (pKe1 = 5.0, pKes2 = 5.9) were obtained for the binding of first and second H2O2 molecules. Another pKa value (pKes1) of 4.2 was obtained from the pH dependence of overall rate constant (ko). The reaction mechanism of catalase was discussed in relation to these ionizable groups.  相似文献   

6.
Acidianus ambivalens is a hyperthermoacidophilic archaeon which grows optimally at approximately 80 degrees C and pH 2.5. The terminal oxidase of its respiratory system is a membrane-bound quinol oxidase (cytochrome aa(3)) which belongs to the heme-copper oxidase superfamily. One difference between this quinol oxidase and a majority of the other members of this family is that it lacks the highly-conserved glutamate (Glu(I-286), E. coli ubiquinol oxidase numbering) which has been shown to play a central role in controlling the proton transfer during reaction of reduced oxidases with oxygen. In this study we have investigated the dynamics of the reaction of the reduced A. ambivalens quinol oxidase with O(2). With the purified enzyme, two kinetic phases were observed with rate constants of 1.8&z.ccirf;10(4) s(-1) (at 1 mM O(2), pH 7.8) and 3. 7x10(3) s(-1), respectively. The first phase is attributed to binding of O(2) to heme a(3) and oxidation of both hemes forming the 'peroxy' intermediate. The second phase was associated with proton uptake from solution and it is attributed to formation of the 'oxo-ferryl' state, the final state in the absence of quinol. In the presence of bound caldariella quinol (QH(2)), heme a was re-reduced by QH(2) with a rate of 670 s(-1), followed by transfer of the fourth electron to the binuclear center with a rate of 50 s(-1). Thus, the results indicate that the quinol donates electrons to heme a, followed by intramolecular transfer to the binuclear center. Moreover, the overall electron and proton-transfer kinetics in the A. ambivalens quinol oxidase are the same as those in the E. coli ubiquinol oxidase, which indicates that in the A. ambivalens enzyme a different pathway is used for proton transfer to the binuclear center and/or other protonatable groups in an equivalent pathway are involved. Potential candidates in that pathway are two glutamates at positions (I-80) and (I-83) in the A. ambivalens enzyme (corresponding to Met(I-116) and Val(I-119), respectively, in E. coli cytochrome bo(3)).  相似文献   

7.
Deuterium kinetic isotope effects (KIE) were measured, and proton inventory plots were constructed, for the rates of reactions in the photocycles of wild-type bacteriorhodopsin and several site-specific mutants. Consistent with earlier reports from many groups, very large KIEs were observed for the third (and largest) rise component for the M state and for the decay of the O state, processes both linked to proton transfers in the extracellular region. The proton inventory plots (ratio of reaction rates in mixtures of H(2)O and D(2)O to that in H(2)O vs mole fraction of D(2)O) were approximately linear for the first and second M rise components and for M decay, as well as for O decay, indicating that the rates of these reactions are limited by simple proton transfer. Uniquely, the third rise component of M (and in the D96N mutant also a fourth rise component) exhibited a strongly curved proton inventory plot, suggesting that its rate, which largely accounts for the rate of deprotonation of the retinal Schiff base, depends on a complex multiproton process. This curvature is observed also in the E194Q, E204Q, and Y57F mutants but not in the R82A mutant. From these findings, and from the locations of bound water in the extracellular region in the crystal structure of the protein [Luecke, Schobert, Richter, Cartailler, and Lanyi (1999) J. Mol. Biol. 291, 899-911], we suspect that the effects of deuterium substitution on the formation of the M state originate from cooperative rearrangements of the extensively hydrogen-bonded water molecules 401, 402, and 406 near Asp-85 and Arg-82.  相似文献   

8.
Stopped-flow techniques were utilized to investigate the kinetics of the reaction of lignin peroxidase compounds I and II (LiPI and LiPII) with veratryl alcohol (VA). All rate data were collected from single turnover experiments under pseudo first-order conditions. The reaction of LiPI with VA strictly obeys second-order kinetics over the pH range 2.72-5.25 as demonstrated by linear plots of the pseudo first-order rate constants versus concentrations of VA. The second-order rate constants are strongly dependent on pH and range from 2.62 x 10(6) M-1 s-1 (pH 2.72) to 1.45 x 10(4) M-1 s-1 (pH 5.25). The reaction of LiPII and VA exhibits saturation behavior when the observed pseudo first-order rate constants are plotted against VA concentrations. The saturation phenomenon is quantitatively explained by the formation of a 1:1 LiPII-substrate complex. Results of kinetic and rapid scan spectral analyses exclude the formation of a LiPII-VA cation radical complex. The first-order dissociation rate constant and the equilibrium dissociation constant for the LiPII reaction are also pH dependent. Binding of VA to LiPII is controlled by a heme-linked ionizable group of pKa approximately 4.2. The pH profiles of the second-order rate constants for the LiPI reaction and of the first-order dissociation constants for the LiPII reaction both demonstrate two pKa values at approximately 3.0 and approximately 4.2. Protonated oxidized enzyme intermediates are most active, suggesting that only electron transfer, not proton uptake from the reducing substrate, occurs at the enzyme active site. These results are consistent with the one-electron oxidation of VA to an aryl cation radical by LiPI and LiPII.  相似文献   

9.
J J Driscoll  D J Kosman 《Biochemistry》1987,26(12):3429-3436
Solvent and solvent proton dependent steps involved in the mechanism of the enzyme galactose oxidase have been examined. The deuterium kinetic solvent isotope effect (KSIE) on the velocity of the galactose oxidase catalyzed oxidation of methyl beta-galactopyranoside by O2 was measured. Examination of the thermodynamic activation parameters for the reaction indicated that the isotope effect was attributable to a slightly less favorable delta H value, consistent with a KSIE on proton transfer. A detailed kinetic analysis was performed, examining the effect of D2O on the rate of reaction over the pH range 4.8-8.0. Both pL-rate profiles exhibited bell-shaped curves. Substitution of D2O as solvent shifted the pKes values for the enzymic central complex: pKes1 from 6.30 to 6.80 and pKes2 from 7.16 to 7.35. Analysis of the observed shifts in dissociation constants was performed with regard to potential hydrogenic sites. pKes1 can be attributed to a histidine imidazole, while pKes2 is tentatively assigned to a Cu2+-bound water molecule. A proton inventory was performed (KSIE = +1.55); the plot of kcat vs. mole fraction D2O was linear, indicating the existence of a single solvent-derived proton involved in a galactose oxidase rate-determining step (or steps). The pH dependence of CN- inhibition was also examined. The Ki-pH profile indicated that a group ionization, with pKa = 7.17, modulated CN- inhibition; Ki was at a minimum when this group was in the protonated state. The inhibition profile followed the alkaline limit of the pH-rate profile for the enzymic reaction, suggesting that the group displaced by CN- was also deprotonating above pH 7.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Flock U  Watmough NJ  Adelroth P 《Biochemistry》2005,44(31):10711-10719
The respiratory nitric oxide reductase (NOR) from Paracoccus denitrificans catalyzes the two-electron reduction of NO to N(2)O (2NO + 2H(+) + 2e(-) --> N(2)O + H(2)O), which is an obligatory step in the sequential reduction of nitrate to dinitrogen known as denitrification. NOR has four redox-active cofactors, namely, two low-spin hemes c and b, one high-spin heme b(3), and a non-heme iron Fe(B), and belongs to same superfamily as the oxygen-reducing heme-copper oxidases. NOR can also use oxygen as an electron acceptor; this catalytic activity was investigated in this study. We show that the product in the steady-state reduction of oxygen is water. A single turnover of the fully reduced NOR with oxygen was initiated using the flow-flash technique, and the progress of the reaction monitored by time-resolved optical absorption spectroscopy. Two major phases with time constants of 40 micros and 25 ms (pH 7.5, 1 mM O(2)) were observed. The rate constant for the faster process was dependent on the O(2) concentration and is assigned to O(2) binding to heme b(3) at a bimolecular rate constant of 2 x 10(7) M(-)(1) s(-)(1). The second phase (tau = 25 ms) involves oxidation of the low-spin hemes b and c, and is coupled to the uptake of protons from the bulk solution. The rate constant for this phase shows a pH dependence consistent with rate limitation by proton transfer from an internal group with a pK(a) = 6.6. This group is presumably an amino acid residue that is crucial for proton transfer to the catalytic site also during NO reduction.  相似文献   

11.
A lysine instead of the usual carboxyl group is in place of the internal proton donor to the retinal Schiff base in the light-driven proton pump of Exiguobacterium sibiricum (ESR). The involvement of this lysine in proton transfer is indicated by the finding that its substitution with alanine or other residues slows reprotonation of the Schiff base (decay of the M intermediate) by more than 2 orders of magnitude. In these mutants, the rate constant of the M decay linearly decreases with a decrease in proton concentration, as expected if reprotonation is limited by the uptake of a proton from the bulk. In wild type ESR, M decay is biphasic, and the rate constants are nearly pH-independent between pH 6 and 9. Proton uptake occurs after M formation but before M decay, which is especially evident in D2O and at high pH. Proton uptake is biphasic; the amplitude of the fast phase decreases with a pKa of 8.5 ± 0.3, which reflects the pKa of the donor during proton uptake. Similarly, the fraction of the faster component of M decay decreases and the slower one increases, with a pKa of 8.1 ± 0.2. The data therefore suggest that the reprotonation of the Schiff base in ESR is preceded by transient protonation of an initially unprotonated donor, which is probably the ϵ-amino group of Lys-96 or a water molecule in its vicinity, and it facilitates proton delivery from the bulk to the reaction center of the protein.  相似文献   

12.
The factors determining the pH dependence of the formation and decay of the O photointermediate of the bacteriorhodopsin (bR) photocycle were investigated in the wild-type (WT) pigment and in the mutants of Glu-194 and Glu-204, key residues of the proton release group (PRG) in bR. We have found that in the WT the rate constant of O --> bR transition decreases 30-fold upon decreasing the pH from 6 to 3 with a pKa of about 4.3. D2O slows the rise and decay of the O intermediate in the WT at pH 3.5 by a factor of 5.5. We suggest that the rate of the O --> bR transition (which reflects the rate of deprotonation of the primary proton acceptor Asp-85) at low pH is controlled by the deprotonation of the PRG. To test this hypothesis, we studied the E194D mutant. We show that the pKa of the PRG in the ground state of the E194D mutant, when Asp-85 is protonated, is increased by 1.2 pK units compared to that of the WT. We found a similar increase in the pKa of the rate constant of the O --> bR transition in E194D. This provides further evidence that the rate of the O --> bR transition is controlled by the PRG. In a further test, the E194Q mutation, which disables the PRG and slows proton release, almost completely eliminates the pH dependence of O decay at pHs below 6. A second phenomenon we investigated was that in the WT at neutral and alkaline pH the fraction of the O intermediate decreases with pKa 7.5. A similar pH dependence is observed in the mutants in which the PRG is disabled, E194Q and E204Q, suggesting that the decrease in the fraction of the O intermediate with pKa ca. 7.5 is not controlled by the PRG. We propose that the group with pKa 7.5 is Asp-96. The slowing of the reprotonation of Asp-96 at high pH is the cause of the decrease in the rate of the N --> O transition, leading to the decrease in the fraction of O.  相似文献   

13.
Proton and electron transfer events during the reaction of solubilized fully reduced bovine heart cytochrome c oxidase with molecular oxygen were investigated using the flow-flash technique. Time-resolved spectral changes resulting from ligand binding and electron transfer events were detected simultaneously with pH changes in the bulk. The kinetics and spectral changes in the visible region (450-750 nm) were probed by optical multichannel detection, allowing high spectral resolution on time scales from 50 ns to 50 ms. Experiments were carried out in the presence and absence of pH-sensitive dyes (carboxyfluorescein at pH 6.5, phenol red at pH 7.5, and m-cresol purple at pH 8.5) which permitted separation of spectral changes due to proton transfer from those caused by ligand binding and electron transfer. The transient spectra recorded in the absence of dye were analyzed by singular-value decomposition and multiexponential fitting. Five apparent lifetimes (0.93 microseconds, 10 microseconds, 36 microseconds, 90 microseconds, and 1.3 ms at pH 7.5) could consistently be distinguished and provided a basis for a reaction mechanism consistent with our most recent kinetic model [Sucheta, A., Szundi, I., and Einarsdóttir, O. (1999) Biochemistry 37, 17905-17914]. The dye response indicated that proton uptake occurred concurrently with the two slowest electron transfer steps, in agreement with previous results based on single-wavelength detection [Hallén, S., and Nilsson, T. (1992) Biochemistry 31, 11853-11859]. The stoichiometry of the proton uptake reactions was approximately 1.3 +/- 0.3, 1.4 +/- 0.3, and 1.6 +/- 0.5 protons per enzyme at pH 6.5, 7.5, and 8.5, respectively. The electron transfer between heme a and CuA was limited by proton uptake on a 90 microseconds time scale. We have established the lower limit of the true rate constant for the electron transfer between CuA and heme a to be approximately 2 x 10(5) s-1.  相似文献   

14.
Changes in pH during the reactions of the fully reduced and mixed-valence cytochrome oxidase with molecular oxygen have been followed in flow-flash experiments, using the pH indicator phenol red. Solubilized enzyme as well as enzyme reconstituted into phospholipid vesicles has been studied. With the solubilized enzyme, a biphasic uptake of one proton from the medium was observed, whereas the reconstituted enzyme gave release of 1.3 protons to the extravesicular medium. It is concluded from these results that a total of two to three protons are taken up during oxidation of the fully reduced enzyme. Kinetic analysis suggests that the proton uptake is initiated by the transfer of the third electron to the oxygen binding site. A reaction scheme that integrates proton transfers and oxygen chemistry is presented.  相似文献   

15.
Superoxide reductase (SOR) is a metalloenzyme that catalyzes the reduction of O2*- to H2O2 and provides an antioxidant mechanism in some anaerobic and microaerophilic bacteria. Its active site contains an unusual mononuclear ferrous center (center II). Protonation processes are essential for the reaction catalyzed by SOR, since two protons are required for the formation of H2O2. We have investigated the acido-basic and pH dependence of the redox properties of the active site of SOR from Desulfoarculus baarsii, both in the absence and in the presence of O2*-. In the absence of O2*-, the reduction potential and the absorption spectrum of the iron center II exhibit a pH transition. This is consistent with the presence of a base (BH) in close proximity to the iron center which modulates its reduction properties. Studies of mutants of the closest charged residues to the iron center II (E47A and K48I) show that neither of these residues are the base responsible for the pH transitions. However, they both interact with this base and modulate its pKa value. By pulse radiolysis, we confirm that the reaction of SOR with O2*- involves two reaction intermediates that were characterized by their absorption spectra. The precise step of the catalytic cycle in which one protonation takes place was identified. The formation of the first reaction intermediate, from a bimolecular reaction of SOR with O2*-, does not involve proton transfer as a rate-limiting step, since the rate constant k1 does not vary between pH 5 and pH 9.5. On the other hand, the rate constant k2 for the formation of the second reaction intermediate is proportional to the H+ concentration in solution, suggesting that the proton arises directly from the solvent. In fact, BH, E47, and K48 have no role in this step. This is consistent with the first intermediate being an iron(III)-peroxo species and the second one being an iron(III)-hydroperoxo species. We propose that BH may be involved in the second protonation process corresponding to the release of H2O2 from the iron(III)-hydroperoxo species.  相似文献   

16.
Saccharopine dehydrogenase [N6-(glutaryl-2)-L-lysine:NAD oxidoreductase (L-lysine forming)] catalyzes the final step in the alpha-aminoadipate pathway for lysine biosynthesis. It catalyzes the reversible pyridine nucleotide-dependent oxidative deamination of saccharopine to generate alpha-Kg and lysine using NAD+ as an oxidizing agent. The proton shuttle chemical mechanism is proposed on the basis of the pH dependence of kinetic parameters, dissociation constants for competitive inhibitors, and isotope effects. In the direction of lysine formation, once NAD+ and saccharopine bind, a group with a pKa of 6.2 accepts a proton from the secondary amine of saccharopine as it is oxidized. This protonated general base then does not participate in the reaction again until lysine is formed at the completion of the reaction. A general base with a pKa of 7.2 accepts a proton from H2O as it attacks the Schiff base carbon of saccharopine to form the carbinolamine intermediate. The same residue then serves as a general acid and donates a proton to the carbinolamine nitrogen to give the protonated carbinolamine. Collapse of the carbinolamine is then facilitated by the same group accepting a proton from the carbinolamine hydroxyl to generate alpha-Kg and lysine. The amine nitrogen is then protonated by the group that originally accepted a proton from the secondary amine of saccharopine, and products are released. In the reverse reaction direction, finite primary deuterium kinetic isotope effects were observed for all parameters with the exception of V2/K(NADH), consistent with a steady-state random mechanism and indicative of a contribution from hydride transfer to rate limitation. The pH dependence, as determined from the primary isotope effect on DV2 and D(V2/K(Lys)), suggests that a step other than hydride transfer becomes rate-limiting as the pH is increased. This step is likely protonation/deprotonation of the carbinolamine nitrogen formed as an intermediate in imine hydrolysis. The observed solvent isotope effect indicates that proton transfer also contributes to rate limitation. A concerted proton and hydride transfer is suggested by multiple substrate/solvent isotope effects, as well as a proton transfer in another step, likely hydrolysis of the carbinolamine. In agreement, dome-shaped proton inventories are observed for V2 and V2/K(Lys), suggesting that proton transfer exists in at least two sequential transition states.  相似文献   

17.
J Heberle  D Oesterhelt    N A Dencher 《The EMBO journal》1993,12(10):3721-3727
Surface bound pH indicators were applied to study the proton transfer reactions in the mutant Asp85-->Glu of bacteriorhodopsin in the native membrane. The amino acid replacement induces a drastic acceleration of the overall rise of the M intermediate. Instead of following this acceleration, proton ejection to the extracellular membrane surface is not only two orders of magnitude slower than M formation, it is also delayed as compared with the wild-type. This demonstrates that Asp85 not only accepts the proton released by the Schiff's base but also regulates very efficiently proton transfer within the proton release chain. Furthermore, Asp85 might be the primary but is not the only proton acceptor/donor group in the release pathway. The Asp85-->Glu substitution also affects the proton reuptake reaction at the cytoplasmic side, although Asp85 is located in the proton release pathway. Proton uptake is slower in the mutant than in the wild-type and occurs during the lifetime of the O intermediate. This demonstrates a feed-back mechanism between Asp85 and the proton uptake pathway in bacteriorhodopsin.  相似文献   

18.
In wild-type bacteriorhodopsin light-induced proton release occurs before uptake at neutral pH. In contrast, in mutants in which R82 is replaced by a neutral residue (as in R82A and R82Q), only a small fraction of the protons is released before proton uptake at neutral pH; the major fraction is released after uptake. In R82Q the relative amounts of the two types of proton release, "early" (preceding proton uptake) and "late" (following proton uptake), are pH dependent. The main conclusions are that 1) R82 is not the normal light-driven proton release group; early proton release can be observed in the R82Q mutant at higher pH values, suggesting that the proton release group has not been eliminated. 2) R82 affects the pKa of the proton release group both in the unphotolyzed state of the pigment and during the photocycle. In the wild type (in 150 mM salt) the pKa of this group decreases from approximately 9.5 in the unphotolyzed pigment to approximately 5.8 in the M intermediate, leading to early proton release at neutral pH. In the R82 mutants the respective values of pKa of the proton release group in the unphotolyzed pigment and in M are approximately 8 and 7.5 in R82Q (in 1 M salt) and approximately 8 and 6.5 in R82K (in 150 mM KCl). Thus in R82Q the pKa of the proton release group does not decrease enough in the photocycle to allow early proton release from this group at neutral pH. 3) Early proton release in R82Q can be detected as a photocurrent signal that is kinetically distinct from those photocurrents that are due to proton movements from the Schiff base to D85 during M formation and from D96 to the Schiff base during the M-->N transition. 4) In R82Q, at neutral pH, proton uptake from the medium occurs during the formation of O. The proton is released during the O-->bacteriorhodopsin transition, probably from D85 because the normal proton release group cannot deprotonate at this pH. 5) The time constant of early proton release is increased from 85 microseconds in the wild type to 1 ms in R82Q (in 150 mM salt). This can be directly attributed to the increase in the pKa of the proton release group and also explains the uncoupling of proton release from M formation. 6) In the E204Q mutant only late proton release is observed at both neutral and alkaline pH, consistent with the idea that E204 is the proton release group. The proton release is concurrent with the O-->bacteriorhodopsin transition, as in R82Q at neutral pH.  相似文献   

19.
Cytochrome c-554 of the ammonia-oxidizing chemolithoautotropic bacteria is thought to mediate electron transfer from hydroxylamine oxidoreductase to a terminal oxidase and/or to ammonia monooxygenase. The cytochrome has four c hemes which interact magnetically and have the same redox potential. We report that the kinetics of reduction of ferric cytochrome c-554 by dithionite or the oxidation of ferrous cytochrome c-554 by O2 or H2O2 are complex and multiphasic. Transient rapid-scan difference spectra indicate discrete maxima at approximately 418 nm, 425 nm and 432 nm. Absorbance changes at all three difference maxima appear to occur in all kinetic phases, although not in equal amounts for each wavelength. Reduction by 20 mM dithionite was biphasic. At pH 7.5 the first phase, which involved approximately 50% of the total absorbance change, had a rate constant (20 degrees C) of 140 s-1 and energy of activation of 20 kJ X mol-1. The slow phase had a rate constant 0.43 s-1 and a relatively high energy of activation, 87 kJ X mol-1, suggesting that a change in protein configuration accompanied the reaction. As the pH of the solution increased, the rate constant for both phases decreased and the fraction of absorbance change in the rapid phase increased. Oxidation of ferrous cytochrome c-554 by O2 involved a discrete rapid phase with a rate constant of 14 s-1, accounting for 6% of the absorbance. The remainder of the reaction was multiphasic with rate constants in the range 0.1-0.01 s-1. With H2O2 as the oxidant, the rapid phase involved 39% of the change in absorbance with a rate constant of 19 s-1. The remainder of the reoxidation was multiphasic with rate constants ranging over 0.4-0.01 s-1.  相似文献   

20.
C Grubmeyer  H Teng 《Biochemistry》1999,38(22):7355-7362
L-Histidinol dehydrogenase catalyzes the biosynthetic oxidation of L-histidinol to L-histidine with sequential reduction of two molecules of NAD. Previous isotope exchange results had suggested that the oxidation of histidinol to the intermediate histidinaldehyde occurred 2-3-fold more rapidly than overall catalysis. In this work, we present kinetic isotope effects (KIE) studies at pH 9.0 and at pH 6.7 with stereospecifically mono- and dideuterated histidinols. The data at pH 9.0 support minimal participation of the first hydride transfer and substantial participation of the second hydride transfer in the overall rate limitation. Stopped-flow experiments with protiated histidinol revealed a small burst of NADH production with stoichiometry of 0.12 per subunit, and 0.25 per subunit with dideuterated histidinol, indicating that the overall first half-reaction was not significantly faster than the second reaction sequence. Results from kcat and kcat/KM titrations with histidinol, NAD, and the alternative substrate imidazolyl propanediol demonstrated an essential base with pKa values between 7.7 and 8.4. In KIE experiments performed at pH 6.7 or with a coenzyme analogue at pH 9. 0, the first hydride transfer became more rate limiting. Kinetic simulations based on rate constants estimated from this work fit well with a mechanism that includes a relatively fast, and thermodynamically unfavorable, hydride transfer from histidinol and a slower, irreversible second hydride transfer from a histidinaldehyde derivative. Thus, although the chemistry of the first hydride transfer is fast, both partial reactions participate in the overall rate limitation.  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