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 156 毫秒
1.
Thermodynamic equilibria and electron transfer kinetics involving the quinone acceptor complex in reaction centers from Rhodopseudomonas sphaeroides were investigated. We focussed on reactions involving the two-electron states QA Qn and QAQ~-, described by the scheme DQAQa~-D +X,~~A- , ~~a- ~k~ .~D+ "r~~ AK~'La2- - k~2~ k~lk O~ (2)D+~D The equilibrium partitioning between QA Q n and QAQ 2n- was determined spectroscopically from either the concentration of oxidized cytochrome c or the concentration of semiquinone after successive flashes of light.At pH < 9.5, QAQ2n - is stabilized relative to QAQn, while for pH > 9.5, QAQB is energetically favored.The reduction of QA, to form QAQ~, is not associated with a protonation step (pK< 8). However, the reduction of Q~, to form the final state QAQ~-, is accompanied by an uptake of a proton (pK >/10.7). The preferential interaction of a proton with QAQ2n - provides the driving force for the forward electron transfer.The shift toward the photochemically inactive state QAQa with increasing pH may serve as a feedback mechanism in photosynthetic organisms to limit the rise in intracellular pH. The electron-transfer rate constants were determined from the observed kinetics and the equilibria between the states QAQ2n - and QA Q n. The forward rate constant z-.A~2n~ was approximately proportional to the proton concentration, whereas kta2A~ depended only weakly on pH. The recombination kinetics of D +QAQ2n- was biphasic. The slow rate agreed with the predicted charge recombination via the intermediate state D +QAQff; the fast rate may be due to the recombination from a separate (conformational) state. The results of this work were combined with those of a previous study on reactions involving the one-electron precursor states QAQa and QAQn(Kleinfeld, D., Okamura, M.Y., and Feher, G. (1984) Biochim. Biophys. Acta 766, 126-140). The overall sequence for the protonation of the reaction center in response to successive reductions of the accept or complex involves the uptake of one proton for each electron transferred to QB- This sequential uptake initiates the formation of a proton gradient across the cell membrane.  相似文献   

2.
Xu Q  Gunner MR 《Biochemistry》2002,41(8):2694-2701
Both large- and small-scale conformational changes are needed as proteins carry out reactions. However, little is known about the identity, energy of, and barriers between functional substates on protein reaction coordinates. In isolated bacterial photosynthetic reaction centers, the electron transfer from the reduced primary quinone, Q(A)(-), to the secondary quinone, Q(B), is rate limited by conformational changes at low pH and by proton binding at high pH. The kinetics and thermodynamics of this reaction were determined between 200 and 300 K from pH 6 to pH 10.5. A model with two substates of the reactant, P(+)Q(A)(-)Q(B), one protonated (state A) and one unprotonated (alpha), and one state of the product, P(+)Q(A)Q(B)(-) (B), was able to simulate the dependence of the rate on temperature and pH fairly well. The equilibrium between the three states were measured in situ at each temperature. Proton binding (alpha to A transition) has a favorable DeltaH and unfavorable DeltaS as does the conformational changes required for electron transfer at low pH (A to B). The pK for the A to alpha transition is 9.7 at room temperature, consistent with previous measurements, and equivalent to 13.5 at 200 K. The activation barriers were determined for each transition. Both the alpha to A and the A to B transitions are limited primarily by the activation enthalpy with modest DeltaS.  相似文献   

3.
A quantitative model for the damping of oscillations of the semiquinone absorption after successive light flashes is presented. It is based on the equilibrium between the states Q(A)-Q(B) and Q(A) Q(-B). A fit of the model to the experimental results obtained for reaction centers from Rhodopseudomonas sphaeroides gave a value of α = [Q(A)-Q(B)I/(IQ(A)-Q(Bl)+ [Q(A)Q(-B)I) = 0.065 +/- 0.005 (T= 21°C, pH 8).  相似文献   

4.
Li J  Takahashi E  Gunner MR 《Biochemistry》2000,39(25):7445-7454
The electron transfer from the reduced primary quinone (Q(A)(-)) to the secondary quinone (Q(B)) can occur in two phases with a well-characterized 100 micros component (tau(2)) and a faster process occurring in less than 10 micros (tau(1)). The fast reaction is clearly seen when the native ubiquinone-10 at Q(A) is replaced with naphthoquinones. The dependence of tau(1) on the free-energy difference between the P(+)Q(A)(-)Q(B) and P(+)Q(A)Q(B)(-) states (-) and on the pH was measured using naphthoquinones with different electrochemical midpoint potentials as Q(A) in Rhodobacter sphaeroides reaction centers (RCs) and in RCs where - is changed by mutation of M265 in the Q(A) site from Ile to Thr (M265IT). Q(B) was ubiquinone (UQ(B)) in all cases. Electron transfer was measured by using the absorption differences of the naphthosemiquinone at Q(A) and the ubisemiquinone at Q(B) between 390 and 500 nm. As - was changed from -90 to -250 meV tau(1) decreased from 29 to 0.2 micros. The free-energy dependence of tau(1) provides a reorganization energy of 850 +/- 100 meV for the electron transfer from Q(A)(-) to Q(B). The slower reaction at tau(2) is free-energy independent, so processes other than electron transfer determine the observed rate. The fraction of the reaction at tau(1) increases with increasing driving force and is 100% of the reaction when - is approximately 100 meV more favorable than in the native RCs with ubiquinone as Q(A). The fast phase, tau(1), is pH independent from pH 6 to 11 while tau(2) slows above pH 9. As the Q(A) isoprene tail length is increased from 2 to 10 isoprene units the fraction at tau(1) decreases. However, tau(1), tau(2), and the fraction of the reaction in each phase are independent of the tail length of UQ(B).  相似文献   

5.
Rapid-scan Fourier transform infrared (FTIR) difference spectroscopy was used to investigate the electron transfer reaction Q(A-)Q(B)-->Q(A)Q(B-) (k(AB)(1)) in mutant reaction centers of Rhodobacter sphaeroides, where Asp-L210 and/or Asp-M17 have been replaced with Asn. Mutation of both residues decreases drastically k(AB)(1)), attributed to slow proton transfer to Glu-L212, which becomes rate limiting for electron transfer to Q(B) [M.L. Paddock et al., Biochemistry 40 (2001) 6893]. In the double mutant, the FTIR difference spectrum recorded during the time window 4-29 ms following a flash showed peaks at 1670 (-), 1601 (-) and 1467 (+) cm(-1), characteristic of Q(A) reduction. The time evolution of the spectra shows reoxidation of Q(A-) and concomitant reduction of Q(B) with a kinetics of about 40 ms. In native reaction centers and in both single mutants, formation of Q(B-) occurs much faster than in the double mutant. Within the time resolution of the technique, protonation of Glu-L212, as characterized by an absorption increase at 1728 cm(-1) [E. Nabedryk et al., Biochemistry 34 (1995) 14722], was found to proceed with the same kinetics as reduction of Q(B) in all samples. These rapid-scan FTIR results support the model of proton uptake being rate limiting for the first electron transfer from Q(A-) to Q(B) and the identification of Glu-L212 as the main proton acceptor in the state Q(A)Q(B-).  相似文献   

6.
The pathway for proton transfer to Q(B) was studied in the reaction center (RC) from Rhodobacter sphaeroides. The binding of Zn(2+) or Cd(2+) to the RC surface at His-H126, His-H128, and Asp-H124 inhibits the rate of proton transfer to Q(B), suggesting that the His may be important for proton transfer [Paddock, M. L., Graige, M. S., Feher, G. and Okamura, M. Y. (1999) Proc. Natl. Acad. Sci. U.S.A. 96, 6183-6188]. To assess directly the role of the histidines, mutant RCs were constructed in which either one or both His were replaced with Ala. In the single His mutant RCs, no significant effects were observed. In contrast, in the double mutant RC at pH 8.5, the observed rates of proton uptake associated with both the first and the second proton-coupled electron-transfer reactions k(AB)(()(1)()) [Q(A)(-)(*)Q(B)-Glu(-) + H(+) --> Q(A)(-)(*)Q(B)-GluH --> Q(A)Q(B)(-)(*)-GluH] and k(AB)(()(2)()) [Q(A)(-)(*)Q(B)(-)(*) + H(+) --> Q(A)(-)(*)(Q(B)H)(*) --> Q(A)(Q(B)H)(-)], were found to be slowed by factors of approximately 10 and approximately 4, respectively. Evidence that the observed changes in the double mutant RC are due to a reduction in the proton-transfer rate constants are provided by the observations: (i) k(AB)(1) at pH approximately pK(a) of GluH became biphasic, indicating that proton transfer is slower than electron transfer and (ii) k(AB)(2) became independent of the driving force for electron transfer, indicating that proton transfer is the rate-limiting step. These changes were overcome by the addition of exogenous imidazole which acts as a proton donor in place of the imidazole groups of His that were removed in the double mutant RC. Thus, we conclude that His-H126 and His-H128 facilitate proton transfer into the RC, acting as RC-bound proton donors at the entrance of the proton-transfer pathways.  相似文献   

7.
8.
The reaction center (RC) from Rhodobacter sphaeroides uses light energy to reduce and protonate a quinone molecule, Q(B) (the secondary quinone electron acceptor), to form quinol, Q(B)H2. Asp-L210 and Asp-M17 have been proposed to be components of the pathway for proton transfer [Axelrod, H. L., Abresch, E. C., Paddock, M. L., Okamura, M. Y., and Feher, G. (2000) Proc. Natl. Acad. Sci. U.S.A. 97, 1542-1547]. To test the importance of these residues for efficient proton transfer, the rates of the proton-coupled electron-transfer reaction k(AB)(2) (Q(A-*)Q(B-*) + H+ <==>Q(A-*)Q(B)H* --> Q(A)Q(B)H-) and its associated proton uptake were measured in native and mutant RCs, lacking one or both Asp residues. In the double mutant RCs, the k(AB)(2) reaction and its associated proton uptake were approximately 300-fold slower than in native RCs (pH 8). In contrast, single mutant RCs displayed reaction rates that were < or =3-fold slower than native (pH 8). In addition, the rate-limiting step of k(AB)(2) was changed from electron transfer (native and single mutants) to proton transfer (double mutant) as shown from the lack of a dependence of the observed rate on the driving force for electron transfer in the double mutant RCs compared to the native or single mutants. This implies that the rate of the proton-transfer step was reduced (> or =10(3)-fold) upon replacement of both Asp-L210 and Asp-M17 with Asn. Similar, but less drastic, differences were observed for k(AB)(1), which at pH > or =8 is coupled to the protonation of Glu-L212 [(Q(A-*)Q(B))-Glu- + H+ --> (Q(A)Q(B-*)-GluH]. These results show that the pathway for proton transfer from solution to reduced Q(B) involves both Asp-L210 and Asp-M17, which provide parallel branches to the proton-transfer pathway and through their electrostatic interaction have a cooperative effect on the proton-transfer rate. A possible mechanism for the cooperativity is discussed.  相似文献   

9.
Ishikita H  Knapp EW 《Biochemistry》2005,44(45):14772-14783
In photosystem II (PSII), the redox properties of the non-heme iron complex (Fe complex) are sensitive to the redox state of quinones (Q(A/)(B)), which may relate to the electron/proton transfer. We calculated the redox potentials for one-electron oxidation of the Fe complex in PSII [E(m)(Fe)] based on the reference value E(m)(Fe) = +400 mV at pH 7 in the Q(A)(0)Q(B)(0) state, considering the protein environment in atomic detail and the associated changes in protonation pattern. Our model yields the pH dependence of E(m)(Fe) with -60 mV/pH as observed in experimental redox titration. We observed significant deprotonation at D1-Glu244 in the hydrophilic loop region upon Fe complex oxidation. The calculated pK(a) value for D1-Glu244 depends on the Fe complex redox state, yielding a pK(a) of 7.5 and 5.5 for Fe(2+) and Fe(3+), respectively. To account for the pH dependence of E(m)(Fe), a model involving not only D1-Glu244 but also the other titratable residues (five Glu in the D-de loops and six basic residues near the Fe complex) seems to be needed, implying the existence of a network of residues serving as an internal proton reservoir. Reduction of Q(A/B) yields +302 mV and +268 mV for E(m)(Fe) in the Q(A)(-)Q(B)(0) and Q(A)(0)Q(B)(-) states, respectively. Upon formation of the Q(A)(0)Q(B)(-) state, D1-His252 becomes protonated. Forming Fe(3+)Q(B)H(2) by a proton-coupled electron transfer process from the initial state Fe(2+)Q(B)(-) results in deprotonation of D1-His252. The two EPR signals observed at g = 1.82 and g = 1.9 in the Fe(2+)Q(A)(-) state of PSII may be attributed to D1-His252 with variable and fixed protonation, respectively.  相似文献   

10.
Ginet N  Lavergne J 《Biochemistry》2001,40(6):1812-1823
The equilibrium and kinetic parameters for the binding of various inhibitors to the Q(B) pocket of the bacterial reaction center were investigated in chromatophores from Rhodobacter capsulatus and Rhodobacter sphaeroides. By monitoring the near-IR absorption changes specific to Q(A)(-) and Q(B)(-), we measured the fraction of inhibited centers in the dark and the kinetics and extent of inhibitor displacement after one flash due to the formation of the Q(A)Q(B)(-) state. The inhibitor release rate was much faster for triazines and o-phenanthroline (t(1/2) in the 50 ms to 1 s range) than for stigmatellin (t(1/2) approximately 20 s). For inhibitors with a rapid release rate, the fast phase of P(+) decay observed in the absence of secondary donor reflects the competition between P(+)Q(A)(-) recombination and inhibitor release: it is thus faster than the P(+)Q(A)(-) recombination, and its relative extent is smaller than the fraction of initially inhibited centers. At appropriate inhibitor concentrations, one can have almost total binding in the dark and almost total inhibitor displacement after one flash. Under such conditions, a pair of closely spaced flashes resets the two-electron gate in a single state (Q(A)Q(B)(-)), irrespective of the initial state. The apparent dissociation constant of terbutryn was significantly increased (by a factor of 4-7) in the presence of Q(A)(-), in agreement with the conclusion of Wraight and co-workers [Stein, R. R., et al. (1984) J. Cell. Biochem. 24, 243-259]. We suggest that this effect is essentially due to a tighter binding of ubiquinone in the Q(A)(-) state.  相似文献   

11.
The oxidized primary electron donor in photosystem II, P(680)(+), is reduced in several phases, extending over 4 orders of magnitude in time. Especially the slower phases may reflect the back-pressure exerted by water oxidation and provide information on the reactions involved. The kinetics of secondary electron-transfer reactions in the microseconds time range after charge separation were investigated in oxygen-evolving thylakoids suspended in H2O or D2O. Flash-induced changes of chlorophyll fluorescence yield and electric field-induced recombination luminescence were decomposed into contributions from oxidation states S(0), S(1), S(2), and S(3) of the oxygen-evolving complex and interpreted in terms of stabilization kinetics of the initial charge-separated state S(j)Y(Z)P(680)(+)Q(A)(-)Q(B). In approximately 10% of the centers, only charge recombination took place. Otherwise, no static heterogeneity was involved in the microsecond reduction of P(680)(+) by Y(Z) (stabilization) or Q(A)(-) (recombination). The recombination component in active centers occurs mainly upon charge separation in S(3), and, in the presence of D2O, in S(2) as well and is tentatively attributed to the presence of Y(Z)(ox)S(j-1) in equilibrium with Y(Z)S(j). A 20-30 micros stabilization occurs in all S-states, but to different extents. Possible mechanisms for this component are discussed. D2O was found to decrease: (i) the rate of the reaction Y(Z)(ox)S(1) to Y(Z)S(2), (ii) the equilibrium constant between P680(+)Y(Z)S(2) and P(680)Y(Z)(ox)S(2), (iii) the rate of the slow phase of P(680)(+) reduction for the S(3) --> S(0) transition, and (iv) the rate of electron transfer from Q(A)(-) to Q(B) /Q(B)(-). The increased 'miss probability' in D2O is due to (iii).  相似文献   

12.
The secondary quinone acceptor, Q(B), has been studied in photosystem II (PSII) isolated from Thermosynechococcus (T.) elongatus. Thermoluminescence indicated that Q(B) was present in this preparation. An EPR signal observed at low temperature at g = 1.9 was attributed to Fe2+ Q(B)- on the basis of the characteristic period-of-two variations in its intensity depending on the number of laser flashes given at 20 degrees C. When samples showing the Fe2+ Q(B)- signal were illuminated at 77 K, an EPR signal at g = 1.66 appeared with an amplitude proportional to that of the Fe2+ Q(B)- signal. This signal is attributed to the Q(A)- Fe2+ Q(B)- state. While these attributions have been made previously in PSII from other origins, they have remained relatively tentative since the characteristic period-of-two oscillations of Q(B) had not previously been observed. The flash experiments indicated that more than one exchangeable plastoquinone is associated with the isolated PSII. The g = 1.66 signal from the Q(A)- Fe2+ Q(B)- state was used to study the temperature dependence of electron transfer between the two quinones. Electron transfer occurred in half of the centers (after 30 s incubation) at -28 degrees C for Q(A)- to Q(B) but at -58 degrees C for Q(A)- to Q(B)-. This marked difference for the two electron transfer reactions indicates different types of rate-limiting reactions. In the better studied but homologous system, the purple bacterial reaction center, the Q(A)- to Q(B) step is limited by a gating process, while the Q(A)- to Q(B)- step is limited by protonation events. Similar reactions in PSII could give rise to the observed temperature dependence.  相似文献   

13.
To gain new insights into the function of photosystem II (PSII) herbicides DCMU (a urea herbicide) and bromoxynil (a phenolic herbicide), we have studied their effects in a better understood system, the bacterial photosynthetic reaction center of the terbutryn-resistant mutant T4 of Blastochloris (Bl.) viridis. This mutant is uniquely sensitive to these herbicides. We have used redox potentiometry and time-resolved absorption spectroscopy in the nanosecond and microsecond time scale. At room temperature the P(+)(*)Q(A)(-)(*) charge recombination in the presence of bromoxynil was faster than in the presence of DCMU. Two phases of P(+)(*)Q(A)(-)(*) recombination were observed. In accordance with the literature, the two phases were attributed to two different populations of reaction centers. Although the herbicides did induce small differences in the activation barriers of the charge recombination reactions, these did not explain the large herbicide-induced differences in the kinetics at ambient temperature. Instead, these were attributed to a change in the relative amplitude of the phases, with the fast:slow ratio being approximately 3:1 with bromoxynil and approximately 1:2 with DCMU at 300 K. Redox titrations of Q(A) were performed with and without herbicides at pH 6.5. The E(m) was shifted by approximately -75 mV by bromoxynil and by approximately +55 mV by DCMU. As the titrations were done over a time range that is assumed to be much longer than that for the transition between the two different populations, the potentials measured are considered to be a weighted average of two potentials for Q(A). The influence of the herbicides can thus be considered to be on the equilibrium of the two reaction center forms. This may also be the case in photosystem II.  相似文献   

14.
M S Graige  M L Paddock  G Feher  M Y Okamura 《Biochemistry》1999,38(35):11465-11473
A proton-activated electron transfer (PAET) mechanism, involving a protonated semiquinone intermediate state, had been proposed for the electron-transfer reaction k(2)AB [Q(A)(-)(*)Q(B)(-)(*) + H(+) <--> Q(A)(-)(*)(Q(B)H)(*) --> Q(A)(Q(B)H)(-)] in reaction centers (RCs) from Rhodobacter sphaeroides [Graige, M. S., Paddock, M. L., Bruce, M. L., Feher, G., and Okamura, M. Y. (1996) J. Am. Chem. Soc. 118, 9005-9016]. Confirmation of this mechanism by observing the protonated semiquinone (Q(B)H)(*) had not been possible, presumably because of its low pK(a). By replacing the native Q(10) in the Q(B) site with rhodoquinone (RQ), which has a higher pK(a), we were able to observe the (Q(B)H)(*) state. The pH dependence of the semiquinone optical spectrum gave a pK(a) = 7.3 +/- 0.2. At pH < pK(a), the observed rate for the reaction was constant and attributed to the intrinsic electron-transfer rate from Q(A)(-)(*) to the protonated semiquinone (i.e., k(2)AB = k(ET)(RQ) = 2 x 10(4) s(-)(1)). The rate decreased at pH > pK(a) as predicted by the PAET mechanism in which fast reversible proton transfer precedes rate-limiting electron transfer. Consequently, near pH 7, the proton-transfer rate k(H) > 10(4) s(-)(1). Applying the two step mechanism to RCs containing native Q(10) and taking into account the change in redox potential, we find reasonable values for the fraction of (Q(B)H)(*) congruent with 0.1% (consistent with a pK(a)(Q(10)) of approximately 4.5) and k(ET)(Q(10)) congruent with 10(6) s(-)(1). These results confirm the PAET mechanism in RCs with RQ and give strong support that this mechanism is active in RCs with Q(10) as well.  相似文献   

15.
The kinetics of the charge recombination D+QA-----DQA was used to probe the protonation of the primary acceptor in reaction centers from Rhodopseudomonas sphaeroides, in which the native ubiquinone was replaced by anthraquinone. We found that QA- is stabilized by the rapid (t less than 10(-2) s) binding of a proton, with a pK of 9.8. The distance between QA- and the proton binding site was estimated to be larger than approximately 5 A.  相似文献   

16.
O A Gopta  A Y Semenov  D A Bloch 《FEBS letters》2001,499(1-2):116-120
An electrometric technique was used to investigate the effect of coenzyme Q(10) (UQ), substitution by decylubiquinone (dQ) at the Q(B) binding site of reaction centers (UQ-RC and dQ-RC, respectively) on the electrogenic proton transfer kinetics upon Q(B) reduction in Rhodobacter sphaeroides chromatophores. Unlike dQ-RC, the kinetics of the second flash-induced proton uptake in UQ-RC clearly deviated from the mono-exponential one. The activation energy (about 30 kJ/mol) and the pH profile of the kinetics in dQ-RC were similar to those in UQ-RC, with the power law approximation used in the latter case. The interpretation of the data presumed the quinone translocation between the two binding positions within the Q(B) site. It is proposed that the native isoprenyl side chain (in contrast to decyl chain) favors the equilibrium binding of neutral quinone at the redox-active 'proximal' position, but causes a higher barrier for the hydroquinone movement from 'proximal' to 'distal' position.  相似文献   

17.
Nabedryk E  Paddock ML  Okamura MY  Breton J 《Biochemistry》2005,44(44):14519-14527
In the photosynthetic reaction center (RC) from the purple bacterium Rhodobacter sphaeroides, proton-coupled electron-transfer reactions occur at the secondary quinone (Q(B)) site. Several nearby residues are important for both binding and redox chemistry involved in the light-induced conversion from Q(B) to quinol Q(B)H(2). Ser-L223 is one of the functionally important residues located near Q(B). To obtain information on the interaction between Ser-L223 and Q(B) and Q(B)(-), isotope-edited Q(B)(-)/Q(B) FTIR difference spectra were measured in a mutant RC in which Ser-L223 is replaced with Ala and compared to the native RC. The isotope-edited IR fingerprint spectra for the C=O [see text] and C=C [see text] modes of Q(B) (Q(B)(-)) in the mutant are essentially the same as those of the native RC. These findings indicate that highly equivalent interactions of Q(B) and Q(B)(-) with the protein occur in both native and mutant RCs. The simplest explanation of these results is that Ser-L223 is not hydrogen bonded to Q(B) or Q(B)(-) but presumably forms a hydrogen bond to a nearby acid group, preferentially Asp-L213. The rotation of the Ser OH proton from Asp-L213 to Q(B)(-) is expected to be an important step in the proton transfer to the reduced quinone. In addition, the reduced quinone remains firmly bound, indicating that other distinct hydrogen bonds are more important for stabilizing Q(B)(-). Implications on the design features of the Q(B) binding site are discussed.  相似文献   

18.
The temperature dependence of charge recombination from the P+QA- and from the P+QB- states produced by a flash was studied in reaction centers isolated from the photosynthetic thermophilic bacterium Chloroflexus aurantiacus. P designates the primary electron donor; QA and QB the primary and secondary quinone electron acceptors respectively. In QB-depleted reaction centers the rate constant (kAP) for P+QA- recombination was temperature independent between 0-50 degrees C (17.6 +/- 0.7 s-1 at pH 8 and pH 10). The same value was obtained in intact membranes in the presence of o-phenanthroline. Upon lowering the temperature from 250 K to 160 K, kAP increased by a factor of two and remained constant down to 80 K. The overall temperature dependence of kAP was consistent with an activationless process. Ubiquinone (UQ-3) and different types of menaquinone were used for QB reconstitution. In UQ-3 reconstituted reaction centers charge recombination was monoexponential (rate constant k = 0.18 +/- 0.03 s-1) and temperature independent between 5-40 degrees C. In contrast, in menaquinone-3- and menaquinone-4-reconstituted reaction centers P+ rereduction following a flash was markedly biphasic and temperature dependent. In menaquinone-6-reconstituted reaction centers a minor contribution from a third kinetic phase corresponding to P+QA- charge recombination was detected. Analysis of these kinetics and of the effects of the inhibitor o-phenanthroline at high temperature suggest that in detergent suspensions of menaquinone-reconstituted reaction centers a redox reaction removing electrons from the quinone acceptor complex competes with charge recombination. Instability of the semiquinone anions is more pronounced when QB is a short-chain menaquinone. From the temperature dependence of P+ decay the activation parameters for the P+QB- recombination and for the competing side oxidation of the reduced menaquinone acceptor have been derived. For both reactions the activation enthalpies and entropies change markedly with menaquinone chain length but counterbalance each other, resulting in activation free energies at ambient temperature independent of the menaquinone tail. When reaction centers are incorporated into phospholipid vesicles containing menaquinone-8 a temperature-dependent, monophasic, o-phenanthroline-sensitive recombination from the P+QB- state is observed, which is consistent with the formation of stable semiquinone anions. This result seems to indicate a proper QB functioning in the two-subunit reaction center isolated from Chlorflexus aurantiacus when the complex is inserted into a lipid bilayer.  相似文献   

19.
Time-resolved photovoltage measurements on destacked photosystem II membranes from spinach with the primary quinone electron acceptor Q(A) either singly or doubly reduced have been performed to monitor the time evolution of the primary radical pair P680(+)Pheo(-). The maximum transient concentration of the primary radical pair is about five times larger and its decay is about seven times slower with doubly reduced compared with singly reduced Q(A). The possible biological significance of these differences is discussed. On the basis of a simple reversible reaction scheme, the measured apparent rate constants and relative amplitudes allow determination of sets of molecular rate constants and energetic parameters for primary reactions in the reaction centers with doubly reduced Q(A) as well as with oxidized or singly reduced Q(A). The standard free energy difference DeltaG degrees between the charge-separated state P680(+)Pheo(-) and the equilibrated excited state (Chl(N)P680)* was found to be similar when Q(A) was oxidized or doubly reduced before the flash (approximately -50 meV). In contrast, single reduction of Q(A) led to a large change in DeltaG degrees (approximately +40 meV), demonstrating the importance of electrostatic interaction between the charge on Q(A) and the primary radical pair, and providing direct evidence that the doubly reduced Q(A) is an electrically neutral species, i.e., is doubly protonated. A comparison of the molecular rate constants shows that the rate of charge recombination is much more sensitive to the change in DeltaG degrees than the rate of primary charge separation.  相似文献   

20.
In reaction centers of Rhodobacter sphaeroides, site-directed mutagenesis has implicated several acidic residues in the delivery of protons to the secondary quinone (Q(B)) during reduction to quinol. In a double mutant (Asp(L210) --> Asn + Asp(M17) --> Asn) that is severely impaired in proton transfer capability over a wide pH range, proton transfer was "rescued" by added weak acids. For low pK(a) acids the total concentration of salt required near neutral pH was high. The ionic strength effect of added salts stimulated the rate of proton-coupled electron transfer at pH < 7, but decreased it at pH > 7.5, indicating an effective isoelectric point between these limits. In this region, a substantial rate enhancement by weak acids was clearly evident. A Br?nsted plot of activity versus pK(a) of the rescuing acids was linear, with a slope of -1, and extrapolated to a diffusion-limited rate at pK(a)(app) approximately 1. However, the maximum rate at saturating concentrations of acid did not correlate with pK(a), indicating that the acid and anion species compete for binding, both with weak affinity. This model predicts that pK(a)(app) corresponds to a true pK(a) = 4-5, similar to that for a carboxylic acid or Q(B)(-), itself. Only rather small, neutral acids were active, indicating a need to access a small internal volume, suggested to be a proton channel to the Q(B) domain. However, the on-rates were near the diffusion limit. The implications for intraprotein proton transfer pathway design are discussed.  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