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.
J Messinger  U Wacker  G Renger 《Biochemistry》1991,30(31):7852-7862
The effect of redox-active amines NH2R (R = OH or NH2) on the period-four oscillation pattern of oxygen evolution has been analyzed in isolated spinach thylakoids as a function of the redox state Si (i = 0, ..., 3) of the water oxidase. The following results were obtained: (a) In dark-adapted samples with a highly populated S1 state, NH2R leads via a dark reaction sequence to the formal redox state "S-1"; (b) the reaction mechanism is different between the NH2R species; NH2OH acts as a one-electron donor, whereas NH2NH2 mainly functions as a two-electron donor, regardless of the interacting redox state Si (i = 0, ..., 3). For NH2NH2, the modified oxygen oscillation patterns strictly depend upon the initial ratio [S0(0)]/[S1(0)] before the addition of the reductant; while due to kinetic reasons, for NH2OH this dependence largely disappears after a short transient period. (c) The existence of the recently postulated formal redox state "S-2" is confirmed not only in the presence of NH2NH2 [Renger, G., Messinger, J., & Hanssum, B. (1990) in Current Research in Photosynthesis (Baltscheffsky, M., Ed.) Vol. 1, pp 845-848, Kluwer, Dordrecht] but also in the presence of NH2OH. (d) Activation energies, EA, of 50 kJ/mol were determined for the NH2R-induced reduction processes that alter the oxygen oscillation pattern from dark-adapted thylakoids. (e) Although marked differences exist between NH2OH and NH2NH2 in terms of the reduction mechanism and efficiency (which is about 20-fold in favor of NH2OH), both NH2R species exhibit the same order of rate constants as a function of the redox state Si in the nonperturbed water oxidase: kNH2R(S0) greater than kNH2R(S1) much less than kNH2R(S2) much greater than kNH2R(S3) The large difference between S2 and S3 in their reactivity toward NH2R is interpreted to indicate that a significant change in the electronic configuration and nuclear geometry occurs during the S2----S3 transition that makes the S3 state much less susceptible to NH2R. The implications of these findings are discussed with special emphasis on the possibility of complexed peroxide formation in redox state S3 postulated previously on the basis of theoretical considerations [Renger, G. (1978) in Photosynthetic Water Oxidation (Metzner, H., Ed.) pp 229-248, Academic Press, London].  相似文献   

2.
Electron paramagnetic resonance (EPR) analyses (g = 2 region) and optical spectrophotometric analyses of P680+ were made of NH2OH-extracted photosystem II (PSII) membranes after various durations of weak-light photoinhibition, in order to identify the sites of damage responsible for the observed kinetic components of the loss of electron transport [Blubaugh, D.J., & Cheniae, G.M. (1990) Biochemistry 29, 5109-5118]. The EPR spectra, recorded in the presence of K3Fe(CN)6, gave evidence for rapid (t1/2 = 2-3 min) and slow (t1/2 = 3-4) losses of formation of the tyrosyl radicals YZ+ and YD+, respectively, and the rapid appearance (t1/2 = 0.8 min) of a 12-G-wide signal, centered at g = 2.004, which persisted at 4 degrees C in subsequent darkness in rather constant abundance (approximately 1/2 spin per PSII). This latter EPR signal is correlated with quenching of the variable chlorophyll a fluorescence yield and is tentatively attributed to a carotenoid (Car) cation. Exogenous reductants (NH2OH greater than or equal to NH2NH2 greater than DPC much greater than Mn2+) were observed to reduce the quencher, but did not reverse other photoinhibition effects. An additional 10-G-wide signal, tentatively attributed to a chlorophyll (Chl) cation, is observed during illumination of photoinhibited membranes and rapidly decays following illumination. The amplitude of formation of the oxidized primary electron donor, P680+, was unaffected throughout 120 min of photoinhibition, indicating no impairment of charge separation from P680, via pheophytin (Pheo), to the first stable electron acceptor, QA. However, a 4-microsecond decay of P680+, reflecting YZ----P680+, was rapidly (t1/2 = 0.8 min) replaced by an 80-140 microsecond decay, presumably reflecting QA-/P680+ back-reaction. Photoinhibition caused no discernible decoupling of the antenna chlorophyll from the reaction center complex. We conclude that the order of susceptibility of PSII components to photodamage when O2 evolution is impaired is Chl/Car greater than YZ greater than YD much greater than P680, Pheo, QA.  相似文献   

3.
M Sivaraja  J Tso  G C Dismukes 《Biochemistry》1989,28(24):9459-9464
EPR studies have revealed that removal of calcium using citric acid from the site in spinach photosystem II which is coupled to the photosynthetic O2-evolving process produces a structural change in the manganese cluster responsible for water oxidation. If done in the dark, this yields a modified S1' oxidation state which can be photooxidized above 250 K to form a structurally altered S2' state, as seen by formation of a "modified" multiline EPR signal. Compared to the "normal" S2 state, this new S2'-state EPR signal has more lines (at least 25) and 25% narrower 55Mn hyperfine splittings, indicative of disruption of the ligands to manganese. The calcium-depleted S2' oxidation state is greatly stabilized compared to the native S2 oxidation state, as seen by a large increase in the lifetime of the S2' EPR signal. Calcium reconstitution results in the reduction of the oxidized tyrosine residue 161YD+ (Em approximately 0.7-0.8 V, NHE) within the reaction center D1 protein in both the S1' and S2' states, as monitored by its EPR signal intensity. We attribute this to reduction by Mn. Thus a possible structural role which calcium plays is to bring YD+ into redox equilibrium with the Mn cluster. Photooxidation of S2' above 250 K produces a higher S state (S3 or S4) having a new EPR signal at g = 2.004 +/- 0.003 and a symmetric line width of 163 +/- 3 G, suggestive of oxidation of an organic donor, possibly an amino acid, in magnetic contact with the Mn cluster. This EPR signal forms in a stoichiometry of 1-2 relative to YD+.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
It is shown that the hydrazine-induced transition of the water-oxidizing complex (WOC) to super-reduced S-states depends on the presence of bicarbonate in the medium so that after a 20 min treatment of isolated spinach thylakoids with 3 mM NH(2)NH(2) at 20 degrees C in the CO(2)/HCO(3)(-)-depleted buffer the S-state populations are: 42% of S(-3), 42% of S(-2), 16% of S(-1) and even formal S(-4) state is reached, while in the presence of 2 mM NaHCO(3), the same treatment produces 30% of S(-3), 38% of S(-2), and 32% of S(-1) and there is no indication of the S(-4) state. Bicarbonate requirement for the oxygen-evolving activity, very low in untreated thylakoids, considerably increases upon the transition of the WOC to the super-reduced S-states, and the requirement becomes low again when the WOC returns back to the normal S-states using pre-illumination. The results are discussed as a possible indication of ligation of bicarbonate to manganese ions within the WOC.  相似文献   

5.
I Vass  S Styring 《Biochemistry》1991,30(3):830-839
The effect of protonation events on the charge equilibrium between tyrosine-D and the water-oxidizing complex in photosystem II has been studied by time-resolved measurements of the EPR signal IIslow at room temperature. The flash-induced oxidation of YD by the water-oxidizing complex in the S2 state is a monophasic process above pH 6.5 and biphasic at lower pHs, showing a slow and a fast phase. The half-time of the slow phase increases from about 1 s at pH 8.0 to about 20 s at pH 5.0, whereas the half-time of the fast phase is pH independent (0.4-1 s). The dark reduction of YD+ was followed by measuring the decay of signal IIslow at room temperature. YD+ decays in a biphasic way on the tens of minutes to hours time scale. The minutes phase is due to the electron transfer to YD+ from the S0 state of the water-oxidizing complex. The half-time of this process increases from about 5 min at pH 8.0 to 40 min at pH 4.5. The hours phase of YD+ has a constant half-time of about 500 min between pH 4.7 and 7.2, which abruptly decreases above pH 7.2 and below pH 4.7. This phase reflects the reduction of YD+ either from the medium or by an unidentified redox component of PSII in those centers that are in the S1 state. The titration curve of the half-times for the oxidation of YD reveals a proton binding with a pK around 7.3-7.5 that retards the electron transfer from YD to the water-oxidizing complex. We propose that this monoprotic event reflects the protonation of an amino acid residue, probably histidine-190 on the D2 protein, to which YD is hydrogen bonded. The titration curves for the oxidation of YD and for the reduction of YD+ show a second proton binding with pK approximately 5.8-6.0 that accelerates the electron transfer from YD to the water-oxidizing complex and retards the process in the opposite direction. This protonation most probably affects the water-oxidizing complex. From the measured kinetic parameters, the lowest limits for the equilibrium constants between the S0YD+ and the S1YD as well as between the S1YD+ and S2YD states were estimated to be 5 and 750-1000, respectively.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

6.
D G Cole  R G Yount 《Biochemistry》1992,31(27):6186-6192
The properties of divalent metal.ADP.vanadate (V(i)) complexes of the 6S extended and 10S folded conformations of gizzard myosin before and after UV irradiation have been studied. The half-lives of both 6S and 10S myosin.MgADP.V(i) complexes in the dark at 0 degrees C are on the order of 2 weeks. Brief irradiation with UV light, however, photomodified the enzyme as suggested by changes in the NH(4+)-, K(+)-, and Ca(2+)-ATPase activities, and destabilized the complexes. The 6S complex, when irradiated, released ADP and V(i) rapidly (t1/2 less than or equal to 1 min) as has been observed in comparable experiments with skeletal myosin subfragment 1 (S1) [Grammer et al. (1988) Biochemistry 27, 8408-8415]. The irradiated 10S complex released approximately 20% of the ADP and V(i) rapidly (t1/2 less than or equal to 1 min), but the remainder stayed trapped, possibly as the vanadyl (VO2+).ADP complex, for much longer times (t1/2 approximately 8 h). The site of photomodification was sought by reducing both photomodified 6S and 10S myosin with NaB3H4. Amino acid composition analyses identified [3H]serine as the only labeled residue(s), suggesting that the hydroxymethyl group of serine had been oxidized to an aldehyde as shown previously for photomodified skeletal myosin S1 [Cremo et al. (1989) J. Biol. Chem. 264, 6608-6611]. The 29-kDa NH2-terminal tryptic peptide from the heavy chain was found to contain essentially all of the [3H]serine. Preparations of 6S and 10S [3H]myosin were digested exhaustively with trypsin. An identical [3H]peptide was purified from each preparation and its sequence determined to be Glu169-Asp-Gln-Ser-Ile-Leu-(Cys)-Thr-Gly-[3H]Ser-Gly-Ala-Gly-Ly s183.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

7.
Transient States of Adenylate Cyclase in Brain Membranes   总被引:3,自引:1,他引:2  
Basal activity of adenylate cyclase from the amygdala of sheep brain and the neostriatum of turkey brain decays in two phases at 37 degrees C. The first phase is rapid (t1/2 = 2.3 +/- 0.3 min) and results in the loss of 60-70% of basal activity. The second phase is slow (t1/2 approximately 100 min) during which time the catalytic units denature irreversibly. The GTP analogue guanosine-5' (beta-gamma imino) triphosphate (p[NH]ppG) prevents the rapid decay by stabilizing the enzyme at its initial level of activity and also reactivates the enzyme to initial levels during or immediately following the early phase, indicating that denaturation of neither the guanylnucleotide units nor the catalytic units causes the rapid decline in basal activity. Activation by p[NH]ppG is rapid at 37 degrees C, but the binding of p[NH]ppG to the guanylnucleotide subunit also occurs at nonactivatory temperatures. This is determined by the protection of catalytic units from thermal or N-ethylmaleimide inactivation after extensive washing. Thus, at 25 degrees C all of the catalytic units can be stabilized by saturating p[NH]ppG concentrations. At 0 degree C, 35% of the catalytic units can be stabilized by saturating p[NH]ppG concentrations within 30 s. The half-saturation constant for the binding of p[NH]ppG at 0 degree C is identical to that derived in an assay at 37 degrees C, or after an incubation of the membranes for 10 min at 45 degrees C, when the process of thermal denaturation is 80% complete (K1/2 approximately 3 +/- 2 microM).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

8.
The nitrogenase-catalyzed H(2) evolution and acetylene-reduction reactions using Ti(III) and dithionite (DT) as reductants were examined and compared under a variety of conditions. Ti(III) is known to make the all-ferrous Fe protein ([Fe(4)S(4)](0)) and lowers the amount of ATP hydrolyzed during nitrogenase catalysis by approximately 2-fold. Here we further investigate this behavior and present results consistent with the Fe protein in the [Fe(4)S(4)](0) redox state transferring two electrons ([Fe(4)S(4)](2+)/[Fe(4)S(4)](0)) per MoFe protein interaction using Ti(III) but transferring only one electron ([Fe(4)S(4)](2+)/[Fe(4)S(4)](1+)) using DT. MoFe protein specific activity was measured as a function of Fe:MoFe protein ratio for both a one- and a two-electron transfer reaction, and nearly identical curves were obtained. However, Fe protein specific activity curves as a function of MoFe:Fe protein ratio showed two distinct reactivity patterns. With DT as reductant, typical MoFe inhibition curves were obtained for operation of the [Fe(4)S(4)](2+)/[Fe(4)S(4)](1+) redox couple, but with Ti(III) as reductant the [Fe(4)S(4)](2+)/[Fe(4)S(4)](0) redox couple was functional and MoFe inhibition was not observed at high MoFe:Fe protein ratios. With Ti(III) as reductant, nitrogenase catalysis produced hyperbolic curves, yielding a V(max) for the Fe protein specific activity of about 3200 nmol of H(2) min(-1) mg(-1) Fe protein, significantly higher than for reactions conducted with DT as reductant. Lag phase experiments (Hageman, R. V., and Burris, R. H. (1978) Proc. Natl. Acad. Sci. U. S. A. 75, 2699-2702) were carried out at MoFe:Fe protein ratios of 100 and 300 using both DT and Ti(III). A lag phase was observed for DT but, with Ti(III) product formation, began immediately and remained linear for over 30 min. Activity measurements using Av-Cp heterologous crosses were examined using both DT and Ti(III) as reductants to compare the reactivity of the [Fe(4)S(4)](2+)/[Fe(4)S(4)](1+) and [Fe(4)S(4)](2+)/[Fe(4)S(4)](0) redox couples and both were inactive. The results are discussed in terms of the Fe protein transferring two electrons per MoFe protein encounter using the [Fe(4)S(4)](2+)/[Fe(4)S(4)](0) redox couple with Ti(III) as reductant.  相似文献   

9.
Fang CH  Chiang KA  Hung CH  Chang K  Ke SC  Chu HA 《Biochemistry》2005,44(28):9758-9765
Ammonia is an inhibitor of water oxidation and a structural analogue for substrate water, making it a valuable probe for the structural properties of the possible substrate-binding site on the oxygen-evolving complex (OEC) in photosystem II (PSII). By using the NH(3)-induced upshift of the 1365 cm(-)(1) IR mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum and the NH(3)-modified S(2) state EPR signals of PSII as spectral probes, we found that ethylene glycol has clear effects on the binding properties of the NH(3)-specific site on the OEC. Our results show that in PSII samples containing 30% (v/v) ethylene glycol, the affinity of the NH(3)-specific binding site on the OEC is estimated to be more than 10 times lower than that in PSII samples containing 0.4 M sucrose. In addition, our results show that the NH(3)-induced upshift of the 1365 cm(-)(1) IR mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum is dependent on the concentration of ethylene glycol, but not dependent on the concentration of sucrose (up to 1.5 M) or methanol (up to 5.4 M). By comparing the concentration dependence of sucrose and ethylene glycol on NH(3)-induced spectral change and also by comparing the sucrose and ethylene glycol data at similar concentrations ( approximately 1 M), we conclude that ethylene glycol has a clear effect on the NH(3)-induced spectral changes. Furthermore, our results also show that ethylene glycol alters the steric requirement of the amine effect on the upshift of the 1365 cm(-)(1) mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum. In PSII samples containing 30% (v/v) ethylene glycol, only NH(3), not other bulkier amines (e.g., Tris, AEPD, and CH(3)NH(2)), has a clear effect on the upshift of the 1365 cm(-)(1) mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum; in contrast, in PSII samples containing 0.4 M sucrose, both NH(3) and CH(3)NH(2) have a clear effect. On the basis of the results mentioned above, we propose that ethylene glycol acts directly or indirectly to decrease the affinity or limit the accessibility of NH(3) and CH(3)NH(2) to the NH(3)-specific binding site on the OEC in PSII. Finally, we also applied the same approach to test whether methanol is able to compete with ammonia on its binding site on the OEC. We found that 4% (v/v) methanol does not have any significant effect on the NH(3)-induced upshift of the 1365 cm(-)(1) mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum and the NH(3)-modified S(2) state g = 2 multiline EPR signal. Our results suggest that methanol is unable to compete with NH(3) upon binding to the Mn site of the OEC that gives rise to the altered S(2) state g = 2 multiline EPR signal.  相似文献   

10.
Havelius KG  Styring S 《Biochemistry》2007,46(26):7865-7874
The photosystem II (PSII) reaction center contains two redox active tyrosines, YZ and YD, situated on the D1 and D2 proteins, respectively. By illumination at 5 K, oxidation of YZ in oxygen-evolving PSII can be observed as induction of the Split S1 EPR signal from YZ* in magnetic interaction with the CaMn4 cluster, whereas oxidation of YD can be observed as the formation of the free radical EPR signal from YD*. We have followed the light induced induction at 5 K of the Split S1 signal between pH 4-8.5. The formation of the signal, that is, the oxidation of YZ, is pH independent and efficient between pH 5.5 and 8.5. At low pH, the split signal formation decreases with pKa approximately 4.7-4.9. In samples with chemically pre-reduced YD, the pH dependent competition between YZ and YD was studied. Only YZ was oxidized below pH 7.2, but at pH above 7.2, the oxidation of YD became possible, and the formation of the Split S1 signal diminished. The onset of YD oxidation occurred with pKa approximately 8.0, while the Split S1 signal decreased with pKa approximately 7.9 demonstrating that the two tyrosines compete in this pH interval. The results reflect the formation and breaking of hydrogen bonds between YZ and D1-His190 (HisZ) and YD and D2-His190 (HisD), respectively. The oxidation of respective tyrosine at 5 K demands that the hydrogen bond is well-defined; otherwise, the low-temperature oxidation is not possible. The results are discussed in the framework of recent literature data and with respect to the different oxidation kinetics of YZ and YD.  相似文献   

11.
D C May  E M Ross 《Biochemistry》1988,27(13):4888-4893
When reconstituted phospholipid vesicles that contain purified beta-adrenergic receptors and the GTP-binding regulatory protein Gs were preincubated with agonist before the addition of guanosine 5'-O-(3-thiotriphosphate) (GTP gamma S), the typical receptor-stimulated GTP gamma S binding reaction was preceded by an even more rapid burst of GTP gamma S binding. This burst was studied in detail at 0 degree C. The rate of the burst was second order in nucleotide and Gs [k assoc approximately 2 X 10(7) (M.min)-1], consistent with diffusion-controlled binding. The magnitude of the burst was always less than the number of receptors present and was roughly linear with receptor number when similarly prepared vesicles were compared. There was no obvious quantitative correlation between the burst and the amount of Gs. The species that gave rise to the burst formed with t1/2 approximately 15 min at 0 degree C in the presence of agonist and decayed by approximately 3 min upon addition of antagonist or detergent. Formation and decay of this species was much faster at at 30 degrees C. The data suggest that a complex of agonist, receptor, and Gs that is primed for the rapid binding of guanine nucleotide can form and be analyzed in reconstituted vesicles.  相似文献   

12.
The purpose of this study was to determine the influence of endurance-type exercise training on alterations of the ammonia content of blood in exercising humans. Seven females and four males trained 6 days/wk for 7 wk alternating days of continuous cycling (40 min) and interval running (five 5-min bouts). The NH3 content of blood was determined before and during cycle ergometer (CE) exercise (4 min) at power outputs (PO) of 119, 172, and 241 W pretraining and of 163, 230, and 271 W posttraining. These PO for each occasion represent relative work loads of approximately 65, 90, and 115% of peak CE maximum O2 uptake (PCE VO2), respectively. Training increased (P less than 0.05) PCE VO2 approximately 32% (2.72 +/- 0.25 to 3.56 +/- 0.29 l/min or 38.5 +/- 1.9 to 51.2 +/- 2.3 ml X kg-1 X min-1). Both pre- and posttraining the NH3 content of blood increased (P less than 0.05) with increasing intensity of exercise. Training did not influence the measure of these responses during exercise at the same relative intensity. During exercise at the same absolute PO, approximately 168 or 235 W, however, increases in blood NH3 were less (P less than 0.05) after training. The results indicate that the magnitude of increase in blood NH3 during exercise is determined by the energy requirement of the absolute work load, relative to an individual's aerobic power.  相似文献   

13.
Vavilin DV  Vermaas WF 《Biochemistry》2000,39(48):14831-14838
The lumenal CD-loop region of the D2 protein of photosystem II contains residues that interact with the primary electron donor P680 and the redox active tyrosyl residue Y(D). Photosystem II properties were studied in a number of photoautotrophic mutants of Synechocystis sp. PCC 6803, most of which carried combinatorial mutations in residues 164-170, 179-186, or 187-194 of the D2 protein. To facilitate characterization of photosystem II properties in the mutants, the CD-loop mutations were introduced into a photosystem I-less background. According to variable fluorescence decay measurements in DCMU-treated cells, charge recombination of Q(A)(-) with the donor side was faster in the majority of mutants (t(1/2) = 45-140 ms) than in the control (t(1/2) = 180 ms). However, in one mutant (named C7-3), the decay of Q(A)(-) was 2 times slower than in the control (t(1/2) = 360 ms). The decay half-time of each mutant correlated with the yield of the Q-band of thermoluminescence (TL) emitted due to S(2)Q(A)(-) charge recombination. The C7-3 mutant had the highest TL intensity, whereas no Q-band was detected in the mutants with fast Q(A)(-) decay (t(1/2) = 45-50 ms). The correlated changes in the rate of recombination and in TL yield in these strains suggest the existence of a nonradiative pathway of charge recombination between Q(A)(-) and the donor side. This may involve direct electron transfer from Q(A)(-) to P680(+) in a way not leading to formation of excited chlorophyll. Many mutations in the CD-loop appear to increase the equilibrium P680(+) concentration during the lifetime of the S(2)Q(A)(-) state, for example, by making the midpoint potential of the P680(+)/P680 redox couple more negative. The nonradiative charge recombination pathway involves a low activation energy and is less temperature-dependent than the formation of excited P680 that leads to TL emission. Therefore, during the TL measurements in these mutants, the S(2)Q(A)(-) state can recombine nonradiatively before temperatures are reached at which radiative charge recombination becomes feasible. The results presented here highlight the presence of two charge recombination pathways and the importance of the CD-loop of the D2 protein in determination of the energy gap between the P680(+)S(1) and P680S(2) states.  相似文献   

14.
Chu HA  Feng YW  Wang CM  Chiang KA  Ke SC 《Biochemistry》2004,43(34):10877-10885
Light-induced Fourier transform infrared difference spectroscopy has been applied to studies of ammonia effects on the oxygen-evolving complex (OEC) of photosystem II (PSII). We found that NH(3) induced characteristic spectral changes in the region of the symmetric carboxylate stretching modes (1450-1300 cm(-1)) of the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectra of PSII. The S(2) state carboxylate mode at 1365 cm(-1) in the S(2)Q(A)(-)/S(1)Q(A) spectrum of the controlled samples was very likely upshifted to 1379 cm(-1) in that of NH(3)-treated samples; however, the frequency of the corresponding S(1) carboxylate mode at 1402 cm(-1) in the same spectrum was not significantly affected. These two carboxylate modes have been assigned to a Mn-ligating carboxylate whose coordination mode changes from bridging or chelating to unidentate ligation during the S(1) to S(2) transition [Noguchi, T., Ono, T., and Inoue, Y. (1995) Biochim. Biophys. Acta 1228, 189-200; Kimura, Y., and Ono, T.-A. (2001) Biochemistry 40, 14061-14068]. Therefore, our results show that NH(3) induced significant structural changes of the OEC in the S(2) state. In addition, our results also indicated that the NH(3)-induced spectral changes of the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are dependent on the temperature of the FTIR measurement. Among the temperatures we measured, the strongest effect was seen at 250 K, a lesser effect was seen at 225 K, and little or no effect was seen at 200 K. Furthermore, our results also showed that the NH(3) effects on the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are dependent on the concentrations of NH(4)Cl. The NH(3)-induced upshift of the 1365 cm(-1) mode is apparent at 5 mM NH(4)Cl and is completely saturated at 100 mM NH(4)Cl concentration. Finally, we found that CH(3)NH(2) has a small but clear effect on the spectral change of the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectrum of PSII. The effects of amines on the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectra (NH(3) > CH(3)NH(2) > AEPD and Tris) are inverse proportional to their size (Tris approximately AEPD > CH(3)NH(2) > NH(3)). Therefore, our results showed that the effects of amines on the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII are sterically selective for small amines. On the basis of the correlations between the conditions (dependences on the excitation temperature and NH(3) concentration and the steric requirement for the amine effects) that give rise to the NH(3)-induced upshift of the 1365 cm(-)(1) mode in the S(2)Q(A)(-)/S(1)Q(A) spectrum of PSII and the conditions that give rise to the altered S(2) state multiline EPR signal, we propose that the NH(3)-induced upshift of the 1365 cm(-1) mode is caused by the binding of NH(3) to the site on the Mn cluster that gives rise to the altered S(2) state multiline EPR signal. In addition, we found no significant NH(3)-induced change in the S(2)Q(A)(-)/S(1)Q(A) FTIR difference spectrum at 200 K. Under this condition, the OEC gives rise to the NH(3)-stabilized g = 4.1 EPR signal and a suppressed g = 2 multiline EPR signal. Our results suggest that the structural difference of the OEC between the normal g = 2 multiline form and the NH(3)-stabilized g = 4.1 form is small.  相似文献   

15.
The interaction of recA protein with single-stranded (ss) phi X174 DNA has been examined by means of a nuclease protection assay. The stoichiometry of protection was found to be 1 recA monomer/approximately 4 nucleotides of ssDNA both in the absence of a nucleotide cofactor and in the presence of ATP. In contrast, in the presence of adenosine 5'-O-(thiotriphosphate) (ATP gamma S) the stoichiometry was 1 recA monomer/approximately 8 nucleotides. No protection was seen with ADP. In the absence of a nucleotide cofactor, the binding of recA protein to ssDNA was quite stable as judged by equilibration with a challenge DNA (t1/2 approximately 30 min). Addition of ATP stimulated this transfer (t1/2 approximately 3 min) as did ADP (t1/2 approximately 0.2 min). ATP gamma S greatly reduced the rate of equilibration (t1/2 greater than 12 h). Direct visualization of recA X ssDNA complexes at subsaturating recA protein concentrations using electron microscopy revealed individual ssDNA molecules partially covered with recA protein which were converted to highly condensed networks upon addition of ATP gamma S. These results have led to a general model for the interaction of recA protein with ssDNA.  相似文献   

16.
As rice can use both nitrate (NO3-) and ammonium (NH4+), we have tested the hypothesis that the shift in the pattern of cultivars grown in Jiangsu Province reflects the ability of the plants to exploit NO3- as a nitrogen (N) source. Four rice cultivars were grown in solution culture for comparison of their growth on NO3- and NH4+ nitrogen sources. All four types of rice,Xian You 63 (XY63), Yang Dao 6 (YD), Nong Keng 57 (NK) and Si You 917 (SY917), grew well and produced similar amounts of shoot biomass with 1 mmol/L NH4+ as the only N source.However, the roots of NK were significantly smaller in comparison with the other cultivars. When supplied with 1 mmol/L NO3-, YD produced the greatest biomass; while NK achieved the lowest growth among the four cultivars. Electrophysiological measurements on root rhizodermal cells showed that the NO3--elicited changes in membrane potential (ΔEm) of these four rice cultivars were significantly different when exposed to low external NO3- (<1 mmol/L); while they were very similar at high external NO3- (10 mmol/L). The root cell membrane potentials of YD and XY63 were more responsive to low external NO3- than those of NK and SY917. The ΔEm values for YD and XY63 rhizodermal cells were almost the same at both 0.1 mmol/L and 1 mmol/L NO3-;while for the NK and SY917 the values became larger as the external NO3- increased. For YD cultivar, ΔEm was measured over a range of NO3- concentrations and a Michaelis-Menten fit to the data gave a Km value of 0.17 mmol/L. Net NO3- uptake depletion kinetics were also compared and for some cultivars (YD and XY63) a single-phase uptake system with first order kinetics best fitted the data; while other cultivars (ND and SY917) showed a better fit to two uptake systems. These uptake systems had two affinity ranges: one had a similar Km in all the cultivars (0.2 mmol/L); the other much higher affinity system (0.03 mmol/L) was only present in NK and SY917. The expression pattern of twelve different NO3- transporter genes was tested using specific primers, but only OsNRT1.1 and OsNRT2. 1 expression could be detected showing significant differences between the four rice cultivars. The results from both the physiological and molecular experiments do provide some support for the hypothesis that the more popular rice cultivars grown in Jiangsu Province may be better at using NO3- as an N source.  相似文献   

17.
Detailed comparative studies of flash induced oxygen evolution patterns in thylakoids from the thermophilic cyanobacterium Synechococcus elongatus (S. elongatus; also referred to as Thermosynechococcus elongatus) and from spinach led to the following results: (i) the miss parameter alpha of S. elongatus thylakoids exhibits a pronounced temperature dependence with a minimum of 7% at 25 degrees C and values of 17 and 10% at 3 and 35 degrees C, respectively, while for spinach thylakoids alpha decreases continuously from 18% at 35 degrees C down to 8% at 3 degrees C; (ii) at all temperatures, the double hit probability beta exceeds in S. elongatus the corresponding values of spinach by an increment Delta beta of about 3%; (iii) at 20 degrees C the slow relaxation of the oxidation states S(2) and S(3) is about 15 and 30 times, respectively, slower in S. elongatus than in spinach, while the reduction of these S states by tyrosine Y(D) is 2-3 times faster; (iv) the reaction S(0)Y(D)(ox) --> S(1)Y(D) is slower by a factor of 4 in S. elongatus as compared to spinach; and (v) the activation energies of S state dark relaxations in S. elongatus are all within a factor of 1.5 as compared to the previously reported values from spinach thylakoids [Vass, I., Deak, Z., and Hideg, E. (1990) Biochim. Biophys. Acta 1017, 63-69; Messinger, J., Schr?der, W. P., and Renger, G. (1993) Biochemistry 32, 7658-7668], but the difference between the activation energies of the slow S(2) and S(3) decays is significantly larger in S. elongatus than in spinach. These results are discussed in terms of differences between cyanobacteria and higher plants on the acceptor side of PSII and a shift of the redox potential of the couple Y(D)/Y(D)(ox). The obtained data are also suitable to address questions about effects of the redox state of Y(D) on the miss probability and the possibility of an S state dependent miss parameter.  相似文献   

18.
The GTPase activity of the stimulatory guanine nucleotide-binding regulatory protein (Gs) of hormone-sensitive adenylate cyclase was investigated using purified rabbit hepatic Gs and either [alpha-32P]- or [gamma-32P] GTP as substrate. The binding of [35S]guanosine 5'-O-(thiotriphosphate) (GTP gamma S) was used to quantitate the total concentration of Gs. 1) GTPase activity was a saturable function of the concentration of GTP, with Km = 0.3 microM. MgCl2 monotonically increased the activity. The maximum observed turnover number was about 1.5 min-1. 2) During steady-state hydrolysis, 20-40% of total Gs could be trapped as a Gs-GDP complex and 1-2% could be trapped as Gs-GTP. The hydrolysis of Gs-GTP to Gs-GDP occurred with t 1/2 less than or equal to 5 s at 30 degrees C and t 1/2 approximately 1 min at 0 degrees C. Hydrolysis of Gs-GTP was inhibited by 1.0 mM EDTA in the absence of added Mg2+. 3) The rate of formation of Gs-GDP and the initial GTPase rate varied in parallel as functions of the concentrations of either GTP or MgCl2 (above 0.1 mM Mg2+). The ratio of the rate of accumulation of Gs-GDP to the GTPase rate was constant at 0.3-0.4. 4) The rate of dissociation of assayable Gs-GDP was biphasic. The initial phase accounted for 60-80% of total assayable Gs-GDP and was characterized by a t 1/2 of about 1 min. 5) Lubrol 12A9 potently inhibited the GTPase reaction and the dissociation of Gs-GDP in parallel, and inhibition of product release may account for the inhibition of steady-state hydrolysis. 6) The beta and gamma subunits of Gs markedly inhibited the dissociation of GDP from Gs in contrast to their ability to stimulate the dissociation of GTP gamma S. 7) GDP, GTP gamma S, and guanyl-5'-yl imidodiphosphate (Gpp(NH)p) competitively inhibited the accumulation of Gs-GDP. GTP gamma S and Gpp(NH)p inhibited the GTPase reaction noncompetitively, GDP displayed mixed inhibition, and Pi did not inhibit. These data are interpretable in terms of the coexistence of two specific mechanistic pathways for the overall GTPase reaction.  相似文献   

19.
We investigated the effects of the cellular redox state on nerve growth factor (NGF)-induced neuronal differentiation and its signaling pathways. Treatment of PC12 cells with buthionine sulfoximine (BSO) reduced the levels of GSH, a major cellular reductant, and enhanced NGF-induced neuronal differentiation, activation of AP-1 and the NGF receptor tyrosine kinase, TrkA. Conversely, incubation of the cells with a reductant, N-acetyl-L-cysteine (NAC), inhibited NGF-induced neuronal differentiation and AP-1 activation. Consistent with the suppression, NAC inhibited NGF-induced activation of TrkA, formation of receptor complexes comprising TrkA, Shc, Grb2, and Sos, and activation of phospholipase Cgamma and phosphatidylinositol 3-kinase. Biochemical analysis suggested that the cellular redox state regulates TrkA activity through modulation of protein tyrosine phosphatases (PTPs). Thus, cellular redox state regulates signaling pathway of NGF through PTPs, and then modulates neuronal differentiation.  相似文献   

20.
A spliced leader contributes the mature 5'ends of many mRNAs in trans-splicing organisms. Trans-spliced metazoan mRNAs acquire an m3(2,2,7)GpppN cap from the added spliced leader exon. The presence of these caps, along with the typical m7GpppN cap on non-trans-spliced mRNAs, requires that cellular mRNA cap-binding proteins and mRNA metabolism deal with different cap structures. We have developed and used an in vitro system to examine mRNA degradation and decapping activities in nematode embryo extracts. The predominant pathway of mRNA decay is a 3' to 5' pathway with exoribonuclease degradation of the RNA followed by hydrolysis of resulting mRNA cap by a scavenger (DcpS-like) decapping activity. Direct decapping of mRNA by a Dcp1/Dcp2-like activity does occur, but is approximately 15-fold less active than the 3' to 5' pathway. The DcpS-like activity in nematode embryo extracts hydrolyzes both m7GpppG and m3(2,2,7)GpppG dinucleoside triphosphates. The Dcp1/Dcp2-like activity in extracts also hydrolyzes these two cap structures at the 5' ends of RNAs. Interestingly, recombinant nematode DcpS differs from its human ortholog in its substrate length requirement and in its capacity to hydrolyze m3(2,2,7)GpppG.  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