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 62 毫秒
1.
The kinetics of the metarhodopsin I-II reaction have been measured over a wide range of temperatures (1-37C ) and pH values (4.5-8) with suspensions containing fragments of bovine rod outer segments. It was found that for all conditions the occurrence of metarhodopsin II could be described by two independent first-order processes. The fast component: slow component amplitude ratio depends upon pH and temperature. The kinetics of the lumi-metarhodopsin I reaction show the same pH dependence for the fast component: slow component amplitude ratio as the one observed for the metarhodopsin II signals. All the results observed could be described with the assumption that rhodopsin itself exists in two conformational states before bleaching which are in a pH and temperature-dependent equilibrium. This hypothesis is confirmed by its ability to explain some apparently anomalous observations in the literature.  相似文献   

2.
Spectrophotometric measurements of metarhodopsin II appearance are made on five different kinds of rhodopsin preparations. Although the preparations differ greatly in their rhodopsin: phospholipid ratio, the meta II kinetics in all of them are strikingly similar in certain respects. Meta II appearance kinetics in all of the preparations are best described by two and only two exponentials. The ratio of these two rates is always about 5. The fast fraction: slow fraction ratio depends upon temperature. These fractions are reversibly convertible in the dark, and are interconverted on a time-scale which is long compared to the meta II appearance rate. It is shown that the kinetics of the earlier step in the bleaching sequence, viz., lumi- meta I, is also described by double exponentials. Again the ratio of rates is ca. 5 and the fast-slow fractions correspond to those found in the meta I meta II step. It is proposed that these facts support an hypothesis for the existence of two conformeric states of rhodopsin which are in thermal equilibrium. Thermodynamic parameters associated with this proposed equilibrium are presented.  相似文献   

3.
The kinetics of the metarhodopsin (meta) I → metarhodopsin II reaction have been studied by flash photolysis in two different types of preparations of bovine rhodopsin: (i) digitonin-solubilized rod outer segment (ROS) membranes with a molar ratio of phospholipid to rhodopsin of approximately 90, and (ii) digitonin-solubilized phospholipid-free rhodopsin with a molar ratio of phospholipid to rhodopsin of less than 0.2. At 20 °C the kinetics in both preparations are multiexponential, but four terms are required to fit the data with the solubilized membranes, whereas only two are required with the phospholipid-free preparation. Thus, phospholipid removal simplifies the kinetics of the meta I → meta II reaction, but the resulting preparation still does not show first-order kinetics. The ratio of the time constants of these two components with detergent-solubilized phospholipid-free rhodopsin was nearly equal to the values found with ROS particles, rhodopsin-phospholipid recombinants and intact rabbit eyes. This suggests a common origin for these two components in all these preparations and appears to exclude heterogeneity in bound phospholipid as the basis of these two-component kinetics.  相似文献   

4.
We have previously described [H, Kühn et al. (1981) Proc. Natl Acad. Sci. USA, 78, 6873-6877] a light-induced scattering change ('binding signal') associated with a stoichiometric binding between photoexcited rhodopsin and a peripheral membrane protein, the GTP-binding protein, in bovine rod outer segment suspensions. We have attempted here to identify the rhodopsin intermediate R* which is responsible for this interaction, by studying its dependence on pH, temperature and ionic strength. The results strongly suggest that the active state is metarhodopsin II (M II). 1. The initial phase of the binding signal is slightly slower than the formation of metarhodopsin II (2-37 degrees C, pH 5.5-9). 2. The kinetics of the decay of the active rhodopsin state are similar to those of the metarhodopsin II leads to metarhodopsin III transition (37 degrees C, pH 7.3). 3. All conditions which lead to light-induced binding of the GTP-binding protein to R* also lead to the formation of M II. At 2 degrees C, pH 8.3, in particular where no M II is formed in the absence of GTP-binding protein, binding signals and light-induced attachment of the GTP-binding protein to the membrane are still observed. Consistently, addition of GTP-binding protein to a suspension of extracted membranes bleached at 2 degrees C (pH 8.3) shifts the metarhodopsin I in equilibrium metarhodopsin II equilibrium towards metarhodopsin II. The shift is reversed by GTP, which dissociates the rhodopsin--GTP-binding protein complex. 4. At low ionic strength, where the GTP-binding protein is soluble in the dark (instead of being associated to the membrane as in the above experiments) M II still induces the binding whereas M I does not, indicating a much lower affinity of the GTP-binding protein for MI.  相似文献   

5.
The fluorescence decay of chlorophyll in spinach thylakoids was measured as a function of the degree of closure of Photosystem II reaction centers, which was set for the flowed sample by varying either the preillumination by actinic light or the exposure of the sample to the exciting pulsed laser light. Three exponential kinetic components originating in Photosystem II were fitted to the decays; a fourth component arising from Photosystem I was determined to be negligible at the emission wavelength of 685 nm at which the fluorescence decays were measured. Both the lifetimes and the amplitudes of the components vary with reaction center closure. A fast (170–330 ps) component reflects the trapping kinetics of open Photosystem II reaction centers capable of reducing the plastoquinone pool; its amplitude decreases gradually with trap closure, which is incompatible with the concept of photosynthetic unit connectivity where excitation energy which encounters a closed trap can find a different, possibly open one. For a connected system, the amplitude of the fast fluorescence component is expected to remain constant. The slow component (1.7–3.0 ns) is virtually absent when the reaction centers are open, and its growth is attributable to the appearance of closed centers. The middle component (0.4–1.7 ns) with approximately constant amplitude may originate from centers that are not functionally linked to the plastoquinone pool. To explain the continuous increase in the lifetimes of all three components upon reaction center closure, we propose that the transmembrane electric field generated by photosynthetic turnover modulates the trapping kinetics in Photosystem II and thereby affects the excited state lifetime in the antenna in the trap-limited case.Abbreviations DCMU 3-(3,4-dichlorophenyl)-1,1-dimethylurea - HEPES 4-(2-hydroxyethyl)-1-piperazineethane sulfonic acid - PQ plastoquinone - PSI and PSII Photosystem I and II - QA and QB primary and secondary quinone acceptor of PSII  相似文献   

6.
Suspensions of vesicles from cattle rod disc membranes are flash illuminated at varying pH and temperature; the light induced proton uptake and spectral change associated with the formation of metarhodopsin II are measured. Metarhodopsin II is shown to be diprotonated at 3°C, and to exist in at least two forms at higher temperature: the diprotonated form observed at 3°C(mainly at acid pH), and an unprotonated form (mainly at alkaline pH), which is found to be in temperature dependent equilibrium with metarhodopsin I and/or protonated metarhodopsin II.  相似文献   

7.
J H Parkes  P A Liebman 《Biochemistry》1984,23(21):5054-5061
The kinetics of the relaxation of bleached bovine rod disk membrane suspensions from metarhodopsin I into the equilibrium between metarhodopsins I and II were determined at pHs between 5.9 and 8.1 and at temperatures between -1 and 15 degrees C. From these data, thermodynamic equations were generated by two-way linear regression that simultaneously describe the functional dependence on pH and temperature of the pseudo-first-order and true forward rate constants, the reverse and observed rate constants, and the equilibrium constant. Using these equations, we obtained the thermodynamic parameters and the apparent net proton uptake for the transitions from metarhodopsin I to metarhodopsin II and from metarhodopsin I to the activated intermediate. The reversibility of this equilibrium and the effect of aging of the preparation on the measured rate constants were investigated.  相似文献   

8.
A. Melis  A.P.G.M. Thielen 《BBA》1980,589(2):275-286
In the present study we used three types of Nicotiana tabacum, cv John William's Broad Leaf (the wild type and two mutants, the yellow-green Su/su and the yellow Su/su var. Aurea) in order to correlate functional properties of Photosystem II and Photosystem I with the structural organization of their chloroplasts. The effective absorption cross-section of Photosystem II and Photosystem I centers was measured by means of the rate constant of their photoconversion under light-limiting conditions. In agreement with earlier results (Okabe, K., Schmid, G.H. and Straub, J. (1977) Plant Physiol. 60, 150–156) the photosynthetic unit size for both System II and System I in the two mutants was considerably smaller as compared to the wild type. We observed biphasic kinetics in the photoconversion of System II in all three types of N. tabacum. However, the photoconversion of System I occurred with monophasic and exponential kinetics. Under our experimental conditions, the effective cross-section of Photosystem I was comparable to that of the fast System II component (α centers). The relative amplitude of the slow System II component (β centers) varied between 30% in the wild type to 70% in the Su/su var. Aurea mutant. The increased fraction of β centers is correlated with the decreased fraction of appressed photosynthetic membranes in the chloroplasts of the two mutants. As a working hypothesis, it is suggested that β centers are located on photosynthetic membranes directly exposed to the stroma medium.  相似文献   

9.
The reconstitution of Neurospora apotyrosinase was studied in the presence of Cu(I) or Cu(II) ions. The kinetics and the mechanism of reactivation were found to differ markedly for the two metal ions. Thus the reconstitution with Cu(I) was found to be very fast and complete following an all or none process; in contrast the reaction with Cu(II) proved to be rather slow and incomplete with the two Cu(II) ions binding with different rate constants.  相似文献   

10.
The fast fluorescence decay kinetics of two photosynthetic mutants of corn (Zea mays) have been compared with those of normal corn. The fluorescence of normal corn can be resolved into three exponential decay components of lifetime 900–1500 ps (slow), 300–500 ps (middle) and 50–120 ps (fast), the yields of which are affected by light intensity and Mg2+ levels. The Photosystem II-(PS II)-defective mutant hcf-3 has similar decay lifetimes (approx. 1200, 450 and 100 ps) but is not affected by light intensity, reflecting the absence of PS II charge recombination. However, yields do respond to Mg2+ in a fashion typical of normal corn, which may be correlated with the presence of normal levels of light-harvesting chlorophyll a + b complex (LHCP). The PS I mutant hcf-50 also shows three-component decay kinetics. In conjunction with the results on the LHCP-deficient mutant of barley presented in a recent paper (Karukstis, K.K. and Sauer, K. (1984) Biochim. Biophys. Acta 766, 148–155), these data suggest that the slow component of normal chloroplasts is kinetically controlled by the decay processes of the LHCP and that the energy comes from one of two sources: (a) charge recombination in the reaction centre or (b) energy transferred within or between LHCP units only. The fast component appears to originate from both PS I and PS II. The complex response of the middle component to cations and light intensity, and its presence in all of the mutants, suggests that it also may have multiple origins.  相似文献   

11.
《BBA》1985,807(2):155-167
The time-resolved fluorescence emission and excitation spectra of Chlorella vulgaris cells have been measured by single-photon timing with picosecond resolution. In a three-exponential analysis the time-resolved excitation spectra recorded at 685 and 706 nm emission wavelength with closed PS II reaction centers show large variations of the preexponential factors of the different decay components as a function of wavelength. At λem = 685 nm the major contribution to the fluorescence decay originates from two components with life-times of 2.1–2.4 and 1.2–1.3 ns. A short-lived component with life-times of 0.1–0.16 ns of relatively small amplitude is also found. When the emission is detected at 706 nm, the short-lived component with a life-time of less than 0.1 ns predominates. Time-resolved emission spectra using λexc = 630 or λexc = 652 nm show a spectral peak of the two longer-lived components at about 680–685 nm, whereas the fast component is red-shifted as compared to the others and shows a maximum at about 690 nm. The emission spectrum observed upon excitation at 696 nm with closed PS II reaction centers shows a large increase in the amplitude of the fast component with a lifetime of 80–100 ps as compared to that at 630 nm excitation. At almost open Photosystem II (PS II) reaction centers (F0), the life-time of the fast component decreased from 150–160 ps at 682 nm to less than 100 ps at 720 nm emission wavelength. We conclude that at least two pigment pools contribute to the fast component. One is attributed to PS II and the other to Photosystem I (PS I). They have life-times of approx. 180 ps and 80 ps, respectively. The 80 ps (PS I) contribution has a spectral maximum slightly below 700 nm, whereas the 180 ps (PS II) spectrum peaks at 680–685 nm. The spectra of the middle decay component τm and its sensitivity to inhibitors of PS II suggest that this component is not preferentially related to LHC II but arises mainly from Chl a pigments probably associated with a second type of PS II centers. The amplitudes of the fast (180 ps, PS II) component and the long-lived decay show an opposite dependence on the state of the PS II centers and confirm our earlier conclusion that the contribution of PS II to the fast component probably disappears at the Fmax state (Haehnel W., Holzwarth, A.R. and Wendler, J. (1983) Photochem. Photobiol. 34, 435–443). Our data are discussed in terms of α,β-heterogeneity in PS II centers.  相似文献   

12.
Using the whole cell voltage-clamp technique and a Cl free and Na free Ba methane sulfonate solution, stage V and VI Xenopus oocytes demonstrated a Ba current (endogenous component) with a peak amplitude average of 6 nA (6 ± 2 nA). When oocytes were injected with crustacean skeletal muscle mRNA, an additional component of IBa could be detected (exogenous IBa). The latter current could be distinguished from the native one by several electrophysiological means: a peak amplitude average of 90 nA (90 ± 4 nA), activation potential threshold, steady state inactivation properties and sensitivity to Ca blockers. As shown by Jdaïâa and Guilbault in crustacean skeletal muscle fibres, exogenous IBa could be divided into two components: a “fast component” and a “slow component” probably passing through two types of Ca channels (fast and slow) since the peak Ba current voltage relationship was biphasic and the fast component of exogenous IBa was less sensitive than the slow to nifedipine. The features of the newly synthesized channels incorporated in the Xenopus oocyte membrane suggest that they may be associated with fast and slow channels, previously described in many preparations, particularly in crustacean skeletal muscle fibres.  相似文献   

13.
Summary— Using the whole cell voltage-clamp technique and a Cl free and Na free Ba methane sulfonate solution, stage V and VI Xenopus oocytes demonstrated a Ba current (endogenous component) with a peak amplitude average of 6 nA (6 ± 2 nA). When oocytes were injected with crustacean skeletal muscle mRNA, an additional component of IBa could be detected (exogenous IBa). The latter current could be distinguished from the native one by several electrophysiological means: a peak amplitude average of 90 nA (90 ± 4 nA), activation potential threshold, steady state inactivation properties and sensitivity to Ca blockers. As shown by Jdaïâa and Guilbault in crustacean skeletal muscle fibres, exogenous IBa could be divided into two components: a “fast component” and a “slow component” probably passing through two types of Ca channels (fast and slow) since the peak Ba current voltage relationship was biphasic and the fast component of exogenous IBa was less sensitive than the slow to nifedipine. The features of the newly synthesized channels incorporated in the Xenopus oocyte membrane suggest that they may be associated with fast and slow channels, previously described in many preparations, particularly in crustacean skeletal muscle fibres.  相似文献   

14.
The absorbance change at 505 nm was used to monitor the kinetics of violaxanthin deepoxidation in isolated pea (Pisum sativum) chloroplasts under dark conditions at various pH values. In long-term measurements (65 min) a fast and a slow exponential component of the 505-nm absorbance change could be resolved. The fast rate constant was up to 10 times higher than the slow rate constant. The asymptote value of the fast kinetic component was twice that of the slow component. The pH dependency of the parameters of the fast kinetic component was analyzed from pH 5.2 to pH 7.0. It was found that the asymptote value dropped slightly with increasing pH. The rate constant was zero at pH values greater than 6.3 and showed maximum values at pH values less than 5.8. Hill plot analysis revealed a strong positive cooperativity for the pH dependency of the fast rate constant (Hill coefficient nH = 5.3). The results are discussed with respect to published activity curves of violaxanthin deepoxidation.  相似文献   

15.
New experimental data and a quantitative theoretical treatment are given for the kinetics of the thermal folding transition of ribonuclease A at pH 3.0. A three-species mechanism is used as a starting point for the analysis: U1 (slow) in equilibrium U2(fast) in equilibrium N, where U1 and U2 are two forms of the unfolded enzyme with markedly different rates of refolding and N is the native enzyme. This mechanism is based on certain facts established in previous studies of refolding. The kinetics of unfolding and refolding show two phases a fast phase and a slow phase, over a range of temperatures extending above the transition midpoint, Tm. The three-species mechanism can be used in this range. At higher temperatures a new much faster kinetic phase is also observed corresponding to the transient formation of a new intermediate (I). Although the general solution for a four-species mechanism is complex it is not difficult to extend the three-species analysis for the special case found here, in which the fast reaction (I in equilibrium N) is well separated from the other two reactions. At temperatures below the transition zone the slow phase of refolding becomes kinetically complex. No attempt has been made to extend the analysis to include this effect. The basic test of the three-state analysis is the prediction as a function of temperature of alpha2, the relative amplitude of the fast phase, both for unfolding and refolding. At temperatures above Tm for which the three-state analysis must be extended to include the new intermediate I, a crresponding quanitity alpha2(cor) is predicted and compared with measured values. Data used in the three-state prediction are values of tau2 and tau1, the time constants of the fast and slow kinetic phases, plus a single value of alpha2 measured when tau2 and tau1 are well separated. The observed and predicted values of alpha2 agree within experimental error. The analysis predicts correctly that, for these experiments, alpha2 should have the same value in unfolding as in refolding in the final conditions. The analysis also predicts satisfactorily the equilibrium transition curve from kinetic data alone. Four striking properties of the kinetics are explained or correlated by the analysis: (a) the drop in alpha2 to a minimum near Tm as well as the delayed rise in alpha2 above Tm;(b) the vanishing of alpha1 above the transition zone; (c) the sharp drop in tau1 inside the transition zone followed by a partial leveling off outside this zone; and (d) the passage of tau2 through a maximum near Tm. Through a comparison of observed and predicted values of alpha2, the analysis also rules out the alternative three-species mechanism U1 (slow) in equilibrium N (fast) in equilibrium U2. Finally, the temperature dependence of the amplitude for the fast reaction (I in equilibrium N) is discussed; the behavior of I is like that of U2 and I may be an unfolded species populated at equilibrium...  相似文献   

16.
Olfactory responses at the receptor level have been thoroughly described in Drosophila melanogaster by electrophysiological methods. Single sensilla recordings (SSRs) measure neuronal activity in intact individuals in response to odors. For sensilla that contain more than one olfactory receptor neuron (ORN), their different spontaneous spike amplitudes can distinguish each signal under resting conditions. However, activity is mainly described by spike frequency.Some reports on ORN response dynamics studied two components in the olfactory responses of ORNs: a fast component that is reflected by the spike frequency and a slow component that is observed in the LFP (local field potential, the single sensillum counterpart of the electroantennogram, EAG). However, no apparent correlation was found between the two elements.In this report, we show that odorant stimulation produces two different effects in the fast component, affecting spike frequency and spike amplitude. Spike amplitude clearly diminishes at the beginning of a response, but it recovers more slowly than spike frequency after stimulus cessation, suggesting that ORNs return to resting conditions long after they recover a normal spontaneous spike frequency. Moreover, spike amplitude recovery follows the same kinetics as the slow voltage component measured by the LFP, suggesting that both measures are connected.These results were obtained in ab2 and ab3 sensilla in response to two odors at different concentrations. Both spike amplitude and LFP kinetics depend on odorant, concentration and neuron, suggesting that like the EAG they may reflect olfactory information.  相似文献   

17.
Shmuel Malkin  Jim Barber 《BBA》1978,502(3):524-541
1. Using a phosphoroscope, delayed luminescence and prompt chlorophyll fluorescence from isolated chloroplasts have been compared during the induction period.2. Two distinct decay components of delayed luminescence were measured a “fast” component (from ≈1 ms to ≈6 ms) and a “slow” component (at ≈6 ms).3. The fast luminescence component often did not correlate with the fluorescence changes while the slow component significantly changed its intensity during the induction period in a manner which could usually be linearly correlated with variable portion of the fluorescence yield change.4. This correlation was evident after preillumination with far-red light or after allowing a considerable time for dark relaxation.5. The close relationship between the slow luminescence component and variable fluorescence yield was observed with a large range of light intensities and also in the presence of 3(3,4-dichlorophenyl)-1,1-dimethylurea which considerably changes the fluorescence induction kinetics.6. Valinomycin and other antibiotics reduced the amplitude of the 6 ms (slow) luminescence without affecting its relation with the fluorescence induction suggesting possibly that a constant electrical gradient exist in the dark or formed very rapidly in the light, which effects the emission intensity.7. Changes in salt levels of suspending media equally affected the amplitude of both delayed luminescence and variable fluorescence under conditions when the reduction of Q is maximal and constant.8. The results are discussed in terms of several models. It is concluded that the model of independent Photosystem II units together with photosynthetic back reaction concept is incompatible with the data. Other alternative models (the “lake” model and photosynthetic back reaction; recombination of charges in the antenna chlorophyll; the “W” hypothesis) were in closer agreement with the results.  相似文献   

18.
It is presently unclear how the fast and slow components of pulmonary oxygen uptake (VO(2)) kinetics would be altered by body posture during heavy exercise [i.e., above the lactate threshold (LT)]. Nine subjects performed transitions from unloaded cycling to work rates representing moderate (below the estimated LT) and heavy exercise (VO(2) equal to 50% of the difference between LT and peak VO(2)) under conditions of upright and supine positions. During moderate exercise, the steady-state increase in VO(2) was similar in the two positions, but VO(2) kinetics were slower in the supine position. During heavy exercise, the rate of adjustment of VO(2) to the 6-min value was also slower in the supine position but was characterized by a significant reduction in the amplitude of the fast component of VO(2), without a significant slowing of the phase 2 time constant. However, the amplitude of the slow component was significantly increased, such that the end-exercise VO(2) was the same in the two positions. The changes in VO(2) kinetics for the supine vs. upright position were paralleled by a blunted response of heart rate at 2 min into exercise during supine compared with upright heavy exercise. Thus the supine position was associated with not only a greater amplitude of the slow component for VO(2) but also, concomitantly, with a reduced amplitude of the fast component; this latter effect may be due, at least in part, to an attenuated early rise in heart rate in the supine position.  相似文献   

19.
Ionic currents of enzymatically dispersed type I and type II cells of the carotid body have been studied using the whole cell variant of the patch-clamp technique. Type II cells only have a tiny, slowly activating outward potassium current. By contrast, in every type I chemoreceptor cell studied we found (a) sodium, (b) calcium, and (c) potassium currents. (a) The sodium current has a fast activation time course and an activation threshold at approximately -40 mV. At all voltages inactivation follows a single exponential time course. The time constant of inactivation is 0.67 ms at 0 mV. Half steady state inactivation occurs at a membrane potential of approximately -50 mV. (b) The calcium current is almost totally abolished when most of the external calcium is replaced by magnesium. The activation threshold of this current is at approximately -40 mV and at 0 mV it reaches a peak amplitude in 6-8 ms. The calcium current inactivates very slowly and only decreases to 27% of the maximal value at the end of 300-ms pulses to 40 mV. The calcium current was about two times larger when barium ions were used as charge carriers instead of calcium ions. Barium ions also shifted 15-20 mV toward negative voltages the conductance vs. voltage curve. Deactivation kinetics of the calcium current follows a biphasic time course well fitted by the sum of two exponentials. At -80 mV the slow component has a time constant of 1.3 +/- 0.4 ms whereas the fast component, with an amplitude about 20 times larger than the slow component, has a time constant of 0.16 +/- 0.03 ms. These results suggest that type I cells have predominantly fast deactivating calcium channels. The slow component of the tails may represent the activity of a small population of slowly deactivating calcium channels, although other possibilities are considered. (c) Potassium current seems to be mainly due to the activity of voltage-dependent potassium channels, but a small percentage of calcium-activated channels may also exist. This current activates slowly, reaches a peak amplitude in 5-10 ms, and thereafter slowly inactivates. Inactivation is almost complete in 250-300 ms. The potassium current is reversibly blocked by tetraethylammonium. Under current-clamp conditions type I cells can spontaneously fire large action potentials. These results indicate that type I cells are excitable and have a variety of ionic conductances. We suggest a possible participation of these conductances in chemoreception.  相似文献   

20.
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   

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