首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The photovoltage generation by halorhodopsin from Halobacterium salinarum (shR) was examined by adsorbing shR-containing membranes onto a thin polymer film. The photovoltage consisted of two major components: one with a sub-millisecond range time constant and the other with a millisecond range time constant with different amplitudes, as previously reported. These components exhibited different Cl(-) concentration dependencies (0.1-9 M). We found that the time constant for the fast component was relatively independent of the Cl(-) concentration, whereas the time constant for the slow component increased sigmoidally at higher Cl(-) concentrations. The fast and the slow processes were attributed to charge (Cl(-)) movements within the protein and related to Cl(-) ejection, respectively. The laser photolysis studies of shR-membrane suspensions revealed that they corresponded to the formation and the decay of the N intermediate. The photovoltage amplitude of the slow component exhibited a distorted bell-shaped Cl(-) concentration dependence, and the Cl(-) concentration dependence of its time constant suggested a weak and highly cooperative Cl(-)-binding site(s) on the cytoplasmic side (apparent K(D) of approximately 5 M and Hill coefficient > or =5). The Cl(-) concentration dependence of the photovoltage amplitude and the time constant for the slow process suggested a competition between spontaneous relaxation and ion translocation. The time constant for the relaxation was estimated to be >100 ms.  相似文献   

2.
《BBA》1987,893(2):320-332
The primary charge separation in Photosystem I of pea chloroplasts was measured as a photovoltage in the pico- and nanosecond time range by applying laser flashes at 532 nm of variable energy and different duration (12 ns and 30 ps, respectively). Contributions to the photovoltage from Photosystem II was eliminated by addition of 3-(3,4-dichlorophenyl)-1,1-dimethylurea and preillumination. The dependence of the photovoltage amplitude on the excitation energy could be described by an exponential saturation law when the excitation flash had a duration of 12 ns. Nearly the same dependence was found when the excitation source was the train of a mode-locked laser (approx. ten 30-ps flashes spaced by 7 ns; highest energy of a single flash, 80 μJ / cm−2). Even with single 30-ps flashes the photovoltage was only slightly smaller than the one elicited by 12-ns flashes of the same energy. These findings demonstrate that trapping of excitation energy by the reaction center of Photosystem I is much more effective than losses by annihilation and other loss processes. The photovoltage yield was nearly independent of the fraction of closed traps, thus demonstrating that the absorption cross section of Photosystem I is not altered by the closing of its reaction centers. By recording the rise time of the photovoltage with our highest time resolution we found that the trapping rate of the excitation energy in Photosystem I depended on the energy of the 30-ps flashes: at low excitation energies (less than 1014 photons / cm2 per pulse) trapping occurred within 90 ± 15 ps and at high excitation energy (1015 photons / cm2 per pulse) trapping and charge stabilization occurred within the time resolution of the apparatus, i.e., up to 50 ps. The trapping rate at low energies is in agreement with the one determined by fluorescence decay kinetics. Up to 50 ns there was no further detectable electrogenic phase (neither forward nor backward reactions). This demonstrates that all the electrogenicity, produced by the charge separation, takes place in less than 50 ps.  相似文献   

3.
H.-W. Trissl  A. Der  P. Ormos  L. Keszthelyi 《BBA》1984,765(3):288-294
Flash-induced photovoltages were measured with metal electrodes in two experimental systems of purple membranes oriented by an electric field. One system consisted of a suspension of purple membranes cooled to 80 K. The photovoltage evoked by a xenon flash lamp displayed a single phase with a fast rise and a slow RC-decay. The signal shape is consistent with a fast charge separation occurring before the decay of the K-intermediate. The other system consisted of purple membranes embedded and stabilized in polyacrylamide gel. At room temperature, the photovoltage, evoked by a 10 ns laser flash, displayed a negative phase in the submicrosecond range and a slower positive one. The shape of the signals were altered in a complex manner by the stray capacitance and the ionic strength. The rise-time of the negative phase was approx. 14 and approx. 40 ns at ionic strengths of 10 and 1 mM, respectively. The initial peak amplitudes of the photovoltage from both experimental systems depended on the external capacitance in an inverse manner, indicating that both experimental systems were not impedance-matched. The evaluation of kinetic data of molecular reactions from measurements of the photovoltage is discussed.  相似文献   

4.
Processes accompanying the quenching of the fluorescent probe 4"-dimethylaminochalcone by hydroxyl groups of the proton-donor solvent 1-butanol have been studied. The kinetics of the deactivation of the excited state of 4"-dimethylaminochalcone has been monitored from the transition absorption spectra at a time resolution of 50 fs and fluorescence decay at a time resolution of 30 ps. The data obtained allow thinking that the next picture occurs in 1-butanol. At first stage, the 4"-dimethylaminochalcone molecule in its ground state forms a hydrogen bond with an alcohol molecule. At the second stage, the absorption of light quantum and corresponding rise of the dipole moment of 4"-dimethylaminochalcone take place, the initially existing hydrogen bond is retained. The third stage consists in the rearrangement of the 4"-dimethylaminochalcone solvation shell formed by alcohol dipole molecules due to an increase of the dipole of moment 4"-dimethylaminochalcone; this rearrangement takes an energy of about 24 kJ/mol, the arrangement time constant is close to 40 ps; the initial hydrogen bond is retained. The fourth stage involves processes that lead to fluorescence quenching; their time constant is about 200 ps. Taking into account that the quenching is a much slower process than the relaxation of the solvation shell, it was supposed that the quenching is not a direct consequence of the solvation shell relaxation or the existence of the hydrogen bond formed prior to excitation. Then the fluorescence quenching of 4"-dimethylaminochalcone can be accomplished through some other processes that are observed in other fluorescent molecules: (a) rearrangement of the initial hydrogen bond from a conformation that cannot quench the fluorescence of 4"-dimethylaminochalcone to a more "effective" conformation, (b) charge transfer between the excited of molecule 4"-dimethylaminochalcone and alcohol, or (c) solvent-induced twist of the 4"-dimethylaminochalcone amino group (its withdrawal from the molecule plane) by the action of the solvent.  相似文献   

5.
A study was made of the processes associated with the quenching of 4″-dimethylaminochalcone (DMAC) fluorescence by proton-donor solvent (1-butanol). The kinetics of deactivation of the DMAC excited state was assessed by transient absorption spectra with a time resolution about 50 fs and by fluorescence decay with ~30-ps resolution. The following sequence of events could thus be envisaged: (i) the DMAC molecule in the ground state (prior to excitation) makes a hydrogen bond with an alcohol molecule; (ii) absorption of a light quantum causes a corresponding increase of the DMAC dipole moment; the H-bond is retained; (iii) the solvation shell formed by alcohol dipoles is reorganized in response to the raise of the DMAC dipole moment, with an energy expenditure about 24 kJ/mol and a time constant about 40 ps; the initial H-bond is still retained; (iv) processes leading to fluorescence quenching occur with an effective time constant of nearly 200 ps. Since quenching is far slower than solvate rearrangement, one can suppose that it is not a direct consequence of shell relaxation or prior H-bonding. Thus, DMAC fluorescence quenching may involve different processes observed with other aromatic molecules: H-bond rearrangement from a nonquenching to a more ‘efficient’ conformation, charge transfer between the excited molecule and alcohol, or solvent-induced out-of-plane twist of the DMAC amino group.  相似文献   

6.
Excitation energy trapping and charge separation in Photosystem II were studied by kinetic analysis of the fast photovoltage detected in membrane fragments from peas with picosecond excitation. With the primary quinone acceptor oxidized the photovoltage displayed a biphasic rise with apparent time constants of 100–300 ps and 550±50 ps. The first phase was dependent on the excitation energy whereas the second phase was not. We attribute these two phases to trapping (formation of P-680+ Phe-) and charge stabilization (formation of P-680+ QA -), respectively. A reversibility of the trapping process was demonstrated by the effect of the fluorescence quencher DNB and of artificial quinone acceptors on the apparent rate constants and amplitudes. With the primary quinone acceptor reduced a transient photoelectric signal was observed and attributed to the formation and decay of the primary radical pair. The maximum concentration of the radical pair formed with reduced QA was about 30% of that measured with oxidized QA. The recombination time was 0.8–1.2 ns.The competition between trapping and annihilation was estimated by comparison of the photovoltage induced by short (30 ps) and long (12 ns) flashes. These data and the energy dependence of the kinetics were analyzed by a reversible reaction scheme which takes into account singlet-singlet annihilation and progressive closure of reaction centers by bimolecular interaction between excitons and the trap. To put on firmer grounds the evaluation of the molecular rate constants and the relative electrogenicity of the primary reactions in PS II, fluorescence decay data of our preparation were also included in the analysis. Evidence is given that the rates of radical pair formation and charge stabilization are influenced by the membrane potential. The implications of the results for the quantum yield are discussed.Abbreviations DCBQ 2,6-dichloro-p-benzoquinone - DCMU 3-(3,4-dichlorophenyl)-1,1-dimethylurea - DNB m-dinitrobenzene - PPBQ phenyl-p-benzoquinone - PS I photosystem I of green plants - PS II photosystem II of green plants - PSU photosynthetic unit - P-680 primary donor of PS II - Phe intermediary pheophytin acceptor of PS II - QA primary quinone acceptor of PS II - RC reaction center  相似文献   

7.
The photovoltaic signal associated with the primary photochemical event in an oriented bacteriorhodopsin film is measured by directly probing the electric field in the bacteriorhodopsin film using an ultrafast electro-optic sampling technique. The inherent response time is limited only by the laser pulse width of 500 fs, and permits a measurement of the photovoltage with a bandwidth of better than 350 GHz. All previous published studies have been carried out with bandwidths of 50 GHz or lower. We observe a charge buildup with an exponential formation time of 1.68 +/- 0.05 ps and an initial decay time of 31.7 ps. Deconvolution with a 500-fs Gaussian excitation pulse reduces the exponential formation time to 1.61 +/- 0.04 ps. The photovoltaic signal continues to rise for 4.5 ps after excitation, and the voltage profile corresponds well with the population dynamics of the K state. The origin of the fast photovoltage is assigned to the partial isomerization of the chromophore and the coupled motion of the Arg-82 residue during the primary event.  相似文献   

8.
Charge pulse experiments performed on the peat-bog alga Eremosphaera viridis revealed an unusual voltage relaxation behaviour. Injection of charge pulses of 1 microseconds duration resulted in an immediate charging of the membranes (time constant of the order of 40 ns). Nevertheless, the potential-measuring microelectrode recorded an exponential increase in membrane voltage with a time constant of about 1.3 ms. The maximum voltage value was recorded after about 3 ms, followed by an exponential decay with a time constant of about 9.6 ms. This biphasic time course was independent of the amplitude of the injected charge and of the location of the impaled microelectrodes in the vacuole. Centrifuged cells in which the chloroplasts and the other organelles were pelleted in one part of the cells showed the same electrical response. Electrical breakdown of the cell membranes resulted in the disappearance of the biphasic voltage response. In this case only the decaying relaxation process could be recorded with a time constant of 3 ms. After resealing of the membranes the original biphasic relaxation response was restored. Increasing concentrations of KCl in the bathing medium reduced both time constants almost correspondingly. The experimental findings were evaluated with an electrical equivalent circuit. Theoretical analysis with reference to the experimental data suggested that the delayed voltage response of the potential-recording electrode resulted from a membrane seal across the tip of this electrode. The resistance of this seal was calculated to be about 400 M omega. The specific resistances and capacitances of tonoplast and plasmalemma membranes were calculated from the decaying part of the biphasic relaxation curves. The average values were found to be 2.58 omega.m2 and 5 mF.m-2. The investigations reported here suggest that charge pulse experiments can be generally used for the detection of membrane and cytoplasmic material clogging of the tip of intracellular microelectrodes, a problem with which most electrophysiologists are faced when interpreting data obtained from impaled microelectrodes.  相似文献   

9.
Kinetics of the gel to liquid-crystalline phase transition of dipalmitoylphosphatidylcholine vesicle membrane was studied by the stopped-flow technique with turbidity detection. The observed change in turbidity was well characterized by a single-exponential decay curve with relaxation time in the millisecond range, although the existence of a faster process than the dead-time of the stopped-flow apparatus was inferred from the amplitude analysis. Relaxation times were determined as functions of 1-hexanol concentration and temperature just below phase transition. From the analysis based on the theories of nonequilibrium relaxation, it is concluded that the phase transition induced by 1-hexanol is governed by a nonlocalized fluctuation mechanism. The anesthetic-induced nonequilibrium state is unstable rather than metastable.  相似文献   

10.
The time course of structural changes accompanying the transition from the M412 intermediate to the BR568 ground state in the photocycle of bacteriorhodopsin (BR) from Halobacterium halobium was studied at room temperature with a time resolution of 15 ms using synchrotron radiation X-ray diffraction. The M412 decay rate was slowed down by employing mutated BR Asp96Asn in purple membranes at two different pH-values. The observed light-induced intensity changes of in-plane X-ray reflections were fully reversible. For the mutated BR at neutral pH the kinetics of the structural alterations (tau 1/2 = 125 ms) were very similar to those of the optical changes characterizing the M412 decay, whereas at pH 9.6 the structural relaxation (tau 1/2 = 3 s) slightly lagged behind the absorbance changes at 410 nm. The overall X-ray intensity change between the M412 intermediate and the ground state was about 9% for the different samples investigated and is associated with electron density changes close to helix G, B and E. Similar changes (tau 1/2 = 1.3-3.6 s), which also confirm earlier neutron scattering results on the BR568 and M412 intermediates trapped at -180 degrees C, were observed with wild type BR retarded by 2 M guanidine hydrochloride (pH 9.4). The results unequivocally prove that the tertiary structure of BR changes during the photocycle.  相似文献   

11.
P. Ormos  L. Reinisch  L. Keszthelyi 《BBA》1983,722(3):471-479
The time behavior of flash-induced charge movements during the first steps in the bacteriorhodopsin photocycle was measured on a suspension of purple membranes oriented by an electric field. The experiments were done in the temperature range 80–278 K. During the formation of the intermediate K, two negative (with respect to the direction of the proton pump) components of the response signal are well resolved with time constants τ1 < 3 μs and τ2 ? 150 μs at 200 K. The distances of the charge displacements responsible for the electric signals are estimated. On the basis of the results the two components are assigned to two steps in the trans-cis isomerization of the retinal. A third negative component appears at higher temperatures which is related by time constant measurements to the K → L transition.  相似文献   

12.
Excitation energy transfer in monomeric and trimeric forms of photosystem I (PSI) from the cyanobacterium Synechocystis sp. PCC 6803 in solution or immobilized on FTO conducting glass was compared using time-resolved fluorescence. Deposition of PSI on glass preserves bi-exponential excitation decay of ~4–7 and ~21–25 ps lifetimes characteristic of PSI in solution. The faster phase was assigned in part to photochemical quenching (charge separation) of excited bulk chlorophylls and in part to energy transfer from bulk to low-energy (red) chlorophylls. The slower phase was assigned to photochemical quenching of the excitation equilibrated over bulk and red chlorophylls. The main differences between dissolved and immobilized PSI (iPSI) are: (1) the average excitation decay in iPSI is about 11 ps, which is faster by a few ps than for PSI in solution due to significantly faster excitation quenching of bulk chlorophylls by charge separation (~10 ps instead of ~15 ps) accompanied by slightly weaker coupling of bulk and red chlorophylls; (2) the number of red chlorophylls in monomeric PSI increases twice—from 3 in solution to 6 after immobilization—as a result of interaction with neighboring monomers and conducting glass; despite the increased number of red chlorophylls, the excitation decay accelerates in iPSI; (3) the number of red chlorophylls in trimeric PSI is 4 (per monomer) and remains unchanged after immobilization; (4) in all the samples under study, the free energy gap between mean red (emission at ~710 nm) and mean bulk (emission at ~686 nm) emitting states of chlorophylls was estimated at a similar level of 17–27 meV. All these observations indicate that despite slight modifications, dried PSI complexes adsorbed on the FTO surface remain fully functional in terms of excitation energy transfer and primary charge separation that is particularly important in the view of photovoltaic applications of this photosystem.  相似文献   

13.
The temperature dependence of the photovoltage developed by a model membrane containing bacteriorhodopsin (BR) is studied. The model membrane is formed by first coating a thin Teflon sheet with lipid and then fusing BR vesicles to it. The time course of the photoresponse is resolved down to 1 microsecond. The photoresponse is taken to be a sum of exponentials. Exponential time constants and amplitudes are determined by an analysis of the photoresponse with a photovoltage vs. log time plot, correlation filter, and nonlinear least-squares routine. The photovoltage is taken to be the sum of three exponentials but only two of the three time constants are resolved. Both are temperature dependent and indicate a thermally activated transport process. The corresponding activation energies are 55 kJ/mol and 62 kJ/mol. Since the photovoltage is proportional to charge times displacement the corresponding charge displacements are 11 and 34 A assuming a total displacement of 45 A. The remaining exponential term corresponds to a small negative transient in the photovoltage that has a rise time less than 1 microsecond even at -20 degrees C. The calculated charge displacement is estimated to be less than 2 A.  相似文献   

14.
15.
V Sharma  K Stebe  J C Murphy    L Tung 《Biophysical journal》1996,71(6):3229-3241
The effect of a nontoxic, nonionic block co-polymeric surface active agent, poloxamer 188, on electroporation of artificial lipid membranes made of azolectin, was investigated. Two different experimental protocols were used in our study: charge pulse and voltage clamp. For the charge pulse protocol, membranes were pulsed with a 10-micronsecond rectangular voltage waveform, after which membrane voltage decay was observed through an external 1-M omega resistance. For the voltage clamp protocol the membranes were pulsed with a waveform that consisted of an initial 10-microsecond rectangular phase, followed by a negative sloped ramp that decayed to zero in the subsequent 500 microseconds. Several parameters characterizing the electroporation process were measured and compared for the control membranes and membranes treated with 1.0 mM poloxamer 188. For both the charge pulse and voltage clamp experiments, the threshold voltage (amplitude of initial rectangular phase) and latency time (time elapsed between the end of rectangular phase and the onset of membrane electroporation) were measured. Membrane conductance (measured 200 microseconds after the initial rectangular phase) and rise time (tr; the time required for the porated membrane to reach a certain conductance value) were also determined for the voltage clamp experiments, and postelectroporation time constant (PE tau; the time constant for transmembrane voltage decay after onset of electroporation) for the charge pulse experiments. The charge pulse experiments were performed on 23 membranes with 10 control and 13 poloxamer-treated membranes, and voltage pulse experiments on 49 membranes with 26 control and 23 poloxamer-treated membranes. For both charge pulse and voltage clamp experiments, poloxamer 188-treated membranes exhibited a statistically higher threshold voltage (p = 0.1 and p = 0.06, respectively), and longer latency time (p = 0.04 and p = 0.05, respectively). Also, poloxamer 188-treated membranes were found to have a relatively lower conductance (p = 0.001), longer time required for the porated membrane to reach a certain conductance value (p = 0.05), and longer postelectroporation time constant (p = 0.005). Furthermore, addition of poloxamer 188 was found to reduce the membrane capacitance by approximately 4-8% in 5 min. These findings suggest that poloxamer 188 adsorbs into the lipid bilayers, thereby decreasing their susceptibility to electroporation.  相似文献   

16.
We have performed time-resolved fluorescence measurements on photosystem II (PSII) containing membranes (BBY particles) from spinach with open reaction centers. The decay kinetics can be fitted with two main decay components with an average decay time of 150 ps. Comparison with recent kinetic exciton annihilation data on the major light-harvesting complex of PSII (LHCII) suggests that excitation diffusion within the antenna contributes significantly to the overall charge separation time in PSII, which disagrees with previously proposed trap-limited models. To establish to which extent excitation diffusion contributes to the overall charge separation time, we propose a simple coarse-grained method, based on the supramolecular organization of PSII and LHCII in grana membranes, to model the energy migration and charge separation processes in PSII simultaneously in a transparent way. All simulations have in common that the charge separation is fast and nearly irreversible, corresponding to a significant drop in free energy upon primary charge separation, and that in PSII membranes energy migration imposes a larger kinetic barrier for the overall process than primary charge separation.  相似文献   

17.
Synchrotron radiation was used to follow the time course of the transitions, induced by temperature jump, in Escherichia coli membranes and their lipid extracts isolated from a fatty acid auxotroph grown with different fatty acids. We measured the relaxation times associated with the phase transitions as well as with the conformational transition of the hydrocarbon chains and observed different behavior as a function of chemical composition. Relaxation times of about 1-2 s were found at a hexagonal to lamellar phase transition and within a lamellar phase whose parameters display important variations with temperature when the conformational transition takes place. On the other hand, no delay was observed for a phase transition where large lipid or water diffusion was not needed. We have shown that phase transitions and conformational transitions are, to a large extent, uncoupled and that the relaxation times corresponding to the latter transition could be related to the size of the ordered domains. In all cases, the order to disorder conformational transition is more rapid than the disorder to order transition. Finally, the relaxation times of the disorder to order transition observed with the membranes and with their lipid extracts were found to be strongly correlated, indicating that the proteins do not play a role in this transition.  相似文献   

18.
Absorbance changes induced by 25-ps laser flashes were measured in membranes of Heliobacterium chlorum at 15 K. Absorbance difference spectra, measured at various times after the flash showed negative bands in the Qy region at 812, 793 and 665 nm. The first of these bands was attributed to the formation of excited singlet states of a long-wavelength form of antenna bacteriochlorophyll g (BChl g 808). Absorbance changes of shorter wavelength absorbing antenna BChls g were at least an order of magnitude smaller, indicating rapid excitation energy transfer (i.e. within the time resolution of the apparatus) from these BChls to BChl g 808. Excited BChl g 808 showed a bi-exponential decay with time constants of 50 and 200 ps. The bands at 793 and 665 nm may be attributed to the primary charge separation and reflect the photooxidation of the primary electron donor P-798 and photoreduction of a primary electron acceptor absorbing near 670 nm, presumably a BChl c or Chl a-like pigment. The bleaching of this pigment reversed with a time constant of 300 ps at 15 K and of 800 ps at 300 K. This indicates that electron transfer from the primary to the secondary electron acceptor is approximately 2.5 times faster at 15 K than at room temperature.Abbreviations BChl bacteriochlorophyll - FWHM full width at half maximum - P-798 primary electron donor - Tris tris(hydroxymethyl)amino methane  相似文献   

19.
The relaxation kinetics of aqueous lipid dispersions after a pressure jump (p-jump) was investigated using time-resolved pressure perturbation calorimetry (PPC). Analysis of the calorimetric response curves by deconvolution with the instrumental response function gives information about slow processes connected with the lipid phase transition. The lipid transition from the gel to the liquid-crystalline state was found to be a multi-step process with relaxation constants in the seconds range resolvable by time-resolved PPC and faster processes with relaxation times shorter than ca. 5 s that could not be resolved by the instrument. The faster processes comprise ca. 50% of the total heat uptake at the transition midpoint. This is the first calorimetric measurement showing the multi-step nature of the transition. The results are in good agreement with data obtained with other detection methods and with molecular modelling experiments describing the transition as a multi-step process with nucleation and growth steps.  相似文献   

20.
J. Lavorel 《BBA》1973,325(2):213-229
The decay of luminescence in the 6–600-μs range following a microsecond flash has been studied in Chlorella. The decay is highly polyphasic; three kinetic components are outlined, in confirmation of the results of K. L. Zankel (1971, Biochim. Biophys. Acta 245, 373–385).Extrapolation of the decay to zero dark time suggests that a unique metastable species C?+, resulting from photochemical charge separation in the System II reaction center, is the substrate of the recombination reaction which gives rise to luminescence.The fast (5–10 μs) and medium (50–70 μs) phases of the decay denote different stabilization steps, preceding relaxation of the centers by electron and proton transduction to the photosynthetic chain.NH2OH specifically inhibits the fast phase and enhances the medium phase. This effect is explained by assuming that the fast phase results from electron transfer from the water splitting system Z to the oxidized primary donor Y.3-(3,4-Dichlorophenyl)-1,1-dimethylurea (DCMU), in the presence of NH2OH elicits another fast phase. It is believed that DCMU affords a parasitic stabilization of C?+ by forming a complex with Q?.  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