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.
Picosecond laser spectroscopic analysis was applied to determine how many intermediates existed in the primary photochemical process of trans-bacteriorhodopsin (light-adapted bacteriorhodopsin) at room temperature (18°C) and to calculate their absorption spectra. Irradiation of bacteriorhodopsin with a laser pulse (wavelength, 532 nm; pulse width, 25 ps) yielded the K intermediate (K) which was produced through a precursor, having an absorption maximum (λmax) longer than that of K. K was stable during a picosecond time range (50–900 ps). The λmax was located at 610 nm and the extinction coefficient (?max) was 0.92-times that of bacteriorhodopsin. The same K intermediate was produced from bacteriorhodopsin even when it was excited with a high-energy pulse by which a saturation effect was induced. A transient difference spectrum measured at 150 ns after the excitation of bacteriorhodopsin was different in shape from that of the K intermediate, suggesting that an intermediate was formed by thermal decay of K. This intermediate, tentatively called the KL intermediate (KL), had a λmax at 596 nm and an ?max 0.80-times that of bacteriorhodopsin. KL decayed to the L intermediate (L) with a time constant of 2.2 μs. L has a λmax at 543 nm and an ?max 0.66-times that of bacteriorhodopsin.  相似文献   

2.
Excitation of bacteriorhodopsin (BR) in its β absorption band drives a photocycle identical in the millisecond range, to that excited in the α band of BR. The relative contribution of the two transition dipoles distinguished in the β band to the initiation of the photocycle was established by photoselection experiments. Having this information the orientation of the chromophoric plane was specified by electric dichroism measurements.  相似文献   

3.
Cross-Correlation Functions for a Neuronal Model   总被引:5,自引:1,他引:4       下载免费PDF全文
Cross-correlation functions, RXY(t,τ), are obtained for a neuron model which is characterized by constant threshold θ, by resetting to resting level after an output, and by membrane potential U(t) which results from linear summation of excitatory postsynaptic potentials h(t). The results show that: (1) Near time lag τ = 0, RXY(t,τ) = fU [θ-h(τ), t + τ] {h′(τ) + EU [u′(t + τ)]} for positive values of this quantity, where fU(u,t) is the probability density function of U(t) and EU [u′(t + τ)] is the mean value function of U′(t + τ). (2) Minima may appear in RXY(t,τ) for a neuron subjected only to excitation. (3) For large τ, RXY(t,τ) is given approximately by the convolution of the input autocorrelation function with the functional of point (1). (4) RXY(t,τ) is a biased estimator of the shape of h(t), generally over-estimating both its time to peak and its rise time.  相似文献   

4.
Decay-associated fluorescence spectra of the green alga Scenedesmus obliquus have been measured by single-photon timing with picosecond resolution in various states of light adaptation. The data have been analyzed by applying a global data analysis procedure. The amplitudes of the decay-associated spectra allow a determination of the relative antenna sizes of the photosystems. We arrive at the following conclusions: (a) The fluorescence kinetics of algal cells with open PS II centers (F0 level) have to be described by a sum of three exponential components. These decay components are attributed to photosystem (PS) I (τ ≈ 85 ps, λmaxem ≈ 695-700 nm), open PS II α-centers (τ ≈ 300 ps, λmaxem = 685 nm), and open PS II β-centers (τ ≈ 600 ps, λmaxem = 685 nm). A fourth component of very low amplitude (τ ≈ 2.2-2.3 ns, λmaxem = 685 nm) derives from dead chlorophyll. (b) At the Fmax level of fluorescence there are also three decay components. They originate from PS I with properties identical to those at the F0 level, from closed PS II α-centers (τ ≈ 2.2 ns, λmaxem = 685 nm) and from closed PS β-centers (τ ≈ 1.2 ns, λmaxem = 685 nm). (c) The major effect of light-induced state transitions on the fluorescence kinetics involves a change in the relative antenna size of α- and β-units brought about by the reversible migration of light-harvesting complexes between α-centers and β-centers. (d) A transition to state II does not measurably increase the direct absorption cross-section (antenna size) of PS I. Our data can be rationalized in terms of a model of the antenna organization that relates the effects of state transitions and light-harvesting complex phosphorylation with the concepts of PS II α,β-heterogeneity. We discuss why our results are in disagreement with those of a recent lifetime study of Chlorella by M. Hodges and I. Moya (1986, Biochim. Biophys. Acta., 849:193-202).  相似文献   

5.
Fiscus EL 《Plant physiology》1986,80(3):752-759
Volume (Jv) and solute (Js) fluxes through Phaseolus root systems were observed over a 24-hour period. The volume flux was varied in a pressure chamber by altering the hydrostatic pressure in 10 steps, from 0 to 0.41 megapascals. All root systems showed strong diurnal peaks in volume flux. The five transport coefficients (σ, ω, Js*, Lp, and π*) were estimated from a nonlinear least squares algorithm. Analysis of the data revealed that all the coefficients exhibited a diurnal rhythm. When the total differential of the volume flux was considered it was possible to show that the diurnal changes in volume flux were due to a complex interaction between the diurnally shifting coefficients with the role of each highly dependent on the level of volume flux. At low volume fluxes, ω, Js*, and π* accounted for nearly all the diurnal change in volume flux. At high volume fluxes, however, the major influence shifted to Lp and π*, while ω and Js* became relatively unimportant. Thus, π* was the only coefficient of interest across the entire range of Jv and appeared to be the single most important one in determining the diurnal rhythm of Jv under conditions of a constant applied pressure.  相似文献   

6.
Components of Sodium and Chloride Flux Across Toad Bladder   总被引:4,自引:0,他引:4       下载免费PDF全文
The effect of transepithelial potential difference (ψ) on Na and Cl flux across toad bladder was assessed by measuring isotopic flux between identical media at various values of ψ. The contribution of edge damage to ionic permeability was eliminated, resulting in relatively high spontaneous ψ (-97 ±4 mv) and low electrical conductance g. Bidirectional Na fluxes were measured simultaneously. Unidirectional Cl fluxes were measured in paired hemibladders at ψ = 0 mv or -97 mv. Net Na flux JNa, at ψ = 0 mv, was slightly less than short-circuit current (SCC). At ψ = -97 mv, JNa averaged 17% of SCC, and was sometimes zero. ΔJNa/Δψ (= g+) averaged 60% of g between -97 mv and +75 mv; at -150 mv, g+ fell, indicating rectification. Analysis of unidirectional Na fluxes indicates low passive conductance (1.5 μmho/mg wet weight), a bidirectional, electrically neutral flux of approximately 0.13 μa/mg, and relatively large conductance of the active transport path at ψ ≥ -97 mv. The absence of appreciable transstimulation of serosal (S)-to-mucosal (M) Na flux (in response to increasing mucosal Na concentration) indicates that the electrically neutral flux is not exchange diffusion in the usual sense. Analysis of Cl fluxes indicates similar values for passive conductance and neutral flux, suggesting linked neutral flux of Na and Cl. Either the electromotive force of the Na pump E, its conductance ga, or both are strong functions of ψ. The product of these two quantities, Ega, is a measure of the “transport capacity” at any given value of ψ, independent of the direct effect of ψ on JNa through the pump path. Ega varies with ψ. Hence estimation of the net Na flux or current at any one value of ψ, including ψ = 0, fails to reveal the maximal transport capacity of the pump, its resting electromotive force (when JNa = 0 through the pump), or the dependence of transport capacity on potential.  相似文献   

7.
Lifetime of the excited state (τ) of bacteriochlorophyll (BChl) in photosynthetic bacteria, measured with a mode-locked argon laser (oscillating at 488 nm; mode locked at 56 MHz) as light source, ranged from 0.3 to 2.5 nsec. These τ values are reported with a precision of ±0.1 nsec. The value of τ at high exciting light intensity (I) was two to three times that at low intensity. For young cultures of green bacterium Chloropseudomonas ethylicum, τ ranged from 0.5 (low I) to 1.0 nsec (high I); for those of the purple bacterium Rhodospirillum rubrum, from 0.4 (low I) to 1.0 nsec (high I); and for those of the BChl b-containing Rhodopseudomonas viridis, from 1.0 (low I) to 2.5 nsec (high I). These data provide information regarding the efficiencies of the photochemical process in these bacteria. Quantum yield (ø) of BChl fluorescence, calculated from ø = τ/τ0 (where τ0 is the intrinsic lifetime of fluorescence), ranges from 2-6% at low intensities to 6-14% at high intensities.  相似文献   

8.
The excited state kinetics of three different allophycocyanin (AP) complexes has been studied by picosecond fluorescence spectroscopy. Both the fluorescence kinetics and the decay-associated fluorescence spectra of the different complexes can be understood on the basis of a structural model for AP which uses (a) an analogy to the known x-ray determined structure of C-phycocyanin, (b) the biochemical analogies of AP and C-phycocyanin, and (c) the biochemical composition of AP-B (AP-681). A model is developed that describes the excited state kinetics as a mixture of internal conversion processes within a coupled exciton pair and energy transfer processes between exciton pairs. We found excited state relaxation times in the range of 13 ps (AP with linker peptide) up to 66 ps (AP-B). The trimeric aggregates AP 660 and AP 665 show one fast relaxation component each, as was expected on the basis of their symmetry properties. The lower symmetry of AP-B (AP-681) gives rise to two fast lifetime components (τ1 = 23 ps and τ2 = 66 ps) which are attributed to internal conversion and/or energy transfer between excitonic states formed by the coupling of symmetrically and spectrally nonequivalent chromophores. It is proposed that the internal conversion between exciton states of strongly coupled chromophores fulfills the requirements of the small energy gap limit. Thus, internal conversion rates in the order of tens of picoseconds are feasible. The influence of the interaction of the linker peptide on the properties of the AP trimer are manifested in the fluorescence kinetics. Lack of the linker peptide in AP 660 gives rise to a heterogeneity in the chromophore conformations and chromophore-chromophore interactions.  相似文献   

9.
Photochemical reaction dynamics of the primary events in recombinant bacteriorhodopsin (bRrec) was studied by femtosecond laser absorption spectroscopy with 25-fs time resolution. bRrec was produced in an Escherichia coli expression system. Since bRrec was prepared in a DMPC–CHAPS micelle system in the monomeric form, its comparison with trimeric and monomeric forms of the native bacteriorhodopsin (bRtrim and bRmon, respectively) was carried out. We found that bRrec intermediate I (excited state of bR) was formed in the range of 100 fs, as in the case of bRtrim and bRmon. Further processes, namely the decay of the excited state I and the formation of intermediates J and K of bRrec, occurred more slowly compared to bRtrim, but similarly to bRmon. The lifetime of intermediate I, judging from the signal of ΔA ESA(470-480 nm), was 0.68 ps (78%) and 4.4 ps (22%) for bRrec, 0.52 ps (73%) and 1.7 ps (27%) for bRmon, and 0.45 ps (90%) and 1.75 ps (10%) for bRtrim. The formation time of intermediate K, judging from the signal of ΔA GSA(625-635 nm), was 13.5 ps for bRrec, 9.8 ps for bRmon, and 4.3 ps for bRtrim. In addition, there was a decrease in the photoreaction efficiency of bRrec and bRmon as seen by a decrease in absorbance in the differential spectrum of the intermediate K by ~14%. Since photochemical properties of bRrec are similar to those of the monomeric form of the native protein, bRrec and its mutants can be considered as a basis for further studies of the mechanism of bacteriorhodopsin functioning.  相似文献   

10.
The unidirectional uptake of sodium across the outer surface of the isolated frog skin (J12Na) was measured in the presence of transepithelial potential difference (Δψ) ranging from +100 to -100 mV. With a sodium concentration of 115 mM in the bathing solutions J12Na increases significantly when the spontaneous Δψ is reduced to zero by short-circuiting the skin. With an Na concentration of 6 mM a progressive increase J12Na can be observed when Δψ is decreased in several steps from +100 to -100 mV (serosal side positive and negative, respectively). The observed change J12Na amounts to a fraction only of that predicted from the shift in Δψ. The results suggest that under open circuit conditions the potential step across the outside surface is at most one half of Δψ and that the resistance across the outside and inside barrier of the skin is ohmic. This is in agreement with measurements of intracellular potentials in the frog skin and with resistance measurements carried out in the toad skin. The data strongly support the view that the saturating component of Jψ proceeds via a charged carrier system. Exposure to negative values of Δψ of 50 mV or more for times of 24 min or more result in a marked reduction of J12Na which shows only partial or no reversibility.  相似文献   

11.
The flash-induced charge movements during the photocycle of light-adapted bacteriorhodopsin in purple membranes attached to a black lipid membrane were investigated under voltage clamp and current clamp conditions. Signal registration ranged from 200 ns to 30 s after flash excitation using a logarithmic clock, allowing the equally weighted measurement of the electrical phenomena over eight decades of time. The active pumping signals were separated from the passive system discharge on the basis of an equivalent circuit analysis. Both measuring methods were shown to yield equivalent results, but the charge translocation could be accurately monitored over the whole time range only under current clamp conditions. To describe the time course of the photovoltage signals a model based on distributed kinetics was found to be more appropriate than discrete first order processes suggesting the existence of conformational substates with distributed activation energies. The time course of the active charge displacement is characterised by a continuous relaxation time spectrum with three broad peaks plus an unresolved fast transient (<0.3 μs) of opposite polarity. The time constants and relative amplitudes (in brackets) derived from the peak rate constants and relative areas of the three bands are: τ1 = 32 μs (20%), τ2 = 0.89 ms (15%) and τ3 = 18 ms (65%) at 25°C in 150 mM KCl at pH7. The Arrhenius plots of the peak rate constants were linear yielding activation energies of EA1 = 57 kJ/mol, EA2 = 52 kJ/mol, and EA3 = 44 kJ/mol. The electrical signal at 890 μs has no counterpart in the photocycle of bacteriorhodopsin suspensions. Fits with a sum of exponentials required 5 to 6 components and were not reproducible. Analysis of photoelectrical signals with continuous relaxation time spectra gave equally good fits with fewer parameters and were well reproducible.  相似文献   

12.
Fast photosignals (FPS) with R1 and R2 components were measured in retinas of cattle, rat, and frog within a temperature range of 0° to 60°C. Except for temperatures near 0°C the signal rise of the R1 component was determined by the duration of the exciting flash. The kinetics of the R2 component and the meta transition of rhodopsin in the cattle and rat retina were compared. For the analysis of the FPS it is presupposed that the signal is produced by light-induced charges on the outer segment envelope membrane that spread onto the whole plasma membrane of the photoreceptor cell. To a good approximation, this mechanism can be described by a model circuit with two distinct capacitors. In this model, the charging capacitance of the pigmented outer segment envelope membrane and the capacitance of the receptor's nonpigmented plasma membrane are connected via the extra- and intracellular electrolyte resistances. The active charging is explained by two independent processes, both with exponential rise (R1 and R2), that are due to charge displacements within the pigmented envelope membrane. The time constant τ2 of the R2 membrane charging process shows a strong temperature dependence that of the charge redistribution, τr, a weak one. In frog and cattle retinas the active charging is much slower within a large temperature range than the passive charge redistribution. From the two-capacitor model it follows for τr « τ2 that the rise of the R2 component is determined by τr, whereas the decay is given by τ2. For the rat retina, however, τ2 approaches τr at physiological temperatures and becomes <τr above 45°C. In this temperature range where τ2 ≈ τr, both processes affect rise and decay of the photosignal. The absolute values of τr are in good accordance with the known electric parameters of the photoreceptors. At least in the cattle retina, the time constant τ2 is identical with that of the slow component of the meta II formation. The strong temperature dependence of the meta transition time gives rise to the marked decrease of the R2 amplitude with falling temperature. As the R1 rise could not be fully time resolved the signal analysis does not yield the time constant τ1 of the R1 generating process. It could be established, however, within the whole temperature range that the decay of the R1 component is determined by τr. Using an extended model that allows for membrane leakage, we show that in normal ringer solution the membrane time constant does not influence the signal time-course and amplitude.  相似文献   

13.
In studies of gating currents of rabbit cardiac Ca channels expressed as α1C2a or α1C2a2δ subunit combinations in tsA201 cells, we found that long-lasting depolarization shifted the distribution of mobile charge to very negative potentials. The phenomenon has been termed charge interconversion in native skeletal muscle (Brum, G., and E. Ríos. 1987. J. Physiol. (Camb.). 387:489–517) and cardiac Ca channels (Shirokov, R., R. Levis, N. Shirokova, and E. Ríos. 1992. J. Gen. Physiol. 99:863–895). Charge 1 (voltage of half-maximal transfer, V1/2 ≃ 0 mV) gates noninactivated channels, while charge 2 (V1/2 ≃ −90 mV) is generated in inactivated channels. In α1C2a cells, the available charge 1 decreased upon inactivating depolarization with a time constant τ ≃ 8, while the available charge 2 decreased upon recovery from inactivation (at −200 mV) with τ ≃ 0.3 s. These processes therefore are much slower than charge movement, which takes <50 ms. This separation between the time scale of measurable charge movement and that of changes in their availability, which was even wider in the presence of α2δ, implies that charges 1 and 2 originate from separate channel modes. Because clear modal separation characterizes slow (C-type) inactivation of Na and K channels, this observation establishes the nature of voltage-dependent inactivation of L-type Ca channels as slow or C-type. The presence of the α2δ subunit did not change the V1/2 of charge 2, but sped up the reduction of charge 1 upon inactivation at 40 mV (to τ ≃ 2 s), while slowing the reduction of charge 2 upon recovery (τ ≃ 2 s). The observations were well simulated with a model that describes activation as continuous electrodiffusion (Levitt, D. 1989. Biophys. J. 55:489–498) and inactivation as discrete modal change. The effects of α2δ are reproduced assuming that the subunit lowers the free energy of the inactivated mode.  相似文献   

14.
Studies were made of the dependence of the rate of oxygen consumption, Jr, on the electrical potential difference, Δψ, across the frog skin. After the abolition of sodium transport by ouabain the basal oxygen consumption was independent of Δψ. In fresh skins Jr was a linear function of Δψ over a range of at least ±70 mv. Treatment with aldosterone stimulated the short-circuit current, Io, and the associated rate of oxygen consumption, Jro, and increased their stability; linearity was then demonstrable over a range of ±160 mv. Brief perturbations of Δψ (±30–200 mv) did not alter subsequent values of Io. Perturbations for 10 min or more produced a "memory" effect both with and without aldosterone: accelerating sodium transport by negative clamping lowered the subsequent value of Io; positive clamping induced the opposite effect. Changes in Jro were more readily detectable in the presence of aldosterone; these were in the same direction as the changes in Io. The linearity of Jr in Δψ indicates the validity of analysis in terms of linear nonequilibrium thermodynamics—brief perturbations of Δψ appear to produce no significant effect on either the phenomenological coefficients or the free energy of the metabolic driving reaction. Hence it is possible to evaluate this free energy.  相似文献   

15.
We investigate the hydration state of horse-heart cytochrome c (hh cyt c) in the unfolding process induced by trifluoroacetic acid (TFA). The conformation of hh cyt c changes from the native (N) state (2.9 < pH < 6.0) to the acid-unfolded (UA) state (1.7 < pH < 2.0) to the acid-induced molten globule (A) state (pH ∼1.2). Hydration properties of hh cyt c during this process are measured at 20°C by high-resolution dielectric relaxation (DR) spectroscopy, UV-vis absorbance, and circular dichroism spectroscopy. Constrained water of hh cyt c is observed at every pH as an ∼5-GHz Debye component (DC) (DR time, τD ∼30 ps) and its DR amplitude (DRA) is increased by 77% upon N-to-UA transition, when pH changes from 6.0 to 2.0. Even in the N state, the DRA of the constrained-water component is found to be increased by 22% with decreasing pH from 6.0 to 2.9, suggesting an increase in the accessible surface area of native hh cyt c. Moreover, hypermobile water around native hh cyt c is detected at pH 6.0 as a 19-GHz DC (τD ∼ 8.4 ps < τDW = 9.4 ps), but is not found at other pH values. The DRA signal of constrained water is found to return to the pH 2.9 (N-state) level upon UA-to-A transition. Fast-response water (slightly slower than bulk) around A-state hh cyt c is detected at pH 1.2, and this suggests some accumulation of TFA ions around the peptide chain. Thus, this high-resolution DR spectroscopy study reveals that hh cyt c exhibits significant hydration-state change in the TFA-unfolding process.  相似文献   

16.
The τ subunit of Escherichia coli DNA polymerase III holoenzyme interacts with the α subunit through its C-terminal Domain V, τC16. We show that the extreme C-terminal region of τC16 constitutes the site of interaction with α. The τC16 domain, but not a derivative of it with a C-terminal deletion of seven residues (τC16Δ7), forms an isolable complex with α. Surface plasmon resonance measurements were used to determine the dissociation constant (KD) of the α−τC16 complex to be ~260pM. Competition with immobilized τC16 by τC16 derivatives for binding to α gave values of KD of 7μM for the α−τC16Δ7 complex. Low-level expression of the genes encoding τC16 and τC167, but not τC16Δ11, is lethal to E. coli. Suppression of this lethal phenotype enabled selection of mutations in the 3′ end of the τC16 gene, that led to defects in α binding. The data suggest that the unstructured C-terminus of τ becomes folded into a helix–loop–helix in its complex with α. An N-terminally extended construct, τC24, was found to bind DNA in a salt-sensitive manner while no binding was observed for τC16, suggesting that the processivity switch of the replisome functionally involves Domain IV of τ.  相似文献   

17.
Undoped and Er-doped NaY(WO4)2 disordered single crystals have been grown by the Czochralski technique. The specific heat and thermal conductivity (κ) of these crystals have been characterized from T = 4 K to 700 K and 360 K, respectively. It is shown that κ exhibits anisotropy characteristic of single crystals as well as a κ(T) behavior observed in glasses, with a saturation mean free phonon path of 3.6 Å and 4.5 Å for propagation along a and c crystal axes, respectively. The relative energy positions and irreducible representations of Stark Er3+ levels up to 4G7/2 multiplet have been determined by the combination of experimental low (<10 K) temperature optical absorption and photoluminescence measurements and simulations with a single-electron Hamiltonian including both free-ion and crystal field interactions. Absorption, emission and gain cross sections of the 4I13/24I15/2 laser related transition have been determined at 77 K. The 4I13/2 Er3+ lifetime (τ) was measured in the temperature range of 77–300 K, and was found to change from τ (77K) ≈ 4.5 ms to τ (300K) ≈ 3.5 ms. Laser operation is demonstrated at 77 K and 300 K by resonantly pumping the 4I13/2 multiplet at λ≈1500 nm with a broadband (FWHM≈20 nm) diode laser source perfectly matching the 77 K crystal 4I15/24I13/2 absorption profile. At 77 K as much as 5.5 W of output power were obtained in π-polarized configuration with a slope efficiency versus absorbed pump power of 57%, the free running laser wavelength in air was λ≈1611 nm with the laser output bandwidth of 3.5 nm. The laser emission was tunable over 30.7 nm, from 1590.7 nm to 1621.4 nm, for the same π-polarized configuration.  相似文献   

18.
The kinetic diversity of burst openings responsible for the persistent Na+ current (INaP) in entorhinal cortex neurons was examined by separately analyzing single bursts. Although remarkable kinetic variability was observed among bursts in terms of intraburst opening probability and mean open and closed times, the values of time constants describing intraburst open times (τo(b)s) and closed times (τc(b)s) were distributed around well-defined peaks. At −40 mV, τo(b) peaks were found at ~0.34 (τo(b)1) and 0.77 (τo(b)2) ms, and major τc(b) peaks were found at ~0.24 (τc(b)1) and 0.54 (τc(b)2) ms. In ~80% of the bursts two preferential gating modes were found that consisted of a combination of either τo(b)1 and τc(b)2 (“intraburst mode 1”), or τo(b)2 and τc(b)1 (“intraburst mode 2”). Individual channels could switch between different gating modalities, but normally tended to maintain a specific gating mode for long periods. Mean burst duration also displayed considerable variability. At least three time constants were found to describe burst duration, and the frequencies at which each of the corresponding “bursting states” occurred varied in different channels. Short-lasting bursting states were preferentially associated with intraburst mode 1, whereas very-long-lasting bursts tended to gate according to mode 2 only or other modes that included considerably longer mean open times. These results show that INaP channels can generate multiple intraburst open and closed states and bursting states, but these different kinetic states tend to combine in definite ways to produce a limited number of prevalent, well-defined gating modalities. Modulation of distinct gating modalities in individual Na+ channels may be a powerful form of plasticity to influence neuronal excitability and function.  相似文献   

19.
《BBA》1985,810(1):94-105
Picosecond absorbance difference spectra at a number of delay times after a 35 ps excitation pulse and kinetics of absorbance changes were measured in chromatophores of the photosynthetic purple bacterium Rhodospirillum rubrum after chemical oxidation of the primary electron donor P-875. Kinetics and spectra were measured of the excited singlet states of carotenoid and bacteriochlorophyll a and also of the triplet state of the carotenoid. The excited singlet state of carotenoid, produced by direct excitation at 532 nm, is characterized by a bleaching of the ground state absorption bands in the region 450–490 nm and by an absorbance increase with a maximum near 570 nm. Its lifetime was calculated to be 0.6 ± 0.1 ps in vitro and less than 1 ps in vivo. The triplet state of carotenoid in vivo is formed within 100 ps after direct carotenoid excitation via a pathway that does not involve excited states of bacteriochlorophyll. Singlet excitation of a bacteriochlorophyll a molecule causes the bleaching of its Qx and Qy absorption bands, and is probably associated with blue shifts of the Qy absorption band of about six neighboring bacteriochlorophyll molecules. Upon increasing the excitation density, the average lifetime of the singlet excitations on bacteriochlorophyll decreased from about 350 ps to about 10 ps or less. The results are in quantitative agreement with the known effect of singlet-singlet annihilation upon the fluorescence yield, and furthermore show that no bacteriochlorophyll or carotenoid triplet formation is associated with this annihilation.  相似文献   

20.
Pressure volume curves for Alternanthera philoxeroides (Mart.) Griseb. (alligator weed) grown in 0 to 400 millimolar NaCl were used to determine water potential (Ψ), osmotic potential (ψs), turgor potential (ψp) and the bulk elastic modulus (ε) of shoots at different tissue water contents. Values of ψs decreased with increasing salinity and tissue Ψ was always lower than rhizosphere Ψ. The relationship between ψp and tissue water content changed because ε increased with salinity. As a result, salt-stressed plants had larger ranges of positive turgor but smaller ranges of tissue water content over which ψp was positive. To our knowledge, this is the first report of such a salinity effect on ε in higher plants. These increases in ε with salinity provided a mechanism by which a large difference between plant Ψ and rhizosphere Ψ, the driving force for water uptake, could be produced with relatively little water loss by the plant. A time-course study of response after salinization to 400 millimolar NaCl showed Ψ was constant within 1 day, ψs and ψp continued to change for 2 to 4 days, and ε continued to change for 4 to 12 days. Changes in ε modified the capacity of alligator weed to maintain a positive water balance and consideration of such changes in other species of higher plants should improve our understanding of salt stress.  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