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.
Slow charge movement in mammalian skeletal muscle   总被引:11,自引:5,他引:6       下载免费PDF全文
Voltage-dependent charge movements were measured in the rat omohyoid muscle with the three-microelectrode voltage-clamp technique. Contraction was abolished with hypertonic sucrose. The standard (ON-OFF) protocol for eliciting charge movements was to depolarize the fiber from -90 mV to a variable test potential (V) and then repolarize the fiber to -90 mV. The quantity of charge moved saturated at test potentials of approximately 0 mV. The steady state dependence of the amount of charge that moves as a function of test potential could be well fitted by the Boltzmann relation: Q = Qmax/(1 + exp[-(V - V)/k]), where Qmax is the maximum charge that can be moved, V is the potential at which half the charge moves, and k is a constant. At 15 degrees C, these values were Qmax = 28.5 nC/microF, V = -34.2 mV, and k = 8.7 mV. Qmax, k, and V exhibited little temperature dependence over the range 7-25 degrees C. "Stepped OFF" charge movements were elicited by depolarizing the fiber from -90 mV to a fixed conditioning level that moved nearly all the mobile charge (0 mV), and then repolarizing the fiber to varying test potentials. The sum of the charge that moved when the fiber was depolarized directly from -90 mV to a given test potential and the stepped OFF charge that moved when the fiber was repolarized to the same test potential had at all test potentials a value close to Qmax for that fiber. In nearly all cases, the decay phase of ON, OFF, and stepped OFF charge movements could be well fitted with a single exponential. The time constant, tau decay, for an ON charge movement at a given test potential was comparable to tau decay for a stepped OFF charge movement at the same test potential. Tau decay had a bell-shaped dependence on membrane potential: it was slowest at a potential near V (the midpoint of the steady state charge distribution) and became symmetrically faster on either side of this potential. Raising the temperature from 7 to 15 degrees C caused tau decay to become faster by about the same proportion at all potentials, with a Q10 averaging 2.16. Raising the temperature from 15 to 25 degrees C caused tau decay to become faster at potentials near V, but not at potentials farther away.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
The voltage dependence of light-induced proton pumping was studied with bacteriorhodopsin (bR) from Halobacterium salinarum, expressed in the plasma membrane of oocytes from Xenopus laevis in the range -160 mV to +60 mV at different light intensities. Depending on the applied field, the quenching effect by blue light, which bypasses the normal photo and transport cycle, is drastically increased at inhibiting (negative) potentials, and is diminished at pump current increasing (positive) potentials. At any potential, two processes with different time constants for the M --> bR decay of approximately 5 ms (tau1) and approximately 20 ms (tau2) are obtained. At pump-inhibiting potentials, a third, long-lasting process with tau3 approximately 300 ms at neutral pH is observed. The fast processes (tau1, tau2) can be assigned to the decay of M2 in the normal pump cycle, i.e., to the reprotonation of the Schiff base via the cytoplasmic side, whereas tau3 is due to the decay of M1 without net pumping, i.e., the reprotonation of the Schiff base via the extracellular side. The results are supported by determination of photocurrents induced by bR on planar lipid films. The pH dependence of the slow decay of M1 is fully in agreement with the interpretation that the reprotonation of the Schiff base occurs from the extracellular side. The results give strong evidence that an externally applied electrical field changes the ratio of the M1 and the M2 intermediate. As a consequence, the transport cycle branches into a nontransporting cycle at negative potentials. This interpretation explains the current-voltage behavior of bR on a new basis, but agrees with the isomerisation, switch, transfer model for vectorial transport.  相似文献   

3.
M-current is a time- and voltage-dependent potassium current which is suppressible by muscarinic receptor activation. We have used curve fitting and noise analysis to determine if macroscopic M-currents deviate from a previously predicted simple two-state kinetic scheme. The M-current was best described by three kinetically distinct components: 'fast' (tau 0), 'intermediate' (tau 1) and 'slow' (tau 2) time constants. The 'fast' (tau 0) and 'intermediate' (tau 1) components were identified from the spectra of M-current noise at potentials positive to the cells' resting membrane potential. The 'intermediate' (tau 1) and 'slow' (tau 2) components were seen by curve fitting M-current deactivation currents. The 'intermediate' (tau 1) time constant was voltage dependent (decreasing e-fold in 23 mV), but voltage dependence of the 'fast' (tau 0) and 'slow' (tau 2) components was not obvious. All kinetic components were sensitive to muscarine, with the 'intermediate' (tau 1) and 'slow' (tau 2) being equally so. These data suggest that all components may derive from the same channel population, and that the M-channel may have at least four kinetic states.  相似文献   

4.
The effect of low pH on the kinetics of Na channel ionic and gating currents was studied in frog skeletal muscle fibers. Lowering external pH from 7.4 to 5.0 slows the time course of Na current consistent with about a +25-mV shift in the voltage dependence of activation and inactivation time constants. Similar shifts in voltage dependence adequately describe the effects of low pH on the tail current time constant (+23.3 mV) and the gating charge vs. voltage relationship (+22.1 mV). A significantly smaller shift of +13.3 mV described the effect of pH 5.0 solution on the voltage dependence of steady state inactivation. Changes in the time course of gating current at low pH were complex and could not be described as a shift in voltage dependence. tau g, the time constant that describes the time course of the major component of gating charge movement, was slowed in pH 5.0 solution by a factor of approximately 3.5 for potentials from -60 to +45 mV. We conclude that the effects of low pH on Na channel gating cannot be attributed simply to a change in surface potential. Therefore, although it may be appropriate to describe the effect of low pH on some Na channel kinetic properties as a "shift" in voltage dependence, it is not appropriate to interpret such shifts as a measure of changes in surface potential. The maximum gating charge elicited from a holding potential of -150 mV was little affected by low pH.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
The rapid inward sodium current in spherical clusters of 11-d-old embryonic chick heart cells, ranging in size between 65 and 90 micron diameter, was studied using the two-microelectrode voltage-clamp technique. Using these preparations, it was possible to resolve the activation phase of the rapid inward current for potentials negative to -25 mV at 37 degrees C. The rapid inward current exhibited a voltage and time dependence similar to that observed in other excitable tissues. It was initiated at potential steps more positive than -45 mV. The magnitude of the current reached its maximum value at a potential of approximately -20 mV. The measured reversal potential was that predicted by the Nernst equation for sodium ions. The falling phase of the current followed a single exponential time-course with a time constant of inactivation, tau h, ranging between 2.14 ms at -40 mV and 0.18 ms at -5 mV. The time constant of inactivation, tau h, determined by a single voltage-step protocol was compared to the constant, tau c, determined by a double voltage-step protocol and no significant different between the two constants of inactivation was found. Furthermore, the time constants of inactivation and reactivation at the same potential in the same preparation were similar. The results of this study demonstrate that the sodium current of heart cells recorded at 37 degrees C can be described by Hodgkin-Huxley kinetics with speeds approximately four times faster than the squid giant axon at 15 degrees C.  相似文献   

6.
Normal muscle has a resting potential of -85 mV, but in a number of situations there is depolarization of the resting potential that alters excitability. To better understand the effect of resting potential on muscle excitability we attempted to accurately simulate excitability at both normal and depolarized resting potentials. To accurately simulate excitability we found that it was necessary to include a resting potential-dependent shift in the voltage dependence of sodium channel activation and fast inactivation. We recorded sodium currents from muscle fibers in vivo and found that prolonged changes in holding potential cause shifts in the voltage dependence of both activation and fast inactivation of sodium currents. We also found that altering the amplitude of the prepulse or test pulse produced differences in the voltage dependence of activation and inactivation respectively. Since only the Nav1.4 sodium channel isoform is present in significant quantity in adult skeletal muscle, this suggests that either there are multiple states of Nav1.4 that differ in their voltage dependence of gating or there is a distribution in the voltage dependence of gating of Nav1.4. Taken together, our data suggest that changes in resting potential toward more positive potentials favor states of Nav1.4 with depolarized voltage dependence of gating and thus shift voltage dependence of the sodium current. We propose that resting potential-induced shifts in the voltage dependence of sodium channel gating are essential to properly regulate muscle excitability in vivo.  相似文献   

7.
Single ventricular myocytes of adult mice were prepared by enzymatic dissociation for voltage clamp experiments with the one suction pipette dialysis method. After blocking the Na current by 10(-4) mol/l TTX early outward currents (IEO) with incomplete inactivation could be elicited by clamping from -50 mV to test potentials (VT) positive to -30 mV. Interfering Ca currents were very small (less than 0.6 nA at VT = 0 mV). The approximation of IEO by the q4r-model showed a pronounced decrease in the time constant of activation (tau q) to more positive potentials. At 50 ms test pulses the time course of the incomplete inactivation could be described by two exponentials and a constant. The time constant of the fast exponential (tau r1) showed a slight decline towards more positive test potentials (8.1 +/- 1.0 ms at -10 mV; 5.8 +/- 1.2 ms at +50 mV, mean +/- SD, n = 5) whereas the time constant of the slow exponential (tau r2) was voltage independent (41.1 +/- 7.9 ms, mean +/- SD, n = 5). The contributions of the fast exponential and the pedestal increased towards positive test potentials. The Q10 value for the time constants of activation and fast inactivation was 2.36 +/- 0.19 and 2.51 +/- 0.09 (mean +/- SD, n = 3), respectively. After an initial delay the recovery of IEO at a recovery potential of -50 mV could be fitted monoexponentially with a time constant of 16.3 +/- 2.9 ms (mean +/- SD, n = 3). The time course of the onset of inactivation determined with the double pulse protocol was slower than the decay at the same potential, and could be described as sum of a fast (tau = 18.4 +/- 6.0 ms) and a slow (tau = 62.1 +/- 19.9ms, mean +/- SD, n = 3) exponential. IEO could be blocked completely by 1 mmol/l 4-aminopyridine at potentials up to +20 mV. Stronger depolarizations had an unblocking effect.  相似文献   

8.
External pH (pH(o)) modifies T-type calcium channel gating and permeation properties. The mechanisms of T-type channel modulation by pH remain unclear because native currents are small and are contaminated with L-type calcium currents. Heterologous expression of the human cloned T-type channel, alpha1H, enables us to determine the effect of changing pH on isolated T-type calcium currents. External acidification from pH(o) 8.2 to pH(o) 5.5 shifts the midpoint potential (V(1/2)) for steady-state inactivation by 11 mV, shifts the V(1/2) for maximal activation by 40 mV, and reduces the voltage dependence of channel activation. The alpha1H reversal potential (E(rev)) shifts from +49 mV at pH(o) 8.2 to +36 mV at pH(o) 5.5. The maximal macroscopic conductance (G(max)) of alpha1H increases at pH(o) 5.5 compared to pH(o) 8.2. The E(rev) and G(max) data taken together suggest that external protons decrease calcium/monovalent ion relative permeability. In response to a sustained depolarization alpha1H currents inactivate with a single exponential function. The macroscopic inactivation time constant is a steep function of voltage for potentials < -30 mV at pH(o) 8.2. At pH(o) 5.5 the voltage dependence of tau(inact) shifts more depolarized, and is also a more gradual function of voltage. The macroscopic deactivation time constant (tau(deact)) is a function of voltage at the potentials tested. At pH(o) 5.5 the voltage dependence of tau(deact) is simply transposed by approximately 40 mV, without a concomitant change in the voltage dependence. Similarly, the delay in recovery from inactivation at V(rec) of -80 mV in pH(o) 5.5 is similar to that with a V(rec) of -120 mV at pH(o) 8.2. We conclude that alpha1H is uniquely modified by pH(o) compared to other calcium channels. Protons do not block alpha1H current. Rather, a proton-induced change in activation gating accounts for most of the change in current magnitude with acidification.  相似文献   

9.
K(+) currents through ERG (ether-à-go-go related gene) channels were recorded in whole-cell voltage clamped NG108-15 neuroblastomaxglioma hybrid cells. The channels were fully activated by low holding potential (V(H)=-20 mV) and long depolarizing prepulses. Hyperpolarizing pulses elicited inward currents which deactivated after reaching a peak. Lowering [Ca(2+)](o) from 5 to 1. 5 or 0.5 mM decreased tau(-1), the rate constant of deactivation. The effect can be explained by a shift of the tau(-1)(V) curve to more negative potentials caused by an increase in surface charge density. Plotting tau(-1) against [Ca(2+)](o) for different potentials yielded straight lines; their slope was independent of potential at -140 to -120 mV and decreased at more positive potentials. The time to peak curve and the maximum of the steady-state inward current were also shifted to more negative potentials. In addition, peak ERG inward current increased. Raising [Ca(2+)](o) from 5 to 10 mM accelerated deactivation and decreased the peak current. 5 mM Ba(2+) affected tau(-1) similarly and inhibited peak current more strongly whereas 5 mM Mg(2+) was less potent. As found by Faravelli et al. (J. Physiol. 496 (1996) 13), bath solutions devoid of divalent cations (0 Ca(2+), 0 Mg(2+), 0.1 or 1.1 mM EGTA) abolished deactivation almost completely. The phenomenon was seen with bath containing either 40 or 6.5 mM K(+). Its occurrence was favored by raising the temperature to 34 degrees C. It suggests a particular requirement of channel closing for Ca(2+).  相似文献   

10.
The effects of deuterium oxide (D2O) and temperature on the properties of endplate channels were studied in voltage-clamped muscle fibers from the frog Rana pipiens. Studies were performed at temperatures of 8, 12, 16, and 20 degrees C. The single channel conductance (gamma) and mean channel lifetime (tau) were calculated from fluctuation analysis of the acetylcholine-induced end-plate currents. The reversal potential was determined by interpolation of the acetylcholine-induced current-voltage relation. The mean reversal potential was slightly more negative in D2O Ringer's (-7.9 +/- 0.1 mV [+/- SEM]) compared with H2O Ringer's (-5.2 +/- 0.6 mV, P less than 0.01). The single channel conductance was decreased in D2O. This decrease was greater than could be accounted for by the increased viscosity of D2O solutions, and the amount of the decrease was greater at higher temperatures. For example, gamma was 38.4 +/- 1.3 pS (+/- SEM) in H2O Ringer's and 25.7 +/- 1.0 pS in D2O Ringer's for a holding potential of -70 mV at 12 degrees C. The mean channel lifetime was significantly shorter in D2O, and the effect was greater at lower temperatures. There was not a strong effect of solvent on the temperature dependence of gamma. On the other hand, the temperature dependence of the reciprocal mean channel lifetime, alpha (where alpha = 1/tau), was strongly dependent upon the solvent. The single channel conductances showed no demonstrable voltage dependence over the range of -90 to -50 mV in both solvents. The reciprocal mean channel lifetime showed a voltage dependence, which could be described by the relation alpha = B exp(AV). The slope A was not strongly affected by either temperature or the solvent. On the other hand, the intercept B was a strong function of temperature and was weakly dependent upon the solvent, with most values greater in D2O. The D2O effects on alpha were what would be expected if they were due to the properties of D2O as a solvent (solvent isotope effects), while the D2O effects on gamma must also include the exchange of D for H in the vicinity of the selectivity filter (primary and/or secondary kinetic isotope effects).  相似文献   

11.
Inactivation of excitation-contraction coupling was examined in extensor digitorum longus (EDL) and soleus muscle fibers from rats injected daily with tri-iodothyronine (T3, 150 micrograms/kg) for 10-14 d. Steady-state activation and inactivation curves for contraction were obtained from measurements of peak potassium contracture tension at different surface membrane potentials. The experiments tested the hypothesis that noninactivating tension is a "window" tension caused by the overlap of the activation and inactivation curves. Changes in the amplitude and voltage dependence of noninactivating tension should be predicted by the changes in the activation and inactivation curves, if noninactivating tension arises from their overlap. After T3 treatment, the area of overlap increased in EDL fibers and decreased in soleus fibers and the overlap region was shifted to more negative potentials in both muscles. Noninactivating tension also appeared at more negative membrane potentials after T3 treatment in both EDL and soleus fibers. The effects of T3 treatment were confirmed with a two microelectrode voltage-clamp technique: at the resting membrane potential (-80 mV) contraction in response to a brief test pulse required less than normal depolarization in EDL, but more than normal depolarization in soleus fibers. After T3 treatment, the increase in contraction threshold at depolarized holding potentials (attributed to inactivation) occurred at more depolarized holding potentials in EDL, or less depolarized holding potentials in soleus. The changes in contraction threshold could be accounted for by the effects of T3 on the activation and inactivation curves. In conclusion, (a) T3 appeared to affect the expression of both activation and inactivation characteristics, but the activation effects could not be cleanly distinguished from T3 effects on the sarcoplasmic reticulum and contractile proteins, and (b) the experiments provided evidence for the hypothesis that the noninactivating tension is a steady-state "window" tension.  相似文献   

12.
Chloride-dependent action potentials were elicited from embryonic skeletal muscle fibers of the chick during the last week of in ovo development. The duration of the action potentials was extremely long (greater than 8 sec). The action potentials were reversibly blocked by the stilbene derivative, SITS, a specific blocker of chloride permeability. Using patch clamp pipettes, in which the intracellular chloride concentration was controlled and with other types of ion channels blocked, the membrane potential at the peak of the action potential closely coincided with the chloride equilibrium potential calculated from the Nernst equation. These data indicate that activation of a chloride-selective conductance underlies the long duration action potential. The occurrence of the chloride-dependent action potential was found to increase during embryonic development. The percentage of fibers that displayed the action potential increased from approximately 20% at embryonic day 13 to approximately 70% at hatching. Chloride-dependent action potentials were not found in adult fibers. The voltage and time-dependent currents underlying the action potential were recorded under voltage clamp using the whole-cell version of the patch pipette technique. The reversal potential of the currents was found to shift with the chloride concentration gradient in a manner predicted by the Nernst equation, and the currents were blocked by SITS. These data indicate that chloride ions were the charge carriers. The conductance was activated by depolarization and exhibited very slow activation and deactivation kinetics.  相似文献   

13.
With the aid of a microelectrophysiological method data were obtained on the difference of the fibers of the inferior rectus muscle of a frog by the type of electrical activity. The greater part of the fibers displayed a continuous tonic activity in the form of polymorphous postsynaptic potentials. Individual fibers responded by a series of action potentials to depolarization created by the introduction of a microelectrode; such reaction was characteristic of the skeletal transitional fibers. The rest of the fibers were inactive, at least some of them with a high membrane potential could be referred to the phasic system.  相似文献   

14.
(1) Single myelinated nerve fibers of Rana esculenta were treated with the steroidal alkaloid batrachotoxin, and Na+ currents and Na+-current fluctuations were measured near the resting potential under voltage-clamp conditions. Between test pulses the fibres were held at hyperpolarizing membrane potentials. (2) The spectral density of Na+-current fluctuations was fitted by the sum of a 1/f component and a Lorentzian function. The time constant tau c = 1/(2 pi fc) obtained from the corner frequency fc of the Lorentzian function approximately agreed with the activation time constant tau m of the macroscopic currents. (3) The conductance gamma of a single Na+ channel modified by batrachotoxin was calculated from the integral of the Lorentzian function and the steady-state Na+ current. At the resting potential V = 0 we obtained gamma - 1.6 pS, higher gamma-values of 3.2 and 3.45 pS were found at V = --8 and --16 mV, respectively. (4) The conductance of a modified Na+ channel is significantly lower than the values 6.4 to 8.85 pS reported in the literature for normal Na+ channels. Hence, our experiments are in agreement with the view that batrachotoxin acts in an 'all-or-none' manner on Na+ channels and creates a distinct population of modified channels.  相似文献   

15.
Excitatory postsynaptic currents (EPSCs) have been studied in voltage- clamped bullfrog sympathetic ganglion B cells. The EPSC was small, rose to a peak within 1-3 ms, and then decayed exponentially over most of its time-course. For 36 cells at --50 mV (21-23 degrees C), peak EPSC size was --6.5 +/- 3.5 nA (mean +/- SD), and the mean decay time constant tau was 5.3 +/- 0.9 ms. tau showed a small negative voltage dependence, which appeared independent of temperature, over the range -- 90 to --30 mV; the coefficient of voltage dependence was --0.0039 +/- 0.0014 mV-1 (n = 29). The peak current-voltage relationship was linear between --120 and --30 mV but often deviated from linearity at more positive potentials. The reversal potential determined by interpolation was approximately --5 mV. EPSC decay tau had a Q10 = 3. The commonly used cholinesterase inhibitors, neostigmine and physostigmine, exhibited complex actions at the ganglia. Neostigmine (1 X 10(-5)M) produced a time-dependent slowing of EPSC decay without consistent change in EPSC size. In addition, the decay phase often deviated from a single exponential function, although it retained its negative voltage dependence. With 1 x 10(-6) M physostigmine, EPSC decay was slowed by the decay phase remained exponential. At higher concentrations of physostigmine, EPSC decay was markedly prolonged and was composed of at least two decay components. High concentrations of atropine (10(-5) to 10(-4) M) produced complex alterations in EPSC decay, creating two or more exponential components; one decay component was faster and the other was slower than that observed in untreated cells. These results suggest that the time-course of ganglionic EPSC decay is primarily determined by the kinetics of the receptor-channel complex rather than hydrolysis or diffusion of transmitter away from the postsynaptic receptors.  相似文献   

16.
Transgenic mice have been increasingly utilized to investigate the molecular mechanisms of cardiac arrhythmias, yet the rate dependence of the murine action potential duration and the electrical restitution curve (ERC) remain undefined. In the present study, 21 isolated, Langendorff-perfused, and atrioventricular node-ablated mouse hearts were studied. Left ventricular and left atrial action potentials were recorded using a validated miniaturized monophasic action potential probe. Murine action potentials (AP) were measured at 30, 50, 70, and 90% repolarization (APD(30)-APD(90)) during steady-state pacing and varied coupling intervals to determine ERCs. Murine APD showed rate adaptation as well as restitution properties. The ERC time course differed dramatically between early and late repolarization: APD(30) shortened with increasing S1-S2 intervals, whereas APD(90) was prolonged. When fitted with a monoexponential function, APD(30) reached plateau values significantly faster than APD(90) (tau = 29 +/- 2 vs. 78 +/- 6 ms, P < 0.01, n = 12). The slope of early APD(90) restitution was significantly <1 (0.16 +/- 0.02). Atrial myocardium had shorter final repolarization and significantly faster ERCs that were shifted leftward compared with ventricular myocardium. Recovery kinetics of intracellular Ca(2+) transients recorded from isolated ventricular myocytes at 37 degrees C (tau = 93 +/- 4 ms, n = 18) resembled the APD(90) ERC kinetics. We conclude that mouse myocardium shows AP cycle length dependence and electrical restitution properties that are surprisingly similar to those of larger mammals and humans.  相似文献   

17.
Inactivation viewed through single sodium channels   总被引:17,自引:12,他引:5       下载免费PDF全文
Recordings of the sodium current in tissue-cultured GH3 cells show that the rate of inactivation in whole cell and averaged single channel records is voltage dependent: tau h varied e-fold/approximately 26 mV. The source of this voltage dependence was investigated by examining the voltage dependence of individual rate constants, estimated by maximum likelihood analysis of single channel records, in a five-state kinetic model. The rate constant for inactivating from the open state, rather than closing, increased with depolarization, as did the probability that an open channel inactivates. The rate constant for closing from the open state had the opposite voltage dependence. Both rate constants contributed to the mean open time, which was not very voltage dependent. Both open time and burst duration were less than tau h for voltages up to -20 mV. The slowest time constant of activation, tau m, was measured from whole cell records, by fitting a single exponential either to tail currents or to activating currents in trypsin-treated cells, in which the inactivation was abolished. tau m was a bell-shaped function of voltage and had a voltage dependence similar to tau h at voltages more positive than -35 mV, but was smaller than tau h. At potentials more negative than about -10 mV, individual channels may open and close several times before inactivating. Therefore, averaged single channel records, which correspond with macroscopic current elicited by a depolarization, are best described by a convolution of the first latency density with the autocorrelation function rather than with 1 - (channel open time distribution). The voltage dependence of inactivation from the open state, in addition to that of the activation process, is a significant factor in determining the voltage dependence of macroscopic inactivation. Although the rates of activation and inactivation overlapped greatly, independent and coupled inactivation could not be statistically distinguished for two models examined. Although rates of activation affect the observed rate of inactivation at intermediate voltages, extrapolation of our estimates of rate constants suggests that at very depolarized voltages the activation process is so fast that it is an insignificant factor in the time course of inactivation. Prediction of gating currents shows that an inherently voltage-dependent inactivation process need not produce a conspicuous component in the gating current.  相似文献   

18.
Li J  Takahashi E  Gunner MR 《Biochemistry》2000,39(25):7445-7454
The electron transfer from the reduced primary quinone (Q(A)(-)) to the secondary quinone (Q(B)) can occur in two phases with a well-characterized 100 micros component (tau(2)) and a faster process occurring in less than 10 micros (tau(1)). The fast reaction is clearly seen when the native ubiquinone-10 at Q(A) is replaced with naphthoquinones. The dependence of tau(1) on the free-energy difference between the P(+)Q(A)(-)Q(B) and P(+)Q(A)Q(B)(-) states (-) and on the pH was measured using naphthoquinones with different electrochemical midpoint potentials as Q(A) in Rhodobacter sphaeroides reaction centers (RCs) and in RCs where - is changed by mutation of M265 in the Q(A) site from Ile to Thr (M265IT). Q(B) was ubiquinone (UQ(B)) in all cases. Electron transfer was measured by using the absorption differences of the naphthosemiquinone at Q(A) and the ubisemiquinone at Q(B) between 390 and 500 nm. As - was changed from -90 to -250 meV tau(1) decreased from 29 to 0.2 micros. The free-energy dependence of tau(1) provides a reorganization energy of 850 +/- 100 meV for the electron transfer from Q(A)(-) to Q(B). The slower reaction at tau(2) is free-energy independent, so processes other than electron transfer determine the observed rate. The fraction of the reaction at tau(1) increases with increasing driving force and is 100% of the reaction when - is approximately 100 meV more favorable than in the native RCs with ubiquinone as Q(A). The fast phase, tau(1), is pH independent from pH 6 to 11 while tau(2) slows above pH 9. As the Q(A) isoprene tail length is increased from 2 to 10 isoprene units the fraction at tau(1) decreases. However, tau(1), tau(2), and the fraction of the reaction in each phase are independent of the tail length of UQ(B).  相似文献   

19.
We have used X-ray fiber diffraction to probe the structure of fibers of tau and tau fragments. Fibers of fragments from the microtubule binding domain had a cross beta-structure that closely resembles that reported both for neurofibrillary tangles found in Alzheimer's disease brain and for fibrous lesions from other protein folding diseases. In contrast, fibers of full-length tau had a different, more complex structure. Despite major differences at the molecular level, all fiber types exhibited very similar morphology by electron microscopy. These results have a number of implications for understanding the etiology of Alzheimer's and other tauopathic diseases. The morphology of the peptide fibers suggests that the region in tau corresponding to the peptides plays a critical role in the nucleation of fiber assembly. The dramatically different structure of the full length tau fibers suggests that some region in tau has enough inherent structure to interfere with the formation of cross beta-fibers. Additionally, the similar appearance by electron microscopy of fibrils with varying molecular structure suggests that different molecular arrangements may exist in other samples of fibers formed from tau.  相似文献   

20.
Glycerol-extracted rabbit psoas muscle fibers were impaled with KCl-filled glass microelectrodes. For fibers at rest-length, the potentials were significantly more negative in solutions producing relaxation than in solutions producing either rigor or contraction; further the potentials in the latter two cases were not significantly different. For stretched fibers, with no overlap between thick and thin filaments, the potentials did not differ in the rigor, the relaxation, or the contraction solutions. The potentials measured from fibers in rigor did not vary significantly with the sarcomere length. For relaxed fibers, however, the potential magnitude decreased with increasing sarcomere length. The difference between the potentials measured for rigor and relaxed fibers exhibited a nonlinear relationship with sarcomere length. The potentials from calcium-insensitive fibers were less negative in both the rigor and the relaxation solutions than those from normal fibers. When calcium-insensitive fibers had been incubated in Hasselbach and Schneider's solution plus MgCl2 or Guba-Straub's solution plus MgATP the potentials recorded upon impalement were similar in the rigor and the relaxation solution to those obtained from normal fibers in the relaxed state. It is concluded that the increase in the negative potential as the glycerinated fiber goes from rigor to relaxation may be due to an alteration in the conformation of the contractile proteins in the relaxed state.  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