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 0 毫秒
1.
J R Glenney  P Kaulfus  K Weber 《Cell》1981,24(2):471-480
We have studied the mechanism of Ca++-dependent restriction of actin filament length by villin, one of the major actin-associated proteins of intestinal microvilli microfilament bundles. Villin acts, even at a ratio of 1 to 1000 with respect to actin, very efficiently as a Ca++-dependent nucleation factor on actin assembly. This gives rise to unidirectional assembly, with the morphologically defined "barbed" end of the resulting filament being capped. Consequently, at steady state treadmilling of actin monomers through the filament is inhibited. Increase of the villin-to-actin ratio enhances the number of nucleated filaments necessarily shorter in length. This results finally in nonsedimentable F actin and a low molecular weight complex of one villin and three monomeric actins, which itself is a potent nucleator. Thus restriction of actin assembly by villin is not due to a direct inhibition of assembly but arises as the consequence of strongly enhanced nucleation followed by unidirectional elongation at the pointed end of the nucleated filaments. In addition, in the presence of Ca++-villin, but not the villin-actin complex, seems able to "break" or "sever" preformed F actin filaments. Thus a variety of cellular phenomena-nucleation, unidirectional assembly, filament end capping, nonpolymerizable actin and F actin bundles-can be observed in vitro in a two-protein component system modulated by the concentration of free Ca++.  相似文献   

2.
We have re-examined the Ca(++)-dependent interaction of an intestinal microvillar 95- kdalton protein (MV-95K) and actin using the isolated acrosomal process bundles from limulus sperm. Making use of the processes as nuclei for assembling actin filaments, we quantitatively and qualitatively examined MV-95K’s effect on filament assembly and on F- actin, both in the presence and in the absence of Ca(++). The acrosomal processes are particularly advantageous for this approach because they nucleate large numbers of filaments, they are extremely stable, and their morphology can be used to determine the polarity of any nucleated filaments. When filament nucleation was initiated in the presence of MV-95K and the absence of Ca(++), there was biased filament assembly from the bundle ends. The calculated elongation rates from both the barbed and pointed filament ends were virtually indistinguishable from control preparations. In the presence of Ca(++), MV-95K completely inhibited filament assembly from the barbed filament end without affecting the initial rate of assembly from the pointed filament end. The inhibition of assembly results from MV-95K binding to and capping the barbed filament end, thereby preventing monomer addition. This indicates that, while MV-95K is a potent nucleator of actin assembly, it is also a potent inhibitor of actin filament elongation. To examine the effects of MV-95K on F-actin in the presence of Ca(++), we developed an assay where MV-95K is added to filaments previously assembled from acrosomal processes without causing filament breakage during mixing. These results clearly demonstrated that rapid filament shortening by MV-95K results through a mechanism of disrupting intrafilament monomer-monomer interactions. Finally, we show that tropomyosin-containing actin filaments are insensitive to cutting, but not to capping, by MV-95K in the presence of Ca(++).  相似文献   

3.
Temperature-dependence of protein dynamics can provide information on details of the free energy landscape by probing the characteristics of the potential responsible for the fluctuations. We have investigated the temperature-dependence of picosecond to nanosecond backbone dynamics at carbonyl carbon sites in chicken villin headpiece subdomain protein using a combination of three NMR relaxation rates: 13C′ longitudinal rate, and two cross-correlated rates involving dipolar and chemical shift anisotropy (CSA) relaxation mechanisms, 13C′/13C′-13Cα CSA/dipolar and 13C′/13C′–15N CSA/dipolar. Order parameters have been extracted using the Lipari-Szabo model-free approach assuming a separation of the time scales of internal and molecular motions in the 2–16°C temperature range. There is a gradual deviation from this assumption from lower to higher temperatures, such that above 16°C the separation of the time scales is inconsistent with the experimental data and, thus, the Lipari-Szabo formalism can not be applied. While there are variations among the residues, on the average the order parameters indicate a markedly steeper temperature dependence at backbone carbonyl carbons compared to that probed at amide nitrogens in an earlier study. This strongly advocates for probing sites other than amide nitrogen for accurate characterization of the potential and other thermodynamics characteristics of protein backbone.  相似文献   

4.
Abstract. The distribution of the mRNA encoding for villin, the major actin-binding protein of intestinal brush border, was studied during the differentiation of mouse intestinal epithelial cells and compared to the distribution of the protein. In situ hybridization using a cRNA clone specific for villin indicated that the distribution of the mRNA did not fully parallel that of the protein, although the overall labeling pattern for mRNA and protein along the crypt-villus axis was similar. While villin was present in equal amounts in all cells along the villi, villin-specific mRNA was mainly accumulated in the cells at the villus base, the area of the epithelium where terminal differentiation takes place and where the brush border is formed.  相似文献   

5.
The pH dependences of activities of homogenous hydrogenases of Thiocapsa roseopersicina and Desulfomicrobium baculatum in the reaction of hydrogen uptake in solution in the presence of benzyl viologen and the pH dependences of catalytic currents of hydrogen oxidation by electrodes on which these hydrogenases were immobilized were compared. Maximal activities of the hydrogenases from T. roseopersicina and D. baculatum in the reaction hydrogen uptake in solution were observed at pH 9.5 and 8.5, respectively. However, the steady-state current caused by catalytic uptake of hydrogen was maximal for the T. roseopersicina hydrogenase-containing electrode at pH 5.5-6.5 under overvoltage of 30-60 mV, whereas for electrodes with D. baculatum hydrogenase it was maximal at pH 6.0-6.5. Analysis of these data suggests that pH-dependent changes in the hydrogenase activities in solution during hydrogen uptake are due not only to the effect of proton concentration on the enzyme conformation or protonation of certain groups of the enzyme active center, but they are rather indicative of changes in free energy of the reaction accompanying changes in pH.  相似文献   

6.
This paper examines, in mouse spleen lymphocytes, the effect of anti-immunoglobulin (anti-Ig) on the cytoplasmic free calcium concentration, [Ca2+]i, measured with the fluorescent indicator quin2, and the relationship of [Ca2+]i to the capping of surface Ig. Anti-Ig causes a rapid rise of [Ca2+], which precedes capping. Assuming that only those 40-50% of the cells which can bind anti-Ig (the B cells) undergo a [Ca2+]i response, [Ca2+]i in those cells approaches 500 nM. It declines to resting levels over many minutes, roughly paralleling the formation of caps. Part of the [Ca2+]i signal is due to stimulated influx across the plasma membrane, since in Ca2+-free medium, anti-Ig gives a smaller and shorter [Ca2+]i rise. The amplitude of this reduced transient now varies inversely with quin2 content, as if some 0.25 mmol Ca per liter of cells was released into the cytoplasm from internal stores. These stores are probably sequestered in organelles since A23187 in Ca2+-free medium also causes a transient [Ca2+]i rise after which anti-Ig has no effect. These organelles seem not to be mitochondria because uncouplers have hardly any effect on [Ca2+]i. Though anti-Ig normally raises [Ca2+]i before causing capping, there seems to be no causal link between the two events. Cells in Ca2+-free medium whose stores have been emptied by A23187, still cap with anti-Ig even though there is no [Ca2+]i rise. Cells loaded with quin2 in the absence of external Ca2+ still cap anti-Ig normally even though their [Ca2+]i remains steady at below 30 nM, four times lower than normal resting [Ca2+]i.  相似文献   

7.
The biochemical properties of villin purified from the brush borders of chicken and rat small intestines were compared, with emphasis on their physical properties and their Ca++-dependent interaction with actin. Like chicken villin, rat villin exists as two isoforms present in equimolar concentrations; the rat isoforms are slightly more acidic than those of chicken villin (6.08 and 6.11 versus 6.26 and 6.34). Rabbit antisera raised against either villin crossreacted with the other one. Like the avian protein, rat villin bundled F-actin at calcium concentrations below 0.1 microM. Above approximately 1 microM calcium, it accelerated the rate of actin assembly and restricted filament lengths of F-actin formed either during coassembly with villin or by addition of villin to preformed filaments. The threshold calcium concentration required for effective severing of preformed filaments was approximately tenfold higher than that required for restricting lengths during coassembly. The extent of filament shortening was proportional to the amount of villin present. At a fixed villin concentration, filament length decreased with increasing [Ca++] over a broad range from 10(-7)-10(-4) M. In general, the mean filament lengths and the dispersion about the mean value were lower in samples where filaments were coassembled with villin than when villin was added to preformed filaments.  相似文献   

8.
Macrophage capping protein (MCP) is a Ca(2+)-sensitive protein which reversibly blocks the barbed ends of actin filaments but does not sever preformed actin filaments. The human cDNA for MCP has been cloned and sequenced. The derived amino acid sequence predicts a polypeptide of 38.4 kDa. Human MCP expressed in Escherichia coli using a pET12a vector was functionally identical to the native protein purified from rabbit alveolar macrophages with respect to Ca2+ sensitivity and ability to block monomer exchange at the barbed end of actin filaments. Sequence comparison with other actin-binding protein sequences indicates that MCP is a member of the gelsolin/villin family of barbed end blocking proteins. Unlike gelsolin, this protein has a limited tissue distribution being detected primarily in macrophages where it was abundant, representing 0.9-1% of the total cytoplasmic protein. Northern blot analysis of U937 and HL60 cells differentiated to macrophage-like cells demonstrated that MCP message increases to 2.6 and greater than 7 times initial levels, respectively. Human MCP displays a 93% amino acid sequence identity with two recently described mouse proteins, gCap39 and Mbh1. Its abundance in macrophages and the corresponding increases in mRNA levels upon promyelocyte and monocyte development into macrophages indicate that MCP may play an important role in macrophage function.  相似文献   

9.
Arrhenius plots of succinate oxidase activity in intact beef heart mitochondria show a clear transition from a low to a high activation energy at 27°C. This temperature is significantly higher than that observed for ATPase (17°C). Arrhenius plots of succinate-cytochromec reductase and cytochromec oxidase also show anomalous curves; while the latter has a breakpoint (at 26°C) only when assayed manometrically, the former has a break at only 20°C.The succinoxidase activity of lipid-deficient mitochondria depends upon addition of exogenous phospholipids. Unsaturated phospholipids are more active than saturated phospholipids but the latter become very effective in restoration of succinoxidase at increasing temperatures. It is suggested that a liquid-crystalline state of the phospholipids is required for correct binding to the lipid-depleted membrane and for restoration of respiratory activity. The is no clear correlation between the above mentioned effects in lipid deficient mitochondria and the transitions in the Arrhenius plots of intact mitochondria.  相似文献   

10.
The villin headpiece (HP67) is a 67 residue, monomeric protein derived from the C-terminal domain of villin. Wild-type HP67 (WT HP67) is the smallest fragment of villin that retains strong in vitro actin-binding activity. WT HP67 is made up of two subdomains, which form a tightly packed interface. The C-terminal subdomain of WT HP67, denoted HP35, is rich in helical structure, folds in isolation, and has been widely used as a model system for folding studies. In contrast, very little is known about the folding of the intact villin headpiece domain. Here, NMR, CD and H/2H amide exchange measurements are used to follow the pH, thermal and urea-induced unfolding of WT HP67 and a mutant (HP67 H41Y) in which a buried conserved histidine in the N-terminal subdomain, His41, has been mutated to Tyr. Although most small proteins display two-state equilibrium unfolding, the results presented here demonstrate that unfolding of the villin headpiece is a multistate process. The presence of a folded N-terminal subdomain is shown to stabilize the C-terminal subdomain, increasing the midpoints of the thermal and urea-induced unfolding transitions and increasing protection factors for H/2H exchange. Histidine 41 has been shown to act as a pH-dependent switch in wild-type HP67: the N-terminal subdomain is unfolded when His41 is protonated, while the C-terminal subdomain remains folded irrespective of the protonation state of His41. Mutation of His41 to Tyr eliminates the segmental pH-dependent unfolding of the headpiece. The mutation stabilizes both domains, but folding is still multistate, indicating that His41 is not solely responsible for the unusual equilibrium unfolding behavior of villin headpiece domain.  相似文献   

11.
J J Marengo  C Hidalgo    R Bull 《Biophysical journal》1998,74(3):1263-1277
The calcium dependence of ryanodine-sensitive single calcium channels was studied after fusing with planar lipid bilayers sarcoendoplasmic reticulum vesicles isolated from excitable tissues. Native channels from mammalian or amphibian skeletal muscle displayed three different calcium dependencies, cardiac (C), mammalian skeletal (MS), and low fractional open times (low Po), as reported for channels from brain cortex. Native channels from cardiac muscle presented only the MS and C dependencies. Channels with the MS or low Po behaviors showed bell-shaped calcium dependencies, but the latter had fractional open times of <0.1 at all [Ca2+]. Channels with C calcium dependence were activated by [Ca2+] < 10 microM and were not inhibited by increasing cis [Ca2+] up to 0.5 mM. After oxidation with 2,2'-dithiodipyridine or thimerosal, channels with low Po or MS dependencies increased their activity. These channels modified their calcium dependencies sequentially, from low Po to MS and C, or from MS to C. Reduction with glutathione of channels with C dependence (native or oxidized) decreased their fractional open times in 0.5 mM cis [Ca2+], from near unity to 0.1-0.3. These results show that all native channels displayed at least two calcium dependencies regardless of their origin, and that these changed after treatment with redox reagents.  相似文献   

12.
Isolated rat and mouse extensor digitorum longus (EDL) and soleus muscles were studied under isometric and isotonic conditions at temperatures from approximately 8 degrees -38 degrees C. The rate constant for the exponential rise of tension during an isometric tetanus had a Q10 of approximately 2.5 for all muscles (corresponding to an enthalpy of activation, delta H = 66 kJ/mol, if the rate was determined by a single chemical reaction). The half-contraction time, contraction time, and maximum rate of rise for tension in an isometric twitch and the maximum shortening velocity in an isotonic contraction all had a similar temperature dependence (i.e., delta H approximately 66 kJ/mol). The Mg++ ATPase rates of myofibrils prepared from rat EDL and soleus muscles had a steeper temperature dependence (delta H = 130 kJ/mol), but absolute rates at 20 degrees C were lower than the rate of rise of tension. This suggests that the Mg++ ATPase cycle rate is not limiting for force generation. A substantial fraction of cross-bridges may exist in a resting state that converts to the force-producing state at a rate faster than required to complete the cycle and repopulate the resting state. The temperature dependence for the rate constant of the exponential decay of tension during an isometric twitch or short tetanus (and the half-fall time of a twitch) had a break point at approximately 20 degrees C, with apparent enthalpy values of delta H = 117 kJ/mol below 20 degrees C and delta H = 70 kJ/mol above 20 degrees C. The break point and the values of delta H at high and low temperatures agree closely with published values for the delta H of the sarcoplasmic reticulum (SR) Ca++ ATPase. Thus, the temperature dependence for the relaxation rate of a twitch or a short tetanus is consistent with that for the reabsorption rate of Ca++ into the SR.  相似文献   

13.
F1-ATPase, a water-soluble portion of the enzyme ATP synthase, is a rotary molecular motor driven by ATP hydrolysis. To learn how the kinetics of rotation are regulated, we have investigated the rotational characteristics of a thermophilic F1-ATPase over the temperature range 4-50°C by attaching a polystyrene bead (or bead duplex) to the rotor subunit and observing its rotation under a microscope. The apparent rate of ATP binding estimated at low ATP concentrations increased from 1.2 × 106 M−1 s−1 at 4°C to 4.3 × 107 M−1 s−1 at 40°C, whereas the torque estimated at 2 mM ATP remained around 40 pN·nm over 4-50°C. The rotation was stepwise at 4°C, even at the saturating ATP concentration of 2 mM, indicating the presence of a hitherto unresolved rate-limiting reaction that occurs at ATP-waiting angles. We also measured the ATP hydrolysis activity in bulk solution at 4-65°C. F1-ATPase tends to be inactivated by binding ADP tightly. Both the inactivation and reactivation rates were found to rise sharply with temperature, and above 30°C, equilibrium between the active and inactive forms was reached within 2 s, the majority being inactive. Rapid inactivation at high temperatures is consistent with the physiological role of this enzyme, ATP synthesis, in the thermophile.  相似文献   

14.
We have studied the inactivation of high-voltage-activated (HVA), omega- conotoxin-sensitive, N-type Ca2+ current in embryonic chick dorsal root ganglion (DRG) neurons. Voltage steps from -80 to 0 mV produced inward Ca2+ currents that inactivated in a biphasic manner and were fit well with the sum of two exponentials (with time constants of approximately 100 ms and > 1 s). As reported previously, upon depolarization of the holding potential to -40 mV, N current amplitude was significantly reduced and the rapid phase of inactivation all but eliminated (Nowycky, M. C., A. P. Fox, and R. W. Tsien. 1985. Nature. 316:440-443; Fox, A. P., M. C. Nowycky, and R. W. Tsien. 1987a. Journal of Physiology. 394:149-172; Swandulla, D., and C. M. Armstrong. 1988. Journal of General Physiology. 92:197-218; Plummer, M. R., D. E. Logothetis, and P. Hess. 1989. Neuron. 2:1453-1463; Regan, L. J., D. W. Sah, and B. P. Bean. 1991. Neuron. 6:269-280; Cox, D. H., and K. Dunlap. 1992. Journal of Neuroscience. 12:906-914). Such kinetic properties might be explained by a model in which N channels inactivate by both fast and slow voltage-dependent processes. Alternatively, kinetic models of Ca-dependent inactivation suggest that the biphasic kinetics and holding-potential-dependence of N current inactivation could be due to a combination of Ca-dependent and slow voltage- dependent inactivation mechanisms. To distinguish between these possibilities we have performed several experiments to test for the presence of Ca-dependent inactivation. Three lines of evidence suggest that N channels inactivate in a Ca-dependent manner. (a) The total extent of inactivation increased 50%, and the ratio of rapid to slow inactivation increased approximately twofold when the concentration of the Ca2+ buffer, EGTA, in the patch pipette was reduced from 10 to 0.1 mM. (b) With low intracellular EGTA concentrations (0.1 mM), the ratio of rapid to slow inactivation was additionally increased when the extracellular Ca2+ concentration was raised from 0.5 to 5 mM. (c) Substituting Na+ for Ca2+ as the permeant ion eliminated the rapid phase of inactivation. Other results do not support the notion of current-dependent inactivation, however. Although high intracellular EGTA (10 mM) or BAPTA (5 mM) concentrations suppressed the rapid phase inactivation, they did not eliminate it. Increasing the extracellular Ca2+ from 0.5 to 5 mM had little effect on this residual fast inactivation, indicating that it is not appreciably sensitive to Ca2+ influx under these conditions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

15.
Villin headpiece (HP67) is a small, autonomously-folding domain that has become a model system for understanding the fundamental tenets governing protein folding. In this communication, we explore the role that Leu61 plays in the structure and stability of the construct. Deletion of Leu61 results in a completely unfolded protein that cannot be expressed in Escherichia coli. Omission of only the aliphatic leucine side chain (HP67 L61G) perturbed neither the backbone conformation nor the orientation of local hydrophobic side chains. As a result, a large, solvent-exposed hydrophobic pocket, a negative replica of the leucine side-chain, was created on the surface. The loss of the hydrophobic interface between leucine 61 and the hydrophobic pocket destabilized the construct by ~3.3 kcal/mol. Insertion of a single glycine residue immediately before Leu61 (HP67 L61[GL]) was also highly destabilizing and had the effect of altering the backbone conformation (α-helix to π-helix) in order to precisely preserve the wild-type position and conformation of all hydrophobic residues, including Leu61. In addition to demonstrating that the hydrophobic side-chain of Leu61 is critically important for the stability of villin headpiece, our results are consistent with the notion that the precise interactions present within the hydrophobic core, rather than the hydrogen bonds that define the secondary structure, specify a protein's fold.  相似文献   

16.
Summary Villin, a 95-kD cytoskeletal protein selectively expressed in the microvilli of some absorptive cells was localized immunohistochemically in the oviduct and the seminiferous excretory ducts of the mouse. Villin was found in the proximal part of the oviduct, comprising the preampulla, ampulla, and part of the isthmus. Distal to the isthmus the oviductal cells lining the junctura and the intrauterine colliculus tubaris were devoid of villin. No villin could be detected in the uterine cells.Ductuli efferentes, connecting the rete testis with the epididymis were the only portion of the male seminiferous ductal system expressing villin. The cells lining the epididymis and the vas deferens were devoid of villin. These data show that villin is selectively expressed in male and female reproductive systems and that it is limited to anatomically defined proximal portions of the reproductive ducts.  相似文献   

17.
The thermostable 36‐residue subdomain of the villin headpiece (HP36) is the smallest known cooperatively folding protein. Although the folding and internal dynamics of HP36 and close variants have been extensively studied, there has not been a comprehensive investigation of side‐chain motion in this protein. Here, the fast motion of methyl‐bearing amino acid side chains is explored over a range of temperatures using site‐resolved solution nuclear magnetic resonance deuterium relaxation. The squared generalized order parameters of methyl groups extensively spatially segregate according to motional classes. This has not been observed before in any protein studied using this methodology. The class segregation is preserved from 275 to 305 K. Motions detected in Helix 3 suggest a fast timescale of conformational heterogeneity that has not been previously observed but is consistent with a range of folding and dynamics studies. Finally, a comparison between the order parameters in solution with previous results based on solid‐state nuclear magnetic resonance deuterium line shape analysis of HP36 in partially hydrated powders shows a clear disagreement for half of the sites. This result has significant implications for the interpretation of data derived from a variety of approaches that rely on partially hydrated protein samples.  相似文献   

18.
The Semliki Forest virus (SFV) replicase protein nsP1 has methyltransferase (MT) and guanylyltransferase-like (GT) activities, which are involved in the capping of viral mRNAs. MT catalyzes the transfer of the methyl group from S-adenosylmethionine (AdoMet) to position 7 of GTP, and this reaction is followed by GT-catalyzed formation of the covalent complex m7GMP-nsP1. These reactions are virus specific and thus potential targets for inhibitors of virus replication. We have mutated residues of SFV nsP1, which are conserved in related proteins of the large alphavirus-like superfamily. Mutations of D64, D90, R93, C135, C142, and Y249 to alanine destroyed or greatly reduced the MT activity of nsP1. All MT-negative mutants lost also the GT activity, confirming that methylation of GTP is an essential prerequisite for the synthesis of the covalent guanylate complex. Mutation of H38 prevented the GT reaction without destroying MT activity. Conservation of residues essential for both reactions in the alphavirus-like superfamily implies that they use a capping mechanism similar to that for the alphaviruses. Residues D64 and D90 were necessary for AdoMet binding, as measured by UV cross-linking. Secondary structure predictions of nsP1 and other proteins of the superfamily place these residues in positions corresponding to AdoMet-binding sites of cellular methyltransferases, suggesting that they all may be structurally related.  相似文献   

19.
The polyphosphoinositides phosphatidylinositol 4-monophosphate (PIP) and phosphatidylinositol 4,5-bisphosphate (PIP2) inactivate the actin filament-severing proteins villin and gelsolin and dissociate them from monomeric and polymeric actin. A potential polyphosphoinositide- (PPI) binding site of human plasma gelsolin regulating filament severing has been localized to the region between residues 150-169 and to the corresponding region in villin which occurs in the second of six homologous domains present in both proteins. Synthetic peptides based on these sequences bind tightly to both PIP and PIP2, in either micelles or bilayer vesicles, compete with gelsolin for binding to PPIs, and dissociate gelsolin-PIP2 complexes, restoring severing activity to the protein. These peptides also bind with moderate affinity to F-actin, suggesting that inactivation of the severing function of the intact proteins by PPIs results from competition between actin and PPIs for a critical binding site on gelsolin-villin. The PPI-binding peptides contain numerous basic amino acids, but their effects on PPIs are far greater than those of Arg or Lys oligomers, a highly basic peptide derived from the calmodulin-binding site of myristoylated, alanine-rich kinase C substrate protein, or the 5-kDa actin-binding protein thymosin beta-4, suggesting that specific aspects of the primary and secondary structure of these basic peptides are important for their interaction with the acidic headgroups of PPIs. In addition to elucidating the structure of PIP2-binding sites in gelsolin, the results describe a sensitive assay for phosphoinositide-binding molecules based on their ability to prevent inhibition of gelsolin function.  相似文献   

20.
The voltage dependence of carbachol-induced desensitization has been analyzed in potassium-depolarized frog sartorius muscle preparations with voltage clamp techniques over a wide voltage range (-120 to +40 mV). Desensitization developed exponentially at all voltages with tau, the time constant of desensitization onset, varying as a logarithmic function of membrane voltage. The voltage dependence of tau remained in calcium-deficient solutions and was not altered by elevating either the level of extracellular or intracellular calcium. We have analyzed our results according to a simple sequential kinetic scheme in which the rate-limiting step in the development of desensitization is a transition of the receptor channel complex from the activated conducting state to a desensitized, nonconducting state. We conclude (a) that the observed voltage sensitivity of desensitization primarily resides in the voltage dependence of this transition, and (b) the kinetics of activation appear to have a greater influence on the observed rate of desensitization than on its voltage dependence. The magnitude of the voltage dependence suggests that a greater change in free energy is required for the transition to the desensitized state than for the transition between the open and closed states of the receptor channel complex.  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