首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
The origins of the initial mutations in sporadic retinoblastoma were explored using polymorphic markers from chromosome 13q. The paternal chromosome was maintained in 3 of 3 informative bilateral tumors which had undergone reduction to homozygosity for regions of this chromosome. The paternal chromosome was maintained in 7 of 8 informative unilateral tumors which likewise demonstrated a reduction of homozygosity. These data are in contrast to previously published studies of chromosome retention in unilateral retinoblastoma [Dryja, T. P., Mukai, S., Petersen, R., Rapaport, J. M., Walton, D., and Yandel, D. W. Nature (Lond.), 339: 556-558, 1989; Zhu, Z., Dunn, J. M., Phillips, R. A., Goddard, A. D., Paton, K. E., Becker, A., and Gallie, B. L. Nature (Lond.), 340: 312-313, 1989] and provide the first evidence that genomic imprinting may play a role in this disease.  相似文献   

2.
Time-resolved neutron diffraction on retinal rod outer segments are performed to reinvestigate the origin of the light-induced structural change observed by Saibil et al. (Saibil, H., M. Chabre, and D. L. Worcester, 1976, Nature (Lond.), 262:266-270). Photoactivating rhodopsin triggers in rods a cascade of GTP-dependent and transducin-mediated reactions controlling cyclic-GMP hydrolysis. Infrared light-scattering studies (Kühn, H., N. Bennett, M. Michel-Villaz, and M. Chabre, 1981, Proc. Natl. Acad. Sci. USA, 78:6873-6877; Vuong, T. M., M. Chabre, and L. Stryer, 1984, Nature (Lond.), 311:659-661) demonstrated the existence of structural changes that correspond to this cascade rather than to rhodopsin photoactivation. We thus look for neutron diffraction changes of similar origins. With 1-min time resolution, intensity changes are observed mainly for orders 2 and 4. The illumination and GTP dependence of these changes indicates an involvement of transducin. Without GTP, they are linear with the amount of photoexcited rhodopsin, saturate at 10% photolysis, and thus correlate well with the light-scattering "binding signal." With GTP, light sensitivity is higher and saturation occurs below 0.5% photolysis, as for the "dissociation signal" of light scattering. In both cases, lattice compressions of 0.2-0.3% are observed. With 4-s time resolution the intensity change with GTP present precedes the lattice compression. The fast intensity change is probably due to the displacement of transducin alpha-subunits away from the disc membrane and the slower lattice shrinkage to an osmotic readjustment of the rod.  相似文献   

3.
The patch-clamp technique of Mueller (1975, Ann. N.Y. Acad. Sci., 274:247-264) and Neher and Sakmann (1976, Nature (Lond.), 260:799-802) was modified to be suitable for single-channel measurements in lipid bilayers at potentials up to 500 mV. This method was used to study gramicidin A single-channel current-voltage characteristics. It was found that the sublinear current-voltage behavior normally observed at low permeant ion concentrations and rather low potentials (V less than or equal to 200 mV) continues to be seen all the way up to 500 mV. This phenomenon is characteristic of the low permeant ion situation in which the channel is far from saturation, and implies that the overall rate constant for association between ion and channel is very weakly, if at all, voltage dependent. The magnitude of the single channel currents at 500 mV is consistent with the notion that the aqueous convergence conductance is a significant factor in determining the permeability characteristics of the gramicidin A channel.  相似文献   

4.
Ca channel gating during cardiac action potentials.   总被引:4,自引:2,他引:2       下载免费PDF全文
How do Ca channels conduct Ca ions during the cardiac action potential? We attempt to answer this question by applying a two-microelectrode technique, previously used for Na and K currents, in which we record the patch current and the action potential at the same time (Mazzanti, M., and L. J. DeFelice. 1987. Biophys. J. 12:95-100, and 1988. Biophys. J. 54:1139-1148; Wellis, D., L. J. DeFelice, and M. Mazzanti. 1990. Biophys. J. 57:41-48). In this paper, we also compare the action currents obtained by the technique with the step-protocol currents obtained during standard voltage-clamp experiments. Individual Ca channels were measured in 10 mM Ca/1 Ba and 10 mM Ba. To describe part of our results, we use the nomenclature introduced by Hess, P., J. B. Lansman, and R. W. Tsien (1984. Nature (Lond.). 311:538-544). With Ba as the charge carrier, Ca channel kinetics convert rapidly from long to short open times as the patch voltage changes from 20 to -20 mV. This voltage-dependent conversion occurs during action potentials and in step-protocol experiments. With Ca as the charge carrier, the currents are brief at all voltages, and it is difficult to define either the number of channels in the patch or the conductance of the individual channels. Occasionally, however, Ca-conducting channels spontaneously convert to long-open-time kinetics (in Hess et al., 1984, notation, mode 2).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
Glial cells in the nervous system are believed to reduce changes of extracellular potassium concentration ([K+]o), caused by neural activity, by carrying out spatial buffering of potassium. In the case of retinal glial cells (Müller cells), light-evoked increases of [K+]o within the retina are reduced by K ions flowing through the Müller cell to the vitreous fluid of the eye. We have calculated the optimal way to distribute the potassium conductance of the Müller cell to maximize spatial buffering to the vitreous fluid. The best distribution is with half the potassium conductance in the outer part of the cell, where K+ enters, and half the conductance in the vitreal endfoot, where K+ leaves the cell. This calculated distribution is very different from the actual distribution measured by Newman (1984, Nature [Lond.], 309: 155-157), where only 6% of the Müller cell conductance is in the outer cell and 94% is in the endfoot. The experimentally observed distribution gives less than a quarter of the spatial buffering that would be produced by the optimal distribution. The possible advantages of this arrangement are discussed.  相似文献   

6.
U-937 cells, a monocytic line derived from a human histiocytic lymphoma, were induced for human tumor necrosis factor (TNF) secretion into the medium and were used for the preparation of TNF mRNA. Biological activity of the latter was quantified in a Xenopus laevis oocyte injection system. TNF mRNA was enriched by gradient centrifugation and this size-fractionated mRNA was used for synthesis of cDNA and inserted into the unique PstI site of pAT153. A recombinant plasmid containing human TNF cDNA was selected by colony hybridization using an internal fragment of a mouse TNF cDNA clone [Fransen, L., Mueller, R., Marmenout, A., Tavernier, J., Van der Heyden, J., Kawashima, E., Chollet, A., Tizard, R., Van Heuverswyn, H., Van Vliet, A., Ruysschaert, M. R. & Fiers, W. (1985) Nucleic Acids Res. 13, 4417-4429] as a probe. The sequence of this human TNF cDNA is in agreement with the one published by Pennica et al. [Pennica, D., Nedwin, G. E., Hayflick, J. S., Seeburg, P. H., Derynck, R., Palladino, M. A., Kohr, W. J., Aggarwal, B. B. & Goeddel, D. V. (1984) Nature (Lond.) 312, 724-729]. The 157-amino-acid-long mature sequence is about 80% homologous to mouse TNF and its hydrophilicity plot is also very similar, in spite of the apparent species specificity of TNF. In contrast to mouse TNF, it contains no potential N-glycosylation site. When compared to other cytokines, like IFN-beta, IFN-gamma, or IL-2, there is a remarkably high preference for G X C pairs in the third-letter positions. Expression of the TNF cDNA in monkey COS cells or in Escherichia coli gives rise to a protein having similar biological and serological properties as natural human TNF. A human genomic clone was also identified and sequenced; it was found to be in good agreement with the one recently published by Shirai et al. [Shirai, T., Yamaguchi, H., Ito, H., Todd, C. W. & Wallace, R. B. (1985) Nature (Lond.) 313, 803-806], except for some differences in the introns and 5'-untranslated region.  相似文献   

7.
The cellular oncogene c-myc has been implicated in the regulation of growth of normal and neoplastic cells. Recently, it was suggested that c-myc gene expression may control the G0----G1-phase transition in normal lymphocytes that were stimulated to enter the cell cycle by the lectin concanavalin A (ConA). Here we describe the effects of purified recombinant interleukin 2 (rIL2) and of ConA on levels of c-myc mRNA in the noncytolytic murine T-cell clone L2. In contrast to resting (G0) primary cultures of lymphocytes, quiescent L2 cells have a higher RNA content than resting splenocytes and express receptors for interleukin 2 (IL2). Resting L2 cells are therefore best regarded as early G1-phase cells. Purified rIL2 was found to stimulate the rapid accumulation of c-myc mRNA in L2 cells. Levels of c-myc mRNA became maximal within 1 h and declined gradually thereafter. In contrast, ConA induced slower accumulation of c-myc mRNA in L2 cells, with increased levels of c-myc mRNA becoming detectable 4 to 8 h after stimulation. Experiments with the protein synthesis inhibitor cycloheximide demonstrated that the increase in levels of c-myc mRNA that were induced by ConA was a direct effect of this lectin and not secondary to IL2 production. Cyclosporin A, an immunosuppressive agent, markedly reduced the accumulation of c-myc mRNA that was induced by ConA but only slightly diminished the accumulation of c-myc mRNA that was induced by rIL2. Taken together, these data provide evidence that (i) c-myc gene expression can be regulated by at least two distinct pathways in T lymphocytes, only one of which is sensitive to cyclosporine A, and (ii) the accumulation of c-myc mRNA can be induced in T cells by IL2 during the G1 phase of the cell cycle.  相似文献   

8.
Hexabrachion proteins in embryonic chicken tissues and human tumors   总被引:19,自引:11,他引:8       下载免费PDF全文
Cell cultures of chicken embryo and human fibroblasts produce a large extracellular matrix molecule with a six-armed structure that we called a hexabrachion (Erickson, H. P., and J. L. Iglesias, 1984, Nature (Lond.), 311:267-269. In the present work we have determined that the myotendinous (M1) antigen described by M. Chiquet and D. M. Fambrough in chicken tissues (1984, J. Cell Biol., 98:1926-1936), and the glioma mesenchymal extracellular matrix protein described by Bourdon et al. in human tumors (Bourdon, M. A., C. J. Wikstrand, H. Furthmayr, T. J. Matthews, and D. D. Bigner, 1983, Cancer Res. 43:2796-2805) have the structure of hexabrachions. We also demonstrate that the M1 antigen is present in embryonic brain, where it was previously reported absent, and have purified hexabrachions from brain homogenates. The recently described cytotactin (Grumet, M., S. Hoffman, K. L. Crossin, and G. M. Edelman, 1985, Proc. Natl. Acad. Sci. USA, 82:8075-8079) now appears to be identical to the chicken hexabrachion protein. In a search for functional roles, we looked for a possible cell attachment activity. A strong, fibronectin-like attachment activity was present in (NH4)2SO4 precipitates of cell supernatant and sedimented with hexabrachions in glycerol gradients. Hexabrachions purified by antibody adsorption, however, had lost this activity, suggesting that it was due to a separate factor associated with hexabrachions in the gradient fractions. The combined information in the several, previously unrelated studies suggests that hexabrachions may play a role in organizing localized regions of extracellular matrix. The protein is prominently expressed at specific times and locations during embryonic development, is retained in certain adult tissues, and is reexpressed in a variety of tumors.  相似文献   

9.
This study demonstrates possible ways to estimate the rate constants of reaction kinetic models for ion transport from steady-state current-voltage data as measured at various substrate concentrations. This issue is treated theoretically by algebraic reduction and extension of a reaction kinetic four-state model for uniport. Furthermore, an example for application is given; current-voltage data from an open K+ selective channel (Schroeder, J.I., R. Hedrich, and J.M. Fernandez, 1984, Nature (Lond.), 312:361-362) supplemented by some new data have been evaluated. The analysis yields absolute numerical estimates of the 14 rate constants of a six-state model, which is discussed in a wider context.  相似文献   

10.
Low conductance sodium channels in canine cardiac Purkinje cells.   总被引:4,自引:0,他引:4       下载免费PDF全文
Low conductance sodium (Na) channels have been observed in nerve, skeletal muscle, and cardiac cells. In cardiac tissues the higher amplitude, more commonly observed Na channel was first investigated in detail by Cachelin et al. (Cachelin, A.B., J.E. de Peyer, S. Kokubun, and H. Reuter, 1983, J. Physiol. (Lond.), 340:389-402). They also reported low amplitude Na channel events. We have studied this low conductance Na channel in single canine cardiac Purkinje cells using cell-attached patches. Patch pipette solutions contained either 140 or 280 mM NaCl, and cells were bathed in a solution of 150 mM KCl to bring their resting potential close to zero. In 140 mM Na+, during steps to -50 mV, the lower and higher openings had amplitudes of 0.57 +/- 0.2 and 1.2 +/- 0.2 pA (means +/- SD of Gaussian fits). In 280 mM Na+ at -50 mV, amplitudes were 0.72 +/- 0.2 and 1.55 +/- 0.2 pA. Over a substantial voltage range, the lower events had amplitudes of about one-third that of the higher events. The frequency of the low conductance openings varied in different patches from zero to 22% of total openings. Histograms of open durations and latencies at several voltages suggested no difference in kinetics between the two channel events. The behavior of the low conductance channels was more consistent with a second population of channels rather than a second open state.  相似文献   

11.
Isolated Ca currents in cultured dorsal root ganglion (DRG) cells were studied using the patch clamp technique. The currents persisted in the presence of 30 microM tetrodotoxin (TTX) or when external Na was replaced by choline. They were fully blocked by millimolar additions of Cd2+ and Ni2+ to the bath. Two components of an inward-going Ca current were observed. In 5 mM external Ca, a current of small amplitude, turned on already during steps changes to -60 mV membrane potential, leveled off at -30 mV to a value of approximately 0.2 nA. A second, larger current component, which resembled the previously described Ca current in other cells, appeared at more positive voltages (-20 to -10 mV) and had a maximum approximately 0 mV. The current component activated at the more negative membrane potentials showed the stronger dependence on external Ca. The presence of a time- and a voltage-dependent activation was indicated by the current's sigmoidal rise, which became faster with increased depolarization. Its tail currents were generally slower than those associated with the Ca currents of larger amplitude. From -60 mV holding potential, the maximum obtainable amplitude of the low depolarization-activated current was only one-tenth of that achieved from a holding potential of -90 mV. Voltage-dependent inactivation of this current component was fast compared with that of the other component. The properties of this low voltage-activated and fully inactivating Ca current suggest it is the same as the inward current that has been postulated in several central neurons (Llinas, R., and Y. Yarom, 1981, J. Physiol. (Lond.), 315:569-584), which produce depolarizing potential waves and burst-firing only when membrane hyperpolarization precedes.  相似文献   

12.
Potassium channels are membrane-spanning proteins with several transmembrane segments and a single pore region where ion conduction takes place (Biggin, P. C., Roosild, T., and Choe, S. (2000) Curr. Opin. Struct. Biol. 4, 456-461; Doyle, D. A., Morais Cabral, J., Pfuetzner, R. A., Kuo, A., Gulbis, J. M., Cohen, S. L., Chait, B. T., and MacKinnon, R. (1998) Science 280, 69-77). TOK1, a potassium channel identified in the yeast Saccharomyces cerevisiae, was the first described member from a growing new family of potassium channels with two pore domains in tandem (2P) (Ketchum, K. A., Joiner, W. J., Sellers, A. J., Kaczmarek, L. K., and Goldstein, S. A. (1995) Nature 376, 690-695). In an attempt to understand the relative contribution of each one of the 2P from TOK1 to the functional properties of this channel, we split and expressed the pore domains separately or in combination. Expression of the two domains separately rescued a potassium transport-deficient yeast mutant, suggesting that each domain forms functional potassium-permeable channels in yeast. In Xenopus laevis oocytes expression of each pore domain resulted in the appearance of unique inwardly rectifying cationic channels with novel gating and pharmacological properties. Both pore domains were poorly selective to potassium; however, upon co-expression they partially restored TOK1 channel selectivity. The single channel conductance was different in both pore domains with 7 +/- 1 (n = 12) and 15 +/- 2 (n = 12) picosiemens for the first and second domain, respectively. In light of the known structure of the Streptomyces lividans KcsA potassium channel pore (see Doyle et al. above), these results suggest a novel non-four-fold-symmetric architecture for 2P potassium-selective channels.  相似文献   

13.
Vincristine-resistant (VCR) Chinese hamster ovary (CHO) cells have been established by stepwise selection in increasing concentrations of vincristine. These cells exhibit multidrug cross-resistance to a number of drugs that have no structural or functional similarities. Cytogenetic analyses of resistant cells revealed the presence of double minutes and expanded chromosomal segments, thus implicating gene amplification as a possible mechanism of resistance. An amplified DNA segment isolated from other multidrug cross-resistant CHO cell lines (Roninson, I. B., H. T. Abelson, D. E. Housman, N. Howell, and A. Varshavsky, 1984, Nature (Lond.), 309:626-628) is also amplified in our VCR lines. This DNA segment was used as a probe to screen a cosmid library of VCR genomic DNA, and overlapping clones were retrieved. All of these segments, totaling approximately 45 kilobases (kb), were amplified in VCR cells. Using in situ hybridization, we localized the amplification domain to the long arm of CHO chromosome 1 or Z1. Northern hybridization analysis revealed that a 4.3-kb mRNA was encoded by this amplified DNA domain and was over-produced in the VCR cells. Suggestions for the involvement of these amplified DNA segments in the acquisition of multidrug cross-resistance in animal cells are also presented.  相似文献   

14.
Axoplasmic vesicles were purified and observed to translocate on isolated microtubules in an ATP-dependent, trypsin-sensitive manner, implying that ATP-binding polypeptides essential for force generation were present on the vesicle surface. To identify these proteins [alpha 32P]8-azidoadenosine 5'-triphosphate ([alpha 32P]8-N3ATP), a photoaffinity analogue of ATP, was used. The results presented here identify and characterize a vesicle-associated polypeptide having a relative molecular mass of 292 kD that bound [alpha 32P]8-N3ATP. The incorporation of label is ultraviolet light-dependent and ATP-sensitive. Moreover, the 292-kD polypeptide could be isolated in association with vesicles or microtubules, depending on the conditions used, and the data indicate that the 292-kD polypeptide is similar to mammalian brain microtubule-associated protein 2 (MAP 2) for the following reasons: The 292-kD polypeptide isolated from either squid axoplasm or optic lobe cross-reacts with antiserum to porcine brain MAP 2. Furthermore, it purifies with taxol-stabilized microtubules and is released with salt. Based on these characteristics, the 292-kD polypeptide is distinct from the known force-generating molecules myosin and flagellar dynein, as well as the 110-130-kD kinesin-like polypeptides that have recently been described (Brady, S. T., 1985, Nature (Lond.), 317:73-75; Vale, R. D., T. S. Reese, and M. P. Sheetz, 1985b, Cell, 42:39-50; Scholey, J. M., M. E. Porter, P. M. Grissom, and J. R. McIntosh, 1985, Nature (Lond.), 318:483-486). Because the 292-kD polypeptide binds ATP and is associated with vesicles that translocate on purified MAP-free microtubules in an ATP-dependent fashion, it is therefore believed to be involved in vesicle-microtubule interactions that promote organelle motility.  相似文献   

15.
cDNAs for human interleukin-2 receptor were recently cloned and sequenced (Leonard et al., 1984, Nature 311, 626-631; Nikaido et al., 1984, Nature 311, 631-635; Cosman et al., Nature 312, 768-771). In the studies reported here, we describe the expression of a cDNA clone for the human interleukin-2 receptor in E. coli using an "open reading frame" expression vector pMR100. The inserted cDNA was expressed in E. coli transformants as a tripartite fusion polypeptide fused to the lambda cI protein at its amino terminus and to beta-galactosidase at its carboxy terminus. We demonstrate that the bacterially produced IL-2 receptor protein can bind to IL-2.  相似文献   

16.
Synaptic vesicle recycling in synapsin I knock-out mice   总被引:3,自引:0,他引:3       下载免费PDF全文
《The Journal of cell biology》1996,134(5):1219-1227
The synapsins are a family of four neuron-specific phosphoproteins that have been implicated in the regulation of neurotransmitter release. Nevertheless, knock-out mice lacking synapsin Ia and Ib, family members that are major substrates for cAMP and Ca2+/ Calmodulin (CaM)-dependent protein kinases, show limited phenotypic changes when analyzed electrophysiologically (Rosahl, T.W., D. Spillane, M. Missler, J. Herz, D.K. Selig, J.R. Wolff, R.E. Hammer, R.C. Malenka, and T.C. Sudhof. 1995. Nature (Lond.). 375: 488-493; Rosahl, T.W., M. Geppert, D. Spillane, D., J. Herz, R.E. Hammer, R.C. Malenka, and T.C. Sudhof. 1993. Cell. 75:661-670; Li, L., L.S. Chin, O. Shupliakov, L. Brodin, T.S. Sihra, O. Hvalby, V. Jensen, D. Zheng, J.O. McNamara, P. Greengard, and P. Andersen. 1995. Proc. Natl. Acad. Sci. USA. 92:9235- 9239; see also Pieribone, V.A., O. Shupliakov, L. Brodin, S. Hilfiker- Rothenfluh, A.J. Czernik, and P. Greengard. 1995. Nature (Lond.). 375:493-497). Here, using the optical tracer FM 1-43, we characterize the details of synaptic vesicle recycling at individual synaptic boutons in hippocampal cell cultures derived from mice lacking synapsin I or wild-type equivalents. These studies show that both the number of vesicles exocytosed during brief action potential trains and the total recycling vesicle pool are significantly reduced in the synapsin I- deficient mice, while the kinetics of endocytosis and synaptic vesicle repriming appear normal.  相似文献   

17.
Taxol, a microtubule stabilizing drug, induces the formation of numerous microtubule asters in the cytoplasm of mitotic cells (De Brabander, M., G. Geuens, R. Nuydens, R. Willebrords, J. DeMey. 1981. Proc. Natl. Acad. Sci. USA. 78:5608-5612). The center of these asters share with spindle poles some characteristics such as the presence of centrosomal material and calmodulin. We have recently reproduced the assembly of taxol asters in a cell-free system (Buendia, B., C. Antony, F. Verde, M. Bornens, and E. Karsenti. 1990. J. Cell Sci. 97:259-271) using extracts of Xenopus eggs. In this paper, we show that taxol aster assembly requires phosphorylation, and that they do not grow from preformed centers, but rather by a reorganization of microtubules first crosslinked into bundles. This process seems to involve sliding of microtubules along each other and we show that cytoplasmic dynein is required for taxol aster assembly. This result provides a possible functional basis to the recent findings, that dynein is present in the spindle and enriched near spindle poles (Pfarr, C. M., M. Cove, P. M. Grissom, T. S. Hays, M. E. Porter, and J. R. McIntosh. 1990. Nature (Lond.). 345:263-265; Steuer, E. R., L. Wordeman, T. A. Schroer, and M. P. Sheetz. 1990. Nature (Lond.). 345:266-268).  相似文献   

18.
M-1 cells, derived from a microdissected cortical collecting duct of a transgenic mouse, grown to confluence on a permeable support, develop a lumen-negative amiloride-sensitive transepithelial potential, reabsorb sodium, and secrete potassium. Electron micrographs show morphological features typical of principal cells in vivo. Using the patch clamp technique distinct differences are detected in whole-cell membrane current and voltage (Vm) between single M-1 cells 24 h after seeding vs cells grown to confluence. (a) Under control conditions (pipette: KCl- Ringer; bath: NaCl-Ringer) Vm averages -42.7 +/- 3.4 mV in single cells vs -16.8 +/- 4.1 mV in confluent cells. Whole-cell conductance (Gcell) in confluent cells is 2.6 times higher than in single cells. Cell capacitance values are not significantly different in single vs confluent M-1 cells, arguing against electrical coupling of confluent M- 1 cells. (b) In confluent cells, 10(-4)-10(-5) M amiloride hyperpolarizes Vm to -39.7 +/- 3.0 mV and the amiloride-sensitive fractional conductance of 0.31 shows a sodium to potassium selectivity ratio of approximately 15. In contrast, single cells express no significant amiloride-sensitive conductance. (c) In single M-1 cells, Gcell is dominated by an inwardly rectifying K-conductance, as exposure to high bath K causes a large depolarization and doubling of Gcell. The barium-sensitive fraction of Gcell in symmetrical KCl-Ringer is 0.49 and voltage dependent. (d) In contrast, neither high K nor barium in the apical bath affect confluent M-1 cells, showing that confluent cells lack a significant apical K conductance. (e) Application of 500 microM glibenclamide reduces whole-cell currents in both single and confluent M-1 cells with a glibenclamide-sensitive fractional conductance of 0.71 and 0.83 in single and confluent cells, respectively. Glibenclamide inhibition occurs slower in confluent M-1 cells than in single cells, suggesting a basolateral action of this lipophilic drug on ATP-sensitive basolateral K channels in M-1 cells. (f) A component of the whole-cell conductance in M-1 cells appears as a deactivating outward current during large depolarizing voltage pulses and is abolished by extracellular chloride removal. The deactivating chloride current averages 103.6 +/- 16.1 pA/cell, comprises 24% of the outward current, and decays with a time constant of 179 +/- 13 ms. The outward to inward conductance ratio obtained from deactivating currents and tail currents is 2.4, indicating an outwardly rectifying chloride conductance.  相似文献   

19.
Mechanisms underlying action potential generation in the newt olfactory receptor cell were investigated by using the whole-cell version of the patch-clamp technique. Isolated olfactory cells had a resting membrane potential of -70 +/- 9 mV. Injection of a depolarizing current step triggered action potentials under current clamp condition. The amplitude of the action potential was reduced by lowering external Na+ concentration. After a complete removal of Na+, however, cells still showed action potentials which was abolished either by Ca2+ removal or by an application of Ca2+ channel blocker (Co2+ or Ni2+), indicating an involvement of Ca2+ current in spike generation of newt olfactory receptor cells. Under the voltage clamp condition, depolarization of the cell to -40 mV from the holding voltage of -100 mV induced a fast transient inward current, which consisted of Na+ (INa) and T-type Ca2+ (ICa.T) currents. The amplitude of ICa,T was about one fourth of that of INa. Depolarization to more positive voltages also induced L-type Ca2+ current (ICa,L). ICa,L was as small as a few pA in normal Ringer solution. The activating voltage of ICa,T was approximately 10 mV more negative than that of INa. Under current clamp, action potentials generated by a least effective depolarization was almost completely blocked by 0.1 mM Ni2+ (a specific T-type Ca2+ channel blocker) even in the presence of Na+. These results suggest that ICa,T contributes to action potential in the newt olfactory receptor cell and lowers the threshold of spike generation.  相似文献   

20.
The effects of quinidine on single inward rectifier K channels were investigated in cell-attached patches with 4.5 mM pipette potassium concentrations. Under these conditions, the single-channel slope conductance of the predominant conductance level of the inward rectifier channels was 3.9 +/- 0.3 pS at membrane potentials between -75 and -150 mV. Quinidine reversibly decreased the likelihood of channel opening to the main conductance level without reducing the single-channel conductance, and also reduced the probability of channel opening to subconducting levels. Quinidine had no significant effects on the channel open times, and the inhibition of channel opening was only slightly voltage dependent over the range of membrane potentials investigated. Quinidine induced a complete cessation of channel openings for brief periods (up to 2 min), suggesting that quinidine promoted occupancy of a state from which opening was less likely. Occasional long periods (up to an hour) with an absence of channel activity were also observed but quinidine did not appear to promote this behavior. The data suggest that quinidine decreases the ability of the channel to enter both main and subconducting states. By binding to a particular closed conformation of the channel, quinidine could reduce the likelihood of channel opening. The main features of these observations could be accounted for using the three-state kinetic model proposed by Sakmann, B. and G. Trube (1984b. J. Physiol. [Lond.]. 347:659-683.) with quinidine binding to the middle closed 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