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 39 毫秒
1.
The cellular electrical activity of diaphragm from F1B normal and BIO 14.6 dystrophic hamsters has been investigated using microelectrodes. Resting membrane potentials and action potentials were recorded from control muscles and from muscles exposed to 2,4-dinitrophenol. The action potentials of normal and dystrophic diaphragms were similar in amplitude and configuration. Treatment with 2,4-dinitrophenol caused the action potential amplitude of both diaphragms to decline by similar amounts. The control resting membrane potential of diaphragm from dystrophic hamsters is not significantly different from that of normal hamsters. Treatment with 2,4-dinitrophenol caused a linear decrease in the resting membrane potentials of both groups of muscles. Dystrophic muscle, however, showed a more rapid decline in excitability when exposed to 2,4-dinitrophenol. This suggests that adenosine triphosphate production in dystrophic muscle is partially inhibited as has been suggested by other workers.  相似文献   

2.
Comparisons between electrotronic potentials and certain predicted curves allow the identification of the membrane potential at which the sodium and potassium currents are switched on in frog sartorius. The activation potentials (the membrane potentials at which the ionic currents are great enough to be resolved by the method) are functions of the resting potential and time but not of ionic concentration. In the normal fiber, the activation potential for sodium lies nearer the resting potential and depolarizations set off sodium currents and action potentials. Below a resting potential of 55 to 60 mv. sodium activation is lost and conduction is impossible. A tenfold increase of calcium concentration lowers (moves further from the resting potential) the sodium activation potential by 20 to 25 mv. whereas the potassium activation potential is lowered by only 15 mv. Certain consequences of this are seen in the behavior of the muscle cell when it is stimulated with long duration shock.  相似文献   

3.
Ion channel proteins are required for both the establishment of resting membrane potentials and the generation of action potentials. Hundreds of mutations in genes encoding voltage-gated ion channels responsible for action potential generation have been found to cause severe neurological diseases. In contrast, the roles of voltage-independent “leak” channels, important for the establishment and maintenance of resting membrane potentials upon which action potentials are generated, are not well established in human disease. UNC80 is a large component of the NALCN sodium-leak channel complex that regulates the basal excitability of the nervous system. Loss-of-function mutations of NALCN cause infantile hypotonia with psychomotor retardation and characteristic facies (IHPRF). We report four individuals from three unrelated families who have homozygous missense or compound heterozygous truncating mutations in UNC80 and persistent hypotonia, encephalopathy, growth failure, and severe intellectual disability. Compared to control cells, HEK293T cells transfected with an expression plasmid containing the c.5098C>T (p.Pro1700Ser) UNC80 mutation found in one individual showed markedly decreased NALCN channel currents. Our findings demonstrate the fundamental significance of UNC80 and basal ionic conductance to human health.  相似文献   

4.
Temperature increases cause a regular and reproducible increase in the frequency of generation of pacemaker potentials in most Aplysia neurons specialized for this type of activity which can only be explained as a direct stimulating effect of temperature upon the ionic mechanisms responsible for pacemaker potentials. At the same time all cells in the visceral ganglion undergo a membrane potential hyperpolarization of approximately 1–2 mv/°C warmed. In spite of the marked variation in resting membrane potential the critical firing threshold remains at a constant membrane potential level at all temperatures in the absence of accommodative changes. The temperature-frequency curves of all types of cells are interpreted as a result of the interaction between the effects of temperature on the pacemaker-generating mechanism and resting membrane potential. Previous observations on the effects of temperature on excitability of mammalian neurons suggest that other types of neurons may undergo similar marked shifts in resting membrane potential with temperature variation.  相似文献   

5.
Transmembrane potentials in the crayfish giant axon have been investigated as a function of the concentration of normally occurring external cations. Results have been compared with data already available for the lobster and squid giant axons. The magnitude of the action potential was shown to be a linear function of the log of the external sodium concentration, as would be predicted for an ideal sodium electrode. The resting potential is an inverse function of the external potassium concentration, but behaves as an ideal potassium electrode only at the higher external concentrations of potassium. Decrease in external calcium results in a decrease in both resting potential and action potential; an increase in external calcium above normal has no effect on magnitude of transmembrane potentials. Magnesium can partially substitute for calcium in the maintenance of normal action potential magnitude, but appears to have very little effect on resting potential. All ionic effects studied are completely reversible. The results are in generally good agreement with data presently available for the lobster giant axon and for the squid giant axon.  相似文献   

6.
The delayed and asynchronous firing of chromatolytic motoneurons in response to group I afferent volleys is shown to be evoked monosynaptically, there being an abnormally long and variable delay between onset of monosynaptic action and generation of impulse discharge. Intensity of monosynaptic excitatory action is reduced, and considerable variability in the form of successively evoked postsynaptic potentials is often observed. No evidence has been found for the development of excitatory group I polysynaptic pathways. Reduction in responsiveness of finer dendrites is indicated by the feeble "d" response evoked by an antidromic volley in a chromatolytic motor nucleus. Antidromic impulses appear to invade the cell bodies and coarse dendrites, but die out at points short of the normal extent of dendritic invasion. Vigorous firing of Renshaw cells can be elicited by antidromic volleys. Chromatolytic motoneurons appear to maintain reasonably normal resting membrane potentials, but are more susceptible to damage than are normal cells. Action potentials are large and usually overshoot the resting potential level. Post spike potentials are similar to those of normal cells except for a less prominent, or absent, early phase of depolarisation. In contrast with the reduced responsiveness of peripheral dendrites, there is a lowered threshold for antidromic and segmental reflex synaptic activation of the more central regions, probably the cell bodies and nearby coarse dendrites, of motoneurons undergoing chromatolysis.  相似文献   

7.
《Biophysical journal》1998,75(3):1483-1490
The effect of ionic strength on the macroscopic and microscopic redox potentials and the heme environment of cytochrome c3 from Desulfovibrio vulgaris Miyazaki F have been investigated by NMR and electrochemical methods. The redox potentials of this tetraheme protein are found to be ionic strength-dependent. Especially, the microscopic redox potentials of hemes 2 and 3 at the fourth reduction step increase significantly with increasing ionic strength, which is in contradiction to the theoretical expectation. The coordinated imidazole proton signals are unaffected by ionic strength. However, the methyl and propionate proton signals of hemes 1 and 4 showed significant ionic strength dependencies that are distinct from those for hemes 2 and 3. This heme classification is the same as that found in the ionic strength dependencies of the microscopic redox potentials at the fourth reduction step. Furthermore, the effect of ionic strength on the electrostatic potentials at the heme irons has been examined on the theoretical basis. The electrostatic potential at heme 4 changes up to 1 M ionic strength, which was not expected from the observations reported on cytochromes so far. These results are discussed in connection with the reported anomalous ionic strength dependency of the reduction rate of cytochrome c3.  相似文献   

8.
Electrophysiology of an Insect Heart   总被引:1,自引:0,他引:1       下载免费PDF全文
Bioelectric activity in single cells of the moth myocardium has been measured in situ with intracellular microelectrodes with particular reference to the bizarre ionic medium which bathes the tissues. Resting potentials averaged 47 mv, inside negative with respect to outside, despite a value of 11 mv calculated on the basis of transmembrane potassium concentration gradients. Action potentials overshoot as much as 12 mv in the virtual absence of extracellular sodium. Two "types" of action potentials have been recorded; one that resembles vertebrate atrial action potentials is found in the cephalic region of the tubular heart, and the other, similar in contour to vertebrate ventricular action potentials, is found in the areas posterior to the first abdominal segment. Histological sections indicate no structural differences between the two areas. Typical cardiac pacemaker type potentials occur but are not topographically localized. The effects of the omission from the perfusion fluid of the four major cationic constituents, Na+, K+, Ca++ and Mg++, on resting and action potentials may be summarized as: no effect, hyperpolarization, prolonged repolarization, and depolarization, respectively.  相似文献   

9.
Interpreting channel behavior in patches requires an understanding of patch structure and dynamics, especially in studies of mechanosensitive channels. High resolution optical studies show that patch formation occurs via blebbing that disrupts normal membrane structure and redistributes in situ components including ion channels. There is a 1-2 μm region of the seal below the patch where proteins are excluded and this may consist of extracted lipids that form the gigaseal. Patch domes often have complex geometries with inhomogeneous stresses due to the membrane-glass adhesion energy (Ea), cytoskeletal forces, and possible lipid subdomains. The resting tension in the patch dome ranges from 1-4 mN/m, a significant fraction of the lytic tension of a bilayer (∼10 mN/m). Thus, all patch experiments are conducted under substantial, and uneven, resting tension that may alter the kinetics of many channels. Ea seems dominated by van der Waals attraction overlaid with a normally repulsive Coulombic force. High ionic strength pipette saline increased Ea and, surprisingly, increased cytoskeletal rigidity in cell-attached patches. Low pH pipette saline also increased Ea and reduced the seal selectivity for cations, presumably by neutralizing the membrane surface charge. The seal is a negatively charged, cation selective, space with a resistance of ∼7 gigohm/μm in 100 mM KCl, and the high resistivity of the space may result from the presence of high viscosity glycoproteins. Patches creep up the pipette over time with voltage independent and voltage dependent components. Voltage-independent creep is expected from the capillary attraction of Ea and the flow of fresh lipids from the cell. Voltage-dependent creep seems to arise from electroosmosis in the seal. Neutralization of negative charges on the seal membrane with low pH decreased the creep rate and reversed the direction of creep at positive pipette potentials.  相似文献   

10.
On the resting potential of isolated frog sympathetic neurons   总被引:4,自引:0,他引:4  
S W Jones 《Neuron》1989,3(2):153-161
One of the oldest questions of electrophysiology, the origin of the resting potential, has yet to be answered satisfactorily for most cells. Isolated frog sympathetic neurons, studied with whole-cell recording, generally have resting potentials of approximately -75 mV with an input resistance of approximately 300 M omega. These properties are not expected from the M-type K+ current (IM) or from other ionic currents previously described in these cells. In the -60 to -110 M mV voltage region, at least three currents are present: an inwardly rectifying current (IQ), a resting current with little voltage sensitivity carried at least in part by K+, and a (Na+,K+)ATPase pump current. The resting K+ current, not IM or IQ is the primary ionic current near the resting potential under these conditions. The electrogenic pump contributes an additional approximately 10 mV of hyperpolarization.  相似文献   

11.
Increased calcium influx in dystrophic muscle   总被引:16,自引:0,他引:16  
We examined pathways which might result in the elevated resting free calcium [( Ca2+]i) levels observed in dystrophic mouse (mdx) skeletal muscle fibers and myotubes and human Duchenne muscular dystrophy myotubes. We found that mdx fibers, loaded with the calcium indicator fura-2, were less able to regulate [Ca2+]i levels in the region near the sarcolemma. Increased calcium influx or decreased efflux could lead to elevated [Ca2+]i levels. Calcium transient decay times were identical in normal and mdx fibers if resting [Ca2+]i levels were similar, suggesting that calcium-sequestering mechanisms are not altered in dystrophic muscle, but are slowed by the higher resting [Ca2+]i. The defect appears to be specific for calcium since resting free sodium levels and sodium influx rates in the absence of Na+/K(+)-ATPase activity were identical in normal and dystrophic cells when measured with sodium-binding benzofuran isophthalate. Calcium leak channels, whose opening probabilities (Po) were voltage independent, could be the major calcium influx pathway at rest. We have shown previously that calcium leak channel Po is significantly higher in dystrophic myotubes. These leak channels were selective for calcium over sodium under physiological conditions. Agents that increased leak channel activity also increased [Ca2+]i in fibers and myotubes. These results suggest that increased calcium influx, as a result of increased leak channel activity, could result in the elevated [Ca2+]i in dystrophic muscle.  相似文献   

12.
Electrical activity in the fertilized egg of the tunicate Clavelina was studied with microelectrode recording and voltage clamp techniques. The resting potential could assume either of two stable values (approximately ?70 or ?30 mV) and could be shifted between these values by direct current stimulation. Spontaneous shifts between two stable resting potentials were also seen. Egg cells produced action potentials spontaneously and in response to depolarizing stimuli. Inward currents were carried by both Na and Ca ions and a prominent outward potassium current was seen with depolarization to voltages above ?15 mV. The steady-state current-voltage relationship (I–V curve) of the membrane showed two voltages where the net membrane current equaled zero: approximately ?35 and ?70 mV. Between these two voltages, membrane current was inward and carried by noninactivating Na and Ca currents. Inward rectification, which was blocked by external Rb, occurred at voltages below ?70 mV. The voltage dependence of inward rectification is thought by the authors to be important for establishing the more negative resting potential; it is also thought the presence of inward current which does not inactivate completely at voltages more negative than about ?20 mV is an important determinant of the more depolarized resting potential.  相似文献   

13.
Myofibrillar protein degradation was measured in 4-week-old normal (line 412) and genetically muscular-dystrophic (line 413) New Hampshire chickens by monitoring the rates of 3-methylhistidine excretion in vivo and in vitro. A method of perfusing breast and wing muscles was developed and the rate of 3-methylhistidine release in vitro was measured between 30 and 90min of perfusion. During this perfusion period, 3-methylhistidine release from the muscle preparation was linear, indicating that changes in 3-methylhistidine concentration of the perfusate were the result of myofibrillar protein degradation. Furthermore, the viability of the perfused muscle was maintained during this interval. After 60min of perfusion, ATP, ADP and creatine phosphate concentrations in pectoral muscle were similar to muscle freeze-clamped in vivo. Rates of glucose uptake and lactate production were constant during the perfusion. In dystrophic-muscle preparations, the rate of 3-methylhistidine release in vitro (nmol/h per g of dried muscle) was elevated 2-fold when compared with that in normal muscle. From these data the fractional degradation rates of myofibrillar protein in normal and dystrophic pectoral muscle were calculated to be 12 and 24% respectively. Daily 3-methylhistidine excretion (nmol/day per g body wt.) in vivo was elevated 1.35-fold in dystrophic chickens. Additional studies revealed that the anti-dystrophic drugs diphenylhydantoin and methylsergide, which improve righting ability of dystrophic chickens, did not alter 3-methylhistidine release in vitro. This result implies that changes in myofibrillar protein turnover are not the primary lesion in avian muscular dystrophy. From tissue amino acid analysis, the myofibrillar 3-methylhistidine content per g dry weight of muscle was similar in normal and dystrophic pectoral muscle. More than 96% of the 3-methylhistidine present in pectoral muscle was associated with the myofibrillar fraction. Dystrophic myofibrillar protein contained significantly less 3-methylhistidine (nmol/g of myofibrillar protein) than protein from normal muscle. This observation supports the hypothesis that there may be a block in the biochemical maturation and development of dystrophic muscle after hatching. Free 3-methylhistidine (nmol/g wet wt.) was elevated in dystrophic muscle, whereas blood 3-methylhistidine concentrations were similar in both lines. In summary, the increased myofibrillar protein catabolism demonstrated in dystrophic pectoral muscle correlates with the increased lysosomal cathepsin activity in this tissue as reported by others.  相似文献   

14.
1. A column procedure for the purification of creatine kinase from normal and dystrophic mouse muscle is described. 2. The native enzymes are indistinguishable by various physical criteria and have mol.wt. about 80000. 3. The purified enzyme from dystrophic muscle is only half as active as the normal, contains only one thiol group readily alkylated by iodoacetamide instead of two and has one less free thiol group/mol. 4. Michaelis constants for MgATP and creatine are the same for both preparations. 5. The inhibitor constant for ADP at pH9·0 is different in the two enzymes and this may account for the different degrees of inhibition observed in vitro with the drug Laevadosin. 6. The enzyme from dystrophic muscle is protected by an equilibrium mixture of substrates against inhibition by iodoacetamide to a greater extent than the normal enzyme. 7. `Fingerprinting' suggests one peptide difference between creatine kinases from normal and dystrophic muscle. 8. The possibility that this finding represents the primary lesion in dystrophy is discussed.  相似文献   

15.
A modified Hodgkin & Huxley (1952) model for axons was used to simulate smooth muscle action potentials. The modifications were such as to match our own experimental results and published data on the passive and active behavior of smooth muscle.A brief account of the modifications introduced to the HH model is as follows. The resting ionic conductances were obtained from the data of Casteels (1969). Chloride conductance was replaced by an ad hoc leakage conductance (g?L) in order to obtain a resting membrane resistance of about 11 kΩcm2. The ionic equilibrium potentials were according to Kao & Nishiyama (1969). The rate constants m, n and h have similar form to those in axons, but their different numerical values produce action potentials that match the duration of the smooth muscle action potential (about 16 ms) at half its maximum amplitude. The effective membrane capacitance was taken as 2.5 μF/cm2.The results obtained by implementing those smooth muscle parameters in the HH formulation include: (a) a membrane potential that matches the main characteristics of experimentally recorded action potentials in uterine smooth muscle and guinea-pig taenia-coli, and (b) a propagated action potential which, on a cable diameter of 5 μm (similar to the diameter of a single smooth muscle cell), has a speed of propagation within the range of the values experimentally recorded in smooth muscle. This observed velocity of propagation is not compatible with a large cable and it is concluded that “functional units” are not required to sustain propagation of action potentials in smooth muscle.  相似文献   

16.
Studies have shown that there is an abnormality in the thymus of dystrophic mice with respect to age-dependent thymus weight changes and altered morphology (T. DeKretser and B. Livett, Nature (London), 263, 682, 1976). Recently, others have shown that natural killer (NK) cells can lyse cells of a large, immature, rapidly dividing cell subpopulation within the thymus of normal young (3 weeks of age) mice (M. Hansson, K. Karre, R. Kiessling, J. Roder, B. Anderson, and P. Hayry, J. Immunol., 123, 765, 1979). The NK susceptibility of dystrophic mouse thymocytes as targets was therefore studied. Spleen cells from normal (+/+) and dystrophic (dy2Jdy2J) male C57BL/6J mice 8–10 weeks old were passed over nylon wool and the nonadherent cells were incubated with 51Cr-labeled YAC-1 lymphoma target cells or thymocytes in a 51Cr-release assay. Spleen cells from dystrophic mice killed twofold more YAC-1 target cells than did spleen cells from normal mice. Thymocytes from 3- to 4-week-old dystrophic mice were three to four times more susceptible to NK lysis by dystrophic mouse spleen cells as compared with normal mouse spleen cells. Spleen cells from dystrophic mice had the same NK activity against dystrophic and normal mouse thymocytes as targets. Normal mouse spleen cells killed three- to fourfold more dystrophic mouse thymocytes than that of normal mouse thymocytes as targets. Target cellbinding studies revealed that conjugate-forming cells from nylon nonadherent dystrophic mouse spleen cells were found to be two- to fourfold greater than for normal mouse spleen cells using YAC-1 tumor cells as targets. The number of lymphocytes bound per YAC-1 target cell ranged from 2 to 5 for dystrophic mouse spleen cells as compared with 1 to 2 for the normal control group. Using both normal and dystrophic mouse thymocytes as targets, the conjugate-forming cells from dystrophic mouse spleen cells were also found to be twofold greater than in the normal control group. Cold target inhibition studies revealed that the natural killing of dystrophic mouse thymocytes was due to a YAC-1-reactive NK cell. Effector cell depletion studies using monoclonal anti-Thy-1.2 plus complement treatment and plastic petri dish adherence also revealed that the natural killing of dystrophic mouse thymocytes was not due to either T lymphocytes or macrophages. Taken together, these results show an increase in NK-sensitive thymocyte targets in dystrophic mice, in combination with an increase in splenic NK activity.  相似文献   

17.
The effect of pentylenetetrazol on the metacerebral neuron of Helix pomatia   总被引:1,自引:0,他引:1  
The effects of Pentylenetetrazol (PTZ) on the metacerebral giant cell (MCC) of the snail, Helix pomatia were studied. Actions on membrane resistance, time constant, resting and action potentials, outward and inward ionic currents were examined. Superfusion with PTZ in concentrations of 25 to 50 mmol/l, induced a gradually evolving convulsive state, which could be studied by intracellular recording from the MCCs. In the pre-convulsive state an acceleration of the spontaneous activity developed and was followed by paroxysmal depolarization shifts (PDSs), in the convulsive phase. PTZ prolonged the membrane time constant by about 10 percent, but this could not be traced back to alterations in membrane resistance or capacity. The resting membrane potential was not significantly altered; the action potentials were prolonged by slowing down of both the rising and decaying phases. The outward potassium currents were repressed by PTZ in a voltage dependent manner. The decrease of the IA current became more pronounced at increasingly positive command pulses, while IK was relieved from depression especially at longer pulse durations. Inward currents were isolated with the aid of suppression of outward currents by 50 mmol/l TEA. Under these conditions sodium currents, measured in calcium deficient Ringer solution were moderately depressed, while the calcium currents, examined during sodium-free superfusion, were mildly enhanced by PTZ. It is concluded that PTZ effects on ionic conductances, on membrane parameters, on the resting potential and ionic currents explain only modifications of spike potentials occurring in the convulsive state and do not account for the PDS, the central phenomenon of the convulsive electrographic activity, at least in this thoroughly examined type of neuron.  相似文献   

18.
The four human Duchenne dystrophic isoenzymes (M-M, M-B, B-B, from the muscle and B-B from the brain) of ATP-creatine transphosphorylase (S. A. Kuby, H. J. Keutel, K. Okabe, H. K. Jacobs, F. Ziter, D. Gerber, and F. H. Tyler, 1977, J. Biol. Chem.252, 8382–8390) have now been compared physically and chemically with their normal human counterparts (viz., with the three isoenzymes, M-M, M-B, B-B, 2). All isoenzymes proved to be composed of two noncovalently linked polypeptide chains, by sedimentation equilibrium analyses in the presence and absence of disruptive agents. In the presence of 2-mercaptoethanol at 0.16(Γ/2), pH 7.8, the two native muscle types yielded identical values for s20,w, concentration dependencies, and molecular weight, and similarly for the brain types (from the brain). But the human brain type proved to be slightly heavier than the muscle type (viz. 88,400 vs 85,900). All of the isoenzymes showed similar electrophoretic behavior between their several counterparts between pH 5–8, except perhaps between pH 8–10, where small differences appeared. The three native normal human isoenzymes, as well as the dystrophic human isoenzymes (M-M from the muscle and B-B from the brain) all contain 2 reactive sulfhydryl groups per mole or 1 per polypeptide chain of these two-chain proteins, which may be titrated with 5,5′-dithiobis(2-nitrobenzoic acid) (Nbs2); and under acidic conditions, quantitative titrations with 4,4′-dithiodipyridine yield a total of 10 -SH groups per mole of each brain type and 8 -SH groups per mole of muscle type, in the case of man, dystrophic man, calf, and rabbit. The kinetics of reactions between Nbs2 and the sulfhydryl groups of all three normal human isoenzymes and two dystrophic human isoenzymes have been measured under several sets of denaturing conditions. A comparison of their reactive calculated second-order velocity constants reveal significant differences between these three normal human isoenzymes, but the ksecond order values for the reactions of the sulfhydryl groups of the dystrophic M-M and B-B with Nbs2, when compared with their normal counterparts, gave identical values in the presence of 7.3 m urea or 1.8% laurylsulfate, from which it may be inferred that very similar, if not identical, environments surround these two sets of sulfhydryl groups. A comparison of the amino acid compositions of the normal human muscle type and brain type with the human dystrophic M-M and B-B (from the brain) reveal essentially identical values for the muscle types but nearly identical values for the brain types, with a few differences. Their respective tryptic peptide maps have been compared of the S-carboxy-methylated proteins (alkylated with iodo[2-14C]acetic acid at the two exposed -SH groups per mole). Thus, the muscle types, normal and dystrophic, yield identical maps, but the brain types nearly identical maps, with a few significant differences. Isolation of the tryptic tridecapeptide from the S-carboxymethylated normal human and dystrophic human dimeric muscle-type ATP-creatine transphosphorylases, labeled at the single exposed SH group per polypeptide chain with iodo[2-14C]acetate, yielded the following sequence for both proteins: ValLeuThrCys(CH2COOH)ProSerAsnLeuGlyThr GlyLeuArg [where Cys(CH2COOH) is S-carboxymethyl cysteine]. This sequence showed remarkable homology with a few other equivalent peptides reported to be derived from the exposed SH group of other ATP-creatine transphosphorylases. In conclusion, there does not appear to be a mutation in the structural genes for the muscle-type creatine kinases detectable by the analyses presented here. However, the brain types warrant further investigation.  相似文献   

19.
Abstract: Acetylcholinesterase (AChE) and pseudocholinesterase (°ChE) were analysed in the blood plasma of developing chickens, both normal and those with inherited muscular dystrophy. The amounts and the molecular forms of each were examined. °ChE concentration rises in the plasma of normal and dystrophic chicks at the end of embryonic development and is maintained after hatching at a constant, relatively high level, accounting for 90-95% of total cholinesterase activity in normal plasma. This level is maintained in normal and dystrophic chickens. In embryonic plasma of both normal and dystrophic chicks, on the other hand, the levels of AChE are higher than those of °ChE. Immediately after hatching the AChE level decreases rapidly in normal plasma, reaching a very low level by 2-3 weeks ex ovo. The AChE level in plasma from dystrophic birds, although less than normal from day 19 in ovo to 2 weeks ex ovo, subsequently increases to peak around 4 months at levels 15-20-fold of those in normal birds. There is virtually no enzyme of either type in the erythrocytes of normal or dystrophic chickens. The changes of AChE in plasma were correlated with the alterations of AChE in dystrophic fast-twitch muscles, suggesting that the latter pool is a precursor of the plasma AChE. Both the AChE and the °ChE in plasma exist in multiple molecular forms, which are similar to certain of those found previously in the muscles of these birds. The major form (60-80%) of both enzymes in the plasma is the M form (sedimentation coefficient ≥11 S) in all cases, but it is accompanied by certain other forms. In no case is there any of the heaviest form (H2, 19-20 S) of AChE or of °ChE found in normal and dystrophic muscle, which is attached at the synapses in normal muscle. The pattern of forms of plasma °ChE is constant at all ages, and in normal and dystrophic chickens. The pattern of forms of AChE in the plasma, in contrast, varies with age and with dystrophy in a characteristic manner. The sedimentation coefficients and the amounts of the enzymes in fast-twitch muscle of dystrophic animals are compared with those of the plasma forms, and an interpretation is given of the characteristic patterns of AChE and of χE in their blood.  相似文献   

20.
Thresholds and Plateaus in the Hodgkin-Huxley Nerve Equations   总被引:9,自引:1,他引:8       下载免费PDF全文
Phase space methods and an analog computer are used to analyze the Hodgkin-Huxley non-linear differential equations for the squid giant axon membrane. V is the membrane potential, m the Na+ activation, h the Na+ inactivation, and n the K+ activation. V and m change rapidly, relative to h and n. The (V, m) phase plane of a reduced system of equations, with h and n held constant at their resting values, has three singular points: a stable resting point, a threshold saddle point, and a stable excited point. When h and n are allowed to vary, recovery and refractoriness result from the movement with subsequent disappearance of the threshold and excited points. Multiplying the time constant of n by 100 or more, and that of h by one-third, reproduces the experimental plateau action potentials obtained with tetraethylammonium by Tasaki and Hagiwara, including the phenomena of abolition and of refractoriness of the plateau duration. The equations have, transiently, two stable states, as found in the real axon by these authors. Since the theoretical membrane conductance curves differ significantly from the experimental ones, further experimental analysis of ionic currents with tetraethylammonium is needed to decide whether the Hodgkin-Huxley model can be generalized to explain these experiments completely.  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