首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
S M de Waegh  V M Lee  S T Brady 《Cell》1992,68(3):451-463
Studies in Trembler and control mice demonstrated that myelinating Schwann cells exert a profound influence on axons. Extensive contacts between myelin and axons have been considered structural. However, demyelination decreases neurofilament phosphorylation, slow axonal transport, and axonal diameter, as well as significantly increasing neurofilament density. In control sciatic nerves with grafted Trembler nerve segments, these changes were spatially restricted: they were confined to axon segments without normal myelination. Adjacent regions of the same axons had normal diameters, neurofilament phosphorylation, cytoskeletal organization, and axonal transport rates. Close intercellular contacts between myelinating Schwann cells and axons modulate a kinase-phosphatase system acting on neurofilaments and possibly other substrates. Myelination by Schwann cells sculpts the axon-altering functional architecture, electrical properties, and neuronal morphologies.  相似文献   

2.
A previous study by McLean and co-workers reported that regenerating axons of the rabbit vagus nerve were unable to sustain axonal transport in vitro for several months after nerve injury. In contrast, we found that sensory axons of the rat sciatic nerve were able to transport 3H-labeled protein into their regenerating portions distal to the site of injury within a week after injury when placed in vitro. Transport in vitro was not significantly less than transport in axons maintained in vivo for the same period. Transport occurred in the medium that was used by the McLean group, but was significantly reduced in calcium-free medium. When axon regeneration was delared, only small amounts of activity were present in the nerve distal to the site of injury, showing that labeled protein normally present in that part of the nerve was associated with axons and was not a result of local precursor uptake by nonneural elements in the sciatic nerve. We were not able to explain the failure of McLean and co-workers to demonstrate transport in vitro in regenerating vagus nerve, but we conclude that there is no general peculiarity of growing axons that makes them unable to sustain transport in vitro.  相似文献   

3.
Spatial organization of axonal microtubules   总被引:9,自引:8,他引:1       下载免费PDF全文
《The Journal of cell biology》1984,99(4):1289-1295
Several workers have found that axonal microtubules have a uniform polarity orientation. It is the "+" end of the polymer that is distal to the cell body. The experiments reported here investigate whether this high degree of organization can be accounted for on the basis of structures or mechanisms within the axon. Substantial depolymerization of axonal microtubules was observed in isolated, postganglionic sympathetic nerve fibers of the cat subjected to cold treatment; generally less than 10% of the original number of microtubules/micron 2 remained in cross section. The number of cold stable MTs that remained was not correlated with axonal area and they were also found within Schwann cells. Microtubules were allowed to repolymerize and the polarity orientation of the reassembled microtubules was determined. In fibers from four cats, a majority of reassembled microtubules returned with the original polarity orientation. However, in no case was the polarity orientation as uniform as the original organization. The degree to which the original orientation returned in a fiber was correlated with the number of cold-stable microtubules in the fiber. We suggest that stable microtubule fragments serve as nucleating elements for microtubule assembly and play a role in the spatial organization of neuronal microtubules. The extremely rapid reassembly of microtubules that we observed, returning to near control levels within the first 5 min, supports microtubule elongation from a nucleus. However, in three of four fibers examined this initial assembly was followed by an equally rapid, but transient decline in microtubule number to a value that was significantly different than the initial peak. This observation is difficult to interpret; however, a similar transient peak has been reported upon repolymerization of spindle microtubules after pressure induced depolymerization.  相似文献   

4.
In peripheral nerves, Schwann cells form the myelin sheath that insulates axons and allows rapid propagation of action potentials. Although a number of regulators of Schwann cell development are known, the signaling pathways that control myelination are incompletely understood. In this study, we show that Gpr126 is essential for myelination and other aspects of peripheral nerve development in mammals. A mutation in Gpr126 causes a severe congenital hypomyelinating peripheral neuropathy in mice, and expression of differentiated Schwann cell markers, including Pou3f1, Egr2, myelin protein zero and myelin basic protein, is reduced. Ultrastructural studies of Gpr126-/- mice showed that axonal sorting by Schwann cells is delayed, Remak bundles (non-myelinating Schwann cells associated with small caliber axons) are not observed, and Schwann cells are ultimately arrested at the promyelinating stage. Additionally, ectopic perineurial fibroblasts form aberrant fascicles throughout the endoneurium of the mutant sciatic nerve. This analysis shows that Gpr126 is required for Schwann cell myelination in mammals, and defines new roles for Gpr126 in axonal sorting, formation of mature non-myelinating Schwann cells and organization of the perineurium.  相似文献   

5.
A period of electrical activity in unmyelinated nerve fibers is followed by a post-tetanic hyperpolarization (PTH), generated by the hyperactivity of the electrogenic Na(+)-K(+) pump. In order to protect the membrane potential against these strong hyperpolarizations, different types of axonal inward currents are activated during the PTH. We investigated in the rabbit vagus nerve one of these currents, which was activated by carbamylcholine (CCh). We observed that the effect of CCh on the PTH amplitude could be blocked or reversed with scopolamine. Moreover, the PTH amplitude increased when scopolamine alone was added to the perfusate, indicating that an endogenous muscarinic agonist was liberated in the preparation during the period of electrical activity. This CCh-activated current was TEA but not Ba(2+) or Cs(+) sensitive. It has been shown previously that muscarinic acetylcholine receptors (mAChRs) in the rabbit vagus nerve are located on the axonal but not glial membrane and that Schwann cells express several types of purinergic receptors, which activation evoke Ca(2+) transients in Schwann cells. We hypothesise that during electrical activity axons release a transmitter, presumably ATP. This transmitter evoke in the neighbouring Schwann cells a Ca(2+)-dependent liberation of a endogenous muscarinic agonist, which in turn activates a TEA-sensitive inward current in axons. We suggest that the major purpose of this mechanism is the control of the membrane potential during and after a period of intense electrical activity when the Na(+)-K(+) pump generates a robust PTH.  相似文献   

6.
Tullidinol, a neurotoxin extracted from the Karwinskia humboldtiana fruit, dissolved in peanut oil was injected into the right sciatic nerve of adult cats. The contralateral sciatic nerve received an equivalent volume of peanut oil alone. The fast axonal transport of labeled ([3H]Leucine) protein was studied in sensory and motor axons of both sciatic nerves. The radioactive label was pressure injected either into the L7 dorsal root ganglion or the ventral region of the same spinal cord segment. Several days after the toxin injection, the cat limped and the Achilles tendon reflex was nearly absent in the right hind limb. The amount of transported label was decreased distal to the site of toxin injection. Proximal to this site, the transported material was dammed. Sensory and motor axons showed similar changes. In addition, the toxin produced demyelination and axonal degeneration. Axonal transport and the structure of the axons were normal in the contralateral nerve. Both, Schwann cells and axons of the right sciatic nerve showed globular inclusions, presumably oil droplets containing the toxin. We conclude that Schwann cells and axons as well are tullidinol targets.Departamento de Química. Centro de Investigación y de Estudios Avanzados del IPN.Special issue dedicated to Dr. Sidney Ochs.  相似文献   

7.
Injury to the rat sciatic nerve leads to the induction of nerve growth factor (NGF) receptors on the denervated Schwann cells and their disappearance on the regenerating axons of the axotomized, normally NGF-sensitive sensory and sympathetic neurons. This disappearance in the axonal expression and retrograde transport of NGF receptors is associated with a similarly dramatic reduction in the axonal uptake and retrograde transport of NGF following axotomy and during regeneration. In view of the massive NGF synthesis occurring in the injured nerve, these results suggest that, while sensory and sympathetic neurons are the primary targets of NGF in the normal peripheral nervous system, the denervated Schwann cells may become its primary target in the aftermath of nerve injury.  相似文献   

8.
To identify the structures to be rapidly transported through the axons, we developed a new method to permit local cooling of mouse saphenous nerves in situ without exposing them. By this method, both anterograde and retrograde transport were successfully interrupted, while the structural integrity of the nerves was well preserved. Using radioactive tracers, anterogradely transported proteins were shown to accumulate just proximal to the cooled site, and retrogradely transported proteins just distal to the cooled site. Where the anterogradely transported proteins accumulated, the vesiculotubular membranous structures increased in amount inside both myelinated and unmyelinated axons. Such accumulated membranous structures showed a relatively uniform diameter of 50--80 nm, and some of them seemed to be continuous with the axonal smooth endoplasmic reticulum (SER). Thick sections of nerves selectively stained for the axonal membranous structures revealed that the network of the axonal SER was also packed inside axons proximal to the cooled site. In contrast, large membranous bodies of varying sizes accumulated inside axons just distal to the cooled site, where the retrogradely transported proteins accumulated. These bodies were composed mainly of multivesicular bodies and lamellated membranous structures. When horseradish peroxidase was administered in the distal end of the nerve, membranous bodies showing this activity accumulated, together with unstained membranous bodies. Hence, we are led to propose that, besides mitochondria, the membranous components in the axon can be classified into two systems from the viewpoint of axonal transport: "axonal SER and vesiculotubular structures" in the anterograde direction and "large membranous bodies" in the retrograde direction.  相似文献   

9.
Microtubules and the capacity of the system for rapid axonal transport   总被引:1,自引:0,他引:1  
Current information favors the view that microtubules are required for rapid axonal transport of proteins and organelles but are normally present in surplus. Different types of axons tolerate losses of between 35 and 65% of their microtubules during exposure to low temperatures or antimitotic drugs before transport is impaired. Greater losses of microtubules are associated with progressive and marked failure of transport. The normal surplus of microtubules may explain why adrenergic axons of rabbit peroneal nerve have spare transport capacity, which enables them to transport between two and three time as much material as they do ordinarily. Spare capacity for transport is diminished or absent when nerves are incubated at temperatures that lead to a partial loss of microtubules. These observations are considered in the light of the hypothesis that the local density of microtubules determines the maximal local concentration of material that can be carried by rapid transport along vertebrate axons.  相似文献   

10.
Axonal Transport of Neuropeptides in the Cervical Vagus Nerve of the Rat   总被引:4,自引:2,他引:2  
Accumulations of the neuropeptides substance P (SP), somatostatin (ST), and vasoactive intestinal polypeptide (VIP) proximal to a crush in the cervical vagus nerve of the rat have been measured using sensitive radioimmunoassays. Each of the peptides was rapidly transport towards the peripheral terminals of vagal afferent fibres, with average rates of flow ranging from 0.8 to 2.7 mm h-1. In the rabbit vagus nerve, SP was transported with an average rate of 4 mm h-1, which is more than double the rate for this peptide in the rat. Double crush experiments in rabbit vagus nerves indicated that the rapidly transported proportion of the total content of SP in the nerve free was about 34%. From this, the rate of transport of SP in the rapidly transported pool in the rabbit vagus nerve can be calculated to be 12 mm h-1 (280 mm day-1). Since such double crush experiments were not possible in the rat, it is not clear whether the different average rates of transport of SP in the rat and the rabbit reflect real differences in the rate of rapid transport in the two species. In common with rapid axonal transport of other neurotransmitters, the transport of SP and ST in the rat vagus nerve was blocked by colchicine, a drug that disrupts microtubules.  相似文献   

11.
Abstract: Biochemical methods were used to study the time course of transport of choline phospholipids (labeled by the injection of [3H]choline into the ventral horn of the lumbar spinal cord) in rat sciatic nerve. Autoradiographic methods were used to localize the transported lipid within motor axons. Transported phospholipid, primarily phosphatidylcholine, present in the nerve at 6 h, continued to accumulate over the following 12 days. No discrete waves of transported lipid were observed (a small wave of radioactive phospholipid moving at the high rate would have been missed); the amounts of radioactive lipid increased uniformly along the entire sciatic nerve. In light-microscope autoradiographs, a class of large-caliber axons, presumably motor axons, retained the labeled lipid. Some lipid, even at 6 h, was seen within the myelin sheaths. Later, the labeling of the myelin relative to axon increased. The continued accumulation of choline phospholipids in the axons probably signifies their prolonged release from cell bodies and their retention in various axonal membranes, including the axolemma. The build-up of these phospholipids in myelin probably represents their transfer from the axons to the myelin sheaths surrounding them. When nerves are crushed and allowed to regenerate for 6 or 12 days, choline phospholipids transported during these times enter the regenerating nerve. In light and electron microscope autoradiographs, transported lipid was seen to be localized primarily in the regenerating axons. However, grains overlay the adjacent Schwann cell cytoplasm, indicating transported lipids were transferred from the regenerating axons to the associated Schwann cells. In addition, some cells not associated with growing axons were labeled, suggesting that phosphatidylcholine and possibly acetylcholine, carried to the regenerating axons by axonal transport, were actively metabolized in the terminal, with released choline label being used by other cells. These results demonstrate that axonal transport supplies mature and growing axons and their glial cells with choline phospholipids.  相似文献   

12.
The optic nerve, as a part of the central nervous system (CNS), has been used to study axonal transport for decades. The present study has concentrated on the axonal transport of synaptic vesicle proteins in the optic nerve, using the “stop-flow/nerve crush” method. After blocking fast axonal transport, distinct accumulations of synaptic vesicle proteins developed during the first hour after crush-operation and marked increases were observed up to 8 h postoperative. Semiquantitative analysis, using cytofluorimetric scanning (CFS) of immunoincubated sections, revealed that the ratio between distal accumulations (organelles in retrograde transport) and proximal accumulations (organelles in anterograde transport) was much higher (up to 80–90%) for the transmembrane proteins than that for surface adsorbed proteins (only 10–20%). The pattern of axonal transport in the optic nerve was comparable to that in the sciatic nerve. However, clathrin and Rab3a immunoreactivities were accumulated in much lower amounts than that in the sciatic nerve. Most synaptic vesicle proteins were colocalized in the axons proximal to the crush. A differential distribution of synaptobrevin I and II, however, was observed in the optic nerve axons; synaptobrevin I was present in large-sized axons, while synaptobrevin II immunoreactivity was present in most axons, including the large ones. The two isoforms were, thus, partially colocalized. The results demonstrate that (1) cytofluorimetric scanning techniques could be successfully used to study axonal transport not only in peripheral nerves, but also in the CNS; (2) synaptic vesicles are transported with fast axonal transport in this nerve; and (3) some differences were noted compared with the sciatic nerve, especially for Rab3a and clathrin. © 1997 John Wiley & Sons, Inc. J Neurobiol 32: 237–250, 1997.  相似文献   

13.
Control of axonal caliber by neurofilament transport   总被引:30,自引:13,他引:17  
The role of neurofilaments, the intermediate filaments of nerve cells, has been conjectural. Previous morphological studies have suggested a close relationship between neurofilament content and axonal caliber. In this study, the regenerating neuron was used as a model system for testing the hypotheses that neurofilaments are intrinsic determinants of axonal caliber, and that neurofilament content is controlled by the axonal transport of neurofilaments. This system was chosen because previous studies had shown that, after axotomy, axonal caliber was reduced within the proximal stump of the regenerating nerve and, because the relative amount of neurofilament protein undergoing axonal transport in regenerating axons was selectively reduced. The relationship between axonal caliber and neurofilament number was examined in a systematic fashion in both regenerating and control motor axons in rat L5 ventral root. Reconstruction of the spatial and temporal sequences of axonal atrophy in the proximal stump after axotomy showed that reductions in axonal caliber were first detected in the most proximal region of the root and subsequently progressed in a proximal-to-distal direction at a rate of 1.7 mm/day, which is identical to the rate of neurofilament transport in these neurons. Quantitative ultrastructural studies showed that these reductions in caliber correlated with a proportional decrease in the number of axonal neurofilaments but not microtubules. These results support the hypotheses that neurofilament content is a major intrinsic determinant of axonal caliber and that neurofilament content is controlled by the axonal transport of neurofilaments. On this basis, we suggest a role for neurofilaments in the control of axonal volume.  相似文献   

14.
Schwann cells as regulators of nerve development.   总被引:15,自引:0,他引:15  
Myelinating and non-myelinating Schwann cells of peripheral nerves are derived from the neural crest via an intermediate cell type, the Schwann cell precursor [K.R. Jessen, A. Brennan, L. Morgan, R. Mirsky, A. Kent, Y. Hashimoto, J. Gavrilovic. The Schwann cell precursor and its fate: a study of cell death and differentiation during gliogenesis in rat embryonic nerves, Neuron 12 (1994) 509-527]. The survival and maturation of Schwann cell precursors is controlled by a neuronally derived signal, beta neuregulin. Other factors, in particular endothelins, regulate the timing of precursor maturation and Schwann cell generation. In turn, signals derived from Schwann cell precursors or Schwann cells regulate neuronal numbers during development, and axonal calibre, distribution of ion channels and neurofilament phosphorylation in myelinated axons. Unlike Schwann cell precursors, Schwann cells in older nerves survive in the absence of axons, indicating that a significant change in survival regulation occurs. This is due primarily to the presence of autocrine growth factor loops in Schwann cells, present from embryo day 18 onwards, that are not functional in Schwann cell precursors. The most important components of the autocrine loop are insulin-like growth factors, platelet derived growth factor-BB and neurotrophin 3, which together with laminin support long-term Schwann cell survival. The paracrine dependence of precursors on axons for survival provides a mechanism for matching precursor cell number to axons in embryonic nerves, while the ability of Schwann cells to survive in the absence of axons is an absolute prerequisite for nerve repair following injury. In addition to providing survival factors to neurones and themselves, and signals that determine axonal architecture, Schwann cells also control the formation of peripheral nerve sheaths. This involves Schwann cell-derived Desert Hedgehog, which directs the transition of mesenchymal cells to form the epithelium-like structure of the perineurium. Schwann cells thus signal not only to themselves but also to the other cellular components within the nerve to act as major regulators of nerve development.  相似文献   

15.
Abstract— The transport, distribution and turnover of choline O -acetyltransferase (ChAc, EC 2.3.1.6) and acetylcholinesterase (AChE, EC 3.1.1.7) in the vagus and hypoglossal nerves were studied in adult rabbits. The enzymes accumulated proximally and distally to single and double ligatures on both nerves and thus indicated both a proximo-distal and retrograde flow of the enzymes. Double ligature experiments indicated that only 5–20 per cent of the enzymes were mobile in the axon. The rate of accumulation of both enzymes above a single ligature corresponded to the slow rate of axonal flow provided that all the enzymes were mobile, but to an intermediate or fast flow if only a small part of the enzymes was transported. The distribution of ChAc along the hypoglossal neurons was studied and only 2 per cent of ChAc was confined to cell bodies, 42 per cent was localized to the main hypoglossal nerve trunks and 56 per cent to the preterminal axons and axon terminals in the tongue. The ratio of AChE to ChAc was about 3 in the hypoglossal nerve and 32 in the vagus nerve.
Transection of the hypoglossal nerve was followed by a decrease in the activity of ChAc in the hypoglossal nucleus and nerve and in the axons and their terminals in the tongue. The activity of AChE decreased in the hypoglossal nucleus and nerve but not in the tongue. The half-life of ChAc in preterminal axons and terminals of the hypoglossal nerve was estimated to be 16-21 days from the results obtained on transport, axotomy and distribution of the enzyme. Intracisternal injection of colchicine inhibited the cellulifugal transport of both enzymes and led to an increase in enzyme activity in the hypoglossal nucleus.  相似文献   

16.
Schwann cell extracellular matrix molecules and their receptors   总被引:10,自引:0,他引:10  
The major cellular constituents of the mammalian peripheral nervous system are neurons (axons) and Schwann cells. During peripheral nerve development Schwann cells actively deposit extracellular matrix (ECM), comprised of basal lamina sheets that surround individual axon-Schwann cell units and collagen fibrils. These ECM structures are formed from a diverse set of macromolecules, consisting of glyco-proteins, collagens and proteoglycans. To interact with ECM, Schwann cells express a number of integrin and non-integrin cell surface receptors. The expression of many Schwann cell ECM proteins and their receptors is developmentally regulated and, in some cases, dependent on axonal contact. Schwann cell ECM acts as an organizer of peripheral nerve tissue and strongly influences Schwann cell adhesion, growth and differentiation and regulates axonal growth during development and regeneration.  相似文献   

17.
Fansa H  Keilhoff G  Wolf G  Schneider W 《Plastic and reconstructive surgery》2001,107(2):485-94; discussion 495-6
Bioengineering is considered to be the laboratory-based alternative to human autografts and allografts. It ought to provide "custom-made organs" cultured from patient's material. Venous grafts and acellular muscle grafts support axonal regeneration only to a certain extent because of the lack of viable Schwann cells in the graft. We created a biologic nerve graft in the rat sciatic nerve model by implanting cultured Schwann cells into veins and acellular gracilis muscles, respectively. Autologous nerve grafts and veins and acellular muscle grafts without Schwann cells served as controls. After 6 and 12 weeks, regeneration was assessed clinically, histologically, and morphometrically. The polymerase chain reaction analvsis showed that the implanted Schwann cells remained within all the grafts. The best regeneration was seen in the control; after 12 weeks the number of axons was increased significantly compared with the other grafts. A good regeneration was noted in the muscle-Schwann cell group, whereas regeneration in both of the venous grafts and the muscle grafts without Schwann cells was impaired. The muscle-Schwann cell graft showed a systematic and organized regeneration including a proper orientation of regenerated fibers. The venous grafts with Schwann cells showed less fibrous tissue and disorganization than the veins without Schwann cells, but failed to show an excellent regeneration. This might be attributed to the lack of endoneural-tube-like components serving as scaffold for the sprouting axon. Although the conventional nerve graft remains the gold standard, the implantation of Schwann cells into an acellular muscle provides a biologic graft with basal lamina tubes as pathways for regenerating axons and the positive effects of Schwann cells producing neurotrophic and neurotropic factors, and thus, supporting axonal regeneration.  相似文献   

18.
The effect of in vitro exposure of bullfrog spinal nerves to 0.2 mM chlorimipramine on the density of axonal microtubules was studied in an attempt to clarify the mechanism by which chlorimipramine inhibits fast axonal transport. A 17-h exposure to chlorimipramine reduced the density of microtubules in unmyelinated axons by only 18%; this microtubular loss does not reach the upper limit of the range of microtubule reduction associated with inhibition of fast axonal transport. A 23-h exposure to chlorimipramine, which had decreased microtubular density in unmyelinated axons by 40% in a previous study, did not decrease microtubular density in myelinated axons in the present study. These results rule out microtubular destruction as the mechanism responsible for inhibition of fast orthograde axonal transport by chlorimipramine, and greatly reduce the likelihood that microtubular destruction plays a significant role in the inhibition of fast retrograde transport by chlorimipramine.  相似文献   

19.
Developing neurons express a motor protein called kinesin-5 (also called kif11 or Eg5) which acts as a 'brake' on the advance of the microtubule array during axonal growth. Pharmacological inhibition of kinesin-5 causes the developing axon to grow at a faster rate, retract less and grow past cues that would otherwise cause it to turn. Here we demonstrate that kinesin-5 is also expressed in adult neurons, albeit at lower levels than during development. We hypothesized that inhibiting kinesin-5 might enable adult axons to regenerate better and to overcome repulsive molecules associated with injury. Using adult mouse dorsal root ganglion neurons, we found that anti-kinesin-5 drugs cause axons to grow faster and to cross with higher frequency onto inhibitory chondroitin sulfate proteoglycans. These effects may be due in part to changes in the efficiency of microtubule transport along the axonal shaft as well as enhanced microtubule entry into the distal tip of the axon. Effects observed with the drugs are further enhanced in some cases when they are used in combination with other treatments known to enhance axonal regeneration. Collectively, these results indicate that anti-kinesin-5 drugs may be a useful addition to the arsenal of tools used to treat nerve injury.  相似文献   

20.
Microtubule (MT) number, axonal area, and MT density were examined in unmyelinated axons of rat cervical vagus nerve. Study of nerve regions proximal (1-5 mm) and distal (35-40 mm) to the nodosum ganglion in controls (incubation at 37 degrees C for 1 h) showed that the number of MT per axon is significantly less in distal than in proximal nerve regions. Cooling (incubation at 0 degree C for 1 h) caused a significant reduction in the number of MT per axon in both nerve regions. The unmyelinated axons from both nerve regions showed a comparable reduction in MT number by cooling, indicating that axonal MT stability to cold was not significantly different between these two nerve regions. In these nerves no detectable changes were found in cross-axonal area of unmyelinated axons between distal and proximal nerve regions. In another experimental series, in distal nerve regions (35-40 mm from the nodosum ganglion) the number of MT was not further reduced in nerves incubated at 0 degree C by increasing the incubation time. Similar results were obtained from colchicine treated nerves (incubation at 37 degrees C, with 10 mM colchicine for 1 and 2 h). Distal nerve regions (35-40 mm from the nodosum ganglion) showed a similar reduction in the number of MT per axon when nerves were incubated at 0 degree C or with colchicine, suggesting that this drug, as well as cold, may be affecting a similar population of axonal MT, i.e., MT susceptible to anti-MT agents. These results indicate that approximately one-half of the axonal MT are stable to cold as well as to colchicine in rat unmyelinated axons.  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