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 46 毫秒
1.
The capacity of CNS neurons for axonal regrowth after injury decreases as the age of the animal at time of injury increases. After spinal cord lesions at birth, there is extensive regenerative growth into and beyond a transplant of fetal spinal cord tissue placed at the injury site. After injury in the adult, however, although host corticospinal and brainstem-spinal axons project into the transplant, their distribution is restricted to within 200 micron of the host/transplant border. The aim of this study was to determine if the administration of neurotrophic factors could increase the capacity of mature CNS neurons for regrowth after injury. Spinal cord hemisection lesions were made at cervical or thoracic levels in adult rats. Transplants of E14 fetal spinal cord tissue were placed into the lesion site. The following neurotrophic factors were administered at the site of injury and transplantation: brain-derived neurotrophic factor (BDNF), neurotrophin-3 (NT-3), neurotrophin-4 (NT-4), ciliary-derived neurotrophic factor (CNTF), or vehicle alone. After 1-2 months survival, neuroanatomical tracing and immunocytochemical methods were used to examine the growth of host axons within the transplants. The neurotrophin administration led to increases in the extent of serotonergic, noradrenergic, and corticospinal axonal ingrowth within the transplants. The influence of the administration of the neurotrophins on the growth of injured CNS axons was not a generalized effect of growth factors per se, since the administration of CNTF had no effect on the growth of any of the descending CNS axons tested. These results indicate that in addition to influencing the survival of developing CNS and PNS neurons, neurotrophic factors are able to exert a neurotropic influence on injured mature CNS neurons by increasing their axonal growth within a transplant.  相似文献   

2.
The injured adult mammalian spinal cord shows little spontaneous recovery after injury. In the present study, the contribution of projections in the dorsal half of the spinal cord to functional loss after adult spinal cord injury was examined, together with the effects of transgenic cellular delivery of neurotrophin-3 (NT-3) on morphological and functional disturbances. Adult rats underwent bilateral dorsal column spinal cord lesions that remove the dorsal corticospinal projections or underwent more extensive resections of the entire dorsal spinal cord bilaterally that remove corticospinal, rubrospinal, and cerulospinal projections. Long-lasting functional deficits were observed on a motor grid task requiring detailed integration of sensorimotor skills, but only in animals with dorsal hemisection lesions as opposed to dorsal column lesions. Syngenic primary rat fibroblasts genetically modified to produce NT-3 were then grafted to acute spinal cord dorsal hemisection lesion cavities. Up to 3 months later, significant partial functional recovery occurred in NT-3-grafted animals together with a significant increase in corticospinal axon growth at and distal to the injury site. These findings indicate that (1) several spinal pathways contribute to loss of motor function after spinal cord injury, (2) NT-3 is a neurotrophic factor for the injured corticospinal projection, and (3) functional deficits are partially ameliorated by local cellular delivery of NT-3. Lesions of the corticospinal projection may be necessary, but insufficient in isolation, to cause sensorimotor dysfunction after spinal cord injury in the rat.  相似文献   

3.
Recent studies show that neuropeptide Y acts indirectly, via release of a neurotrophic factor(s) from the spinal cord, to increase the neurite outgrowth of dissociated adult rat dorsal root ganglion cells. This study examines further the neuropeptide Y-induced increase in neurite outgrowth. To characterize the factor(s) mediating the neuropeptide Y-induced increase in neurite outgrowth, we have examined whether antisera to either nerve growth factor or neurotrophin-3 influence the neuropeptide Y-induced increase in neurite outgrowth. Spinal cord slices were incubated with media alone or in combination with 10 nM neuropeptide Y for 2 h at 37 degrees C. The supernatant of spinal cord incubated with neuropeptide Y significantly enhanced the neurite outgrowth of normal dorsal root ganglion cells. Antiserum against nerve growth factor had no effect on the trophic actions of the supernatant. Antiserum against neurotrophin-3, however, significantly attenuated the increase in neurite outgrowth. Consistent with this finding, neurotrophin-3 also increased the percentage of cells with neurites. Transganglionic labelling of A-fibres with choleragenoid-horseradish peroxidase in animals treated intrathecally with neurotrophin-3 for 14 days via an osmotic pump showed that the area of choleragenoid-horseradish peroxidase label expanded into lamina II. In comparison, saline-treated animals had no label in lamina II. In addition, neurotrophin-3-treated animals also had a significant decrease in mechanical nociceptive threshold. The results suggest that neuropeptide Y acts via neurotrophin-3 to mediate an increase in neurite outgrowth of dorsal root ganglion cells. These results have important implications for the mechanisms underlying neuropathic pain.  相似文献   

4.
We replaced the mouse neurotrophin-3 gene with the Escherichia coli-derived lacZ gene by means of homologous recombination. The mice with this mutation were useful models for studying the distribution of neurotrophin-3 expression in vivo, because visualization by 5-bromo-4-chloro-3-indoyl-beta-D-galactopyranoside (X-Gal) staining was simple and rapid compared with in situ hybridization or immunohistochemistry. Whole-mount staining of mutant embryos at embryonic day 10 revealed that lacZ, a reporter for the neurotrophin-3 gene, was expressed in the mesencephalon, mandibular arch and somites. In the embryos at days 13-17, lacZ was markedly expressed in the peripheral target tissues of sensory and sympathetic neurons. We also found that spinal motor neurons and sensory neurons in trigeminal and dorsal root ganglia express lacZ. Some of these X-Gal staining regions overlapped with the sites expressing trkC, a high-affinity receptor for neurotrophin-3. The distribution of X-Gal staining in heterozygotes and homozygotes was similar to that of neurotrophin-3 messenger RNA detected by in situ hybridization. However, there was less lacZ expression in the dorsal root ganglia of homozygotes than neurotrophin-3 expression in wild-type mice. These results suggest that the neurotrophin-3 produced in the dorsal root ganglia also plays a role in the survival of some of the neurotrophin-3-positive neurons and that the local mode of neurotrophic activity is widely distributed.  相似文献   

5.
We compared the effects of peripheral nerve grafts, embryonic spinal cord transplants and brain-derived neurotrophic factor (BDNF) on the survival and axon regeneration of adult rat spinal motor neurons undergoing retrograde degeneration after ventral root avulsion. Following implantation into the dorsolateral funiculus of the injured spinal cord segment, neither a peripheral nerve graft nor a combination of peripheral nerve graft with embryonic spinal cord transplant could prevent the retrograde motor neuron degeneration induced by ventral root avulsion. However, intrathecal infusion of BDNF promoted long-term survival of the lesioned motor neurons and induced abundant motor axon regeneration from the avulsion zone along the spinal cord surface towards the BDNF source. A combination of ventral root reconstitution and BDNF treatment might therefore be a promising means for the support of both motor neuron survival and guided motor axon regeneration after ventral root lesions.  相似文献   

6.
7.
This study examined the effects of nerve growth factor, brain-derived neurotrophic factor, neurotrophin-3 and neurotrophin-4/5 on substance P levels in dorsal root ganglia of the quail shortly after ganglia formation (stage 26, embryonic day 4.5), during the middle of development (stage 33, embryonic day 7.5) and during late development (stage 44, embryonic day 14). It has already been shown that nerve growth factor increases levels of substance P during the middle and late stages of development, and that messenger RNA for the neurotrophin receptors, trkA, trkB and trkC is present at all of these stages. Dorsal root ganglia were isolated, rinsed with defined medium to dilute endogenous neurotrophins and exposed to one of the neurotrophins for either 4 or 20 h. Substance P levels were quantitated using enzyme immunoassay. None of the neurotrophins had any effect on substance P levels in dorsal root ganglia obtained at stage 26 after either a 4 or 20 h exposure time. Nerve growth factor, brain-derived neurotrophic factor, neurotrophin-3 and neurotrophin-4/5 all significantly increased levels of substance P after either a 4 h or 20 h incubation in ganglia obtained at stages 33 and 44. The effects of nerve growth factor and neurotrophin-3 were specific: increases in substance P were completely blocked by simultaneous exposure to antibodies against either nerve growth factor or neurotrophin-3. The absence of any effect of neurotrophins on substance P expression during early development was unexpected, since dorsal root ganglia exhibit substantial levels of substance P and receptors for the neurotrophins are present and are apparently functional. It was also surprising that brain-derived neurotrophic factor, neurotrophin-3 and neurotrophin-4/5 induced increases in substance P levels during the middle and late stages of development, since substance P was thought to be exclusively localized to small TrkA neurons in dorsal root ganglia. However, immunocytochemical examination of dorsal root ganglia at stages 33 and 44 revealed substance P-like immunoreactivity in larger neurons as well as in small neurons. The results of this study have shown that different cellular responses to neurotrophins, such as effects on survival and/or peptide expression, may be acquired with differing temporal patterns not strictly related to expression of their receptors. Further, the regulation of neuropeptide synthesis in dorsal root ganglia is not due to any one neurotrophic factor. and the factors that regulate expression during early development are still unknown.  相似文献   

8.
Hepatocyte growth factor (HGF) has been shown to function as a potent mitogen for a variety of cells, transducing its signal through the c-met tyrosine kinase receptor. Ciliary neurotrophic factor (CNTF) is a cytokine that has been shown to promote survival of motor neurons. We show here that c-met mRNA is present in the embryonic rat spinal cord. Peak expression of c-met (at E14) coincides with the period of naturally occurring cell death in motor neurons, suggesting a possible role of HGF in the regulation of this process. Utilizing a neuron-enriched culture system, we established that HGF, like CNTF, stimulates choline acetyltransferase (CAT) activity in motor neurons. When co-administered to motor neuron cultures, saturating concentrations of HGF and CNTF produced a synergistic increase in CAT levels. We show that this synergy reflects enhanced motor neuron survival. Exposure of motor neuron cultures to the cytostatic agent vincristine markedly decreased CAT levels; co-treatment with HGF and CNTF (but not either factor alone) restored CAT activity to control levels. Our findings indicate that HGF is a survival factor for motor neurons, that it acts synergistically with CNTF, and that HGF and CNTF can together be neuroprotective in the face of vincristine toxicity.  相似文献   

9.
Epigenetic factors are known to influence the differentiation of neocortical neurons. The present study analyses the role of spontaneous bioelectrical activity (SBA) and neurotrophic factors on the expression of neuropeptide Y (NPY) in rat visual cortical neurons using organotypic monocultures prepared from newborn animals and in situ hybridization to detect the NPY messenger ribonucleic acid (mRNA). Spontaneously active cortex cultures display NPY mRNA expression in about 7% of all cortical neurons from 10 days in vitro (DIV) on. Blocking the SBA by chronic application of 10 mM Mg2+ for 3-30 DIV reduces the percentage of NPY neurons to about 2%. Allowing an initial phase of SBA (1-20 DIV) followed by an SBA blockade (for 21-50 DIV) results in 2% labelled neurons, indicating a dramatic reduction of NPY mRNA expression in the absence of SBA. Surprisingly, the reverse experiment (a period of SBA blockade for 1-20 DIV followed by a period of SBA recovery for 21-40 DIV) does not cause an upregulation of NPY mRNA expression. However, allowing cultures to differentiate as spontaneously active cultures, then applying a transient period of SBA blockade which is followed by a second period of SBA, does rescue the NPY mRNA expression in 7% of the cortical neurons. We conclude that SBA is a main trigger for NPY mRNA expression and it is particularly important during an early postnatal period of differentiation. We then analysed whether neurotrophic factors known to modulate cortical neuropeptide expression are able to do so in the absence of SBA. Supplementing chronically blocked cultures with the neurotrophins, brain-derived neurotrophic factor (BDNF), neurotrophin-3 (NT-3), neurotrophin-4/5 (NT-4/5) and the cytokine, leukaemia inhibitory factor (LIF), reveals that BDNF and NT-3 are unable to increase the percentage of NPY neurons. In contrast, LIF and NT-4/5 increase the percentage of NPY neurons to 4 and 6-7%, respectively. Moreover, neurons treated with NT-4/5 display a very high level of NPY mRNA expression in somata and in the dendritic trees. The data suggest a complex interplay and a hierarchy of epigenetic factors in regulating the neurochemical architecture of the developing neocortex.  相似文献   

10.
To identify production sites and action targets of neurotrophins during neurogenesis, we investigated immunoreactivities of neurotrophins and their tyrosine kinase receptors in the cerebral cortex of rat embryos. Two sets of ligand-receptor systems, brain-derived neurotrophic factor/TrkB and neurotrophin-3/TrkC, were expressed simultaneously in Cajal-Retzius, subplate neurons and ventricular multipotent stem cells at embryonic days 13 and 15. Intraventricular administration of brain-derived neurotrophic factor or neurotrophin-3 at embryonic day 16 markedly modulated microtubule-associated protein II and/or Hu protein expression in different ways in the cortical plate cells by embryonic day 20. These observations indicate the involvement of autocrine and/or local paracrine action of brain-derived neurotrophic factor and/or neurotrophin-3 during formation of the cerebral cortex.  相似文献   

11.
The capacity of embryonic spinal cord tissue in the repair of injured structure of spinal cord has been noted for years. In order to investigate the embryonic spinal cord graft in the repair of motor function of injured spinal cord, the embryonic spinal cord tissue was transplanted to the hemisection cavity in spinal cord in adult rat. One hundred adult Wistar Rats were used to simulate the hemisectional injury of spinal cord by drilling 2-3 mm cavity in lumbar enlargement. Sixty rats were treated with rat embryonic spinal cord tissue grafting while the other forty were chosen as control. The outcome was evaluated according the combined behavioural score (CBS) and motor evoked potential (MEP) in the 1, 2, 4 and 12 weeks. The grafting group was superior to the control as assessed by CBS (P < 0.05), especially within 4 weeks. (P < 0.01). The restoration of the latent peak of early wave(P1, N1) was better in the grafting group, too. This suggested that embryonic spinal cord graft could improve the recovery of motor function of injured spinal cord in adult rat. The effect of the embryonic spinal cord tissue graft might be concerned with its secretion of several kinds of neurotrophic factors, nerve growth factor, nerve transmitted factor, or adjustment of hormone.  相似文献   

12.
In the absence of descending spinal and supraspinal afferent inputs, neurons in the developing lumbar spinal cord of the chick embryo undergo regressive changes including cellular atrophy and degeneration between embryonic days 10 and 16. There are significant decreases in the number of motoneurons, interneurons, and sensory (dorsal root ganglion) neurons. Although there are several possible explanations for how afferents might regulate the maintenance of neuronal viability, we have focused attention on the putative role of neurotrophic agents in these events. Previous studies have shown that specific tissue extracts (e.g., muscle, brain), soluble proteins, growth factors, and trophic agents can promote the in vitro and in vivo survival of avian motoneurons during the period of natural cell death (embryonic days 6-10). Several of these agents were also effective following deafferentation. These included brain extract (BEX), muscle extract (MEX), conditioned medium from astrocyte cultures (ACM), as well as the following neurotrophic agents: nerve growth factor (NGF), brain-derived neurotrophic factor (BDNF), neurotrophin-3 (NT-3), S-100, insulin-like growth factor-I (IGF-I), ciliary neurotrophic factor (CNTF), platelet-derived growth factor (PDGF), basic fibroblast growth factor (bFGF), and leukemia inhibitory factor (CDF/LIF). Both transforming growth factor-beta (TGF-beta) and acidic fibroblast growth factor (aFGF) were ineffective. Although considerable more work is needed to determine which (and how) specific CNS-derived trophic agents regulate motoneuron survival, the present results are consistent with the notion that neurotrophic agents released from or modulated by synaptic inputs to target neurons promote neuronal differentiation and survival in the CNS.  相似文献   

13.
Developmental cell death in the nervous system usually is controlled by the availability of target-derived trophic factors. It is well established that dorsal root ganglia (DRG) neurons require the presence of their peripheral target for survival, but because of their central projections, it is possible that the spinal cord also may be required. Before examining this possibility in rat embryos, we first used terminal deoxynucleotidyl transferase-mediated biotinylated UTP nick end labeling (TUNEL) to determine that thoracic DRG cell death occurred from embryonic day 15 (E15) to E18. To determine the target requirements of DRG neurons, we used organotypic slice cultures of E15 thoracic trunk segments. After peripheral target removal, essentially all DRG neurons disappeared within 5 d. In contrast, after removal of the spinal cord, approximately half of the DRG neurons survived for at least 8 d. Hence, some E15 DRG neurons could survive without the spinal cord. However, those DRG neurons that died after spinal cord ablation apparently required trophic factors from both central and peripheral targets, because the presence of only one of these tissues was not adequate by itself to support this cell group. Addition of neurotrophin-3 (NT-3) to the culture medium rescued some DRG neurons after CNS removal, suggesting a possible role for NT-3 in vivo. In other experiments, cultures were established from older (E16) embryos, and essentially all neurons survived after spinal cord ablation, even without added factors. These and other experiments indicated that approximately 65% of DRG neurons are transiently dependent on the CNS early in development.  相似文献   

14.
It has been shown that abnormalities in axonal transport occur in several mouse models with motoneuron degeneration and also in the human disease amyotrophic lateral sclerosis. In this report, we have examined the potential of neurotrophic factors to act on axonal transport properties in a mouse mutant, progressive motor neuronopathy (pmn). This mouse mutant has been characterized as a "dying-back" motoneuronopathy, with a loss of motoneuron cell bodies and motor fibers. Retrograde transport to the spinal cord motoneurons was determined using fluorescent tracers either injected into the gastrocnemius muscle or applied directly onto the cut sciatic nerve. Because the rate of retrograde labeling was significantly reduced in the pmn, we examined the potential of neurotrophic factors to compensate for the impairment. Ciliary neurotrophic factor (CNTF), brain-derived neurotrophic factor (BDNF), and neurotrophin-3 (NT-3) but not glial-derived neurotrophic factor (GDNF) or nerve growth factor (NGF) were capable of significantly improving the rate of labeling. The differential effects of these factors agree with previous studies showing that molecules that promote cell survival do not necessarily compensate for axonal deficiency. Because impairment of axonal properties appears as an early event in motoneuron pathology, our results may have important clinical implications in the treatment of motoneuron diseases.  相似文献   

15.
Following spinal cord injury, projection neurons are frequently axotomized and many of the cells subsequently die. One goal in spinal injury research is to preserve damaged neurons so that ultimately they are accessible to regeneration-promoting strategies. Here we ask if neurotrophin treatment can prevent atrophy and death of axotomized sensory projection neurons. In adult rats, a hemisection was made in the thoracic spinal cord and axotomized neurons were retrogradely labelled with Fluoro-Gold. Four distinct populations of cells were identified in the lumbar spinal cord, and both numbers and sizes of labelled cells were assessed at different time points postlesion. A progressive and significant degeneration was observed over time with severe atrophy apparent in all cell populations and significant cell loss evident by 4 weeks postlesion. This time point was used to assess neurotrophin effects. Hemisected rats were treated with either neurotrophin 3 (NT-3) or brain-derived neurotrophic factor (BDNF, 12 microg/day for each), or a vehicle solution, delivered continuously to the lesion site via an osmotic minipump. Treatment with NT-3, but not BDNF, completely reversed cell atrophy in three of the four cell populations and also induced a significant increase in the number of surviving cells. In situ hybridization experiments showed trkB and trkC mRNA to be expressed in the majority of ascending spinal projection neurons, suggesting that these cells should be responsive to both BDNF and NT-3. However, only NT-3 treatment was neuroprotective, indicating that BDNF may not have reached the cell bodies of injured neurons. These results demonstrate that NT-3 may be of benefit in preventing the secondary cell loss that occurs following spinal injury.  相似文献   

16.
17.
Oligodendrocytes develop from a subpopulation of precursor cells within the ventral ventricular zone of the spinal cord. The molecular cues that direct this spatially and temporally restricted event seem to originate in part from structures ventral to and within the spinal cord. Here, we present evidence that the family of ligands termed neuregulins are necessary for the normal generation of mouse spinal cord oligodendrocytes. Oligodendrocytes mature in spinal cord explants from wild-type mice and mice heterozygotic for a null mutation in the neuregulin gene (NRG +/-) in a temporal sequence of developmental events that replicates that observed in vivo. However, in spinal cord explants derived from mice lacking neuregulin (NRG -/-), oligodendrocytes fail to develop. Addition of recombinant neuregulin to spinal cord explants from NRG -/- mice rescues oligodendrocyte development. In wild-type spinal cord explants, inhibitors of neuregulin mimic the inhibition of oligodendrocyte development that occurs in NRG -/- explants. In embryonic mouse spinal cord, neuregulins are present in motor neurons and the ventral ventricular zone where they likely exert their influence on early oligodendrocyte precursor cells.  相似文献   

18.
Growth factors are considered pivotal for the development, maintenance, and function of mesencephalic dopaminergic neurons. Recent studies have identified a plethora of growth factors which support the survival and differentiation of embryonic dopaminergic neurons. However, the exact cellular targets of these growth factors, and, thus, their precise mechanisms of action, remain largely unknown. To identify these cellular targets, we analysed, at the single cell level, growth factor-induced c-fos expression in dissociated mesencephalic cell cultures derived from a fos-lac Z transgenic mouse line. Pharmacological interference with cell-cell communication was utilized to control for direct growth factor effects. beta-Galactosidase-expressing cells were phenotypically characterized by immunocytochemistry to specific neural cell markers. Glia cell line-derived neurotrophic factor, basic fibroblast growth factor, brain-derived neurotrophic factor, and neurotrophin-3 directly induced Fos expression in differently sized, yet overlapping, populations of tyrosine hydroxylase-immunoreactive dopaminergic neurons. In an additional subpopulation of dopaminergic neurons, neurotrophin-3 induced fos-lac Z expression indirectly through a glutamate-mediated activation of N-methyl-D-aspartate receptors. Consistent with their proposed glial-mediated mode of action, transforming growth factor alpha and platelet-derived growth factor induced Fos expression predominantly in glia but only in a very small number of dopaminergic neurons. These findings demonstrate that individual dopaminergic neurons represent the direct targets of different sets of extracellular growth factors. Our findings further establish that growth factors affect dopaminergic neurons by indirect mechanisms which require specific cell-cell communication. These data also suggest a potential role for growth factors in the establishment of the morphological and functional diversity of midbrain dopaminergic neurons.  相似文献   

19.
The possibility that brain derived neurotrophic factor (BDNF) and insulin like growth factor-1 (IGF) induced neuroprotection is influenced by mechanisms involving nitric oxide was examined in a rat model of focal spinal cord injury. BDNF or IGF-I (0.1 microgram/10 microliters in phosphate buffer saline) was applied topically 30 min before injury on the exposed spinal cord followed by repeated doses of growth factors immediately before and 30 min after injury. Thereafter application of BDNF or IGF was carried out at every 1 h interval until sacrifice. Five hours after injury, the tissue pieces from the T9 segment were processed for nNOS immunostaining, edema and cell injury. Untreated injured rats showed a profound upregulation of nNOS which was most pronounced in the nerve cells of the ipsilateral side. A marked increase in the blood-spinal cord barrier (BSCB) permeability to 125I-albumin, water content and cell injury in these perifocal segments was also found. Pretreatment with BDNF and IGF significantly reduced the upregulation of nNOS in the spinal cord. This effect of the growth factors was most pronounced in the contralateral side. Rats treated with these neurotrophic factors showed much less signs of BSCB damage, edema and cell injury. These results suggest that BDNF and IGF pretreatment is neuroprotective in spinal cord injury and that these neurotrophic factors have the capacity to down regulate nNOS expression following trauma to the spinal cord. Our data provide new experimental evidences which suggest that BDNF and IGF may exert their potential neuroprotective effects probably via regulation of NOS activity.  相似文献   

20.
Dynorphin, an endogenous opioid, may contribute to secondary nervous tissue damage following spinal cord injury. The temporal and spatial distribution of preprodynorphin (PPD) mRNA expression in the injured rat spinal cord was examined by in situ hybridization. Rats were subjected to traumatic spinal cord injury at the T13 spinal segment using the weight-drop method. Motor function of these rats was evaluated by their ability to maintain their position on an inclined plane. Two double-labeling experiments revealed that increased PPD mRNA and dynorphin peptide expression were found exclusively in dorsal horn neurons. Neurons exhibiting an increase in the level of PPD mRNA were concentrated in the superficial laminae and the neck of dorsal horn within several spinal segments from the epicenter of the injury at 24 and 48 h after injury. A number of neurons showing increased PPD mRNA were found in gray matter adjacent to the injury areas. Segments caudal to the injury site exhibited a long-lasting elevation of PPD mRNA in neurons, compared to the rostral segments. The number of neurons expressing PPD mRNA in each rat was significantly positively correlated with its motor dysfunction. These findings suggest that increased expression of dynorphin mRNA and peptide in dorsal horn neurons occurs after traumatic spinal cord injury. This also supports the hypothesis that the dynorphin has a pathological role in secondary tissue damage and neurological dysfunction after spinal cord injury.  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