波谱学杂志, 2023, 40(2): 158-168 doi: 10.11938/cjmr20223036

研究论文

实验性脊髓损伤后大脑血管可塑性的磁共振血管尺寸成像研究

田雨1,2,3, 周臣2,3, 张亚男2,3,4, 王鹏4, 张彩云4, 宋天玮2,3,4, 钱俊超,1,2,3,*

1.安徽医科大学基础医学院,安徽 合肥 230032

2.中国科学院合肥肿瘤医院,安徽 合肥 230031

3.中国科学院合肥物质科学研究院,健康与医学技术研究所,医学物理与技术安徽省重点实验室,安徽 合肥 230031

4.滨州医学院,山东 烟台 264003

In vivo MR Vessel Size Imaging of Brain Vascular Plasticity After Experimental Spinal Cord Injury

TIAN Yu1,2,3, ZHOU Chen2,3, ZHANG Yanan2,3,4, WANG Peng4, ZHANG Caiyun4, SONG Tianwei2,3,4, QIAN Junchao,1,2,3,*

1. School of Basic Medical Sciences, Anhui Medical University, Hefei 230032, China

2. Hefei Cancer Hospital, Chinese Academy of Sciences, Hefei 230031, China

3. Anhui Province Key Laboratory of Medical Physics and Technology, Institute of Health and Medical Technology, Hefei Institutes of Physical Science, Chinese Academy of Sciences, Hefei 230031, China

4. Binzhou Medical University, Yantai 264003, China

通讯作者: *E-mail:qianjunchao@hmfl.ac.cn.

收稿日期: 2022-11-22   网络出版日期: 2023-01-29

基金资助: 国家自然科学基金资助项目(U1932158); 国家自然科学基金资助项目(81871085); 国家自然科学基金资助项目(82271519); 山东省自然科学基金肿瘤防治联合基金重点项目(ZR2019LZL018); 安徽省自然科学基金杰出青年项目(2208085J10); 中国科学院合肥大科学中心协同创新培育基金重要项目(2019HSC-CIP003); 中国博士后科学基金项目(2019M652403); 山东省博士后创新项目(202002048)

Corresponding authors: *E-mail:qianjunchao@hmfl.ac.cn.

Received: 2022-11-22   Online: 2023-01-29

摘要

本文应用7 T高场磁共振血管尺寸成像技术研究了大鼠脊髓半切损伤后大脑血管的可塑性.通过感兴趣区域分析发现,损伤4周后,与损伤部位同侧锥体相比,对侧锥体区域的平均血管直径、微血管密度和血管尺寸指数显著增加,提示对侧皮质脊髓束白质区域血管生成或活化.该结果经由免疫荧光实验证实,损伤4周后,对侧锥体区域血小板内皮细胞粘附分子和胶质纤维酸性蛋白的染色强度也显著增加.以上结果表明磁共振血管尺寸成像技术可为脊髓损伤后大脑相关区域的血管生成提供有价值的信息,并有可能成为诊断脊髓损伤患者脑部血管病变的新工具.

关键词: 大脑血管可塑性; 血管尺寸成像; 大鼠; 脊髓损伤; 高场磁共振成像

Abstract

In vivo magnetic resonance vessel size imaging at 7 T high field was used to explore the changes in cerebral vascular plasticity after spinal cord hemisection injury in rats. Region of interest (ROI) analysis showed that four weeks after injury, mean vessel diameter (mVD), microvascular density (Density) and vessel size index (VSI) values were significantly increased in the contralateral side compared with the ipsilateral pyramid of the injury site, suggesting angiogenesis or vascular activation in the white matter region of the contralateral corticospinal tract (CST). Immunofluorescence results showed that the intensity of staining for platelet endothelial cell adhesion molecule 1 (CD31) and glial fibrillary acidic protein (GFAP) also increased significantly in the contralateral pyramid four weeks after injury. These results suggested that magnetic resonance vessel size imaging could provide valuable information on neovascularization in brain after spinal cord injury and may be a novel tool to diagnose brain vascular pathologies in spinal cord injury patients.

Keywords: brain vascular plasticity; vessel size imaging; rats; spinal cord injury; high field magnetic resonance imaging

PDF (1229KB) 元数据 多维度评价 相关文章 导出 EndNote| Ris| Bibtex  收藏本文

本文引用格式

田雨, 周臣, 张亚男, 王鹏, 张彩云, 宋天玮, 钱俊超. 实验性脊髓损伤后大脑血管可塑性的磁共振血管尺寸成像研究[J]. 波谱学杂志, 2023, 40(2): 158-168 doi:10.11938/cjmr20223036

TIAN Yu. In vivo MR Vessel Size Imaging of Brain Vascular Plasticity After Experimental Spinal Cord Injury[J]. Chinese Journal of Magnetic Resonance, 2023, 40(2): 158-168 doi:10.11938/cjmr20223036

引言

脊髓损伤(spinal cord injury,SCI)会导致沿皮质脊髓和其他神经束的神经元轴突变性[1-4].然而,有研究表明SCI后,人和大鼠的大脑发生可塑性变化,可形成新的神经回路,从而在不同程度上恢复感觉和运动功能[5-8].虽然大脑神经可塑性变化的过程已被广泛研究,但具体机制仍不清楚.轴突芽生、突触重组和神经新生是影响神经回路重建的关键因素[9,10],而损伤部位相关区域的血管生成或重建可提供更多的氧气和营养物质,从而有助于轴突再生和神经新生[11,12].已有研究[3]表明,SCI后,大鼠大脑的锥体和大脑脚等区域可观察到星形胶质细胞的增殖和活化.激活的星形胶质细胞可刺激和招募成纤维细胞和内皮细胞,从而促进损伤区域血管生成[13].此外,SCI后,激活的星形胶质细胞将表达肝细胞生长因子(hepatocyte growth factor,HGF)和间质表皮转换因子(c-mesenchymal-epithelial transition factor,c-Met),促进包括脊髓在内的大鼠、小鼠及人体各种组织和器官的血管生成[14-16].因此,我们推测星形胶质细胞可能在SCI后的大脑相关脑区血管生成过程中发挥重要作用.

高场磁共振血管尺寸成像(magnetic resonance vessel size imaging)可用于检测大脑中微米级的微小血管[17].它是一种通过确定平均血管直径(mean vessel diameter,mVD)、微血管密度(Density)和血管尺寸指数(vessel size index,VSI)等参数来评估血管生成和形态变化的定量成像方法[18,19].其中,mVD和Density分别代表局部血管网络中微血管的大小和密度;VSI是一个定量指标,能够反映微血管直径的分布,有助于监测体内微血管的扩张和收缩[19,20].Ielacqua等[21]采用高场动态磁敏感对比磁共振成像(dynamic susceptibility contrast-magnetic resonance imaging,DSC-MRI)联合注射造影剂的方法,获得了脑淀粉样变性小鼠的弛豫位移指数Q(relaxation shift index Q),从而计算出大脑微血管密度.已有研究[22]表明,磁共振血管尺寸成像得到的mVD、Density和VSI与组织学的估计值具有良好的相关性,可用于评估大鼠脑胶质瘤的血管生成.在另一项研究中,Xu等[19]利用9.4 T高场磁共振血管尺寸成像,研究了阿尔茨海默病(Alzheimer’s disease,AD)转基因小鼠的脑血管大小和密度与AD进展的关系.高场磁共振血管尺寸成像可为传统MRI提供补充信息,但利用该技术对大鼠SCI后大脑血管可塑性变化的研究至今未见报道.

本文利用7 T高场磁共振血管尺寸成像技术研究了大鼠SCI后沿皮质脊髓束(corticospinal tract,CST)区域(经锥体、大脑脚、内囊和皮质的通路)的血管生成,并通过组织学观察进行证实.

1 实验方法

1.1 仪器与试剂

7 T小动物MRI扫描仪(Bruker公司);冰冻切片机(Leica公司,型号:CM3050S);正置荧光显微镜(Leica公司,型号:DM6 B);超小超顺磁性氧化铁(ultrasmall superparamagnetic iron oxide,USPIO)造影剂(上海羧菲生物医药有限公司).胶质纤维酸性蛋白(glial fibrillary acidic protein,GFAP)是星形胶质细胞中间丝的主要成分,是星形胶质细胞活化的常用标志物[23].血小板内皮细胞粘附分子(platelet endothelial cell adhesion molecule 1,CD31)是来自免疫球蛋白大家族的130KD I型跨膜糖蛋白,是血管生成的标志物[24].因此,选取GFAP抗体(Millipore公司,货号:MAB360)和CD31抗体(Abcam公司,货号:ab281583)分别标记星形胶质细胞和新生血管.相应的荧光二抗为山羊抗小鼠IgG抗体(Invitrogen公司,货号:A-11005)和山羊抗兔IgG抗体(Abcam公司,货号:ab150077).

1.2 大鼠脊髓半切模型的建立

本实验所涉及的动物实验程序均经由中国科学院合肥物质科学研究院实验动物伦理委员会批准.实验选取6只雄性Sprague-Dawley大鼠(购买于安徽医科大学动物实验中心),年龄约8周龄,体重在200~250 g范围内.实验开始前均在温度为22~26 ℃、相对湿度为50%~70%、12 h昼夜交替照明的环境中饲养一周. 采用3.5%剂量的异氟烷混合氧气诱导麻醉后,将大鼠俯卧放置在实验台上.胸髓T9节段半切手术的术中使用1.5%~2.0%剂量的异氟烷使大鼠保持麻醉状态.备皮消毒后,切开背部皮肤,分离皮下组织以暴露 T9-T10棘突和椎板.在T9处行背侧单侧椎板切除术,充分暴露脊髓.以后正中静脉为界线,用眼科显微剪刀横断右半侧脊髓[3].术后,将大鼠放置在单独的笼子里,可以自由获得水和食物.每天人工排空膀胱至少三次,直至大鼠排尿反射恢复.

1.3 体内磁共振血管尺寸成像

SCI术后4周,利用7 T小动物MRI扫描仪进行磁共振血管尺寸成像.选择4周的时间点与先前功能磁共振成像(functional MRI,fMRI)发现大脑有多个区域被激活的研究相吻合[25].采用3.5%剂量的异氟烷混合氧气诱导麻醉后,将大鼠俯卧放置在扫描床上.扫描期间,使用1.5% ~2.0%剂量的异氟烷使大鼠保持麻醉状态,在尾静脉中放置导管以注射造影剂,并使用生理监控设备监测其呼吸频率和体温.

T2*加权图像(T2*WI)采用2D梯度回波(gradient-echo,GE)序列快速小角度激发(fast low-angle shot,FLASH)成像获得,扫描参数如下:回波时间(echo time,TE)= 3.5 ms;重复时间(repetition time,TR)= 800 ms;翻转角= 50˚;层厚为1 mm;层数为24;视野范围(field of vision,FOV)= 25 mm×25 mm;矩阵=256×256,平均次数为2.T2加权图像(T2WI)采用自旋回波(spin-echo,SE)序列获得,扫描参数如下:TE = 35 ms;TR = 4 500 ms;层厚为1 mm;层数为24;FOV = 25 mm×25 mm;矩阵= 256×256,平均次数为2.弥散图像采用多次激发自旋平面回波成像(echo-planar imaging,EPI)序列获得,扫描参数如下:TE = 25 ms;TR = 3 000 ms;激发次数为8;层厚为1 mm;层数为24;FOV = 25 mm×25 mm;矩阵=128×128;平均次数为2.在注射USPIO[26]之前和之后5 min,通过重复SE和GE序列扫描获得横向弛豫率变化$\Delta {{R}_{2}}$$\Delta R_{2}^{*}$.此外,表观弥散系数(apparent diffusion coefficient,ADC)图像从30个方向的弥散加权自旋回波图像(b = 800 s/mm2)和参考图像(b ≈ 0 s/mm2)中获得.造影剂(0.02 mL/kg)的注射速度为0.02 mL/s.每只动物的MRI扫描时间约为2 h.

1.4 数据处理

本文使用ImageJ软件分析MRI数据集来量化本研究中观察到的皮质脊髓束的变化.

首先使用ImageJ软件在T2WI图像上勾勒出除皮肤和颅骨以外的大鼠脑模板.SCI会导致损伤部位以外的沿皮质脊髓束的变性[27],故而选取双侧大脑锥体(pyramid,PY)、大脑脚(cerebral peduncle,CP)、内囊(internal capsule,IC)和皮质(cortex,CTX)作为感兴趣区域(regions of interest,ROIs).并将T2WI上勾画的所有ROIs转移到mVDDensity和VSI图像上.

基于先前的描述计算每个ROI的和$\Delta {{R}_{2}}$$\Delta R_{2}^{*}$[21]

$\Delta {{R}_{2}}=\frac{1}{\text{TE}}\ln \left( \frac{{{S}_{\text{pre}}}}{{{S}_{\text{post}}}} \right)$
$\Delta R_{2}^{*}=\frac{1}{\text{TE}}\ln \left( \frac{S_{\text{pre}}^{*}}{S_{\text{post}}^{*}} \right)$

其中${{S}_{\text{pre}}}$${{S}_{\text{post}}}$代表SE图像造影前和造影后的信号强度,$S_{\text{pre}}^{*}$$S_{\text{post}}^{*}$代表GE图像造影前和造影后的信号强度.

每个ROI内的mVD、Density和VSI参考文献[19,21,22]中给出的计算公式:

$\text{mVD}=\frac{\Delta R_{2}^{*}}{\Delta {{R}_{2}}}$
$\text{Density}=329\frac{{{(\Delta {{R}_{2}})}^{3}}}{{{(\Delta R_{2}^{*})}^{2}}}$
$\text{VSI}=0.424{{\left( \frac{\text{ADC}}{\gamma \Delta \chi {{B}_{0}}} \right)}^{1/2}}{{\left( \frac{\Delta R_{2}^{*}}{\Delta R_{2}^{{}}} \right)}^{3/2}}$

其中,ADC是指在梯度系统的30个方向上得到的ADC的平均值;B0代表静磁场强度;γ代表质子的旋磁比;由血管内的造影剂引起的血管外和血管内的磁感应强度差异$\Delta \chi $被认为等于0.57 ppm[19].

1.5 免疫荧光实验

MRI数据采集完成后,分别使用生理盐水和4%多聚甲醛对大鼠进行心脏灌流,断头取脑,将脑组织放在4%多聚甲醛中固定.之后将完全固定的脑组织浸入30%蔗糖溶液中,待组织沉底后利用冰冻切片机进行切片,切片厚度为20 μm.选取与T2加权图像大致相同层面的脑片进行后续处理.

脑片经0.5% TritonX-100透膜后,用5%山羊血清封闭.用GFAP(1:250)和CD31(1:200)抗体4℃孵育过夜后,与相应的荧光二抗(山羊抗小鼠IgG抗体和山羊抗兔IgG抗体)避光孵育60 min,然后采用正置荧光显微镜进行观察.通过ImageJ软件来评估荧光的分布情况和表达强度,从而量化大鼠脑组织单位面积内GFAP和CD31蛋白表达水平.

1.6 数据分析

使用GraphPad Prism 9.0软件进行统计学分析,所有数据均以平均值±标准差(mean±SD)表示.大鼠脊髓半切损伤部位的对侧和同侧水平之间的统计学差异采用配对t检验进行分析,p<0.05表示差异有统计学意义.

2 结果与讨论

2.1 体内磁共振血管尺寸成像

注射USPIO前后T2WI的对比分析如图1所示,可观察到注射USPIO后上矢状窦内信号降低.损伤4周后大鼠大脑的代表性磁共振图像的对比分析如图2所示,我们发现同侧和对侧锥体区域的mVD、Density和VSI表现出明显差异.定量结果(表S1)显示,损伤4周后,与同侧锥体相比,对侧锥体的mVD、Density和VSI分别增加了65%±33%(p=0.002 6)、160%±104%(p=0.001)和83%±12%(p<0.000 1).然而,同侧和对侧大脑皮质、内囊和大脑脚区域的mVD、Density和VSI无显著性差异(图3).

图1

图1   注射USPIO后,T2WI图像中上矢状窦内信号降低.白色箭头指向上矢状窦

Fig. 1   Reduced signal in the superior sagittal sinus after USPIO injection on T2-weighted images. White arrows point to the superior sagittal sinus


图2

图2   脊髓半切损伤4周后,大鼠脑的代表性的不同层的磁共振图像.在T2WI图像上勾画ROIs,并转移到mVD、Density和VSI图像上.CTX:皮质,包括初级运动皮质(M1);IC:内囊;CP:大脑脚;PY:锥体

Fig. 2   Representative magnetic resonance images of different layers of a rat brain after spinal cord hemisection injury for four weeks. The ROIs were delineated on T2-weighted images and transferred to the mVD, Density and VSI images. CTX: cortex including primary motor cortex (M1); IC: internal capsule; CP: cerebral peduncle; PY: pyramid


图3

图3   脊髓半切损伤4周后,6只大鼠同侧与对侧脑区的各感兴趣区域的mVD、Density和VSI定量分析.CTX:皮层,包括初级运动皮质(M1);IC:内囊;CP:大脑脚;PY:锥体.柱状图由平均值±标准差表示,其中**表示p<0.01

Fig. 3   Quantitative analysis of mVD, Density, and VSI for each region of interest in the ipsilateral and contralateral brain regions of six rats after spinal cord hemisection injury for four weeks. CTX: cortex including primary motor cortex (M1); IC: internal capsule; CP: cerebral peduncle; PY: pyramid. Data were represented by mean ± standard deviation, ** p<0.01


2.2 新生血管和星形胶质细胞的免疫荧光检测

为了进一步证实MRI实验的结果,我们用免疫荧光染色标记新生血管和星形胶质细胞.对侧锥体区域免疫荧光染色结果与以往的研究[3]一致,即在同一区域观察到星形胶质细胞的增殖和活化.损伤4周后,与同侧锥体相比,对侧锥体区域 CD31和GFAP染色强度显著增加.然而,同侧和对侧大脑皮质、内囊和大脑脚区域的CD31和GFAP染色强度无显著性差异(图4).定量结果(表S2)显示,损伤4周后,与同侧锥体相比,对侧锥体区域 CD31和GFAP染色强度分别增加了116%±24%(p<0.000 1)和169%±49%(p=0.000 1)(图5).

图4

图4   脊髓半切损伤4周后,大鼠代表性脑区的免疫荧光染色.白色箭头指向反应性星形胶质细胞.CTX:皮质,包括初级运动皮质(M1);IC:内囊;CP:大脑脚;PY:锥体

Fig. 4   Immunofluorescence staining of representative brain region of a rat after spinal cord hemisection injury for four weeks. White arrows point reactive astrocytes. CTX: cortex including primary motor cortex (M1); IC: internal capsule; CP: cerebral peduncle; PY: pyramid


图5

图5   脊髓半切损伤4周后,6只大鼠同侧和对侧脑区CD31和GFAP荧光强度定量分析.CTX:皮质,包括初级运动皮质(M1);IC:内囊;CP:大脑脚;PY:锥体.柱状图由平均值±标准差表示,**表示p<0.01

Fig. 5   Quantitative analysis of immunofluorescence intensity of CD31 and GFAP in the ipsilateral and contralateral brain regions of 6 rats after spinal cord hemisection injury for four weeks. CTX: cortex including primary motor cortex (M1); IC: internal capsule; CP: cerebral peduncle; PY: pyramid. Data were represented by mean ± standard deviation, ** p<0.01


2.3 讨论

在这项研究中,我们利用7 T高场磁共振血管尺寸成像技术,研究了大鼠脊髓半切损伤后沿皮质脊髓束的大脑相关脑区血管可塑性,并通过免疫荧光实验进行证实.

SCI导致损伤部位以外的沿皮质脊髓束的变性[27],以及进一步的轴突芽生和血管新生等可塑性变化.基于磁共振血管尺寸成像的结果,我们发现与损伤部位同侧锥体相比,损伤4周后的对侧锥体表现出血管活化.另外,免疫荧光染色结果显示,同一区域激活的星形胶质细胞也显著增加.星形胶质细胞是中枢神经系统数量最多、分布最广的细胞[28],在维持中枢神经系统稳态和功能方面发挥重要作用[29].SCI后,星形胶质细胞活化增生形成胶质瘢痕[30],通过分泌抑制性因子,如硫酸软骨素蛋白聚糖(chondroitin sulfate proteoglycans,CSPGs)[31],抑制少突胶质前体细胞(oligodendrocyte progenitor cells,OPCs)分化[32]、轴突再生和再髓鞘化[33].然而,另一些研究[34]表明,星形胶质细胞增多和瘢痕形成并不总是对SCI有害的.在SCI的急性期,激活的星形胶质细胞会迅速迁移到损伤区域,从而限制炎症细胞的浸润,加强病变区域的收缩[35],抑制神经毒性炎症从周围组织向中枢神经实质扩散[36].同样,SCI后,星形胶质细胞可清除细胞外过量的谷氨酸,从而减弱兴奋性神经毒性损伤[3738].除了隔离毒性微环境外,一些研究[13]表明,中枢神经系统损伤后,反应性星形胶质细胞还可通过刺激和招募成纤维细胞和内皮细胞加速损伤区域血管生成.大鼠SCI两周后,在损伤区域发现反应性星形胶质细胞与新生血管共定位[39].另一项研究[14]发现,SCI 1周后,损伤区域周围激活的星形细胞中HGF和c-Met的表达同时上调,发挥多种神经营养作用.在SCI后的恢复阶段(亚急性期和慢性期),HGF可减弱炎症细胞浸润,保护神经元并促进损伤区域血管生成[16,40]. 此外,在中枢神经系统损伤后,激活的星形胶质细胞也可分泌血管内皮生长因子(vascular endothelial growth factor,VEGF)来增强内皮细胞的增殖,从而促进血管生成[41,42].小鼠索曼中毒三个月后,在受损脑区观察到伴随GFAP表达增加的血管生成[43],这与我们关于星形胶质细胞和血管生成之间关系的假说一致.因此,我们推测SCI后对侧锥体区域的血管生成可能与激活的星形胶质细胞有关.另一方面,星形胶质细胞也是血-脊髓屏障的重要组成部分,在维持血-脊髓屏障的正常结构和功能方面起着不可或缺的作用[44]. Whetstone等[45]发现,在SCI后血-脊髓屏障的恢复阶段,星形胶质细胞邻近的新生血管上葡萄糖转运蛋白-1(glucose transporter-1,GLUT-1)的表达量增加,相反,其他不表达GLUT-1的新生血管可能没有正常的屏障功能,这也说明星形胶质细胞在SCI后血管的重塑和功能恢复中发挥重要作用.此外,星形胶质细胞通过在突触发生、突触成熟中发挥重要作用来调节突触可塑性,是大脑神经可塑性的关键调节因子[46].例如,星形胶质细胞可以通过分泌神经营养因子,如富含半胱氨酸的酸性分泌蛋白(secreted protein acidic and rich in cysteine,SPARC),促进神经元的生长、控制突触形成并调节突触发生[47].星形胶质细胞通过释放各种神经活性物质(例如谷氨酸、GABA、D-丝氨酸和ATP等),作用于突触前和突触后元件来调控突触传递[48].

本研究具有一定的局限性,比如只在损伤4周后的时间点观察相关磁共振图像的变化.我们没有进行纵向研究的主要原因是本研究的重点是利用活体无创的血管尺寸成像技术研究脊髓损伤后大脑血管可塑性变化.因此,本研究并不能真正代表长期纵向研究.而纵向研究对于客观评价血管生成的变化至关重要.我们将在后续的研究中解决这些问题.

3 结论

本文通过7 T高场磁共振血管尺寸成像技术证明了大鼠脊髓半切损伤4周后,对侧皮质脊髓束白质区域(主要是锥体区域)的mVD、Density和VSI和同侧相比显著增加.此外,在同一区域发现更多激活的星形胶质细胞和新生血管.以上结果表明,磁共振血管尺寸成像可为SCI后大脑相关区域血管生成提供有价值的信息,并有可能成为诊断SCI患者脑部血管病变的新工具.

致谢

感谢“国家自然科学基金资助项目(U1932158,81871085,82271519);山东省自然科学基金肿瘤防治联合基金重点项目(ZR2019LZL018);安徽省自然科学基金杰出青年项目(2208085J10);中国科学院合肥大科学中心协同创新培育基金重要项目(2019HSC-CIP003);中国博士后科学基金项目(2019M652403);山东省博士后创新项目(202002048)”对本研究的支持.

利益冲突

附件材料

表S1 脊髓半切损伤4周后,6只大鼠同侧与对侧脑区的各感兴趣区域的mVD、Density和VSI的值.

表S2 脊髓半切损伤4周后,6只大鼠同侧和对侧脑区CD31和GFAP荧光强度定量分析.

(可在《波谱学杂志》期刊官网 http://magres.wipm.ac.cn获取)

参考文献

WRIGLEY P J, GUSTIN S M, MACEY P M, et al.

Anatomical changes in human motor cortex and motor pathways following complete thoracic spinal cord injury

[J]. Cereb Cortex, 2009, 19(1): 224-232.

DOI:10.1093/cercor/bhn072      PMID:18483004      [本文引用: 1]

A debilitating consequence of complete spinal cord injury (SCI) is the loss of motor control. Although the goal of most SCI treatments is to re-establish neural connections, a potential complication in restoring motor function is that SCI may result in anatomical and functional changes in brain areas controlling motor output. Some animal investigations show cell death in the primary motor cortex following SCI, but similar anatomical changes in humans are not yet established. The aim of this investigation was to use voxel-based morphometry (VBM) and diffusion tensor imaging (DTI) to determine if SCI in humans results in anatomical changes within motor cortices and descending motor pathways. Using VBM, we found significantly lower gray matter volume in complete SCI subjects compared with controls in the primary motor cortex, the medial prefrontal, and adjacent anterior cingulate cortices. DTI analysis revealed structural abnormalities in the same areas with reduced gray matter volume and in the superior cerebellar cortex. In addition, tractography revealed structural abnormalities in the corticospinal and corticopontine tracts of the SCI subjects. In conclusion, human subjects with complete SCI show structural changes in cortical motor regions and descending motor tracts, and these brain anatomical changes may limit motor recovery following SCI.

ILVESMäKI T, KOSKINEN E, BRANDER A, et al.

Spinal cord injury induces widespread chronic changes in cerebral white matter

[J]. Hum Brain Mapp, 2017, 38(7): 3637-3647.

DOI:10.1002/hbm.23619      PMID:28429407      [本文引用: 1]

Traumatic spinal cord injuries (SCIs) lead to axonal damage at the trauma site, as well as disconnections within the central nervous system. While the exact mechanisms of the long-term pathophysiological consequences of SCIs are not fully understood, it is known that neuronal damage and degeneration are not limited to the direct proximity of the trauma. Instead, the effects can be detected even in the cerebrum. We examined SCI-induced chronic brain changes with a case-control design using 32 patients and 70 control subjects. Whole-brain white matter (WM) tracts were assessed with diffusion tensor imaging (DTI). In addition, we analysed associations between DTI metrics and several clinical SCI variables. Whole-brain analyses were executed by tract-based spatial statistics (TBSS), with an additional complementary atlas-based analysis (ABA). We observed widespread, statistically significant (P ≤ 0.01) changes similar to neural degeneration in SCI patients, both in the corticospinal tract (CST) and beyond. In addition, associations between DTI metrics and time since injury were found with TBSS and ABA, implying possible long-term post-injury neural regeneration. Using the ABA approach, we observed a correlation between SCI severity and DTI metrics, indicating a decrease in WM integrity along with patient sensory or motor scores. Our results suggest a widespread neurodegenerative effect of SCI within the cerebrum that is not limited to the motor pathways. Furthermore, DTI-measured WM integrity of chronic SCI patients seemed to improve as time elapsed since injury. Hum Brain Mapp 38:3637-3647, 2017. © 2017 Wiley Periodicals, Inc.© 2017 Wiley Periodicals, Inc.

HE G, QIAN J.

Phase imaging of axonal integrity of cranial corticospinal tract in experimental spinal cord injury at 9.4T

[J]. Microsc Res Tech, 2017, 80(9): 1009-1017.

DOI:10.1002/jemt.v80.9      URL     [本文引用: 4]

COHEN-ADAD J, EL MENDILI M M, LEHéRICY S, et al.

Demyelination and degeneration in the injured human spinal cord detected with diffusion and magnetization transfer MRI

[J]. Neuroimage, 2011, 55(3): 1024-1033.

DOI:10.1016/j.neuroimage.2010.11.089      URL     [本文引用: 1]

MOXON K A, OLIVIERO A, AGUILAR J, et al.

Cortical reorganization after spinal cord injury: always for good?

[J]. Neuroscience, 2014, 283: 78-94.

DOI:10.1016/j.neuroscience.2014.06.056      PMID:24997269      [本文引用: 1]

Plasticity constitutes the basis of behavioral changes as a result of experience. It refers to neural network shaping and re-shaping at the global level and to synaptic contacts remodeling at the local level, either during learning or memory encoding, or as a result of acute or chronic pathological conditions. 'Plastic' brain reorganization after central nervous system lesions has a pivotal role in the recovery and rehabilitation of sensory and motor dysfunction, but can also be "maladaptive". Moreover, it is clear that brain reorganization is not a "static" phenomenon but rather a very dynamic process. Spinal cord injury immediately initiates a change in brain state and starts cortical reorganization. In the long term, the impact of injury - with or without accompanying therapy - on the brain is a complex balance between supraspinal reorganization and spinal recovery. The degree of cortical reorganization after spinal cord injury is highly variable, and can range from no reorganization (i.e. "silencing") to massive cortical remapping. This variability critically depends on the species, the age of the animal when the injury occurs, the time after the injury has occurred, and the behavioral activity and possible therapy regimes after the injury. We will briefly discuss these dependencies, trying to highlight their translational value. Overall, it is not only necessary to better understand how the brain can reorganize after injury with or without therapy, it is also necessary to clarify when and why brain reorganization can be either "good" or "bad" in terms of its clinical consequences. This information is critical in order to develop and optimize cost-effective therapies to maximize functional recovery while minimizing maladaptive states after spinal cord injury. Copyright © 2014 IBRO. Published by Elsevier Ltd. All rights reserved.

JURKIEWICZ M T, MIKULIS D J, MCILROY W E, et al.

Sensorimotor cortical plasticity during recovery following spinal cord injury: a longitudinal fMRI study

[J]. Neurorehabil Neural Repair, 2007, 21(6): 527-538.

DOI:10.1177/1545968307301872      URL     [本文引用: 1]

Background. Although the consequences of spinal cord injury (SCI) within the spinal cord and peripheral nervous system have been studied extensively, the influence of SCI on supraspinal structures during recovery remains largely unexplored. Objective. To assess temporal changes in cortical sensorimotor representations beginning in the subacute phase following SCI and determine if an association exists between the plastic changes within cortical sensorimotor areas and recovery of movement postinjury. Methods. Functional magnetic resonance imaging (fMRI) was used to study 6 SCI patients for 1 year, beginning shortly postinjury, and 10 healthy control individuals. During fMRI, individuals performed a simple self-paced wrist extension motor task. Recovery of movement was assessed using the American Spinal Injury Association (ASIA) Standard Neurological Classification of SCI. Results. In the subacute period post-SCI, during impaired movement, little task-related activation within the primary motor cortex (M1) was present, whereas activation in associated cortical sensorimotor areas was more extensive than in controls. During motor recovery, a progressive enlargement in the volume of movement-related M1 activation and decreased activation in associated cortical sensorimotor areas was seen. When movement was performed with little to no impairment, the overall pattern of cortical activation was similar to that observed in control individuals. Conclusions. This study provides the first report of the temporal progression of cortical sensorimotor representational plasticity during recovery following traumatic SCI in humans and suggests an association between movement-related fMRI activation and motor recovery postinjury. These findings have implications on current and future rehabilitative interventions for patients with SCI.

FASSETT H J, TURCO C V, EL-SAYES J, et al.

Alterations in Motor Cortical representation of muscles following incomplete spinal cord injury in humans

[J]. Brain Sci, 2018, 8(12).

[本文引用: 1]

WANG L, CHEN N.

MRI research progresses of motor imagery on brain activity and network reorganization in patients with spinal cord injury

[J]. Chin J Med Imaging Technol, 2019, 35(10):1586-1589.

[本文引用: 1]

王玲, 陈楠.

运动想象对脊髓损伤患者大脑活动和脑网络重塑MR研究进展

[J]. 中国医学影像技术, 2019, 35(10): 1586-1589.

[本文引用: 1]

TSUJIOKA H, YAMASHITA T.

Neural circuit repair after central nervous system injury

[J]. Int Immunol, 2021, 33(6): 301-309.

DOI:10.1093/intimm/dxaa077      PMID:33270108      [本文引用: 1]

Central nervous system injury often causes lifelong impairment of neural function, because the regenerative ability of axons is limited, making a sharp contrast to the successful regeneration that is seen in the peripheral nervous system. Nevertheless, partial functional recovery is observed, because axonal branches of damaged or undamaged neurons sprout and form novel relaying circuits. Using a lot of animal models such as the spinal cord injury model or the optic nerve injury model, previous studies have identified many factors that promote or inhibit axonal regeneration or sprouting. Molecules in the myelin such as myelin-associated glycoprotein, Nogo-A or oligodendrocyte-myelin glycoprotein, or molecules found in the glial scar such as chondroitin sulfate proteoglycans, activate Ras homolog A (RhoA) signaling, which leads to the collapse of the growth cone and inhibit axonal regeneration. By contrast, axonal regeneration programs can be activated by many molecules such as regeneration-associated transcription factors, cyclic AMP, neurotrophic factors, growth factors, mechanistic target of rapamycin or immune-related molecules. Axonal sprouting and axonal regeneration largely share these mechanisms. For functional recovery, appropriate pruning or suppressing of aberrant sprouting are also important. In contrast to adults, neonates show much higher sprouting ability. Specific cell types, various mouse strains and different species show higher regenerative ability. Studies focusing on these models also identified a lot of molecules that affect the regenerative ability. A deeper understanding of the mechanisms of neural circuit repair will lead to the development of better therapeutic approaches for central nervous system injury.© The Japanese Society for Immunology. 2020. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com.

LIU J, YANG X, JIANG L, et al.

Neural plasticity after spinal cord injury

[J]. Neural Regen Res, 2012, 7(5): 386-391.

DOI:10.3969/j.issn.1673-5374.2012.05.010      PMID:25774179      [本文引用: 1]

Plasticity changes of uninjured nerves can result in a novel neural circuit after spinal cord injury, which can restore sensory and motor functions to different degrees. Although processes of neural plasticity have been studied, the mechanism and treatment to effectively improve neural plasticity changes remain controversial. The present study reviewed studies regarding plasticity of the central nervous system and methods for promoting plasticity to improve repair of injured central nerves. The results showed that synaptic reorganization, axonal sprouting, and neurogenesis are critical factors for neural circuit reconstruction. Directed functional exercise, neurotrophic factor and transplantation of nerve-derived and non-nerve-derived tissues and cells can effectively ameliorate functional disturbances caused by spinal cord injury and improve quality of life for patients.

YU S, YAO S, WEN Y, et al.

Angiogenic microspheres promote neural regeneration and motor function recovery after spinal cord injury in rats

[J]. Sci Rep, 2016, 6: 33428.

DOI:10.1038/srep33428      PMID:27641997      [本文引用: 1]

This study examined sustained co-delivery of vascular endothelial growth factor (VEGF), angiopoietin-1 and basic fibroblast growth factor (bFGF) encapsulated in angiogenic microspheres. These spheres were delivered to sites of spinal cord contusion injury in rats, and their ability to induce vessel formation, neural regeneration and improve hindlimb motor function was assessed. At 2-8 weeks after spinal cord injury, ELISA-determined levels of VEGF, angiopoietin-1, and bFGF were significantly higher in spinal cord tissues in rats that received angiogenic microspheres than in those that received empty microspheres. Sites of injury in animals that received angiogenic microspheres also contained greater numbers of isolectin B4-binding vessels and cells positive for nestin or beta III-tubulin (P < 0.01), significantly more NF-positive and serotonergic fibers, and more MBP-positive mature oligodendrocytes. Animals receiving angiogenic microspheres also suffered significantly less loss of white matter volume. At 10 weeks after injury, open field tests showed that animals that received angiogenic microspheres scored significantly higher on the Basso-Beattie-Bresnahan scale than control animals (P < 0.01). Our results suggest that biodegradable, biocompatible PLGA microspheres can release angiogenic factors in a sustained fashion into sites of spinal cord injury and markedly stimulate angiogenesis and neurogenesis, accelerating recovery of neurologic function.

POTENTE M, GERHARDT H, CARMELIET P.

Basic and therapeutic aspects of angiogenesis

[J]. Cell, 2011, 146(6): 873-887.

DOI:10.1016/j.cell.2011.08.039      PMID:21925313      [本文引用: 1]

Blood vessels form extensive networks that nurture all tissues in the body. Abnormal vessel growth and function are hallmarks of cancer and ischemic and inflammatory diseases, and they contribute to disease progression. Therapeutic approaches to block vascular supply have reached the clinic, but limited efficacy and resistance pose unresolved challenges. Recent insights establish how endothelial cells communicate with each other and with their environment to form a branched vascular network. The emerging principles of vascular growth provide exciting new perspectives, the translation of which might overcome the current limitations of pro- and antiangiogenic medicine.Copyright © 2011 Elsevier Inc. All rights reserved.

LI X, LI M, TIAN L, et al.

Reactive astrogliosis: implications in spinal cord injury progression and therapy

[J]. Oxid Med Cell Longev, 2020, 2020: 9494352.

[本文引用: 2]

SHIMAMURA M, SATO N, SATA M, et al.

Expression of hepatocyte growth factor and c-Met after spinal cord injury in rats

[J]. Brain Res, 2007, 1151: 188-194.

PMID:17425951      [本文引用: 2]

Since hepatocyte growth factor (HGF) plays a pivotal role in the development of the central nervous system and pathological conditions, we examined the long-term changes in the mRNA and protein expression of HGF and its receptor c-Met after spinal cord injury (SCI) in rats. HGF mRNA was significantly increased from 7 days after SCI in the injured segment, and the peak was at 7 days after SCI as assessed by real-time RT-PCR. Importantly, c-met mRNA expression was up-regulated from 1 day after SCI, and reached a peak at 14 days after SCI. Although up-regulation of HGF and c-met mRNA expression in the injured segment gradually decreased, the increased expression level persisted until 56 days after SCI. Consistent with HGF mRNA expression, HGF protein level was significantly increased mainly in the injured region, which persisted until 56 days after SCI. Immunohistochemistry showed that most of GFAP-positive reactive astrocytes expressed HGF and c-Met both on 14 days and 56 days after SCI. Staining with the mitotic indicator, bromodeoxyuridine (BrdU), revealed that a small number of BrdU-incorporated cells were co-localized with HGF/GFAP-positive or c-Met/GFAP-positive cells both on 14 and 56 days. These data suggest that HGF and c-Met were up-regulated mainly in the reactive astrocytes around the injured region in the subacute to chronic stage of spinal cord injury. Since HGF plays a critical role in neurotrophic activity, activation of the HGF/c-Met signaling system might be involved in the process of post-traumatic regeneration.

NAKAMURA T, MIZUNO S.

The discovery of hepatocyte growth factor (HGF) and its significance for cell biology, life sciences and clinical medicine

[J]. Proc Jpn Acad Ser B Phys Biol Sci, 2010, 86(6): 588-610.

DOI:10.2183/pjab.86.588      URL     [本文引用: 1]

KITAMURA K, IWANAMI A, NAKAMURA M, et al.

Hepatocyte growth factor promotes endogenous repair and functional recovery after spinal cord injury

[J]. J Neurosci Res, 2007, 85(11): 2332-2342.

PMID:17549731      [本文引用: 2]

Many therapeutic interventions using neurotrophic factors or pharmacological agents have focused on secondary degeneration after spinal cord injury (SCI) to reduce damaged areas and promote axonal regeneration and functional recovery. Hepatocyte growth factor (HGF), which was identified as a potent mitogen for mature hepatocytes and a mediator of inflammatory responses to tissue injury, has recently been highlighted as a potent neurotrophic and angiogenic factor in the central nervous system (CNS). In the present study, we revealed that the extent of endogenous HGF up-regulation was less than that of c-Met, an HGF receptor, during the acute phase of SCI and administered exogenous HGF into injured spinal cord using a replication-incompetent herpes simplex virous-1 (HSV-1) vector to determine whether HGF exerts beneficial effects and promotes functional recovery after SCI. This treatment resulted in the significant promotion of neuron and oligodendrocyte survival, angiogenesis, axonal regrowth, and functional recovery after SCI. These results suggest that HGF gene delivery to the injured spinal cord exerts multiple beneficial effects and enhances endogenous repair after SCI. This is the first study to demonstrate the efficacy of HGF for SCI.Copyright 2007 Wiley-Liss, Inc.

CHRISTOFORIDIS G A, YANG M, KONTZIALIS M S, et al.

High resolution ultra high field magnetic resonance imaging of glioma microvascularity and hypoxia using ultra-small particles of iron oxide

[J]. Invest Radiol, 2009, 44(7): 375-383.

DOI:10.1097/RLI.0b013e3181a8afea      PMID:19448552      [本文引用: 1]

This study assessed whether ultra-small particles of iron oxide (USPIO) intravascular contrast agent could enhance visualization of tumor microvascularity in F98 glioma bearing rats by means of ultra high field (UHF) high-resolution gradient echo (GRE) magnetic resonance imaging (MRI). In an effort to explain differences in visualization of microvascularity before and after USPIO administration, hypoxia and vessel diameters were assessed on corresponding histopathologic sections.F98 glioma cells were implanted stereotactically into the brains of syngeneic Fischer rats. Based on clinical criteria, rats were imaged 1 to 2 days before their death with and without USPIO contrast on an 8 Tesla MRI. To identify hypoxic regions of the brain tumor by immunohistochemical staining, a subset of animals also received a nitroimidazole-based hypoxia marker, EF5, before euthanasia. These sections then were compared with noncontrast enhanced MR images. The relative caliber of tumor microvasculature, compared with that of normal brain, was analyzed in a third group of animals.After USPIO administration, UHF high-resolution GRE MRI consistently predicted increased microvascular density relative to normal gray matter when correlated with histopathology. The in-plane visibility of glioma microvascularity in 22 rats increased by an average of 115% and signal intensity within the tumor decreased by 13% relative to normal brain. Tumor microvascularity identified on noncontrast MR images matched hypoxic regions identified by immunohistochemical staining with a sensitivity of 83% and specificity of 89%. UHF GRE MRI was able to resolve microvessels less than 20 micro in diameter, although differences in tumor vessel size did not consistently account for differences in visualization of microvascularity.USPIO administration significantly enhanced visualization of tumor microvascularity on gradient echo 8 T MRI and significantly improved visualization of tumor microvascularity. Microvascularity identified on precontrast images is suspected to be partly associated with hypoxia.

XU C, SCHMIDT W U, VILLRINGER K, et al.

Vessel size imaging reveals pathological changes of microvessel density and size in acute ischemia

[J]. J Cereb Blood Flow Metab, 2011, 31(8): 1687-1695.

DOI:10.1038/jcbfm.2011.38      URL     [本文引用: 1]

The aim of this study was to test the feasibility of vessel size imaging with precise evaluation of apparent diffusion coefficient and cerebral blood volume and to apply this novel technique in acute stroke patients within a pilot group to observe the microvascular responses in acute ischemic tissue. Microvessel density-related quantity Q and mean vessel size index ( VSI) were assessed in 9 healthy volunteers and 13 acute stroke patients with vessel occlusion within 6hours after symptom onset. Our results in healthy volunteers matched with general anatomical observations. Given the limitation of a small patient cohort, the median VSI in the ischemic area was higher than that in the mirrored region in the contralateral hemisphere ( P &lt; 0.05). Decreased Q was observed in the ischemic region in 2 patients, whereas no obvious changes of Q were found in the remaining 11 patients. In a patient without recanalization, the VSI hyperintensity in the subcortical area matched well with the final infarct. These data reveal that different observations of microvascular response in the acute ischemic tissue seem to emerge and vessel size imaging may provide useful information for the definition of ischemic penumbra and have an impact on future therapeutic approaches.

XU X, MENG T, WEN Q, et al.

Dynamic changes in vascular size and density in transgenic mice with Alzheimer's disease

[J]. Aging (Albany NY), 2020, 12(17): 17224-17234.

[本文引用: 5]

JIANG J J, ZHAO L Y, WANG C Y, et al.

Initial application of magnetic resonance vessel size imaging in cerebral glioma and meningioma

[J]. Radiol Practic, 2014, 29(7): 770-773.

[本文引用: 1]

江晶晶, 赵凌云, 王承缘, .

磁共振血管大小成像在脑胶质瘤和脑膜瘤中的初步应用

[J]. 放射学实践, 2014, 29(7): 770-773.

[本文引用: 1]

IELACQUA G D, SCHLEGEL F, FüCHTEMEIER M, et al.

Magnetic resonance Q mapping reveals a decrease in microvessel density in the arcAβ mouse model of cerebral amyloidosis

[J]. Front Aging Neurosci, 2015, 7: 241.

DOI:10.3389/fnagi.2015.00241      PMID:26834622      [本文引用: 3]

Aterations indensity and morphology of the cerebral microvasculature have been reported to occur in Alzheimer's disease patients and animal models of the disease. In this study we compared magnetic resonance imaging (MRI) techniques for the irutility to detectage-dependent changes of the cerebralvasculature in the arc Ab mouse model of cerebralamyloidosis. Dynamic susceptibility contrast (DSC)-MRI was performed by tracking the passage of a super paramagnetic iron oxide nanoparticle in the brain with dynamic gradientechoplanarimaging(EPI). From this measurements relative cerebral blood volume rCBV(DSC) and relative cerebral blood flow (rCBF) were estimated. For the same animal maps of the relaxation shift index Q were computed from high resolution gradientecho and spine chodata that were acquired before and after superparamagnetic iron oxide(SPIO )nanoparticle injection. Q-values were used to derive estimates of microves seldensity. The change in the relaxation rates Delta R*(2) obtained from pre- and Delta*(2) post-contrast gradientecho data was used for the alternative determination of rCBV rCBV(Delta R*(2)). L near mixed effects modeling found no significant association between rCBV(DSC),rCBV(Delta R*(2)),rCBF,and Q with genotype in 1-montholdmicecompared to age-matchednon-transgenic litter mates(NTLs) for any of the evaluated brain regions. In 24-month old mice there was a significant association for rCBV(DSC) with genotype in the cerebralcortex, and for rCBV(Delta R*(2)) in the cerebralcortex and cerebellum. For rCBF there was a significant association in the cerebellum but not in other brain regions. Q-values in the olfactory bulb, cerebral cortex, striatum, hippocampus, and cerebellumin 24-month old mice were significantly associated with genotype. In those regions Q-values were reduced between 11 and 26% in arc Ab mice comparedtoage-matched NTLs. Vessel staining with CD31 immunohistochemistry confirmed are duction of microvessel density in the old arc Ab mice. We further demonstrated a region-specific association between parenchymaland vascularde position of beta-amyloidand decreased vasculardensity, without a correlation with the amount of Abde position.We found that Q mapping was more suitable than the he modynamicread-outs to detectamyloid-related de generation of the cerebral microvasculature.

LEMASSON B, VALABLE S, FARION R, et al.

In vivo imaging of vessel diameter, size, and density: a comparative study between MRI and histology

[J]. Magn Reson Med, 2013, 69(1): 18-26.

DOI:10.1002/mrm.24218      PMID:22431289      [本文引用: 2]

The aim of this study was to compare magnetic resonance imaging (MRI) and histological estimates of the mean vessel diameter (mVD), the vessel density (Density), and the vessel size index (VSI) obtained in the same tumor-bearing animals. Twenty-seven rats bearing intracranial glioma (C6 or RG2) were imaged by MRI. Changes in transverse relaxations (ΔR 2* and R(2)) were induced by the injection of an iron-based contrast agent and were mapped using a multi gradient-echo spin-echo sequence. Then, brain vascular network was studied ex vivo by histology. Three regions of interest were drawn in apparently normal tissue (neocortex and striatum) and in the tumor. In vivo mVD(MRI), Density(MRI), and VSI(MRI) were measured; ex vivo, mVD(histo), Density(histo), and VSI(histo) were quantified on the same animals. MRI and histology measurements differed by -15 to 26%. A positive correlation was found between MRI and histology for mVD, Density, and VSI counterparts (R(2) = 0.62, 0.50, 0.73, respectively; P < 0.001 in all cases). This study indicates that MRI and histology yields well correlated the estimates of mVD, Density, and VSI. VSI is the closest MRI estimate to histology. As Density and mVD or VSI provide complementary information, it is worth computing them to characterize angiogenesis beyond blood volume fraction.Copyright © 2012 Wiley Periodicals, Inc.

HOL E M, PEKNY M.

Glial fibrillary acidic protein (GFAP) and the astrocyte intermediate filament system in diseases of the central nervous system

[J]. Curr Opin Cell Biol, 2015, 32: 121-130.

DOI:10.1016/j.ceb.2015.02.004      PMID:25726916      [本文引用: 1]

Glial fibrillary acidic protein (GFAP) is the hallmark intermediate filament (IF; also known as nanofilament) protein in astrocytes, a main type of glial cells in the central nervous system (CNS). Astrocytes have a range of control and homeostatic functions in health and disease. Astrocytes assume a reactive phenotype in acute CNS trauma, ischemia, and in neurodegenerative diseases. This coincides with an upregulation and rearrangement of the IFs, which form a highly complex system composed of GFAP (10 isoforms), vimentin, synemin, and nestin. We begin to unravel the function of the IF system of astrocytes and in this review we discuss its role as an important crisis-command center coordinating cell responses in situations connected to cellular stress, which is a central component of many neurological diseases. Copyright © 2015 Elsevier Ltd. All rights reserved.

RAZAVI S M, YAHYAABADI R.

Comparative study of correlation between angiogenesis markers (CD31) and Ki67 marker with behavior of aggressive and nonaggressive central giant cell granuloma with immunohistochemistry technique

[J]. Asian Pac J Cancer Prev, 2018, 19(8): 2279-2283.

[本文引用: 1]

RAMU J, BOCKHORST K H, MOGATADAKALA K V, et al.

Functional magnetic resonance imaging in rodents: Methodology and application to spinal cord injury

[J]. J Neurosci Res, 2006, 84(6): 1235-1244.

PMID:16941500      [本文引用: 1]

Functional MRI (fMRI) on spinal cord-injured rodents at 4 and 8 weeks post injury (PI) is described. The paradigm for fMRI, based on electrical stimulation of rat paws, was automated using an in-house designed microprocessor-based controller that was interfaced to a stimulator. The MR images were spatially normalized to the Paxinos and Watson atlas using publicly available digital images of the cryosections. In normal uninjured animals, the activation was confined to the contralateral somatosensory cortex. In contrast, in injured animals, extensive activation, which included structures such as ipsilateral cortex, thalamus, hippocampus, and the caudate putamen, was observed at 4 and 8 weeks PI. Quantitative cluster analysis was carried out to calculate the volumes and centers of activation in individual brain structures. Based on this analysis, significant increase in activation between 4 and 8 weeks was observed only in the ipsilateral caudate putamen and thalamus. These studies suggest extensive and ongoing brain reorganization in spinal cord-injured animals.Copyright 2006 Wiley-Liss, Inc.

SHEN Y M, ZHENG W L, CHENG Y-C N, et al.

USPIO high resolution neurovascular imaging in a rat stroke model of transient middle cerebral artery occlusion

[J]. Chinese J Magn Reson, 2014, 31(1): 20-31.

[本文引用: 1]

沈伊民, 郑伟丽, CHENG Y-C N, .

大鼠中风模型的超小氧化铁粒子神经血管成像

[J]. 波谱学杂志, 2014, 31(1): 20-31.

[本文引用: 1]

JIRJIS M B, VEDANTAM A, BUDDE M D, et al.

Severity of spinal cord injury influences diffusion tensor imaging of the brain

[J]. J Magn Reson Imaging, 2016, 43(1): 63-74.

DOI:10.1002/jmri.24964      PMID:26094789      [本文引用: 2]

The purpose of this study was to determine whether DTI changes in the brain induced by a thoracic spinal cord injury are sensitive to varying severity of spinal contusion in rats.A control, mild, moderate, or severe contusion injury was administered over the eighth thoracic vertebral level in 32 Sprague-Dawley rats. At 11 weeks postinjury, ex vivo DTI of the brain was performed on a 9.4T Bruker scanner using a pulsed gradient spin-echo sequence.Mean water diffusion in the internal capsule regions of the brain and pyramid locations of the brainstem were correlated with motor function (r(2)  = 0.55). Additionally, there were significant differences between injury severity groups for mean diffusivity and fractional anisotropy at regions associated with the corticospinal tract (P = 0.05).These results indicate that DTI is sensitive to changes in brain tissue as a consequence of thoracic SCI.© 2015 Wiley Periodicals, Inc.

LIDDELOW S A, BARRES B A.

Reactive astrocytes: production, function, and therapeutic potential

[J]. Immunity, 2017, 46(6): 957-967.

DOI:S1074-7613(17)30234-0      PMID:28636962      [本文引用: 1]

Astrocytes constitute approximately 30% of the cells in the mammalian central nervous system (CNS). They are integral to brain and spinal-cord physiology and perform many functions important for normal neuronal development, synapse formation, and proper propagation of action potentials. We still know very little, however, about how these functions change in response to immune attack, chronic neurodegenerative disease, or acute trauma. In this review, we summarize recent studies that demonstrate that different initiating CNS injuries can elicit at least two types of "reactive" astrocytes with strikingly different properties, one type being helpful and the other harmful. We will also discuss new methods for purifying and investigating reactive-astrocyte functions and provide an overview of new markers for delineating these different states of reactive astrocytes. The discovery that astrocytes have different types of reactive states has important implications for the development of new therapies for CNS injury and diseases.Copyright © 2017 Elsevier Inc. All rights reserved.

SOFRONIEW M V.

Reactive astrocytes in neural repair and protection

[J]. Neuroscientist, 2005, 11(5): 400-407.

DOI:10.1177/1073858405278321      PMID:16151042      [本文引用: 1]

Reactive astrocytosis occurs prominently in response to all forms of CNS injury or disease. The functions of reactive astrocytes are not well understood, and both harmful and beneficial activities have been attributed to these cells. The basic process of reactive astrocytosis is conserved in vertebrate evolution, suggesting fitness-enhancing benefits, but scar-forming reactive astrocytes are often regarded as uniformly detrimental to clinical outcome, in particular, when implicated as inhibitors of axon regeneration. Transgenic mouse models are providing new means to study the activities of reactive astrocytes after CNS insults in vivo. Recent studies point toward roles for reactive astrocytes in restricting inflammation and protecting neurons and oligodendrocytes, thereby helping to limit tissue degeneration and preserve function after CNS injury.

OKADA S, HARA M, KOBAYAKAWA K, et al.

Astrocyte reactivity and astrogliosis after spinal cord injury

[J]. Neurosci Res, 2018, 126: 39-43.

DOI:S0168-0102(17)30592-8      PMID:29054466      [本文引用: 1]

After traumatic injuries of the central nervous system (CNS), including spinal cord injury (SCI), astrocytes surrounding the lesion become reactive and typically undergo hypertrophy and process extension. These reactive astrocytes migrate centripetally to the lesion epicenter and aid in the tissue repair process, however, they eventually become scar-forming astrocytes and form a glial scar which produces axonal growth inhibitors and prevents axonal regeneration. This sequential phenotypic change has long been considered to be unidirectional and irreversible; thus glial scarring is one of the main causes of the limited regenerative capability of the CNS. We recently demonstrated that the process of glial scar formation is regulated by environmental cues, such as fibrotic extracellular matrix material. In this review, we discuss the role and mechanism underlying glial scar formation after SCI as well as plasticity of astrogliosis, which helps to foster axonal regeneration and functional recovery after CNS injury.Copyright © 2017 Elsevier Ireland Ltd and Japan Neuroscience Society. All rights reserved.

HUSSEIN R K, MENCIO C P, KATAGIRI Y, et al.

Role of chondroitin sulfation following spinal cord injury

[J]. Front Cell Neurosci, 2020, 14: 208.

DOI:10.3389/fncel.2020.00208      PMID:32848612      [本文引用: 1]

Traumatic spinal cord injury produces long-term neurological damage, and presents a significant public health problem with nearly 18,000 new cases per year in the U.S. The injury results in both acute and chronic changes in the spinal cord, ultimately resulting in the production of a glial scar, consisting of multiple cells including fibroblasts, macrophages, microglia, and reactive astrocytes. Within the scar, there is an accumulation of extracellular matrix (ECM) molecules-primarily tenascins and chondroitin sulfate proteoglycans (CSPGs)-which are considered to be inhibitory to axonal regeneration. In this review article, we discuss the role of CSPGs in the injury response, especially how sulfated glycosaminoglycan (GAG) chains act to inhibit plasticity and regeneration. This includes how sulfation of GAG chains influences their biological activity and interactions with potential receptors. Comprehending the role of CSPGs in the inhibitory properties of the glial scar provides critical knowledge in the much-needed production of new therapies.Copyright © 2020 Hussein, Mencio, Katagiri, Brake and Geller.

WANG Y, CHENG X, HE Q, et al.

Astrocytes from the contused spinal cord inhibit oligodendrocyte differentiation of adult oligodendrocyte precursor cells by increasing the expression of bone morphogenetic proteins

[J]. J Neurosci, 2011, 31(16): 6053-6058.

DOI:10.1523/JNEUROSCI.5524-09.2011      PMID:21508230      [本文引用: 1]

Promotion of remyelination is an important therapeutic strategy to facilitate functional recovery after traumatic spinal cord injury (SCI). Transplantation of neural stem cells (NSCs) or oligodendrocyte precursor cells (OPCs) has been used to enhance remyelination after SCI. However, the microenvironment in the injured spinal cord is inhibitory for oligodendrocyte (OL) differentiation of NSCs or OPCs. Identifying the signaling pathways that inhibit OL differentiation in the injured spinal cord could lead to new therapeutic strategies to enhance remyelination and functional recovery after SCI. In the present study, we show that reactive astrocytes from the injured rat spinal cord or their conditioned media inhibit OL differentiation of adult OPCs with concurrent promotion of astrocyte differentiation. The expression of bone morphogenetic proteins (BMP) is dramatically increased in the reactive astrocytes and their conditioned media. Importantly, blocking BMP activity by BMP receptor antagonist, noggin, reverse the effects of active astrocytes on OPC differentiation by increasing the differentiation of OL from OPCs while decreasing the generation of astrocytes. These data indicate that the upregulated bone morphogenetic proteins in the reactive astrocytes are major factors to inhibit OL differentiation of OPCs and to promote its astrocyte differentiation. These data suggest that manipulation of BMP signaling in the endogenous or grafted NSCs or OPCs may be a useful therapeutic strategy to increase their OL differentiation and remyelination and enhance functional recovery after SCI.

YIU G, HE Z.

Glial inhibition of CNS axon regeneration

[J]. Nat Rev Neurosci, 2006, 7(8): 617-627.

DOI:10.1038/nrn1956      PMID:16858390      [本文引用: 1]

Damage to the adult CNS often leads to persistent deficits due to the inability of mature axons to regenerate after injury. Mounting evidence suggests that the glial environment of the adult CNS, which includes inhibitory molecules in CNS myelin as well as proteoglycans associated with astroglial scarring, might present a major hurdle for successful axon regeneration. Here, we evaluate the molecular basis of these inhibitory influences and their contributions to the limitation of long-distance axon repair and other types of structural plasticity. Greater insight into glial inhibition is crucial for developing therapies to promote functional recovery after neural injury.

YANG T, DAI Y, CHEN G, et al.

Dissecting the dual role of the glial scar and scar-forming astrocytes in spinal cord injury

[J]. Front Cell Neurosci, 2020, 14: 78.

DOI:10.3389/fncel.2020.00078      PMID:32317938      [本文引用: 1]

Recovery from spinal cord injury (SCI) remains an unsolved problem. As a major component of the SCI lesion, the glial scar is primarily composed of scar-forming astrocytes and plays a crucial role in spinal cord regeneration. In recent years, it has become increasingly accepted that the glial scar plays a dual role in SCI recovery. However, the underlying mechanisms of this dual role are complex and need further clarification. This dual role also makes it difficult to manipulate the glial scar for therapeutic purposes. Here, we briefly discuss glial scar formation and some representative components associated with scar-forming astrocytes. Then, we analyze the dual role of the glial scar in a dynamic perspective with special attention to scar-forming astrocytes to explore the underlying mechanisms of this dual role. Finally, taking the dual role of the glial scar into account, we provide several pieces of advice on novel therapeutic strategies targeting the glial scar and scar-forming astrocytes.Copyright © 2020 Yang, Dai, Chen and Cui.

OKADA S, NAKAMURA M, KATOH H, et al.

Conditional ablation of Stat3 or Socs3 discloses a dual role for reactive astrocytes after spinal cord injury

[J]. Nat Med, 2006, 12(7): 829-834.

DOI:10.1038/nm1425      PMID:16783372      [本文引用: 1]

In the injured central nervous system (CNS), reactive astrocytes form a glial scar and are considered to be detrimental for axonal regeneration, but their function remains elusive. Here we show that reactive astrocytes have a crucial role in wound healing and functional recovery by using mice with a selective deletion of the protein signal transducer and activator of transcription 3 (Stat3) or the protein suppressor of cytokine signaling 3 (Socs3) under the control of the Nes promoter-enhancer (Nes-Stat3(-/-), Nes-Socs3(-/-)). Reactive astrocytes in Nes-Stat3(-/-) mice showed limited migration and resulted in markedly widespread infiltration of inflammatory cells, neural disruption and demyelination with severe motor deficits after contusive spinal cord injury (SCI). On the contrary, we observed rapid migration of reactive astrocytes to seclude inflammatory cells, enhanced contraction of lesion area and notable improvement in functional recovery in Nes-Socs3(-/-) mice. These results suggest that Stat3 is a key regulator of reactive astrocytes in the healing process after SCI, providing a potential target for intervention in the treatment of CNS injury.

SOFRONIEW M V.

Astrocyte barriers to neurotoxic inflammation

[J]. Nat Rev Neurosci, 2015, 16(5): 249-263.

DOI:10.1038/nrn3898      PMID:25891508      [本文引用: 1]

Astrocytes form borders (glia limitans) that separate neural from non-neural tissue along perivascular spaces, meninges and tissue lesions in the CNS. Transgenic loss-of-function studies reveal that astrocyte borders and scars serve as functional barriers that restrict the entry of inflammatory cells into CNS parenchyma in health and disease. Astrocytes also have powerful pro-inflammatory potential. Thus, astrocytes are emerging as pivotal regulators of CNS inflammatory responses. This Review discusses evidence that astrocytes have crucial roles in attracting and restricting CNS inflammation, with important implications for diverse CNS disorders.

LIU D, XU G Y, PAN E, et al.

Neurotoxicity of glutamate at the concentration released upon spinal cord injury

[J]. Neuroscience, 1999, 93(4): 1383-1389.

DOI:10.1016/s0306-4522(99)00278-x      PMID:10501463      [本文引用: 1]

Damage caused by administering glutamate into the spinal cord was characterized histologically. Glutamate destroyed neurons for several hundred micrometers around the administering microdialysis fiber. At 24 h after treatment, significant (P = 0.036) loss of neurons was observed (75%) relative to control (47%) near the fiber when glutamate was administered for 1 h at a concentration outside the fiber approximating the maximum glutamate released upon spinal cord injury. Significant loss of neurons (P = 0.006, 0.022) was also caused by administering a combination of glutamate at about its average concentration released upon injury over the 1 h period of administration in combination with the maximum aspartate concentration released upon injury. This work provides a direct demonstration that the concentrations of excitatory amino acids released upon spinal cord injury are neurotoxic. The destruction of neurons by exposure to excitatory amino acids when there is also substantial loss of neurons simply from the presence of the microdialysis fiber may reflect sensitization of neurons to excitotoxicity by stress.

MARAGAKIS N J, DYKES-HOBERG M, ROTHSTEIN J D.

Altered expression of the glutamate transporter EAAT2b in neurological disease

[J]. Ann Neurol, 2004, 55(4): 469-477.

PMID:15048885      [本文引用: 1]

Functional studies suggest that up to 95% of all glutamate transport is handled by the glutamate transporter EAAT2. Amino and C-terminal antibodies demonstrate that under normal conditions EAAT2 is specific to astrocytes. A truncated splice variant of EAAT2, known as EAAT2b, also has been identified in astrocytes and some neurons. In vitro studies suggest EAAT2b transports glutamate similar to EAAT2, although the contribution of EAAT2b to normal clearance of extracellular glutamate is unknown. To investigate EAAT2b biology in pathological conditions, we examined the cellular and regional distribution of EAAT2b in amyotrophic lateral sclerosis. Using epitope-specific, affinity purified antibodies, we found that EAAT2b tissue levels were increased by more than twofold in amyotrophic lateral sclerosis motor cortex, whereas EAAT2 levels were decreased by up to 95%. EAAT2b distribution in normal human cortex was largely confined to the neuropil-like EAAT2, with occasional faint neuronal expression. In contrast, amyotrophic lateral sclerosis motor cortex had an obvious qualitative increase in neuropil EAAT2b staining and a drastic increase in neuronal soma and dendritic EAAT2b immunostaining. Despite these increases in EAAT2b immunostaining, functional transporter studies demonstrated a large loss of EAAT2 function. These studies clearly document altered regulation and splicing of the dominant glutamate transporter EAAT2 under conditions of neurological stress.

CASELLA G T, MARCILLO A, BUNGE M B, et al.

New vascular tissue rapidly replaces neural parenchyma and vessels destroyed by a contusion injury to the rat spinal cord

[J]. Exp Neurol, 2002, 173(1): 63-76.

PMID:11771939      [本文引用: 1]

Blood vessels identified by laminin staining were studied in uninjured spinal cord and at 2, 4, 7, and 14 days following a moderate contusion (weight drop) injury. At 2 days after injury most blood vessels had been destroyed in the lesion epicenter; neurons and astrocytes were also absent, and few ED1+ cells were seen infiltrating the lesion center. By 4 days, laminin associated with vessel staining was increased and ED1+ cells appeared to be more numerous in the lesion. By 7 days after injury, the new vessels formed a continuous cordon oriented longitudinally through the lesion center. ED1+ cells were abundant at this time point and were found in the same area as the newly formed vessels. Astrocyte migration from the margins of the lesion into the new cordon was apparent. By 14 days, a decrease in the number of vessels in the lesion center was observed; in contrast, astrocytes were more prominent in those areas. In addition to providing a blood supply to the lesion site, protecting the demise of the newly formed vascular bridge might provide an early scaffold to hasten axonal regeneration across the injury site.(c) 2002 Elsevier Science.

YAMANE K, MISAWA H, TAKIGAWA T, et al.

Multipotent neurotrophic effects of hepatocyte growth factor in spinal cord injury

[J]. Int J Mol Sci, 2019, 20(23) : 6078.

DOI:10.3390/ijms20236078      URL     [本文引用: 1]

Spinal cord injury (SCI) results in neural tissue loss and so far untreatable functional impairment. In addition, at the initial injury site, inflammation induces secondary damage, and glial scar formation occurs to limit inflammation-mediated tissue damage. Consequently, it obstructs neural regeneration. Many studies have been conducted in the field of SCI; however, no satisfactory treatment has been established to date. Hepatocyte growth factor (HGF) is one of the neurotrophic growth factors and has been listed as a candidate medicine for SCI treatment. The highlighted effects of HGF on neural regeneration are associated with its anti-inflammatory and anti-fibrotic activities. Moreover, HGF exerts positive effects on transplanted stem cell differentiation into neurons. This paper reviews the mechanisms underlying the therapeutic effects of HGF in SCI recovery, and introduces recent advances in the clinical applications of HGF therapy.

MENEGHINI V, PEVIANI M, LUCIANI M, et al.

Delivery platforms for CRISPR/Cas9 genome editing of glial cells in the central nervous system

[J]. Front Genome Ed, 2021, 3: 644319.

DOI:10.3389/fgeed.2021.644319      URL     [本文引用: 1]

LIBERTO C M, ALBRECHT P J, HERX L M, et al.

Pro-regenerative properties of cytokine-activated astrocytes

[J]. J Neurochem, 2004, 89(5): 1092-1100.

PMID:15147501      [本文引用: 1]

The prevailing view of the astrocytic response to injury is that reactive astrocytes impede the regenerative process by forming scar tissue. As the levels of many cytokines dramatically increase following CNS insult and as this increase in cytokine expression precedes the production of the glial scar, a long-standing view has been that cytokines diminish neuronal survival and regeneration by stimulating the formation of astrogliotic scar tissue. However, there is a wealth of data indicating that cytokines "activate" astrocytes, and that cytokine-stimulated astrocytes can promote the recovery of CNS function. Supporting evidence demonstrates that cytokine-activated astrocytes produce energy substrates and trophic factors for neurons and oligodendrocytes, act as free radical and excess glutamate scavengers, actively restore the blood-brain barrier, promote neovascularization, restore CNS ionic homeostasis, promote remyelination and also stimulate neurogenesis from neural stem cells. Accordingly, a re-assessment of cytokine-activated astrocytes is necessary. Here, we review studies that promote the thesis that cytokines elicit potent neuroprotective and regenerative responses from astrocytes.

COLLOMBET J M, FOUR E, FAUQUETTE W, et al.

Soman poisoning induces delayed astrogliotic scar and angiogenesis in damaged mouse brain areas

[J]. Neurotoxicology, 2007, 28(1): 38-48.

DOI:10.1016/j.neuro.2006.07.011      URL     [本文引用: 1]

BARTANUSZ V, JEZOVA D, ALAJAJIAN B, et al.

The blood-spinal cord barrier: morphology and clinical implications

[J]. Ann Neurol, 2011, 70(2): 194-206.

DOI:10.1002/ana.22421      PMID:21674586      [本文引用: 1]

The blood-spinal cord barrier (BSCB) is the functional equivalent of the blood-brain barrier (BBB) in the sense of providing a specialized microenvironment for the cellular constituents of the spinal cord. Even if intuitively the BSCB could be considered as the morphological extension of the BBB into the spinal cord, evidence suggests that this is not so. The BSCB shares the same principal building blocks with the BBB; nevertheless, it seems that morphological and functional differences may exist between them. Dysfunction of the BSCB plays a fundamental role in the etiology or progression of several pathological conditions of the spinal cord, such as spinal cord injury, amyotrophic lateral sclerosis, and radiation-induced myelopathy. This review summarizes current knowledge of the morphology of the BSCB, the methodology of studying the BSCB, and the potential role of BSCB dysfunction in selected disorders of the spinal cord, and finally summarizes therapeutic approaches to the BSCB.Copyright © 2011 American Neurological Association.

WHETSTONE W D, HSU J Y, EISENBERG M, et al.

Blood-spinal cord barrier after spinal cord injury: relation to revascularization and wound healing

[J]. J Neurosci Res, 2003, 74(2): 227-239.

PMID:14515352      [本文引用: 1]

Spinal cord injury produces prominent disruption of the blood-spinal cord barrier. We have defined the blood-spinal cord barrier breakdown to the protein luciferase (61 kDa) in the acutely injured murine spinal cord and during revascularization. We show that newly formed and regenerating blood vessels that have abnormal permeability exhibit differential expression of the glucose-1 transporter (Glut-1), and that its expression is dependent on astrocytes. There was overt extravasation of luciferase within the first hour after injury, a period that coincided with marked tissue disruption within the epicenter of the lesion. Although there was a significant reduction in the number of blood vessels relative to controls by 24 hr after injury, abnormal barrier permeability remained significantly elevated. A second peak of abnormal barrier permeability at 3-7 days postinjury coincided with prominent revascularization of the epicenter. The barrier to luciferase was restored by 21 days postinjury and vascularity was similar to that of controls. During wound-healing process, the cord was reorganized into distinct domains. Between 14 and 21 days postinjury, each domain consisted primarily of nonneuronal cells, including macrophages. Astrocytes were limited characteristically to the perimeter of each domain. Only blood vessels affiliated closely with astrocytes in the perimeter expressed Glut-1, whereas blood vessels within each domain of the repairing cord did not express it. Together, these data demonstrate that both injured and regenerating vessels exhibit abnormal permeability and suggest that Glut-1 expression during revascularization is dependent on the presence of astrocytes.Copyright 2003 Wiley-Liss, Inc.

WAHIS J, HENNES M, ARCKENS L, et al.

Star power: the emerging role of astrocytes as neuronal partners during cortical plasticity

[J]. Curr Opin Neurobiol, 2021, 67: 174-182.

DOI:10.1016/j.conb.2020.12.001      PMID:33360483      [本文引用: 1]

Plasticity is a fundamental property of neuronal circuits, allowing them to adapt to alterations in activation. Generally speaking, plasticity has been viewed from a 'neuron-centric' perspective, with changes in circuit function attributed to alterations in neuronal excitability, synaptic strength or neuronal connectivity. However, it is now clear that glial cells, in particular astrocytes, are key regulators of neuronal plasticity. This article reviews recent progress made in understanding astrocyte function and attempts to summarize these functions into a coherent framework that positions astrocytes as central players in the plasticity process.Copyright © 2020 The Authors. Published by Elsevier Ltd.. All rights reserved.

CHUNG W S, ALLEN N J, EROGLU C.

Astrocytes control synapse formation, function, and elimination

[J]. Cold Spring Harb Perspect Biol, 2015, 7(9): a020370.

DOI:10.1101/cshperspect.a020370      URL     [本文引用: 1]

WANG Y, FU A K Y, IP N Y.

Instructive roles of astrocytes in hippocampal synaptic plasticity: neuronal activity-dependent regulatory mechanisms

[J]. Febs J, 2022, 289(8): 2202-2218.

DOI:10.1111/febs.v289.8      URL     [本文引用: 1]

/