Advances in Clinical Medicine
Vol. 11  No. 04 ( 2021 ), Article ID: 42052 , 13 pages
10.12677/ACM.2021.114290

SMARCAL1在维持复制叉、端粒稳定性中的作用及Schimke免疫骨性发育不良的研究进展

王媛媛1*,石淑娟2,刘涛3,陈蓉1*

1山东潍坊市妇幼保健院,儿科,山东 潍坊

2山东青岛妇女儿童医院,儿科,山东 潍坊

3山东潍坊市市直机关医院,检验科,山东 潍坊

收稿日期:2021年3月27日;录用日期:2021年4月22日;发布日期:2021年4月29日

摘要

SMARCAL1也称为HARP,是一种ATP依赖退火解旋酶,可在DNA损伤期间稳定复制叉。该基因中的突变是造成Schimke免疫性骨发育不良(SIOD)的原因,SIOD是一种常染色体隐性遗传疾病,以生长功能障碍、肾脏损害、T细胞免疫缺陷为表征。我们总结了SMARCAL1在应对DNA复制应激过程中对DNA修复,维持端粒和复制叉稳定上的主要作用、SMARCAL1基因突变导致的疾病表型、癌症预测、SIOD的诊疗进展等方面进行总结。

关键词

DNA复制,稳定性,SMARCAL1,端粒,SIOD

The Role of SMARCAL1 in Maintaining Replication Fork and Telomere Stability and the Research Progress of SIOD

Yuanyuan Wang1*, Shujuan Shi2, Tao Liu3, Rong Chen1*

1Department of Pediatric, Shandong Weifang Maternal and Child Health Hospital, Shandong Weifang

2Department of Pediatric, Shandong Qingdao Women's and Children's Hospital, Shandong Weifang

3Clinical laboratory, Shandong Weifang Municipal Government Hospital, Shandong Weifang

Received: Mar. 27th, 2021; accepted: Apr. 22nd, 2021; published: Apr. 29th, 2021

ABSTRACT

SMARCAL1, called HARP, is an ATP-dependent annealing helicase that stably replicates forks during DNA damage. Mutations in this gene are responsible for immune bone dysplasia (SIOD), SIOD is an autosomal recessive inherited disease characterized by growth dysfunction, kidney damage, and T cell immunodeficiency. We summarized the main role of SMARCAL1 in DNA repair and telomere and replication fork stability in response to DNA replication stress, disease phenotypes caused by SMARCAL1 gene mutations, cancer prediction, and diagnosis and treatment progress of SIOD.

Keywords:DNA Replication, Stability, SMARCAL1, Telomere, SIOD

Copyright © 2021 by author(s) and Hans Publishers Inc.

This work is licensed under the Creative Commons Attribution International License (CC BY 4.0).

http://creativecommons.org/licenses/by/4.0/

1. 引言

SNF2 (非发酵蔗糖2)是一种存在于从酵母到人类的ATP依赖染色质重塑酶家族。这个家族的成员参与基因转录、DNA重组、细胞周期调控、DNA甲基化和DNA损伤修复等各种过程。SNF2家族蛋白包含一种由7个保守序列构成的类解旋酶ATP酶结构域,和在许多DNA和RNA解旋酶中发现的序列有相似性 [1]。1986年首次从小牛胸腺中分离出SMARCAL1 [2]。从线虫到人类,不同物种中都有SMARCAL1的存在。在人和小鼠的所有组织中均见Smarcal1基因表达,如在人免疫系统中,Smarcal1在单核细胞中表达量为1.25‰、B淋巴细胞表达量为1.58‰、CD4 + T细胞表达量为1.25‰、CD8 + T细胞表达量为1.99‰、NK细胞中的表达量为0.1‰;在内分泌系统的胰腺细胞中表达量为0.1‰、前列腺细胞中表达量约为0.02‰;在睾丸细胞中表达量则为0.16‰ [3] - [10]。Smarcal1基因突变与Schimke免疫骨性发育不良(Schimke immuno-osseous dysplasia, SIOD)这个综合征密切相关 [11] [12]。所以说SIOD是一种罕见的累及多系统、进行性加重的常染色体遗传病,主要表现为骨骼发育不良造成的生长迟缓、局灶节段性肾小球硬化(FSGS)最终发展为肾衰竭、T细胞免疫缺陷、脑发育受损等 [6] [13] [14] [15] [16]。除此之外,部分SIOD患者还表现有角膜混浊、动脉粥样硬化、中风、偏头痛、甲状腺功能减退、骨髓衰竭等症状 [6] [13] [16]。

很多研究结果显示,SMARCAL1在维持基因组稳定性和停滞的DNA复制叉的活化中起到作用 [17] [18] [19]。

本文将从SMARCAL1的结构与功能、SMARCAL1在DNA损伤部位应答中所起的作用,以及在端粒的完整性维护过程中的作用、SMARCAL1活性调节与基因组稳定性的关系、SMARCAL1基因突变引起的疾病及基因–表型相关性、SMARCAL1相关的癌症预测、等方面进行总结。

2. SMARCAL1的结构与功能

人Smarcal1基因位于2q34-q36,含有17个外显子,编码由954个氨基酸残基组成的蛋白质 [20]。SMARCAL1 N末端包括一个RPA (复制蛋白A)相互作用的序列,之后是2个串联的HARP结构域 [13] [21] [22];其中解旋酶结构域位于它的C端,具有ATP酶的活性,并被115个氨基酸的长序列连接成两个RecA类的结构域。与RPA作用的序列位于N-末端;与“退火解旋活性”有关的结构域位于239-307与331-4002区间的“HARP”结构域 [23]。C-末端是解旋酶的结构域,有ATP酶催化活性(由115个氨基酸残基组成的“RecA”结构域)和SWI/SNF“核小体重塑蛋白”结构 [6]。SMARCAL1的ATP依赖DNA退火解旋酶含有2个HARP结构域组成的ATPase [17] [24] [25]。在停滞的复制叉重塑、端粒DNA完整性维护、S期细胞周期关卡通路的激活、利用NHEJ机制修复DNA双链断裂损伤等过程中,当DNA出现损伤时,与单链DNA结合的单链DNA结合蛋白RPA32识别SMARCAL1 N端的RPA作用结构域,同时招募SMARCAL1到dsDNA-ssDNA的单链DNA一侧 [3] [6] [17] [26] [27] [28] [29]。

3. SMARCAL1在DNA损伤应答中的作用

3.1. SMARCAL1通过与RPA的相互作用被招募至DNA损伤部位

RPA是由RPA1、RPA2和RPA3形成的异三聚体复合物,在复制、重组和修复过程中结合ssDNA从而防止DNA二级结构的形成 [30]。当存在或不存在DNA损伤的情况下,RPA都可以覆盖并保护ssDNA。RPA作为一种用于招募DNA修复酶的支架蛋白被募集到DNA损伤部位。SMARCAL1与RPA在DNA损伤部位共同定位,且SMARCAL1与RPA有直接的相互作用 [31] [32] [33] [34]。SMARCAL1的RPA结合序列与其他DNA修复蛋白中的序列非常相似,其中包括复制叉保护复合物成员TIPIN和复制应激反应的重要参与者RAD52 [35] - [40]。SMARCAL1在其N端区域有一个RPA2相互作用的序列。该序列与一个可以结合RPA2的α螺旋相对应 [35]。SMARCAL1和RPA间相互作用的破坏造成无法正常的把SMARCAL1招募到DNA的损伤部位。

3.2. SMARCAL1参与DNA损伤应答

SMARCAL1在ATR、ATM和DNA-PK检测点的激酶磷酸化,在参与DNA损伤应答中被活化。有研究表明了SMARCAL1在DNA双链断裂(DSB)位点的作用。在接受辐照的U2OS细胞中,在DNA损伤部位,SMARCAL1被招募后识别DNA末端,并且与γH2AX和RAD51病灶共同定位 [41]。具体步骤是使用线性化的质粒结合在链亲和素珠包被的珠子上,可以提取出结合在DNA上的蛋白质,然后通过质谱比色法鉴定这些蛋白质。仅在一端生物素化的线性DNA分子结合链亲和素珠后,呈现出一个类似于DNA双链断裂的游离DNA末端。相反,如果DNA的两端均被生物素化,则两端均被珠覆盖。将这种技术与非洲爪蟾卵提取物一起使用,研究发现,只有一个DNA末端被生物素化时,SMARCAL1才被有效招募到DNA上,而当两个末端都被生物素化时,则SMARCAL1则未被被招募 [31] [41]。表明了进行加工了的游离DNA末端,以形成单链DNA (ssDNA),才具备招募SMARCAL1的条件。在直接结合试验中,也与以上的结果一致,SMARCAL1对没有加工的的DNA末端不具有高亲和力 [32]。

3.3. SMARCAL1应对DNA复制应激的应答

有研究表明,敲除SMARCAL1的细胞对诱导复制应激的药物(例如羟基脲(HU)、蚜虫碱或喜树碱)表现出超敏性 [35]。在凝胶迁移试验中,与ssDNA或双链DNA相比,SMARCAL1对分叉DNA结构的亲和力更高。当蛋白与分叉的DNA结合时,激发了SMARCAL1的ATP酶活性。但在使用部分双链DNA作为底物的解旋酶测定中,SMARCAL1无法显示出解链活性。与之相反,在部分解链的质粒DNA用作底物的退火解旋酶测定中,在RPA和ATP存在的情况下,SMARCAL1能够重绕ssDNA链。这些结果表明SMARCAL1是ATP驱动的退火解旋酶,它可以退火互补的RPA结合ssDNA [42]。

3.4. SMARCAL1对新生的DNA链进行退火调节停滞的复制叉的回退

SMARCAL1除了上述两个功能外还可以通过对新生的DNA链进行退火来调节停滞的复制叉的回退 [43]。RecG是一种具有3’至5’极性的DNA解旋酶,可回退停滞的复制叉并完成霍利迪中间体(Holliday intermediates)的分支迁移。RecG与SSB (单链DNA结合蛋白)相互作用,这种相互作用可在DNA上稳定RecG并促进RecG的复制叉的回退 [44]。回退的DNA复制叉是一种类似于霍利迪(Holliday)连接体的四向DNA结构。这些DNA结构可以通过切割产生单端DNA,然后可以用于启动DNA的重组修复机制 [45]。回退的DNA复制叉的切割是通过一种与SLX4突变蛋白相互作用的内切酶MUS81的调解 [46] [47]。MUS81的耗竭对正常细胞没有影响,但却会阻止缺乏SMARCAL1的细胞中γH2AX的聚集 [32]。从而证实,SMARCAL1可保护停滞的复制叉避免核酸酶异常加工,因为核酸酶异常加工可能导致基因组不稳定,所以SMARCAL1在维持基因组稳定性方面有着很重要的作用。

3.5. SMARCAL1在复制叉重启中起特定作用,而不是经典的同源重组过程

质谱实验证实WRN解旋酶是一种SMARCAL1相互作用蛋白 [48]。WRN是一种解旋酶/核酸外切酶,它在DNA修复中有多种功能,可以保持基因组脆弱部位的稳定性 [49]。RPAS作为一种支架蛋白,调解MARCAL1和WRN的相互作用,而SMARCAL1和WRN在复制叉修复中均具有重叠作用,又各自独立,有区别 [48]。研究报告指出,在MMC引起的复制叉停滞后,WRN与RAD52共同定位,而WRN的活性受RAD52调节 [50]。总而言之,这些数据表明SMARCAL1、WRN和RAD52可以保护和修复受损的复制叉,但是这些蛋白质的确切的底物和调解方式尚无实验明确。SMARCAL1还可以在体外催化分支迁移。此外,基因转换报告(HDR-GFP)分析检测到,有效的基因转化并不需要SMARCAL1 [13]。这些数据表明SMARCAL1在复制叉重启中起特定作用,而不是经典的同源重组过程。

4. SMARCAL1在端粒完整性维护过程的作用

端粒在S期的异常结构会干扰DNA在这些部位的复制 [51]。因为有富含G序列的存在,才有可能形成G-四链体结构。这些结构阻碍了复制叉前进,造成复制叉停滞和/或复制叉崩溃 [52]。此外,端粒由双链六聚体重复序列TTAGGG组成,TTAGG可分解成阻碍DNA复制的T环。在某些细胞中,除了常见的TTAGGG重复序列外,端粒还包含其他类型的六聚体重复序列,例如TCAGGG或TTCGGG [53] [54] [55] [56]。在缺乏端粒酶的情况下,酵母和哺乳动物细胞均依靠一种同源重组通路(BIR)调解端粒的维持,促进端粒表型的的延伸(ALT)。聚合酶δ是来调节ALT和BIR参与保守的DNA复制合成 [57] [58] [59]。延伸的端粒细胞有长而异构的端粒。这些细胞的另一个特征是C环的存在,即端粒DNA的额外染色体环,它可以是一部分单链结构,并且可以作为ALT活性的标记物 [43] [60]。

造成端粒激活和维持的确切机制尚不明确。推测可能是因为SMARCAL1在停滞的复制叉处重塑了染色质,延伸的端粒中富含SMARCAL1,所以它对端粒起着重要作用。有研究证实,缺少SMARCAL1会加大端粒DNA损伤以及C环的增加,C环的丰度与SMARCAL1的缺失程度有关 [61]。另一项研究表明,延伸的端粒中富含SMARCAL1,而缺少SMARCAL1会影响端粒长度 [62]。还有研究显示,SMARCAL1在端粒中的作用与其他应答DNA复制应激的解旋酶无关 [23] [61] [63]。

5. SMARCAL1活性调节与基因组稳定性的关系

不同类型的DNA损伤会触发激活不同的DNA修复蛋白。DNA损伤应答的主要调解剂是激酶ATR、ATM和DNA-PK,它们让各种下游DNA修复酶磷酸化。SMARCAL1会导致依赖HU、IR或UV以ATR,ATM和DNA-PK方式诱导的DNA损伤后磷酸化。目前已经确定了包括S173、S652和S919在内的多个磷酸化部位。SMARCAL1被招募到DNA损伤部位,这与这些部位的磷酸化无关。然而,仿磷酸化的S652D突变在体外损伤了其DNA依赖ATP酶和分支迁移活性 [64]。因此,该部位的磷酸化主要是由ATR诱导,它限制了SMARCAL1对停滞复制叉的异常处理,防止复制叉崩溃。相反,在S889磷酸化未应激细胞中也很明显,它激活了DNA依赖ATP酶和SMARCAL1的分支迁移活性。因此,调解SMARCAL1的正常水平对于确保基因组稳定性也很重要,因为SMARCAL1的耗竭或过表达都会加大DNA损伤 [65]。

6. SMARCAL1基因突变引起的疾病及基因-表型相关性的研究

SMARCAL1基因的突变会引起常染色体隐性遗传疾病,即(SIOD) [66],世界范围内SIOD的发病率约为1/300万~1/100万 [13]。SMARCAL1表达于全身多个组织及器官,包括骨骼、肾脏、胸腺、甲状腺、神经及血液等。因此SMARCAL1基因变异可导致全身多脏器系统功能异常,主要表型为骨骼、肾脏和免疫系统受损 [44]。

6.1. 脊柱发育不良

几乎所有的SIOD患者都有身材矮小的表型,主要表现为短躯干矮小,颈短,腰椎前凸、腹部突出 [67] [68]。研究统计SIOD男性患者成人身高约为135-156cm,女患者成人身高约97.5~142.5 cm [67]。

SMARCALl是目前已知的唯一能导致SIOD的基因,然而并不是所以的患儿都可以检测到SMARCALl的突变 [68]。Hunter [68] 等人研究了33名SIOD患者的SMARCAL1基因发现有66%的患者是SMARCAL1基因突变,剩下34%的患者未检测到此突变。分析这33名患者的骨骼X线片基因有问题的患者脊柱发育不良(SEDT)基本上仅局限于骨盆、脊柱、股骨骺 [29]。还发现少部分的成人患者年轻时患有骨质疏松症和髋关节病 [68]。在没有可检测到SMARCAL1突变的患者中,大部分患者SEDT放射学表现与突变患者无明显区别。因此,脊柱发育不良不能用来判断具有SEDT的SIOD患者是否具有SMARCAL1突变。

所有的患者都具有正常的生长激素水平。迄今为止,未见SIOD患者垂体前叶功能缺失的文献报道 [8]。具有SMARCAL1突变的SIOD患者的特征性骨骼特征 [68] 是:可能存在的宽蝶鞍;椎骨的扁平化及髋部的异常逐年恶化;横向移位的股骨头骨骺,小的,发育不全的基底髂骨,向上倾斜的髋臼;没有分割缺陷的扁平椎体;髋关节病和椎骨骨质减少(儿童期,青春期和成年早期)。

6.2. 肾脏疾病

SMARCAL1在肾脏发育期间的所有细胞中均有表达,在成熟人肾中肾小管上皮细胞及集合管细胞中表达 [69]。大量蛋白尿是SIOD患者的最早期表现,SIOD肾病最常见的病理结果为局灶性节段性肾小球硬化(FSGS),大多对治疗的药物包括糖皮质激素、环磷酰胺、他克莫司、环孢菌素A等免疫抑制剂治疗无反应,最终发展为肾衰(ESRD),在未成年时死于肾衰竭 [67] [70]。研究表明SIOD的FSGS与肾小球中NOTCH受体、配体的表达的明显增加密切相关 [7]。增加的NOTCH信号传导是FSGS的已知明确致病机制 [71] [72]。

Morimoto [73] 等人发现SIOD病人肾脏中的Wnt和Notch信号通路的组分和标志物的表达增加,SIOD病人肾小球中未磷酸化的β-连环蛋白和Notch1细胞内结构域的表达水平增加。研究发现增加的Wnt和Notch活性增加是由SMARCAL1缺乏引起的,并且为肾脏病理FSGS的致病原因,导致大多数SIOD患者起初的大量蛋白尿为表型的肾病 [73]。对其他以大量蛋白尿为表型的肾小球病的研究发现,Wnt [74] [75] [76] [77] 和Notch信号传导的增加 [71] [72] [78] [79] 导致足细胞功能障碍。Wnt和Notch信号传导对于肾脏发育至关重要,但是在出生后肾脏的肾小球中却无法检测到 [79] [80] [81]。

6.3. T细胞免疫缺乏

原发性免疫缺陷病是由免疫系统内在缺陷引起的异质性疾病组,常表现为反复病毒、细菌或真菌感染 [82]。在正常的人T细胞发育中,T细胞谱系的祖细胞来自骨髓,胸腺中分化产生不成熟的CD4和CD8细胞。然后离开胸腺并进入外周成熟。在胸腺内T细胞完成克隆多样性,如果SIOD患者的外周的T细胞是单克隆或寡克隆 [83],提示T细胞发育缺陷。T细胞缺乏导致约80% SIOD的患者淋巴细胞的减少,存在的T细胞主要是记忆T细胞,与胸腺对T细胞的产生减少一致 [29]。B细胞计数通常是正常或略升高。T细胞缺乏的发生与SIOD患者T细胞中白细胞介素7受体α的缺乏有关 [29] [84]。IL-7及其受体系统在早期T细胞发育中起着重要作用。在淋巴细胞发育过程中,编码免疫球蛋白和T细胞受体抗原结合域的功能基因需要经过与NHEJ类似的V(D)J重组才能形成 [85] [86]。与NHEJ类似,V(D)J重排也需产生DNA双链断裂,并由NHEJ机制完成断链末端的连接 [85] [86]。而SMARCAL1的突变常影响NHEJ在V(D)J重排重组的连接效率,这可能是SIOD患者常见T细胞免疫缺陷的原因之一 [83] [87]。

T细胞免疫缺陷增加机会性感染的风险,如卡氏肺孢子虫的感染,大部分的SIOD患者反复感染各种病毒(包括水痘-带状疱疹病毒、单纯疱疹病毒、巨细胞病毒)、细菌和真菌 [13] [20]。T细胞缺乏而导致的反复感染是SIOD死亡的主要原因。5.4其他系统异常也有部分SIOD患者会合并中枢神经系统(CNS)的症状,伴有高血压和动脉粥样硬化。会出现短暂性神经系统发作、偏头痛样头痛或短暂性脑缺血发作 [57]。短暂性脑缺血发作的源头可能是加重的动脉粥样硬化并伴有严重的高血压 [1] [13]。大多数患者的智力是正常,少数患者出现精神运动发育迟缓。少数SIOD患者出现角膜混浊、牙齿畸形和毛发稀疏 [88]。Kilic [57] 等人也提出血管炎症和血管反应的发生与SMARCAL1突变导致免疫的紊乱有关。有研究证明SIOD患者中ELN的表达明显降低 [89],ELN基因编码的是弹性蛋白前体,弹性蛋白前体对维持动脉壁的完整性有非常重要的作用。三个SIOD患者的死后动脉组织病理学分析显示弹性蛋白纤维的分裂和碎裂 [8] [89]。弹性蛋白减少导致动脉壁平滑肌细胞的增殖从而导致内膜的增生 [90]。SIOD主动脉中弹性蛋白表达的减少可能与ELN转录减少以及转录后ELN mRNA衰变增加有关 [90]。Haffner [91] 等报道了一名严重SIOD的6岁男孩,表现为波动性偏瘫和癫痫发作、剧烈的偏头痛、暂时性脑缺血发作;发作初始、期间血管造影、磁共振成像均正常;发作后可见灌注和动脉狭窄减少,多个可逆限制扩散区域。这一系列症状和影像学表现提示患儿为可逆性的脑血管收缩综合征。

6.4. SIOD患者基因型与表型的相关性

根据人类基因突变数据库(HGMD)报道,目前已发现60多种SMARCAL1基因突变。突变类型包括错义突变、无义突变、剪切突变、微小插入或缺失、交叉缺失等,其中错义和无义突变最多,达30多种 [23]。早期研究表明SIOD影响的个体中存在基因-表型相关性:严重SIOD患者表现出两个无义、移码或剪接突变,而轻度受影响的个体具有错义突变 [20]。缺失、无义和移码突变通常导致SMARCAL1的mRNA和蛋白质的表达缺失,而错义突变可能改变亚细胞定位、酶活性及蛋白质水平。导致SIOD发生的Smarcal1基因突变多为双等位基因功能缺失、错义突变、插入/缺失(insertion and deletion, Ins/Del)、大片段缺失以及SMAR-CAL1 mRNA拼接错误 [14] [92]。上述基因改变常出现在SMARCAL1的RecA样结构域I中,由于突变影响了SMARCAL1的ATP酶活性,故常见疾病的严重程度与突变体SMARCAL1所表现出的ATP酶活性成反比 [93]。SMARCAL1的突变与SIOD患者的染色体不稳定相关,这表明该表型可能是该疾病的重要特征 [94]。在另一篇最新论文中,染色质变化也与SIOD相关。SMARCAL1还可能结合染色质而直接影响基因表达。果蝇的SMARCAL1直系同源物Marcal1与trxG和PcG相互作用,而Marcal1缺陷影响着染色质的结构和基因表达 [95]。SMARCAL1等位基因缺失、无义或移码突变常见于重症患者。重症SIOD患者的症状在孕期外显,表现为胎儿生长迟缓、甲状腺功能减退症、骨髓衰竭、短暂性脑缺血发作、中风和肾衰竭等,一般死于5岁前。SMARCAL1等位基因错义突变患者症状较轻,多数发病较晚,常见8~13岁以后发病,数年后进展至肾衰 [23] [96]。Lipska [95] 等人通过分析来自28个家族的34名SIOD患者的肾脏相关基因型-表型,未发现肾病病程的基因型-表型相关性,但发现肾组织对基因组不稳定性的累积效应具有高度敏感性。研究表明SIOD疾病的严重程度与SMARCAL1的活性成反比 [97]。然而,除了基因变异可以影响基因表达及临床表现外,许多其他遗传因素或环境因素也可以影响基因表达,改变临床表型 [89]。

7. SMARCAL1基因突变与癌症易感性

虽然报道过几例SIOD患者伴随非霍奇金淋巴瘤 [98],但SIOD患者的癌症易感性并不高,这可能是因为SIOD患者的免疫缺陷与T细胞功能缺陷有关 [99]。在小鼠实验中,SMARCAL1的失活导致对DNA损伤因子如伊立替康(CPT-11)的超敏反应[107]。在另一项研究中,去除SMARCAL1基因的小鼠无法表达RPA结合结构域(第一个HARP域和第二个HARP域的一部分 [100] ),从而对癌症的发展形成阻力。因此,当接触低剂量的IR时,野生型小鼠因为T细胞祖细胞中DNA突变的积累而发展为T细胞淋巴瘤,而缺少SMARCAL1 N末端结构域的小鼠则表现出肿瘤形成的延迟,这可能是因为缺少SMARCAL1,增加了T细胞祖细胞对DNA损伤的易感性并诱导祖细胞凋亡 [100]。

8. SIOD的诊疗进展

SIOD的诊断是在依靠先证者的临床特征和放射学表现中建立起来的 [7]。如果通过基因明确SMARCAL1的双等位基因存在致病变异,即使临床特征轻微或不典型,也可诊断SIOD [4]。对于具有以下特征的个体,应怀疑SIOD:1) 身材矮小(99%),表现为短躯干、短颈、腰椎前凸;2) 脊柱发育不良(75%),同前叙述;3) 肾病,几乎所有的SIOD患者都会出现蛋白尿且逐渐演变成终末肾衰,83%的患者肾脏病理为FSGS;4) T细胞缺乏(76%),CD4和CD8细胞都减少,比例正常 [7]。

Saraiva等 [23] 依据疾病出现的时间和疾病的严重程度将SIOD分为早发严重型和迟发缓和型两种类型。若SIOD患者在胎儿期就出现宫内发育迟缓,生后生长迟缓则归为早发严重型,其余为迟发缓和型。两种类型的临床表现没有明显差异,但早发严重型的肾病最终全部发展为终末期肾病。迟发缓和型的肾病经过合理的替代治疗可存活到成人期 [23]。Boerkoel等 [5] 发现SIOD患者中孕33周前的早产儿的存活年龄明显下降;约2/3的患儿在2岁前即出现生长发育延迟;一半以上患者在15岁前死亡。

SIOD治疗目前无特异治疗,主要是针对合并症及并发症的对症处理及预防,对于慢性肾衰竭可行腹膜透析或者血液透析或肾移植,对于免疫缺陷可行骨髓移植,或联合移植来提高患者存活率 [5]。T细胞免疫缺陷,若发生反复口腔疱疹感染或带状疱疹,可预防性使用抗病毒药物;进行疫苗接种以预防肺孢子菌肺炎。对自身免疫紊乱者进行免疫调节治疗。粒细胞集落刺激因子治疗中性粒细胞减少症。改善血流量或降低凝血能力药物以治疗短暂性脑缺血发作或中风。定期监测骨骼疾病的发展,每年检测肾脏功能、免疫和血液学状况。SIOD预后差,多数患者15岁前死亡,部分死于肾衰竭、各种感染及脑栓塞,只有少数患者可存活至成年 [100]。

基金项目

山东省医药卫生科技发展计划项目(2019WSA07025),项目负责人:王媛媛。

潍坊市卫生健康委员会科研项目(WFWSJK-2020-292),项目负责人:王媛媛。

文章引用

王媛媛,石淑娟,刘 涛,陈 蓉. SMARCAL1在维持复制叉、端粒稳定性中的作用及Schimke免疫骨性发育不良的研究进展
The Role of SMARCAL1 in Maintaining Replication Fork and Telomere Stability and the Research Progress of SIOD[J]. 临床医学进展, 2021, 11(04): 1383-1395. https://doi.org/10.12677/ACM.2021.114290

参考文献

  1. 1. Eisen, J.A., Sweder, K.S. and Hanawalt, P.C. (1995) Evolution of the SNF2 Family of Proteins: Subfamilies with Distinct Sequences and Functions. Nucleic Acids Research, 23, 2715-2723. https://doi.org/10.1093/nar/23.14.2715

  2. 2. Hockensmith, J.W., Wahl, A.F., Kowalski, S. and Bambara, R.A. (1986) Purification of a Calf Thymus DNA-Dependent Adenosinetriphosphatase That Prefers a Primer-Template Junction Effector. Biochemistry, 25, 7812-7821. https://doi.org/10.1021/bi00372a005

  3. 3. Coleman, M.A., Eisen, J.A. and Mohrenweiser, H.W. (2000) Cloning and Characterization of HARP/SMARCAL1: A Prokaryotic HepA-Related SNF2 Helicase Protein from Human and Mouse. Genomics, 65, 274-282. https://doi.org/10.1006/geno.2000.6174

  4. 4. Kakar, S., Fang, X., Lubkowska, L., Zhou, Y.N., Shaw, G.X., Wang, Y.X., Kashlev, M. and Ji, X. (2015) Allosteric Activation of Bacterial Swi2/Snf2 (Switch/Sucrose Non-Fermentable) Protein RapA by RNA Polymerase: Biochemical and Structural Studies. Journal of Biological Chemistry, 290, 23656-23669. https://doi.org/10.1074/jbc.M114.618801

  5. 5. Hargreaves, D.C. and Crabtree, G.R. (2011) ATP-Dependent Chromatin Remodeling: Genetics, Genomics and Mechanisms. Cell Research, 21, 396-420. https://doi.org/10.1038/cr.2011.32

  6. 6. Morimoto, M. (2016) Characterization of the Disease Pathogenesis of Schimke Immuno-Osseous Dysplasia. University of British Columbia, Vancouver, 27-46.

  7. 7. Zivicnjak, M., Franke, D., Zenker, M., Hoyer, J., Lücke, T., Pape, L. and Ehrich, J.H. (2009) SMARCAL1 Mutations: A Cause of Prepubertal Idiopathic Steroid-Resistant Nephritic Syndrome. Pediatric Research, 65, 564-568. https://doi.org/10.1203/PDR.0b013e3181998a74

  8. 8. Dekel, B., Metsuyanim, S., Goldstein, N., Pode-Shakked, N., Kovalski, Y., Cohen, Y., Davidovits, M. and Anikster, Y. (2008) Schimke Immuno-Osseous Dysplasia: Expression of SMARCAL1 in Blood and Kidney Provides Novel Insight into Disease Phenotype. Pediatric Research, 63, 398-403. https://doi.org/10.1203/PDR.0b013e31816721cc

  9. 9. Haokip, T.D., Kumari, R. and Muthuswami, R. (2011) Human SMARCAL1—A Member of the SWI2/SNF2 Family-Is Required for Cell Division. The FASEB Journal, 25, 3515-3524.

  10. 10. Boerkoel, C.F., O’Neill, S., Andre, J.L., Benke, P.J., Bogdanovic, R., Bulla, M., Burguet, A., Cockfield, S., Cordeiro, I., Ehrich, J.H., et al. (2000) Manifestations and Treatment of Schimke Immuno-Osseous Dysplasia: 14 New Cases and a Review of the Literature. European Journal of Pediatrics, 159, 1-7. https://doi.org/10.1007/s004310050001

  11. 11. Spranger, J., Hinkel, G.K., Stoss, H., Thoenes, W., Wargowski, D. and Zepp, F. (1991) Schimke Immuno-Osseous Dysplasia: A Newly Recognized Multisystem Disease. The Journal of Pediatrics, 119, 64-72. https://doi.org/10.1016/S0022-3476(05)81040-6

  12. 12. Baradaran-Heravi, A., Lange, J., Asakura, Y., Cochat, P., Massella, L. and Boerkoel, C.F. (2013) Bone Marrow Transplantation in Schimke Immuno-Osseous Dysplasia. American Journal of Medical Genetics Part A, 161, 2609-2613. https://doi.org/10.1002/ajmg.a.36111

  13. 13. Severino, M., Giacomini, T., Verrina, E., Prato, G. and Rossi, A. (2018) Reversible Cerebral Vasoconstriction Complicating Cerebral Atherosclerotic Vascular Disease in Schimke Immune-Osseous Dysplasia. Neuroradiology, 60, 885-888. https://doi.org/10.1007/s00234-018-2052-y

  14. 14. Zhang, L., Fan, S., Liu, H. and Huang, C. (2012) Targeting SMARCAL1 as a Novel Strategy for Cancer Therapy. Biochemical and Biophysical Research Communications, 427, 232-235. https://doi.org/10.1016/j.bbrc.2012.09.060

  15. 15. Carroll, C., Hunley, T.E., Guo, Y. and Cortez, D. (2015) A Novel Splice Site Mutation in SMARCAL1 Results in Aberrant Exon Definition in a Child with Schimke Immunoosseous Dysplasia. American Journal of Medical Genetics Part A, 167, 2260-2264. https://doi.org/10.1002/ajmg.a.37146

  16. 16. Prato, G., Grandis, E.D., Mancardi, M.M., Croci, C., Pisciotta, D., Uccella, S., Costanzo, C., Severino, S., Tortora, D., Pavanello, M. and Venesellividovits, E. (2018) Schimke Immuno-Osseous Dysplasia: A Peculiar EEG Pattern. Neuropediatrics, 49, S1-S12. https://doi.org/10.1055/s-0038-1653933

  17. 17. Bansbach, C.E., Betous, R., Lovejoy, C.A., Glick, G.G. and Cortez, D. (2009) The Annealing He-licase SMARCAL1 Maintains Genome Integrity at Stalled Replication Forks. Genes & Development, 23, 2405-2414. https://doi.org/10.1101/gad.1839909

  18. 18. Ciccia, A., Bredemeyer, A.L., Sowa, M.E., Terret, M.E., Jallepalli, P.V., Harper, J.W. and Elledge, S.J. (2009) The SIOD Disorder Protein SMARCAL1 Is an RPA-Interacting Protein Involved in Replication Fork Restart. Genes & Development, 23, 2415-2425. https://doi.org/10.1101/gad.1832309

  19. 19. Baradaranheravi, A., Cho, K.S., Tolhuis, B., Sanyal, M., Morozova, O., Morimoto, M., et al. (2012) Penetrance of Biallelic SMARCAL1 Mutations Is Associated with Environmental and Genetic Disturbances of Gene Expression. Human Molecular Genetics, 21, 2572-2587. https://doi.org/10.1093/hmg/dds083

  20. 20. Bokenkamp, A., dejong, M., van Wijk, J.A., Block, D., van Hagen, J.M. and Ludwig, M. (2005) R561C Missense Mutation in the SMARCAL1 Gene Associated with Mild Schimke Immuno-Osseous Dysplasia. Pediatric Nephrology, 20, 1724-1728. https://doi.org/10.1007/s00467-005-2047-x

  21. 21. Flaus, A., Martin, D.M., Barton, G.J. and Owen-Hughes, T. (2006) Identification of Multiple Distinct Snf2 Subfamilies with Conserved Structural Motifs. Nucleic Acids Research, 34, 2887-2905. https://doi.org/10.1093/nar/gkl295

  22. 22. Ghosal, G., Yuan, J. and Chen, J. (2011) The HARP Domain Dictates the Annealing Helicase Activity of HARP/SMARCAL1. EMBO Reports, 12, 574-580. https://doi.org/10.1038/embor.2011.74

  23. 23. Saraiva, J.M., Dinis, A., Resende, C., Faria, E., Gomes, C., Jorge Correia, A., et al. (1999) Schimke Immuno-Osseous dysplasia: Case Report and Review of 25 Patients. Journal of Medical Genetics, 36, 786-789. https://doi.org/10.1136/jmg.36.10.786

  24. 24. Hauk, G. and Bowman, G.D. (2011) Structural Insights into Regulation and Action of SWI2/SNF2 ATPases. Current Opinion in Structural Biology, 21, 719-727. https://doi.org/10.1016/j.sbi.2011.09.003

  25. 25. Osman, F. and Whitby, M.C. (2007) Exploring the Roles of Mus81-Eme1/Mms4 at Perturbed Replication Forks. DNA Repair, 6, 1004-1017. https://doi.org/10.1016/j.dnarep.2007.02.019

  26. 26. Han, J.J., Song, Z.T., Sun, J.L., Yang, Z.T., Xian, M.J., Wang, S., Sun, L. and Liu, J.X. (2018) Chromatin Remodeling Factor CHR18 Interacts with Replication Protein RPA1A to Regulate the DNA Replication Stress Response in Arabidopsis. New Phytologist, 220, 476-487. https://doi.org/10.1111/nph.15311

  27. 27. Postow, L., Woo, E.M., Chait, B.T. and Funabiki, H. (2009) Identification of SMARCAL1 as a Component of the DNA Damage Response. Journal of Biological Chemistry, 284, 35951-35961. https://doi.org/10.1074/jbc.M109.048330

  28. 28. 芦广庆, 段金志, 张昱. 哺乳动物DNA连接酶在DNA双链断裂修复通路中的作用[J]. 遗传, 2016, 38(2): 178-179.

  29. 29. Sanyal, M., Morimoto, M., Baradaran-Heravi, A., Choi, K., Kambham, N., Jensen, K., et al. (2015) Lack of IL7Rα Expression in T Cells Is a Hallmark of T-Cell Immunodeficiency in Schimke Immune Osseous Dysplasia (SIOD). Clinical Immunology, 161, 355-365. https://doi.org/10.1016/j.clim.2015.10.005

  30. 30. Wold, M.S. (1997) Replication Protein A: A Heterotrimeric Single-Stranded DNA-Binding Protein Required for Eukaryotic DNA Metabolism. Annual Review of Biochemistry, 66, 61-92. https://doi.org/10.1146/annurev.biochem.66.1.61

  31. 31. Postow, L., Ghenoiu, C., Woo, E.M., Krutchinsky, A.N., Chait, B.T. and Funabiki, H. (2008) Ku80 Removal from DNA through Double Strand Break-Induced Ubiquitylation. Journal of Cell Biology, 182, 467-479. https://doi.org/10.1083/jcb.200802146

  32. 32. Bansal, R., Arya, V., Sethy, R., Rakesh, R. and Muthuswami, R. (2018) RecA-Like Domain 2 of DNA-Dependent ATPase A Domain, a SWI2/SNF2 Protein, Mediates Conformational Integrity and ATP Hydrolysis. Bioscience Reports, 38, Article ID: BSR20180568. https://doi.org/10.1042/BSR20180568

  33. 33. Poole, L.A. and Cortez, D. (2017) Functions of SMARCAL1, ZRANB3, and HLTF in Maintaining Genome Stability. Critical Reviews in Biochemistry and Molecular Biology, 52, 696-714. https://doi.org/10.1080/10409238.2017.1380597

  34. 34. 文雅蕾, 吕柯孬, 徐小康, 张欣, 丁良, 潘学峰. 退火解旋酶SMARCAL1在维持基因组稳定中的作用与机制[J]. 遗传, 2019, 41(12): 1084-1098.

  35. 35. Xie, S., Lu, Y., Jakoncic, J., Sun, H. and Xia, J. and Qian, C. (2014) Structure of RPA32 Bound to the N-Terminus of SMARCAL1 Redefines the Binding Interface between RPA32 and Interacting Proteins. The FEBS Journal, 281, 3382-3396. https://doi.org/10.1111/febs.12867

  36. 36. Varley, H., Pickett, H.A., Foxon, J.L., Reddel, R.R. and Royle, N.J. (2002) Molecular Characterization of Inter-Telomere and Intra-Telomere Mutations in Human ALT Cells. Nature Genetics, 30, 301-305. https://doi.org/10.1038/ng834

  37. 37. Murfuni, I., Basile, G., Subramanyam, S., Malacaria, E., Bignami, M., Spies, M., Franchitto, A. and Pichierri, P. (2013) Survival of the Replication Checkpoint Deficient Cells Requires MUS81-RAD52 Function. PLoS Genetics, 9, e1003910. https://doi.org/10.1371/journal.pgen.1003910

  38. 38. Aydin, Ö.Z., Vermeulen, W. and Lans, H. (2014) ISWI Chromatin Remodeling Complexes in the DNA Damage Response. Cell Cycle, 13, 3016-3025. https://doi.org/10.4161/15384101.2014.956551

  39. 39. Sugiyama, T. and Kowalczykowski, S.C. (2002) Rad52 Protein Associates with Replication Protein A (RPA)-Single-Stranded DNA to Accelerate Rad51-Mediated Displacement of RPA and Presynaptic Complex Formation. Journal of Biological Chemistry, 277, 31663-31672. https://doi.org/10.1074/jbc.M203494200

  40. 40. Bhowmick, R., Minocherhomji, S. and Hickson, I.D. (2016) RAD52 Facilitates Mitotic DNA Synthesis Following Replication Stress. Molecular Cell, 64, 1117-1126. https://doi.org/10.1016/j.molcel.2016.10.037

  41. 41. Mason, A.C., Rambo, R.P., Greer, B., Pritchett, M., Tainer, J.A., Cortez, D. and Eichman, B.F. (2014) A Structure-Specific Nucleic Acid-Binding Domain Conserved among DNA Repair Proteins. Proceedings of the National Academy of Sciences of the United States of America, 111, 7618-7623. https://doi.org/10.1073/pnas.1324143111

  42. 42. Yusufzai, T. and Kadonaga, J.T. (2008) HARP Is an ATP-Driven Annealing Helicase. Science, 322, 748-750. https://doi.org/10.1126/science.1161233

  43. 43. Havas, K., Whitehouse, I. and Owen-Hughes, T. (2001) ATP-Dependent Chromatin Remodeling Activities. Cellular and Molecular Life Sciences, 58, 673-682. https://doi.org/10.1007/PL00000891

  44. 44. Thoma, N.H., Czyzewski, B.K., Alexeev, A.A., Mazin, A.V., Kowalczykowski, S.C. and Pavletich, N.P. (2005) Structure of the SWI2/SNF2 Chromatin-Remodeling Domain of Eukaryotic Rad54. Nature Structural & Molecular Biology, 12, 350-356. https://doi.org/10.1007/PL00000891

  45. 45. Wild, P. and Matos, J. (2016) Cell Cycle Control of DNA Joint Molecule Resolution. Current Opinion in Cell Biology, 40, 74-80. https://doi.org/10.1016/j.ceb.2016.02.018

  46. 46. Bochar, D.A., Wang, L., Beniya, H., Kinev, A., Xue, Y., Lane, W.S., et al. (2000) BRCA1 Is Associated with a Human SWI/SNF-Related Complex: Inking Chromatin Remodeling to Breast Cancer. Cell, 102, 257-265. https://doi.org/10.1016/S0092-8674(00)00030-1

  47. 47. Patne, K., Rakesh, R., Arya, V., Chanana, U.B., Sethy, R., Swer, P.B. and Muthuswami, R. (2017) BRG1 and SMARCAL1 Transcriptionally Co-Regulate DROSHA, DGCR8 and DICER in Response to Doxorubicin-Induced DNA Damage. Biochimica et Biophysica Acta (BBA)-Gene Regulatory Mechanisms, 1860, 936-951. https://doi.org/10.1016/j.bbagrm.2017.07.003

  48. 48. Maezawa, Y., Onay, T., Scott, R.P., Keir, L.S., Dimke, H., Li, C., et al. (2014) Loss of the Podocyte-Expressed Transcription Factor Tcf21/Pod1 Results in Podocyte Differentiation Defects and FSGS. Journal of the American Society of Nephrology, 25, 2459-2470. https://doi.org/10.1681/ASN.2013121307

  49. 49. Murfuni, I., De Santis, A., Federico, M., Bignami, M., Pichierri, P. and Franchitto, A. (2012) Perturbed Replication Induced Genome Wide or at Common Fragile Sites Is Differently Managed in the Absence of WRN. Carcinogenesis, 33, 1655-1663. https://doi.org/10.1093/carcin/bgs206

  50. 50. Pugliese, G.M., Salaris, F., Palermo, V., Marabitti, V., Morina, N., Rosa, A., et al. (2019) Inducible SMARCAL1 Knockdown in iPSC Reveals a Link between Replication Stress and Altered Expression of Master Differentiation Genes. Disease Models & Mechanisms, 12, dmm039487. https://doi.org/10.1242/dmm.039487

  51. 51. Yusufzai, T., Kong, X., Yokomori, K. and Kadonaga, J.T. (2009) The Annealing Helicase HARP Is Recruited to DNA Repair Sites via an Interaction with RPA. Genes & Development, 23, 2400-2404. https://doi.org/10.1101/gad.1831509

  52. 52. Rhodes, D. and Lipps, H.J. (2015) G-Quadruplexes and Their Regulatory Roles in Biology. Nucleic Acids Research, 43, 8627-8637. https://doi.org/10.1093/nar/gkv862

  53. 53. Yuan, J., Ghosal, G. and Chen, J. (2009) The Annealing Helicase HARP Protects Stalled Replication Forks. Genes & Development, 23, 2394-2399. https://doi.org/10.1101/gad.1836409

  54. 54. Carroll, C., Bansbach, C.E., Zhao, R., Jung, S.Y., Qin, J. and Cortez, D. (2014) Phosphorylation of a C-Terminal Auto-Inhibitory Domain Increases SMARCAL1 Activity. Nucleic Acids Research, 42, 918-925. https://doi.org/10.1093/nar/gkt929

  55. 55. Conomos, D., Stutz, M.D., Hills, M., Neumann, A.A., Bryan, T.M., Reddel, R.R. and Pickett, H.A. (2012) Variant Repeats Are Interspersed Throughout the Telomeres and Recruit Nuclear Receptors in ALT Cells. Journal of Cell Biology, 199, 893-906. https://doi.org/10.1083/jcb.201207189

  56. 56. O’Sullivan, R.J. and Almouzni, G. (2014) Assembly of Telomeric Chromatin to Create Alternative Endings. Trends in Cell Biology, 24, 675-685. https://doi.org/10.1016/j.tcb.2014.07.007

  57. 57. Kilic, S.S., Donmez, O., Sloan, E.A., Elizondo, L.I., Huang, C., Andre, J.L., et al. (2005) Association of Migraine-Like Headaches with Schimke Immuno-Osseous Dysplasia. American Journal of Medical Genetics Part A, 135A, 206-210. https://doi.org/10.1002/ajmg.a.30692

  58. 58. Lydeard, J.R., Jain, S., Yamaguchi, M. and Haber, J.E. (2007) Break-Induced Replication and Telomerase-Independent Telomere Maintenance Require Pol32. Nature, 448, 820-823. https://doi.org/10.1038/nature06047

  59. 59. Dilley, R.L., Verma, P., Cho, N.W., Winters, H.D., Wondisford, A.R. and Greenberg, R.A. (2016) Break-Induced Telomere Synthesis Underlies Alternative Telomere Maintenance. Nature, 539, 54-58. https://doi.org/10.1038/nature20099

  60. 60. Lau, L.M., Dagg, R.A., Henson, J.D., Au, A.Y., Royds, J.A. and Reddel, R.R. (2013) Detection of Alternative Lengthening of Telomeres by Telomere Quantitative PCR. Nucleic Acids Research, 41, e34. https://doi.org/10.1093/nar/gks781

  61. 61. Poole, L.A., Zhao, R., Glick, G.G., Lovejoy, C.A., Eischen, C.M. and Cortez, D. (2015) SMARCAL1 Maintains Telomere Integrity during DNA Replication. Proceedings of the National Academy of Sciences of the United States of America, 112, 14864-14869. https://doi.org/10.1073/pnas.1510750112

  62. 62. Cox, K.E., Marechal, A. and Flynn, R.L. (2016) SMARCAL1 Resolves Replication Stress at ALT Telomeres. Cell Reports, 14, 1032-1040. https://doi.org/10.1016/j.celrep.2016.01.011

  63. 63. Betous, R., Glick, G.G., Zhao, R. and Cortez, D. (2013) Identification and Characterization of SMARCAL1 Protein Complexes. PloS ONE, 8, e63149. https://doi.org/10.1371/journal.pone.0063149

  64. 64. Holsclaw, J.K. and Sekelsky, J. (2017) Annealing of Complementary DNA Sequences during Double-Strand Break Repair in Drosophila Is Mediated by the Ortholog of SMARCAL1. Genetics, 206, 467-480. https://doi.org/10.1534/genetics.117.200238

  65. 65. Gilson, E. and Geli, V. (2007) How Telomeres Are Replicated. Nature Reviews Molecular Cell Biology, 8, 825-838. https://doi.org/10.1038/nrm2259

  66. 66. Rivera, T., Haggblom, C., Cosconati, S. and Karlseder, J. (2017) A Balance between Elongation and Trimming Regulates Telomere Stability in Stem Cells. Nature Structural & Molecular Biology, 24, 30-39. https://doi.org/10.1038/nsmb.3335

  67. 67. Chernin, G., Vega-Warner, V., Schoeb, D.S., Heeringa, S.F., Ovunc, B., Saisawat, P., et al. (2010) Genotype/Phenotype Correlation in Nephrotic Syndrome Caused by WT1 Mutations. Clinical Journal of the American Society of Nephrology, 5, 1655-1662. https://doi.org/10.2215/CJN.09351209

  68. 68. Hunter, K.B., Lucke, T., Spranger, J., Smithson, S.F., Alpay, H., André, J.-L., et al. (2010) Schimke Immunoosseous Dysplasia: Defining Skeletal Features. European Journal of Pediatrics, 169, 801-811. https://doi.org/10.1007/s00431-009-1115-9

  69. 69. Sarin, S., Javidan, A., Boivin, F., Alexopoulou, I., Lukic, D., Svajger, B., et al. (2015) Insights into the Renal Pathogenesis in Schimke Immuno-Osseous Dysplasia: A Renal Histological Characterization and Expression Analysis. Journal of Histochemistry & Cytochemistry, 63, 32-44. https://doi.org/10.1369/0022155414558335

  70. 70. Lucke, T., Kanzelmeyer, N., Baradaran-Heravi, A., Boerkoel, C.F., Burg, M., Ehrich, J.H.H., et al. (2009) Improved Outcome with Immunosuppressive Monotherapy after Renal Transplantation in Schimke-Immuno-Osseous Dysplasia. Pediatric Transplantation, 13, 482-489. https://doi.org/10.1111/j.1399-3046.2008.01013.x

  71. 71. Kakuda, S. and Haltiwanger, R.S. (2017) Deciphering the Fringe-mediated Notch code: identification of activating and Inhibiting Sites Allowing Discrimination between Ligands. Developmental Cell, 40, 193-201. https://doi.org/10.1016/j.devcel.2016.12.013

  72. 72. Murea, M., Park, J.K., Sharma, S., Kato, H., Gruenwald, A., Niranjan, T., et al. (2010) Expression of Notch Pathway Proteins Correlates with Albuminuria, Glomerulosclerosis, and Renal Function. Kidney international, 78, 514-522. https://doi.org/10.1038/ki.2010.172

  73. 73. Morimoto, M., Myung, C., Beirnes, K., Choi, K., Asakura, Y., Bokenkamp, A., et al. (2016) Increased Wnt and Notch Signaling: A Clue to the Renal Disease in Schimke Immuno-Osseous Dysplasia? Orphanet Journal of Rare Diseases, 11, Article No. 149. https://doi.org/10.1186/s13023-016-0519-7

  74. 74. Liu, Z., Chen, S., Boyle, S., Zhu, Y., Zhang, A., Piwnica-Worms, D.R., et al. (2013) The Extracellular Domain of Notch2 Increases Its Cell-Surface Abundance and Ligand Responsiveness during Kidney Development. Developmental Cell, 25, 585-598. https://doi.org/10.1016/j.devcel.2013.05.022

  75. 75. Kato, H., Gruenwald, A., Suh, J.H., Miner, J.H., Barisoni-Thomas, L., Taketo, M.M., et al. (2011) Wnt/Beta-Catenin Pathway in Podocytes Integrates Cell Adhesion, Differentiation, and Survival. The Journal of Biological Chemistry, 286, 26003-26015. https://doi.org/10.1074/jbc.M111.223164

  76. 76. Shkreli, M., Sarin, K.Y., Pech, M.F., Papeta, N., Chang, W., Brockman, S.A., et al. (2011) Reversible Cell-Cycle Entry in Adult Kidney Podocytes through Regulated Control of telomerase and Wnt Signaling. Nature Medicine, 18, 111-119. https://doi.org/10.1038/nm.2550

  77. 77. He, W., Tan, R.J., Li, Y., Wang, D., Nie, J., Hou, F.F., et al. (2012) Matrix metalloprOteinase-7 as a Surrogate Marker Predicts Renal Wnt/β-Catenin Activity in CKD. Journal of the American Society of Nephrology, 23, 294-304. https://doi.org/10.1681/ASN.2011050490

  78. 78. Chau, Y.Y., Brownstein, D., Mjoseng, H., Lee, W.-C., Buza-Vidas, N., Nerlov, C., et al. (2011) Acute Multiple Organ Failure in Adult Mice Deleted for the Developmental Regulator Wt1. PLoS Genet, 7, e1002404. https://doi.org/10.1371/journal.pgen.1002404

  79. 79. Lasagni, L., Ballerini, L., Angelotti, M.L., Parente, E., Sagrinati, C., Mazzinghi, B., et al. (2010) Notch Activation Differentially Regulates Renal Progenitors Proliferation and Differentiation toward the Podocyte Lineage in Glomerular Disorders. Stem Cells, 28, 1674-1685. https://doi.org/10.1002/stem.492

  80. 80. Kato, H. and Susztak, K. (2012) Repair Problems in Podocytes: Wnt, Notch, and Glomerulosclerosis. Seminars in Nephrology, 32, 350-356. https://doi.org/10.1016/j.semnephrol.2012.06.006

  81. 81. Clewing, J.M., Fryssira, H., Goodman, D., Smithson, S.F., Sloan, E.A., Lou, S., et al. (2007) Schimke Immunoosseous Dysplasia: Suggestions of Genetic Diversity. Human Mutation, 28, 273-283. https://doi.org/10.1002/humu.20432

  82. 82. Kobrynski, L., Powell, R.W. and Bowen, S. (2014) Prevalence and Morbidity of Primary Immunodeficiency Diseases, United States 2001-2007. Journal of Clinical Immunology, 34, 954-961. https://doi.org/10.1007/s10875-014-0102-8

  83. 83. Lev, A., Amariglio, N., Levy, Y., Spirer, Z., Anikster, Y., Rechavi, G., et al. (2009) Molecular Assessment of Thymic Capacities in Patients with Schimke Immuno-Osseous Dysplasia. Clinical Immunology, 133, 375-381. https://doi.org/10.1016/j.clim.2009.08.017

  84. 84. Nijland, M.L., Koens, L., Pals, S.T., Berge, I., Bemelman, F.J. and Kersten, M.J. (2018) Clinicopathological Characteristics of T-Cell Non-Hodgkin Lymphoma Arising in Patients with Immunodeficiencies: A Single-Center Case Series of 25 Patients and a Review of the Literature. Haematologica, 103, 486-496. https://doi.org/10.3324/haematol.2017.169987

  85. 85. Mortaz, E., Tabarsi, P., Mansouri, D., Khosravi, A., Garssen, J., Velayati, A., et al. (2016) Cancers Related to Immunodeficiencies: Update and Perspectives. Frontiers in Immunology, 7, Article No. 365. https://doi.org/10.3389/fimmu.2016.00365

  86. 86. Buck, D., Moshous, D., de Chasseval, R., Ma, Y. and le Deist, F., Cavazzana-Calvo, M., Fischer, A., Casanova, J.L., Lieber, M.R. and de Villartay, J.P. (2006) Severe Combined Immune Deficiency and Microcephaly in Siblings with Hypomorphic Mutations in DNA Ligase IV. European Journal of Immunology, 36, 224-235. https://doi.org/10.1002/eji.200535401

  87. 87. Rambo, R.P. and Tainer, J.A. (2010) Bridging the Solution Divide: Comprehensive Structural Analyses of Dynamic RNA, DNA, and Protein Assemblies by Small-Angle X-Ray Scattering. Current Opinion in Structural Biology, 20, 128-137. https://doi.org/10.1016/j.sbi.2009.12.015

  88. 88. Clewing, J.M., Antalfy, B.C., Lücke, T., Najafian, B., Marwedel, K.M., Hori, A., et al. (2007) Schimke Immuno-Osseous Dysplasia: A Clinicopathological Correlation. Journal of Medical Genetics, 44, 122-130. https://doi.org/10.1136/jmg.2006.044313

  89. 89. Morimoto, M., Yu, Z., Stenzel, P., Marietta Clewing, J., Najafian, B., Mayfield, C., et al. (2012) Reduced Elastogenesis: A Clue to the Arteriosclerosis and Emphysematous Changes in Schimke Immuno-Osseous Dysplasia? Orphanet Journal of Rare Diseases, 7, Article No. 70. https://doi.org/10.1186/1750-1172-7-70

  90. 90. Fritze, O., Romero, B., Schleicher, M., Jacob, M.P., Oh, D.-Y., Starcher, B., et al. (2012) Age-Related Changes in the elastic Tissue of the Human Aorta. Journal of Vascular Research, 49, 77-86. https://doi.org/10.1159/000331278

  91. 91. Haffner, D.N., Rollins, N.K. and Dowling, M.M. (2019) Reversible Cerebral Vasoconstriction Syndrome: A Novel Mechanism for Neurological Complications in Schimke Immuno-Osseous Dysplasia. Pediatric Neurology, 92, 67-70. https://doi.org/10.1016/j.pediatrneurol.2018.10.022

  92. 92. Boerkoel, C.F, Takashima, H., John, J., Yan, J., Stankiewicz, P., Rosenbarker, L., et al. (2002) Mutant Chromatin Remodeling Protein SMARCAL1 Causes Schimke Immuno-Osseous Dysplasia. Nature Genetics, 30, 215-220. https://doi.org/10.1038/ng821

  93. 93. Lücke, T., Franke, D., Clewing, J.M., Boerkoel, C.F., Ehrich, J.H., Das, A.M. and Zivicnjak, M. (2006) Schimke Versus Non-Schimke Chronic Kidney Disease: An Anthropometric Approach. Pediatrics, 118, e400-e407. https://doi.org/10.1542/peds.2005-2614

  94. 94. Simon, A.J., Lev, A., Jeison, M., Borochowitz, Z.U., Korn, D., Lerenthal, Y. and Somech, R. (2014) Novel SMARCAL1 bi-Allelic Mutations Associated with a Chromosomal Breakage Phenotype in a Severe SIOD Patient. Journal of Clinical Immunology, 34, 76-83. https://doi.org/10.1007/s10875-013-9957-3

  95. 95. Lucke, T., Billing, H., Sloan, E.A., Boerkoel, C.F., Franke, D., Zimmering, M., et al. (2005) Schimke-Immuno-Osseous Dysplasia: New Mutation with Weak Genotype-Phenotype Correlation in Siblings. American Journal of Medical Genetics Part A, 135, 202-205. https://doi.org/10.1002/ajmg.a.30691

  96. 96. Driscoll, R., Cimprich, K. and Har, A. (2009) Ping on about the DNA Damage Response during Replication. Genes & Development, 23, 2359-2365. https://doi.org/10.1101/gad.1860609

  97. 97. Elizondo, L.I, Huang, C., Northrop, J.L., Deguchi, K., Clewing, J.M., Armstrong, D.L., et al. (2006) Schimke Immuno-Osseous Dysplasia: A Cell Autonomous Disorder? American Journal of Medical Genetics Part A, 140A, 340-348. https://doi.org/10.1002/ajmg.a.31089

  98. 98. Taha, D., Boerkoel, C.F., Balfe, J.W., Khalifah, M., Sloan, E.A., Barbar, M., Haider, A. and Kanaan, H. (2004) Fatal Lymphoproliferative Disorder in a Child with Schimke Immuno-Osseous Dysplasia. American Journal of Medical Genetics Part A, 131A, 194-199. https://doi.org/10.1002/ajmg.a.30356

  99. 99. Bandino, A., Compagnone, A., Bravoco, V., Cravanzola, C., Lomartire, A., Rossetto, C., et al. (2008) Beta-Catenin Triggers Nuclear Factor KappaB-Dependent Up-Regulation of Hepatocyte Inducible Nitric Oxide Synthase. The International Journal of Biochemistry & Cell Biology, 40, 1861-1871. https://doi.org/10.1016/j.biocel.2008.01.029

  100. 100. Puccetti, M.V., Fischer, M.A., Arrate, M.P., Boyd, K.L., Duszynski, R.J., Betous, R., Cortez, D. and Eischen, C.M. (2017) Defective Replication Stress Response Inhibits Lymphomagenesis and Impairs Lymphocyte Reconstitution. Oncogene, 36, 2553-2564. https://doi.org/10.1038/onc.2016.408

  101. 101. Lou, S., Lamfers, P., Mcguire, N. and Boerkoel, C.F. (2002) Longevity in Schimke Immuno-Osseous Dysplasia. Journal of Medical Genetics, 39, 922-925. https://doi.org/10.1136/jmg.39.12.922

  102. NOTES

    *通讯作者。

期刊菜单