RUNX3抑制胃癌转移的作用及分子机制研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
胃癌目前仍是我国死亡人数最多的恶性肿瘤之一,五年生存率仅约20%。转移是肿瘤最具破坏性的本质,也是胃癌患者死亡的主要原因。深刻理解胃癌转移的分子机制对治疗十分重要。Runt结构域转录因子(RUNX1、RUNX2和RUNX3)是一组同源性较高的转录因子,它们与核心结合因子β亚基结合形成异源二聚体,作用于基因的启动子、增强子或沉默子区域,调节基因的表达。RUNX-CBFbeta复合体对基因的调节具有组织和阶段特异性,RUNX蛋白在个体发育中具有重要的作用;在受调节的靶基因中包括许多在肿瘤中发生变化的、与细胞命运相关的基因。
     RUNX3是Runt结构域转录因子家族成员,最近被确定为抑癌基因,在肝癌、胰腺癌、前列腺癌、肺癌中都发现有不同比例的RUNX3表达缺失。研究发现在60%~90%的胃癌病例中存在RUNX3的缺失,与胃癌患者的预后情况呈负相关。随着胃癌分期的进展,缺失比例增高。RUNX3敲除小鼠的胃粘膜上皮细胞发生恶性转化,激活胃癌细胞的RUNX3表达可以显著抑制胃癌细胞的生长和成瘤性。因此RUNX3表达下降或缺失在胃癌的发生和进展中具有重要作用。文献表明,随着胃癌分期进展RUNX3缺失的比例增加。在胃癌肝转移细胞中RUNX3的两等位基因全部缺失,这些均提示RUNX3在胃癌的转移中可能具有重要作用,但RUNX3的表达下降是否与胃癌转移相关还不明确。本课题将对此进行探讨。
     【目的】
     1、明确RUNX3下降与胃癌侵袭、转移是否相关;2、探讨RUNX3抑制胃癌转移的可能分子机制;3、筛选RUNX3调控的下游基因。
     【方法】
     1、利用免疫组化观察RUNX3在转移和无转移的胃癌组织中的表达;2、基因重组的方法构建含有RUNX3特异性SiRNA的表达载体;3、用构建的特异性SiRNA的表达载体和RUNX3的真核表达载体(Paul J Farrell教授赠予)转染胃癌细胞MKN28和SGC7901,经G418筛选获得稳定表达的细胞系。3、利用失巢培养模型、损伤刮擦实验、体外侵袭实验、裸鼠成瘤及Ⅷ因子染色、裸鼠体内转移实验研究RUNX3对胃癌细胞MKN28和SGC7901侵袭转移能力的影响;4、凝胶酶谱法检测转染胃癌细胞上清中分泌MMP2,MMP9的活性变化;5、RT-PCR及Western blot检测MMP2,MMP9,TIMP-1在mRNA和蛋白水平的表达变化;ELISA检测TIMP-1、VEGF在转染细胞培养上清中的水平。6、基因芯片技术筛选RUNX3相关的差异表达基因,并用RT-PCR进行个别验证。7、双荧光素酶报告基因实验检测RUNX3对TIMP-1启动子活性的调控作用。通过染色质免疫共沉淀实验(ChIP)和电泳迁移率实验(EMSA)验证细胞内RUNX3与TIMP-1启动子是否相互作用。
     【结果】
     1、RUNX3在转移的胃癌组织中表达下降,RUNX3可抑制胃癌细胞的转移能力
     利用免疫组化技术检测了83例无转移胃癌和40例有转移胃癌中RUNX3的表达,发现RUNX3在无转移的胃癌中表达的阳性率(62.7% (52/83))高于有转移的胃癌(42.5% (17/40)),免疫组化评分结果显示,RUNX3在无转移的胃癌中表达高于有转移的胃癌(1.25±0.48 versus 0.57±0.26, P<0.05),说明RUNX3更多地表达在无转移的胃癌组织。
     成功构建了含有RUNX3特异性SiRNA的表达质粒pSilencer-RUNX3。选取胃癌细胞SGC7901(表达中等强度RUNX3)和MKN28(无RUNX3的表达),pBK-RUNX3转染细胞后能上调RUNX3的表达,pSilencer- RUNX3转染细胞后能显著抑制RUNX3的表达。pBK-RUNX3转染细胞后,MKN28和SGC7901穿过Transewell小室微孔膜的侵袭能力下降,抑制率分别为40.3%和42.5%,而pSilencer-RUNX3转染细胞后,SGC7901细胞穿过Transewell小室微孔膜的侵袭能力显著增加。损伤刮擦实验表明,RUNX3可明显抑制胃癌细胞的迁移运动能力。MTT实验表明,RUNX3抑制细胞的失巢存活能力,转染pBK-RUNX3的SGC7901细胞和MKN28细胞在失巢培养24小时后,其细胞生存率均明显低于空载体对照组(SGC7901:32.2% vs 55.7%, MKN28:58.8% vs 80.7% P<0.05),而转染pSilencer-RUNX3的SGC7901细胞组的细胞生存率均高于空载体对照组(75.3% vs 53.6% P<0.05)。裸鼠成瘤后瘤体固定切片,Ⅷ因子染色表明转染pBK-RUNX3瘤体中微血管密度(MDV)显著下降(P<0.05),而pSilencer- RUNX3转染细胞后瘤体中MDV增多(P<0.05),说明RUNX3可抑制胃癌新生血管生成。裸鼠体内转移实验结果显示转染pBK-RUNX3后细胞在裸鼠肺部及肝部形成的可见转移结节明显少于对照(SGC7901和MKN28中均为P<0.05)而转染pSilencer-RUNX3后形成的可见转移结节明显多于对照组(P<0.05)。
     2、RUNX3可以通过调节TIMP-1、VEGF、CRKⅡ等分子的表达抑制胃癌细胞侵袭转移的能力
     我们通过凝胶酶谱试验发现RUNX3可抑制胃癌细胞中MMP9的活性(SGC7901和MKN28中均为P<0.05),RUNX3特异性siRNA则增强SGC7901细胞MMP9的活性(P<0.05);而对MMP2的活性无显著影响。RT-PCR和Western blot检测发现MMP2和MMP9的表达水平并无明显变化。因此,RUNX3对MMP9的调节可能不在mRNA和蛋白水平,而可能通过其它分子影响MMP9的活性。TIMP-1是MMP9主要的内源抑制因子,进一步RT-PCR、Western blot及ELISA的结果证实,RUNX3能上调TIMP-1在mRNA和蛋白水平的表达以及胃癌细胞培养上清中TIMP-1的水平,因此RUNX3可通过上调TIMP-1的表达水平,从而抑制MMP9的酶活性;ELISA还发现RUNX3可以抑制胃癌细胞培养上清中VEGF的表达水平,提示RUNX3通过抑制VEGF的表达而减少胃癌的血管生成。
     经过Trizol法抽提SGC7901/ pBK-RUNX3和空载体转染的对照细胞SGC7901/ pBK-CMV两种细胞的总RNA,通过琼脂糖电泳检查及紫外定量分析表明所提取得RNA无明显降解,其质和量均符合芯片检测要求。经荧光探针的制备、纯化及定量,芯片杂交、图像采集和数据分析等步骤得到差异表达的基因。我们以两种细胞信号强度的比值小于0.5或者大于2作为判定标准,结果发现,与对照细胞相比,RUNX3表达上调后,SGC7901/ pBK-RUNX3细胞中有14个基因表达被上调,109个基因表达被下调。
     在基因芯片筛选出的下游基因中,我们通过RT-PCR对个别差异表达得基因进行验证,证实接头分子CRKⅡ的表达在胃癌细胞MKN28和SGC7901中可被RUNX3抑制。
     3、RUNX3作为转录因子增强TIMP-1在转录水平的表达双荧光素酶报告基因实验结果显示, pBK-RUNX3质粒与pGL-TIMP-1共转染MKN28细胞后,细胞内的激发荧光强度显著高于质粒与pBK-CMV空载体转染的MKN28细胞。表明RUNX3可以增强TIMP-1启动子的活性
     ChIP实验发现,设计扩增包含TIMP-1启动子上2个RUNX3结合位点的序列的特异性引物,经RUNX3抗体沉淀后的DNA模版可以扩增出特异性的片段,提示无论是内源性表达的RUNX3还是外源性的RUNX3均可在细胞内与TIMP-1的启动子结合并相互作用。
     进一步EMSA和Supper Shift试验结果显示:SGC7901细胞核蛋白提取物可以使TIMP-1启动子的双链DNA探针的电泳条带滞后,而RUNX3抗体可以使电泳条带进一步滞后。表明RUNX3可以与TIMP-1启动子上的两个保守结合位点结合。
     【结论】
     本研究提供了抑癌基因RUNX3具有抑制胃癌转移作用的证据,同时,我们发现这种作用可能与以下机制有关:RUNX3作为转录因子,增强下游靶基因TIMP-1的转录及表达,抑制MMP9的活性,从而抑制胃癌细胞侵袭能力;RUNX3抑制下游基因CRKⅡ的表达,从而抑制胃癌细胞的运动迁移能力及失巢存活能力;RUNX3抑制下游基因VEGF的表达,从而抑制胃癌血管生成。这些发现有助于RUNX3在胃癌转移的分子治疗和诊断中的应用。
Despite the progress in diagnosis and treatment of gastric cancer, only about 20% of patients survive to 5 years. Metastasis, the most fearful aspect of cancer, is one of the major causes of mortality for gastric cancer patients. In order to improve the treatment of this disease, it is of great importance to clearly understand the molecular mechanism of gastric cancer metastasis. All the RUNX family members, namely, RUNX1, RUNX2 and RUNX3, encode DNA bindingαsubunits which form heterodimers with the commonβsubunit CBFβand act as transcription regulators. It has been realized that all three RUNX family members play important roles in normal developmental processes and carcinogenesis. They regulate the expression of cellular genes through binding to promoters, enhancer or silencer elements, with growing list of targes that including many genes relevant to carcinogenesis.
     Recently, RUNX3, one of Runt-related genes, is taken as a candidate tumor suppressor gene whose deficiency is causally related with gastric cancer. In RUNX3-knock-out mice, the stomach mucosal cells showed characteristics of malignant transformation. Reactivation of RUNX3 in gastric cancer cells inhibited their growth and tumorigenicity. There have emerged numerous reports on gene methylation and silencing of RUNX3 in tumors of many other tissues, such as liver, colon, gallbladder, lung, etc. Furthermore, decrease of RUNX3 protein expression was significantly associated with inferior survival duration of gastric cancer patients. It was reported RUNX3 expression was decreased in about 60% of the analyzed primary human gastric tumors, while the reduced RUNX3 levels rising to nearly 90% among the late stage, representing highly metastatic tumors. These data indicated that RUNX3 deficiency is closely related with gastric cancer metastasis.
     However, the function of RUNX3 in pathogenesis of gastric cancer metastasis remains not fully understood. So in the present work, we investigated whether RUNX3 might supress the invasive and metastatic abilities of gastric cancer cells and explored the relative molecular mechanism involved.
     【Objectives】
     (1) To investigate whether RUNX3 expression level is assocciated with invasion and metastasis of gastric cancer; (2) To examine the possible molecular mechanisms underlying the metastasis-suppressing effect by RUNX3 in gastric cancer; (3) To screen the downstream effectors of RUNX3
     【Methods】
     (1) The expression level of RUNX3 in metastatic gastric cancer and non-metastatic gastric cancer were determined by immunohistochemistry assay. (2) The RUNX3-specific siRNA vector was designed and constructed. (3) The plasmids pBK-RUNX3 (kindly provided by Prof. Paul J Farrell) or pSilencer-RUNX3 was transfected into SGC7901 and MNK28 cells using lipofectamine 2000 reagent. Empty vector of pBK-CMV or pSilencer were used as control. The transfected cells were screened with G418 for 2 months, single cell colonies were obtained by limited dilution method. Semi-quantitative RT-PCR and Western blot were used to detect the RUNX3 expression in the transfected cells for confirmation of transfection. (4) The effects of RUNX3 on the motility, invasive, angiogensis and in vivo metastatic abilities of gastric cancer cell lines SGC7901 and MKN28 were respectively investigated by wound-healing assays, invasion assay, tail vein metastatic assay andⅧfactor staining. (5) The activities of MMP2 and MMP9 in the conditioned medium were visualized by performing gelatin zymography (6) The mRNA and protein expression of MMP2, MMP9 and TIMP-1 were determined by RT-PCR and Western blotting. (7) The TIMP-1 and VEGF protein levels in the culture supernatants were determined using ELISA. (8) The TIMP-1 promoter was amplified by PCR using the genomic DNA from SGC7901 cells as template. The effect on TIMP-1 promoter activity by RUNX3 was evaluated by dual luciferase reporter assay. (9)The direct interaction between RUNX3 and the TIMP-1 promoter in vivo was determined by performing ChIP assay. (10) EMSA and supper shift assay were performed to visualize the binding of RUNX3 to TIMP-1 promoter via the two putative binding sites. (11) The downstream effectors of RUNX3 were screened cDNA microarray.
     【Results】
     1.RUNX3 expression is down-regulated in metastatic gastric cancer. To investigate the relationship between the expression of RUNX3 and metastasis of gastric cancer, we compared expression of RUNX3 in primary sites from 83 patients enduring non-metastatic gastric cancer with those from 40 patients enduring metastatic gastric cancer. The positive rate of RUNX3 expression in non-metastatic gastric cancer was 62.7% (52/83), lower than 42.5% (17/40) in metastatic gastric cancer. When considered staining scores, the average staining score in metastatic gastric cancer was significantly higher than that in non-metastatic gastric cancer (1.25±0.48 versus 0.57±0.26, P<0.05). These results suggested that RUNX3 expression level was negatively related to gastric cancer metastasis.
     2.Reduced RUNX3 expression is associated with increased metastatic potential of gastric cancer cells in vitro and in vivo
     RT-PCR and Western blot suggested that transfecting SGC7901 and MKN28 cells with pBK/RUNX3 increased the RUNX3 expression in the cells. And transfecting pSilencer-RUNX3 decreased the RUNX3 expression in the SGC7901 cells.
     Upregulation of RUNX3 in SGC7901 and MKN28 cells decreased their abilities to migrate on matrigel, and reduced RUNX3 in SGC7901 cells inceased its mobility. The RUNX3-transfected cells also showed significantly lower invasive potency than that of the control. The RUNX3 transfection reduced the number of invasive cells by 42.5% and 40.3% in SGC7901 cells and MKN28 cells, respectively. The difference in invasiveness was found to be statistically significant (P<0.05 in both cell lines). Similarly, inhibition of RUNX3 expression in SGC7901 cells by RUNX3 siRNA constructs promoted the cell invasive potency significantly compared with the control (P<0.05). MTT assay also showed that the survival rates of the detached SGC7901 and MKN28 cells transfected with pBK-RUNX3 were markedly decrease as compared with control group ( P < 0.05) . Survival rate of SGC7901 cells transfected with pSilencer-RUNX3 were significantly increase as compared with control (P < 0.05). The above results suggested that RUNX3 play a negative role in the gastric cancer cell resistant to anoilkis.
     RUNX3 inhibited angiogenesis in xenograft tumor in nude mice (p<0.05). We performed IHC analysis on various tumor sections to assess angiogenesis in mice. The MVD (miro-vessel density) of tumors in mice subcutaneously inoculated with gastric cancer cells trasfected with pBK-RUNX3 were lower than that trasfected with pBK-CMV (6.50±2.13 vs 17.30±8.32, 23.72±9.13 vs 50.25±15.46 respectively in SGC7901 and MKN28, P < 0.05 in both cell lines). And similarly, the MVD in tumor of mice subcutaneously inoculated with SGC7901/pSilencer-RUNX3 were higher than that of SGC7901/pSilencer (28.50±12.73 vs 12.83±4.92 P < 0.05). Tail vein metastatic assay in nude mice was further adopted to examine the suppressing effect on metastatic abilities of gastric cancer cells by RUNX3. Compared with control cells transfected with empty vectors, the injection of the cells transfected with pBK-RUNX3 via a tail vein of athymic nude mice led to significantly less visible tumors in the liver and lung surface (P< 0.05). Conformably, the injection of the SGC7901 cells transfected with pSilencer-RUNX3 led to much more visible tumors in the liver and lung surface than that of control cells (P<0.05). The results above suggest that RUNX3 could suppress the moltility, invasiveness, angiogenesis and metastatic potential of gastric cancer cells.
     3. RUNX3 regulates the expression level of downstream genes that are relevant to tumor metastasis.
     To study the possible role of MMPs in RUNX3-induced inhibition of cell invasion, we first analyzed the regulation on the expression and activity of MMP2 ,9 by RUNX3. The RT-PCR and Western blot were used to examine the mRNA and protein expression of MMP2 ,9 in the gastric cancer cells. We found no significant difference in MMP2 or MMP9 expression between cells with up-regulated RUNX3 (or down- regulated RUNX3) and their respective controls
     At the same time, the activities of MMP2, 9 in the serum-free conditioned medium were determined by gelatin zymography. MMP9 that exhibited collagen-degrading activity was detected in the culture supernatant of the SGC7901 and MKN28 cells. Compared with the control cells, the activity of the MMP9 enzyme in pBK-RUNX3 transfectants was significantly lower (both P<0.05 in SGC7901 and MKN28). And the activity of the MMP9 enzyme in pSilencer-RUNX3 transfectants was significantly higher than that of the control cells (P<0.05). While the activity of MMP2 is relatively low compared with that of MMP9 and showed no significant difference between cells with up-regulated RUNX3 (or down- regulated RUNX3) and their respective controls. Above results indicated that instead of regulating the transcription of MMP9, the enhanced expression of RUNX3 attenuate the matrix degrading activity of MMP9 in gastric cancer cells.
     TIMP1 expression is up-regulated by RUNX3 at transcriptional level. To examine the expression level of the TIMP1 which is considered the specific inhibitor of MMP-9 activity, we performed RT-PCR to detect the mRNA expression level and Western blot to detect the protein expression. We found that transfection of pBK-RUNX3 induced significantly stronger expression of TIMP-1 in SGC7901/pBK-RUNX3 and MKN28/pBK-RUNX3 cells. However, weaker expression of TIMP-1 was observed in cells transfected with pSilencer-RUNX3 as comparing with the control cells.
     The TIMP-1 levels secreted in the conditioned culture medium of gastric cancer cells were detected by performing ELISA We found that TIMP-1 production in the medium of the gastric cells transfected with pBK-RUNX3 was signifiantly higher than that of the respective control cells (16.5 ng/ml vs10.6 ng/ml in SGC7901, P<0.05; 14.9 ng/ml vs 7.2ng/ml in MKN28, P<0.05). Conformably, TIMP-1 level in the medium of SGC7901 /pSilencer -RUNX3 deceased compared with the control cells SGC7901/pSilencer (5.7 ng/ml vs 12.5 ng/ml, P<0.05).
     The VEGF levels secreted in the conditioned culture medium of gastric cancer cells were detected by performing ELISA. The VEGF production in the medium of the gastric cells transfected with pBK-RUNX3 was signifiantly lower than that of the respective control cells (1270±580 pg/millioncell vs 2950±870pg/millioncell in SGC7901, P<0.05; 2060±790 pg/millioncell vs 4300±1130 pg/millioncell in MKN28, P<0.05). VEGF level in the medium of SGC7901/pSilencer-RUNX3 increased when compared with the control cells SGC7901/pSilencer (6500±1730 pg/millioncell vs 3100±1066 pg/millioncell P<0.05).
     The total RNA of SGC7901/pBK-RUNX3 and SGC7901/pBK-CMV were extracted. The quality of extracted RNA was evaluated by agarose electrophoresis and UV absorbance spectroscopy. After microarray hybridization and data normalization, 14 genes were showed up-regulated while 109 genes were down-regulated by RUNX3. Among those genes, the expression level of CRKⅡ was evaluated by RT-PCR. It was confirmed that CRKⅡexpression was reduced by RUNX3, which indicated that inhibiting effect of RUNX3 on gastric cancer cell mobility was at least partially mediated by CRKⅡdownregulation.
     4. RUNX3 regulated TIMP-1 expression in gastric cancer cells by directly interacting with TIMP-1 promoter
     To investigate the possibility that RUNX3 increased TIMP-1 expression by stimulating the TIMP-1 promoter activity, the dual luciferase reporter assay was performed. The transfection of different doses of pBK-RUNX3 plasmid led to a significantly increase of TIMP-1 promoter activity compared with the empty vector transfection. The result indicated that RUNX3 caused transactivation of TIMP-1 promoter and then up-regulated the mRNA and protein expression of TIMP-1.
     Further analysis of TIMP-1 revealed two putative RUNX binding sites was contained in the TIMP-1 promoter (-327 to -321; -66 to -61). Therefore, we tried to determine whether RUNX3 binds to the regions of the TIMP-1 promoter in vivo by performing ChIP assay. The chromatin fragments from SGC7901 cells and MKN28/ pBK-RUNX3 were immunoprecipitated with specific anti-RUNX3 antibody or with normal rabbit serum as a negative control. We found that both exogenously expressed RUNX3 in MKN28/ pBK-RUNX3 cells and endogenous RUNX3 in SGC7901 cells could be recruited to the two sites.
     Moreover, EMSA was conducted to demonstrate whether RUNX3 could bind to the two putative binding sites. The nuclear extracts of SGC7901 cells were prepared. Biotin-labeled probes including the RUNX3-binding site were incubated with the nuclear extracts, and DNA-protein complexes were analyzed with PAGE. For both sites, shifted bands of labeled probes were detected with nuclear extracts from SGC7901 cells. These bands were dramatically inhibited when competitive cold probes were included in the reaction, while the mutated cold probes could not inhibit the specific binding between labeled probes and nuclear protein. The shifted band was supershifted by addition of the anti-RUNX3 antibody. These observations suggest that the band was a specific complex formed by RUNX3 and the probes including the RUNX3-binding sites and RUNX3 did bind to the putative RUNX binding sites inTIMP-1gene promoter.
     【Conclusion】
     We provide evidence that RUNX3 expression was down-regulated in metastatic gastric cancer tissues than metastatic gastric cancer tissues. Our study presented evidence for RUNX3-mediated suppression of gastric cancer metastasis by inhibting the invasiveness, mobility and angiogenesis of cancer cells. We also provided novel molecular mechanism that at least partly contributes to the metastasis inhibiting activity of RUNX3. These data may be helpful for beneficial application of RUNX3 to diagnostics and therapeutics of gastric cancer metastasis.
引文
1. van Wijnen, A.J., et al., 2004. Nomenclature for Runt-related (RUNX) proteins. Oncogene 23, 4209–4210.
    2. Berardi MJ, Sun C, Zehr M, Abildgaard F, Peng J, Speck NA et al. The Ig fold of the core binding factor alpha Runt domain is a member of a family of structurally and functionally related Ig-fold DNA-binding domains. Structure 1999;7:1247–56.
    3. Ghozi MC, Bernstein Y, Negreanu V, Levanon D, Groner Y. 1996. Expression of the human acute myeloid leukemia gene AML1 is regulated by two promoter regions. Proc Natl Acad Sci USA 93:1935–1940.
    4. Zambotti A, Makhluf H, Shen J, Ducy P. 2002. Characterization of an osteoblast-specific enhancer element in the CBFA1 gene. J Biol Chem 277:41497–41506.
    5. Kanzler B, Kuschert SJ, Liu YH, Mallo M. 1998. Hoxa-2 restricts the chondrogenic domain and inhibits bone formation during development of the branchial area. Dev Suppl 125:2587–2597.
    6. Lecka-Czernik B, Gubrij I, Moerman EJ, Kajkenova O, Lipschitz DA, Manolagas SC, Jilka RL. 1999. Inhibition of Osf2/Cbfa1 expression and terminal osteoblast differentiation by PPARgamma2. J Cell Biochem 74:357– 371
    7. Satokata I, Ma L, Ohshima H, Bei M, Woo I, Nishizawa K, Maeda T, Takano Y, Uchiyama M, Heaney S, Peters H, Tang Z, Maxson R, Maas R. 2000. Msx2 deficiency in mice causes pleiotropic defects in bone growth and ectodermal organ formation. Nat Genet 24:391–395.
    8. Tribioli C, Lufkin T. 1999. The murine Bapx1 homeobox gene plays a critical role in embryonic development of the axial skeleton and spleen. Dev Suppl 126:5699–5711.
    9. Bae, S.-C., Takahashi, E.-i., Zhang, Y.W., Ogawa, E., Shigesada, K.,Namba, Y., Satake, M., Ito, Y., 1995. Cloning, mapping and expression of PEBP2aC, a third gene encoding the mammalian runt domain. Gene 159, 245–248.
    10.Levanon, D., Negreanu, V., Bernstein, Y., Bar-Am, I., Avivi, L., Groner, Y., 1994. AML1, AML2, and AML3, the human members of the runt domain gene-family: cDNA structure, expression, and chromosomal localization. Genomics 23, 425–432.
    11.Leiden, J.M., Thompson, C.B., 1994. Transcriptional regulation of T-cell genes during T-cell development. Curr. Opin. Immunol. 6, 231–237.
    12.O’Riordan, M., Grosschedl, R., 2000. Transcriptional regulation of early Blymphocyte differentiation. Immunol. Rev. 175, 94–103.
    13.Bangsow C, Rubins N, Glusman G, Bernstein Y, Negreanu V, Goldenberg D, Lotem J, Ben-Asher E, Lancet D, Levanon D, Groner Y. The RUNX3 gene--sequence, structure and regulated expression. Gene. 2001 Nov 28;279(2):221-32.
    14.Rini D, Calabi F. Identification and comparative analysis of a second runx3 promoter. Gene. 2001 Jul 25;273(1):13-22.
    15.Drissi H, Pouliot A, Koolloos C, Stein JL, Lian JB, Stein GS, van Wijnen AJ. 1,25-(OH)2-vitamin D3suppresses the bone-related Runx2/Cbfa1 gene promoter. Exp Cell Res. 2002 Apr 1;274(2):323-33.
    16.Spender LC, Whiteman HJ, Karstegl CE, Farrell PJ. Transcriptional cross-regulation of RUNX1 by RUNX3 in human B cells. Oncogene. 2005 Mar 10;24 (11):1873-81.
    17.Tanaka, T., et al., 1996. The extracellular signal-regulated kinase pathway phosphorylates AML1, an acute myeloid leukemia gene product, and potentially regulates its transactivation ability. Mol. Cell. Biol. 16, 3967–3979.
    18.Zhang, Y., Biggs, J.R., Kraft, A.S., 2004. Phorbol ester treatment of K562 cells regulates the transcriptional activity of AML1c through phosphorylation. J. Biol. Chem. 279, 53116–53125.
    19.Ito, Y., 1999. Molecular basis of tissue-specific gene expression mediated by the runt domain transcription factor PEBP2/CBF. Genes Cells 4, 685–696
    20.Kanno, T., Kanno, Y., Chen, L.F., Ogawa, E., Kim,W.Y., Ito, Y., 1998. Intrinsic transcriptional activation-inhibition domains of the polyomavirus enhancer binding protein 2/core binding factor alpha subunit revealed in the presence of the beta subunit. Mol. Cell. Biol. 18, 2444–2454.
    21.Imai, Y., et al., 2004. The corepressor mSin3A regulates phosphorylationinduced activation, intranuclear location, and stability of AML1. Mol. Cell. Biol. 24, 1033–1043
    22.Lutterbach, B., Westendorf, J.J., Linggi, B., Isaac, S., Seto, E., Hiebert, S.W.,2000. A mechanism of repression by acute myeloid leukemia-1, the target of multiple chromosomal translocations in acute leukemia. J. Biol. Chem. 275, 651–656.
    23.Huang, G., Shigesada, K., Ito, K., Wee, H.J., Yokomizo, T., Ito, Y., 2001. Dimerization with PEBP2beta protects RUNX1/AML1 from ubiquitinproteasome- mediated degradation. EMBO J. 20, 723–733.
    24.Brooks, P.C., Klemke, R.L., Schon, S., Lewis, J.M., Schwartz, M.A., Cheresh, D.A., 1997. Insulin-like growth factor receptor cooperates with integrin alpha v beta 5 to promote tumor cell dissemination in vivo. J. Clin. Invest. 99, 1390–1398.
    25.Lopaczynski,W., 1999. Differential regulation of signaling pathways for insulin and insulin-like growth factor I. Acta Biochim. Pol. 46, 51–60.
    26.Shelton, J.G., Steelman, L.S., White, E.R., McCubrey, J.A., 2004. Synergy between PI3K/Akt and Raf/MEK/ERK pathways in IGF-1R mediated cell cycle progression and prevention of apoptosis in hematopoietic cells. Cell Cycle 3, 372–379.
    27.Sun, L., Vitolo, M., Passaniti, A., 2001. Runt-related gene 2 in endothelial cells: inducible expression and specific regulation of cell migration and invasion. Cancer Res. 61, 4994–5001.
    28.Qiao, M., Shapiro, P., Kumar, R., Passaniti, A., 2004. Insulin-like growth factor-1 regulates endogenous RUNX2 activity in endothelial cells through a phosphatidylinositol 3-kinase/ERK-dependent and Akt-independent signaling pathway. J. Biol. Chem. 279, 42709–42718.
    29.Fujita, T., et al., 2004. Runx2 induces osteoblast and chondrocyte differentiation and enhances their migration by coupling with PI3K-Akt signaling. J. Cell Biol. 166, 85–95.
    30.Jabs, E.W., et al., 1994. Jackson-Weiss and Crouzon syndromes are allelic with mutations in fibroblast growth factor receptor 2. Nat. Genet. 8, 275–279.
    31.Meyers, G.A., Orlow, S.J., Munro, I.R., Przylepa, K.A., Jabs, E.W., 1995. Fibroblast growth factorreceptor 3 (FGFR3) transmembrane mutation in Crouzon syndrome with acanthosis nigricans. Nat. Genet. 11, 462–464.
    32.Muenke, M., et al., 1994. A common mutation in the fibroblast growth factor receptor 1 gene in Pfeiffer syndrome. Nat. Genet. 8, 269–274.
    33.Reardon,W.,Winter, R.M., Rutland, P., Pulleyn, L.J., Jones, B.M., Malcolm, S., 1994. Mutations in the fibroblast growth factor receptor 2 gene cause Crouzon syndrome. Nat. Genet. 8, 98–103.
    34.Zhou, Y.X., Xu, X., Chen, L., Li, C., Brodie, S.G., Deng, C.X., 2000. A Pro250Arg substitution in mouse Fgfr1 causes increased expression of Cbfa1 and premature fusion of calvarial sutures. Hum. Mol. Genet. 9, 2001–2008.
    35.Kim, H.J., Kim, J.H., Bae, S.C., Choi, J.Y., Kim, H.J., Ryoo, H.M., 2003. The protein kinase C pathway plays a central role in the fibroblast growth factorstimulated expression and transactivation activity of Runx2. J. Biol. Chem. 278, 319–326.
    36.Xiao, G., et al., 2000. MAPK pathways activate and phosphorylate the osteoblast-specific transcription factor, Cbfa1. J. Biol. Chem. 275, 4453–4459.
    37.Xiao, G., Jiang, D., Gopalakrishnan, R., Franceschi, R.T., 2002. Fibroblast growth factor 2 induction of the osteocalcin gene requires MAPK activity and phosphorylation of the osteoblast transcription factor, Cbfa1/Runx2. J. Biol. Chem. 277, 36181–36187.
    38.Dempster, D.W., Cosman, F., Parisien, M., Shen, V., Lindsay, R., 1993. Anabolic actions of parathyroid hormone on bone. Endocr. Rev. 14, 690–709.
    39.Partridge, N.C., Alcorn, D., Michelangeli, V.P., Kemp, B.E., Ryan, G.B., Martin, T.J., 1981. Functional properties of hormonally responsive cultured normal and malignant rat osteoblastic cells. Endocrinology 108, 213–219.
    40.Lindsay, R., et al., 1997. Randomised controlled study of effect of parathyroid hormone on vertebral-bone mass and fracture incidence among postmenopausal women on oestrogen with osteoporosis. Lancet 350, 550–555.
    41.Swarthout, J.T., Doggett, T.A., Lemker, J.L., Partridge, N.C., 2001. Stimulation of extracellular signal-regulated kinases and proliferation in rat osteoblastic cells by parathyroid hormone is protein kinase C-dependent. J. Biol. Chem. 276, 7586–7592.
    42.Jimenez, M.J., Balbin, M., Lopez, J.M., Alvarez, J., Komori, T., Lopez-Otin, C., 1999. Collagenase 3 is a target of Cbfa1, a transcription factor of the runt gene family involved in bone formation. Mol. Cell. Biol. 19, 4431–4442.
    43.Quinn, C.O., Scott, D.K., Brinckerhoff, C.E., Matrisian, L.M., Jeffrey, J.J., Partridge, N.C., 1990. Rat collagenase. Cloning, amino acid sequence comparison, and parathyroid hormone regulation in osteoblastic cells. J. Biol. Chem. 265, 22342–22347.
    44.Massague, J., 1990. The transforming growth factor-beta family. Annu. Rev. Cell Biol. 6, 597–641.
    45.Hogan, B.L., 1996. Bone morphogenetic proteins: multifunctional regulators of vertebrate development. Genes Dev. 10, 1580–1594.
    46.Reddi, A.H., 1994. Bone and cartilage differentiation. Curr. Opin. Genet. Dev. 4, 737–744.
    47.Heldin, C.H., Miyazono, K., ten Dijke, P., 1997. TGF-beta signalling from cell membrane to nucleus through SMAD proteins. Nature 390, 465–471.
    48.Massague, J., 1998. TGF-beta signal transduction. Annu. Rev. Biochem. 67, 753–791.
    49.Zhang, Y., Derynck, R., 1999. Regulation of Smad signalling by protein associations and signalling crosstalk. Trends Cell Biol. 9, 274–279.
    50.Hanai, J., et al., 1999. Interaction and functional cooperation of PEBP2/CBF with Smads. Synergistic induction of the immunoglobulin germline Calpha promoter. J. Biol. Chem. 274, 31577–31582.
    51.Ito, Y., 2004. Oncogenic potential of the RUNX gene family: ‘overview’. Oncogene 23, 4198–4208.
    52.Miyazono, K., Maeda, S., Imamura, T., 2004. Coordinate regulation of cell growth and differentiation by TGF-beta superfamily and Runx proteins. Oncogene 23, 4232–4237.
    53.Ito, Y., Miyazono, K., 2003. RUNX transcription factors as key targets of TGF-beta superfamily signaling. Curr. Opin. Genet. Dev. 13, 43–47.
    54.Lee, K.S., et al., 2000. Runx2 is a common target of transforming growth factor beta1 and bone morphogenetic protein 2, and cooperation between Runx2 and Smad5 induces osteoblast-specific gene expression in the pluripotent mesenchymal precursor cell line C2C12. Mol. Cell. Biol. 20, 8783–8792.
    55.Gallea, S., et al., 2001. Activation of mitogen-activated protein kinase cascades is involved in regulation of bone morphogenetic protein-2-induced osteoblast differentiation in pluripotent C2C12 cells. Bone 28, 491–498.
    56.Hanafusa, H., et al., 1999. Involvement of the p38 mitogen-activated protein kinase pathway in transforming growth factor-beta-induced gene expression. J. Biol. Chem. 274, 27161–27167.
    57.Li, Q.L., et al., 2002. Causal relationship between the loss of RUNX3 expression and gastric cancer. Cell 109, 113–124.
    58.Yang, X.J., 2004. The diverse superfamily of lysine acetyltransferases and their roles in leukemia and other diseases. Nucleic Acids Res. 32, 959–976.
    59.Jin, Y.H., et al., 2004. Transforming growth factor-beta stimulates p300-dependent RUNX3 acetylation, which inhibits ubiquitination-mediated degradation. J. Biol. Chem. 279, 29409–29417.
    60.Zhao, M., Qiao, M., Oyajobi, B.O., Mundy, G.R., Chen, D., 2003. E3 ubiquitin ligase Smurf1 mediates core-binding factor alpha1/Runx2 degradation and plays a specific role in osteoblast differentiation. J. Biol. Chem. 278, 27939–27944.
    61.Kitabayashi, I., Yokoyama, A., Shimizu, K., Ohki, M., 1998. Interaction and functional cooperation of the leukemia-associated factors AML1 and p300 in myeloid cell differentiation. EMBO J. 17, 2994–3004.
    62.Westendorf, J.J., et al., 2002. Runx2 (Cbfa1, AML-3) interacts with histone deacetylase 6 and represses the p21(CIP1/WAF1) promoter. Mol. Cell. Biol. 22, 7982–7992.
    63.Yamaguchi, Y., et al., 2004. AML1 is functionally regulated through p300-mediated acetylation on specific lysine residues. J. Biol. Chem. 279, 15630–15638.
    64.Shen, X., Hu, P.P., Liberati, N.T., Datto, M.B., Frederick, J.P.,Wang, X.F., 1998. TGF-beta-induced phosphorylation of Smad3 regulates its interaction with coactivator p300/CREB-binding protein. Mol.Biol. Cell 9, 3309–3319.
    65.Wee, H.J., Huang, G., Shigesada, K., Ito, Y., 2002. Serine phosphorylation of RUNX2 with novel potential functions as negative regulatory mechanisms. EMBO Rep. 3, 967–974.
    66.Barlev, N.A., et al., 2001. Acetylation of p53 activates transcription through recruitment of coactivators/histone acetyltransferases. Mol. Cell 8, 1243–1254.
    67.Quack, I., et al., 1999. Mutation analysis of core binding factor A1 in patients with cleidocranial dysplasia. Am. J. Hum. Genet. 65, 1268–1278
    68.Kanno, Y., Kanno, T., Sakakura, C., Bae, S.C., Ito, Y., 1998. Cytoplasmic sequestration of the polyomavirus enhancer binding protein 2 (PEBP2)/core binding factor alpha (CBFalpha) subunit by the leukemia-related PEBP2/ CBFbeta-SMMHC fusion protein inhibits PEBP2/CBF-mediated transactivation. Mol. Cell. Biol. 18, 4252–4261.
    69.Zhang, Y.W., et al., 2000. A RUNX2/PEBP2alpha A/CBFA1 mutation displaying impaired transactivation and Smad interaction in cleidocranial dysplasia. Proc. Natl. Acad. Sci. U. S. A. 97, 10549–10554.
    70.Okuda, T.; van Deursen, J.; Hiebert, S. W.; Grosveld, G.; Downing, J. R. : AML1, the target of multiple chromosomal translocations in human leukemia, is essential for normal fetal liver hematopoiesis. Cell 84: 321-330, 1996.
    71.Wang, Q.; Stacy, T.; Binder, M.; Marin-Padilla, M.; Sharpe, A. H.; Speck, N. A. : Disruption of the Cbfa2 gene causes necrosis and hemorrhaging in the central nervous system and blocks definitive hematopoiesis. Proc. Nat. Acad. Sci. 93: 3444-3449, 1996.
    72.Okuda T, Takeda K, Fujita Y, et al. Biological characteristics of the leukemia-associated transcriptional factor AML1 disclosed by hematopoietic rescue of AML1-deficient embryonic stem cells by using a knock-in strategy. Mol Cell Biol 2000;20:319_/28.
    73.Miller JD, Stacy T, Liu PP, Speck NA. Core-binding factor b (CBFb), but not CBFb-smooth muscle myosin heavy chain, rescues definitive hematopoiesis in CBFb-deficient embryonic stem cells. Blood 2001;97:2248_/56
    74.Yokomizo T, Ogawa M, Osato M, et al. Requirement of Runx1/ AML1/PEBP2aB for the generation of haematopoietic cells from endothelial cells. Genes Cells 2001;6:13_/23.
    75.Levanon D, et al. Spatial and temporal expression pattern of Runx3 (Aml2) and Runx1 (Aml1) indicates non-redundant functions during mouse embryogenesis Mechanisms of Development Volume 109, Issue 2 , December 2001, Pages 413-417
    76.Levanon D, Bettoun D, Harris-Cerruti C, Woolf E, Negreanu V, Eilam R, Bernstein Y, Goldenberg D, Xiao C, Fliegauf M, Kremer E, Otto F, Brenner O, Lev-Tov A, Groner Y. The Runx3 transcription factor regulates development and survival of TrkC dorsal root ganglia neurons. EMBO J. 2002 Jul 1;21(13):3454-63.
    77.Inoue K, Ozaki S, Shiga T, Ito K, Masuda T, Okado N, Iseda T, Kawaguchi S, Ogawa M, Bae SC, Yamashita N, Itohara S, Kudo N, Ito Y. Runx3 controls the axonal projection of proprioceptive dorsal root ganglion neurons. Nat Neurosci. 2002 Oct;5(10):946-54.
    78.Inoue K, Ozaki S, Ito K, Iseda T, Kawaguchi S, Ogawa M, Bae SC, Yamashita N, Itohara S, Kudo N, Ito Y. Runx3 is essential for the target-specific axon pathfinding of trkc-expressing dorsal root ganglion neurons. Blood Cells Mol Dis. 2003 Mar-Apr;30(2):157-60.
    79.Otto, F. et al. Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development. Cell 1997;89, 765–771
    80.Inada, M. et al. Maturational disturbance of chondrocytes in Cbfa1-deficient mice. Dev. Dyn. 1999;214, 279–290
    81.Ducy, P. et al. Osf2/Cbfa1: a transcriptional activator of osteoblast differentiation. Cell. 1997;89, 747–754
    82.Miller, J. et al. The core-binding factor beta subunit is required for bone formation and hematopoietic maturation. Nat. Genet. 2002; 32, 645–649
    83.Komori, T. et al. (1997) Targeted disruption of Cbfa1 results in a complete lack of bone formation owing to maturational arrest of osteoblasts. Cell 89, 755–764
    84.Miller, J. et al. (2002) The core-binding factor beta subunit is required for bone formation and hematopoietic maturation. Nat. Genet. 32, 645–649
    85.Takeda, S. et al. (2001) Continuous expression of Cbfa1 in nonhypertrophic chondrocytes uncovers its ability to induce hypertrophic chondrocyte differentiation and partially rescues Cbfa1-deficient mice. Genes Dev. 15, 467–481
    86.Liu, W. et al. (2001) Overexpression of Cbfa1 in osteoblasts inhibits osteoblast maturation and causes osteopenia with multiple fractures. J. Cell Biol. 155, 157–166
    87.Geoffroy, V. et al. (2002) High bone resorption in adult aging transgenic mice overexpressing cbfa1/runx2 in cells of the osteoblastic lineage. Mol. Cell. Biol. 22, 6222–6233
    88.D'Souza, R. N.; Aberg, T.; Gaikwad, J.; Cavender, A.; Owen, M.; Karsenty, G.; Thesleff, I. Cbfa1 is required for epithelial-mesenchymal interactions regulating tooth development in mice. Development 126: 2911-2920, 1999.
    89.Gaikwad, J. S.; Cavender, A.; D'Souza, R. N. Identification of tooth-specific downstream targets of Runx2. Gene 279: 91-97, 2001.
    90.Devlin H, Sloan P. Early bone healing events in the human extraction socket. Int J Oral Maxillofac Surg. 2002 Dec;31(6):641-5.
    91.Mundlos, S.; Otto, F.; Mundlos, C.; Mulliken, J. B.; Aylsworth, A. S.; Albright, S.; Lindhout, D.; Cole, W. G.; Henn, W.; Knoll, J. H. M.; Owen, M. J.; Mertelsmann, R.; Zabel, B. U.; Olsen, B. R. : Mutations involving the transcription factor CBFA1 cause cleidocranial dysplasia. Cell 89: 773-779, 1997.
    92.Aberg T, Cavender A, Gaikwad JS, Bronckers AL, Wang X, Waltimo-Siren J, Thesleff I, D'Souza RN. Phenotypic changes in dentition of Runx2 homozygote-null mutant mice. J Histochem Cytochem. 2004 Jan;52(1):131-9.
    93.Cohen MM Jr. The new bone biology: pathologic, molecular, and clinical correlates. Am J Med Genet A. 2006 Dec 1;140(23):2646-706.
    94.Nishio Y, Dong Y, Paris M, O'Keefe RJ, Schwarz EM, Drissi H. Runx2-mediated regulation of the zincfinger Osterix/Sp7 gene. Gene. 2006 May 10;372:62-70.
    95.Zheng L, Iohara K, Ishikawa M, Into T, Takano-Yamamoto T, Matsushita K, Nakashima M. Runx3 negatively regulates Osterix expression in dental pulp cells. Biochem J. 2007 Mar 13;
    96.Taniuchi I, Osato M, Egawa T, Sunshine MJ, Bae SC, Komori T, Ito Y, Littman DR. Differential requirements for Runx proteins in CD4 repression and epigenetic silencing during T lymphocyte development. Cell. 2002 Nov 27;111(5):621-33
    97.Woolf E, Xiao C, Fainaru O, Lotem J, Rosen D, Negreanu V, Bernstein Y, Goldenberg D, Brenner O, Berke G, Levanon D, Groner Y. Runx3 and Runx1 are required for CD8 T cell development during thymopoiesis. Proc Natl Acad Sci U S A. 2003 Jun 24;100(13):7731-6.
    98.Ehlers M, Laule-Kilian K, Petter M, Aldrian CJ, Grueter B, Wurch A, Yoshida N, Watanabe T, Satake M, Steimle V. Morpholino antisense oligonucleotide-mediated gene knockdown during thymocyte development reveals role for Runx3 transcription factor in CD4 silencing during development of CD4-/CD8+ thymocytes. J Immunol. 2003 Oct 1;171(7):3594-604.
    99.Telfer JC, Hedblom EE, Anderson MK, Laurent MN, Rothenberg EV. Localization of the domains in runx transcription factors required for the repression of CD4 in thymocytes. J Immunol. 2004 Apr 1;172(7):4359-70.
    100. Meng-Jiao Shi and Janet Stavnezer. CBF 3 (AML2) Is Induced by TGF-?1 to Bind and Activate the Mouse Germline Ig Promoter. The Journal of Immunology, 1998, 161: 6751-6760.
    101. Park SR, Lee EK, Kim BC, Kim PH. p300 cooperates with Smad3/4 and Runx3 in TGFbeta1-induced IgA isotype expression. Eur J Immunol. 2003 Dec;33(12):3386-92.
    102. Dominguez-Soto A, Relloso M, Vega MA, Corbi AL, Puig-Kroger A. RUNX3 regulates the activity of the CD11a and CD49d integrin gene promoters. Immunobiology. 2005;210(2-4):133-9.
    103. Puig-Kroger A, Sanchez-Elsner T, Ruiz N, Andreu EJ, Prosper F, Jensen UB, Gil J, Erickson P, Drabkin H, Groner Y, Corbi AL. RUNX/AML and C/EBP factors regulate CD11a integrin expression in myeloid cells through overlapping regulatory elements.Blood. 2003 Nov 1;102(9):3252-61.
    104. Miyoshi, H.; Shimizu, K.; Kozu, T.; Maseki, N.; Kaneko, Y.; Ohki, M. : t(8;21) breakpoints on chromosome 21 in acute myeloid leukemia are clustered within a limited region of a single gene, AML1. Proc. Nat. Acad. Sci. 88: 10431-10434, 1991
    105. Wang J, Hoshino T, Redner RL, Kajigaya S, Liu JM. ETO, fusion partner in t(8;21) acute myeloid leukemia, represses transcription by interaction with the human N-CoR/mSin3/ HDAC1 complex. Proc Natl Acad Sci USA 1998;95:10860_/5.
    106. Lutterbach B, Westendorf JJ, Linggi B, et al. ETO, a target of t(8;21) in acute leukemia, interacts with the N-CoR and mSin3 corepressors. Mol Cell Biol 1998;18:7176_/84.
    107. Gelmetti V, Zhang J, Fanelli M, Minucci S, Pelicci PG, Lazar MA. Aberrant recruitment of the nuclear receptor corepressorhistone deacetylase complex by the acute myeloid leukemia fusion partner ETO. Mol Cell Biol 1998;18:7185_/91.
    108. Kitabayashi I, Ida K, Morohoshi F, et al. The AML1-MTG8 leukemic fusion protein forms a complex with a novel member of the MTG8 (ETO/CDR) family, MTGR1. Mol Cell Biol 1998;18:846_/58.
    109. Shikami M, Miwa H, Nishii K, et al. Low BCL-2 expression in acute leukemia with t(8;21) chromosomal abnormality. Leukemia 1999;13:358_/68.
    110. Burel SA, Harakawa N, Zhou L, Pabst T, Tenen DG, Zhang DE. Dichotomy of AML1-ETO functions: growth arrest versus block of differentiation. Mol Cell Biol 2001;21:5577_/90
    111. Yergeau DA, Hetherington CJ, Wang Q, et al. Embryonic lethality and impairment of haematopoiesis in mice heterozygous for an AML-ETO fusion gene. Nat Genet 1997;15:303_/6.
    112. Okuda T, Cai Z, Yang S, et al. Expression of a knocked-in AML1-ETO leukemia gene inhibits the establishment of normal definitive hematopoiesis and directly generates dysplastic hematopoietic progenitors. Blood 1998;91:3134_/43.
    113. Shimada H, Ichikawa H, Nakamura S, et al. Analysis of genes under the downstream control of the t(8;21) fusion proteinAML1-MTG8: overexpression of the TIS11b (ERF-1, cMG1) gene induces myeloid cell proliferation in response to G-CSF. Blood 2000;96:655_/63.
    114. Hiebert SW, Sun W, Davis, et al. The t(12;21) translocation converts AML1B from an activator to a repressor of transcription. Mol Cell Biol 1996;16:1349_/55.
    115. Uchida H, Downing JR, Miyazaki Y, Frank R, Zhang J, Nimer SD. Three distinct domains in TEL-AML1 are required for transcriptional repression of the IL-3 promoter. Oncogene 1999;18:1015_/22.
    116. Guidez F, Petrie K, Ford AM, et al. Recruitment of the nuclear receptor corepressor N-CoR by the TEL moiety of the childhood leukemia-associated TEL-AML1 oncoprotein. Blood 2000;96:2557_/61.
    117. Wang L, Hiebert SW. TEL contacts multiple co-repressors and specifically associates with histone deacetylase-3. Oncogene 2001;20:3716_/25.
    118. Kim CA, Phillips ML, Kim W, et al. Polimerization of the SAM domain of TEL in leukemogenesis and transcriptional repression. EMBO J 2001;20:4173_/82.
    119. Tanaka T, Mitani K, Kurokawa M, et al. Dual functions of the AML1/Evi-1 chimeric protein in the mechanism of leukemogenesis in t(3;21) leukemias. Mol Cell Biol 1995;15:2383_/92.
    120. Zent CS, Mathieu C, Claxton DF, et al. The chimeric genes AML1/MDS1 and AML1/EAP inhibit AML1B activation at the CSF1R promoter, but only AML1/MDS1 has tumor-promoting properties. Proc Natl Acad Sci USA 1996;93:1044_/8.
    121. Kurokawa M, Mitani K, Irie K, et al. The oncoprotein Evi-1 represses TGF-b signaling by inhibiting Smad3. Nature 1998;394:92_/6.
    122. Norio Asou The role of a Runt domain transcription factor AML1/RUNX1 in leukemogenesis and its clinical implications Critical Reviews in Oncology/Hematology 45 (2003) 129_/150
    123. Osato M, Asou N, Abdalla E, et al. Biallelic and heterozygous point mutations in the runt domain of the AML1/PEBP2aB gene associated with myeloblastic leukemias. Blood 1999;93:1817_/24.
    124. Preudhomme C, Warot-Loze D, Roumier C, et al. High incidence of biallelic point mutations in the Runt domain of the AML1/PEBP2aB gene in M0 acute myeloid leukemia and in myeloid malignancies with acquired trisomy 21. Blood 2000;96:862_/2869.
    125. Britos-Bray M, Ramirez M, Cao W, et al. CBFb-SMMHC, expressed in M4eo acute myeloid leukemia,reduces p53 induction and slows apoptosis in hematopoietic cells exposed to DNAdamaging agents. Blood 1998;92:4344_/52.
    126. Dang J, Inukai T, Kurosawa H, et al. The E2A-HLF oncoprotein activates Grouch-related genes and suppresses Runx1. Mol Cell Biol 2001;21:5935_/45.
    127. Lou J, Cao W, Bernardin F, Ayyanathan K, Rauscher FJ III, Friedman AD. Exogenous cdk4 overcomes reduced cdk4 RNA and inhibition of G1 progression in hematopoietic cells expressing a dominant-negative CBF*/a model for overcoming inhibition of proliferation by CBF oncoproteins. Oncogene 2000;19:2695_/703.
    128. Strom DK, Nip J, Westendorf JJ, et al. Expression of the AML-1 oncogene shortens the G1 phase of the cell cycle. J Biol Chem 2000;275:3438_/45.
    129. Friedman AD, Yang Y, Wang W, Cheng L, Civin CI. Acceleration of G1 cooperates with CBFb-SMMHC to induce acute leukemia in mice. Blood 2001;98:93a_/94a.
    130. Song WJ, Sullivan MG, Legare RD, et al. Haploinsufficiency of CBFA2 causes familial thrombocytopenia with propensity to develop acute myelogenous leukaemia. Nat Genet 1999;23:166_/75.
    131. Michaud J, Feng W, Osato M et al. In vitro analysis of known and novel RUNX1 mutations in dominant familial platelet disorder with predisposition to acute myelogenous leukemia (FPD/AML): implications for mechanisms of pathogenesis. Blood 2002;99
    132. Silva FP, Morolli B, Storlazzi CT, Anelli L, Wessels H, Bezrookove V, Kluin-Nelemans HC, Giphart-Gassler M. Identification of RUNX1/AML1 as a classical tumor suppressor gene. Oncogene. 2003 Jan 30;22(4):538-47
    133. Elagib KE, Racke FK, Mogass M, Khetawat R, Delehanty LL, Goldfarb AN. RUNX1 and GATA-1 coexpression and cooperation in megakaryocytic differentiation. Blood. 2003 Jun 1;101(11):4333-41. Epub 2003 Feb 06.
    134. Fainaru, O., Woolf, E., Lotem, J., Yarmus, M., Brenner, O., Goldenberg, D., Negreanu, V., Bernstein, Y., Levanon, D., Jung, S. & Groner, Y. (2004) EMBO J. 23, 969–979.
    135. Ori Brenner, Ditsa Levanon, Varda Negreanu, Olga Golubkov, Ofer Fainaru, Eilon Woolf, Yoram Groner., Loss of Runx3 function in leukocytes is associated with spontaneously developed colitis and gastric mucosal hyperplasia , PNAS, 2004, 101(9) 16016–16021
    136. Parkes, M. & Jewell, D. (2001) Expert Rev. Mol. Med. 2001, 1–18.
    137. Cho, J. H., Nicolae, D. L., Gold, L. H., Fields, C. T., LaBuda, M. C., Rohal, P. M., Pickles, M. R., Qin, L., Fu, Y., Mann, J. S., et al. (1998) Proc. Natl. Acad. Sci. USA 95, 7502–7507.
    138. Cho, J. H., Nicolae, D. L., Ramos, R., Fields, C. T., Rabenau, K., Corradino, S., Brant, S. R., Espinosa, R., LeBeau, M., Hanauer, S. B., et al. (2000) Hum. Mol. Genet. 9, 1425–1432.
    139. Prokunina L, Castillejo-Lopez C, Oberg F, Gunnarsson I, Berg L, Magnusson V, Brookes AJ, Tentler D, Kristjansdottir H, Grondal G, et al.: A regulatory polymorphism in PDCD1 is associated with susceptibility to systemic lupus erythematosus in humans. Nat Genet 2002, 32:666-669.
    140. John S, Eyre S, Myerscough A, Barrett J, Silman A, Ollier W, Worthington J: Linkage and associationanalysis of candidate genes in rheumatoid arthritis. J Rheumatol 2001, 28:1752- 1755.
    141. Rioux JD, Daly MJ, Silverberg MS, Lindblad K, Steinhart H, Cohen Z, Delmonte T, Kocher K, Miller K, Guschwan S, et al.: Genetic variation in the 5q31 cytokine gene cluster confers susceptibility to Crohn disease. Nat Genet 2001, 29:223-228.
    142. Tokuhiro S, Yamada R, Chang X, Suzuki A, Kochi Y, Sawada T, Suzuki M, Nagasaki M, Ohtsuki M, Ono M, et al.: An intronic SNP in a RUNX1 binding site of SLC22A4, encoding an organic cation transporter, is associated with rheumatoid arthritis. Nat Genet 2003, 35:341-348
    143. Helms C, Cao L, Krueger JG, Wijsman EM, Chamian F, Gordon D, Heffernan M, Daw JA, Robarge J, Ott J, et al.: A putative RUNX1 binding site variant between SLC9A3R1 and NAT9 is associated with susceptibility to psoriasis. Nat Genet 2003.
    144. Mundlos, S.; Otto, F.; Mundlos, C.; Mulliken, J. B.; Aylsworth, A. S.; Albright, S.; Lindhout, D.; Cole, W. G.; Henn, W.; Knoll, J. H. M.; Owen, M. J.; Mertelsmann, R.; Zabel, B. U.; Olsen, B. R. : Mutations involving the transcription factor CBFA1 cause cleidocranial dysplasia. Cell 89: 773-779, 1997
    145. Quack, I.; Vonderstrass, B.; Stock, M.; Aylsworth, A. S.; Becker, A.; Brueton, L.; Lee, P. J.; Majewski, F.; Mulliken, J. B.; Suri, M.; Zenker, M.; Mundlos, S.; Otto, F. : Mutation analysis of core binding factor A1 in patients with cleidocranial dysplasia. Am. J. Hum. Genet. 65: 1268-1278, 1999.
    146. Bergwitz, C.; Prochnau, A.; Mayr, B.; Kramer, F.-J.; Rittierodt, M.; Berten, H.-L.; Hausamen, J.-E.; Brabant, G. Identification of novel CBFA1/RUNX2 mutations causing cleidocranial dysplasia. J. Inherit. Metab. Dis. 24: 648-656, 2001.
    147. Otto, F.; Kanegane, H.; Mundlos, S. : Mutations in the RUNX2 gene in patients with cleidocranial dysplasia. Hum. Mutat. 19: 209-216, 2002.
    148. Zhou, G.; Chen, Y.; Zhou, L.; Thirunavukkarasu, K.; Hecht, J.; Chitayat, D.; Gelb, B. D.; Pirinen, S.; Berry, S. A.; Greenberg, C. R.; Karsenty, G.; Lee, B. CBFA1 mutation analysis and functional correlation with phenotypic variability in cleidocranial dysplasia. Hum. Molec. Genet. 8: 2311-2316, 1999.
    149. Yoshida, T.; Kanegane, H.; Osato, M.; Yanagida, M.; Miyawaki, T.; Ito, Y.; Shigesada, K. Functional analysis of RUNX2 mutations in Japanese patients with cleidocranial dysplasia demonstrates novel genotype-phenotype correlations. Am. J. Hum. Genet. 71: 724-738, 2002. Note: Erratum: Am. J. Hum. Genet. 72: 780 only, 2003.
    150. Tamura G. Promoter methylation status of tumor suppressor and tumor-related genes in neoplastic and non-neoplastic gastric epithelia. Histol Histopathol. 2004 Jan;19(1):221-8.
    151. Waki T, Tamura G, Sato M, Terashima M, Nishizuka S, Motoyama T. Promoter methylation status of DAP-kinase and RUNX3 genes in neoplastic and non-neoplastic gastric epithelia. Cancer Sci. 2003 Apr;94(4):360-4.
    152. Nakase Y, Sakakura C, Miyagawa K, Kin S, Fukuda K, Yanagisawa A, Koide K, Morofuji N, Hosokawa Y, Shimomura K, Katsura K, Hagiwara A, Yamagishi H, Ito K, Ito Y. Frequent loss of RUNX3 gene expression in remnant stomach cancer and adjacent mucosa with special reference to topography Br J Cancer. 2005 Feb 14;92(3):562-9
    153. Kang S, Kim JW, Kang GH, Park NH, Song YS, Kang SB, Lee HP. Polymorphism in folate- andmethionine-metabolizing enzyme and aberrant CpG island hypermethylation in uterine cervical cancer. Gynecol Oncol. 2005 Jan;96(1):173-80
    154. 肖文华, 刘为纹 肝细胞癌RUNX3 基因甲基化与杂合缺失的分析及意义 中华肝脏病杂志2004,Vol 12(4) 227-23
    155. Takahashi T, Shivapurkar N, Riquelme E, Shigematsu H, Reddy J, Suzuki M, Miyajima K, Zhou X, Bekele BN, Gazdar AF, Wistuba II. Aberrant promoter hypermethylation of multiple genes in gallbladder carcinoma and chronic cholecystitis. Clin Cancer Res. 2004 Sep 15;10(18 Pt 1):6126-33
    156. Sakakura C, Hagiwara A, Miyagawa K, Nakashima S, Yoshikawa T, Kin S, Nakase Y, Ito K, Yamagishi H, Yazumi S, Chiba T, Ito Y. Frequent downregulation of the runt domain transcription factors RUNX1, RUNX3 and their cofactor CBFB in gastric cancer. Int J Cancer. 2005 Jan 10;113(2):221-8
    157. Kristie L Durst, Scott W Hiebert. Role of RUNX family members in transcriptional repression and gene silencing. Oncongene. 2004 23,4220-2224
    158. Djuretic IM, Levanon D, Negreanu V, Groner Y, Rao A, Ansel KM.:,Transcription factors T-bet and Runx3 cooperate to activate Ifng and silence Il4 in T helper type 1 cells Nat Immunol. 2007 Feb;8(2):145-53.
    159. Yano T, Ito K, Fukamachi H, Chi XZ, Wee HJ, Inoue K, Ida H, Bouillet P, Strasser A, Bae SC, Ito Y. The RUNX3 tumor suppressor upregulates Bim in gastric epithelial cells undergoing transforming growth factor beta-induced apoptosis. Mol Cell Biol. 2006 Jun;26(12):4474-88
    160. Kim WY, Sieweke M, Ogawa E, et al. Mutual activation of Ets-1 and AML1 DNA binding by direct interaction of their autoinhibitory domains. EMBO J 1999;18:1609_/20.
    161. Kitabayashi I, Yokoyama A, Shimizu K, Ohki M. Interaction and functional cooperation of the leukemia-associated factors AML1 and p300 in myeloid cell differentiation. EMBO J 1998;17:2994_/3004.
    162. Ogryzko VV, Schiltz RL, Russanova V, Howard BH, Nakatani Y. The transcriptional coactivators p300 and CBP are histone acetyltransferases. Cell 1996;87:953_/9.
    163. Bannister AJ, Kouzarides T. The CBP co-activator is a histone acetyltransferase. Nature 1996;384:641_/3.
    164. Chen H, Lin RJ, Schiltz RL, et al. Nuclear receptor coactivator ACTR is a novel histone acetyltransferase and form a multimeric activation complex with P/CAF and CBP/p300. Cell 1997;90:569_/80.
    165. Yasui W, Oue N, Aung PP, Matsumura S, Shutoh M, Nakayama H. Molecular-pathological prognostic factors of gastric cancer: a review. Gastric Cancer. 2005;8:86-94
    166. Yonemura Y, Ninomiya I, Ohoyama S, Kimura H, Yamaguchi A, Fushida S, Kosaka T, Miwa K, Miyazaki I, Endou Y. Expression of c-erbB-2 oncoprotein in gastric carcinoma. Immunoreactivity for c-erbB-2 protein is an independent indicator of poor short-term prognosis in patients with gastric carcinoma.Cancer. 1991;67:2914-2918.
    167. Sanz-Ortega J, Steinberg SM, Moro E, Saez M, Lopez JA, Sierra E, Sanz-Esponera J, Merino MJ. Comparative study of tumor angiogenesis and immunohistochemistry for p53, c-ErbB2, c-myc and EGFras prognostic factors in gastric cancer. Histol Histopathol. 2000;15:455-462.
    168. Kuniyasu H, Yasui W, Yokozaki H, Kitadai Y, Tahara E. Aberrant expression of c-met mRNA in human gastric carcinomas. Int J Cancer. 1993;55:72-75.
    169. Han SU, Lee HY, Lee JH, Kim WH, Nam H, Kim H, Cho YK, Kim MW, Lee KU. Modulation of E-cadherin by hepatocyte growth factor induces aggressiveness of gastric carcinoma. Ann Surg. 2005;242:676-683.
    170. Yasui W, Naka K, Suzuki T, Fujimoto J, Hayashi K, Matsutani N, Yokozaki H, Tahara E. Expression of p27Kip1, cyclin E and E2F-1 in primary and metastatic tumors of gastric carcinoma. Oncol Rep. 1999;6:983-987.
    171. Feakins RM, Nickols CD, Bidd H, Walton SJ. Abnormal expression of pRb, p16, and cyclin D1 in gastric adenocarcinoma and its lymph node metastases: relationship with pathological features and survival. Hum Pathol. 2003;34:1276-1282.
    172. Chen W, Salto-Tellez M, Palanisamy N, Ganesan K, Hou Q, Tan LK, Sii LH, Ito K, Tan B, Wu J, Tay A, Tan KC, Ang E, Tan BK, Tan PH, Ito Y, Tan P. Targets of genome copy number reduction in primary breast cancers identified by integrative genomics. Genes Chromosomes Cancer. 2007;46:288-301.
    173. Sakakura C, Hasegawa K, Miyagawa K, Nakashima S, Yoshikawa T, Kin S, Nakase Y, Yazumi S, Yamagishi H, Okanoue T, Chiba T, Hagiwara A. Possible involvement of RUNX3 silencing in the peritoneal metastases of gastric cancers. Clin Cancer Res. 2005;11:6479-6488.
    174. Sun L, Vitolo M, Passaniti A. Runt-related gene 2 in endothelial cells: inducible expression and specific regulation of cell migration and invasion.Cancer Res. 2001 Jul 1;61(13):4994-5001
    175. Reginato MJ,Mills KR,Paulus JK,et al. Nat Cell Biol,2003,5(8):733- 740.
    176. Zhu Z,Sanchez- Sweatman O,Huang X,et al. Cancer Res,2001,61(4):1707- 1716.
    177. 雷群英,王丽影,戴振宇,等. 生物化学与生物物理学报,2002,34(4):463- 468.
    178. Frankel A,Rosen K,Filmus J,et al. Cancer Res,2001,61(12):4837- 4841.
    179. Swan EA,Jasser SA,Holsinger FC,et al. Oral Oncol,2003,39(7):648- 655.
    180. Frisch S M, Ruoslahti E. Integrins and anoikis. Curr Opin Cell Biol, 1997, 4 (5): 701~706
    181. Harnois C, Demers M J, Bouchard V, Human intestinal epithelial crypt cell survival and death: complex modulations of Bcl-2 homologs by Fak, PI3-K/Akt-1, MEK/Erk, and p38 signaling pathways. J Cell Physiol, 2004, 246 (2): 209~222
    182. Schlaepfer D D, Broome M A, Hunter T. Fibronectin-stimulated signaling from a focal adhesion kinase-c-Src complex: involvement of the Grb2, p130cas, and Nck adaptor proteins. Mol Cell Biol, 1997, 27 (3): 1702~1713
    183. Zelzer E, Glotzer DJ, Hartmann C, Thomas D, Fukai N, Soker S, Olsen BR. Tissue specific regulation of VEGF expression during bone development requires Cbfa1/Runx2.Mech Dev. 2001 Aug;106(1-2):97-106
    184. Van den Steen PE, Proost P, Wuyts A, Van Damme J, Opdenakker G. Neutrophil gelatinase B potentiates interleukin-8 tenfold by aminoterminal processing, whereas it degrades CTAP-III, PF-4, and GRO-alpha and leaves RANTES and MCP-2 intact. Blood. 2000;96:2673-2681.
    185. Huang S, Mills L, Mian B, Tellez C, McCarty M, Yang XD, Gudas JM, Bar-Eli M. Fully humanized neutralizing antibodies to interleukin-8 (ABX-IL8) inhibit angiogenesis, tumor growth, and metastasis of human melanoma. Am J Pathol. 2002;161:125-134.
    186. Bikfalvi A, Gimenez-Gallego G. The control of angiogenesis and tumor invasion by platelet factor-4 and platelet factor-4-derived molecules. Semin Thromb Hemost. 2004;30:137-144.
    187. Dien J, Amin HM, Chiu N, Wong W, Frantz C, Chiu B, Mackey JR, Lai R. Signal transducers and activators of transcription-3 up-regulates tissue inhibitor of metalloproteinase-1 expression and decreases invasiveness of breast cancer. Am J Pathol. 2006;169:633-642.
    188. Stearns ME, Rhim J, Wang M. Interleukin 10 (IL-10) inhibition of primary human prostate cell-induced angiogenesis: IL-10 stimulation of tissue inhibitor of metalloproteinase-1 and inhibition of matrix metalloproteinase (MMP)-2/MMP-9 secretion. Clin Cancer Res. 1999;5:189-196.
    189. Watanabe M, Takahashi Y, Ohta T, Mai M, Sasaki T, Seiki M. Inhibition of metastasis in human gastric cancer cells transfected with tissue inhibitor of metalloproteinase 1 gene in nude mice. Cancer. 1996;77:1676-1680.
    190. Bertrand-Philippe M, Ruddell RG, Arthur MJ, Thomas J, Mungalsingh N, Mann DA. Regulation of tissue inhibitor of metalloproteinase 1 gene transcription by RUNX1 and RUNX2. J Biol Chem. 2004 Jun 4;279(23):24530-9.
    191. David Chodniewicz, Richard L. Klemke Regulation of integrin-mediated cellular responses through assembly of a CAS/Crk scaffold Biochimica. Biophysica Acta 1692 (2004) 63– 76
    192. Louie Lamorte, Isabelle Royal, Monica Naujokas, Morag Park. Crk Adapter Proteins Promote an Epithelial–Mesenchymal-like Transition and Are Required for HGF-mediated Cell Spreading and Breakdown of Epithelial Adherens Junctions. Molecular Biology of the Cell Vol. 13, 1449–1461, May 2002.
    193. Chi XZ, Yang JO, Lee KY, Ito K, Sakakura C, Li QL, Kim HR, Cha EJ, Lee YH, Kaneda A, Ushijima T, Kim WJ, Ito Y, Bae SC. RUNX3 suppresses gastric epithelial cell growth by inducing p21(WAF1/Cip1) expression in cooperation with transforming growth factor {beta}-activated SMAD. Mol Cell Biol. 2005;25:8097-8107.
    194. Yamamura Y, Lee WL, Inoue K, Ida H, Ito Y. RUNX3 cooperates with FoxO3a to induce apoptosis in gastric cancer cells. J Biol Chem. 2006;281:5267-5276.
    195. Guo C, Ding J, Yao L, Sun L, Lin T, Song Y, Sun L, Fan D. Tumor suppressor gene Runx3 sensitizes gastric cancer cells to chemotherapeutic drugs by downregulating Bcl-2, MDR-1 and MRP-1. Int J Cancer. 2005;116:155-160.
    196. Peng Z, Wei D, Wang L, Tang H, Zhang J, Le X, Jia Z, Li Q, Xie K. RUNX3 inhibits the expression of vascular endothelial growth factor and reduces the angiogenesis, growth, and metastasis of human gastric cancer. Clin Cancer Res. 2006;12:6386-6394.

© 2004-2018 中国地质图书馆版权所有 京ICP备05064691号 京公网安备11010802017129号

地址:北京市海淀区学院路29号 邮编:100083

电话:办公室:(+86 10)66554848;文献借阅、咨询服务、科技查新:66554700