慢性乙型肝炎相关肝癌发生、发展过程中的表观遗传学机制研究
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
原发性肝癌(以下简称肝癌)是世界上最常见的恶性肿瘤之一,特别是东南亚及非洲更是常见。在恶性肿瘤引起的死亡位次中肝癌仅次于胃癌、食道癌,居第三位。从世界范围来看,慢性乙型肝炎是肝癌最重要的病因之一。慢性乙型肝炎的分布区域与肝癌的高发区高度重合。尽管近年来早期诊断、手术切除及药物治疗等方面的发展,肝癌患者的5年生存率仍然很低,全世界范围内每年大约有600000肝癌病人死亡。肝癌预后差,一个很重要的原因就是人们对肝癌发生、发展的具体机制目前仍然所知甚少。已报道的研究多半集中于肝癌组织中单个差异表达基因的功能(多数为编码蛋白的基因)及其参与的信号途径调控。然而在肝癌发生、发展过程中复杂的调控网络改变绝不仅仅局限于此。表观遗传学调控方式在慢性乙型肝炎相关肝癌的发生、发展过程充当了重要的调控角色。功能性非编码RNA及蛋白的翻译后修饰等表观遗传遗传调控方式越来越受到研究者的重视。功能性非编码RNA在基因表达中发挥重要的作用,按照它们的大小可分为长链非编码RNA和短链非编码RNA。长链非编码RNA或在基因簇以至于整个染色体水平发挥顺式调节作用,或改变直接相互作用蛋白分子的分子结构而发挥作用,或影响相关编码蛋白基因的剪切、翻译等过程。短链RNA对基因表达进行调控,其可介导mRNA的降解,诱导染色质结构的改变,决定着细胞的分化命运。
     本研究中采用蛋白质组学、基因表达谱芯片等高通量筛选技术辅以经典分子生物学技术、生物信息学分析等手段对慢性乙型肝炎向肝癌进展的过程以及肝癌发生后非编码RNA的调控功能及蛋白翻译后修饰进行了广泛而深入的研究,取得了如下结果:
     第一、发现与慢性乙型肝炎疾病进展及肝癌预后相关的新分子,可做疾病诊断、预后判断的新指标。近年来,人们一直热衷于诊断新标志物的寻找,在肝癌组织中以及病人外周血中一些新诊断标志物的提出,结合着传统指标可以一定程度上提高肝癌早期诊断的灵敏性和准确性。我们的研究结果显示在肝癌组织中上调表达的长链非编码RNA lncRNA-HEIH与肝癌复发概率密切相关,该长链非编码RNA在肝癌组织中的表达量与病人的术后生存时间显著关联,肝癌组织中lncRNA-HEIH表达量越高则预示着病人的术后生存时间可能越短;同时我们的研究结果也表明在肝癌组织中特异性表达的长链非编码RNA谱可以将癌组织和癌旁组织进行准确判别。因此,检测肝癌组织中长链非编码RNA lncRNA-HEIH的表达量结合现有一些指标或者是联合检测若干个长链非编码RNA的表达情况可用于手术病人预后判断的参考指标,帮助临床医生进行决策。根据雌、雄两性别转基因鼠肝脏蛋白质组学的研究结果,我们发现在男、女两性别的慢性乙型肝炎患者血清中Apo A-I氧化修饰的程度存在着明显的差别,可能与不同性别的患者体内抗氧化应激损伤的能力大小有关,可用于慢性乙型肝炎疾病进展程度的监测指标。
     第二、发现一些新分子或者是新的调控方式参与了慢性乙型肝炎相关肝癌的发生、发展过程,可成为肝癌新治疗策略的作用靶点。肝癌之所以治疗效果差,病人预后不理想,很重要的一个原因就是肝癌具有侵袭性,经常出现远处或者是肝脏内的转移。在哺乳细胞的转录产物中除了众人所熟知的编码蛋白RNA,其实包含了更多的具有信号调节功能的非编码RNA,如现在正研究地如火如荼的miRNA。在我们的研究中利用荧光差异蛋白质组学的方法,发现在肝癌组织中上调表达的miRNA-17-5p可以通过激活MAPK p38途径增加HSP27的磷酸化修饰水平而增强肝癌细胞的迁移能力,且这条信号调控通路在临床肝癌样本中也得到验证,为转移性肝癌的判别和治疗提供了新的研究靶点。近年来的文献报道表明一些长度大于200nt的非编码RNA也具有非常复杂的调控功能。我们的研究表明在肝癌组织中上调表达的长链非编码RNA lncRNA-HEIH可以和EZH2相互作用,引导PRC2复合物在基因组中的定位而调控细胞周期相关基因的表达。以往的大部分研究均在事先建立假设的基础上对某一单个的分子或者信号通路进行研究,缺乏整体、全面的手段对慢性乙型肝炎相关肝癌的发生机制进行研究。近年来各种“组学”技术的出现和发展为这一研究领域提供了新的手段。我们运用蛋白质组学的方法对HBV转基因鼠肝脏组织进行了全面、系统的研究,发现了HBV抗原的表达对宿主细胞影响的两个重要方面:①HBV抗原的表达抑制了宿主细胞的抗氧化系统(如谷胱甘肽过氧化物酶1的下调),增加了宿主细胞的氧化应激损伤(血清中DNA氧化产物8-羟基脱氧鸟苷增多,Apo A-I氧化修饰的改变);②HBV抗原的表达扰乱了宿主细胞的脂质代谢。一些脂肪酸的转运蛋白(如脂肪酸结合蛋白5,酰基辅酶A结合蛋白)及Apo A-I在HBV转基因鼠肝脏中的表达或者是修饰发生了明显的异常,为进一步深入的机制研究奠定了基础。
     总之,本研究应用基因芯片、蛋白质组学等手段对参与慢性乙型肝炎相关肝癌发生、发展过程中的功能性非编码RNA调控机制及蛋白翻译后修饰等表观遗传调控方式进行深入而广泛的探讨,为进一步理解该过程中纷繁复杂的信号通路提供了新的视角,为肝癌的早期诊断及治疗提供了新的研究靶点。
Hepatocellular carcinoma (HCC) is one of the most common human cancer worldwide, particularly in Southeast Asia and Africa. Unfortunately, the 5-year survival rate of HCC patient remains poor and ~600 000 HCC patients die each year despite recent advances in surgical techniques and medical treatment. Although previous studies identified many aberrantly expressed protein-coding genes in HCC, the novel molecular markers that can help to early diagnosis and risk assessment are still urgently needed. It is also paramount important to understand the relationships between clinical symptoms and molecular changes in HCC for developing new diagnosis and treatment of HCC and improving the prognosis of diagnosed patients.
     In this study, we used hybridization-based microarrays and two-dimensional electrophoresis proteomic approach to investigate the roles of protein, miRNAs and long non-coding RNAs (lncRNAs) in HCC. Our research has achieved the following results:
     First, we found some novel prognostic biomarkers of chronic hepatitis B (CHB) or HCC in this study. In recent years, identifying the molecular markers correlating with the survival of cancer patients attracts much attention. For HCC, deregulated expression of both protein-coding genes and miRNAs has been suggested to have considerable potential in predicting the prognosis of HCC patients. Using a microarray platform, we identified a long non-coding RNA (termed lncRNA-HEIH), a novel mRNA-like noncoding RNA, as one of the up-regulated genes in HCC. We found that patients with lncRNA-HEIH high expression tumors had an increasing risk of recurrence and significantly reduced overall postoperative survival. Furthermore, univariate and multivariate analyses revealed that lncRNA-HEIH expression was a powerful independent prognostic factor for both recurrence and survival of HCC patients, which is consistent with the results from cell culture and with the notion that lncRNA-HEIH expression may be used as a novel prognostic biomarker of HCC. Apo A-I expression is down-regulated in male and female HBV-transgenic (HBV-Tg) mouse liver and that there is a disordered expression pattern of Apo A-I isoforms in male HBV-Tg mouse liver. We also verified this finding in CHB patient serum. Furthermore, we identified overoxidized Apo A-I mainly resides in basic isoform (isoform 3). Although it is not clear at present whether the occurrence of these modifications has a causal role or simply reflects secondary epiphenomena, the selectively modified Apo A-I isoforms may be considered to be a pathological hallmark that could extend our knowledge of the molecular pathogenesis of CHB. Development of antibodies that specifically recognize the isoforms of Apo A-I may prove to be useful, in combination with other traditional markers, as a more efficient way to evaluate the prognosis of CHB.
     Second, our study increased understanding the relationships between clinical symptoms and molecular changes in HCC. Our findings suggest that the p38 MAPK pathway plays a crucial role in miR-17-5p-induced phosphorylation of heat shock protein 27 (HSP27) and, as a consequence, phosphorylated HSP27 enhances the migration of HCC cells. Our data highlight an important role of miR-17-5p in the proliferation and migration of HCC cells and support the potential application of miR-17-5p in HCC therapy. We also identified non-overlapping signatures of a small number of lncRNAs that are aberrantly expressed in human HCC compared with paired peritumoral tissues. Then we used real-time PCR to validate five lncRNAs whose expression was altered in HCC compared with paired peritumoral tissues. Using loss-of-function and gain-of-function approaches, we found that lncRNA-HEIH plays a key role in cell cycle regulation. We further demonstrated that lncRNA-HEIH bound to enhancer of zeste homolog 2 (EZH2) and that this interaction was required for the repression of EZH2 target genes. These results reveal insights into the molecular regulation mechanisms of HCC cell cycle regulation and lead us to propose that lncRNAs may serve as key regulatory hubs in cancer biology. The present study also provides new insights into the pathogenesis of hepatitis B virus infection, which might not be obtained by studying a specific individual molecule. In addition to a close connection between glutathione peroxidase 1and oxidative stress, our integrated approach highlighted the fact that fatty acid binding 5, Acyl-CoA binding protein and apo A-I could be the key points of lipid metabolism derangement in HBV-Tg mice. Our study also shed light on the physiological difference between HBV-Tg and wild-type mice, which may pave the way for further use of HBV-Tg mice for the study of pathogenesis of hepatitis B virus infection.
     Overall, our finds have diagnostic and therapeutic implications. Understanding the precise molecular mechanisms in HCC will be critical for exploring these potential new strategies for early diagnosis and therapy of HCC.
引文
[1] V Ambros. microRNAs: tiny regulators with great potential. Cell, 2001, 107(7):823-826.
    [2] Y Hayashita, H Osada, Y Tatematsu, et al. A polycistronic microRNA cluster, miR-17-92, is overexpressed in human lung cancers and enhances cell proliferation. Cancer Res, 2005, 65(21):9628-9632.
    [3] S Zhu, H Wu, F Wu, et al. MicroRNA-21 targets tumor suppressor genes in invasion and metastasis. Cell Res, 2008, 18(3):350-359.
    [4] D Chen, MA Farwell, B Zhang. MicroRNA as a new player in the cell cycle. J Cell Physiol, 2010, 225(2):296-301.
    [5] A Hossain, MT Kuo, GF Saunders. Mir-17-5p regulates breast cancer cell proliferation by inhibiting translation of AIB1 mRNA. Mol Cell Biol, 2006, 26(21):8191-8201.
    [6] M Li, J Li, X Ding, et al. microRNA and cancer. AAPS J, 2010, 12(3):309-317.
    [7] TA Farazi, JI Spitzer, P Morozov, et al. miRNAs in human cancer. J Pathol, 2011, 223(2):102-115.
    [8] M Dews, A Homayouni, D Yu, et al. Augmentation of tumor angiogenesis by a Myc-activated microRNA cluster. Nat Genet, 2006, 38(9):1060-1065.
    [9] J Meister, MH Schmidt. miR-126 and miR-126*: new players in cancer. ScientificWorldJournal, 2010, 10:2090-2100.
    [10] G Di Leva, CM Croce. Roles of small RNAs in tumor formation. Trends Mol Med, 2010, 16(6):257-267.
    [11] S Bai, MW Nasser, B Wang, et al. MicroRNA-122 inhibits tumorigenic properties of hepatocellular carcinoma cells and sensitizes these cells to sorafenib. J Biol Chem, 2009, 284(46):32015-32027.
    [12] S Imbeaud, Y Ladeiro, J Zucman-Rossi. Identification of novel oncogenes and tumor suppressors in hepatocellular carcinoma. Semin Liver Dis, 2010, 30(1):75-86.
    [13] Y Ladeiro, G Couchy, C Balabaud, et al. MicroRNA profiling in hepatocellular tumors is associated with clinical features and oncogene/tumor suppressor genemutations. Hepatology, 2008, 47(6):1955-1963.
    [14] J Burchard, C Zhang, AM Liu, et al. microRNA-122 as a regulator of mitochondrial metabolic gene network in hepatocellular carcinoma. Mol Syst Biol, 2010, 6:402.
    [15] D Iliopoulos, K Drosatos, Y Hiyama, et al. MicroRNA-370 controls the expression of microRNA-122 and Cpt1alpha and affects lipid metabolism. J Lipid Res, 2010, 51(6):1513-1523.
    [16] H Xu, JH He, ZD Xiao, et al. Liver-enriched transcription factors regulate microRNA-122 that targets CUTL1 during liver development. Hepatology, 2010, 52(4):1431-1442.
    [17] E Connolly, M Melegari, P Landgraf, et al. Elevated expression of the miR-17-92 polycistron and miR-21 in hepadnavirus-associated hepatocellular carcinoma contributes to the malignant phenotype. Am J Pathol, 2008, 173(3):856-864.
    [18] Y Lu, JM Thomson, HY Wong, et al. Transgenic over-expression of the microRNA miR-17-92 cluster promotes proliferation and inhibits differentiation of lung epithelial progenitor cells. Dev Biol, 2007, 310(2):442-453.
    [19] L He, JM Thomson, MT Hemann, et al. A microRNA polycistron as a potential human oncogene. Nature, 2005, 435(7043):828-833.
    [20] S Volinia, GA Calin, CG Liu, et al. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc Natl Acad Sci U S A, 2006, 103(7):2257-2261.
    [21] H Kutay, S Bai, J Datta, et al. Downregulation of miR-122 in the rodent and human hepatocellular carcinomas. J Cell Biochem, 2006, 99(3):671-678.
    [22] KA O'Donnell, EA Wentzel, KI Zeller, et al. c-Myc-regulated microRNAs modulate E2F1 expression. Nature, 2005, 435(7043):839-843.
    [23] W Kolch, A Pitt. Functional proteomics to dissect tyrosine kinase signalling pathways in cancer. Nat Rev Cancer, 2010, 10(9):618-629.
    [24] DK Nomura, MM Dix, BF Cravatt. Activity-based protein profiling for biochemical pathway discovery in cancer. Nat Rev Cancer, 2010, 10(9):630-638.
    [25] D Baek, J Villen, C Shin, et al. The impact of microRNAs on protein output. Nature, 2008, 455(7209):64-71.
    [26] M Selbach, B Schwanhausser, N Thierfelder, et al. Widespread changes in protein synthesis induced by microRNAs. Nature, 2008, 455(7209):58-63.
    [27] G Van den Bergh, L Arckens. Fluorescent two-dimensional difference gel electrophoresis unveils the potential of gel-based proteomics. Curr Opin Biotechnol, 2004, 15(1):38-43.
    [28] JX Yan, AT Devenish, R Wait, et al. Fluorescence two-dimensional difference gel electrophoresis and mass spectrometry based proteomic analysis of Escherichia coli. Proteomics, 2002, 2(12):1682-1698.
    [29] L Xu, S Chen, RC Bergan. MAPKAPK2 and HSP27 are downstream effectors of p38 MAP kinase-mediated matrix metalloproteinase type 2 activation and cell invasion in human prostate cancer. Oncogene, 2006, 25(21):2987-2998.
    [30] J Landry, J Huot. Modulation of actin dynamics during stress and physiological stimulation by a signaling pathway involving p38 MAP kinase and heat-shock protein 27. Biochem Cell Biol, 1995, 73(9-10):703-707.
    [31] T Hershko, K Korotayev, S Polager, et al. E2F1 modulates p38 MAPK phosphorylation via transcriptional regulation of ASK1 and Wip1. J Biol Chem, 2006, 281(42):31309-31316.
    [32] DR Boutz, P Collins, U Suresh, et al. A two-tiered approach identifies a network of cancer and liver diseases related genes regulated by miR-122. J Biol Chem, 2011.
    [33] M Jovanovic, L Reiter, P Picotti, et al. A quantitative targeted proteomics approach to validate predicted microRNA targets in C. elegans. Nat Methods, 2010, 7(10):837-842.
    [34] GR Yan, SH Xu, ZL Tan, et al. Global identification of miR-373-regulated genes in breast cancer by quantitative proteomics. Proteomics, 2011, 11(5):912-920.
    [35] SH Kang, KW Kang, KH Kim, et al. Upregulated HSP27 in human breast cancer cells reduces Herceptin susceptibility by increasing Her2 protein stability. BMC Cancer, 2008, 8:286.
    [36] WA Golembieski, SL Thomas, CR Schultz, et al. HSP27 mediates SPARC-induced changes in glioma morphology, migration, and invasion. Glia, 2008, 56(10):1061-1075.
    [37] HJ Arts, H Hollema, W Lemstra, et al. Heat-shock-protein-27 (hsp27) expression in ovarian carcinoma: relation in response to chemotherapy and prognosis. Int J Cancer, 1999, 84(3):234-238.
    [38] KL King, AF Li, GY Chau, et al. Prognostic significance of heat shock protein-27 expression in hepatocellular carcinoma and its relation to histologic grading andsurvival. Cancer, 2000, 88(11):2464-2470.
    [39] JM Luk, CT Lam, AF Siu, et al. Proteomic profiling of hepatocellular carcinoma in Chinese cohort reveals heat-shock proteins (Hsp27, Hsp70, GRP78) up-regulation and their associated prognostic values. Proteomics, 2006, 6(3):1049-1057.
    [40] JT Feng, YK Liu, HY Song, et al. Heat-shock protein 27: a potential biomarker for hepatocellular carcinoma identified by serum proteome analysis. Proteomics, 2005, 5(17):4581-4588.
    [41] J Guay, H Lambert, G Gingras-Breton, et al. Regulation of actin filament dynamics by p38 map kinase-mediated phosphorylation of heat shock protein 27. J Cell Sci, 1997, 110 ( Pt 3):357-368.
    [42] E Chang, KS Heo, CH Woo, et al. MK2 SUMOylation regulates actin filament remodeling and subsequent migration in endothelial cells by inhibiting MK2 kinase and HSP27 phosphorylation. Blood, 2011, 117(8):2527-2537.
    [43] C Zheng, Z Lin, ZJ Zhao, et al. MAPK-activated protein kinase-2 (MK2)-mediated formation and phosphorylation-regulated dissociation of the signal complex consisting of p38, MK2, Akt, and Hsp27. J Biol Chem, 2006, 281(48):37215-37226.
    [44] MC Garcia, DM Ray, B Lackford, et al. Arachidonic acid stimulates cell adhesion through a novel p38 MAPK-RhoA signaling pathway that involves heat shock protein 27. J Biol Chem, 2009, 284(31):20936-20945.
    [45] SM Kwon, SA Kim, JH Yoon, et al. Transforming growth factor beta1-induced heat shock protein 27 activation promotes migration of mouse dental papilla-derived MDPC-23 cells. J Endod, 2010, 36(8):1332-1335.
    [46] Z Zhu, X Xu, Y Yu, et al. Silencing heat shock protein 27 decreases metastatic behavior of human head and neck squamous cell cancer cells in vitro. Mol Pharm, 2010, 7(4):1283-1290.
    [1] F Yang, S Yan, Y He, et al. Expression of hepatitis B virus proteins in transgenic mice alters lipid metabolism and induces oxidative stress in the liver. J Hepatol, 2008, 48(1):12-19.
    [2] F Yang, Y Yin, F Wang, et al. miR-17-5p Promotes migration of human hepatocellular carcinoma cells through the p38 mitogen-activated protein kinase-heat shock protein 27 pathway. Hepatology, 2010, 51(5):1614-1623.
    [3] F Yang, Y Yin, F Wang, et al. An altered pattern of liver apolipoprotein A-I isoforms is implicated in male chronic hepatitis B progression. J Proteome Res, 2010, 9(1):134-143.
    [4] JE Wilusz, H Sunwoo, DL Spector. Long noncoding RNAs: functional surprises from the RNA world. Genes Dev, 2009, 23(13):1494-1504.
    [5] J Zhao, BK Sun, JA Erwin, et al. Polycomb proteins targeted by a short repeat RNA to the mouse X chromosome. Science, 2008, 322(5902):750-756.
    [6] RR Pandey, T Mondal, F Mohammad, et al. Kcnq1ot1 antisense noncoding RNA mediates lineage-specific transcriptional silencing through chromatin-level regulation. Mol Cell, 2008, 32(2):232-246.
    [7] X Wang, S Arai, X Song, et al. Induced ncRNAs allosterically modifyRNA-binding proteins in cis to inhibit transcription. Nature, 2008, 454(7200):126-130.
    [8] M Beltran, I Puig, C Pena, et al. A natural antisense transcript regulates Zeb2/Sip1 gene expression during Snail1-induced epithelial-mesenchymal transition. Genes Dev, 2008, 22(6):756-769.
    [9] MA Faghihi, F Modarresi, AM Khalil, et al. Expression of a noncoding RNA is elevated in Alzheimer's disease and drives rapid feed-forward regulation of beta-secretase. Nat Med, 2008, 14(7):723-730.
    [10] A Garen, X Song. Regulatory roles of tumor-suppressor proteins and noncoding RNA in cancer and normal cell functions. Int J Cancer, 2008, 122(8):1687-1689.
    [11] MB Clark, JS Mattick. Long noncoding RNAs in cell biology. Semin Cell Dev Biol, 2011.
    [12] K Li, Y Blum, A Verma, et al. A noncoding antisense RNA in tie-1 locus regulates tie-1 function in vivo. Blood, 115(1):133-139.
    [13] TR Mercer, ME Dinger, SM Sunkin, et al. Specific expression of long noncoding RNAs in the mouse brain. Proc Natl Acad Sci U S A, 2008, 105(2):716-721.
    [14] KC Pang, ME Dinger, TR Mercer, et al. Genome-wide identification of long noncoding RNAs in CD8+ T cells. J Immunol, 2009, 182(12):7738-7748.
    [15] MC Tsai, O Manor, Y Wan, et al. Long noncoding RNA as modular scaffold of histone modification complexes. Science, 2010, 329(5992):689-693.
    [16] Z Yang, L Zhou, LM Wu, et al. Overexpression of Long Non-coding RNA HOTAIR Predicts Tumor Recurrence in Hepatocellular Carcinoma Patients Following Liver Transplantation. Ann Surg Oncol, 2011.
    [17] K Li, Y Blum, A Verma, et al. A noncoding antisense RNA in tie-1 locus regulates tie-1 function in vivo. Blood, 2010, 115(1):133-139.
    [18] AP Bracken, K Helin. Polycomb group proteins: navigators of lineage pathways led astray in cancer. Nat Rev Cancer, 2009, 9(11):773-784.
    [19] GA Calin, CG Liu, M Ferracin, et al. Ultraconserved regions encoding ncRNAs are altered in human leukemias and carcinomas. Cancer Cell, 2007, 12(3):215-229.
    [20] W Yu, D Gius, P Onyango, et al. Epigenetic silencing of tumour suppressor gene p15 by its antisense RNA. Nature, 2008, 451(7175):202-206.
    [21] R Lin, S Maeda, C Liu, et al. A large noncoding RNA is a marker for murine hepatocellular carcinomas and a spectrum of human carcinomas. Oncogene, 2007,26(6):851-858.
    [22] L Li, T Feng, Y Lian, et al. Role of human noncoding RNAs in the control of tumorigenesis. Proc Natl Acad Sci U S A, 2009, 106(31):12956-12961.
    [23] K Panzitt, MM Tschernatsch, C Guelly, et al. Characterization of HULC, a novel gene with striking up-regulation in hepatocellular carcinoma, as noncoding RNA. Gastroenterology, 2007, 132(1):330-342.
    [24] MA Pujana, JD Han, LM Starita, et al. Network modeling links breast cancer susceptibility and centrosome dysfunction. Nat Genet, 2007, 39(11):1338-1349.
    [25] Q Liao, C Liu, X Yuan, et al. Large-scale prediction of long non-coding RNA functions in a coding-non-coding gene co-expression network. Nucleic Acids Res, 2011.
    [26] RR Nayak, M Kearns, RS Spielman, et al. Coexpression network based on natural variation in human gene expression reveals gene interactions and functions. Genome Res, 2009, 19(11):1953-1962.
    [27] LM Topper, MS Campbell, S Tugendreich, et al. The dephosphorylated form of the anaphase-promoting complex protein Cdc27/Apc3 concentrates on kinetochores and chromosome arms in mitosis. Cell Cycle, 2002, 1(4):282-292.
    [28] Q Zhu, G Wani, J Yao, et al. The ubiquitin-proteasome system regulates p53-mediated transcription at p21waf1 promoter. Oncogene, 2007, 26(29):4199-4208.
    [29] F Martinez-Arribas, D Agudo, M Pollan, et al. Positive correlation between the expression of X-chromosome RBM genes (RBMX, RBM3, RBM10) and the proapoptotic Bax gene in human breast cancer. J Cell Biochem, 2006, 97(6):1275-1282.
    [30] RA Gupta, N Shah, KC Wang, et al. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature, 2010, 464(7291):1071-1076.
    [31] AM Khalil, M Guttman, M Huarte, et al. Many human large intergenic noncoding RNAs associate with chromatin-modifying complexes and affect gene expression. Proc Natl Acad Sci U S A, 2009, 106(28):11667-11672.
    [32] TA Paul, J Bies, D Small, et al. Signatures of polycomb repression and reduced H3K4 trimethylation are associated with p15INK4b DNA methylation in AML. Blood, 2010, 115(15):3098-3108.
    [33] Y Kotake, R Cao, P Viatour, et al. pRB family proteins are required for H3K27trimethylation and Polycomb repression complexes binding to and silencing p16INK4alpha tumor suppressor gene. Genes Dev, 2007, 21(1):49-54.
    [34] JY Yao, L Zhang, X Zhang, et al. H3K27 trimethylation is an early epigenetic event of p16INK4a silencing for regaining tumorigenesis in fusion reprogrammed hepatoma cells. J Biol Chem, 2010, 285(24):18828-18837.
    [35] R Aoki, T Chiba, S Miyagi, et al. The polycomb group gene product Ezh2 regulates proliferation and differentiation of murine hepatic stem/progenitor cells. J Hepatol, 2010, 52(6):854-863.
    [36] CE Burd, WR Jeck, Y Liu, et al. Expression of Linear and Novel Circular Forms of an INK4/ARF-Associated Non-Coding RNA Correlates with Atherosclerosis Risk. PLoS Genet, 2010, 6(12):e1001233.
    [37] X Yang, RK Karuturi, F Sun, et al. CDKN1C (p57) is a direct target of EZH2 and suppressed by multiple epigenetic mechanisms in breast cancer cells. PLoS One, 2009, 4(4):e5011.
    [38] DY But, CL Lai, MF Yuen. Natural history of hepatitis-related hepatocellular carcinoma. World J Gastroenterol, 2008, 14(11):1652-1656.
    [39] K Hao, JM Luk, NP Lee, et al. Predicting prognosis in hepatocellular carcinoma after curative surgery with common clinicopathologic parameters. BMC Cancer, 2009, 9:389.
    [40] RT Poon, IO Ng, ST Fan, et al. Clinicopathologic features of long-term survivors and disease-free survivors after resection of hepatocellular carcinoma: a study of a prospective cohort. J Clin Oncol, 2001, 19(12):3037-3044.
    [41] M Huarte, M Guttman, D Feldser, et al. A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell, 2010, 142(3):409-419.
    [42] D Tian, S Sun, JT Lee. The long noncoding RNA, jpx, is a molecular switch for x chromosome inactivation. Cell, 2010, 143(3):390-403.
    [43] S Souquere, G Beauclair, F Harper, et al. Highly-ordered Spatial Organization of the Structural Long Non-coding NEAT1 RNAs Within Paraspeckle Nuclear Bodies. Mol Biol Cell, 2010.
    [44] J Hou, L Lin, W Zhou, et al. Identification of miRNomes in Human Liver and Hepatocellular Carcinoma Reveals miR-199a/b-3p as Therapeutic Target for Hepatocellular Carcinoma. Cancer Cell, 2011, 19(2):232-243.
    [45] J Ji, J Shi, A Budhu, et al. MicroRNA expression, survival, and response to interferon in liver cancer. N Engl J Med, 2009, 361(15):1437-1447.
    [46] B Lu, H Guo, J Zhao, et al. Increased expression of iASPP, regulated by hepatitis B virus X protein-mediated NF-kappaB activation, in hepatocellular carcinoma. Gastroenterology, 2010, 139(6):2183-2194 e2185.
    [47] J Wang, X Liu, H Wu, et al. CREB up-regulates long non-coding RNA, HULC expression through interaction with microRNA-372 in liver cancer. Nucleic Acids Res, 2010, 38(16):5366-5383.
    [48] IJ Matouk, N DeGroot, S Mezan, et al. The H19 non-coding RNA is essential for human tumor growth. PLoS One, 2007, 2(9):e845.
    [49] T Yoshimizu, A Miroglio, MA Ripoche, et al. The H19 locus acts in vivo as a tumor suppressor. Proc Natl Acad Sci U S A, 2008, 105(34):12417-12422.
    [50] R Cao, Y Zhang. The functions of E(Z)/EZH2-mediated methylation of lysine 27 in histone H3. Curr Opin Genet Dev, 2004, 14(2):155-164.
    [51] MY Cai, ZT Tong, F Zheng, et al. EZH2 protein: a promising immunomarker for the detection of hepatocellular carcinomas in liver needle biopsies. Gut, 2011.
    [52] Q Chang, Y Zhang, KJ Beezhold, et al. Sustained JNK1 activation is associated with altered histone H3 methylations in human liver cancer. J Hepatol, 2009, 50(2):323-333.
    [53] Y Chen, MC Lin, H Yao, et al. Lentivirus-mediated RNA interference targeting enhancer of zeste homolog 2 inhibits hepatocellular carcinoma growth through down-regulation of stathmin. Hepatology, 2007, 46(1):200-208.
    [54] M Sasaki, H Ikeda, K Itatsu, et al. The overexpression of polycomb group proteins Bmi1 and EZH2 is associated with the progression and aggressive biological behavior of hepatocellular carcinoma. Lab Invest, 2008, 88(8):873-882.
    [1] C Neuveut, Y Wei, MA Buendia. Mechanisms of HBV-related hepatocarcinogenesis. J Hepatol, 2010, 52(4):594-604.
    [2] Y Matsuzaki, T Chiba, T Hadama, et al. HBV genome integration and genetic instability in HBsAg-negative and anti-HCV-positive hepatocellular carcinoma in Japan. Cancer Lett, 1997, 119(1):53-61.
    [3] K Koike. Hepatitis B virus X gene is implicated in liver carcinogenesis. Cancer Lett, 2009, 286(1):60-68.
    [4] CM Lee, SN Lu, CS Changchien, et al. Age, gender, and local geographic variations of viral etiology of hepatocellular carcinoma in a hyperendemic area for hepatitis B virus infection. Cancer, 1999, 86(7):1143-1150.
    [5] MF Yuen, Y Tanaka, DY Fong, et al. Independent risk factors and predictive score for the development of hepatocellular carcinoma in chronic hepatitis B. J Hepatol, 2009, 50(1):80-88.
    [6] G Fattovich, N Olivari, M Pasino, et al. Long-term outcome of chronic hepatitis B in Caucasian patients: mortality after 25 years. Gut, 2008, 57(1):84-90.
    [7] AB Rogers, EJ Theve, Y Feng, et al. Hepatocellular carcinoma associated with liver-gender disruption in male mice. Cancer Res, 2007, 67(24):11536-11546.
    [8] Y Wang, F Cui, Y Lv, et al. HBsAg and HBx knocked into the p21 locus causes hepatocellular carcinoma in mice. Hepatology, 2004, 39(2):318-324.
    [9] FK Stevenson, CH Ottensmeier, J Rice. DNA vaccines against cancer come of age. Curr Opin Immunol, 2010, 22(2):264-270.
    [10] JH van den Berg, K Oosterhuis, JH Beijnen, et al. DNA vaccination in oncology: current status, opportunities and perspectives. Curr Clin Pharmacol, 2010, 5(3):218-225.
    [11] DA Li, Y He, YJ Guo, et al. Comparative proteomics analysis to annexin B1 DNA and protein vaccination in mice. Vaccine, 2007, 25(5):932-938.
    [12] XY Zi, YC Yao, HY Zhu, et al. Long-term persistence of hepatitis B surface antigen and antibody induced by DNA-mediated immunization results in liver and kidney lesions in mice. Eur J Immunol, 2006, 36(4):875-886.
    [13] K Schwarz, E Meijerink, DE Speiser, et al. Efficient homologous prime-boost strategies for T cell vaccination based on virus-like particles. Eur J Immunol, 2005, 35(3):816-821.
    [14] S Song, Y Wang, Y Zhang, et al. Augmented induction of CD8+ cytotoxic T-cell response and antitumor effect by DCs pulsed with virus-like particles packaging with CpG. Cancer Lett, 2007, 256(1):90-100.
    [15] FX Ding, F Wang, YM Lu, et al. Multiepitope peptide-loaded virus-like particles as a vaccine against hepatitis B virus-related hepatocellular carcinoma. Hepatology, 2009, 49(5):1492-1502.
    [16] J Ai, Y Tan, W Ying, et al. Proteome analysis of hepatocellular carcinoma by laser capture microdissection. Proteomics, 2006, 6(2):538-546.
    [17] S Aldred, T Sozzi, I Mudway, et al. Alpha tocopherol supplementation elevates plasma apolipoprotein A1 isoforms in normal healthy subjects. Proteomics, 2006, 6(5):1695-1703.
    [18] ZH Beg, JA Stonik, JM Hoeg, et al. Human apolipoprotein A-I. Post-translational modification by covalent phosphorylation. J Biol Chem, 1989, 264(12):6913-6921.
    [19] E Contiero, R Ferrari, GM Vaselli, et al. Apolipoprotein AI isoforms in serum determined by isoelectric focusing and immunoblotting. Electrophoresis, 1997, 18(1):122-126.
    [20] XG Lei. Glutathione peroxidase-1 gene knockout on body antioxidant defense in mice. Biofactors, 2001, 14(1-4):93-99.
    [21] XG Lei, WH Cheng. New roles for an old selenoenzyme: evidence from glutathione peroxidase-1 null and overexpressing mice. J Nutr, 2005, 135(10):2295-2298.
    [22] YH Hsieh, IJ Su, HC Wang, et al. Pre-S mutant surface antigens in chronic hepatitis B virus infection induce oxidative stress and DNA damage. Carcinogenesis, 2004, 25(10):2023-2032.
    [23] T Severi, C Ying, JR Vermeesch, et al. Hepatitis B virus replication causes oxidative stress in HepAD38 liver cells. Mol Cell Biochem, 2006, 290(1-2):79-85.
    [24] J Fernandez-Irigoyen, E Santamaria, L Sesma, et al. Oxidation of specific methionine and tryptophan residues of apolipoprotein A-I in hepatocarcinogenesis. Proteomics, 2005, 5(18):4964-4972.
    [25] H Inoue, I Shimizu, G Lu, et al. Idoxifene and estradiol enhance antiapoptoticactivity through estrogen receptor-beta in cultured rat hepatocytes. Dig Dis Sci, 2003, 48(3):570-580.
    [26] T Omoya, I Shimizu, Y Zhou, et al. Effects of idoxifene and estradiol on NF-kappaB activation in cultured rat hepatocytes undergoing oxidative stress. Liver, 2001, 21(3):183-191.
    [1] Y Ladeiro, G Couchy, C Balabaud, et al. MicroRNA profiling in hepatocellular tumors is associated with clinical features and oncogene/tumor suppressor gene mutations. Hepatology, 2008, 47(6):1955-1963.
    [2] LM Li, ZB Hu, ZX Zhou, et al. Serum microRNA profiles serve as novel biomarkers for HBV infection and diagnosis of HBV-positive hepatocarcinoma. Cancer Res, 2010, 70(23):9798-9807.
    [3] M Li, J Li, X Ding, et al. microRNA and cancer. AAPS J, 2010, 12(3):309-317.
    [4] X Zhang, S Liu, T Hu, et al. Up-regulated microRNA-143 transcribed by nuclear factor kappa B enhances hepatocarcinoma metastasis by repressing fibronectinexpression. Hepatology, 2009, 50(2):490-499.
    [5] S Zhu, H Wu, F Wu, et al. MicroRNA-21 targets tumor suppressor genes in invasion and metastasis. Cell Res, 2008, 18(3):350-359.
    [6] MB Clark, JS Mattick. Long noncoding RNAs in cell biology. Semin Cell Dev Biol, 2011.
    [7] CP Ponting, PL Oliver, W Reik. Evolution and functions of long noncoding RNAs. Cell, 2009, 136(4):629-641.
    [8] IA Qureshi, JS Mattick, MF Mehler. Long non-coding RNAs in nervous system function and disease. Brain Res, 2010, 1338:20-35.
    [9] RR Pandey, T Mondal, F Mohammad, et al. Kcnq1ot1 antisense noncoding RNA mediates lineage-specific transcriptional silencing through chromatin-level regulation. Mol Cell, 2008, 32(2):232-246.
    [10] J Zhao, BK Sun, JA Erwin, et al. Polycomb proteins targeted by a short repeat RNA to the mouse X chromosome. Science, 2008, 322(5902):750-756.
    [11] X Wang, S Arai, X Song, et al. Induced ncRNAs allosterically modify RNA-binding proteins in cis to inhibit transcription. Nature, 2008, 454(7200):126-130.
    [12] M Beltran, I Puig, C Pena, et al. A natural antisense transcript regulates Zeb2/Sip1 gene expression during Snail1-induced epithelial-mesenchymal transition. Genes Dev, 2008, 22(6):756-769.
    [13] MA Faghihi, F Modarresi, AM Khalil, et al. Expression of a noncoding RNA is elevated in Alzheimer's disease and drives rapid feed-forward regulation of beta-secretase. Nat Med, 2008, 14(7):723-730.
    [14] E Vafiadaki, DA Arvanitis, SN Pagakis, et al. The anti-apoptotic protein HAX-1 interacts with SERCA2 and regulates its protein levels to promote cell survival. Mol Biol Cell, 2009, 20(1):306-318.
    [15] PP Amaral, MB Clark, DK Gascoigne, et al. lncRNAdb: a reference database for long noncoding RNAs. Nucleic Acids Res, 2010.
    [16] DP Caley, RC Pink, D Trujillano, et al. Long noncoding RNAs, chromatin, and development. ScientificWorldJournal, 10:90-102.
    [17] R Lin, S Maeda, C Liu, et al. A large noncoding RNA is a marker for murine hepatocellular carcinomas and a spectrum of human carcinomas. Oncogene, 2007, 26(6):851-858.
    [18] M Huarte, M Guttman, D Feldser, et al. A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell, 2010, 142(3):409-419.
    [19] AM Khalil, M Guttman, M Huarte, et al. Many human large intergenic noncoding RNAs associate with chromatin-modifying complexes and affect gene expression. Proc Natl Acad Sci U S A, 2009, 106(28):11667-11672.
    [20] MC Tsai, O Manor, Y Wan, et al. Long noncoding RNA as modular scaffold of histone modification complexes. Science, 329(5992):689-693.
    [21] RA Gupta, N Shah, KC Wang, et al. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature, 2010, 464(7291):1071-1076.
    [22] MC Tsai, O Manor, Y Wan, et al. Long noncoding RNA as modular scaffold of histone modification complexes. Science, 2010, 329(5992):689-693.
    [23] V Tripathi, JD Ellis, Z Shen, et al. The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Mol Cell, 2010, 39(6):925-938.
    [24] X Cai, BR Cullen. The imprinted H19 noncoding RNA is a primary microRNA precursor. RNA, 2007, 13(3):313-316.
    [25] WP Tsang, EK Ng, SS Ng, et al. Oncofetal H19-derived miR-675 regulates tumor suppressor RB in human colorectal cancer. Carcinogenesis, 2010, 31(3):350-358.
    [26] K Panzitt, MM Tschernatsch, C Guelly, et al. Characterization of HULC, a novel gene with striking up-regulation in hepatocellular carcinoma, as noncoding RNA. Gastroenterology, 2007, 132(1):330-342.
    [27] J Wang, X Liu, H Wu, et al. CREB up-regulates long non-coding RNA, HULC expression through interaction with microRNA-372 in liver cancer. Nucleic Acids Res, 2010, 38(16):5366-5383.
    [28] H Kutay, S Bai, J Datta, et al. Downregulation of miR-122 in the rodent and human hepatocellular carcinomas. J Cell Biochem, 2006, 99(3):671-678.
    [29] H Xu, JH He, ZD Xiao, et al. Liver-enriched transcription factors regulate microRNA-122 that targets CUTL1 during liver development. Hepatology, 2010, 52(4):1431-1442.
    [30] M Huarte, M Guttman, D Feldser, et al. A large intergenic noncoding RNA induced by p53 mediates global gene repression in the p53 response. Cell, 142(3):409-419.
    [31] J Ji, J Shi, A Budhu, et al. MicroRNA expression, survival, and response tointerferon in liver cancer. N Engl J Med, 2009, 361(15):1437-1447.
    [32] R Taby, JP Issa. Cancer epigenetics. CA Cancer J Clin, 2010, 60(6):376-392.
    [33] CE Burd, WR Jeck, Y Liu, et al. Expression of Linear and Novel Circular Forms of an INK4/ARF-Associated Non-Coding RNA Correlates with Atherosclerosis Risk. PLoS Genet, 2010, 6(12):e1001233.
    [34] KL Yap, S Li, AM Munoz-Cabello, et al. Molecular interplay of the noncoding RNA ANRIL and methylated histone H3 lysine 27 by polycomb CBX7 in transcriptional silencing of INK4a. Mol Cell, 2010, 38(5):662-674.
    [35] W Yu, D Gius, P Onyango, et al. Epigenetic silencing of tumour suppressor gene p15 by its antisense RNA. Nature, 2008, 451(7175):202-206.
    [36] KV Morris, S Santoso, AM Turner, et al. Bidirectional transcription directs both transcriptional gene activation and suppression in human cells. PLoS Genet, 2008, 4(11):e1000258.
    [37] LY Pang, M Scott, RL Hayward, et al. p21 (WAF1) is component of a positive feedback loop that maintains the p53 transcriptional program. Cell Cycle, 2011, 10(6).
    [38] X Fu, L Ravindranath, N Tran, et al. Regulation of apoptosis by a prostate-specific and prostate cancer-associated noncoding gene, PCGEM1. DNA Cell Biol, 2006, 25(3):135-141.
    [39] GO Ifere, GA Ananaba. Prostate cancer gene expression marker 1 (PCGEM1): a patented prostate- specific non-coding gene and regulator of prostate cancer progression. Recent Pat DNA Gene Seq, 2009, 3(3):151-163.
    [40] G Petrovics, W Zhang, M Makarem, et al. Elevated expression of PCGEM1, a prostate-specific gene with cell growth-promoting function, is associated with high-risk prostate cancer patients. Oncogene, 2004, 23(2):605-611.
    [41] V Srikantan, Z Zou, G Petrovics, et al. PCGEM1, a prostate-specific gene, is overexpressed in prostate cancer. Proc Natl Acad Sci U S A, 2000, 97(22):12216-12221.
    [42] WP Tsang, TW Wong, AH Cheung, et al. Induction of drug resistance and transformation in human cancer cells by the noncoding RNA CUDR. RNA, 2007, 13(6):890-898.
    [43] Z Yang, L Zhou, LM Wu, et al. Overexpression of Long Non-coding RNA HOTAIR Predicts Tumor Recurrence in Hepatocellular Carcinoma PatientsFollowing Liver Transplantation. Ann Surg Oncol, 2011.
    [44] F Wang, X Li, X Xie, et al. UCA1, a non-protein-coding RNA up-regulated in bladder carcinoma and embryo, influencing cell growth and promoting invasion. FEBS Lett, 2008, 582(13):1919-1927.
    [45] XS Wang, Z Zhang, HC Wang, et al. Rapid identification of UCA1 as a very sensitive and specific unique marker for human bladder carcinoma. Clin Cancer Res, 2006, 12(16):4851-4858.
    [46] J Hou, L Lin, W Zhou, et al. Identification of miRNomes in Human Liver and Hepatocellular Carcinoma Reveals miR-199a/b-3p as Therapeutic Target for Hepatocellular Carcinoma. Cancer Cell, 2011, 19(2):232-243.
    [47] IJ Matouk, N DeGroot, S Mezan, et al. The H19 non-coding RNA is essential for human tumor growth. PLoS One, 2007, 2(9):e845.
    [48] A Etheridge, I Lee, L Hood, et al. Extracellular microRNA: a new source of biomarkers. Mutat Res, 2011.
    [49] KZ Qu, K Zhang, H Li, et al. Circulating MicroRNAs as Biomarkers for Hepatocellular Carcinoma. J Clin Gastroenterol, 2011, 45(4):355-360.
    [50] G Wang, L Tam, EK Li, et al. Serum and urinary free microRNA level in patients with systemic lupus erythematosus. Lupus, 2011.

© 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