2012 cancer cell --Chemical Genomics Identifies Small-Molecule MCL1 Repressors

合集下载

2012年诺贝尔生理学或医学奖解读——为细胞重新编程

2012年诺贝尔生理学或医学奖解读——为细胞重新编程

获奖者简介---Shinya Yamanaka

1962年9月4 日出生于日本大阪府东大阪市。1987 年, 在神户大学获得医学学位,开始在国立大阪医院做整形外 科实习医生; 但由于觉得自己缺乏成为一名外科医生的天 才,最终决定转做基础研究。1993年在大阪市立大学医 学院获得博士学位后,Yamanaka于1993 ~ 1996 年在美 国加州大学旧金山分校的Gladstone 心血管疾病研究所从 事博士后工作,随后就职于大阪市立大学医学院及日本奈 良科技研究院。目前是日本京都大学诱导多能干细胞研究 和应用中心主任、Gladstone 研究所高级研究员、国际干 细胞联合会的现任理事长。
iPS的临床应用

目前已从多种退行性疾病的患者体内获得 iPS 细胞,包括肌萎缩性侧索硬化症 ( amyotrophic lateral sclerosis,ALS) 、脊髓 肌肉萎缩症( spinal muscular atrophy, SMA) 、帕金森病( Parkinson's disease, PD) 、脊髓损伤、黄斑变性等,用以进行细 胞替代疗法的研究。
回英国,就职于牛津大学动物学系,直至1972 年进入英国剑桥
大学分子生物学系及后来的动物学系工作; 曾任剑桥麦格达伦学 院院长。1989 年他参与创办旨在资助细胞生物学和癌症研究
的剑桥( 后成为英国) 维康信托研究所,并任主席至2001 年。
2004 年该研究所更名为Gurdon 研究所; 目前79 岁的Gurdon 仍 工作于Gurdon研究所。
主要的科学贡献 ----体细胞核移植

1935 年诺贝尔生理学或医学奖获得者,德国科学 家Han Spemann 就曾在1938 年提出细胞核移植 的奇妙设想,即把分化细胞的细胞核转移到一个

国外天然产物化学成分实物库及数据库建设概况

国外天然产物化学成分实物库及数据库建设概况

国外天然产物化学成分实物库及数据库建设概况天然产物是新药发现的重要源泉,天然产物化学成分实物库和数据库的建设对天然产物的研究与开发具有重要意义。

目前国外建设的小分子化合物库多为合成或组合合成分子,天然产物实物获取较困难。

在信息数据库建设方面由于使用标准不同,信息不统一,开发规范、实用、智能型、综合型的大规模天然产物数据库还存在一定困难。

该文就目前国外可以公开查询到的有关天然产物的实物库及数据库建设情况进行了概述和分析,以期对天然产物研究与开发,特别是天然产物化学成分实物库和数据库的建设提供参考。

标签:天然产物;实物库;数据库2014-09-241实物库建设概况国外很多制药公司和研究机构都建有自己的化合物库,如美国辉瑞、德国拜耳、瑞士诺华、英国葛兰素史克、美国国立癌症研究所等,都在以多种方式大力扩建自己的化学成分库,占领新药研发的源头——分子资源,但多不公开共享,其库存成分多为合成或组合合成分子,分子结构多样性较少,其天然分子多从国外如中国大量收购或合作收集。

一方面,由于植物、微生物等天然产物的化学结构独特,一些人工很难合成的化合物在生物体内通过酶的作用就容易形成;另一方面,生物在不断进化的过程中其天然成分大多具有某些生物活性,从中寻找先导化合物比人工合成成功率更高。

因此天然产物备受世界各国医药研发者的青睐。

目前,美国、欧盟、日本、韩国等一些国家和地区的许多医药研究机构都在加紧进行有关天然植物药的研发工作。

不少大型制药公司正尽力把大量的植物物种送入实验室进行大规模筛选,以便从中发现任何可能的生物学功效。

如美国国立癌症研究所通过与世界各地的高校或研究所建立合作关系,收集大量的植物、海洋生物、真菌等样品,建立了其天然产物筛选库,据报道,到2009年末已收集并制备了230 000多个样品<sup>[3]</sup>。

虽然国外目前专门从事天然产物实物库建设的单位不多,但由于在世界各地都有不少从事天然产物的研究和开发的研究单位和公司,且其大多为微生物和海洋天然产物,表1列举了一些国外建有天然产物实物库或可提供天然产物的研究单位或公司。

The Splicing Factor RBM4 Controls

The Splicing Factor RBM4 Controls

Cancer CellArticleThe Splicing Factor RBM4ControlsApoptosis,Proliferation,and Migrationto Suppress Tumor ProgressionYang Wang,1,2,*Dan Chen,3Haili Qian,4Yihsuan S.Tsai,2Shujuan Shao,5Quentin Liu,1Daniel Dominguez,2and Zefeng Wang2,*1Institute of Cancer Stem Cell,Second Affiliated Hospital,Cancer Center,Dalian Medical University,Dalian116044,China2Department of Pharmacology and Lineberger Comprehensive Cancer Center,University of North Carolina,Chapel Hill,NC27599,USA3Department of Pathology,First Affiliated Hospital.Dalian Medical University,Dalian116001,China4State Key Laboratory of Molecular Oncology,Peking Union Medical College and Chinese Academy of Medical Sciences,Beijing100021, China5Key Laboratory of Proteomics of Liaoning Province,Dalian Medical University,Dalian116044,China*Correspondence:yangwang@(Y.W.),zefeng@(Z.W.)/10.1016/r.2014.07.010SUMMARYSplicing dysregulation is one of the molecular hallmarks of cancer.However,the underlying molecular mech-anisms remain poorly defined.Here we report that the splicing factor RBM4suppresses proliferation and migration of various cancer cells by specifically controlling cancer-related splicing.Particularly,RBM4reg-ulates Bcl-x splicing to induce apoptosis,and coexpression of Bcl-xL partially reverses the RBM4-mediated tumor suppression.Moreover,RBM4antagonizes an oncogenic splicing factor,SRSF1,to inhibit mTOR acti-vation.Strikingly,RBM4expression is decreased dramatically in cancer patients,and the RBM4level corre-lates positively with improved survival.In addition to providing mechanistic insights of cancer-related splicing dysregulation,this study establishes RBM4as a tumor suppressor with therapeutic potential and clinical values as a prognostic factor.INTRODUCTIONAs one of the most prevalent mechanisms of gene regulation, alternative splicing(AS)plays a vital role in the intricate regula-tion of protein function,and splicing dysregulation is closely associated with human cancers(David and Manley,2010;Ol-tean and Bates,2013;Venables,2006).Accumulating evidence suggests that aberrant AS elicits control over major hallmarks of cancer,including apoptosis(Schwerk and Schulze-Osthoff, 2005),epithelial-mesenchymal transition(Warzecha et al., 2010),and tumor invasion and metastasis(Ghigna et al., 2008).The‘‘cancerous’’splicing variants of specific genes can serve as molecular markers of cancer(e.g.,CD44and WT1)(Venables et al.,2008)or directly mediate cancer patho-genesis(e.g.BRCA1and p53)(Venables,2006).However,knowledge of the mechanistic details underlying deregulated splicing in cancer is still limited.AS is generally regulated by multiple cis-elements that recruit splicing factors to affect adjacent splice sites(ss)via various mechanisms(Matera and Wang,2014;Matlin et al.,2005; Wang and Burge,2008).Common splicing factors include serine/arginine-rich(SR)proteins that promote splicing by bind-ing to exons but inhibit splicing by binding to introns(Erkelenz et al.,2013;Wang et al.,2013)and heterogeneous nuclear ribo-nucleoproteins(hnRNPs)that positively or negatively control splicing in different pre-mRNA regions(Wang et al.,2012).The expression level,localization,and activity of splicing factors generally determine splicing outcomes in different tissues and cellular conditions.Therefore,altered splicing factor activity is believed to be a main cause of splicing dysregulation incancer374Cancer Cell26,374–389,September8,2014ª2014Elsevier Inc.(Bechara et al.,2013;Shkreta et al.,2013).For example,SRSF1 is a proto-oncogene that controls splicing of several cancer-related genes,including those in the mammalian target of rapa-mycin(mTOR)pathway(Blaustein et al.,2005;Karni et al.,2007). Because splicing dysregulation is one of the molecular hallmarks of cancer(Oltean and Bates,2013),specifically targeting splicing factors opens potential new avenues for cancer therapy(Dehm, 2013).We have previously identified RNA-binding motif4(RBM4)as a binding factor for a group of intronic splicing regulatory ele-ments that control the AS of human genes(Wang et al.,2012). Initially identified by sharing the nuclear import pathway with SR proteins,RBM4shuttles between the cytoplasm and nucleus but is mostly found in nuclear speckles(Lai et al.,2003),where most splicing events occur.RBM4has been shown consistently to control the AS of Tau and a-tropomyosin(Kar et al.,2006;Lin and Tarn,2005).In addition,RBM4has been found to affect translation(Lin and Tarn,2009;Uniacke et al.,2012).Multiple physiological functions have been reported for RBM4,including mediating differentiation of muscle and pancreas cells(Lin et al., 2007;Lin et al.,2013).However,the involvement of RBM4in tumorigenesis has not been reported.Here we systematically analyzed RBM4-mediated changes of the transcriptome and as-sessed the role of RBM4in cancer progression.RESULTSRBM4Is a Sequence-Specific Splicing Inhibitor that Regulates Various AS EventsPreviously we identified several groups of intronic splicing regu-latory elements and their cognate splicing factors(Wang et al., 2012,2013).We demonstrated that,of those factors,RBM4spe-cifically binds to the GTAACG motif to inhibit splicing from in-trons(Wang et al.,2012).In addition,another RBM4binding motif(CGG repeats)was also identified with crosslinking immu-noprecipitation sequencing(Uniacke et al.,2012).Because AS is usually regulated in a context-dependent manner,we sought to examine how RBM4controls splicing when bound to distinct RNA motifs in different pre-mRNA contexts.We generated four splicing reporters with candidate RBM4-binding motifs(GTAACG or CGGCGG)inserted in different re-gions to examine whether RBM4can specifically alter their splicing(Figure1).First,we found that RBM4specifically in-hibited the inclusion of a cassette exon containing its cognate binding sites,whereas the control reporter was not affected(Fig-ure1A).Furthermore,RBM4specifically suppressed the inclu-sion of a reporter exon with a downstream RBM4binding site (Figure1B).These results suggest that RBM4functions as a gen-eral splicing inhibitor to specifically suppress splicing from both exonic and intronic contexts.Such activities are in contrast to DAZAP1,a splicing factor that recognizes the same GTAACG site but functions as a splicing activator(Choudhury et al., 2014).Interestingly,DAZAP1does not affect splicing of exons containing a nearby CGGCGG site(Figures S1A and S1B avail-able online),suggesting a partial overlap of binding specificity and an incomplete functional competition between RBM4and DAZAP1.Using splicing reporters containing RBM4-binding motifs be-tween alternative50ss or30ss,we found that RBM4reduced the use of the downstream50ss(Figure1C)or upstream30ss (Figure1D).The inhibition of distal alternative ss is again sequence-specific because RBM4showed no effect on the con-trol reporters(Figures1C and1D).Consistently,knockdown of RBM4with small hairpin RNA had opposite effects by increasing exon inclusion of the same splicing reporters that contain RBM4-binding sites in various locations(Figures S1C–S1F).In addition, similar results were obtained in a different cell type(e.g.HeLa cells),indicating that the splicing regulation activity of RBM4is not limited to a specific cell line(Figures S1G–S1J).Together, these data demonstrate that RBM4is a general splicing inhibitor that controls different types of AS when specifically binding to pre-mRNA.Like many canonical splicing factors,RBM4has a modular domain configuration.The N terminus contains two RNA recog-nition motifs(RRMs)and a CCHC-type zincfinger that can specifically bind RNAs,whereas the C terminus contains a low-complexity region(i.e.Ala-rich stretches)that can interact with other proteins(Lin and Tarn,2009)(Figure1E).To examine whether RBM4has a modular role in splicing regulation,we fused the full-length N-or C-terminal fragments of RBM4to another RNA binding domain,Pumilio/FBF(PUF)(Wang et al., 2009).We coexpressed the fusion proteins with splicing re-porters containing cognate PUF targets inside an alternative exon(Figure1F)or at a downstream intron(Figure1G)and measured how splicing is affected.As expected,tethering the full-length RBM4to a target site inside an alternative exon sup-pressed exon inclusion.Surprisingly,tethering either the N-or C-terminal domain of RBM4partially inhibited exon inclusion (Figure1F),suggesting that the RNA binding fragment and the low-complexity domain both serve as functional modules. Such an effect is sequence-specific because these fusion pro-teins had no effect on control reporters with a noncognate target. Consistently,the full-length RBM4inhibited exon inclusion when tethered downstream of a cassette exon(Figure1G).Interest-ingly,the N-terminal fragment partially inhibited splicing from an intron,whereas the C terminus showed a slight splicing-inhib-itory activity(Figure1G).Together,the N-terminal RNA-binding fragment and the C-terminal low-complexity domain of RBM4 function cooperatively to control different types of AS events in a sequence-specific manner.Global Regulation of the Transcriptome by RBM4in Cancer-Related GenesTo gain further insights into RBM4-regulated AS events and, thereby,its physiological functions,we conducted high-throughput sequencing of mRNA(mRNA-seq)with H157cells expressing RBM4.With$80million100-nt paired-end reads, we identified473RBM4-regulated AS events with an obvious change of percent-spliced-in(PSI)values(PSI R0.15).Figure2A shows the read tracks of two examples.We found that various types of AS can be regulated by RBM4,including skipped exon(SE),alternative50ss exon(A5E),alternative30ss exon (A3E),retained intron(RI),mutually exclusive exons(MXE),and tandem UTR(TUTR)(Figure2B;Table S1).Subsequent analysis indicated that most of the AS events were negatively regulated by RBM4(decreased PSI value by RBM4expression)(Figure2C), consistent with ourfinding that RBM4suppressed splicing when binding directly to its pre-mRNA targets(Figures1A–1D).Cancer CellRBM4Inhibits Tumorigenesis via Splicing ControlCancer Cell26,374–389,September8,2014ª2014Elsevier Inc.375(legend on next page)Cancer CellRBM4Inhibits Tumorigenesis via Splicing Control376Cancer Cell 26,374–389,September 8,2014ª2014Elsevier Inc.We further analyzed RNA motifs in RBM4-regualted pre-mRNAs by extracting the sequences near the RBM4-regulated SEs or between alternative 50ss of A5E.The relative abundance of RBM4binding motifs (GTAACG and CGGCGG)in these re-gions was compared with control exons unaffected by RBM4(Fairbrother et al.,2002).We found that RBM4-binding motifs are enriched near the SEs or A5Es negatively regulated by RBM4(Figure 2D),consistent with the model that RBM4directly recognizes these pre-mRNAs to control splicing.The AS events apparently promoted by RBM4are likely due to indirect effects because these exons lack known RBM4binding motifs (Figure 2D).When analyzing cellular functions of RBM4-regulated AS events using gene ontology,we found that RBM4affects genes in the RNA processing pathway,including translation control,RNA processing,and the mRNA metabolic process (Figure 2E).Such functional enrichment is not surprising because RBM4is an RNA binding factor known to regulate splicing and translation.Intriguingly,RBM4targets are also enriched with cancer-related functions such as regulation of the NF-k B cascade and cell cy-cle.In addition,several RBM4-regulated AS events were found to regulate the apoptotic pathway.Although this enrichment of apoptosis is slightly below our significance cutoff,the changes of PSI value are fairly large and,therefore,may have significant functional consequences.Many of the RBM4-regulated splicing targets were functionally connected into well linked interaction networks,as judged by the Search Tool for the Retrieval of Inter-acting Genes/Proteins (STRING)(Figure 2F).As expected,two large subgroups of RBM4targets contain genes involved in translation control and RNA processing.Surprisingly,the other subgroup includes many genes involved in cell migration and adhesion (Figure 2F).Taken together,these results suggest that the biological processes affected by RBM4are related to apoptosis,proliferation,migration,and tumorigenesis.We subsequently validated mRNA-seq results by measuring the splicing change of ten newly identified targets that were selected arbitrarily to include genes with a cancer-related func-tion.We confirmed that RBM4either positively or negatively con-trols all endogenous AS events tested (Figure 2G)and that the relative changes of PSIs obtained from RT-PCR are highly corre-lated to those observed by mRNA-seq (Figure S2A;R 2=0.6).These events were also validated in another cell line (HeLa)(Fig-ure S2B),suggesting that RBM4can regulate AS with consistent activity across different cell types.In addition,we found that knockdown of RBM4caused opposite changes of splicing inendogenous RBM4targets,further confirming the reliability of our analyses (Figures S2C and S2D).We also analyzed how RBM4affects global gene expression.We identified 185genes with significant expression change (>2-fold with adjusted p <0.05)(Table S2).These genes are associ-ated significantly with cancer-related functions,as judged by gene ontology (including DNA replication,chemotaxis,cell pro-liferation,response to wounding,cell cycle,and cell migration;Figure 2H),again suggesting that RBM4is involved in cancer cell proliferation and migration.Many RBM4-regulated genes were also connected functionally into a densely linked network that contains genes involved in regulating cell proliferation,wound healing,cell cycle,and DNA damage (Figure 2I).The selected RBM4targets were further validated with real-time RT-PCRs (Figure 2J).Taken together,these data imply that RBM4may be a key regulator of cell proliferation and migration,therefore controlling cancer progression.RBM4Inhibits Cancer Cell Proliferation and Migration To examine this possible role of RBM4in cancer progression,we stably expressed RBM4in a panel of human cancer cells,including H157(lung cancer),MDA-MB-231(breast cancer),SKOV3(ovarian cancer),Panc-1(pancreatic cancer),HepG2(liver cancer),and PC-3(prostate cancer)(Figure S3A).Strikingly,in all cancer cells tested,RBM4inhibited anchorage-dependent or anchorage-independent growth,as judged by colony forma-tion or soft agar assay (Figure 3A).In addition,RBM4inhibited migration of these cells in a wound healing assay (Figure 3B).Together,the inhibition of cancer cell proliferation and migration by RBM4suggests that it may function as a tumor suppressor.We further analyzed how RBM4affects cancer progression using non-small cell lung carcinoma (NSCLC)cells,which repre-sent one of the most prevalent human cancers.The RBM4levels were decreased markedly in a panel of NSCLC cells compared with normal bronchial cells (Figure 3C).Consistently,when re-expressed in a NSCLC cell line,H157,RBM4significantly in-hibited cell growth (Figure 3D;p =0.02by t test).Similar growth inhibition by RBM4was observed in 293T cells (Figures S3B and S3C).Interestingly,although both the N-and C termini of RBM4partially regulate splicing,lung cancer cells expressing either domain (amino acids (aa)1–177or aa 178–364of RBM4)dis-played normal growth rates (Figure 3E),suggesting that both do-mains are required to suppress tumorigenesis.To further assess whether RBM4affects cancer growth in vivo,we determined whether RBM4re-expression can suppressFigure 1.Splicing Regulation by RBM4(A)The RBM4binding sites and a control (GAATTG)were inserted into splicing reporter pGZ3and cotransfected with the RBM4expression vector or an empty vector (mock)into 293T cells.Splicing changes were examined by electrophoresis of RT-PCR products.(B)The same set of sequences as analyzed in (A)was inserted downstream of a cassette exon in the pZW2C reporter to measure splicing changes as in (A).(C and D)The same set of RBM4-binding sequences as analyzed in (A)was inserted into the splicing reporters between two tandem sites with competing 50(C)or 30ss (D),and splicing changes were measured as in (A).(E)Schematics of RBM4domains.The R1R2Z fragment contains two RRM domains and a zinc finger domain.The polyalanine fragment contains a polyalanine stretch.(F and G)Different RBM4fragments were fused to a PUF domain,PUF(3-2),that specifically binds to its target RNA.The fusion proteins were cotransfected with a splicing reporter containing a PUF binding site or a control (Ctl)site in a cassette exon (F)or at a downstream intron (G),and splicing changes were measured as in (A).The arrowhead indicates a nonspecific product (F).In panels measuring changes in splicing,expression of exogenous protein was confirmed by western blot analyses.Tubulin served as a protein loading control.Three independent experiments were conducted,with the mean ±SD of PSIs plotted below the representative gels.*p <0.05as calculated by Student’s t test.See also Figure S1.Cancer CellRBM4Inhibits Tumorigenesis via Splicing ControlCancer Cell 26,374–389,September 8,2014ª2014Elsevier Inc.377(legend on next page)Cancer CellRBM4Inhibits Tumorigenesis via Splicing Control378Cancer Cell 26,374–389,September 8,2014ª2014Elsevier Inc.tumor growth in a xenograft mouse model.We generated H157-luc-RBM4cells and control cells with lentiviral vectors and injected them subcutaneously into the flanks of nude mice (left flank,RBM4;right flank,control).The growth of tumors was measured every 3days for 5weeks,and xenograft tumors were removed for final analysis.Consistent with the in vitro re-sults,cells expressing RBM4developed smaller tumors compared with control cells (Figures 3F and 3G).In addition,the xenograft tumors with RBM4re-expression grew much slower than controls (Figure 3H),suggesting that RBM4substan-tially inhibits cancer progression in vivo.Together,these findings indicate that RBM4is a potent tumor suppressor that inhibits lung cancer progression both in cultured cells and in a tumor xenograft model.RBM4Induces Cancer Cell Apoptosis via Regulating AS of Bcl-xTo determine the mechanisms of how RBM4affects cancer pro-gression,we focused on an RBM4target gene,Bcl-x,an apoptosis regulator that produces two splicing isoforms with opposite functions.By alternative use of 50ss,Bcl-x is spliced as an antiapoptotic isoform (Bcl-xL)or a proapoptotic isoform (Bcl-xS)(Adams and Cory,2007).RBM4expression appeared to shift Bcl-xL into Bcl-xS (Figure 2G).Such a shift requires an entire RBM4because neither the N terminus nor the C terminus can affect Bcl-x splicing by itself (Figure S4A).We identified a po-tential RBM4binding site (CGGCGG)between the two alterna-tive 50ss (Figure 4A),implying that RBM4may control splicing through binding directly to Bcl-x pre-mRNA.Consistently,with an RNA immunoprecipitation assay,we found that RBM4indeed binds directly to the endogenous Bcl-x pre-mRNA but not the control pre-mRNA of another alternatively spliced apoptotic gene (Mcl1)(Figure 4B).Using a splicing reporter containing Bcl-x pre-mRNA,we found that RBM4binding is indeed depen-dent on the CGGCGG site because mutation of this site abol-ished RNA-protein interaction (Figure 4C).Replacing the mutated sequence with the other RBM4-binding site (GTAACG)restored the interaction,confirming that RBM4directly recog-nizes the exon extension region of Bcl-x.In addition to H157cells,an inducible expression of RBM4also shifted splicing of Bcl-x in 293cells (Figure 4D).This shift caused a rapid and robust decrease of Bcl-xL protein,as judgedby western blot analysis (Figure S4B).To determine whether the binding by RBM4is responsible for the observed splicing shift,we cotransfected RBM4with a series of Bcl-x reporters contain-ing various mutations near the alternative 50ss (Figure 4A).We found that RBM4shifted the splicing of the wild-type reporter by reducing Bcl-xL and that such a regulation was not affected by three exonic mutations (mutations 1–3)(Figure 4E).However,the mutation of the RBM4binding site (mut 4)completely abol-ished the splicing regulation through RBM4,indicating that the RBM4binding motif (CGGCGG)is indeed responsible for the Bcl-x splicing switch.Importantly,replacing CGGCGG with another RBM4binding site (mut 5)restored the regulation by RBM4(Figure 4E),suggesting that binding of RBM4to Bcl-x pre-mRNA is sufficient to shift splicing.The two splicing isoforms of Bcl-x have opposite functions in controlling apoptosis (Adams and Cory,2007).Bcl-xL is the pre-dominant isoform in cancer,and RNAi of Bcl-xL has been shown to induce apoptosis in several cancer cell lines (Mercatante et al.,2001;Zhu et al.,2005).We found that expression of RBM4in H157cells substantially reduced the level of Bcl-xL protein,re-sulting in the cleavage of caspase 3and poly-ADP-ribose poly-merase (PARP),two molecular markers of apoptosis (Figure 4F).Consistently,RBM4dramatically increased spontaneous apoptosis,as judged by flow cytometry (Figure 4G;Figure S4C).These results support the model that sequence-specific binding of RBM4to Bcl-x pre-mRNA shifts its splicing from antiapoptotic Bcl-xL to proapoptotic Bcl-xS,thereby promoting cancer cell death.RBM4Suppresses Tumor Progression in Part through Bcl-xBecause RBM4may inhibit cancer proliferation through modu-lating Bcl-x splicing,we next examined whether coexpression of Bcl-xL,but not other similar apoptotic regulators,can overturn the tumor suppressor activity of RBM4.We stably transfected the parental H157line containing re-expressed RBM4with Bcl-xL or another apoptotic inhibitor,Mcl-1(Figure 5A),generating a cell line with a partially restored Bcl-xL/Bcl-xS ratio and reduced PARP cleavage (Figure 5B).We found that cells expressing RBM4/Bcl-xL grew much faster than those ex-pressing RBM4alone,although the growth rate was not fully restored compared with the control (Figure 5C).However,cellsFigure 2.Global Splicing and Transcriptional Regulation by RBM4(A)Examples of alternative exons affected by RBM4.Genes were chosen to represent both an increase and a decrease of PSI,and the numbers of exon junction reads are indicated.(B)Quantification of the different AS events affected by RBM4.(C)The relative fraction of each AS event affected positively or negatively by RBM4.(D)Relative enrichment of the indicated RNA motifs bound by RBM4.Enrichment scores were computed by comparing RBM4-regulated SEs or A5Es with control AS events unaffected by RBM4.AS events with increased or decreased PSI values upon RBM4expression were analyzed separately.(E)Gene ontology of RBM4-regulated AS targets.Fisher exact p values were plotted for each enriched functional category.(F)Functional association network of RBM4-regulated AS targets.The genes in (E)were analyzed using the STRING database,and subgroups are marked according to their functions.(G)Validation of different types of RBM4-regulated AS events by semiquantitative RT-PCR using H157cells transfected with RBM4or control vectors.The mean ±SD of PSIs from three experiments were plotted (p values were calculated by paired Student’s t test).(H)Gene ontology analyses of RBM4-regulated gene expression events.Fisher exact p values were plotted for each category.(I)The functional association networks of RBM4-regulated genes were analyzed using the STRING database,with subgroups marked by their functions.(J)Validation of gene expression changes by real-time RT-PCR.The mean ±SD of relative fold changes from triplicate experiments were plotted,with p values calculated by paired Student’s t test.See also Figure S2and Tables S1and S2.Cancer CellRBM4Inhibits Tumorigenesis via Splicing ControlCancer Cell 26,374–389,September 8,2014ª2014Elsevier Inc.379(legend on next page)Cancer CellRBM4Inhibits Tumorigenesis via Splicing Control380Cancer Cell 26,374–389,September 8,2014ª2014Elsevier Inc.expressing RBM4/Mcl-1showed a similar growth rate compared with cells expressing RBM4alone (Figure 5C),indicating that such phenotypical rescue is specific for Bcl-xL.In addition,can-cer cells expressing RBM4/Bcl-xL migrated significantly faster than cells expressing RBM4alone or RBM4/Mcl-1(Figure 5D),again suggesting that restoring the Bcl-xL level partially reversed the RBM4phenotype.Consistently,the xenograft tumors gener-ated from RBM4/Bcl-xL cells were significantly larger than those from RBM4/vector cells,indicating that reducing the Bcl-xL level is partially responsible for RBM4-mediated tumor suppression in vivo (Figure 5E).This phenotypic rescue is robust and statisti-cally significant,although it could not fully restore tumor progres-sion,probably because of the partial reversal of the Bcl-xL/Bcl-xS ratio (Figure 5B).We further applied a specific Bcl-xL inhibitor (WEHI-539)in cells expressing RBM4and examined its effect on cell growth.Consistent with a previous report (Lessene et al.,2013),WEHI-539did not significantly affect the viability of control cells.However,WEHI-539treatment inhibited the proliferation of RBM4-expressing cancer cells compared with untreated cells (Figures 5F and 5G).Such an apparent synergistic effect may reflect two mechanisms that are not mutually exclusive:(1)Through splicing regulation,RBM4reduces the level of Bcl-xL to the extent where the WEHI-539can have a detectable effect;(2)RBM4inhibits cell proliferation through other mecha-nisms in addition to reducing antiapoptotic Bcl-xL,whereas WEHI-539specifically inhibits Bcl-xL.By targeting parallel pro-survival pathways,the combination of RBM4and WEHI-539syn-ergistically suppressed cancer cell proliferation.Consistently,we found an increased expression of Bcl-xL in lung cancers,breast cancers,and pancreatic cancers,which is correlated inversely to the RBM4level (Figure 5H;Figures S5A and S5B).This finding further supports the hypothesis that RBM4inhibits tumor progression (at least partially)via controlling Bcl-x splicing.RBM4Antagonizes Oncogenic SRSF1to Inhibit mTOR ActivationAlthough our data clearly demonstrate that RBM4suppresses cancer progression by modulating Bcl-x splicing,this may not be the only mechanism because coexpression of Bcl-xL partially reversed the phenotype of RBM4.To eliminate the apoptosis ef-fect,we treated cells with a pan-caspase inhibitor,carboben-zoxy-valyl-alanyl-aspartyl (Z-VAD).We found that,even when the apoptosis was inhibited strongly (Figure 6A),proliferation and migration of cancer cells were still suppressed significantly by RBM4(Figure 6B).This observation suggests that RBM4might also inhibit cancer progression through other mechanisms besides regulating apoptosis.It has been reported previously that the general splicing factor SRSF1functions as a proto-oncogene to transform rodent fibro-blasts (Karni et al.,2007).We found that RBM4interacted with SRSF1in a coimmunoprecipitation assay (Figure S6A).Remark-ably,RBM4can reduce the protein level of SRSF1in a dose-dependent manner (Figure 6C).Such inhibition is specific to SRSF1because two other splicing factors,DAZAP1and hnRNPA1,were not affected (Figure 6C).Similar results were also obtained in a cell line with inducible expression of RBM4(Figure S6B).Since SRSF1is a well characterized oncogenic factor to promote tumorigenesis through multiple pathways (An-czuko´w et al.,2012;Karni et al.,2007),our observation suggests that RBM4may also inhibit cancer progression by antagonizing SRSF1.SRSF1is known to control multiple AS events that promotetumorigenesis (Anczuko´w et al.,2012;Karni et al.,2007).For example,BIN1is a tumor suppressor that binds to MYC (Saka-muro et al.,1996),and SRSF1promotes inclusion of BIN1exon 12a to generate a BIN1+12isoform that lacks tumor suppressor activity (Karni et al.,2007).SRSF1also inhibits the exclusion of exon 11in RON,generating RON D 11,which promotes cellmigration and invasion (Anczuko´w et al.,2012).We examined whether RBM4could affect the splicing of cancer-related SRSF1targets using cells stably expressing SRSF1,RBM4,or SRSF1/RBM4.As expected,RBM4regulated splicing of both BIN1and RON in an opposite fashion as SRSF1,shifting their splicing toward antioncogenic isoforms (Figure 6D;Figure S6C).SRSF1has also been reported to activate the mTOR pathway by increasing phosphorylation of S6K1and 4E-BP1as well as by promoting oncogenic S6K1splicing isoform 2(Karni et al.,2007;Karni et al.,2008).Coexpression of RBM4with SRSF1substan-tially inhibited SRSF1-induced mTOR activation,as judged by the dramatic reduction in the phosphorylation of S6K1and 4E-BP1(Figure 6E).However,phosphorylation of two upstreamFigure 3.RBM4Inhibits Cancer Progression(A)RBM4effects on the proliferation of various cancer cells,including H157,MDA-MB-231,SKOV3,Panc-1,HepG2,and PC-3cells.The cells were stably transfected with RBM4or a vector control and analyzed by colony formation (top panels)or soft agar (bottom panels)assays.All experiments were performed in triplicate,with mean ±SD of relative colony numbers plotted (p values were calculated by Student’s t test).Images of the whole plate are shown in the top panels.Scale bars,100m m.(B)Different cancer cell lines expressing RBM4or a vector control were analyzed by wound healing assay.Percent of wound closure was measured in triplicate experiments,with mean ±SD plotted (p values were calculated by Student’s t test).Scale bar,200m m.(C)Levels of RBM4in the indicated NSCLC cell lines and normal bronchial cells were measured by western blot analysis.(D)H157cells stably expressing RBM4or a vector control were grown for 9days,with cell numbers counted every 2days.The changes of cell numbers were compared to day 0.The mean ±SD from three experiments was plotted.(E)H157cells expressing full-length (FL)RBM4or the N-terminal (N-term)or C-terminal (C-term)fragments of RBM4were analyzed by colony formation assay.Representative pictures of the whole plates from triplicate experiments are shown.The mean ±SD of relative colony numbers were plotted,with p values calculated by Student’s t test.(F)H157-luc-RBM4and control cells were injected subcutaneously into the left and right flanks of seven nude mice.The growth of xenograft tumors was monitored by bioluminescence imaging on days 3and 35,and pictures of two representative mice are shown.(G)Pictures of the tumors removed after 35days.(H)The average sizes of xenograft tumors measured every 3days (n =7,error bars indicate SD,p <0.05by Student’s t test).See also Figure S3.Cancer CellRBM4Inhibits Tumorigenesis via Splicing ControlCancer Cell 26,374–389,September 8,2014ª2014Elsevier Inc.381。

A Functional Genomic Approach Identifies FAL1 as an Oncogenic Long Noncoding RNA that Associates

A Functional Genomic Approach Identifies FAL1 as an Oncogenic Long Noncoding RNA that Associates

Cancer CellArticleA Functional Genomic Approach Identifies FAL1as an Oncogenic Long Noncoding RNA that Associates with BMI1and Represses p21Expression in CancerXiaowen Hu,1,2,3Yi Feng,4Dongmei Zhang,1,2,8Sihai D.Zhao,10Zhongyi Hu,1,2Joel Greshock,4Youyou Zhang,1,2Lu Yang,1,2,9Xiaomin Zhong,1,2,11Li-Ping Wang,5Stephanie Jean,3Chunsheng Li,1,2Qihong Huang,12Dionyssios Katsaros,13Kathleen T.Montone,5Janos L.Tanyi,3Yiling Lu,15Jeff Boyd,14Katherine L.Nathanson,6 Hongzhe Li,7Gordon ls,15and Lin Zhang1,2,3,*1Ovarian Cancer Research Center2Center for Research on Reproduction&Women’s Health3Department of Obstetrics and Gynecology4Abramson Family Cancer Research Institute5Department of Pathology and Laboratory Medicine6Department of Medicine7Department of Biostatistics and EpidemiologyPerelman School of Medicine,University of Pennsylvania,Philadelphia,PA19104,USA8State Key Laboratory of Biotherapy9Department of Obstetrics and GynecologyWest China Medical School,Sichuan University,Chengdu610041,China10Department of Statistics,University of Illinois at Urbana-Champaign,Champaign,IL61820,USA11Center for Stem Biology and Tissue Engineering,Department of Biology,Zhongshan School of Medicine,Sun Yat-sen University, Guangzhou510080,China12Wistar Institute,Philadelphia,PA19104,USA13Department of Obstetrics and Gynecology,University of Turin,Turin10124,Italy14Cancer Genome Institute,Fox Chase Cancer Center,Philadelphia,PA19111,USA15Department of Systems Biology,MD Anderson Cancer Center,Houston,TX7705,USA*Correspondence:linzhang@/10.1016/r.2014.07.009SUMMARYIn a genome-wide survey on somatic copy-number alterations(SCNAs)of long noncoding RNA(lncRNA)in 2,394tumor specimens from12cancer types,we found that about21.8%of lncRNA genes were located in regions with focal SCNAs.By integrating bioinformatics analyses of lncRNA SCNAs and expression with functional screening assays,we identified an oncogene,f ocally a mplified l ncRNA on chromosome1 (FAL1),whose copy number and expression are correlated with outcomes in ovarian cancer.FAL1associates with the epigenetic repressor BMI1and regulates its stability in order to modulate the transcription of a num-ber of genes including CDKN1A.The oncogenic activity of FAL1is partially attributable to its repression of p21.FAL1-specific siRNAs significantly inhibit tumor growth in vivo.INTRODUCTIONCancer genomes are highly disorganized and harbor numerous somatic copy-number alterations(SCNAs)(Beroukhim et al., 2010;Zack et al.,2013).Although the majority of the copy num-ber abnormalities are the consequence of genomic instability,a subset of SCNAs contributes to tumorigenesis.Systemic ana-lyses using large-scale genomic profiles and genome-wide func-tional screening have been successfully applied to identifying cancer-driving SCNA loci that encode proteins(Beroukhim et al.,2010;Zack et al.,2013).However,protein-coding sequences occupy less than2%of the humangenome344Cancer Cell26,344–357,September8,2014ª2014Elsevier Inc.(International Human Genome Sequencing Consortium,2004), and many focal SCNAs in cancer have been mapped to‘‘pro-tein-coding gene desert’’regions(Beroukhim et al.,2010;Zack et al.,2013).Recent advances in high-throughput sequencing technology have revealed that the majority( 70%)of the human genome is transcribed to RNA,generating many thousands of noncoding transcripts(Derrien et al.,2012;Djebali et al.,2012).Long noncoding RNAs(lncRNAs)are operationally defined as RNA transcripts that are larger than200nt but do not appear to have protein-coding potential(Batista and Chang,2013;Gutt-man and Rinn,2012;Karreth and Pandolfi,2013;Lee,2012;Lie-berman et al.,2013;Ørom and Shiekhattar,2013;Prensner and Chinnaiyan,2011;Ulitsky and Bartel,2013).Similar to protein-coding transcripts,the transcription of lncRNAs is subject to typical histone-modification-mediated regulation,and lncRNA transcripts are processed by the canonical spliceosome pared with their protein-coding counterparts,lncRNA genes are composed of fewer exons,under weaker selective constraints during evolution,and in relatively lower abundance. In addition,the expression of lncRNAs is strikingly cell type and tissue specific and,in many cases,even primate specific.To date,most of the well-characterized lncRNAs have been discov-ered serendipitously.The investigations on this small cohort of lncRNAs have demonstrated that these noncoding transcripts can serve as scaffolds or guides to regulate protein-protein or protein-DNA interactions(Engreitz et al.,2013;Gupta et al., 2010;Huarte et al.,2010;Jeon and Lee,2011;Simon et al., 2013;Tsai et al.,2010;Yang et al.,2011,2013b),as decoys to bind proteins(Di Ruscio et al.,2013;Hung et al.,2011;Tripathi et al.,2010,2013)or microRNAs(miRNAs)(Hansen et al., 2013;Memczak et al.,2013;Poliseno et al.,2010;Tay et al., 2011),and as enhancers to influence gene transcription,when transcribed from the enhancer regions(enhancer RNA)(Kim et al.,2010;Li et al.,2013;Wang et al.,2011)or their neighboring loci(noncoding RNA activator)(Lai et al.,2013;Ørom et al., 2010).The biological processes affected by lncRNAs include cell proliferation(Hung et al.,2011;Tripathi et al.,2013),differen-tiation(Guttman et al.,2009;Guttman et al.,2011;Kretz et al., 2013;Loewer et al.,2010;Ulitsky et al.,2011),migration(Gupta et al.,2010;Ling et al.,2013;Ørom et al.,2010;Yang et al., 2013a),immune response(Carpenter et al.,2013;Gomez et al., 2013),and apoptosis(Huarte et al.,2010),all of which have been implicated in tumorigenesis.In addition to being higher de-regulated in tumors(Du et al.,2013;Gupta et al.,2010;Prensner et al.,2011),lncRNAs have been found to act as tumor suppres-sors or oncogenes(Gupta et al.,2010;Ji et al.,2003;Ling et al., 2013;Pasmant et al.,2007;Prensner et al.,2011,2013;Yang et al.,2013b;Yildirim et al.,2013).To characterize the landscape of lncRNA gene SCNAs across cancers,we repurposed the SNP microarray results from a total of2,394tumor specimens taken from12cancer types(Beroukhim et al.,2010)and analyzed the SCNAs of13,870lncRNA gene loci.RESULTSlncRNAs Exhibit Frequent SCNAs in Human CancerWe analyzed the SNP arrays of a total of2,394tumor specimens from12cancer types in the Tumorscape database created by the Broad Institute(Beroukhim et al.,2010)(Table S1and Fig-ure S1A available online).The genomic locations of13,870 lncRNAs(Table S2)were retrieved from an evidence-based lncRNA annotation provide by the GENCODE Consortium(Der-rien et al.,2012),and the SCNA frequency of each lncRNA-containing locus was calculated.This revealed that the more frequently a lncRNA has a copy-number gain in a given tumor type,the less likely it would also have a high frequency of copy-number loss in the same tumor type(Figure S1B).As a result,when we define high-frequency gains or losses as alter-ations that take place in more than25%of specimens from a given tumor type,few lncRNAs had both high-frequency gain and loss in the same type of tumor.Across the12tumor types, there were on average12.0%and7.6%of lncRNAs with high-frequency(i.e.,in>25%of tumors)gain and loss,respectively (Figures1A–1C;Table S3).Although small cell lung cancer had the largest number of high-frequency lncRNA SCNAs,myelopro-liferative disorder had none(Figures1B and1C).Similar to the overall genomic alteration profiles,lncRNA SCNA profiles were cancer-type specific(Figure1B;Figure S1A).Additionally,we analyzed the SNP arrays using a second lncRNA annotation generated by Cabili et al.(2011)(Table S2)and found the lncRNA SCNA frequency and tumor-type specificity were similar to that analyzed with GENCODE annotation(Figures S1C–S1E and Table S3).To further validate thesefindings,we acquired SNP arrays from The Cancer Genome Atlas(TCGA)project and analyzed lncRNA SCNAs in breast cancer.The lncRNA SCNA profiles in breast cancer samples from TCGA data sets were almost identical to those from the Broad Institute database (Figure S1F).Two types of SCNAs are present in cancer genomes:those confined to a small genomic region are termed focal alterations, and those encompassing a large fragment,or even a whole chromosomal arm,are referred as broad(arm-level)alterations. Because focal alterations contain only a handful of genes and often exhibit high-amplitude variation,analyses of these alterations have led to the successful identification of cancer-causing genes(Beroukhim et al.,2010;Du et al.,2013).To screen for lncRNAs that may act as driver genes in tumorigen-esis,we mapped lncRNA loci to158independent focal genomic alteration peaks(76gains and82losses)that have been previously identified(Beroukhim et al.,2010).Totals of 1,064and1,953lncRNAs were located in the regions with focal gains and losses,respectively(Tables S4and S5).Although 995lncRNAs were located in focal SCNA regions where can-cer-associated protein-coding genes reside,we identified 2,022(14.6%)lncRNAs in focal alteration regions that contain no known cancer-associated protein-coding genes(Tables S4 and S5).Importantly,within the top20most significant focal alteration peaks(Beroukhim et al.,2010),we identified 56lncRNAs in focal gain regions and132lncRNAs in focal loss regions(Figure1D).We reasoned that the lncRNAs that demonstrate high-frequency genomic alterations and/or reside in focal alteration loci are candidates for cancer-causing lncRNAs.lncRNAs Are Widely Expressed in Human Cancer Cells Because lncRNAs exert their functions as RNAs,we reasoned that the presence of RNA transcripts in cells should be aCancer CellOncogenic lncRNA FAL1Represses p21ExpressionCancer Cell26,344–357,September8,2014ª2014Elsevier Inc.345prerequisite for a lncRNA to be functional and that alterations in the genomic loci harboring lncRNAs with no detectable RNA transcripts are likely to be passenger events.We profiled 40established cancer cell lines (across five cancer types)from the NCI60cell line panel (Table S6)using a custom 60-mer oligo-nucleotide microarray with a total of 14,262probes for 2,965lncRNAs (an average of 5probes for each lncRNA;Table S7),which were initially identified using the GENCODE annotation (Ørom et al.,2010).Probes for 11,081protein-coding genes were also included in our microarray as controls.Overall,41.7%of the lncRNA and 82.9%of the protein-coding gene probes were detected in 10(25%)or more of the 40cell lines;23.8%of the lncRNA and 4.9%of the protein-coding gene probes were not detected in any cell line (Figure S1G).Among all the lncRNAs studied,about 17.8%were expressed in all 40cancer cell lines.To validate the RNA expression results from mi-croarray,we measured the RNA expression of 6well-known lncRNAs in these cancer cell lines by quantitative RT-PCR (qRT-PCR)and found that there were strong correlations be-tween the RNA expression measured by microarray and by PCR (Figure S1H).These findings demonstrate that lncRNAs are indeed widely expressed in cancers.Together,the cancer-cell-specific RNA expression information and the lncRNAs SCNA in multiple types of tumors can help us narrow down the list of cancer-causing lncRNA candidates by eliminating lncRNAs that do not express in cancer cells.Clinically Guided Genetic Screening Identified FAL1as a Potential Oncogenic lncRNANext,we used the information obtained from the above genomic and transcriptomic analyses to select oncogenic lncRNA candi-dates for functional validation.The three criteria for candidate selection were as follows:(1)the lncRNA copy-number gain is observed in more than 25%of the samples in at least one type of tumors,(2)the lncRNA is located in a focal amplicon,and (3)the RNA expression of the candidate lncRNA is detected in more than 50%of cancer cell lines.The functional readout for the initial screening was in vitro clonogenicity.We hypothesized that short hairpin RNAs (shRNAs)targeting true oncogenic lncRNAs should greatly reduce the clonogenicity of cells,and shRNAs targeting bystander lncRNAs will have no effect.To minimize the possibility of observing off-target effects,we de-signed two independent shRNAs for each lncRNA candidate.In the initial clonogenic screening (Figure 2A),37lncRNA candi-dates were screened,and we found that both shRNAs targeting ENSG00000228126(focally amplified lncRNA on chromosome 1[FAL1]),a lncRNA in a focal amplicon on chromosome 1q21.2(Figures S2A and S2B),significantly reduced the clonogenicity of A2780cells in a dose-dependent pared with FAL1shRNA1,shRNA2was more efficient in knocking down endogenous FAL1expression (Figure 2B)and had a greater ef-fect on inhibiting cell growth and colony formation (Figures 2A and 2B).Similar results were also observed inMDA-MB-231Figure 1.SCNAs of lncRNA in Cancers(A)A genome-wide view of SCNAs in lncRNA-containing loci in cancers.Each track shows the frequency of lncRNA SCNAs in one cancer type.Red indicates gain;blue indicates loss.The outer and inner tracks represent cancer types 1and 12,respectively.(B)Heatmap of SCNA frequencies of lncRNA genomic loci in cancers.Each row represents one lncRNA,ordered by genomic location.Left,frequency of gain (red);right,frequency of loss (blue).(C)Percentages of lncRNAs with significant copy-number alteration (>25%of specimens)in cancers.(D)The lncRNAs and protein-coding genes in the top 20most significant focal gain (left)or loss (right)peaks across cancers.The numbers of protein-coding genes (left)and lncRNAs (right)in each peak are indicated in parentheses.The independent focal genomic alteration peaks and the numbers of protein-coding genes in each peak were previously identified by the Tumorscape Project (Beroukhim et al.,2010).See also Figure S1and Tables S1,S2,S3,S4,S5,S6,and S7.Cancer CellOncogenic lncRNA FAL1Represses p21Expression346Cancer Cell 26,344–357,September 8,2014ª2014Elsevier Inc.and HCT116cells.Next,we validated the oncogenicity of FAL1in seven more cell lines that have a wide range of FAL1expres-sion and various status of FAL1SCNA.With the exception of SKOV3cells,which have normal copy number and low RNA expression of FAL1,all other cell lines were more or less depen-dent on the expression of FAL1for their growth (Figure 2C).Soft-agar assays further demonstrated that the expression of FAL1shRNAs significantly inhibited theanchorage-independentFigure 2.Identification and Validation of FAL1as a Potential Oncogenic lncRNA(A)Representative results from clonogenic shRNA screening for oncogenic lncRNAs in A2780(in 24-well plates).(Bottom)Wells with colonies expressing controls and FAL1hairpins.(B)Relative expression of FAL1(left)and growth curve (right)of A2780cells expressing control and FAL1shRNAs.(C)Growth curves of seven cancer cell lines transfected with control or FAL1siRNAs.The FAL1SCNA status of each cell line is indicated as a blue (gain)or gray (normal)rectangle,and the relative FAL1expression in parental cells is indicated by the intensities of the pink rectangles.(D).Soft-agar assay with cells expressing control and FAL1shRNAs (in 6-well plates).(E)In vivo xenograft tumor growth curves of A2780and MDA-MB-231cells expressing control and FAL1shRNAs.(F)Schematic diagram of the experimental design of testing the oncogenic potential of FAL1.(G)The expression of Myc or Ras in HOSE cells transduced with FAL1alone or in combination with Myc or Ras.(H and I)The representative result of soft-agar assay (H)and the corresponding quantification (I)on control cells and cells expressing FAL1alone or in combination with Myc or Ras.Error bars indicate SD.*p <0.05.See also Figure S2.Cancer CellOncogenic lncRNA FAL1Represses p21ExpressionCancer Cell 26,344–357,September 8,2014ª2014Elsevier Inc.347growth of cancer cells(Figure2D).Next,we demonstrated that the expression of FAL1shRNAs significantly suppressed the growth of subcutaneous tumors formed by A2780or MDA-MB-231cells in nude mice(Figure2E).We examined if FAL1expression is sufficient to promote trans-formation.We forced the expression of full-length FAL1cDNA (Figures S2C and S2D)in two independent batches of primary human ovarian surface epithelial(HOSE)cells and further trans-duced these FAL1-modified cells with Myc or Ras and their cor-responding controls(Figures2F and2G).The oncogenicity of FAL1alone or in combination with Myc or Ras was evaluated in soft-agar assays.Although control HOSE cells form no colony in soft agar,cells expressing FAL1were able to form some col-onies,although compared with those formed with Myc or Ras cells,the FAL1colonies were smaller and in fewer numbers(Fig-ures2F and2G).Intriguingly,HOSE cells expressing FAL1in combination with Myc(or Ras)formed significantly more col-onies than their single-gene expressing counterparts(Figures 2H and2I;Figure S2E–S2Q).In aggregate,by integrating genomic and transcriptomic analysis with functional screening, we have successfully identified FAL1as a potential oncogenic lncRNA.Interestingly,FAL1amplicon also contains a known protein-coding oncogene,MCL1(Beroukhim et al.,2010).We compared the mRNA levels of MCL1andfive other genes within the FAL1 locus in control and FAL1shRNA expressing A2780cells and found that knocking down FAL1did not affect the expression of any of these neighboring genes(Figure S2R).Thisfinding sug-gests that FAL1does not control the transcription of its neigh-boring genes;as such,the function of FAL1is likely independent to regulation of MCL1expression.It has been documented that a cluster of oncogenic lncRNAs,including PCAT-1,CCAT2,and CARLo-5,coamplify with MYC;yet they promote tumor growth via Myc-independent mechanisms(Kim et al.,2014;Ling et al., 2013;Prensner et al.,2011).Expression and SCNA of FAL1Are Associated with Clinical Outcomes in Patients with Ovarian CancerAn in-depth investigation of the SNP arrays revealed that the fre-quency of FAL1copy-number gain was remarkably high(49.7%) in epithelial tumors but much lower in neural(<19%)and hema-tologic(<6%)tumors(Figure3A).Importantly,FAL1gene resides at a significant focal amplicon(Q<0.25)on chromosome1q21.2 in epithelial cancers(Figures3A and3B).To confirm these obser-vations,we measured the copy number of FAL1in99cancer cell lines using quantitative PCR and observed FAL1copy-number gain in46%of the cell lines(Figure3C;Table S8).We then ex-tracted the FAL1RNA expression data from the aforementioned custom RNA array containing40cancer cell lines and found a significant and positive correlation between the genomic copy number and RNA expression of FAL1(R=0.472,p=0.002;Fig-ure3D).It is also worth noting that several cell lines without FAL1 amplification express high-level FAL1RNA.This observation suggests that FAL1RNA overexpression may be a common phenomenon in cancer cells and that mechanisms other than genomic amplification are present to cause FAL1RNA overex-pression in cancer(Figure3D).To evaluate the clinical significance of FAL1in cancer, we characterized its expression and cellular location by in situ hybridization(ISH)using a FAL1-specific probe in a cohort of ovarian cancer specimens(n=181,including53early-stage cases and128late-stage cases;Table S9).A FAL1-positive signal was detected in more than93%of the specimens. Although31.6%of the samples exhibited a strong signal, 37.5%and23.9%had intermediate and weak signals,respec-tively(Figure3E).FAL1-positive samples also exhibited a nu-clear-enriched staining pattern,with a weak signal in cytoplasm. Similar staining patterns were also observed in cancer cell lines. We also characterized subcellular localization of FAL1by cell fractionation followed by qRT-PCR and observed the majority of FAL1RNA in the nuclear(Figure S3).Next,we measured FAL1RNA expression and genomic copy number using qPCR in ovarian tumors and found that both the FAL1RNA expression and genomic copy number in late-stage tumors were signifi-cantly higher than those in early-stage tumors(Figures3F and 3G).Consistent with the observation from cell lines,there was a strong and positive correlation between FAL1RNA expression and its genomic copy number in the ovarian tumor specimens (R=0.577,p<0.001;Figure3H).After stratifying the128late-stage ovarian cancer patients with FAL1RNA expression(cutoff, median expression)or gene amplification status,we found that both higher expression of FAL1RNA and genomic gain of FAL1gene were significantly associated with decreased survival in patients(p<0.0001and p=0.03,respectively;Figure3I). Taken together,these clinicalfindings demonstrated that gene amplification and RNA overexpression of FAL1occur frequently in epithelial cancer and are both associated with tumor progres-sion in ovarian cancer.FAL1Associates with BMI1Protein and RegulatesIts StabilityTo explore the molecular mechanisms underlying the oncogenic activity of FAL1,we sought to use RNA pull-down assay to iden-tify proteins associated with FAL1.Briefly,biotinylated full-length FAL1or antisense transcript(negative control)synthesized by in vitro transcription was incubated with the nuclear lysate from A2780cells,and coprecipitating proteins were isolated with streptavidin-agarose beads(Figure4A).The RNA-associ-ating proteins were resolved on SDS-PAGE gel,and the bands specific to FAL1were identified.BMI1,a37kD core subunit of the polycomb repressive complex1(PRC1)(Schuettengruber et al.,2007),was initially identified as a protein that was present only in FAL1-associated samples.To validate the association between BMI1and FAL1,we subjected the lncRNA-pull-down protein samples to western blot with BMI1antibody.A strong signal was observed in proteins pulled down with FAL1RNA but not in samples bound with either antisense FAL1or an unre-lated fragment of HOTAIR(Figure4B).To further confirm the interaction between FAL1and BMI1,we performed an RNA-immunoprecipitation(RNA-IP)assay,in which the RNA-BMI1 complex was immunoprecipitated using a BMI1antibody.The amount of FAL1RNA in the coprecipitate was then measured by pared with the immunoglobulin G(IgG)-bound sample,the BMI1-antibody-bound complex had a significant in-crease in the amount of FAL1RNA(Figures4C and4D).As nega-tive controls,we also quantified the levels of two unrelated lncRNAs,ENST00000457448and HOTAIR,in the complexes coprecipitated by IgG or the BMI1antibody.No significantCancer Cell Oncogenic lncRNA FAL1Represses p21Expression348Cancer Cell26,344–357,September8,2014ª2014Elsevier Inc.Figure 3.Characterization of FAL1Copy Number and RNA Expression in Cancers(A)SCNAs of FAL1locus in cancers.Focal amplicons were identified by GISTIC analysis (Tumorscape).(B)Copy-number profiles of chromosome 1q from breast and ovarian tumor specimens.Each sample is represented with a vertical line,and the positions of FAL1are noted with black horizontal lines.Red indicates gain;blue indicates loss.(C)Copy numbers of FAL1in cancer cell lines (n =99)were measured by qPCR.(D)A correlation between FAL1gene copy number and RNA expression was observed in 40cell lines.(E)FAL1expression visualized by ISH in ovarian cancer.(F)FAL1expression levels in early-and late-stage ovarian cancer specimens.(G)Copy number of FAL1in the same cohort.(H)A correlation between FAL1copy number and expression was observed in ovarian cancer specimens.(I)Survival curves of late-stage ovarian cancer patients with high and low FAL1RNA expression (top)or different genomic SCNA status (bottom).See also Figure S3and Tables S8and S9.Cancer CellOncogenic lncRNA FAL1Represses p21ExpressionCancer Cell 26,344–357,September 8,2014ª2014Elsevier Inc.349Figure 4.FAL1Associates with the BMI1Protein and Regulates Its Stability(A)A schematic representation of RNA pull-down.(B)Western blot of BMI1expression in 5%input and protein complexes pulled down by FAL1,antisense control,or unrelated control HOTAIR fragment from nuclear extracts.(C)A schematic representation of an RNA immunoprecipitation assay.(D)Results from RNA-IP and subsequent qRT-PCR assays.(Top)Relative quantification of FAL1,HOTAIR ,and ENST00000457448in RNA-protein complexes immunoprecipitated with IgG or BMI1antibodies from nuclear extracts.(Bottom)Representative western blot of BMI1in the corresponding samples.(E)Deletion mapping of BMI1-binding domain in FAL1.(Left)The schematic diagram of full-length and deleted fragments of FAL1;(right top)in vitro transcribed full-length and deleted fragments of FAL1showing correct sizes;(right bottom)western blot of BMI1in protein samples pulled down by different FAL1fragments.(F)Expression of FAL1and BMI1in A2780(left)and MCF-7(right)cells expressing shRNAs targeting these two genes.(G)The expression of Ring1A,Ring1B,and ubiqintination of H2AK119in A2780(left)and MCF7(right)cells expressing control and FAL1shRNAs.(H)The levels of BMI1,Ring1A,and Ring1B in the cytoplasmic fraction,the soluble nuclear fraction,the chromatin-bound insoluble nuclear fraction of A2780cells expressing control and FAL1shRNAs.Tubulin and H3were used as cytoplasmic and chromatin-bound loading controls,respectively.(I)The expression levels of BMI1,Ring1A,and Ring1B in control and FAL1knockdown cells treated with CHX.(J)Western blot (left)and quantification (right)of BMI1expression in control and FAL1knockdown cells treated with vehicle control or MG132.(K)Western blot of BMI1-associated ubiquitination in control and FAL1knockdown cells treated with MG132.Error bars indicate SD.*p <0.05.See also Figure S4.Cancer CellOncogenic lncRNA FAL1Represses p21Expression350Cancer Cell 26,344–357,September 8,2014ª2014Elsevier Inc.enrichment of either RNA was observed in the BMI1complex (Figure4D).Furthermore,using a series of deletion-mapping an-alyses,we identified a116nt region in the middle of the FAL1 transcript(nt296–411)as a major BMI1-binding domain,which is both required and sufficient for FAL1-BMI1association(Fig-ure4E).Taken together,these results demonstrate that BMI1 is a FAL1-associated protein.Next,we explored the molecular consequences of FAL1-BMI1 association.Although downregulation of BMI1mRNA expres-sion via BMI1shRNAs had no effect on FAL1RNA levels,ex-pressing FAL1shRNAs significantly reduced the protein level, but not the mRNA level,of BMI1(Figure4F).The level of two other PRC1core proteins,Ring1A and Ring1B,were similar in control and FAL1-knockdown cells,and the level of ubH2AK119 was much lower in FAL1-knockdown cells than in control cells (Figure4G).Although we detected a weak signal of Ring1B in the FAL1-protein complex from lncRNA pull-down assay,the signal of BMI1in FAL1-protein complex was much stronger than that of Ring1B,and FAL1-mediated pull-down significantly enriched BMI1but not Ring1B protein(Figure S4A).This obser-vation suggests that the FAL1-BMI1association may help spe-cifically stabilize BMI1protein.Additionally,we fractionated con-trol and FAL1-knockdown cells and analyzed the protein levels of BMI1,Ring1A,and Ring1B in the cytoplasm,the soluble nu-clear fraction,and the insoluble,chromatin-bound fraction.As shown in Figure4H,there was a marked decrease of chro-matic-bound BMI1,Ring1A,and Ring1B proteins in FAL1-knockdown cells than in controls.Concomitantly,there was a slight increase of these three PRC1proteins in the soluble nu-clear fraction in FAL1-knockdown cells than in control cells(Fig-ure4H).To further explore the mechanism of FAL1-mediated BMI1regulation,we treated A2780cells with cycloheximide (CHX)and analyzed the stabilities of BMI1,Ring1A and Ring1B in response to FAL1downregulation.Although the half-lives of Ring1A and Ring1B were not significantly affected by FAL1 knockdown,the half-life of BMI1was much shorter in FAL1 knockdown cells than in controls(Figure4I).The half-life of MDM2,a protein unrelated to PRC1complex,was not affected by FAL1knockdown,suggesting that FAL1shRNA expressions does not affect protein half-lives globally(Figure S4B).In agree-ment with this observation,when MG132was added into the cul-ture medium to inhibit proteasome degradation,the endogenous BMI1protein expression in FAL1knockdown cells was signifi-cantly increased and reached a level that was comparable to that in control cells(Figure4J),and higher BMI ubiquitination levels were also observed in FAL1knockdown cells treated with MG132(Figure4K).In aggregate,these observations sug-gested that FAL1expression is important in regulating BMI1pro-tein stability.FAL1overexpression led to higher expression of BMI1protein in HOSE cells(Figure S4C).Consistently,compared with control cells,FAL1-overexpressing cells had higher level of H2AK119 ubiquitination(Figure S4C).Further analysis on different subcel-lular fractions revealed that FAL1-expressing cells had higher BMI1expression in all different fractions than control cells(Fig-ure S4D).Interestingly,in response to FAL1overexpresion,there were also slight increases of Ring1A and Ring1B protein in the whole-cell lysates and in different subcellular fractions.How-ever,the changes in Ring1A/B were to a much lesser extent than that in BMI1(Figure S4C).Together,thesefindings suggest that the primary function of FAL1is to stabilize BMI1,and BMI1 stabilization,we reason,may further stabilize the whole PRC1 complex,therefore causing increases in the levels of other PRC1core proteins.FAL1Regulates the Transcription of a LargeSet of GenesBMI1is part of the PRC1complex,a well-characterized chro-matin-modifying complex that represses the transcription of a wide range of genes(Schuettengruber et al.,2007).Given that FAL1can bind to and stabilize BMI1and that FAL1expression alteration changed the level of H2AK119ubiquitination,we reasoned that FAL1expression alteration may influence BMI1 activity,which in turn can lead to genome-wide alterations in transcription.To test this hypothesis,we analyzed the RNA expression profiles of A2780cells expressing shRNAs targeting either FAL1or BMI1.Two independent shRNA hairpins were used for each target gene to avoid off-target effects.The tran-scription of732genes(represented by1,015probes)was upre-gulated by the expression of both BMI1shRNAs in A2780cells. In support of our hypothesis,we found that knocking down FAL1 induced transcriptional alterations in a wide range of genes, including887genes(represented by1,019probes)whose expression was upregulated by both FAL1shRNAs(Figure5A). Intriguingly,the expression of641of the1,019FAL1-induced probes(62.9%)was also increased by at least one of the BMI1 shRNAs,with285(28%)probes induced by both BMI1shRNAs (Figure5A).Only59(5.8%)probes were upregulated by FAL1 knockdown but downregulated by the expression of at least one BMI1shRNA;within these59probes,only four(0.4%) were downregulated by both BMI1shRNAs(Figure5A).The high degree of similarity between FAL1-and BMI1-mediated transcriptional repression strongly indicates a functional interac-tion between FAL1and BMI1,and the285probes whose expres-sions was upregulated by all four hairpins(Figure5A)may be a common set of target genes shared by FAL1and BMI1.To explore the functional processes that are affected by FAL1-mediated transcriptional regulation,we performed gene onto-logy(GO)analysis on the887genes that were upregulated by the knockdown of FAL1.The most significantly overrepresented biological processes included pathways involved in cell prolifer-ation,death,and survival,as well as cellular movement and pro-tein degradation(Figure5B;Table S10).For example,genes involved in cell-cycle arrest and apoptosis,such as CDKN1A, FAS,BTG2,TP53I3,FBXW7,and CYFIP2,were found to be significantly upregulated by both FAL1and BMI1knockdown in the above array studies.The increased expression of these six target genes was further validated by qRT-PCR(Figures5A and5C).Given that the PRC1complex regulates gene transcrip-tion by binding to promoter regions and modifying chromatin,we examined whether FAL1knockdown affected BMI1occupancy of the promoter regions in these target genes.The effect of FAL1knockdown on the occupancy of BMI1or ubiquitination levels of H2AK119in the target gene promoters was evaluated using a chromatin immunoprecipitation(ChIP)assay followed by qPCR.Among the six target genes tested,BMI1occupancy and ubiquitinated H2AK119were validated in the promoter regions offive genes,and knocking down FAL1significantlyCancer CellOncogenic lncRNA FAL1Represses p21ExpressionCancer Cell26,344–357,September8,2014ª2014Elsevier Inc.351。

2013-cancer cell-Chromatin-Bound IκBα RegulatesPolycomb Target Genes in Differentiation and Cancer

2013-cancer cell-Chromatin-Bound IκBα RegulatesPolycomb Target Genes in Differentiation and Cancer

Cancer CellArticleChromatin-Bound I k B a Regulates a Subset of Polycomb Target Genes in Differentiation and CancerMarı´a Carmen Mulero,1Dolors Ferres-Marco,2Abul Islam,3,4Pol Margalef,1Matteo Pecoraro,5Agustı´Toll,6Nils Drechsel,8Cristina Charneco,8Shelly Davis,9Nicola´s Bellora,3Fernando Gallardo,6Erika Lo ´pez-Arribillaga,1Elena Asensio-Juan,1Vero´nica Rodilla,1Jessica Gonza ´lez,1Mar Iglesias,7Vincent Shih,10M.Mar Alba `,3,11Luciano Di Croce,5,11Alexander Hoffmann,10Shigeki Miyamoto,9Jordi Villa`-Freixa,8,12Nuria Lo ´pez-Bigas,3,11William M.Keyes,5Marı´a Domı´nguez,2Anna Bigas,1,13and Lluı´s Espinosa 1,13,*1Program in Cancer Research,Institut Hospital del Mar d’Investigacions Me`diques (IMIM),Barcelona 08003,Spain 2DevelopmentalNeurobiology,Instituto de Neurociencias de Alicante,CSIC-UMH,Alicante 03550,Spain3Research Program on Biomedical Informatics,Universitat Pompeu Fabra,IMIM-Hospital del Mar,Barcelona 08003,Spain 4Department of Genetic Engineering and Biotechnology,University of Dhaka,Dhaka 1000,Bangladesh 5Gene Regulation,Stem Cells and Cancer,Centre de Regulacio ´Geno `mica (CRG),Barcelona 08003,Spain 6Dermatology Department 7Pathology DepartmentHospital del Mar,Barcelona 08003,Spain8Computational Biochemistry and Biophysics Laboratory,IMIM-Hospital del Mar and Universitat Pompeu Fabra,Barcelona 08003,Spain 9McArdle Laboratory for Cancer Research,University of Wisconsin Carbone Cancer Center,University of Wisconsin-Madison,6159Wisconsin Institute for Medical Research,1111Highland Avenue,Madison,WI 53705,USA 10Signaling Systems Laboratory,UCSD,La Jolla,CA 92093-0375,USA 11Institucio ´Catalana de Recerca i Estudis Avanc ¸ats (ICREA),Barcelona 08003,Spain 12Escola Polite `cnica Superior (EPS),Universitat de Vic,Barcelona 08500,Spain 13These authors contributed equally to this work *Correspondence:lespinosa@imim.es/10.1016/r.2013.06.003SUMMARYHere,we demonstratethat sumoylated and phosphorylated of keratinocytes and interacts with histones H2A and H4at the regulatory region of HOX and IRX genes.Chromatin-bound I k B a modulates Polycomb recruitment and imparts their competence to be activated by TNF a .Mutations in the Drosophila I k B a gene cactus enhance the homeotic phenotype of Polycomb mutants,which is not counteracted by mutations in dorsal/NF-k B .Oncogenic trans-formation of keratinocytes results in cytoplasmic I k B a translocation associated with a massive activation of Hox .Accumulation of cytoplasmic I k B a was found in squamous cell carcinoma (SCC)associated with IKK activation and HOX upregulation.INTRODUCTIONNF-k B plays a crucial role in biological processes,such as native and adaptive immune responses,organ development,cell proliferation,apoptosis,or cancer (Naugler and Karin,2008;Vallabhapurapu and Karin,2009).NF-k B activation de-pends on the IKK-mediated degradation of the NF-k B inhibitors,I k B proteins,that takes place in the cytoplasm and results in the translocation of the NF-k B transcription factor to the nucleus,where it activates gene expression.Recent studies demonstrate the existence of alternative nuclear functions for regulatory ele-ments of the pathway (reviewed in Espinosa et al.,2011),but their biological implications remain poorly understood.Recently,it has been demonstrated that nuclear I k B b binds the promoterCancer Cell 24,151–166,August 12,2013ª2013Elsevier Inc.151角质形成细胞Cancer CellI k B a Is a Modulator of Polycomb Function(legend on next page) 152Cancer Cell24,151–166,August12,2013ª2013Elsevier Inc.of NF-k B target genes following lipopolysaccharide (LPS)stimu-lation to prevent I k B a -mediated inactivation,thereby sustaining cytokine expression in immune cells (Rao et al.,2010).Numerous studies have reported nuclear translocation of I k B a (Aguilera et al.,2004;Arenzana-Seisdedos et al.,1997;Huang and Miyamoto,2001;Wuerzberger-Davis et al.,2011)and various partners for nuclear I k B a ,including histone deacetylases (HDACs)and nuclear corepressors,have been identified (Agui-lera et al.,2004;Espinosa et al.,2003;Viatour et al.,2003).In fibroblasts,nuclear I k B a associates with the promoter of Notch target genes correlating with their transcriptional repression,which is reverted by TNF a (Aguilera et al.,2004).Nevertheless,the mechanisms that regulate association of I k B to the chromatin and its repressive function remain unknown.I k B a -deficient mice die around day 5because of skin inflam-mation associated with high levels of IL1b and IFN-g in the dermis,CD8+T cells,and Gr-1+neutrophils infiltrating the epidermis,as well as altered keratinocyte differentiation (Beg et al.,1995;Klement et al.,1996;Rebholz et al.,2007),similar to keratinocyte-specific I k B a -deficient mice (I k B a k5D /k5D )(Re-bholz et al.,2007).In all cases,disruption of TNF a signaling rescued the skin phenotype (Shih et al.,2009),suggesting that lethality was associated with an excessive inflammatory response,likely due to increased NF-k B activity.However,mice expressing different I k B a mutants that are equally able to repress NF-k B in the skin showed divergent phenotypes.Specifically,mice expressing the nondegradable I k B a mutant,I k B a S32-36A ,developed skin tumors resembling SCC (van Hoger-linden et al.,1999),whereas mice carrying a predominantly nuclear form of I k B a show no overt skin defects (Wuerzberger-Davis et al.,2011).Skin differentiation depends on the correct establishment and maintenance of specific gene expression patterns,including genes of the HOX family,which in the basal progenitor cells are repressed by EZH2,the catalytic subunit of the Polycomb repressive complex 2(PRC2)(Ezhkova et al.,2009,2011).PRC2is composed by EZH2,the WD-repeat protein EED,RbAp48,and the zinc-finger protein SUZ12(Zhang and Reinberg,2001).Methylation of lysine 27on histone H3(H3K27me3)by EZH2imposes gene silencing in part by trig-gering recruitment of PRC1(Cao et al.,2002;Min et al.,2003)and histone deacetylases (HDACs).Here,we investigate analternative function for I k B a in the regulation of skin homeosta-sis,development,and cancer.RESULTSPhosphorylated and Sumoylated I k B a Localizes in the Nucleus of KeratinocytesTo investigate the physiological role for nuclear I k B a ,we per-formed an initial screen to determine its subcellular distribution in human tissues.We found that I k B a localizes in the cytoplasm of most tissues and cell types as expected (Figure S1A available online);yet,a distinctive nuclear staining of I k B a was found in human (Figure 1A)and mouse skin sections (Figures 1A,S1A,and S1C),more prominently in the keratin14+basal layer kerati-nocytes.I k B a distribution became more diffused in the supra-basal layer of the skin and gradually disappeared in the more differentiated cells.Specificity of nuclear I k B a staining was confirmed using skin sections from newborn I k B a -knockout (KO)mice (Figure S1B)and different anti-I k B a antibodies and blocking peptides (Figure S1C).By immunofluorescence (IF)and immunoblot (IB),we detected I k B a protein in both the cyto-plasmic and the nuclear/chromatin fractions of human (Figures 1B and 1C)and mouse (Figure S1D)keratinocytes.Interestingly,nuclear I k B a displayed a shift in its electrophoretic mobility (z 60kDa)detected by different anti-I k B a antibodies,including the anti-phospho-S32-36-I k B a antibody.We next precipitated I k B a from nuclear and cytoplasmic keratinocyte extracts and determined whether this low I k B a mobility was a result of ubiq-uitin or SUMO modifications.We found that nuclear I k B a was specifically recognized by anti-SUMO2/3,but not anti-SUMO1or anti-ubiquitin antibodies (Figure 1D;data not shown).Here-after,we will refer to this nuclear I k B a species as phospho-SUMO-I k B a (PS-I k B a ).By cotransfection of different SUMO plasmids in HEK293T cells,we demonstrated that SUMO2was integrated to HA-I k B a at K21,22(Figure S1E),independently of S32,36phosphorylation (Figure 1E).By subcellular fractionation,we found that most HA-PS-I k B a was distributed in the nucleus of HEK293T cells (data not shown),and both K21,22R and S32,36A I k B a mutants showed reduced association with the chromatin (Figure 1F).These results suggest that phosphorylation and sumoylation are both required for I k B a nuclear functions in vivo.Of note,PS-I k B a levels were always low in HEK293T cells whenFigure 1.Phosphorylated and Sumoylated I k B a Is Found in the Nucleus of Normal Basal Keratinocytes(A)Immunodetection of I k B a (green)in normal human skin and detail of basal layer.B,basal;S,spinous,G,granular;and C,cornified layers of epidermis.Dashed line indicates the dermis interphase.DAPI was used for nuclear staining.(B)IF of I k B a in primary human keratinocytes.(C)Subcellular fractionation of human keratinocytes followed by IB with the indicated antibodies.(D)I k B a was immunoprecipitated from primary murine keratinocyte extracts followed by IB with the indicated antibodies.(E)IB analysis of His-tag precipitates from HEK293T cells transfected with the indicated plasmids.SUMO2is incorporated in I k B a when K21,22are present.(F)HEK293T cells were transfected with the indicated I k B a plasmids and processed following the ChIP protocol to obtain the whole chromatin fraction that was analyzed by IB.(G)IF of I k B a and P-IKK in skin sections.Cells with P-IKK staining do not contain nuclear I k B a .(H)IB analysis of keratinocytes transduced with myc-IKK a EE or control.(I)IF analysis of the indicated differentiation markers in skin sections of WT and I k B a KO newborn mice.(J)IB analysis of indicated proteins in control or Ca 2+-treated murine keratinocytes.Total and nuclear/chromatin fractions are shown.(K)Determination of Filaggrin ,K10,and p63mRNA levels in control and I k B a KD keratinocytes following Ca 2+treatment.Expression levels are relative to Gapdh and compared to control cells.Error bars indicate SD.I k B a protein levels were analyzed by IB.Data correspond to one representative of three experiments.N,nuclear;C,cytoplasmic.See also Figure S1.Cancer CellI k B a Is a Modulator of Polycomb FunctionCancer Cell 24,151–166,August 12,2013ª2013Elsevier Inc.153Cancer CellI k B a Is a Modulator of Polycomb FunctionFigure2.I k B a Binds Histones H2A and H4(A)PD experiment using GST-I k B a and native(lane2)or denatured-renatured(lanes3–4)human keratinocyte nuclear extracts.One representative gel stained with Coomassie blue is shown(n=3).(B)Purification and analysis of B and C bands identified as histones H2A and H4by mass spectrometry.Table indicates the number of peptides identified and their score factor.The highest score is highlighted.(C)Coprecipitation from DSP-crosslinked nuclear extracts from human keratinocytes.(D and E)PD using different GST-H2A proteins and total lysates from HEK293T cells expressing the indicated proteins.(legend continued on next page) 154Cancer Cell24,151–166,August12,2013ª2013Elsevier Inc.compared with keratinocytes,even in overexpression conditions and cell lysates directly obtained under denaturing conditions (see inputs in Figures 1E and S1E).It is well known that IKK activity regulates the cytoplasmic levels of I k B a .By double staining of skin sections,we found that the few cells that were positive for active IKK contained I k B a ,but this I k B a was excluded from the nucleus (Figure 1G).To directly investigate whether IKK regulates subcellular distri-bution of I k B a ,we transduced primary murine keratinocytes with lentiviral IKK a EE .We found that active IKK a induced a decrease in the nuclear levels of PS-I k B a as determined by IB (Figure 1H)and IF (Figure S1G).Additional experiments comparing the effects of both IKK isoforms demonstrated that IKK a EE was more efficient than IKK b EE in decreasing nuclear PS-I k B a levels (60%±5%compared with 16%±9%reduction;p <0.001)(Figure S1F).To directly address whether I k B a was required for normal skin differentiation,we performed IF analysis using different markers comparing I k B a wild-type (WT)and KO newborn skins.Consis-tent with previous reports,I k B a -deficient mice do not show any obvious skin defect at birth with a normal K5-positive basal layer,although we observed a slight reduction in the thickness of the K10-positive suprabasal epidermal layer.Most importantly,I k B a mutant skins showed a severe reduction of the more differ-entiated layer of cells identified by the accumulation of filaggrin granules (Figure 1I).This is a cause for impaired barrier function (Palmer et al.,2006).Next,we aimed to distinguish between cell-autonomous and non-cell-autonomous effects of I k B a defi-ciency by using an in vitro system for keratinocyte differentiation induced by high Ca 2+exposure (Hennings et al.,1980).In this model,we found that keratinocyte differentiation was associated with a decrease in both I k B a and PS-I k B a levels and activation of nuclear IKK (Figures 1J and S1H).Notably,knockdown (KD)of I k B a disturbs in vitro keratinocyte differentiation as indicated by the impaired K10and filaggrin induction in response to Ca 2+,which was accompanied by sustained expression of the progenitor marker p63(Figure 1K).Together,these results strongly suggest that I k B a plays a cell-autonomous function in skin differentiation.I k B a Directly Binds to the N-Terminal Tail of Histones H2A and H4To further investigate the mechanisms underlying nuclear I k B a functions,we searched for nuclear proteins that directly asso-ciate with I k B a .Using GST-I k B a and human keratinocyte nuclear extracts in pull-down (PD)experiments,we isolated pro-teins of estimated molecular weights of 15and 14kDa that were identified by mass spectrometry as histones H2A and H4(Fig-ures 2A and 2B).Interaction between histones H2A and H4and I k B a was further confirmed by coprecipitation of endoge-nous proteins from keratinocyte nuclear extracts.Of note,the NF-k B subunit p65was absent from nuclear I k B a precipitates but coprecipitated in the cytoplasmic fraction (Figure 2C).By PD assays,we determined the specificity of I k B a binding compared to other I k B homologs (Figure 2D)and mapped the I k B a -binding domain of histone H2A to be between amino acids 2and 35(Figure 2E).Preincubation of I k B a with p65prevented its association with histones (Figure S2A),suggestive of mutually exclusive complexes.Comparative sequence analysis of the I k B a -binding region of histone H2A (AA1–36)and the homologous region of H4revealed the presence of a motif (3KXXXK/R)that was absent from other histone and nonhistone proteins (Figure 2F).To further study I k B a binding specificity,we screened a histone peptide array using nuclear HA-I k B a expressed in HEK293T cells as bait.We found that I k B a bound to peptides containing AA11–30of his-tone H4,but not the corresponding region of histone H3.Most importantly,binding of I k B a to H4was prevented by the combi-nation K12/K16Ac and K20Ac or Me2(Figure 2G).Because the equivalent peptides from histone H2A were not included in the array,we performed parallel precipitation experiments using biotin-tagged peptides (AA5–23)of histone H2A and H4(Figures 2H and 2I).We found that I k B a association was prevented by K12Ac,K16Ac,and K20me2of histone H4or the equivalent modifications of the H2A peptide (Figure 2H)and also when all K/R residues in the 3KXXXK motif were changed into A (Figure 2I).Of note that in these experiments histone-bound HA-I k B a was mostly identified as a nonsumoylated band,which opens the possibility that posttranslation modifications are not essential to mediate this interaction in vitro.However,parallel binding experiments using keratinocyte extracts,PS-I k B a ,showed a preferential binding to the histone peptides compared with the cytoplasmic 37kDa I k B a form (Figure S2B).Together,these re-sults strongly suggest that only PS-I k B a can bind the chromatin,but in HEK293T cells this molecule is then desumoylated in vivo or as a consequence of the experimental processing.To gain further insights into the molecular basis of I k B a binding to histones,we completed the structure of I k B a obtained from the Protein Data Bank (ID code 1IKN),which lacked part of the ankyrin repeat (AR)1,using RAPPER (Depristo et al.,2005)and performed docking studies with AutoDock Vina (Trott and Olson,2010)of the histone H4peptide,GKGGAKRHRKV,that contains most of the KXXXK domain.Docking calculations showed two deep pockets for K interaction in I k B a located between ARs 1-2and 2-3and an additional shallower patch between AR3and 4.Overall,the peptide bound in a clearly nega-tive region on the I k B a surface (Figure 2J),with higher affinity than the modified peptide that was acetylated in the first and(F)ClustalW alignment of the conserved KXXXK/R motifs in the N terminus of histones H2A and H4.Conserved K and R residues are in green.Red triangle indicates the last AA included in GST-H2A 2-35.(G)Histone peptide array hybridized with nuclear HEK293T extracts expressing HA-I k B a .One informative area of the blot image and the relative binding of selected peptides are shown.(H)Coprecipitation of cytoplasmic and nuclear HA-I k B a expressed in HEK293T cells with the indicated histone H2A and H4peptides.(I)HA-I k B a was precipitated using the indicated histone H2A peptide,the K/A mutant,or scrambled peptide.In (G),(H),and (I),cell lysates were denatured-renatured previous to the precipitation to disrupt preformed complexes.(J)Model for binding of the histone H4peptide (unmodified or modified)to consecutive ankyrin repeats of I k B a (3KXXXK).See also Figure S2.Cancer CellI k B a Is a Modulator of Polycomb FunctionCancer Cell 24,151–166,August 12,2013ª2013Elsevier Inc.155Cancer CellI k B a Is a Modulator of Polycomb FunctionFigure3.Analysis and Identification of I k B a Target Genes(A)Developmentally related genes selected from I k B a targets identified in ChIP-seq analysis.Fold change over the random background is indicated.(B)Functional enrichment of target genes with p value cutoff%10À5based on gene ontology(GO)as extracted from Ensembl database using GiTools.Enriched categories are represented in heatmap with the indicated color-coded p value scale.(legend continued on next page) 156Cancer Cell24,151–166,August12,2013ª2013Elsevier Inc.second K residues (K12and K16).We experimentally validated that ARs of I k B a participate in histone binding because the I k B aD 55-106mutant,lacking part of AR1,failed to bind GST-H2A (Figure S2C).Similarly,this association was prevented by 1%deoxycholate (Figure S2D),as described for interactions involving the ARs of I k B a (Baeuerle and Baltimore,1988;Savi-nova et al.,2009).I k B a Is Specifically Recruited to the Regulatory Regions of Developmental GenesTo identify putative PS-I k B a target genes,we performed chromatin immunoprecipitation sequencing (ChIP-seq)using chromatin extracts from primary human keratinocytes and anti-I k B a antibody.We identified 2,778enriched peaks,correspond-ing to 2,433Ensembl genes that were significantly enriched with p values %10À5.Gene ontology analysis showed that a signifi-cant proportion of genes participate in biological processes associated with embryonic development and cell differentiation.I k B a targets included genes of the HOX and IRX families,ASCL4,CDX2,NEUROD4,OLIG3,and NEURL ,among others (Figures 3A and 3B).Annotation of the peak genomic positions to the closest gene demonstrated that many peaks were positioned immediately after the transcription start site (TSS),with a sharp decrease near the transcription termination site (TTS)(Figure 3C),whereas others were located far from promoter regions.Some of the latter overlapped with regions enriched in H3K4me1,a his-tone mark associated with enhancer regions (data not shown).Randomly selected I k B a targets were confirmed by conventional chromatin immunoprecipitation (ChIP)using primers flanking the regions identified in the ChIP-seq experiment (Figure 3D).Consistent with its overall effects on I k B a levels,sustained Ca 2+treatment caused the loss of I k B a from all tested gene pro-moters (Figure 3E).Similarly,short treatment with TNF a released chromatin-bound I k B a in keratinocytes,as we previously found in fibroblasts (Aguilera et al.,2004).However,we did not detect a general effect of TNF a on PS-I k B a levels,but we consistently found a partial redistribution of PS-I k B a to the soluble nuclear fractions (Figure S3A).Next,we investigated whether TNF a and Ca 2+modulated HOX and IRX transcription in keratinocytes.All tested I k B a targets (n =12)were robustly induced by TNF a treatment (up to 12-fold)following different kinetics (Figures 3F)and to a lesser extent (up to 3-fold)by Ca 2+treatment (Fig-ure S3B)or I k B a KD (Figure S3C).Interestingly,1hr of TNF a treatment prevented Ca 2+-induced differentiation of murine keratinocytes (Figure S3D),supporting the notion that PS-I k B a integrates inflammatory signals with skin homeostasis (see Dis-cussion ).We also tested whether p65participated in HOX or IRX gene activation by TNF a .By ChIP analysis,we did not find anyrecruitment of p65to I k B a target genes after TNF a treatment,in contrast to a canonical NF-k B target gene promoter (Fig-ure S3E).However,we detected low amounts of p65at HOX genes under basal conditions that might contribute to gene repression (Dong et al.,2008),although the function of chromatin-bound p65at regions distant from the TSS of both NF-k B targets and nontargets is unresolved.Binding of p65to HOX and IRX was reduced after TNF a treatment,suggesting that p65was redistributed from noncanonical to canonical NF-k B targets once activated.Silencing of HOX genes in keratinocytes involves PRC2and its core component the H3K27methyltransferase EZH2(Ezhkova et al.,2009;Mejetta et al.,2011).To explore a putative associa-tion between I k B a and PRC2function,we crossed our list of 2,433I k B a targets with available ChIP-seq data from keratino-cytes.Approximately 50%of I k B a targets corresponded to genes enriched for the H3K27me3mark (Figures 3G and S3F),although I k B a targeted only 13%of the H3K27trimethylated genes.Most importantly,genomic sequences occupied by I k B a essentially overlapped with those regions containing high H3K27me3levels (Figure 3G).We also found a statistically signif-icant overlap (p <10À16)between I k B a target genes and PRC tar-gets in ES cells (Birney et al.,2007;Ku et al.,2008)(Figure S3G).I k B a Interacts with and Regulates Association of PRC2to Target Genes in Response to TNF aIn the mass spectrometry analysis of proteins that associate with GST-I k B a ,we identified a few peptides corresponding to chro-matin modifiers,such as EZH2and SUZ12,and SIN3A (Figure S4A).Specificity of I k B a interactions with PRC2elements,but also I k B a association to the PRC1protein BMI1,was confirmed by PD assays (Figure S4B).SUZ12was able to interact with nuclear I k B a ,whereas p65specifically associated with cyto-plasmic I k B a in the IP experiments (Figure 4A).Importantly,exogenous wild-type I k B a ,but not an I k B a mutant that failed to bind histones,facilitated the association of SUZ12to GST-H4(Figure 4B).Moreover,ChIP experiments demonstrated that TNF a treatment induced the dissociation of SUZ12from I k B a target regions,but not non-I k B a targets (Figure 4C).Sequential ChIP experiments demonstrated that I k B a and SUZ12simultaneously bound to I k B a target genes (Figure 4D).To test the functional relevance of I k B a in PRC-mediated repression,we used WT murine embryonic fibroblast (MEFs),which expressed detectable levels of PS-I k B a (Figure 4E)and I k B a KO MEFs.By ChIP-on-chip experiments using three different I k B a antibodies,we confirmed that several Hox genes were also targets of I k B a in MEFs (Table S1).By ChIP,we found that SUZ12and EZH2bound I k B a targets efficiently in WT MEFs(C)Graphs show the relative distance to the nearest ChIPed region,3kb upstream and downstream of the RefSeq gene’s TSS and TTS.(D)Validation of the identified DNA regions (À222to À200for HOXA10,À9,380to À9,360for HOXB2,+6,166to +6,186for HOXB5,+4,451to +4,471for HOXB3,and À18,820to À18,800for IRX3)by conventional ChIP.Amplification of 2kb distant regions was used as negative controls.(E)ChIP analysis of I k B a after 20min of TNF a or 48hr Ca 2+treatments.In (D)and (E),graphs represent mean enrichment relative to nonspecific immunoglobulin G (IgG)(n =2).(F)Expression levels of I k B a target genes following TNF a treatment analyzed by qRT-PCR.Gene represented is in black,whereas genes following the same kinetics are indicated in red.(G)ChIP-seq profiles of endogenous I k B a occupancy in three enriched loci (HOXA ,HOXB ,and IRX5)and one negative locus (JARID1B/KDM5B )compared to H3K27me3(from the UW ENCODE Project)in keratinocytes.(D–F)Bars represent mean,and error bars indicate SD.See also Figure S3.Cancer CellI k B a Is a Modulator of Polycomb FunctionCancer Cell 24,151–166,August 12,2013ª2013Elsevier Inc.157Figure 4.I k B a Interacts with and Regulates Association of PRC2in Response to TNF a(A)IB analysis of I k B a precipitates from nuclear and cytoplasmic primary murine keratinocyte extracts.Five percent of the input and 25%of the IP was loaded in all cases except for detection of I k B a input that represents 0.5%.(B)PD using GST-H4and cell lysates from HEK293T expressing different combinations of HA-I k B a and SUZ12.(C)Relative recruitment assessed by ChIP of SUZ12to different genes 40min after TNF a in primary murine keratinocytes.(legend continued on next page)Cancer CellI k B a Is a Modulator of Polycomb Function158Cancer Cell 24,151–166,August 12,2013ª2013Elsevier Inc.but only weakly in I k B a KO MEFs (Figure 4F,time 0).In WT cells,TNF a treatment induced a significant but temporary release of SUZ12and EZH2from these loci,which peaked after 30–60min of treatment (Figure 4F).The binding of PRC2proteins at Hox genes inversely correlated with the expression of these Hox genes (Figure 4G).In contrast,I k B a KO cells failed to activate Hox transcription in response to TNF a (Figure 4G),which is consistent with a defective release of PRC2proteins (Figure 4F).Unexpectedly,we did not detect changes in H3K27me3levels in these Hox genes upon TNF a treatment at any of the time points studied (20min,2hr,and 7hr)(data not shown),likely reflecting the high stability of this histone modifica-tion (De Santa et al.,2007).Supporting the possibility that activa-tion of I k B a targets is independent of the enzymatic activity of EZH2,a 24hr treatment with the EZH2inhibitor DZNep does not affect Hoxb8or Irx3messenger RNA (mRNA)levels in kera-tinocytes (data not shown).Together,these results suggest that transcriptional repression-activation of these genes does not strictly depend on EZH2enzymatic activity but rather PRC2release might modulate the dissociation of PRC1or HDACs (van der Vlag and Otte,1999)that associate with more dynamic chromatin modifications.In agreement with this possibility,his-tone H3is rapidly acetylated following TNF a treatment at different Hox gene promoters (Figure S4C).To further study the involvement of NF-k B in the regulation of I k B a targets by TNF a ,we attempted to use different mutant MEFs,including the p65,Ikk a ,Ikk b ,and the triple p65;p50;c-Rel KO.We found that TNF a induced Hox and Irx expression in both p65and Ikk b KO cells (Figure S4D)suggesting that it was NF-k B independent.However,specific mutants contained vari-able levels of I k B a and PS-I k B a (Figures S4E and S4F),which make a more accurate quantitative analysis unproductive.Inter-estingly,phosphorylation of nuclear I k B a was not reduced in the Ikk a or Ikk b KO cells (Figures S4F),indicating that other kinases are involved in generating PS-I k B a .Only triple KO cells,which essentially lacked I k B a (Figure S4G)and the Ikk a -deficient MEFs,showed a strong defect on Hox and Irx transcription (Figures S4D and S4H).To better understand the contribution of NF-k B to Hox regulation,we next performed luciferase re-porter assays measuring the ability of different I k B a mutant pro-teins to repress a Hoxb8-promoter construct compared with a reporter containing three consensus sites for NF-k B (3x k B).We found that WT I k B a and the nuclear I k B a NES mutant (Huang et al.,2000)significantly repressed both promoters.However,mutations that affect chromatin-association (Figure 1F)pre-vented Hoxb8repression but still inhibited the expression of the 3x k B reporter (Figure 4H).Consistently,Hoxb8mRNA levels were significantly reduced in the skin of mice expressing I k B a NES(Wuerzberger-Davis et al.,2011)(Figure 4I).Moreover,these animals showed an expansion of the K14-positive basal layer of keratinocytes containing nuclear I k B a (Figure 4J),associated with increased proliferation measured by ki67staining and impaired differentiation as indicated by the reduced thickness of the suprabasal K10-expressing layer (Figure 4K).Genetic Interaction between I k B a /cactus and polycomb in DrosophilaPolycomb group (PcG)I k B and NF-k B proteins are conserved from flies to humans.In addition,Drosophila contains one Hox cluster,compared with four clusters in vertebrates,which facili-tates studying genetic interactions.We first confirmed that the single Drosophila I k B homolog,cactus (cact)(Geisler et al.,1992),maintained the capacity to associate with histones (Figure 5A).By IF,we detected colocalization of cactus and Polycomb (Pc),a PRC1protein that is essential for the repressive PRC2function,in specific bands of polytene chromosomes (Figure 5B).Most of the cactus staining overlapped with Pc,but only a few Pc-positive bands contained cactus.Based on our mammalian data,our first attempt was to generate single PRC2mutants and combine them with cactus -deficient mutants.All mutants were tested in heterozygosis because homozygous mutations in cactus or PcG genes are lethal.We found that heterozygous mutations in PRC2genes (e.g.,the null mutations of E (z ))as well as the composed E (z )and cact exhibit no overt homeotic phenotypes.In contrast,hetero-zygous Pc mutants exhibit a variety of characteristic homeotic transformations,including the partial transformation of the sec-ond and third (mesothoracic and metathoracic)legs toward first (also known as prothoracic)legs that in males are characterized by the presence of sex combs.Modifications of this phenotype (also called ‘‘extra sex comb’’)have been extensively used as a functional assay to validate new PcG proteins in vivo.Thus,we tested the effect of reducing cact on Pc -induced homeotic transformation using 12recessive mutations of cact and two independently generated Pc alleles.All cact mutant alleles,but more prominently cact 1,enhanced the ‘‘extra sex comb’’pheno-type of Pc mutations (Pc 3and Pc XT109)(Figure 5C;Table S2).Because cells lacking cact exhibit massive nuclear localization of the transcription factor dorsal (dl,Drosophila NF-k B/Rel/p65ortholog)during postembryonic stages (Lemaitre et al.,1995),we tested whether enhancement of homeotic defect of Pc by cact mutations is due to increased dorsal activity.Because sta-bility of cact is under NF-k B/Dorsal control (Kubota and Gay,1995),we anticipated that for phenotypes due to increased dorsal,dl mutations would counteract cact mutations,whereas for phenotypes independent of dorsal,reducing dl should yield(D)Sequential ChIP using the indicated combinations of antibodies.An analysis of two different Hox regulatory regions is shown.(E)IB analysis of WT fibroblasts showing the presence of cytoplasmic and nuclear I k B a .(F)Relative chromatin binding of PRC2and I k B a in WT and I k B a KO MEFs treated with TNF a .ChIP values were normalized by IgG precipitation.(G)Relative levels of the indicated genes in WT and I k B a KO MEFs.(H)Luciferase assays to determine the effect of different I k B a constructs on the activity of HoxB8compared to the 3x k B reporter.Lower panels show expression levels of different constructs.(I)Expression levels of HoxB8in the skin of WT and I k B a NES/NES mice by qRT-PCR (n =2).(J and K)Analysis of skin sections from 7-to 8-week-old WT and I k B a NES/NES mice by IF.K14labels the basal layer keratinocytes (J).Immunostaining of ki67,the suprabasal marker K10,and filaggrin (K).Throughout the figure,bars represent mean,and error bars indicate SD.See also Figure S4and Table S1.Cancer CellI k B a Is a Modulator of Polycomb FunctionCancer Cell 24,151–166,August 12,2013ª2013Elsevier Inc.159。

专家点评CellRes刘兵兰雨团队解密人类造血干细胞起源

专家点评CellRes刘兵兰雨团队解密人类造血干细胞起源

专家点评CellRes刘兵兰雨团队解密人类造血干细胞起源造血干细胞(hematopoietic stem cell,HSC)能够在体内产生所有类型的血液细胞,并且通过自我更新和多系分化维持个体整个生命周期的血液系统功能。

虽然HSC的发生过程在斑马鱼和小鼠等动物模型已经被充分揭示,但受限于研究技术的缺乏和研究材料的稀缺,目前对人类早期胚胎造血发育的认识仍十分有限。

研究人员通过异种移植的功能分析,确认了人类的HSC于胚胎卡内基阶段(Carnegie stage,CS)14(妊娠后32天)最早出现在主动脉-性腺-中肾(aorta–gonad–mesonephros,AGM)区域。

与胎肝和脐带血中HSC类似,AGM区的HSC也具有CD45 CD34 表型【1,2】(图1)。

图1. 人类胚胎造血发育事件概览【2】功能学和组织学实验证实,人类胚胎的第一个HSC产生于背主动脉腹侧壁【3,4】。

鉴于人类HSC与血管内皮细胞存在着紧密的时空联系,并且共享一些表面标志【4】,因此推测人类HSC与小鼠类似,均起源于生血内皮细胞(hemogenic endothelial cell,HEC)【5】。

目前研究认为,HEC表达内皮基因和生血特异的转录因子RUNX1,但不表达典型造血表面标志如CD43和CD45【6,7】。

HEC的特化是血管内皮细胞选择HSC命运的关键步骤,然而,由于HEC数量稀少、发育过程转瞬即逝,对于内皮造血转化的认识,尤其是对HEC的精准识别,成为造血发育研究领域的重点和难点。

鉴于临床的巨大需求和血液细胞来源有限性之间的矛盾,血液系统尤其是HSC的再生研究显得尤为重要。

研究者一直致力于利用多能干细胞高效诱导功能性HSC,但罕见成功的报道,表明对于HSC胚胎发生的时序特征及调控机制的认识仍有待深入,HSC的再生研究依然任重道远【8】。

2019年9月9日,解放军总医院第五医学中心刘兵研究组与暨南大学基础医学院兰雨研究组合作在Cell Research杂志在线发表了题为Tracing the first hematopoietic stem cell generation in human embryo by single-cell RNA sequencing的研究论文。

Resveratrol Rescues SIRT1-Dependent Adult Stem Cell Decline and Alleviates Proge

Resveratrol Rescues SIRT1-Dependent Adult Stem Cell Decline and Alleviates Proge

Cell MetabolismArticleResveratrol Rescues SIRT1-DependentAdult Stem Cell Decline and AlleviatesProgeroid Features in Laminopathy-Based ProgeriaBaohua Liu,1,2,4,*Shrestha Ghosh,1Xi Yang,1Huiling Zheng,1,4,5Xinguang Liu,4,5Zimei Wang,1Guoxiang Jin,1Bojian Zheng,3Brian K.Kennedy,4,6Yousin Suh,4,7Matt Kaeberlein,4,8Karl Tryggvason,9and Zhongjun Zhou1,2,*1Department of Biochemistry,Li Ka Shing Faculty of Medicine2Shenzhen Institute of Research and Innovation3Department of Microbiology,Li Ka Shing Faculty of MedicineThe University of Hong Kong,Hong Kong4Institute of Aging Research,Guangdong Medical College,Dongguan523808,China5Key Laboratory for Medical Molecular Diagnostics of Guangdong Province,Dongguan523808,China6Buck Institute for Research on Aging,Novato,CA98945,USA7Departments of Medicine and Genetics,Albert Einstein College of Medicine,Bronx,NY10461,USA8Department of Pathology,University of Washington,Seattle,WA98195,USA9Medical Biochemistry and Biophysics,Karolinska Institute,Stockholm17177,Sweden*Correspondence:ppliew@hku.hk(B.L.),zhongjun@hkucc.hku.hk(Z.Z.)/10.1016/j.cmet.2012.11.007SUMMARYAbnormal splicing of LMNA gene or aberrant pro-cessing of prelamin A results in progeroid syndrome.Here we show that lamin A interacts with and acti-vates SIRT1.SIRT1exhibits reduced associationwith nuclear matrix(NM)and decreased deacetylaseactivity in the presence of progerin or prelaminA,leading to rapid depletion of adult stem cells(ASCs)in Zmpste24À/Àmice.Resveratrol enhancesthe binding between SIRT1and A-type lamins toincreases its deacetylase activity.Resveratrol treat-ment rescues ASC decline,slows down body weightloss,improves trabecular bone structure and mineraldensity,and significantly extends the life span in Zmpste24À/Àmice.Our data demonstrate lamin A as an activator of SIRT1and provide a mechanisticexplanation for the activation of SIRT1by resveratrol.The link between conserved SIRT1longevity path-way and progeria suggests a stem cell-based andSIRT1pathway-dependent therapeutic strategy forprogeria.INTRODUCTIONLamin A,encoded by the LMNA locus,is a major component of the nuclear matrix(NM),afilamentous nucleoskeleton important for nuclear structure maintenance.Chromatin and other proteins dynamically associate with the NM to regulate various nuclear activities,including replication,gene transcription,DNA repair, and chromatin organization(Tsutsui et al.,2005).NM copurifies with a majority of the nuclear histone deacetylase(HDAC) activity,thus regulating chromatin structure(Hendzel et al., 1991).A de novo G608G mutation in LMNA results in alternative which is the predominant cause of Hutchinson-Gilford Progeria Syndrome(HGPS),a severe form of early-onset premature aging (Eriksson et al.,2003).Mice deficient for Zmpste24,a metallopro-teinase responsible for prelamin A maturation,manifest many of the progeroid features resembling HGPS patients(Penda´s et al.,2002).We and others previously have shown that HGPS skinfibroblasts and mouse embryonicfibroblasts(MEFs) derived from Zmpste24À/Àembryos undergo early senescence, attributable to genomic instability and hyperactivation of the p53pathway,and that reduction of the prelamin A level in Zmpste24À/Àmice by Lmna heterozygosity ameliorates proge-roid phenotypes and significantly extends life span(Liu et al., 2005;Varela et al.,2005).Human cells engineered to express progerin exhibited premature senescence(Kudlow et al., 2008).One of the hallmarks of progeria cells is a misshaped nucleus,which leads to disorganized heterochromatin(Scaffidi and Misteli,2005)and mislocalized nuclear proteins(Krishnan et al.,2011;Liu et al.,2008;Manju et al.,2006;Scaffidi and Misteli,2008).Reducing membrane-bound prelamin A or pro-gerin by farnesyl transferase inhibitor rescues the nuclear shape abnormality and ameliorates progeroid features(Fong et al., 2006;Gordon et al.,2012).NAD+-dependent protein deacetylase SIRT1,the closest homolog of Saccharomyces cerevisiae Sir2(silent information regulator2),regulates various metabolic pathways(Haigis and Sinclair,2010).Loss of SIRT1causes defective gametogenesis, heart and retinal abnormalities,genomic instability,small body size,and reduced survival in mice(Cheng et al.,2003;McBurney et al.,2003;Wang et al.,2008)and abolishes many beneficial effects of dietary restriction(DR)(Chen et al.,2005).Although life span extension in yeast,worms,andflies by ectopic Sir2is under debate(Burnett et al.,2011;Lombard et al.,2011;Viswa-nathan and Guarente,2011),transgenic mice with additional copies of SIRT1show phenotypes resembling DR(Alcendor et al.,2007;Banks et al.,2008;Bordone et al.,2007;Herranz et al.,2010;Pfluger et al.,2008).SIRT1activator resveratrol in-span in rodents (Baur et al.,2006;Howitz et al.,2003;Milne et al.,2007;Wood et al.,2004).Though how resveratrol activates SIRT1is still unclear (Villalba et al.,2012),many of its in vivo benefits are dependent on SIRT1(Baur,2010).SIRT1deacety-lates a variety of proteins and regulates genomic integrity,inflammatory response,adipogenesis,mitochondrial biogen-esis,and stress resistance (Lavu et al.,2008).Of note,SIRT1deacetylates Foxo3a to enhance stress resistance through MnSOD,catalase,and Gadd45a ,etc.(Brunet et al.,2004).SIRT1is highly expressed in embryonic stem cells (ESCs),but is reduced in differentiated cells through a process mediated by miRNAs (Saunders et al.,2010).SIRT1is required for mainte-nance of self-renewal of ESCs via modulating p53cellular distri-bution and Nanog expression (Han et al.,2008).The hematopoi-etic differentiation of ESCs is defective,and the number and function of hematopoietic progenitor cells decline in SIRT1À/Àand SIRT1+/Àmice (Ou et al.,2011).When cultured under 5%oxygen,both SIRT1À/Àand SIRT1+/Àhematopoietic progenitor cells exhibit defective proliferation compared with wild-type cells (Mantel et al.,2008).Alternate splicing events also occur at the wild-type LMNA locus (Scaffidi and Misteli,2006).That the number of cells ex-pressing progerin increases in normal aged individuals (McClin-tock et al.,2007)and that telomere shortening or dysfunction activates progerin production (Cao et al.,2011)raise the possi-bility that progerin may contribute to normal aging (Burtner and Kennedy,2010),possibly through modulating the activity of longevity/antiaging proteins.Given the essential roles of NM inFigure 1.SIRT1Interacts with Lamin A(A)FLAG-SIRT1and lamin A were ectopically ex-pressed in HEK293cells.By western blotting,lamin A was detected in anti-FLAG immunopre-cipitates;FLAG-SIRT1was detected in anti-lamin A/C immunoprecipitates.(B)In total cell lysates of HEK293cells,SIRT1was pulled down by anti-lamin A/C immunoprecipi-tates and lamin A was pulled down by anti-SIRT1immunoprecipitates.(C)Representative immunofluorescence confocal microscopy of SIRT1and lamin A/C in human fibroblasts.The majority of nuclear SIRT1coloc-alizes with lamin A in the nuclear interior (arrows).Scale bar,5m m.(D)Representative confocal microscopy showing colocalization of EGFP-SIRT1and DsRed-lamin A in human fibroblast cells.Scale bar,10m m.(E)Lamin A but not lamin C was pulled down in anti-FLAG-SIRT1immunoprecipitates in HEK293cells.ties of SIRT1,we tested the potential effect of prelamin A on SIRT1.We found that lamin A directly interacts with and serves as an activator of SIRT1on the NM;prelamin A and progerin exhibit significantly reduced binding capacity to SIRT1in vivo,leading to a rapid decline of ASCs in Zmpste24À/Àmice.Resvera-trol increases the binding of SIRT1withA-type lamins and thus enhances its deacetylase activity,restores ASC population,ameliorates progeroid features,and extends life span in Zmpste24À/Àmice.RESULTSLamin A Interacts with SIRT1To test the potential involvement of SIRT1in progeria,we first examined the potential interaction between lamin A and SIRT1by coimmunoprecipitation in HEK293cells expressing ectopic FLAG-SIRT1.As shown in Figure 1A,lamin A was pulled down in the anti-FLAG immunoprecipitates,while FLAG-SIRT1was detected in the anti-lamin A/C immunoprecipitates.The interac-tion between endogenous SIRT1and lamin A was confirmed in HEK293cells,bone marrow stromal cells (BMSCs),and MEFs,where anti-SIRT1immunoprecipitates pulled down lamin A and vice versa (Figures 1B,S1A,and S1B).Immunofluorescence confocal microscopy showed that lamin A localizes both on the nuclear envelope and in the interior nucleoplasm.A signifi-cant portion of the nuclear SIRT1colocalized with nucleoplasmic lamin A in the nuclear interior in human fibroblasts (Figure 1C).Consistently,ectopic EGFP-SIRT1and DsRed-lamin A coex-isted in the nuclear interior (Figure 1D).This interaction seems specific to nuclear SIRT1,as neither cytoplasmic SIRT2nor mitochondrial SIRT5was detected in the anti-GFP-lamin A immunoprecipitates (Figure S1C).Alternative splicing of LMNA gives rise to different A-type lam-ins,of which lamin A and C are the most abundant (Lin and Wor-Cell MetabolismLamin A-Dependent Activation SIRT1by Resveratrola specific98aa carboxyl tail and lamin C has a unique6aa carboxyl tail(Liu and Zhou,2008).Although the level of lamin C was much higher than A in HEK293cells,lamin C was hardly de-tected in the anti-SIRT1immunoprecipitates(Figure1B),indi-cating that lamin A likely interacts with SIRT1via its C-terminal domain.This notion was further confirmed by coimmunoprecipi-tation in HEK293cells expressing FLAG-SIRT1together with either lamin A or lamin C.As shown in Figure1E,lamin A was detected in the anti-FLAG-SIRT1immunoprecipitates,whereas lamin C was negligible.Taken together,these data suggest that lamin A interacts with SIRT1through its carboxyl terminus. Lamin A Is an Activator of SIRT1The NM association of deacetylase activity suggests the exis-tence of potential SIRT1activators on the NM.In mammalian cells,lysine acetyltransferase p300and SIRT1mediate the acet-ylation and deacetylation of p53on residue K382(Gu and Roeder,1997).A BioMol SIRT1Fluorimetric Drug Discovery Kit (BSDK)utilizesfluorophore-conjugated acetyl p53peptide to determine SIRT1activity.To test the hypothesis,the in vitro de-acetylase activity of recombinant human SIRT1(rhSIRT1)deter-mined by BSDK assay was quantified in the presence or absence of NM derived from wild-type cells.The deacetylase activity of rhSIRT1was enhanced by approximately3-fold in the presence of NM compared with the control without NM(Figure S1D),sug-gesting potential SIRT1activator(s)associated with the NM. These data,together with the fact that lamin A interacts with SIRT1,prompted us to further examine whether lamin A acts as an activator of SIRT1.We found that SIRT1physically inter-acts with lamin A,as rhSIRT1was pulled down by recombinant human lamin A(rhLamin A)in the test tube(Figure2A).Addition-rhSIRT1toward BSDK acetyl p53peptide was increased in a lamin A dose-dependent manner(Figure2B).Lamin A-stimu-lated SIRT1deacetylase activity was completely abolished by SIRT1inhibitor Suramin Sodium.We also tested the effect of rhLamin A on the native target of SIRT1,i.e.,acetyl p53.Full-length acetyl FLAG-p53was purified by anti-FLAG immunopre-cipitation in HEK293cells ectopically expressing FLAG-p53 and HA-p300.SIRT1deacetylation assay was performed as described in Experimental Procedures.As shown in Figure2C, an approximate40%decrease in the acetylation level of FLAG-p53was observed in the presence of rhLamin A-rhSIRT1 complex,compared with rhSIRT1only.Moreover,the synthetic peptide(LA-80)containing the carboxyl80aa of lamin A protein increased SIRT1deacetylase activity in a dose-dependent manner in either BSDK assay(Figure2D)or the assay using puri-fied full-length acetyl p53protein(Figure2E).Notably,the activa-tion of the rhSIRT1activity by LA-80was saturated when the molar ratio between LA-80and rhSIRT1was close to1:1(Figures 2D and2E).Collectively,these data suggest that lamin A serves as a SIRT1activator.Resveratrol Stimulates SIRT1Deacetylase Activity ina Lamin A-Dependent MannerResveratrol,a potential SIRT1activator,has been reported to en-hance health span in a range of age-related diseases.However, independent studies have revealed that resveratrol activates SIRT1toward thefluorophore-conjugated synthetic p53peptide rather than its unconjugated native targets(Borra et al.,2005;Dai et al.,2010;Kaeberlein et al.,2005;Pacholec et al.,2010). Consistent with the published data,we also found that resvera-trol didn’t enhance SIRT1deacetylase activity toward full-lengthmin A Is an Activator of SIRT1(A)Recombinant human SIRT1(rhSIRT1)waspulled down by anti-lamin A immunoprecipitates intest tubes containing rhSIRT1and recombinanthuman lamin A(rhLamin A).(B)RhSIRT1deacetylase activity was determinedby BioMol SIRT1Fluorimetric Drug Discovery Kit(BSDK)in the presence or absence of rhLamin A.Data represent mean±SEM,n=3.*p<0.05,**p<0.01,rhLamin A+rhSIRT1versus rhSIRT1only.***The molar ratios of rhLamin A to rhSIRT1are0.5,1.0,2.0,and4.0,respectively.(C)Acetyl FLAG-p53was incubated with rhSIRT1in the presence or absence of rhLamin A.FLAG-p53acetylation was detected by western blottingwith anti-acetyl lysine antibodies.Relative level ofacetylated p53was quantified by ImageJ.*Themolar ratio of rhLamin A to rhSIRT1is1.(D)RhSIRT1deacetylase activity was determinedby BioMol SIRT1Fluorimetric Drug Discovery Kit(BSDK)in the presence or absence of LA-80(synthetic peptide of carboxyl80aa of lamin A).Data represent mean±SEM,n=3.**p<0.01,LA-80+rhSIRT1versus rhSIRT1only.(E)Acetyl FLAG-p53was incubated with rhSIRT1in the presence of various amount of LA-80.FLAG-p53acetylation was detected by western blottingwith anti-acetyl lysine antibodies(left).Relativelevel of acetylated p53was quantified by ImageJ(right).Cell Metabolism Lamin A-Dependent Activation SIRT1by Resveratrolpresence of rhLamin A,resveratrol significantly enhanced SIRT1deacetylase activity toward its native target,acetyl p53(Fig-ure 3A,right panel).rhLamin A alone enhances rhSIRT1deace-tylase activity independent of resveratrol (Figures 3B,left panel),and resveratrol further enhances the activation of rhSIRT1de-acetylase activity mediated by rhLamin A (Figures 3B,right panel).Further investigation revealed that resveratrol enhanced the association between rhSIRT1and rhLamin A both in the test tube (Figure 3C)and in HEK293cells examined by coimmu-noprecipitation (Figures 3D and 3E).SIRT1Is Mislocalized in Progeroid CellsIt is widely accepted that the unprocessed C-terminal tail in pro-gerin or prelamin A is responsible for the progeroid features in HGPS and progeria mouse models.Given that lamin A interacts with SIRT1via its C-terminal domain,we examined whether the interaction between SIRT1and prelamin A or progerin is reduced compared to lamin A.We performed coimmunoprecipitation in HEK293cells expressing FLAG-SIRT1together with one of the A-type lamins,i.e.,wild-type lamin A,or prelamin A or progerin.As shown in Figures 4A and 4B,significantly less prelamin A and progerin were pulled down by anti-FLAG antibodies,compared with lamin A,though comparable or higher levels of prelamin A and progerin were present in the input.These data suggest that SIRT1preferentially interacts with lamin A,whereas prela-min A or progerin has significantly reduced association withFigure 3.Resveratrol Activates SIRT1in a Lamin A-Dependent Manner(A)A representative western blot showing FLAG-p53acetylation using anti-acetyl lysine antibodies.Acetyl FLAG-p53was incubated with rhSIRT1in the presence or absence of rhLamin A and resveratrol.(B)Acetyl FLAG-p53was incubated with rhSIRT1and rhLamin A in the presence or absence of re-sveratrol.FLAG-p53acetylation was detected by western blotting with anti-acetyl lysine antibodies.Relative level of acetylated p53was quantified by ImageJ.*The molar ratios of rhLamin A to rhSIRT1are 0.5,1.0,and 2.0,respectively.(C)A representative western blot showing rhSIRT1pulled down by anti-lamin A/C antibody in the presence or absence of resveratrol.(D)A representative western blot showing that treatment with increasing concentrations of re-sveratrol resulted in increased interaction between SIRT1and lamin A.FLAG-SIRT1and lamin A were cotransfected into HEK293cells.Cells were treated with resveratrol followed by anti-FLAG immunoprecipitation before being subjected to western blotting.(E)Quantification of (D).Data represent mean ±SEM,n =3.*p <0.05.Since lamin A is one of the major com-ponents of NM,we further investigated the association of SIRT1with the NM by subcellular fractionation.SIRT1À/Àcells were utilized as a negative control forthe specific staining of SIRT1protein.NM-associated KAP-1(KRAB-associated protein 1)(Goodarzi et al.,2008)and chro-matin-bound MCM3(Me´ndez and Stillman,2000)served as controls for the purity of the subcellular fractionation.As ex-pected,KAP-1was resistant to MNase digestion and remained in the NM fraction (P20),whereas the majority of MCM3was released into the nucleoplasmic and chromatic fraction (S20)after MNase treatment (Figure S2A).Consistent with its interac-tion with lamin A,SIRT1protein was enriched in the NM fraction (P20)in MEFs (Figure S2B).Since prelamin A has less binding capacity to SIRT1compared with lamin A and SIRT1is highly expressed in stem cells (Saunders et al.,2010),we examined SIRT1localization in Zmpste24À/Àcells by subcellular fraction-ation in multipotent BMSCs.NM-associated SIRT1was largely reduced in Zmpste24À/ÀBMSCs compared to wild-type con-trols (Figure 4C,right panel),though total nuclear proportion of SIRT1was comparable (Figures 4C,left panel).The reduction in NM-associated SIRT1appeared specific,because SIRT6,CBP acetyltransferase,and Foxo3a were not significantly affected in Zmpste24À/Àcells.The NM-associated SIRT1was also reduced in HGPS dermal fibroblasts,including HG143,HG188,HG164,and HG122,compared to either healthy F2-S fibroblasts or dermal fibroblasts harboring nonprogeria LMNA mutations,i.e.,R453W in Emery Dreifuss Muscular Dystrophy (EDMD),R482W in Familial Lipodystrophy (FLPD),and R401C in EDMD (Liu and Zhou,2008)(Figures S2C and S2D).Though Cell MetabolismLamin A-Dependent Activation SIRT1by ResveratrolFigure 4.Mislocalization of SIRT1in Progeroid Cells(A)HEK293cells were transiently transfected with FLAG-SIRT1together with one of the A-type lamins,i.e.,wild-type lamin A,unprocessible prelamin A,and progerin.Western blotting was performed to determine levels of A-type lamins in anti-FLAG immunoprecipitates.Representative immunoblot shows that significantly less prelamin A/progerin was pulled down by anti-FLAG antibody compared with wild-type lamin A,even though comparable or higher level of prelamin A/progerin was found in the inputs.(B)Quantification of (A).Data represent mean ±SEM,n =3.**p <0.01.(C)A representative immunoblot showing various proteins in nuclear (Nu,P1)and nuclear matrix (NM,P20)fractions in BMSCs.NM-associated SIRT1was significantly reduced in Zmpste24À/ÀBMSCs,whereas levels of Sirt6,CBP,Foxo3a,histone H3,and b -actin were comparable between wild-type and Zmpste24À/ÀBMSCs in NM fraction.Total nuclear level of SIRT1was not changed.Cell MetabolismLamin A-Dependent Activation SIRT1by Resveratrollines,the percentage of NM-associated SIRT1was consistently reduced.Moreover,ectopic expression of prelamin A and pro-gerin caused remarkable disassociation of SIRT1from the NM,while the nuclear level of SIRT1was hardly affected in HEK293cells (Figures 4E and 4F).Taken together,these observations suggest that prelamin A or progerin compromises the proper NM localization of SIRT1.SIRT1Deacetylase Activity Is Compromised in Progeroid CellsTo further assess the functional significance of mislocalization of SIRT1,we applied BSDK assay to compare the rhSIRT1activa-tion by NM prepared from wild-type or Zmpste24À/ÀBMSCs.Consistently,the deacetylase activity of rhSIRT1was enhanced approximately 3-fold in the presence of NM from wild-type BMSCs (Figure 5A).In contrast,the NM from the Zmpste24À/ÀBMSCs showed a significantly reduced stimulatory effect on rhSIRT1deacetylase activity.Cytoplasm did not significantly enhance rhSIRT1activity.We next examined downstream path-way(s)of SIRT1in progeria cells.SIRT1deacetylates Foxo3a and upregulates its transcriptional activity,thus promoting expression of antioxidant enzymes such as MnSOD and cata-lase in response to oxidative stress (Brunet et al.,2004).Consis-tent with the mislocalization of SIRT1and significantly less acti-vation of rhSIRT1in the test tube,Foxo3a was hyperacetylated in Zmpste24À/ÀBMSCs,and the levels of catalase and Gadd45a were reduced by approximately 40%in Zmpste24À/Àmice rela-tive to wild-type controls (Figures 5B and 5C).The increase in Foxo3a acetylation is likely the result of decreased SIRT1deace-tylase activity in vivo,as neither total nuclear SIRT1level nor NM association of CBP,the acetyltransferase for Foxo3a,was changed in Zmpste24À/ÀBMSCs (Figures 4C and 4D).Support-ing this idea,ectopic expression of either prelamin A or progerin increased the acetylation of FOXO3A and reduced the ex-pression of catalase,MnSOD,and GADD45a in HEK293cells (Figures 5D and 5E).Consistent with elevated levels of acetyl Foxo3a,levels of acetyl FLAG-p53are also significantly higher in HEK293cells ectopically expressing prelamin A and progerin compared with those expressing wild-type lamin A (Figure 5F).As SIRT1interacts with lamin A and thus associates with NM,and resveratrol increases the binding between lamin A and SIRT1,we asked whether resveratrol enhances NM association of SIRT1.As shown in Figures S3A and S3B,in wild-type and Zmpste24À/ÀBMSCs incubated with different concentrations of resveratrol,NM-associated SIRT1was increased compared to untreated controls.The ability of resveratrol to stimulate the NM association of SIRT1was also observed in test tube.When equal amounts of rhSIRT1were incubated with the insoluble NM fraction from either wild-type or Zmpste24À/Àcells sus-pended in BSDK assay buffer,significantly less NM-bound rhSIRT1was found in precipitates from Zmpste24À/ÀNM relative to wild-type NM precipitates in the absence of resveratrol (Fig-ure 5G).The presence of resveratrol enhanced the associa-tion of rhSIRT1with NM derived from both wild-type andZmpste24À/Àcells.Consistently,rhSIRT1level in the superna-tant underwent a compensatory reduction (Figure 5G).Collec-tively,these data suggest that resveratrol could enhance the interaction between lamin A and SIRT1to increase the NM asso-ciation of SIRT1.Resveratrol Treatment Rescues ASC Decline in Zmpste24–/–MiceAs resveratrol enhances SIRT1deacetylase activity by increasing its binding to lamin A,we decided to examine the effects of re-sveratrol treatment on the early senescence in Zmpste24À/ÀMEFs.However,no obvious difference in b -galactosidase activity was observed between resveratrol-treated and saline-treated Zmpste24À/ÀMEFs (Figure S4A),and resveratrol did not reduce the elevated levels of p16Ink4a in Zmpste24À/ÀMEFs (Figure S4B).Progerin and prelamin A have been previously linked to defects in mesenchymal stem cells (MSCs)and in hair follicle progenitor cells in Zmpste24À/Àmice (Espada et al.,2008;Scaffidi and Misteli,2008).Consistently,the number of BMSCs was significantly reduced in Zmpste24À/Àmice compared with wild-type controls at 4months of age (Figures S5A).Zmpste24À/ÀBMSCs in culture showed compromised colony-forming capacity (Figure S5B),reduced proliferation (Figure S5C),and a dramatically increased cellular senescence (Figures S5D).Similarly,an early decline in mononucleated cells (MNCs)and hematopoietic stem cells (HSCs,Lineage ÀFlt3ÀSca-1+cKit high )was observed in Zmpste24À/Àmice (Figures S5E–S5G),such that by 4months of age HSC levels fell to less than half of that of wild-type controls.HSC transplantation experiments showed that the self-renewal defects were cell-intrinsic (Figure S5H).As SIRT1is highly expressed in stem cells and is critical for maintaining stem cell self-renewal and function,we then tested the effects of resveratrol on ASCs.Interestingly,resveratrol enhanced the colony-forming capacity in Zmpste24À/ÀBMSCs in a dose-dependent manner (Figures 6A and 6B).The treatment increased the binding of SIRT1to prelamin A,downregulated acetylation level of Foxox3a,and upregulated the expression of Gadd45a and catalase (Figures 6C and 6D).Moreover,the rescue effect of resveratrol is SIRT1dependent,as knocking down SIRT1attenuated its effect on Zmpste24À/ÀBMSCs (Figures 6D–6F).Knocking down SIRT1abolished the stimulating effect of resveratrol on the expression of Gadd45a and catalase (Figure 6D).Knockdown of SIRT1decreased colony-forming capacity (Figures 6E and 6F),whereas ectopic SIRT1increased the colony-forming capacity of Zmpste24À/ÀBMSCs to levels comparable to that of wild-type BMSCs (Figures 6G and 6H).These data suggest that BMSC decline in Zmpste24À/Àmice is attributable largely to impaired SIRT1function,which can be rescued by resveratrol.Resveratrol Alleviates Progeroid Features and Extends Life Span in Zmpste24–/–MiceThe SIRT1-dependent rescue of BMSC colony-forming capacity in vitro prompted us to ask whether resveratrol could rescue the(E)Lamin A,unprocessible prelamin A,or progerin was stably expressed in HEK293cells.Subcellular fractionation and western blotting were performed to determine the NM-associated SIRT1.A representative western blotting shows that while NM association of SIRT1was reduced in prelamin A-and progerin-transfected cells compared with wild-type lamin A,the levels of Foxo3a and b -catenin remained unchanged.Cell MetabolismLamin A-Dependent Activation SIRT1by ResveratrolFigure promised Foxo3a and p53Pathways in Progeroid Cells(A)RhSIRT1deacetylase activity was determined in the presence of cytoplasmic or NM fraction.The relative increase in deacetylase activity after addition of rhSIRT1to the assay buffer,cytoplasm,or NM was determined and plotted.The NM from wild-type BMSCs potentiated rhSIRT1deacetylase activity,whereas the stimulating capacity of NM from Zmpste24À/ÀBMSCs was greatly compromised.Data represent mean ±SEM,n =3.*p <0.05.(B)A representative western blot showing hyperacetylation of Foxo3a in Zmpste24À/ÀBMSCs using anti-acetyl lysine antibodies in anti-Foxo3a immunopre-cipitates.(C)Quantification of (B).Data represent mean ±SEM,n =3.*p <0.05,**p <0.01.(D)Upper:A representative western blot showing acetylation of Foxo3a using anti-acetyl lysine antibodies in anti-Foxo3a immunoprecipitates in HEK293cells ectopically expressing lamin A or prelamin A or progerin.Lower:Expression of catalase,MnSOD,and GADD45a in the input.(E)Quantification of (D).Data represent mean ±SEM,n =3.*p <0.05,**p <0.01.(F)A representative western blot showing acetylation level of ectopic FLAG-p53in HEK293cells transfected with different lamin A in the presence or absence of resveratrol.Relative level of acetylated FLAG-p53was quantified by ImageJ.(G)Resveratrol enhanced the association of rhSIRT1with NM in test tube.Recombinant hSIRT1was incubated with NM fraction prepared from wild-type or Zmpste24À/ÀBMSCs in the presence and absence of resveratrol (10m M)in a similar way as the SIRT1deacetylase activity assay was performed.Insoluble NM Cell MetabolismLamin A-Dependent Activation SIRT1by Resveratrol。

2012 CANCER CELL DNA

2012  CANCER  CELL    DNA

Cancer CellArticleDNA Methylation Screening IdentifiesDriver Epigenetic Events of Cancer Cell SurvivalDaniel D.De Carvalho,1,3Shikhar Sharma,1,2,3Jueng Soo You,1Sheng-Fang Su,1,2Phillippa C.Taberlay,1Theresa K.Kelly,1Xiaojing Yang,1Gangning Liang,1and Peter A.Jones1,*1Department of Urology,Biochemistry and Molecular Biology2Program in Genetic,Molecular and Cellular BiologyUniversity of Southern California/Norris Comprehensive Cancer Center Keck School of Medicine,University of Southern California,Los Angeles,CA90089-9181,USA3These authors contributed equally to this work*Correspondence:pjones@DOI10.1016/r.2012.03.045SUMMARYCancer cells typically exhibit aberrant DNA methylation patterns that can drive malignant transformation. Whether cancer cells are dependent on these abnormal epigenetic modifications remains elusive.We used experimental and bioinformatic approaches to unveil genomic regions that require DNA methylation for survival of cancer cells.First,we surveyed the residual DNA methylation profiles in cancer cells with highly impaired DNA methyltransferases.Then,we clustered these profiles according to their DNA methylation status in primary normal and tumor tissues.Finally,we used gene expression meta-analysis to identify regions that are dependent on DNA methylation-mediated gene silencing.We further showed experimentally that these genes must be silenced by DNA methylation for cancer cell survival,suggesting these are key epigenetic events associated with tumorigenesis.INTRODUCTIONDuring tumorigenesis cancer cells acquire,through a multistep process,a new set of properties that allows them to overcome physiological homeostasis.These properties include unlimited proliferation potential,self-sufficiency in growth signals,resis-tance to antiproliferative and apoptotic signals and immune system evasion,among others(Hanahan and Weinberg,2000, 2011).These alterations,on the other hand,contribute to a process known as the stress phenotype of cancer(Luo et al., 2009),which includes DNA damage/replication stress,proteo-toxic stress,mitotic stress,metabolic stress,and oxidative stress.To survive the tumorigenic process,a cancer cell undergoes several modifications to its genomic circuitry,such as activating mutations in oncogenes and aberrant activation of nononco-genic pathways.These adaptations lead to oncogene addiction (Weinstein,2002)and nononcogene addiction(Solimini et al.,2007),respectively.Because of this aberrant circuitry,cancer cells become hypersensitive to the effects of classic tumor suppressor genes(TSGs)(Luo et al.,2009;Weinstein,2002) and,potentially,to genes that can inhibit the nononcogenic signaling pathways that cancer cells rely on to survive. Changes in the cancer cell transcriptome can be driven by genetic and epigenetic alterations(Baylin and Ohm,2006;Jones and Baylin,2007).DNA methylation is an epigenetic process that can heritably change gene expression without altering the DNA sequence.In normal somatic cells,most DNA methylation occurs at CpG dinucleotides within CpG poor sequences, whereas CpG-rich sequences,also known as CpG islands,are usually unmethylated(Sharma et al.,2010).DNA methylation is a vital mechanism of epigenetic gene silencing,playing key roles in X chromosome inactivation,genomic imprinting,embry-onic development,silencing of repetitive elements and germ cell-specific genes,differentiation,and maintenance of pluripo-tency(De Carvalho et al.,2010;Meissner,2010;Robertson, Cancer Cell21,655–667,May15,2012ª2012Elsevier Inc.655HCT116DKO1LDHAL6B 200bpADAM2200bpARMCX1 MEOX2HCT116DKO10.66 (1.0) 0.44 (0.9) 0.97 (1.0)0.74 (0.87)1.0 (1.0) 0.63 (1.0) 0.96 (1.0)0.55 (0.83)DD N A M e t h y l a t i o n P r o b e sA B CFigure 1.Clustering of DNMT-Deficient Cells Identifies Three Classes of Putative Driver Genes Marked by DNA Methylation(A)One-dimensional hierarchical clustering using Euclidean distance and average linkage was performed with the $24,000Infinium DNA methylation probes located outside of repeats or known SNPs in HCT116wild-type,DKO8,and DKO1cell lines.Each row represents a probe;each column represents a sample.The b value (level of DNA methylation)for each probe is represented with a color scale as shown in the key.(B)k Means (K =4)clustering of the 566Infinium DNA methylation probes that maintain DNA methylation in DKO1sample (a b value of at least 0.6and a difference between HCT116and DKO1smaller than 0.2)in (A)for 10TCGA samples (n =4normal colon and n =6primary colon adenocarcinoma).(C)Heat map of 566Infinium DNA methylation probes in 32normal tissues retaining the probe order from (B).Primary normal bladder (n =4),sperm (n =1),and primary normal TCGA kidney (n =15),lung (n =4),and ovary (n =8).WGA DNA was used as a negative control for DNA methylation.(D)Bisulfite-sequencing validation of Infinium DNA methylation data from two regions (LDHAL6B and ADAM2)from the somatic-specific DNA methylation cluster and two regions (ARMCX1and MEOX2)from the cancer-specific DNA methylation cluster.Arrow indicates the position of the Infinium probe.Empty and filledCancer CellDNA Methylation Drivers656Cancer Cell 21,655–667,May 15,2012ª2012Elsevier Inc.2005).Besides these physiological roles,deregulated DNA methylation can also be a major driver of pathological conditions,including neurological and autoimmune diseases,as well as cancer (Kelly et al.,2010;Portela and Esteller,2010;Taberlay and Jones,2011).During tumorigenesis,global DNA methylation patterns change,resulting in hypomethylation of non-CpG islands and hypermethylation of CpG islands (Sharma et al.,2010).DNA hypermethylation has been shown to result in abnormal silencing of several TSGs in most types of cancer (Jones and Baylin,2002,2007).Recently,several efforts to examine the cancer methylome,utilizing genome-wide techniques,have revealed that a large number of genes exhibit aberrant DNA methylation profiles in cancer (Figueroa et al.,2010;Irizarry et al.,2009).These changes can be used to stratify subtypes of cancers (Figueroa et al.,2010;Noushmehr et al.,2010)and to predict cancer outcomes (Portela and Esteller,2010),other uses.Distinguishing which genes play key ‘‘driver’’roles via DNA methylation-mediated gene silencing in cancer initiation,progression,and maintenance and those genes that are only ‘‘passengers’’in the tumorigenic process would be extremely useful in more targeted epigenetic therapies (Kelly et al.,2010).However,making this distinction has proven extremely difficult due to the large number of differentially DNA-methylated genes in human cancers (Kalari and Pfeifer,2010).We,and others,have suggested that cancer cells may become addicted to an aberrant epigenetic landscape,espe-cially with respect to DNA methylation (Baylin and Ohm,2006;Kelly et al.,2010).However,as of yet,and to our knowledge,there is no direct evidence for such an addiction.Furthermore,mining the thousands of genomic regions that are de novo DNA methylated in cancer and identifying those required for cancer cell survival have proven extremely challenging (Kalari and Pfeifer,2010).Here,we describe an approach to identify driver epigenetic events associated with cancer cell survival.Our findings pave the way for new generations of epigenetic therapies,which target the genes cancer cells rely on being silenced by DNA methylation in order to survive.RESULTSIdentification of the Minimum DNA Methylation Profile Required for Cancer Cell SurvivalWe hypothesized that cancer cells depend on DNA methylation of a few key regions for survival and that these regions would preferentially maintain methylation when artificially reducing global DNA methylation.To test this hypothesis,we profiled HCT116colon cancer cells and HCT116cells with a genetic disruption of DNMT3B and DNMT1(DKO)(Rhee et al.,2002).This genetic disruption led to a complete knockout of DNMT3B and a truncated DNMT1transcript,expressed at very low levels (Egger et al.,2006;Rhee et al.,2002;Spada et al.,2007).For this study we used two DKO subclones,DKO8and DKO1,which retain approximately 45%and 5%of the HCT116wild-type global DNA methylation levels,respectively (Rhee et al.,2002;Sharma et al.,2011).It is important to note that a further reduc-tion of DNMT1levels,by RNAi,in cells with a genetic disruption of DNMT1results in demethylation and a massive reduction of cell viability and immediate induction of cell death (Spada et al.,2007),suggesting that DNA methylation is required for cancer cell survival.We profiled promoter DNA methylation of HCT116,DKO8,and DKO1cell lines using the Illumina Infinium platform (HumanMethylation27)and observed a reduction in global DNA methylation levels in DKO8cells compared to HCT116wild-type cells and an even greater reduction in DKO1cells (Fig-ure 1A),consistent with previous data (Rhee et al.,2002;Sharma et al.,2011).Surprisingly,we found a collection of 566CpG sites,spanning 490genes that despite the strong impairment in DNA methyltransferase activity,still retained a high level of DNA methylation in DKO1cells,with a b value higher than 0.6(see Table S1available online for gene/probe list).These regions were also highly methylated in HCT116and DKO8cells,and none showed a difference in their b values greater than 0.2among the three cell lines.Next,we sought to identify whether there was a cancer-specific DNA methylation profile at these regions that maintained DNA methylation even in DKO1cells,which would potentially include important putative targets for epigenetic therapy.To accomplish this,we first compared the DNA methylation levels of the 566CpG sites that retained DNA methylation in DKO1cells to the DNA methylation profile of 6primary colon adenocar-cinoma tissue samples and 4normal colon tissue samples obtained from The Cancer Genome Atlas (TCGA)ing k means clustering,we identified 92CpG sites,spanning 77genes that were unmethylated in normal colon and became hypermethylated in colon adenocarcinoma (Figure 1B;Table S1),consistent with a cancer-specific methylation profile.We further compared these data to DNA methylation data of several normal tissues including sperm,bladder,kidney,lung,and ovary,which allowed us to identify clusters of gene regions highly enriched for somatic tissue-specific DNA methylation.Such genes were methylated in the somatic tissues analyzed and unmethylated in germ cells.This somatic tissue-specific cluster comprised 99CpG sites,spanning 83genes (Figure 1C;Table S1).Furthermore,we also identified genes that exhibit cell culture-specific DNA methylation,such that these regions are methylated in all cell lines analyzed but unmethylated in primary tissues (Figure 1C;Table S1).This cell culture-specific cluster comprised 29CpG sites,spanning 25genes.We focused only on these three groups because of their differential DNA methylation profiles.We speculate that the remaining 346CpG sites might be regions that are more prone to methylation,remaining a good target for residual DNA methylation activity without functional relevance or,alternatively,may have a tissue-specific expression profile,being unmethylated only incircles denote unmethylated and methylated CpG sites,respectively.Each horizontal row represents one sequenced DNA clone.The number on the right represents the mean DNA methylation score of each region,and the number in the parentheses represents the mean DNA methylation score of the specific Infinium CpG site.See also Figure S1and Table S1.Cancer CellDNA Methylation DriversCancer Cell 21,655–667,May 15,2012ª2012Elsevier Inc.657specific cell types that were not surveyed in this study.Whole Genome Amplified(WGA)DNA served as a negative control(Fig-ure1C)to confirm that the regions identified as being methylated in DKO1cells were not false positives due to technical problems with the specific Infinium probes.The distribution of probes,relative to transcription start sites (TSSs),in the cancer-specific and somatic tissue-specific clusters was found to be very similar to the distribution of the array itself,whereas the distribution in the cell culture-specific cluster tended to be slightly more concentrated at the TSS (Figure S1A).However,it should be noted that there was no association between distance to TSS and methylation cluster (somatic,cancer,cell line)as assessed by one-way ANOVA (p>0.05).We selected genomic regions from each cluster to validate the Infinium-based DNA methylation data using bisulfite sequencing.All of the sequences analyzed showed high levels of DNA methylation in HCT116wild-type and DKO1cells(Fig-ure1D),with the CpG site surveyed by the Infinium platform presenting a maximum difference between their b values of 0.17in DKO1when compared to HCT116wild-type.These results demonstrate that even though DKO cells are globally DNA hypomethylated(Rhee et al.,2002;Sharma et al.,2011), the residual DNA methylation is focal and site specific,support-ing the hypothesis that there is a functional role for some of the retained DNA methylation.To further demonstrate the importance of the DNA methylation that is retained in the three identified clusters,we treated DKO1 cells with1m M of5-Aza-20-deoxycytidine or PBS for24hr.After treatment,we allowed at least two population doublings(5days) for demethylation to occur and then analyzed changes in the DNA methylation profile by the Illumina Infinium array.As ex-pected,most of the regions in the three previously identified clusters were resistant to demethylation,with only eight regions from the somatic cluster,five regions from the cancer cluster, and one region from the cell culture cluster presenting a differ-ence in the b value greater than0.2(Figure S1B).These regions we considered false positives and excluded them from subse-quent analysis.Residual Methylation in DKO1Cannot Be Explainedby an Inherent Susceptibility to DNA MethylationOur working hypothesis is that the artificial impairment of DNA methyltransferase machinery in DKO1cells will induce a strong selective pressure for any remaining DNA methylation to be maintained at the regions necessary for cancer cell survival.An equally plausible hypothesis is that the residual methylation reflects an inherent tendency for some genes to remain methyl-ated.Indeed,previous studies suggest that certain genomic regions are more prone to DNA methylation(Este´cio et al., 2010;Ohm et al.,2007;Schlesinger et al.,2007;Widschwendter et al.,2007).Therefore,these regions may remain better targets for residual DNA methylation activity.To directly test this alternative hypothesis,we used two known approaches to predict whether a gene is more prone to DNA methyltransferase activity in cancer cells:one based on its chromatin structure (Ohm et al.,2007;Schlesinger et al.,2007;Widschwendter et al.,2007),and another based on its genomic architecture (Este´cio et al.,2010).Genes marked by H3K27me3in embryonic stem cells(ESCs) are known to be predisposed to DNA methylation in cancer cells (Ohm et al.,2007;Schlesinger et al.,2007;Widschwendter et al., 2007).Indeed,we found that the H3K27me3status in ESCs can accurately predict the methylation levels in the wild-type HCT116cells(Figure2A).Then,we tested whether methyla-tion-prone regions(H3K27me3positive in ESCs)would preferen-tially retain methylation in DKO1cells,compared to HCT116 cells.If this hypothesis was correct,we should observe an enrichment of methylation-prone genes in the cohort of genes that is methylated in DKO1cells because they would be better targets for residual DNA methyltransferase activity.Yet,there was no such enrichment(Figure2B).Rather,we observed a slight decrease in the proportion of methylation-prone genes that retain methylation in DKO1cells.These data suggest that genes found to retain DNA methylation in DKO1cells are not simply predisposed to DNA methylation in cancer cells.To further test this hypothesis,we performed a similar analysis using a previously published algorithm to predict whether a genomic region is prone,intermediate,or resistant to DNA methylation in cancer cells,based on its genomic architecture (Este´cio et al.,2010).In our test this algorithm accurately pre-dicted methylation levels in HCT116cells(Figure2C).Similar to the previous analysis,if the genes that maintain methylation in DKO1cells were simply more prone to DNA methyltransferase activity,one would expect an enrichment of methylation-prone genes in the pool of genes that retains methylation in DKO1cells. Again,we could notfind such enrichment(Figure2D).Taken together,these data suggest that the targets of residual DNA methylation in DKO1cells are not dictated by an inherent predisposition to DNA methylation based on either the chromatin structure or the genomic architecture.Thesefindings further support our original hypothesis that these loci retain methylation due to a functional selection pressure.Validation of the Findings in Other Types of Cancerand Association with Gene ExpressionWe next validated ourfindings in a larger test set of colon adeno-carcinoma and normal colon ing DNA methylation data available from TCGA,we observed a significant increase in DNA methylation in the majority of the CpG sites identified with cancer-specific DNA methylation in168primary colon adenocarcinoma samples relative to16normal colon samples (Figure3A;Wilcoxon rank sum test followed by a FDR correction, p<0.05).We next extended ourfindings to determine whether this cancer-specific DNA methylation profile was unique to colon adenocarcinoma or if it could also be observed in other tumor ing DNA methylation data available from TCGA,we analyzed DNA methylation from19lung adenocarcinoma samples against4normal samples.Again,the same pattern emerged,where the identified CpG sites presented an overall significant gain of DNA methylation in the tumor samples(Fig-ure3C;Wilcoxon rank sum test followed by a FDR correction, p<0.05).Indeed,most of the genes statistically determined to be hypermethylated in lung adenocarcinoma were also hyper-methylated in colon adenocarcinoma(Figure3E).These data indicate that the CpG sites we identified as cancer specific are frequently hypermethylated in other types of human cancer relative to the normal cell counterparts,suggesting that theseCancer CellDNA Methylation Drivers658Cancer Cell21,655–667,May15,2012ª2012Elsevier Inc.regions might have a more fundamental role in tumorigenesis,such as cell survival.Because DNA methylation of CpG islands located in promoter regions is well known to be correlated with gene silencing (Cedar and Bergman,2009;Jones and Baylin,2007;Portela and Estel-ler,2010),we investigated the expression state of the genes identified using independent data sets.We selected two micro-array data sets from the Gene Expression Omnibus database (/geo/):colon adenocarcinoma against normal colon (GSE 8671)(Sabates-Bellver et al.,2007);and lung adenocarcinoma against normal lung (GSE7670)(Su et al.,2007).We found an inverse correlation between DNA methylation and gene expression when we analyzed the gene expression data of 32normal colon samples and 25colon tumor samples.The majority of the genes subject to cancer-specific DNA methylation displayed decreased gene expression in colon cancer samples compared to normal colon (Figure 3B;t test followed by a FDR correction,p <0.05).We also observed that some genes showed a similarly low level of expression in both samples,probably due to an epigenetic switch in the silencing mechanism where the gene was already silenced in the normal sample by another epigenetic mechanism and became de novo DNA methylated in cancer (Gal-Yam et al.,2008).Moreover,we found a similar gene expression pattern in lung adenocarcinoma,where most of the cancer-specific DNAmethylation genes displayed decreased gene expression in the tumor when compared to the correspondent normal tissue (Figure 3D;t test followed by a FDR correction,p <0.05).Again,most of the genes statistically repressed in lung adenocarci-noma were also repressed in colon adenocarcinoma (Figure 3F).These data further suggest that there is a functional relevance of identified DNA methylation.Altogether,by combining gene expression with DNA methyla-tion data,we identified regions that are candidates for DNA methylation-mediated gene silencing.Moreover,the gene expression data corroborate our cluster analysis by using a different method and independent data sets to demonstrate biological differences in the gene clusters we identified.Spontaneous Loss of DNA Methylation at the Identified Genomic Regions Is Associated with Cell DeathDKO1cells have highly impaired DNA methyltransferase machinery due to the absence of DNMT3B,very low protein levels of DNMT3A,and low levels of a truncated DNMT1(Egger et al.,2006;Sharma et al.,2011).As a consequence of the impaired DNA methyltransferase machinery,the global DNA methylation level in this cell line is very low,with most of the genes that were methylated in the parental HCT116cells losing this methylation in DKO1cells.Therefore,we hypothesize that DKO1cells would be under a constant selective pressure to maintain the residual DNA methylation at key regions necessary for this cancer cell to survive.HCT116H3K27me3positiveH3K27me3negative PcG Marking in ESC:M E T H Y L A T I O N(m e a n b e t a -v a l u e +/-S E M )ABH3K27me3positiveH3K27me3negativePcG Marking in ESC:n=7972n=498n=2766F r e q u e n c y (%)CDMethylation-proneMethylation-Intermediate Methylation-resistant Prediction Group:n=6436n=407n=2221F r e q u e n c y (%)HCT116Methylation-intermediateMethylation-proneMethylation-resistant Prediction Group:M E T H Y L A T I O N(m e a n b e t a -v a l u e +/-S E M )Figure 2.Residual Methylation in DKO1Is Not Caused by an Inherent Susceptibility to DNA Methylation(A)Validation of H3K27me3status in ESCs as a predictive method for DNA methylation in HCT116cells.Methylation status of $27,000CpG sites was determined by Infinium.A t test with Mann-Whitney U posttest was performed.Data represent the mean ±SEM.(B)Frequency of probes marked by H3K27me3in ESCs in the cohort of DNA-methylated probes (b value >0.6)in HCT116,DKO8,and DKO1cells.(C)Validation of the predictive method based on genomic architecture (Este´cio et al.,2010)in HCT116cells.Methylation status of $27,000CpG sites was determined by Infinium.One-way ANOVA with Kruskal-Wallis test as performed.Data represent the mean ±SEM.(D)Frequency of methylation-prone genes in the cohort of DNA-methylated genes (b value >0.6)in HCT116,DKO8,and DKO1cells.Cancer CellDNA Methylation DriversCancer Cell 21,655–667,May 15,2012ª2012Elsevier Inc.659Figure 3.Validation of CpG Sites Identified with Cancer-Specific DNA Methylation Using Independent Data Sets and Association with Gene Repression(A)Volcano plot of the CpG loci identified as cancer-specifically methylated in colon adenocarcinoma (normal n =16,cancer n =168)from TCGA Data Portal.The b value difference in DNA methylation between the tumor samples and the correspondent normal samples is plotted on the x axis,and the p value for a FDR-corrected Wilcoxon rank sum test of differences between the tumor and correspondent normal samples (À13log10scale)is plotted on the y axis.Probes that are significantly hypermethylated (FDR adjusted p <0.05)in tumors are shown in red.(B)Volcano plot gene expression data of cancer-specific DNA-methylated genes.Gene expression data were obtained from GEO (GSE 8671)from primary normal colon (n =32)and primary colon cancer (n =25).For the volcano plots,gene expression fold change between the normal tissues and the tumor tissues isCancer CellDNA Methylation Drivers660Cancer Cell 21,655–667,May 15,2012ª2012Elsevier Inc.We next investigated whether DKO1cells exhibit a higher basal level of cell death than HCT116wild-type cells.When quantifying cell death by measuring the externalization of phosphatidylserine (PS)using annexin V by flow cytometry,we observed at least four times more spontaneous cell death in DKO1than in the parental HCT116cells (Figure 4A).This suggests that DKO1cells are indeed under constant selective pressure,probably because during cell division,some daughter cells lose DNA methylation at key regions due to the impaired DNA methyltransferase activity in DKO1cells (Egger et al.,2006;Spada et al.,2007),and consequently,they cannot survive.We took advantage of the increased rates of spontaneous cell death in DKO1cells to further test our hypothesis that cancer cells depend on constant DNA methylation of these regions in order to ing cell sorting,we first separated DKO1cells into two populations:annexin V positive (early spon-taneous apoptosis),and annexin V negative (viable cells)(Fig-ure 4A).These two populations have distinct morphologies,with annexin V-positive cells in the range of lower Forward Scatter (FSC)and higher Side Scatter (SSC),a characteristic feature of apoptotic cells (Darzynkiewicz et al.,1992),compared to annexin V-negative cells (Figure 4B).We then compared the DNA methylation levels of EYA4and IRAK3gene promoter regions in early apoptotic and viable cells.We have previously defined these genes as harboring cancer-specific DNA methylation and differential expression in cancer versus normal cells.Furthermore,these genes were in the top tier for significantly hypermethylated genes,and for gene repres-sion,in colon and lung adenocarcinoma.In addition we also compared the DNA methylation levels of SYCP3and ADAM2gene promoter regions between early apoptotic and viable cells.These genes were identified as having somatic cell-specific DNA methylation and differential gene expression between somatic and germ cells (data not shown).In agreement with our hypoth-esis that DNA methylation-induced silencing of these regions is required for survival,early apoptotic cells showed at least a 27%reduction in DNA methylation in all four regions analyzed,with some specific CpG sites having as much as 80%reduction in DNA methylation (Figures 4C and S2A).Because degradation of cellular mRNA is an early apoptosis-induced event (Del Prete et al.,2002),we could not reliably measure whether this deme-thylation was associated with re-expression of these genes in the dying cells.In contrast,DNA degradation is a late apoptotic event,which allowed us to study the DNA methylation status during the first steps of apoptosis.An alternative hypothesis is that global demethylation in DKO1cells,due to impaired DNA methyltransferase activity,leads to genomic instability and cell death.To test this hypoth-esis,we measured the global DNA methylation levels of the early spontaneous apoptotic and viable cells and did not find a global reduction in DNA methylation (Figure S2B),further sug-gesting that demethylation of these specific genes lead to cell death.In addition to test whether apoptosis itself could cause demethylation of these regions,we treated HCT116cells with 0.2m M of Staurosporine (STS),a drug known to induce cell death by blocking protein kinases (Manns et al.,2011).Next,we sorted viable and STS-induced dead cells and did not observe any difference in DNA methylation of these candidate regions (Figure S2C).Altogether,this strongly suggests that demethylation of these regions is causing cell death rather than the other way around.These data,together with our previous data showing that DKO1cells have reduced cell viability when this low level of DNMT1,and consequently the DNA methylation level,is further reduced by RNAi (Egger et al.,2006;Spada et al.,2007),and that complete knockout of the maintenance DNMT1leads to massive cell death (Chen et al.,2007),demonstrate that these cells are under constant selective pressure to retain DNA methylation at these key regions that we identified here in order to survive.Functional ValidationWe sought to further demonstrate that re-expression of genes whose DNA methylation is critical for cancer cell survival leads to increased cell death.We cloned the cDNA of six genes from the cancer cluster (IRAK3,P2RY14,CDO1,BCHE ,ESX1,and ARMCX1),two from the somatic cluster (ADAM2and SYCP3),and one from the cell line cluster (STEAP4)into the pLJM1lenti-viral vector to individually reexpress these genes in HCT116and RKO colon carcinoma cell lines (Figures S3A and S3B).We observed that expression of each of these genes decreased cell viability in both HCT116and RKO cells (Figures 5A and 5B).We also re-expressed NOX4as a control gene (Figures 5A and 5B).NOX4was heavily methylated in HCT116(b value of 0.95)and completely demethylated in DKO1(b value of 0.007),sug-gesting that DNA methylation-mediated repression of this gene is not necessary for DKO1survival.It is important to note that these ten genes have a low relative expression in RKO (Fig-ure S3B)and a very high basal DNA methylation level in this cell line (Figure S3E).plotted on the x axis,and the p value for a FDR-corrected t test of differences between the normal and the tumor tissues (À13log10scale)is plotted on the y axis.Probes that are significantly (p <0.05)downregulated in tumor tissues are shown in red.(C)Volcano plot of the CpG loci identified as cancer-specifically methylated in lung adenocarcinoma (normal n =4,cancer n =19)from TCGA Data Portal.The b value difference in DNA methylation between the tumor samples and the correspondent normal samples is plotted on the x axis,and the p value for a FDR-corrected Wilcoxon rank sum test of differences between the tumor and correspondent normal samples (À13log10scale)is plotted on the y axis.Probes that are significantly hypermethylated (FDR adjusted p <0.05)in tumors are shown in red.(D)Volcano plot gene expression data of cancer-specific DNA-methylated genes.Gene expression data were obtained from GEO (GSE7670)from primary lung adenocarcinoma (n =27)and primary lung (n =30).For the volcano plots,gene expression fold change between the normal tissues and the tumor tissues is plotted on the x axis,and the p value for a FDR-corrected t test of differences between the normal and the tumor tissues (À13log10scale)is plotted on the y axis.Probes that are significantly (p <0.05)downregulated in tumor tissues are shown in red.(E)Venn diagram showing the overlap between the genes statistically hypermethylated in colon adenocarcinoma (n =50;FDR adjusted p <0.05)and lung adenocarcinoma (n =33;FDR adjusted p <0.05).(F)Venn diagram showing the overlap between the genes statistically repressed in colon adenocarcinoma (n =44;FDR adjusted p <0.05)and lung adeno-carcinoma (n =25;FDR adjusted p <0.05).Cancer CellDNA Methylation DriversCancer Cell 21,655–667,May 15,2012ª2012Elsevier Inc.661。

  1. 1、下载文档前请自行甄别文档内容的完整性,平台不提供额外的编辑、内容补充、找答案等附加服务。
  2. 2、"仅部分预览"的文档,不可在线预览部分如存在完整性等问题,可反馈申请退款(可完整预览的文档不适用该条件!)。
  3. 3、如文档侵犯您的权益,请联系客服反馈,我们会尽快为您处理(人工客服工作时间:9:00-18:30)。

these proteins (e.g. point mutations, copy number abnormalities, or chromosomaltranslocation), or by other mechanisms (e.g. epigenetic mechanisms or upstream oncogenicmutations). Despite this central importance in the development and maintenance of cancer,few apoptosis-targeted therapeutics have reached clinical evaluation.Of particular importance is the BCL2 family of proteins. Highly conserved from worm tohuman, these proteins control the activation of downstream caspases, which are the majoreffectors of apoptosis. The BCL2 family can be divided into three main subclasses, definedin part by the homology shared within four conserved regions termed BCL2 homology (BH)domains (Adams and Cory, 2007; Danial and Korsmeyer, 2004). The “multidomain” pro-apoptotic members BAX and BAK possess BH1-3 domains, and together constitute arequisite gateway to the intrinsic apoptosis pathway (Lindsten et al., 2000; Wei et al., 2001).In contrast, the pro-apoptotic proteins, such as BIM, PUMA and NOXA, share homologyonly within the BH3 amphipathic α-helical death domain, prompting the title “BH3-only”.Anti-apoptotic family members such as BCL2, BCL-xL and MCL1 show conservation in allfour BH domains. The BH1, BH2 and BH3 domains of those proteins are in close proximityand create a hydrophobic pocket that can accommodate the BH3 domain of a pro-apoptoticmember (Danial and Korsmeyer, 2004; Petros et al., 2004).Despite overwhelming genetic and functional evidence implicating the BCL2-familyproteins as therapeutic targets, effective therapeutic inhibitors of these proteins have beendifficult to develop. Elegant NMR-based structural biology efforts led to development of the small-molecule BCL2/BCL-xL inhibitor ABT-737 (Oltersdorf et al., 2005) and its analogABT-263, now in early clinical trials (Tse et al., 2008). While it is expected that ABT-263or related compounds will have clinical activity in BCL2- or BCL-xL-dependent tumors, itis clear that many tumors do not depend on these proteins, but rather rely on other anti-apoptotic factors such as MCL1 (Lin et al., 2006; van Delft et al., 2006).MCL1 has only recently been recognized as an important therapeutic target in cancer.MCL1 is highly expressed in a variety of human cancers (Krajewska et al., 1996a;Krajewska et al., 1996b). Its expression has been linked to tumor development (Zhou et al.,2001) and resistance to anti-cancer therapies. For example, over-expression of MCL1 is amajor resistance mechanism for the experimental BCL2/BCL-xL inhibitor ABT-737 (Chenet al., 2007; Keuling et al., 2009; van Delft et al., 2006), and MCL1 has been similarlyimplicated in the resistance of non-BCL2-family-targeted therapy (Wei et al., 2006).Importantly, we recently reported that amplification of the MCL1 locus is one of the mostfrequent somatic genetic events in human cancer, further pointing to its centrality in thepathogenesis of malignancy (Beroukhim et al., 2010). While the development of MCL1inhibitors has been of considerable interest, no such inhibitors have yet reached the clinic. Aparticularly promising strategy, however, was recently reported by Walensky andcolleagues, whereby ‘stapled’ helical MCL1 BH3 peptides function as effective MCL1inhibitors in pre-clinical models (Stewart et al., 2010). Whether such stapled peptides willmake for effective clinical therapeutics remains to be established. Furthermore, nobiomarkers for patient selection have been discovered for MCL1 inhibitors. Therefore, weused a chemical genomic strategy to identify MCL1-downregulating small-molecules and todiscover biomarkers of MCL1 dependency.RESULTSGene-expression-based high-throughput screen identifies small-molecules repressing MCL1 expressionMCL1 is frequently amplified in human cancers (Beroukhim et al., 2010), and is highlyexpressed across a panel of 729 human cancer cell lines (Figure S1A). We hypothesized thatNIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscriptit might be possible to discover small-molecules that decrease MCL1 expression, thereby activating the apoptosis cascade in MCL1-dependent tumors. We therefore developed anassay to profile the mRNA levels of MCL1 and 48 other apoptosis-related genes using the Luminex bead-based method (Hieronymus et al., 2006; Peck et al., 2006) (Figure 1A, Table S1). We profiled many apoptosis-related genes in addition to MCL1 in order to identifycompounds that preferentially repress MCL1 while preserving expression of the pro-apoptotic factors.We carried out a pilot screen using MCF7 breast cancer cells treated with 2,922 small-molecule compounds including 530 FDA approved drugs. We used MCF7 cells, which are deficient in caspase-3, to avoid identifying compounds that repress MCL1 expressionthrough feedback apoptosis mechanisms. We also performed the assay at an early time point (8 hours post-treatment) for this reason. We counter-screened against compounds thatcaused significant cell death at 8 hours using an LDH viability assay, reasoning that such compounds must not be acting by classical apoptosis-inducing mechanisms.Twenty-four compounds (0.8%) decreased MCL1 expression at least two-fold (Figure 1B).All 24 compounds reduced MCL1 expression more than any of the other 48 apoptosis-related genes assayed, suggesting at least some degree of preferential activity against MCL1.We selected 14 commercially available compounds for further testing. Seven of theseexhibited significant dose-related repression of MCL1 expression. The 7 compoundsincluded the natural product triptolide, the transcription inhibitors 5,6-dichlorobenzimidazole riboside (DRB) and actinomycin D, the kinase inhibitor 5-iodotubercidin, and the anthracyclines doxorubicin, daunorubicin and epirubicin. Despite having distinct reported mechanisms of action (Table S2), treatment of these compoundsresulted in decreased MCL1 expression in multiple cell lines, suggesting a commonmechanism of MCL1 repression across cancer types (Figure S1B).Small molecules that repress MCL1 share transcriptional profilesWe compared genome-wide expression profiles of cells following treatment with candidate compounds to det1ermine whether they shared a common mechanism of action. Weperformed genome-wide gene expression profiling in MCF7 cells following treatment with triptolide and actinomycin D. The expression changes induced by triptolide and actinomycinD were highly similar (R 2=0.85), suggesting that, like actinomycin D, triptolide likelyfunctions as a transcriptional inhibitor (Figure 1C). Consistent with this observation,triptolide was recently reported to bind to XPB, a subunit of TFIIH (Titov et al., 2011), and inhibit phosphorylation of the C-terminal tail of RNA polymerase II, which results intranscriptional inhibition (Leuenroth and Crews, 2008).Using the Connectivity Map database containing expression profiles of 1,366 compounds (/cmap ) (Lamb et al., 2006), the triptolide induced profile showed a high degree of similarity to both doxorubicin and daunorubicin (ranked 1 and 2 of 1,366,respectively, using Spearman correlation) (Figure 1D). The anti-cancer effect ofanthracyclines has long been attributed to inhibition of DNA topoisomerase II (Desmedt et al., 2011; Moretti et al., 2009). However, the DNA topoisomerase II inhibitor etoposideinduced a transcriptional profile distinct from that induced by triptolide (Figure 1D). Taken together, these results strongly suggest that the compounds that emerged from our MCL1-repression screen, including the anthracyclines, function as global transcriptional repressors.We therefore refer to them as Transcriptional Repressor (TR) compounds.Strikingly, the TR compounds showed dramatic preferential activity against MCL1compared to the rest of the transcriptome. For example, MCL1 was in the top 0.05 percentile of triptolide-repressed genes (Figure 1E-1G), and the MCL1 transcript was repressed moreNIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscriptthan 5-fold within 2 hours of treatment (Figure 1E). On the contrary, none of the otherBCL2 family genes was repressed more than two fold. Consistent with the reported shorthalf-life of MCL1 protein (30 minutes) (Adams and Cooper, 2007), inhibition of MCL1mRNA caused a rapid decrease in MCL1 protein levels that occurred prior to PARPcleavage, a marker for caspase activation (Figure S1C).TR compounds share a pattern of cell killing and can be rescued by physiologically relevant levels of MCL1Based on the shared mechanisms suggested above, we hypothesized that if MCL1 repression is a biologically-relevant target of TR compounds, then these compounds should induceapoptosis in the same cancer cell lines. We therefore measured caspase activation and cell viability of 74 non-small cell lung cancer (NSCLC) and 33 breast cancer cell lines following treatment with actinomycin D, doxorubicin, triptolide, and flavopiridol. Flavopiridol haspreviously been reported to repress MCL1 expression via inhibition of CDK9 (Chen et al.,2005).Responses to the TR compounds were highly correlated when measured both by caspaseactivation and cell viability (Pearson r >0.82 and 0.93, respectively, when compared totriptolide) (Figure 2A-B). As expected, cell viability was highly correlated with caspaseactivation for each TR compound (Figure 2C, Pearson r >0.59), indicating that the TRcompounds impair cell viability via apoptosis. By contrast, compounds that kill cells viadifferent mechanisms, such as methotrexate and etoposide, demonstrated different patterns of cytotoxicity (Figure 2A-B, S2A). Despite the fact that TR compounds repress theexpression of many genes, ectopic expression of physiological levels of MCL1 rescued cells from TR compound treatment (Figure 2D-F). In contrast, ectopic expression of MCL1 had no such rescue effect for other classes of compounds, such as methotrexate (Figure 2D-F).If TRs block global transcription, we hypothesized that combination treatment with TRcompounds would counteract the effects of compounds that kill cells by inducing theexpression of pro-apoptotic proteins. The proteasome inhibitor bortezomib inducesapoptosis through the induction of the pro-apoptotic protein NOXA (Gomez-Bougie et al.,2007; Voortman et al., 2007). As predicted, treatment with the TR compounds doxorubicin,actinomycin D or triptolide rescued cells from the apoptotic effects of bortezomib whereas treatment with the non-TR compound etoposide had no effect (Figure S2B-F). Similarly, the TR compounds were able to rescue cells from the histone deacetylase (HDAC) inhibitorvorinostat (Figure S2G), which kills cells via the induction of the pro-apoptotic proteinsBMF and NOXA (Wiegmans et al., 2011).MCL1 knockdown phenocopies TR compoundsIn order to determine whether MCL1 repression explains the activity of TR compounds, we tested whether their effects could be phenocopied by knockdown of MCL1. We treated 17breast and 16 NSCLC cancer cell lines representing different levels of sensitivity to TRcompounds with each of the five most effective shRNAs selected from a library of 60 anti-MCL1 shRNAs (Figure 3A). The response to the 5 MCL1 shRNAs was highly correlated (R 2>0.64 for breast cell lines and R 2>0.55 for NSCLC cell lines) (Figure 3B). Ectopicexpression of MCL1 with a heterologous 3’ UTR at physiologically relevant levels was able to rescue cells from the 2 MCL1 shRNAs targeting the 3’ UTR of MCL1, but not the 3MCL1 shRNAs targeting the coding region of MCL1 (Figure 3C), indicating that theircellular effects are most likely due to MCL1 repression as opposed to off-target effects.In addition, we generated shRNAs against BCL-xL to test whether MCL1-dependent cells were sensitive to knockdown of other anti-apoptotic genes. The responses to the 5 most NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscripteffective BCL-xL shRNAs (out of the 24 shRNAs tested) were highly correlated (Figure 3D-E, S3A), but these responses did not correlate with response to the MCL1 shRNAs (R 2=0.002) (Figure 3F, S3B).Impaired viability induced by doxorubicin was strongly correlated with the effects of MCL1shRNAs (R 2=0.80 for breast cancer cells (Figure 3G) and R 2=0.74 for NSCLC). Conversely,doxorubicin sensitivity did not correlate with the effects of shRNAs targeting BCL-xL (R 2=0.0001 for breast cancer cell lines) (Figure 3H). Furthermore, doxorubicin did not induce additional significant cell death after MCL1 knockdown, consistent with MCL1repression being a major effector of doxorubicin action (Figure 3I-J). Triptolide yielded similar results, suggesting that this is a general property of TR compounds (Figure S3C).Taken together, these results further support the notion that a subset of tumor cells aredependent upon MCL1 for survival, and that TR compounds act largely via MCL1repression.Discovering predictive biomarkers of MCL1 essentialityWe next sought to discover biomarkers that are predictive of MCL1 essentiality bycomparing TR compound sensitivities with genomic data. Such biomarkers would prove useful for the prediction of sensitivity to any present or future MCL1 inhibitors. Wedeveloped an analytical method to infer groups of compounds that induce sensitivity in similar cancer genetic subtypes and infer predictive biomarkers of sensitivity to eachcompound group. Briefly, the method uses an expectation-maximization (EM) algorithm and iterates until convergence between clustering groups of compounds based on thesimilarity of their response profiles and uses an elastic net algorithm to infer a predictive model for each group based on its genetic features (Lee et al., 2009). The method further employs a bootstrapping procedure to obtain a parsimonious model containing only robustly predictive features (Figure 4A, also see Supplemental Experimental Procedures for details).We examined the genetic features (copy number and expression data for >18,000 genes, and mutation data for 34 genes) across 72 cell lines for which we had TR compound sensitivity measurements. To ensure that our predicted biomarkers were specific to sensitivity induced by the TR compounds, we also performed dose-response measurements on 37 additional control compounds (Table S2). The algorithm identified a cluster of compounds consisting of all of the TR compounds (actinomycin D, doxorubicin, flavopiridol, and triptolide), as well as 3 additional compounds (puromycin, emetine, and anisomycin) that function asglobal repressors of protein translation (Figure 4B). Similar to MCL1 mRNA, the extremely short half-life of MCL1 protein likely explains the selective effects of protein translation inhibitors on MCL1 activity.The predictive model of sensitivity to the group of transcriptional and translationalrepressors contained only a single feature, corresponding to mRNA expression of BCL-xL .Specifically, low expression of BCL-xL was associated with sensitivity and high expression of BCL-xL was associated with resistance to compounds that repress MCL1 expression. The half-life of BCL-xL protein is much longer than that of MCL1 (Figure S1C), consistent with its ability to prevent apoptosis induced by transcriptional and translational inhibitors. Also consistent with this observation, sensitivity to MCL1 shRNAs anti-correlated with BCL-xL mRNA levels in the 17 breast cancer cell lines (R 2=0.57) (Figure 4C).We next sought to derive a computational model for the causal interactions that explain how MCL1 and BCL-xL influence sensitivity to TR compounds. We applied the ARACNEreverse engineering algorithm (Basso et al., 2005; Margolin et al., 2006), which is designed to deconvolute direct and indirect interactions among a set of covariates, and derived a network of direct interactions among variables corresponding to gene expression and copyNIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscriptnumber of MCL1 and BCL-xL and sensitivity to TR compounds. We used as input to the algorithm a matrix of values across the panel of 72 cell lines, corresponding to normalized expression and copy number of MCL1 and BCL-xL , as well as sensitivity to the TRcompounds, computed as the average of normalized IC50 values across all TR compounds.This approach yielded a model in which expression of BCL-xL was indeed the directpredictor of sensitivity to TRs (Figure 4D). As expected, gene expression of BCL-xL and MCL1 was directly influenced by the copy number of the respective genes (Figure 4E-F).Interestingly, the model indicated an epistatic relationship between MCL1 copy number and BCL-xL expression. MCL1 copy number was negatively correlated with BCL-xLexpression (Figure 4G), suggesting that MCL1 amplification may decrease the selective pressure requiring BCL-xL for inhibition of apoptosis.Sequestration of pro-apoptotic proteins by MCL1 and BCL-xLThe above data suggested that breast and lung cancer cells with low expression of BCL-xL rely on MCL1 to sequester pro-apoptotic proteins. Upon repressing of MCL1 protein levels,pro-apoptotic proteins might be released from MCL1 and cause downstream caspaseactivation and apoptosis. BIM binds to all anti-apoptotic proteins (Merino et al., 2011). In a panel of 19 NSCLC cell lines, in cells expressing low levels of BCL-xL, depletion of MCL1by immunoprecipitation resulted in depleting nearly the entirety of BIM (Figure 5A-B). In contrast, in cells expressing high levels of BCL-xL, only a small fraction of BIM wassequestered by MCL1 (Figure 5A-B). Furthermore, when BCL-xL was over-expressed in cells that normally have low levels of BCL-xL, the fraction of BIM bound by MCL1decreased significantly (Figure 5C). These experiments demonstrate a shuttling of BIMsequestration between MCL1 and BCL-xL, depending on their relative expression levels. To explore whether the release of BIM from MCL1 explains the apoptotic effect of MCL1-repressing TR compounds, we repeated the MCL1-BIM co-immunoprecipitationexperiments under conditions of TR treatment. Surprisingly, despite the TR compounds triptolide or flavopiridol significantly reducing MCL1 levels, the majority of BIM protein remained bound to the residual MCL1 (Figure S4A-B). In addition, BIM knockdown by shRNA did not abrogate the sensitivity to TR compounds (Figure S7C-G), although we cannot exclude the possibility that more complete BIM knock-down might have a more dramatic effect.Because BIM seemed unlikely to be the principle pro-apoptotic mediator of MCL1repression, we considered other candidate proteins. MCL1 co-immunoprecipitationexperiments showed that while the majority of PUMA, BAK and BAX proteins were not bound by MCL1 (Figure 5A, S4A), significant amounts of PUMA and BAK were pulled down by MCL1, and overexpression of BCL-xL disrupted this interaction (Figure 5C,D).MCL1-bound PUMA decreased after triptolide-mediated MCL1 repression, but this result is best explained by triptolide’s concomitant repression of PUMA expression (Figure 5D). To test the possibility that BAK release from MCL1 explains the TR effect, we used Bak −/−MEFs to determine contribution of Bak in TR compound-induced apoptosis. Bak deletion nearly completely rescued cells from TRs but did not protect cells from the non-TRcompound trichostatin A (Figure 5E-H). BAX and BAK are both multidomain pro-apoptotic BCL2 family proteins. However, BAK proved unique in that we did not detect MCL1-BAX interaction in co-immunoprecipitation experiments (Figure 5C,D), and Bax −/− cells were not rescued from TR compounds (Figure 5E-G). Taken together, our data suggest that MCL1protects cells from cell death at least in part through sequestration of BAK, and thissequestration is diminished with TR compound-mediated MCL1 repression.NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author ManuscriptBCL-xL predicts MCL1 dependency in vivoAn important question in developing biomarkers of MCL1 dependency is whether resistance mechanisms observed in vitro hold in vivo , where tumor-microenvironment interactions are known to modulate apoptotic mechanisms (Bewry et al., 2008; Buggins et al., 2010). We therefore examined the in vivo response of two NSCLC cell lines grown as xenografts in NOD-SCID mice. H1437 cancer cells express low levels of BCL-xL and are sensitive to triptolide (as a prototype MCL1 repressor) in vitro . HCC15 cells, in contrast, express high levels of BCL-xL and are triptolide-resistant in vitro . This pattern of sensitivity persisted in vivo . Triptolide significantly attenuated the growth of the H1437 NSCLC cancer model (Figure 6A-B). By contrast, in the HCC15 xenograft model, triptolide did not significantly affect tumor volume or survival of the mice (Figure 6C-D). Western blotting of whole tumor lysates demonstrated that treatment with triptolide decreased MCL1 protein abundance and increased PARP cleavage in the H1437 xenograft model (Figure 6E), indicated thattriptolide repressed MCL1 expression and induced apoptosis in vivo .Our model predicts that patients with high levels of BCL-xL expression are resistant to TRs.To test this hypothesis, we investigated the relationship between BCL-xL gene expression and clinical response to neoadjuvant treatment with the anthracycline epirubicin in 114estrogen receptor-negative breast cancer patients for which it was determined whether a complete pathological response (pCR) was achieved (Desmedt et al., 2011). BCL-xLshowed significant differential expression between patients who achieved pCR and those who did not (Figure 6F). As previously reported, expression of topoisomerase 2A (TOP2A)did not correlate with response to epirubicin (Figure S5), consistent with our finding that anthracyclines kill tumor cells via a transcriptional repressive mechanism rather than via a topoisomerase inhibitory mechanism as has been generally assumed.BCL-xL is a functional determinant of MCL1 dependencyWe next investigated whether BCL-xL was simply a marker of MCL1 dependency or it was a functional determinant of response. Overexpression of BCL-xL in MCL1-dependent lines protected them from apoptosis induced by MCL1 shRNAs or TR compounds (Figure 7A-C),but not by other cytotoxic agents such as methotrexate (Figure 7B-C), suggesting a specific effect for TR compounds. Conversely, BCL-xL knockdown conferred sensitivity in cell lines otherwise resistant to TR compounds. Cell lines resistant to treatment with TRcompounds (using doxorubicin as a representative example) were sensitive to combined treatment with BCL-xL shRNAs (Figure 7D,G), and cell lines resistant to treatment with MCL1 shRNAs were sensitive to combined treatment with the BCL-xL inhibitor ABT-263(Figure 7E,G). The viability of cells treated with BCL-xL shRNAs was highly correlated with viability after treatment with the BCL-xL inhibitor ABT-263, and combined treatment of cells with ABT-263 and BCL-xL shRNAs did not yield synergistic effects (Figure 7F-G).The above data suggest that TR compounds would exhibit a synergistic effect when used in combination with BCL-xL inhibitors. We treated a panel of 74 NSCLC cancer cell lines with a 42-point dose-response matrix (6 concentrations of triptolide or Actinomycin D, and 7 concentrations of ABT-263 or ABT-737). We examined the synergy between TRcompounds and BCL-xL inhibitors for each cell line by computing the excess growthinhibition over the Bliss independence model for each combination of compoundconcentrations (Figure 8A-C). Cell lines that were highly sensitive to TR compoundsshowed no evidence of synergy when treated in combination with ABT-737. Cell lines that were resistant to TR compounds and to BCL-xL inhibitors were sensitive to the combination (Figure 8A-C).NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author ManuscriptA synergy score was computed for each combination experiment in each of the 74 NSCLC cell lines by summing the excess over Bliss independence across all dose combinations. The synergy score was averaged over the 4 combination experiments, performed by pairing triptolide or actinomycin D with ABT-263 or ABT-737. This synergy score was highly correlated with expression of BCL-xL (Figure 8D), suggesting that high expression of BCL-xL determines the synergistic relationship between TR compounds and BCL-xL inhibitory compounds, and that resistance to TR compounds, caused by high expression of BCL-xL ,can be overcome by treating in combination with BCL-xL inhibitors. Consistent with this notion, ABT-263 released BAK from BCL-xL (Figure 8E).DISCUSSIONAt an accelerating pace, the genomic characterization of human cancer is elucidating the molecular basis of the disease. Recent large-scale analyses of gene copy number in cancer demonstrated that the genes encoding the BCL2-family proteins MCL1 and BCL-xL are frequent targets of amplification. Low-level MCL1 amplification is particularly notable,representing one of the most common copy number abnormalities in all of human cancer (Beroukhim et al., 2010). In support of a functionally important role of MCL1, numerous studies have elucidated the critical role of MCL1 in preventing tumor cell death (Warr and Shore, 2008).Using a multiplexed Luminex bead-based assay, we screened for compounds that reduced MCL1 expression while preserving the expression of proapoptotic genes. Although the compounds that emerged from this screen were general transcriptional repressor (TR)compounds (as opposed to specifically targeting the MCL1 locus), they preferentially repressed MCL1 because of the short half-life of MCL1 mRNA and protein. Multiple lines of evidence suggest that TR compounds induce apoptosis in cancer cells primarily through repression of MCL1 expression including 1) upon treatment with TR compounds, MCL1protein levels decreased rapidly and preceded caspase activation; 2) ectopic expression of physiological levels of MCL1 rescued cancer cells from TR compounds, despite theexpression of other genes still being repressed; 3) the pattern of TR compound sensitivity across a panel of cancer cell lines closely mirrored the pattern of sensitivity of those cell lines to MCL1 knock-down by RNAi; 4) of over 40,000 genomic features measured, the top feature that predicted sensitivity to TR compounds was the low expression of BCL-xL ,which shares redundant function with MCL1; 5) Ectopic expression of BCL-xL rescued cancer cells from TR compounds; 6) MCL1 repression by TR compounds resulted in the release of, pro-apoptotic protein BAK protein from MCL1; and 7) Bak deficiency protected cells from TR compounds. These results suggest that the mechanism of cell death induced by TR compounds is best explained by MCL1 inhibition.This indicated that some of the widely used chemotherapeutic drugs such as anthracyclines may preferentially repress MCL1 to induce apoptosis in tumor cells. Although the anti-tumor effect of anthracyclines has long been speculated to be related to the drug’s inhibition of DNA topoisomerase II (Desmedt et al., 2011; Moretti et al., 2009), and an association between low TOP2A expression and anthracycline response in ER-negative breast cancer patients has been reported (Martin, 2011), our data suggest that their activity may be largely explained by inhibition of transcription, leading most dramatically to the repression of short-lived MCL1 transcripts. While it is possible that multiple mechanisms of action explain the anti-tumor effects of anthracyclines, at least in the experimental cancer models studied here,anthracycline gene expression consequences most reflected transcriptional inhibition rather than DNA topoisomerase II inhibition. Furthermore, the similar pattern of sensitivity of cell lines to MCL1 knockdown compared to anthracycline treatment is also consistent with an MCL1-mediated transcriptional inhibitory effect. Lastly, our observation that BCL-xLNIH-PA Author ManuscriptNIH-PA Author ManuscriptNIH-PA Author Manuscript。

相关文档
最新文档