Open access peer-reviewed chapter

Post-Transcriptional Control of RNA Expression in Cancer

Written By

Carlos DeOcesano-Pereira, Fernando Janczur Velloso, Ana Claudia Oliveira Carreira, Carolina Simões Pires Ribeiro, Sheila Maria Brochado Winnischofer, Mari Cleide Sogayar and Marina Trombetta-Lima

Submitted: 19 July 2017 Reviewed: 23 October 2017 Published: 21 February 2018

DOI: 10.5772/intechopen.71861

From the Edited Volume

Gene Expression and Regulation in Mammalian Cells - Transcription From General Aspects

Edited by Fumiaki Uchiumi

Chapter metrics overview

1,664 Chapter Downloads

View Full Metrics

Abstract

Approximately 80% of the human genome contains functional DNA, including protein coding genes, non-protein coding regulatory DNA elements and non-coding RNAs (ncRNAs). An altered transcriptional signature is not only a cause, but also a consequence of the characteristics known as the hallmarks of cancer, such as sustained proliferation, replicative immortality, evasion of growth suppression and apoptotic signals, angiogenesis, invasion, metastasis, evasion of immune destruction and metabolic re-wiring. Post-transcriptional events play a major role in determining this signature, which is evidenced by the fact that alternative RNA splicing takes place in more than half of the human genes, and, among protein coding genes, more than 60% contain at least one conserved miRNA-binding site. In this chapter, we will discuss the involvement of post-transcriptional events, such as RNA processing, the action of non-coding RNAs and RNA decay in cancer development, and how their machinery may be used in cancer diagnosis and treatment.

Keywords

  • post-transcriptional control
  • splicing
  • microRNAs
  • long non-coding RNAs
  • mRNA decay

1. Introduction

The word cancer defines a group of diverse diseases, which share unique traits. Tumor cells display mechanisms of sustained proliferation, replicative immortality, evasion of growth suppression and apoptotic signals, angiogenesis, invasion, metastasis, evasion of immune destruction and metabolic re-wiring [1]. These characteristics represent a great challenge to cancer treatment being both a cause and a consequence of an abnormal gene expression profile. Efforts to understand the consequences of these different expression profiles and the mechanisms underlying them contribute to clarify cancer biology and, consequently, to predict response to and optimization of therapeutic approaches [2, 3, 4].

There are several layers of gene expression modulation including epigenetics, transcriptional modulation, RNA expression control, translational regulation and post-translational modifications. All these mechanisms work in an orchestrated manner leading to specific expression signatures and phenotypes. In this chapter, we focus on RNA expression control mechanisms, which take place after RNA polymerase recognition of the gene promoter and start of RNA synthesis, discussing their implications to malignant transformation and cancer progression.

Advertisement

2. mRNA processing

RNA processing takes place after the start of transcription, resulting in a mature mRNA which is able to fulfill its function. This process comprises: 5′-Cap addition, splicing and poly(A) addition. RNA splicing is a process in which portions of the pre-RNA, denominated introns, are excised and the remaining portions (exons) are bound to form the mature RNA. Both cis and trans elements act to recognize exon/intron boundaries and/or to orchestrate the splicing machinery, the spliceosome, a complex of five small nuclear ribonucleoprotein particles (snRNP) and 100–200 non-snRNP proteins which catalyze the splicing reaction [5, 6, 7]. Recognition of the intron/exon boundaries is context-dependent; as a result, a single gene can originate several mature RNAs and, therefore, several proteins with independent or even opposite functions. This alternative splicing (AS) occurs by recognition of the alternative donor or acceptor splice sites, exon inclusion or exclusion, intron incorporation or combinatory mechanisms as mutually exclusive exons and so on. AS stands out as a major source for transcripts and proteins variability, occurring in approximately 59% of human genes [8] and almost 95% of the multi-exon genes [9]. Splicing factor genes are commonly mutated in different types of cancer and several splice variants have already been implicated in cancer development [10].

The splicing profile of a certain tissue changes dramatically when compared with malignant cells with their normal counterparts [11, 12, 13]. This difference may result from mutations or single-nucleotide polymorphisms (SNPs) on acceptor, donor splice sites, enhancing or silencing sequences which lead to alterations in the exon/intron boundary recognition; or due to deregulated expression or change of function mutation in a trans regulator (reviewed in [14, 15]). Serine-rich protein (SRP) and heterologous nuclear ribonuclear particle (hnRNP) are two protein families which are classically involved in splicing modulation by interacting with intronic or exonic enhancer or silencer sequences [16, 17]. The SRSF1 member of the SRP family is one of the most well characterized splice factor, being described as up-regulated in lung [18] and breast cancers [19, 20]. In the breast cancer model, SRSF1 association to a sequence near to a donor splice site usually promotes exon inclusion, while its association in the vicinities of an acceptor splice site leads to exon skipping or inclusion [20]. Important cancer-related gene transcripts, such as Casp9 [21], CD44 [22] and VEGF [23], are among SRSF1 known targets.

Cell survival outcome is a perfect example of the influence of AS in basic cellular mechanisms, with alternative isoforms of several apoptotic-related gene transcripts displaying opposite roles, when compared to their canonical variant, shifting the cell status from apoptosis-prone to the survival state (reviewed in [24]). Upon an apoptotic stimulus, cytochrome C is released from the mitochondria and forms a complex with Apaf-1. The N-terminal portion of Apaf-1 interacts with the N-terminal pro-domain of pro-caspase-9, leading to Caspase-9 activation, which, in turn, activates the Caspase-3 and -7 effector proteases (reviewed in [25]). Caspase-9, a key player in this process, has an alternative-splicing variant in which exclusion of the exon cassette 3, 4, 5 and 6 leads to a protein isoform which lacks part of its large subunit. This Caspase-9b isoform retains the domain which interacts with Apaf-1, but lacks the Caspase-9 catalytic site, thus acting like a dominant negative and inhibiting the apoptotic pathway [26, 27]. The ratio between these two isoforms modulates the propensity of the cells to respond to death stimuli, altering their chemo-sensitivity and, potentially, the treatment’s outcome. Interestingly, while Akt mediates exclusion of the exon cassette via phosphorylation of the RNA splicing factor SRp30a [28]; in this case, SRSF1 interacts with an intronic enhancer site at intron 6 favoring the exon cassette inclusion, which renders the cells more sensitive to chemotherapeutic agents as the combined therapy with daunorubicin and erlotinib [21]. Taking into account that SRSF1 is upregulated in non-small cell lung cancer cells, this case exemplifies the complexity of splicing as an expression regulator and how it can be explored to optimize therapy efficacy.

Another great source of transcripts variability is alternative polyadenylation (APA), since approximately 30% of human mRNAs display alternative polyadenylation sites [29]. Polyadenylation occurs in almost every mammalian transcript, a process in which an endonucleolytic cleavage is catalyzed by polyadenylation machinery proteins, immediately followed by polyadenylation (200–300 nucleotides, on average, in humans) of the 3′-end by poly(A) polymerases (reviewed in [30]). The resulting alternative transcripts will have different sizes, depending on the localization of the alternative poly(A) site, originating alternative 3′-untranslated regions (3′-UTR). Also, more rarely, when polyadenylation occurs inside the open reading frame region, it may originate truncated forms of the translated protein [31]. The 3′-UTR is extremely important to transcripts stability, localization and regulation by trans elements (such as miRNAs and RNA binding proteins), topics to be further discussed in this chapter and which have great implications for cancer development.

A shift in the polyadenylation global pattern occurs in tumor cells, with the proximal poly(A) sites being favored, when compared to their normal counterparts [29]. Also, highly proliferative murine T lymphocytes favor shorter 3′-UTRs, which is also observed in colorectal cancer, but only for certain groups of genes, including those involved in cell cycle, nucleic acid-binding and processing factors. It has been proposed that such shortening would restrict miRNA modulation over the transcripts, increasing their expression [32, 33]. Such a mechanism is observed upon treatment of ER+ breast cancer cells with the proliferation stimulant 17β-estradiol. This treatment leads to APA of the CD6 transcript, which is essential for the start of DNA replication, originating a shorter 3′-UTR. The generated CD6 variant is resistant to repression dependent on its 3′-UTR and is more efficiently translated, correlating with a higher rate of BrdU incorporation by the cells [34].

Curiously, mammalian RNAs can also be post-transcriptionally modified through a process called RNA editing. Well-known cases are the RNA editing enzymes adenosine and cytidine deaminases, which catalyze the conversion of adenine into inosine and of cytosine into uracil, respectively [35]. Adenosine deaminases acting on RNA (ADAR) enzymes act on double-stranded RNA regions, usually the secondary structure of a single mRNA molecule. Through a hydrolytic deamination at C6, ADAR enzymes catalyze adenine conversion into inosine, which pairs with cytosine. Cytidine deaminases are much more specific and different members of the APOBEC3 family are transcriptionally regulated by p53 [36]. Altered RNA editing signatures were found in different types of tumors, such as glioblastoma [37], breast [38] and gastric cancers [39, 40]. If located at a coding region, these editing events may cause a missense mutation. One example is ADAR-1 editing of the Antizyme Inhibitor 1 (AZIN1), which leads to a serine-to-glycine substitution at residue 367 [41]. AZIN1 is an inactive homolog of ornithine decarboxylase (ODC) that competitively binds to antizymes [42]. ADAR-1 editing increases AZIN1 affinity to antizyme, leading to a decrease in ODC antizyme-mediated degradation and promoting polyamines biosynthesis, with consequent cell proliferation and a more aggressive behavior in hepatocellular carcinoma cells [41]. Although editing on consensus splicing sites are rare, ADAR enzymes alter the global splicing pattern of the cell by editing splicing regulatory cis elements and, possibly, indirectly, by altering the activity of trans elements [43, 44].

The interaction of transcripts with long non-coding RNAs (lncRNAs) and microRNAs are important post-transcriptional regulatory mechanisms which will be further addressed in this chapter. RNA edition adds a layer of complexity to this apparatus. It is estimated that over 70% of potential editing sites within long non-coding RNAs may lead to changes in their secondary structure, a feature which is crucial for its target recognition [45]. If the editing takes place in a precursor miRNA, it can lead to alterations in its biosynthesis and target recognition, increasing their range of action [46, 47, 48]. Alterations in the mRNA 3′-UTR may alter its recognition by a specific miRNA or lncRNA [37, 40, 47]. Furthermore, RNA editing may also modulate RNA expression by regulating RNA decay. This is exemplified by the ADAR-1 interaction with the RNA binding protein HuR, which promotes HuR binding to the target transcript, increasing its stability [49].

Advertisement

3. miRNAs

Several RNA-based mechanisms evolved in eukaryotes to modulate gene expression or suppress invading material. In animals, the small non-coding RNAs (18–30 nucleotides) are subdivided into three major classes, namely microRNA (miRNA), small interfering RNA (siRNA) and PIWI-interacting RNA (piRNA). The main purpose of piRNAs are suggested to be silencing of transposable elements in germline cells [45], siRNAs and miRNAs seem to have evolved from an antiviral defense system into an ubiquitous gene expression modulation mechanism [46, 47]. Originally identified in Caenorhabditis elegans [48], miRNAs are the dominating class of small RNAs in most somatic tissues, being highly conserved and repressing the expression of target genes by inhibiting mRNAs translation and/or stability [49, 50]. The latest update of the human miRNA database lists 2588 mature miRNAs, processed out of 1881 precursors [51]. miRNA genes are originally transcribed by RNA polymerase II (Pol II) as a long (typically over 1 kb) primary transcript (pri-miRNA) bearing hairpins, in which miRNA sequences are embedded [52]. Hairpins are cropped by the Drosha nuclear RNase III liberating the stem-loop shaped ~65 nucleotide long precursor miRNA (pre-miRNA) [53]. Upon exporting to the cytoplasm through Exportin 5 (EXP5), pre-miRNAs are cleaved by DICER near the terminal loop, liberating a small RNA duplex [54]. This duplex is subsequently loaded onto RNA-induced silencing complex (RISC), RNP effector complexes containing Argonaut (AGO) proteins. Finally, unwinding of the RNA duplex allows the final single-stranded miRNA to act as a guide for the effector complex [55]. Specific targeting is accomplished by base pairing between mRNA and miRNA, as miRNAs usually guide RISC to 3′UTR regions in target protein-coding transcripts [56], recruiting proteins that lead to target RNA degradation, deadenylation or decay [53]. However, miRNAs may also interact with 5′UTR and coding sequence (CDS) regions, culminating in a range of effects, from translational activation to repression.

More than 60% of human protein-coding genes contain at least one conserved miRNA-binding site [57], encompassing every major cellular functional pathway. Therefore, miRNAs biogenesis needs to be under tight temporal and spatial control, and their deregulation is evidently associated with a wide range of human diseases, including cancer [58]. The first instance of the direct involvement of a miRNA in cancer was uncovered in 2002. A critical region at chromosome 13q14, frequently deleted in chronic lymphocytic leukemia (CLL), was shown to harbor miRNA genes miR-15a and miR-16-1. About 70% of CLL cases have null or reduced expression of these miRNAs, which normally control apoptosis by targeting BCL-2 [59, 60]. The following years revealed a remarkable number of additional examples, establishing the association of miRNAs and cancer to be the norm, rather than the exception. Currently, hundreds of human miRNAs are associated to the onset and progression of several malignancies, including lymphomas, colorectal carcinoma, breast cancer, lung cancer, thyroid cancer and hepatocellular carcinomas [61].

Several miRNAs may be differentially expressed in cancer patients, when compared to normal samples, acting either as oncogenes or tumor suppressors [62] ( Table 1 ). Most often, miRNAs are detected as tumor suppressors, with reduced expression in tumors when compared to normal tissues [63, 64]. These miRNAs have commonly been shown to negatively regulate protein-coding oncogenes. Thus, HER2 and HER3, two oncogenes which are significantly correlated with decreased disease-specific survival in breast cancer patients [65], are suppressed by miR-125a or miR-125b [66]. Additionally, the let-7 family of miRNAs targets several genes associated with cell cycle and cell division, including the RAS oncogene [67]. Inhibition of epidermal growth factor receptor by miR-128b in non-small cell lung cancer (NSCLC) [68] and miR-7 in glioma [69] are additional pertinent examples of miRNAs acting as tumor suppressors. However, several miRNAs have also been found to be overexpressed in cancer, being classified as oncomiRs, often repressing known tumor suppressors. Thus, overexpression of miR-155 and miR-21 is sufficient to induce lymphomagenesis in mice [70, 71].

miRNA Cancer phenotype Target mRNA Cancer association References
miR-15a Tumor suppressor BCL2 Chronic lymphocytic leukemia [59, 60]
miR-16-1 Tumor suppressor BCL2 Chronic lymphocytic leukemia [59, 60]
miR-125a Tumor suppressor HER2/HER3 Breast cancer [66]
miR-125b Tumor suppressor HER2/HER3 Breast cancer [66]
let-7 Tumor suppressor RAS Lung tumor [67]
miR128-b Tumor suppressor EGFR Non-small lung cancer [68]
miR128-b Tumor suppressor EGFR Acute lymphoblastic leukemia [77]
miR-7 Tumor suppressor EGFR Glioma [69]
miR-155 Oncogenic BIC Lymphoma [70, 71]
miR-21 Oncogenic NA Lymphoma [70, 71]
miR-127 Tumor suppressor BCL6 Prostate cancer [75, 76]
miR-372/373 Oncogenic RAS, p53 Testicular germ cell tumor [170]
miR-17 Tumor suppressor c-MYC Large B-cell lymphoma [72, 171]
miR-34 Tumor suppressor P53 Ovarian cancer [73]
miR-210 Tumor suppressor DIMT1 Multiple myeloma [172]
miR-10b Tumor suppressor TIAM1 Gastric cancer [173]
miR-126 Tumor suppressor ADAM9 Breast cancer [174]
miR-335 Tumor suppressor BRCA1 Breast cancer [175]

Table 1.

List of miRNAs involved in cancer and their respective mRNA targets.

Mapping efforts have revealed that many miRNAs are located in fragile regions of the genome, which are deleted, amplified or translocated in cancer, directly altering miRNAs genes expression, hence leading to aberrant expression of downstream target mRNAs [59]. In addition to genomic alterations, miRNA expression is also modulated by tumor suppressor or oncogenic factors, which function as transcriptional activators or repressors to control pre-miRNA transcription. One of the first examples of this interaction is the transcriptional upregulation of the miR-17/92 cluster by the c-myc oncogene product, counterbalancing the apoptotic activity of E2F1 and allowing c-Myc mediated-proliferation [72]. Likewise, p53 stimulates transcription of the miR-34 family, inducing apoptosis and senescence. Loss of p53 function induces downregulation of the miR-34 family in a very high percentage of ovarian cancer patients with a p53 mutation [73]. The expression of miRNA genes may also be indirectly modulated. Aberrant epigenetic changes, such as DNA hypermethylation of tumor suppressor genes, extensive genomic DNA hypomethylation and alteration of histone modification patterns, are a well-known feature of cancer cells. In fact, epigenetic modifications represent another common mechanism related to the alteration of miRNA expression in cancer. Tumor-suppressing miRNAs are usually found to be hypermethylated in cancer, which, in turn, allows overexpression of their oncogenic targets [74]. Thus, epigenetic repression of the tumor-suppressor miR-127 in primary prostate cancer [75] and bladder tumor causes upregulation of its target transcripts, including that of the proto-oncogene BCL6 [76]. A cancer-driving alteration may arise early in the biogenesis of miRNAs, during transcription of the pri-miRNA. For example, a point mutation in miR-128b gene blocks processing of pri-miR-128b and reduces the levels of mature miR-128b, thus leading to glucocorticoid resistance in acute lymphoblastic leukemia (ALL) [77]. Another mechanism which can lead to an aberrant expression of miRNAs and, thus, to cancer, is the altered expression and/or function of the enzymes involved in the biogenesis of microRNAs, such as DROSHA and DICER. Aberrant expression of these proteins affects the biogenesis of all miRNAs in the cell, influencing the regulation of a multitude of genes. Thus, the first heterozygous germline mutations in DICER1 were identified as causing pleuropulmonary blastoma (PPB), a rare pediatric lung tumor that arises during fetal lung development [78]. Likewise, decreased expression of DROSHA and DICER has been found in 39% of ovarian cancer patients [79]. miRNA biogenesis may also be modulated during nuclear translocation by exportin 5 (XPO5). XPO5 mutations in some tumors generate pre-miRNA accumulation in the nucleus, reducing miRNA maturation and availability in the cytoplasm [80]. miRNA processing is orchestrated by a large number of proteins assisting the basic machinery. Several of these modulatory proteins, such as DDX5 and DDX17, were shown to be either directly mutated or to serve as targets for oncoproteins or tumor suppressors, modulating miRNA biogenesis [81].

The functional outcomes of miRNAs deregulation coincide with the hallmarks of malignant cells, namely: (1) self-sufficiency in growth signals (let-7 family), (2) insensitivity to anti-growth signals (miR-17-92 cluster), (3) apoptosis evasion (miR-34a), (4) limitless replicative potential (miR-372/373 cluster), (5) angiogenesis (miR-210) and (6) invasion and metastases (miR-10b). miRNAs have also been shown to regulate the generation of cancer stem cells (CSCs) [82, 83] and epithelial-mesenchymal transition (EMT), paramount for the metastatic process [84]. Thus, as breast cancer cells metastasize, expression of miR-126 and miR-335 is lost. Overexpressing these miRNAs in cancer cells decreases lung and bone metastasis in vivo [85].

The high number of human miRNAs, regulating a wide range of cancer-related processes, renders these small non-coding RNAs an ideal profiling tool. miRNA expression profiles can distinguish not only between normal and cancerous tissue, but also help to discriminate different subtypes of a particular cancer, or even specific oncogenic abnormalities [86], increasing the accuracy of tumor classification. These expression profiles were able to classify tumors according to their tissue of origin with accuracy higher than 90%. miRNAs regulation of cancer progression also allows these molecules to serve as efficient predictors of prognosis, tumor metastasis and therapy selection. Specific miRNA signatures have recently been shown to correlate to metastatic breast and colon tumors, arising as potent biomarkers to predict metastatic outcome. miRNA profiles may also be applied to select for more personalized and efficient therapies and to adjust the therapeutic scheme during treatment to achieve a better outcome. Noteworthy, in ovarian cancer, miRNA signatures are able to predict chemo-resistant tumors, while a polymorphism (SNP34091), which creates a new binding site for miR-191, was suggested as a modulator of tumor chemosensitivity [75].

miRNAs are highly stable molecules present in body fluids including plasma, blood, serum, urine, saliva and milk, being potential cancer biomarkers which may be found in different phases of the tumoral process [87, 88]. Although understanding of how miRNAs are selectively released from cells and how circulating miRNAs are related to disease remains largely unclear, circulating miRNAs may serve as novel diagnostic and prognostic biomarkers for human diseases, including cancer [89].

Advertisement

4. Long non-coding RNAs

Recent studies based on the Encyclopedia of DNA elements (ENCODE) project indicate that more than 80% of the human genome contains functional DNA that includes protein coding genes, non-protein coding regulatory DNA elements and non-coding RNAs (ncRNAs) [90]. Non-coding RNAs is a class of genetic regulators, containing short (<200 nucleotides) and long (>200 nucleotides) transcripts with novel abilities to be used as biomarkers due to their role in disease development and their implications for genomic organization [91, 92]. Short ncRNAs include ribosomal RNAs (rRNAs), transfer RNAs (tRNAs), small nuclear RNAs (snRNAs) and small nucleolar RNAs (snoRNAs). Regulatory long non-coding RNAs (lncRNAs) have been found in a large variety of organisms, ranging from yeasts to mammals, including mice and humans [93]. lncRNAs have emerged as a fundamental molecular class whose members play critical roles in genome regulation and in tissue development and maintenance [92]. Based on their positions relative to the protein coding genes in the genome, lncRNAs can be classified into natural antisense transcripts (NATs), long intronic ncRNAs and long intergenic ncRNAs (lincRNAs) [93].

Recent transcriptional profiling of multiple human tissues, including both normal and tumor samples, has led to the assumption that misregulation of lncRNAs could disrupt these delicate processes and lead to tumorigenesis [94, 95, 96, 97]. These studies have validated the tissue-specific expression of lncRNAs in normal tissues, and have identified large sets of lncRNAs which are aberrantly expressed in either a specific cancer or multiple types of cancer, suggesting these RNAs act as master regulators of gene expression [98, 99]. Differential expression of lncRNAs is increasingly recognized as a hallmark feature in cancer [100]. lncRNAs are a novel class of mRNA-like transcripts, which contribute to cancer development and progression, accelerating cancer cells proliferation, apoptosis, invasion and metastasis [101] ( Table 2 ).

LncRNA Cancer phenotype Molecular mechanism Cancer association References
HOTAIR Oncogenic, promotes metastasis and invasion Interacts with PRC2 and LSD1 complex, promotes silencing of HOX genes in trans epigenetically Overexpressed in liver, breast, lung and pancreatic tumors [109, 176, 177]
GAS5 Tumor suppressor, induces growth arrest and sensitizes cells to apoptosis Inhibits and binds glucocorticoid receptor (GR) from activating target genes Downregulated in breast cancer [178, 179]
H19 Oncogenic, promotes cell proliferation and tumor growth Unknown Breast cancer [180]
MALAT1 Oncogenic, promotes cell proliferation and metastasis Related to alternative splicing and active transcription, regulation of gene expression Overexpressed in lung, breast, pancreatic, colon, prostate and hepatocellular carcinomas [117, 181, 182]
MEG3 Tumor suppressor, inhibits cell proliferation and induces apoptosis Enhancing p53’s transcriptional activity on its target genes. Controls expression of gene loci through recruitment of PRC2 Downregulated in multiple tumor types [183, 184]
PTENP1 Tumor suppressor; Inhibits cell proliferation, migration, invasion and tumor growth Binds and inhibits miRNAs from targeting and repressing PTEN Locus lost in prostate cancer, colon cancer and melanoma [185, 186, 187]
ZFas1 Tumor suppressor and inhibits proliferation Unknown Breast cancer and dysregulated in many types of tumors [128, 188]

Table 2.

List of lncRNAs involved in cancer with their proposed functions.

General mechanisms of lncRNA function implicated in cancer progression are associated with a wide-repertoire of biological processes. Among the main biological pathways, lncRNAs may be involved in epigenetic silencing, splicing regulation, translational control, regulation of apoptosis and cell cycle control [102]. Like protein-coding genes, lncRNAs can function as oncogenes or tumor suppressors. Many lncRNAs shuttle between the nucleus and the cytoplasm, suggesting that they may have dual functions, while others are restricted to the nucleus [103]. In the nucleus, lncRNAs are often part of the nuclear architecture and, in some cases, are critical for maintenance of sub-nuclear structures [104].

lncRNAs bind to and target chromatin regulators allowing connection between RNA and chromatin, acting on the control of gene expression at the transcriptional level [105]. Moreover, several lncRNAs mechanistic themes have emerged, both at the transcriptional and post-transcriptional levels, such as decoys, scaffolds and guides [106]. Examples of the mechanisms of action of some lncRNAs on the control of gene expression and mammalian cells regulation are described below.

HOTAIR (Hox transcript antisense intergenic RNA) is expressed from the HOXC locus and was the first lncRNA shown to be acting in trans. HOTAIR binds to and targets the PRC2 complex to the HOXD locus [107], functioning as an RNA scaffold containing two main functional domains. The 5′ domain of HOTAIR binds PRC2, whereas a 3′ domain binds the LSD1/CoREST/REST H3K4 demethylase complex [108], thus bridging two repressive complexes in order to coordinate their functions in gene silencing. Ectopic HOTAIR expression in epithelial cancer cells induces genome-wide retargeting of PRC2, leading to widespread changes in repressive (H3K27me3) and active (H3K4me3) chromatin markers, resembling those found in embryonic fibroblasts. This results in more invasive and metastatic cells and HOTAIR expression is predictive of cancer survival [109].

lncRNAs can also participate in global cellular behavior by controlling cell growth. The growth-arrest-specific 5 (GAS5) lncRNA sensitizes the cell to apoptosis by regulating the activity of glucocorticoids in response to nutrient starvation [110]. GAS5 binds to the DNA-binding domain (DBD) of the glucocorticoid receptor (GR), where it acts as a decoy, preventing GR interaction with cognate glucocorticoid response elements (GRE). Under normal conditions, GR target genes are involved in apoptosis suppression, such as cellular inhibitor of apoptosis 2 (cIAP2) and inhibit the cell-death executioners caspases 3, 7 and 9 [111]. However, upon growth arrest, GAS5 activation compromises GR ability to bind to the cIAP2 GRE, reducing cIAP2 expression levels, thereby removing its suppressive effect on caspases [110]. GAS5 has also been associated with breast cancer because its transcript levels are significantly reduced, when compared to unaffected normal breast epithelium [110]. Therefore, GAS5 could act as a tumor suppressor if reduced levels of this lncRNA are unable to maintain sufficient caspase activity to activate an appropriate apoptotic response in disease-compromised cells.

H19 is an imprinted gene expressed exclusively from the maternal allele, which maintains silencing of IGF2. H19 is highly expressed in a wide variety of solid tumors. The majority of cancers express high levels of H19 when compared to normal tissues. H19 is generally overexpressed in stromal cells, rarely in tumor epithelial cells and has been found to be associated with the presence of estrogen receptor (ER) and progesterone receptor (PR) [112]. Data indicating both oncogenic and tumor suppressive roles for H19 in different cancers are available [113]. In cancer cell lines, H19 RNA expression is directly regulated by E2F1, promoting cell cycle progression [114].

The lncRNA MALAT1 (metastasis associated in lung adenocarcinoma transcript) was identified in an attempt to characterize transcripts associated with early stage non-small cell lung cancer (NSCLC) [115]. Some studies found that MALAT1 regulates alternative splicing through its interaction with the serine/arginine-rich (SR) family of nuclear phosphoproteins, which are involved in the splicing machinery [116, 117]. Because the SR family of proteins affects the alternative splicing patterns of many pre-mRNAs, its activity must be tightly regulated. Small changes in SR protein concentration or phosphorylation status can upset the fragile balance that controls mRNA variability among different cells and tissue types [118]. Therefore, the lncRNA MALAT1 has been suggested to serve as a fine-tuning mechanism to modulate the activity of SR proteins.

The maternally expressed gene 3 (MEG3) is an imprinted lncRNA located on chromosome 14q32 is expressed exclusively from the maternal allele. MEG3 has been shown to activate p53 and facilitate p53 signaling, including enhancement of p53 binding to target genes [119]. Furthermore, MEG3 regulates genes of the TGF-β pathway through formation of RNA-DNA triplex structures [120]. MEG3 overexpression in meningioma, hepatocellular carcinoma and breast cancer cell lines leads to suppression of cell proliferation [121, 122, 123].

The PTEN (phosphatase and tensin homolog) gene encodes a tumor suppressor that functions by negatively regulating the AKT/PKB signaling pathway [124, 125]. Mutations of this gene constitute a step into the development of many cancers and it is one of the most commonly lost tumor suppressors in human cancer [126]. A highly homologous processed of PTENP1 (phosphatase and tensin Homolog pseudogene 1) is a pseudogene which is associated with the lncRNA class found on chromosome 9, regulating PTEN by both sense and antisense RNAs. This long non-coding RNA acts as a decoy for PTEN, targeting microRNAs and exerting a tumor suppressive activity [125, 127].

The lncRNA Zfas1 (Znfx1 antisense 1) is a transcript antisense to the 5′ end of the protein-coding gene Znfx1, which has functions in epithelial cells and was identified in large-scale studies aimed at isolating differentially expressed genes during mammary development [128]. Zfas1 intronically hosts three C/D box snoRNAs (Snord12, Snord12b and Snord12c) [128] and recently has been associated with ribosomes cancer cells [129].

The highly specific lncRNA expression signatures render them as attractive markers for accurate disease diagnosis and patients prognosis. In addition, advancement of RNA-based therapeutics opens new avenues for lncRNAs as new targets for cancer therapy.

Advertisement

5. mRNA decay

mRNA degradation is an important mechanism for post-transcriptional control of gene expression, controlling both the quality and the abundance of cellular mRNAs. Deadenylation of the mRNA is the default process, often representing a rate-limiting step in cytoplasmic mRNA decay, in which the poly(A) tail of the transcript is degraded through recruitment of deadenylase complexes [130, 131, 132]. In the literature, different deadenylases or poly(A)-specific ribonucleases have been described, namely PARN (poly(A)-specific ribonuclease), Pan2/Pan3 (poly(A) nuclease 2/3) complex and CCR4–NOT (carbon catabolite repression 4) complex [131, 133]. The PARN deadenylase is involved in destabilization of different transcripts related to cell cycle progression and cell proliferation [133, 134], as well as in degradation of oncogenic miRNAs, such as miR-21 [135]. In addition, its expression is altered in different tumors, such as gastric tumors [136] and acute leukemias [137].

Different proteins are able to interact with each other and promote the recruitment of deadenylases to the mRNA poly(A) tail. Members of BTG/Tob family, associated with anti-proliferative activities [138], are able to associate with both Caf1a and Caf1b (enzymatic subunits of the CCR4-NOT complex) [139], and, also, with PABPC1 (cytoplasmic poly(A)-binding protein) [139, 140], promoting mRNA poly(A) tail removal and cytoplasmic mRNA decay. Expression of the BTG/Tob proteins is classically associated with inhibition of cell cycle progression [138]. The Tob/Caf1 complex is also involved in the negative regulation of c-myc proto-oncogene expression by accelerating deadenylation and decay of its mRNA [141]. In addition, BTG2 has been characterized as a p53 transcriptional-target, being an essential component for suppression of Ras-induced transformation by p53 [142]. In agreement, reduced expression of BTG2 and TOB proteins are observed in human samples derived from different types of tumor [143, 144, 145, 146]. On the other hand, interaction of Tob1 with Caf1a (but not with Caf1b) was recently associated with the metastatic phenotype in mouse mammary carcinoma model and the deadenylase activity of Caf1a was shown to be required for promotion of metastatic disease [147]. Using a human breast cancer model, it has also been shown that high expression of either TOB1 or CNOT1 (the scaffold subunit of the CCR4-NOT complex) correlated with poor survival [147] and was associated with poor distant metastasis free survival in breast cancer patients [148]. Interestingly, PABPC1 has also been described as an oncogenic protein in gastric carcinoma. Zhu and collaborators showed that PABPC1 is upregulated in gastric carcinoma tissues, predicting poor survival and inhibits apoptosis by targeting miR-34c [149]. Following shortening of the poly(A) tail, mRNA can either be degraded through the 3′ pathway, by the eukaryotic exosome complex, or, alternatively, by removal of the cap by Dcp2 and exonuclease decay through the 5′ pathway, promoted by exonuclease Xrn1 [130, 131].

AU-rich elements (ARE) are critical cis-acting elements in the 3′-UTRs of a variety of short-lived transcripts. Tristetraprolin (TTP) and human antigen R (HuR) are two important RNA-binding proteins which can bind to AREs in their target mRNAs. TTP promotes deadenylation and degradation of target mRNAs, whereas HuR, as already mentioned, is involved in stabilization of target mRNAs. It has been extensively described that TTP expression is significantly decreased in different types of tumors [150] and that it is involved in cell cycle control, angiogenesis and tumor metastasis [151]. Recently, it has been reported that TTP inhibits the epithelial-mesenchymal transition (EMT) of cancer cells through mRNA degradation of the EMT inducers, specifically, Twist1 and Snail1, and inhibits cell proliferation through downregulation of c-fos, CDC34 and VEGF [152]. Interestingly, TTP appears to bind to AREs and interact with proteins involved in mRNA decay, such as the PM-scl75 exosome component, Xrn1 5′–3′ exonuclease, CCR4deadenylase and Dcp1 decapping enzyme [153], supporting a model in which TTP promotes mRNA decay through the ability to recruit components of the cellular mRNA decay machinery to the target mRNAs. In recent publications, high expression levels of HuR have been correlated with tumor progression and aggressiveness by affecting cell cycle progression, migration, invasion, metastasis and apoptosis in different tumor models [154, 155, 156, 157]. HuR enhances the stability of the human epidermal growth factor receptor 2 (ERBB2/HER-2) mRNA, modulating the estrogen receptor-alpha-positive (ER+) breast cancer cells responsiveness to tamoxifen [158].

In addition, deadenylase complexes could be recruited to the mRNA poly(A) tail through the action of miRNAs. GW182 proteins, which participate of the miRNA-induced silencing complex (miRISC), directly interact with PAN3 and NOT1 subunits, leading to recruitment of the PAN2-PAN3 and CCR4-CAF1-NOT deadenylase complexes to the 3′-UTR of target mRNAs [159]. Also, it has been described that PARN deadenylase binds to the 3′ UTR of p53 mRNA through recruitment mediated by miR-125b-loaded miRISC, promoting p53 mRNA decay [134]. Interestingly, this effect can be reverted by HuR proteins, which bind to the p53 AREs and increase p53 mRNA stability [134].

The deadenylation machinery is also an important target for antitumor agents and anticancer therapy. Cantharidin (an inhibitor of protein phosphatase 2A) inhibits the invasive ability of pancreatic cancer cells, with concomitant deadenylation-dependent degradation of MMP2 mRNA [20]. Resveratrol (3,5,4′-trihydroxystilbene), a naturally occurring compound, induces TPP expression in U87MG human glioma cells and leads to the decay of urokinase plasminogen activator (uPA) and urokinase plasminogen activator receptor (uPAR) mRNAs, promoting suppression of cell growth and inducing apoptosis [160].

Additionally, several mature mRNAs surveillance mechanisms guarantee quality and fidelity to encode a functional protein in a translation-dependent manner. The nonsense-mediated decay (NMD) pathway is the best understood surveillance mechanism; detecting and degrading transcripts which contain premature termination codons (PTCs), avoiding the expression of semi-functional and truncated proteins [161]. The UPF-1 (up-frameshift1) protein, a key component of the NMD mechanism, interacts with both Dcp2 and PARP, linking NMD with the decapping and deadenylation processes [162]. Low expression levels of UPF-1 protein as well as inactivation of UPF-1 function were described in several types of human cancer, suggesting that NMD downregulation is related to tumorigenesis. Decreased levels of UPF-1 were detected in lung adenocarcinoma in comparison to normal tissues, and its downregulation was correlated to poor prognosis and higher histological grade [163]. The pancreatic adenosquamous carcinoma (ASC) is an aggressive tumor which is associated with high metastatic potential and poor prognosis. In these tumors, a mutation that promotes UPF-1 alternative splicing and results in a non-functional UPF-1 protein, has been observed. Inactivation of the NMD pathway promotes selective accumulation of a p53 isoform, which acts in a dominant-negative manner, contributing to tumorigenesis [164].

NMD can also be inhibited by a wide variety of cellular stresses, some of which are associated to the tumoral context [165]. In response to stress events, phosphorylation of the alpha-subunit of the eukaryotic initiation factor 2 (eIF2α) is able to inhibit NMD. It has been described that phospho-eIF2α is necessary for oncogene c-myc-mediated NMD inhibition [106]. Inhibition of NMD by cellular stress promotes stabilization of the SLC7A11 mRNA, which encodes a subunit of the cystine/glutamate aminoacid transport system, leading to increased intracellular levels of cysteine, accelerating the production of glutathione. SLC7A11 is upregulated in hypoxic cells, promotes tumorigenesis and chemotherapy resistance, suggesting that it could be an adaptive response that protects tumor cells against oxidative stress [166]. It has recently been described that NMD regulates the epithelial-mesenchymal transition (EMT) in the lung adenocarcinoma model, by targeting the TGF-β signaling pathway [163]. In addition, the NMD mechanism controls the expression of a novel human E-cadherin variant mRNA produced by alternative splicing. Overexpression of this alternatively spliced E-cadherin variant in MCF-7, breast cancer cells was able to induce EMT by promoting higher expression levels of Twist, Snail, Zeb1 and Slug, with a concomitant decrease in the wild type E-cadherin mRNA levels [167].

Several promising NMD targets mRNAs for cancer therapy have been proposed. The MDM4 protein, which is undetectable in normal tissues, is frequently upregulated in cancer cells, acting by inhibiting the p53 tumor-suppressor function [168]. The abundance of the MDM4 protein is controlled, at least in part, by alternative splicing mechanisms and the NMD pathway. In most normal adult tissues, the lack of exon 6 in the Mdm4-spliced variant leads to the production of an unstable transcript (Mdm4-S), which contains a PTC and is targeted to NMD [168]. On the other hand, the oncogenic splicing-factor SRSF3 supports exon 6 inclusion in the Mdm4 mRNA transcript (full-length Mdm4 variant), which is not efficiently degraded by NMD. Therapeutic strategies which lead to antisense oligonucleotide-mediated (ASO-mediated) Mdm4 exon 6 skipping efficiently decreases MDM4 abundance and inhibits tumor cell growth in melanoma and diffuse large B cell lymphoma models, as well as increases sensitivity to MAPK-targeting therapies [169].

Advertisement

6. Final considerations

Different post-transcriptional mechanisms have been associated with gene expression control, leading to complex transcriptional signatures in cancer. The mechanisms presented in this chapter constitute fine regulators of gene expression which influence multiple and highly relevant pathways in cancer development (summarized in Figure 1 ). Several splicing variants, miRNAs and lncRNAs, have been shown to act as possible oncoRNAs or as tumor suppressors. The functional roles of these RNAs are only beginning to be elucidated providing an uncharted resource for the development of diagnostic methods and novel cancer therapies.

Figure 1.

Schematic representation and key roles of different RNA species in the control of gene expression in mammalian cells. This scheme represents a genomic locus and the main molecular mechanisms associated with the control of gene expression pattern. Proximal control elements are located close to the promoter, while distal elements (called enhancers) may be far away from a gene or even located in an intron. Alternative splicing (AS) generates transcriptome diversity. During AS, cis-acting regulatory elements, present in the pre-mRNA sequence, determine which exons are retained and which exons are spliced out. For an individual pre-mRNA, several alternative exons show different types of alternative-splicing patterns. Addition of 5’ Cap and Poly(A) tail are controlled events which are extremely important for the stability of the mRNA and its transport from the cytoplasm to the nucleus. Non-coding RNAs (ncRNAs) with regulatory functions can act in multiple pathways during the transcription process by controlling specific events which culminate in synthesis of different proteins. Long non-coding RNAs (lncRNAs) target protein complexes to specific genomic loci affecting transcription patterns (transcriptional interference), leading to chromatin modifications (interplay between epigenetic marks, such as DNA methylation and histone acetylation) and DNA polymerase II activity. Advances in transcriptomics have resulted in the discovery of large numbers of ncRNAs (miRNAs e lncRNAs), many of which display the capacity to regulate gene expression at the levels of transcription (control of AS), post-transcription (mRNA editing, mRNA decay and mRNA stability) and translation (translation initiation).

Advertisement

Abbreviations

ADARAdenosine deaminases acting on RNA
AGOArgonaut
Akt/PKBProtein kinase B
Apaf-1Apoptotic protease activating factor 1
APOBECApolipoprotein B Mrna editing enzyme, catalytic polypeptide-like
AREAU-rich elements
ASAlternative splicing
ASCPancreatic adenosquamous carcinoma
ASOAntisense oligonucleotide
AZIN1Antizyme inhibitor 1
BCLB cell lymphoma gene family
BrdUBromodeoxyuridine (5-bromo-2′-deoxyuridine)
BTGBTG anti-proliferation factor
Caf1Chromatin assembly factor-1 complex
CaspCaspase
CCR4C-C motif chemokine receptor 4
CCR4–NOTCarbon catabolite repression 4 complex
CD44CD44 molecule (Indian blood group)
CD6Cluster of differentiation 6
CDC34Cell division cycle 34
CDSCoding DNA sequence
c-fosProto-oncogene c-Fos
cIAP2Cellular inhibitor of apoptosis 2
CLLChronic lymphocytic leukemia
c-MycMyc proto-oncogene
CNOT1CCR4-NOT transcription complex subunit 1
CoRESTREST corepressor 1
CSCsCancer stem cells
DBDDNA-binding domain
Dcp1Decapping protein 1
DDXDEAD-box helixases
DICERDicer 1, ribonuclease III
DROSHADrosha ribonuclease III
E2F1E2F transcription factor 1
eIF2αEukaryotic initiation factor 2
EMTEpithelial-mesenchymal transition
ENCODEEncyclopedia of DNA elements
EREstrogen receptor
ER+ Estrogen receptor-alpha-positive
ERBB2/HERHuman epidermal growth factor receptor 2
EXP5Exportin 5
GAS5Growth-arrest-specific 5
GRGlucocorticoid receptor
GREGlucocorticoid response elements
H19H19, imprinted maternally expressed transcript
H3K4Histone H3 lysine 4
hnRNPHeterologous nuclear ribonuclear particle
HOTAIRHox transcript antisense intergenic RNA
HOXCHomeobox C cluster
HuRHuman antigen R
IGF2Insulin-like growth factor 2
lincRNAsLong intergenic ncRNAs
lncRNAslong non-coding RNAs
LSD1Lysine-specific histone demethylase 1
MALAT1Metastasis associated in lung adenocarcinoma transcript
MAPKmitogen-activated kinase-like protein
MDM4MDM4, p53 regulator
MEG3Maternally expressed gene 3
miRISCmiRNA-induced silencing complex
miRNA/miRmicroRNA
MMP2Matrix metalloproteinase 2
NATsNatural antisense transcripts
ncRNAsNon-coding RNAs
NMDNonsense-mediated decay
NSCLCNon-small cell lung cancer
ODCOrnithine decarboxylase
p53Tumor protein p53
PABPC1Cytoplasmic poly(A)-binding protein
PABPC1Poly(A) binding protein cytoplasmic 1
Pan2/Pan3Poly(A) nuclease 2/3 complex
PARNPoly(A)-specific ribonuclease
piRNAPIWI-interacting RNA
Pol IIRNA polymerase II
PPBPleuropulmonary blastoma
PRProgesterone receptor
PRC2Polycomb repressive complex 2
Pri-miRNAmiRNA primary transcript
PTCsPremature termination codons
PTENPhosphatase and tensin homolog
PTENP1Phosphatase and tensin homolog pseudogene 1
RasHRas proto-oncogene, GTPase
RESTRE1-silencing transcription factor
RISCRNA-induced silencing complex
rRNAsRibosomal RNAs
siRNASmall interfering RNA
SLC7A11Solute carrier family 7 member 11
SlugSnail family transcriptional repressor 2
Snail1Snail family transcriptional repressor 1
snoRNAsSmall nucleolar RNAs
SNPsSingle-nucleotide polymorphisms
snRNAsSmall nuclear RNAs
snRNPSmall nuclear ribonucleoprotein particles
SRPSerine-rich protein
SRSF1Serine and arginine-rich splicing factor 1
TGF-βTransforming growth factor beta 1
TobTransducer of ERBB2
tRNAsTransfer RNAs
TTPTristetraprolin
Twist1Twist family BHLH transcription factor 1
uPAUrokinase plasminogen activator
uPARUrokinase plasminogen activator receptor
UPF-1Up-frameshift1 protein
UTRUntranslated region
VEGFVascular endothelial growth factor
XPO5Exportin 5
Xrn15′–3′ exoribonuclease 1
Zeb1Zinc finger E-box binding homeobox 1
Zfas1Znfx1 sntisense 1
Znfx1Zinc finger NFX1-type containing 1

References

  1. 1. Hanahan D, Weinberg RA. Hallmarks of cancer: The next generation. Cell. 2011;144(5):646-674
  2. 2. Zhang L, Zhou W, Velculescu VE, Kern SE, Hruban RH, Hamilton SR, et al. Gene expression profiles in normal and cancer cells. Science. 1997;276(5316):1268-1272
  3. 3. Bouska A, Bi C, Lone W, Zhang W, Kedwaii A, Heavican T, et al. Adult high grade B-cell lymphoma with Burkitt lymphoma signature: Genomic features and potential therapeutic targets. Blood. 2017
  4. 4. Kuang M, Cheng J, Zhang C, Feng L, Xu X, Zhang Y, et al. A novel signature for stratifying the molecular heterogeneity of the tissue-infiltrating T-cell receptor repertoire reflects gastric cancer prognosis. Scientific Reports. 2017;7(1):7762
  5. 5. Castle JC, Zhang C, Shah JK, Kulkarni AV, Kalsotra A, Cooper TA, et al. Expression of 24,426 human alternative splicing events and predicted cis regulation in 48 tissues and cell lines. Nature Genetics. 2008;40(12):1416-1425
  6. 6. Sharp PA. Splicing of messenger RNA precursors. Science. 1987;235(4790):766-771
  7. 7. McManus CJ, Graveley BR. RNA structure and the mechanisms of alternative splicing. Current Opinion in Genetics & Development. 2011;21(4):373-379
  8. 8. Lander ES, Linton LM, Birren B, Nusbaum C, Zody MC, Baldwin J, et al. Initial sequencing and analysis of the human genome. Nature. 2001;409(6822):860-921
  9. 9. Pan Q, Shai O, Lee LJ, Frey BJ, Blencowe BJ. Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nature Genetics. 2008;40(12):1413-1415
  10. 10. Sveen A, Kilpinen S, Ruusulehto A, Lothe RA, Skotheim RI. Aberrant RNA splicing in cancer; expression changes and driver mutations of splicing factor genes. Oncogene. 2016;35(19):2413-2427
  11. 11. Venables JP, Klinck R, Koh C, Gervais-Bird J, Bramard A, Inkel L, et al. Cancer-associated regulation of alternative splicing. Nature Structural & Molecular Biology. 2009;16(6):670-676
  12. 12. Xi L, Feber A, Gupta V, Wu M, Bergemann AD, Landreneau RJ, et al. Whole genome exon arrays identify differential expression of alternatively spliced, cancer-related genes in lung cancer. Nucleic Acids Research. 2008;36(20):6535-6547
  13. 13. Armero VES, Tremblay MP, Allaire A, Boudreault S, Martenon-Brodeur C, Duval C, et al. Transcriptome-wide analysis of alternative RNA splicing events in Epstein-Barr virus-associated gastric carcinomas. PLoS One. 2017;12(5):e0176880
  14. 14. Silipo M, Gautrey H, Tyson-Capper A. Deregulation of splicing factors and breast cancer development. Journal of Molecular Cell Biology. 2015;7(5):388-401
  15. 15. Srebrow A, Kornblihtt AR. The connection between splicing and cancer. Journal of Cell Science. 2006;119(Pt 13):2635-2641
  16. 16. Venables JP, Koh CS, Froehlich U, Lapointe E, Couture S, Inkel L, et al. Multiple and specific mRNA processing targets for the major human hnRNP proteins. Molecular and Cellular Biology. 2008;28(19):6033-6043
  17. 17. Oltean S, Bates DO. Hallmarks of alternative splicing in cancer. Oncogene. 2014;33(46):5311-5318
  18. 18. Jiang L, Huang J, Higgs BW, Hu Z, Xiao Z, Yao X, et al. Genomic landscape survey identifies SRSF1 as a key Oncodriver in small cell lung cancer. PLoS Genetics. 2016;12(4):e1005895
  19. 19. Anczuków O, Rosenberg AZ, Akerman M, Das S, Zhan L, Karni R, et al. The splicing factor SRSF1 regulates apoptosis and proliferation to promote mammary epithelial cell transformation. Nature Structural & Molecular Biology. 2012;19(2):220-228
  20. 20. Anczuków O, Akerman M, Cléry A, Wu J, Shen C, Shirole NH, et al. SRSF1-regulated alternative splicing in breast cancer. Molecular Cell. 2015;60(1):105-117
  21. 21. Shultz JC, Goehe RW, Murudkar CS, Wijesinghe DS, Mayton EK, Massiello A, et al. SRSF1 regulates the alternative splicing of caspase 9 via a novel intronic splicing enhancer affecting the chemotherapeutic sensitivity of non-small cell lung cancer cells. Molecular Cancer Research. 2011;9(7):889-900
  22. 22. Loh TJ, Moon H, Jang HN, Liu Y, Choi N, Shen S, et al. SR proteins regulate V6 exon splicing of CD44 pre-mRNA. BMB Reports. 2016;49(11):612-616
  23. 23. Amin EM, Oltean S, Hua J, Gammons MV, Hamdollah-Zadeh M, Welsh GI, et al. WT1 mutants reveal SRPK1 to be a downstream angiogenesis target by altering VEGF splicing. Cancer Cell. 2011;20(6):768-780
  24. 24. Schwerk C, Schulze-Osthoff K. Regulation of apoptosis by alternative pre-mRNA splicing. Molecular Cell. 2005;19(1):1-13
  25. 25. Green DR, Reed JC. Mitochondria and apoptosis. Science. 1998;281(5381):1309-1312
  26. 26. Seol DW, Billiar TR. A caspase-9 variant missing the catalytic site is an endogenous inhibitor of apoptosis. The Journal of Biological Chemistry. 1999;274(4):2072-2076
  27. 27. Srinivasula SM, Ahmad M, Guo Y, Zhan Y, Lazebnik Y, Fernandes-Alnemri T, et al. Identification of an endogenous dominant-negative short isoform of caspase-9 that can regulate apoptosis. Cancer Research. 1999;59(5):999-1002
  28. 28. Shultz JC, Goehe RW, Wijesinghe DS, Murudkar C, Hawkins AJ, Shay JW, et al. Alternative splicing of caspase 9 is modulated by the phosphoinositide 3-kinase/Akt pathway via phosphorylation of SRp30a. Cancer Research. 2010;70(22):9185-9196
  29. 29. Lin Y, Li Z, Ozsolak F, Kim SW, Arango-Argoty G, Liu TT, et al. An in-depth map of polyadenylation sites in cancer. Nucleic Acids Research. 2012;40(17):8460-8471
  30. 30. Erson-Bensan AE, Can T. Alternative polyadenylation: Another foe in cancer. Molecular Cancer Research. 2016;14(6):507-517
  31. 31. Elkon R, Ugalde AP, Agami R. Alternative cleavage and polyadenylation: Extent, regulation and function. Nature Reviews. Genetics. 2013;14(7):496-506
  32. 32. Sandberg R, Neilson JR, Sarma A, Sharp PA, Burge CB. Proliferating cells express mRNAs with shortened 3′ untranslated regions and fewer microRNA target sites. Science. 2008;320(5883):1643-1647
  33. 33. Morris AR, Bos A, Diosdado B, Rooijers K, Elkon R, Bolijn AS, et al. Alternative cleavage and polyadenylation during colorectal cancer development. Clinical Cancer Research. 2012;18(19):5256-5266
  34. 34. Akman BH, Can T, Erson-Bensan AE. Estrogen-induced upregulation and 3′-UTR shortening of CDC6. Nucleic Acids Research. 2012;40(21):10679-10688
  35. 35. Baysal BE, Sharma S, Hashemikhabir S, Janga SC. RNA editing in pathogenesis of cancer. Cancer Research. 2017;77(14):3733-3739
  36. 36. Menendez D, Nguyen TA, Snipe J, Resnick MA. The cytidine deaminase APOBEC3 family is subject to transcriptional regulation by p53. Molecular Cancer Research. 2017;15(6):735-743
  37. 37. Tomaselli S, Galeano F, Alon S, Raho S, Galardi S, Polito VA, et al. Modulation of microRNA editing, expression and processing by ADAR2 deaminase in glioblastoma. Genome Biology. 2015;16:5
  38. 38. Anantharaman A, Gholamalamdari O, Khan A, Yoon JH, Jantsch MF, Hartner JC, et al. RNA-editing enzymes ADAR1 and ADAR2 coordinately regulate the editing and expression of Ctn RNA. FEBS Letters. 2017
  39. 39. Chan TH, Qamra A, Tan KT, Guo J, Yang H, Qi L, et al. ADAR-mediated RNA editing predicts progression and prognosis of gastric cancer. Gastroenterology. 2016;151(4):637-650.e10
  40. 40. Zhang L, Yang CS, Varelas X, Monti S. Altered RNA editing in 3′ UTR perturbs microRNA-mediated regulation of oncogenes and tumor-suppressors. Scientific Reports. 2016;6:23226
  41. 41. Chen L, Li Y, Lin CH, Chan TH, Chow RK, Song Y, et al. Recoding RNA editing of AZIN1 predisposes to hepatocellular carcinoma. Nature Medicine. 2013;19(2):209-216
  42. 42. Qiu S, Liu J, Xing F. Antizyme inhibitor 1: A potential carcinogenic molecule. Cancer Science. 2017;108(2):163-169
  43. 43. Solomon O, Oren S, Safran M, Deshet-Unger N, Akiva P, Jacob-Hirsch J, et al. Global regulation of alternative splicing by adenosine deaminase acting on RNA (ADAR). RNA. 2013;19(5):591-604
  44. 44. Goldberg L, Abutbul-Amitai M, Paret G, Nevo-Caspi Y. Alternative splicing of STAT3 is affected by RNA editing. DNA and Cell Biology. 2017;36(5):367-376
  45. 45. Ishizu H, Siomi H, Siomi MC. Biology of PIWI-interacting RNAs: New insights into biogenesis and function inside and outside of germlines. Genes & Development. 2012;26(21):2361-2373
  46. 46. Axtell MJ, Westholm JO, Lai EC. Vive la différence: Biogenesis and evolution of microRNAs in plants and animals. Genome Biology. 2011;12(4):221
  47. 47. Ghildiyal M, Zamore PD. Small silencing RNAs: An expanding universe. Nature Reviews. Genetics. 2009;10(2):94-108
  48. 48. Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391(6669):806-811
  49. 49. Bartel DP. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell. 2004;116(2):281-297
  50. 50. Iorio MV, Croce CM. MicroRNAs in cancer: Small molecules with a huge impact. Journal of Clinical Oncology. 2009;27(34):5848-5856
  51. 51. Kozomara A, Griffiths-Jones S. miRBase: Annotating high confidence microRNAs using deep sequencing data. Nucleic Acids Research. 2014;42(Database issue):D68-D73
  52. 52. Lee Y, Jeon K, Lee JT, Kim S, Kim VN. MicroRNA maturation: Stepwise processing and subcellular localization. The EMBO Journal. 2002;21(17):4663-4670
  53. 53. Huntzinger E, Izaurralde E. Gene silencing by microRNAs: Contributions of translational repression and mRNA decay. Nature Reviews. Genetics. 2011;12(2):99-110
  54. 54. Ketting RF, Fischer SE, Bernstein E, Sijen T, Hannon GJ, Plasterk RH. Dicer functions in RNA interference and in synthesis of small RNA involved in developmental timing in C. elegans. Genes & Development. 2001;15(20):2654-2659
  55. 55. Kawamata T, Tomari Y. Making RISC. Trends in Biochemical Sciences. 2010;35(7):368-376
  56. 56. Bartel DP. MicroRNAs: Target recognition and regulatory functions. Cell. 2009;136(2):215-233
  57. 57. Friedman RC, Farh KK, Burge CB, Bartel DP. Most mammalian mRNAs are conserved targets of microRNAs. Genome Research. 2009;19(1):92-105
  58. 58. Ha M, Kim VN. Regulation of microRNA biogenesis. Nature Reviews. Molecular Cell Biology. 2014;15(8):509-524
  59. 59. Calin GA, Sevignani C, Dumitru CD, Hyslop T, Noch E, Yendamuri S, et al. Human microRNA genes are frequently located at fragile sites and genomic regions involved in cancers. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(9):2999-3004
  60. 60. Cimmino A, Calin GA, Fabbri M, Iorio MV, Ferracin M, Shimizu M, et al. miR-15 and miR-16 induce apoptosis by targeting BCL2. Proceedings of the National Academy of Sciences of the United States of America. 2005;102(39):13944-13949
  61. 61. Di Leva G, Calin GA, Croce CM. MicroRNAs: Fundamental facts and involvement in human diseases. Birth Defects Research. Part C, Embryo Today. 2006;78(2):180-189
  62. 62. Malumbres M. miRNAs versus oncogenes: The power of social networking. Molecular Systems Biology. 2012;8:569
  63. 63. Bucay N, Sekhon K, Yang T, Majid S, Shahryari V, Hsieh C, et al. MicroRNA-383 located in frequently deleted chromosomal locus 8p22 regulates CD44 in prostate cancer. Oncogene. 2017;36(19):2667-2679
  64. 64. Volinia S, Galasso M, Costinean S, Tagliavini L, Gamberoni G, Drusco A, et al. Reprogramming of miRNA networks in cancer and leukemia. Genome Research. 2010;20(5):589-599
  65. 65. Wiseman SM, Makretsov N, TO N, Gilks B, Yorida E, Cheang M, et al. Coexpression of the type 1 growth factor receptor family members HER-1, HER-2, and HER-3 has a synergistic negative prognostic effect on breast carcinoma survival. Cancer. 2005;103(9):1770-1777
  66. 66. Scott GK, Goga A, Bhaumik D, Berger CE, Sullivan CS, Benz CC. Coordinate suppression of ERBB2 and ERBB3 by enforced expression of micro-RNA miR-125a or miR-125b. The Journal of Biological Chemistry. 2007;282(2):1479-1486
  67. 67. Johnson SM, Grosshans H, Shingara J, Byrom M, Jarvis R, Cheng A, et al. RAS is regulated by the let-7 microRNA family. Cell. 2005;120(5):635-647
  68. 68. Weiss GJ, Bemis LT, Nakajima E, Sugita M, Birks DK, Robinson WA, et al. EGFR regulation by microRNA in lung cancer: Correlation with clinical response and survival to gefitinib and EGFR expression in cell lines. Annals of Oncology. 2008;19(6):1053-1059
  69. 69. Kefas B, Godlewski J, Comeau L, Li Y, Abounader R, Hawkinson M, et al. microRNA-7 inhibits the epidermal growth factor receptor and the Akt pathway and is down-regulated in glioblastoma. Cancer Research. 2008;68(10):3566-3572
  70. 70. Costinean S, Zanesi N, Pekarsky Y, Tili E, Volinia S, Heerema N, et al. Pre-B cell proliferation and lymphoblastic leukemia/high-grade lymphoma in E(mu)-miR155 transgenic mice. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(18):7024-7029
  71. 71. Medina PP, Nolde M, Slack FJ. OncomiR addiction in an in vivo model of microRNA-21-induced pre-B-cell lymphoma. Nature. 2010;467(7311):86-90
  72. 72. O’Donnell KA, Wentzel EA, Zeller KI, Dang CV, Mendell JT. c-Myc-regulated microRNAs modulate E2F1 expression. Nature. 2005;435(7043):839-843
  73. 73. Corney DC, Flesken-Nikitin A, Godwin AK, Wang W, Nikitin AY. MicroRNA-34b and MicroRNA-34c are targets of p53 and cooperate in control of cell proliferation and adhesion-independent growth. Cancer Research. 2007;67(18):8433-8438
  74. 74. Di Leva G, Garofalo M, Croce CM. MicroRNAs in cancer. Annual Review of Pathology. 2014;9:287-314
  75. 75. Acunzo M, Romano G, Wernicke D, Croce CM. MicroRNA and cancer—A brief overview. Advances in Biological Regulation. 2015;57:1-9
  76. 76. Saito Y, Liang G, Egger G, Friedman JM, Chuang JC, Coetzee GA, et al. Specific activation of microRNA-127 with downregulation of the proto-oncogene BCL6 by chromatin-modifying drugs in human cancer cells. Cancer Cell. 2006;9(6):435-443
  77. 77. Kotani A, Ha D, Schotte D, den Boer ML, Armstrong SA, Lodish HF. A novel mutation in the miR-128b gene reduces miRNA processing and leads to glucocorticoid resistance of MLL-AF4 acute lymphocytic leukemia cells. Cell Cycle. 2010;9(6):1037-1042
  78. 78. Hill DA, Ivanovich J, Priest JR, Gurnett CA, Dehner LP, Desruisseau D, et al. DICER1 mutations in familial pleuropulmonary blastoma. Science. 2009;325(5943):965
  79. 79. Merritt WM, Lin YG, Han LY, Kamat AA, Spannuth WA, Schmandt R, et al. Dicer, Drosha, and outcomes in patients with ovarian cancer. The New England Journal of Medicine. 2008;359(25):2641-2650
  80. 80. Melo SA, Moutinho C, Ropero S, Calin GA, Rossi S, Spizzo R, et al. A genetic defect in exportin-5 traps precursor microRNAs in the nucleus of cancer cells. Cancer Cell. 2010;18(4):303-315
  81. 81. Lin S, Gregory RI. MicroRNA biogenesis pathways in cancer. Nature Reviews. Cancer. 2015;15(6):321-333
  82. 82. Peter ME. Regulating cancer stem cells the miR way. Cell Stem Cell. 2010;6(1):4-6
  83. 83. Shimono Y, Zabala M, Cho RW, Lobo N, Dalerba P, Qian D, et al. Downregulation of miRNA-200c links breast cancer stem cells with normal stem cells. Cell. 2009;138(3):592-603
  84. 84. Adam L, Zhong M, Choi W, Qi W, Nicoloso M, Arora A, et al. miR-200 expression regulates epithelial-to-mesenchymal transition in bladder cancer cells and reverses resistance to epidermal growth factor receptor therapy. Clinical Cancer Research. 2009;15(16):5060-5072
  85. 85. Tavazoie SF, Alarcón C, Oskarsson T, Padua D, Wang Q, Bos PD, et al. Endogenous human microRNAs that suppress breast cancer metastasis. Nature. 2008;451(7175):147-152
  86. 86. Iorio MV, Croce CM. microRNA involvement in human cancer. Carcinogenesis. 2012;33(6):1126-1133
  87. 87. Cheng L, Sun X, Scicluna BJ, Coleman BM, Hill AF. Characterization and deep sequencing analysis of exosomal and non-exosomal miRNA in human urine. Kidney International. 2014;86(2):433-444
  88. 88. Huang X, Yuan T, Tschannen M, Sun Z, Jacob H, Du M, et al. Characterization of human plasma-derived exosomal RNAs by deep sequencing. BMC Genomics. 2013;14:319
  89. 89. Kosaka N, Iguchi H, Ochiya T. Circulating microRNA in body fluid: A new potential biomarker for cancer diagnosis and prognosis. Cancer Science. 2010;101(10):2087-2092
  90. 90. Thomas DJ, Rosenbloom KR, Clawson H, Hinrichs AS, Trumbower H, Raney BJ, et al. The ENCODE Project at UC Santa Cruz. Nucleic Acids Research. 2007;35(Database):D663-D667
  91. 91. Kapranov P, Willingham AT, Gingeras TR. Genome-wide transcription and the implications for genomic organization. Nature Reviews. Genetics. 2007;8(6):413-423
  92. 92. Morris KV, Mattick JS. The rise of regulatory RNA. Nature Reviews. Genetics. 2014;15(6):423-437
  93. 93. Shao J, Chen H, Yang D, Jiang M, Zhang H, Wu B, et al. Genome-wide identification and characterization of natural antisense transcripts by strand-specific RNA sequencing in Ganoderma lucidum. Scientific Reports. 2017;7(1):5711
  94. 94. Hung T, Chang HY. Long noncoding RNA in genome regulation: Prospects and mechanisms. RNA Biology. 2010;7(5):582-585
  95. 95. Hansji H, Leung EY, Baguley BC, Finlay GJ, Askarian-Amiri ME. Keeping abreast with long non-coding RNAs in mammary gland development and breast cancer. Frontiers in Genetics. 2014;5:379
  96. 96. Malek E, Jagannathan S, Driscoll JJ. Correlation of long non-coding RNA expression with metastasis, drug resistance and clinical outcome in cancer. Oncotarget. 2014;5(18):8027-8038
  97. 97. Necsulea A, Soumillon M, Warnefors M, Liechti A, Daish T, Zeller U, et al. The evolution of lncRNA repertoires and expression patterns in tetrapods. Nature. 2014;505(7485):635-640
  98. 98. Brunner AL, Beck AH, Edris B, Sweeney RT, Zhu SX, Li R, et al. Transcriptional profiling of long non-coding RNAs and novel transcribed regions across a diverse panel of archived human cancers. Genome Biology. 2012;13(8):R75
  99. 99. Rinn JL, Chang HY. Genome regulation by long noncoding RNAs. Annual Review of Biochemistry. 2012;81:145-166
  100. 100. Gutschner T, Diederichs S. The hallmarks of cancer: A long non-coding RNA point of view. RNA Biology. 2012;9(6):703-719
  101. 101. Wapinski O, Chang HY. Long noncoding RNAs and human disease. Trends in Cell Biology. 2011;21(6):354-361
  102. 102. Schmitt AM, Chang HY. Long noncoding RNAs in cancer pathways. Cancer Cell. 2016;29(4):452-463
  103. 103. Engreitz JM, Ollikainen N, Guttman M. Long non-coding RNAs: Spatial amplifiers that control nuclear structure and gene expression. Nature Reviews. Molecular Cell Biology. 2016;17(12):756-770
  104. 104. Geisler S, Coller J. RNA in unexpected places: Long non-coding RNA functions in diverse cellular contexts. Nature Reviews. Molecular Cell Biology. 2013;14(11):699-712
  105. 105. Holoch D, Moazed D. RNA-mediated epigenetic regulation of gene expression. Nature Reviews. Genetics. 2015;16(2):71-84
  106. 106. Villares GJ, Zigler M, Dobroff AS, Wang H, Song R, Melnikova VO, et al. Protease activated receptor-1 inhibits the Maspin tumor-suppressor gene to determine the melanoma metastatic phenotype. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(2):626-631
  107. 107. Rinn JL, Kertesz M, Wang JK, Squazzo SL, Xu X, Brugmann SA, et al. Functional demarcation of active and silent chromatin domains in human HOX loci by noncoding RNAs. Cell. 2007;129(7):1311-1323
  108. 108. Tsai MC, Manor O, Wan Y, Mosammaparast N, Wang JK, Lan F, et al. Long noncoding RNA as modular scaffold of histone modification complexes. Science. 2010;329(5992):689-693
  109. 109. Gupta RA, Shah N, Wang KC, Kim J, Horlings HM, Wong DJ, et al. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature. 2010;464(7291):1071-1076
  110. 110. Kino T, Hurt DE, Ichijo T, Nader N, Chrousos GP. Noncoding RNA gas5 is a growth arrest- and starvation-associated repressor of the glucocorticoid receptor. Science Signaling. 2010;3(107):ra8
  111. 111. Webster JC, Huber RM, Hanson RL, Collier PM, Haws TF, Mills JK, et al. Dexamethasone and tumor necrosis factor-alpha act together to induce the cellular inhibitor of apoptosis-2 gene and prevent apoptosis in a variety of cell types. Endocrinology. 2002;143(10):3866-3874
  112. 112. Adriaenssens E, Dumont L, Lottin S, Bolle D, Lepretre A, Delobelle A, et al. H19 overexpression in breast adenocarcinoma stromal cells is associated with tumor values and steroid receptor status but independent of p53 and Ki-67 expression. The American Journal of Pathology. 1998;153(5):1597-1607
  113. 113. Gabory A, Jammes H, Dandolo L. The H19 locus: Role of an imprinted non-coding RNA in growth and development. BioEssays. 2010;32(6):473-480
  114. 114. Berteaux N, Lottin S, Monte D, Pinte S, Quatannens B, Coll J, et al. H19 mRNA-like noncoding RNA promotes breast cancer cell proliferation through positive control by E2F1. The Journal of Biological Chemistry. 2005;280(33):29625-29636
  115. 115. Ji P, Diederichs S, Wang W, Böing S, Metzger R, Schneider PM, et al. MALAT-1, a novel noncoding RNA, and thymosin beta4 predict metastasis and survival in early-stage non-small cell lung cancer. Oncogene. 2003;22(39):8031-8041
  116. 116. Bernard D, Prasanth KV, Tripathi V, Colasse S, Nakamura T, Xuan Z, et al. A long nuclear-retained non-coding RNA regulates synaptogenesis by modulating gene expression. The EMBO Journal. 2010;29(18):3082-3093
  117. 117. Tripathi V, Ellis JD, Shen Z, Song DY, Pan Q, Watt AT, et al. The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Molecular Cell. 2010;39(6):925-938
  118. 118. Long JC, Caceres JF. The SR protein family of splicing factors: Master regulators of gene expression. The Biochemical Journal. 2009;417(1):15-27
  119. 119. Zhou Y, Zhong Y, Wang Y, Zhang X, Batista DL, Gejman R, et al. Activation of p53 by MEG3 non-coding RNA. The Journal of Biological Chemistry. 2007;282(34):24731-24742
  120. 120. Mondal T, Subhash S, Vaid R, Enroth S, Uday S, Reinius B, et al. MEG3 long noncoding RNA regulates the TGF-β pathway genes through formation of RNA-DNA triplex structures. Nature Communications. 2015;6:7743
  121. 121. Zhang X, Zhou Y, Mehta KR, Danila DC, Scolavino S, Johnson SR, et al. A pituitary-derived MEG3 isoform functions as a growth suppressor in tumor cells. The Journal of Clinical Endocrinology and Metabolism. 2003;88(11):5119-5126
  122. 122. Zhang X, Gejman R, Mahta A, Zhong Y, Rice KA, Zhou Y, et al. Maternally expressed gene 3, an imprinted noncoding RNA gene, is associated with meningioma pathogenesis and progression. Cancer Research. 2010;70(6):2350-2358
  123. 123. Braconi C, Kogure T, Valeri N, Huang N, Nuovo G, Costinean S, et al. microRNA-29 can regulate expression of the long non-coding RNA gene MEG3 in hepatocellular cancer. Oncogene. 2011;30(47):4750-4756
  124. 124. Chen CL, Tseng YW, JC W, Chen GY, Lin KC, Hwang SM, et al. Suppression of hepatocellular carcinoma by baculovirus-mediated expression of long non-coding RNA PTENP1 and MicroRNA regulation. Biomaterials. 2015;44:71-81
  125. 125. Tang J, Ning R, Zeng B, Li Y. Molecular evolution of PTEN Pseudogenes in mammals. PLoS One. 2016;11(12):e0167851
  126. 126. Chen TH, Yen AM, GH W, Chen LS, Chiu YH. Multiple detection modalities and disease natural history of breast cancer. Studies in Health Technology and Informatics. 2007;129(Pt 1):78-81
  127. 127. Welch JD, Baran-Gale J, Perou CM, Sethupathy P, Prins JF. Pseudogenes transcribed in breast invasive carcinoma show subtype-specific expression and ceRNA potential. BMC Genomics. 2015;16:113
  128. 128. Askarian-Amiri ME, Crawford J, French JD, Smart CE, Smith MA, Clark MB, et al. SNORD-host RNA Zfas1 is a regulator of mammary development and a potential marker for breast cancer. RNA. 2011;17(5):878-891
  129. 129. Hansji H, Leung EY, Baguley BC, Finlay GJ, Cameron-Smith D, Figueiredo VC, et al. ZFAS1: A long noncoding RNA associated with ribosomes in breast cancer cells. Biology Direct. 2016;11(1):62
  130. 130. Schoenberg DR, Maquat LE. Regulation of cytoplasmic mRNA decay. Nature Reviews. Genetics. 2012;13(4):246-259
  131. 131. Wahle E, Winkler GS. RNA decay machines: Deadenylation by the Ccr4-not and Pan2-Pan3 complexes. Biochimica et Biophysica Acta. 2013;1829(6-7):561-570
  132. 132. Rambout X, Detiffe C, Bruyr J, Mariavelle E, Cherkaoui M, Brohée S, et al. The transcription factor ERG recruits CCR4-NOT to control mRNA decay and mitotic progression. Nature Structural & Molecular Biology. 2016;23(7):663-672
  133. 133. Balatsos NA, Maragozidis P, Anastasakis D, Stathopoulos C. Modulation of poly(A)-specific ribonuclease (PARN): Current knowledge and perspectives. Current Medicinal Chemistry. 2012;19(28):4838-4849
  134. 134. Zhang X, Devany E, Murphy MR, Glazman G, Persaud M, Kleiman FE. PARN deadenylase is involved in miRNA-dependent degradation of TP53 mRNA in mammalian cells. Nucleic Acids Research. 2015;43(22):10925-10938
  135. 135. Boele J, Persson H, Shin JW, Ishizu Y, Newie IS, Søkilde R, et al. PAPD5-mediated 3′adenylation and subsequent degradation of miR-21 is disrupted in proliferative disease. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(31):11467-11472
  136. 136. Zhang LN, Yan YB. Depletion of poly(A)-specific ribonuclease (PARN) inhibits proliferation of human gastric cancer cells by blocking cell cycle progression. Biochimica et Biophysica Acta. 2015;1853(2):522-534
  137. 137. Maragozidis P, Karangeli M, Labrou M, Dimoulou G, Papaspyrou K, Salataj E, et al. Alterations of deadenylase expression in acute leukemias: Evidence for poly(a)-specific ribonuclease as a potential biomarker. Acta Haematologica. 2012;128(1):39-46
  138. 138. Winkler GS. The mammalian anti-proliferative BTG/Tob protein family. Journal of Cellular Physiology. 2010;222(1):66-72
  139. 139. Ezzeddine N, Chang TC, Zhu W, Yamashita A, Chen CY, Zhong Z, et al. Human TOB, an antiproliferative transcription factor, is a poly(A)-binding protein-dependent positive regulator of cytoplasmic mRNA deadenylation. Molecular and Cellular Biology. 2007;27(22):7791-7801
  140. 140. Stupfler B, Birck C, Séraphin B, Mauxion F. BTG2 bridges PABPC1 RNA-binding domains and CAF1 deadenylase to control cell proliferation. Nature Communications. 2016;7:10811
  141. 141. Ogami K, Hosoda N, Funakoshi Y, Hoshino S. Antiproliferative protein Tob directly regulates c-myc proto-oncogene expression through cytoplasmic polyadenylation element-binding protein CPEB. Oncogene. 2014;33(1):55-64
  142. 142. Boiko AD, Porteous S, Razorenova OV, Krivokrysenko VI, Williams BR, Gudkov AV. A systematic search for downstream mediators of tumor suppressor function of p53 reveals a major role of BTG2 in suppression of Ras-induced transformation. Genes & Development. 2006;20(2):236-252
  143. 143. Kawakubo H, Carey JL, Brachtel E, Gupta V, Green JE, Walden PD, et al. Expression of the NF-kappaB-responsive gene BTG2 is aberrantly regulated in breast cancer. Oncogene. 2004;23(50):8310-8319
  144. 144. Struckmann K, Schraml P, Simon R, Elmenhorst K, Mirlacher M, Kononen J, et al. Impaired expression of the cell cycle regulator BTG2 is common in clear cell renal cell carcinoma. Cancer Research. 2004;64(5):1632-1638
  145. 145. Kawakubo H, Brachtel E, Hayashida T, Yeo G, Kish J, Muzikansky A, et al. Loss of B-cell translocation gene-2 in estrogen receptor-positive breast carcinoma is associated with tumor grade and overexpression of cyclin d1 protein. Cancer Research. 2006;66(14):7075-7082
  146. 146. Möllerström E, Kovács A, Lövgren K, Nemes S, Delle U, Danielsson A, et al. Up-regulation of cell cycle arrest protein BTG2 correlates with increased overall survival in breast cancer, as detected by immunohistochemistry using tissue microarray. BMC Cancer. 2010;10:296
  147. 147. Faraji F, Hu Y, Yang HH, Lee MP, Winkler GS, Hafner M, et al. Post-transcriptional control of tumor cell autonomous metastatic potential by CCR4-NOT Deadenylase CNOT7. PLoS Genetics. 2016;12(1):e1005820
  148. 148. Helms MW, Kemming D, Contag CH, Pospisil H, Bartkowiak K, Wang A, et al. TOB1 is regulated by EGF-dependent HER2 and EGFR signaling, is highly phosphorylated, and indicates poor prognosis in node-negative breast cancer. Cancer Research. 2009;69(12):5049-5056
  149. 149. Zhu J, Ding H, Wang X, Lu Q. PABPC1 exerts carcinogenesis in gastric carcinoma by targeting miR-34c. International Journal of Clinical and Experimental Pathology. 2015;8(4):3794-3802
  150. 150. Brennan SE, Kuwano Y, Alkharouf N, Blackshear PJ, Gorospe M, Wilson GM. The mRNA-destabilizing protein tristetraprolin is suppressed in many cancers, altering tumorigenic phenotypes and patient prognosis. Cancer Research. 2009;69(12):5168-5176
  151. 151. Wang H, Ding N, Guo J, Xia J, Ruan Y. Dysregulation of TTP and HuR plays an important role in cancers. Tumour Biology. 2016;37(11):14451-14461
  152. 152. Yoon NA, Jo HG, Lee UH, Park JH, Yoon JE, Ryu J, et al. Tristetraprolin suppresses the EMT through the down-regulation of Twist1 and Snail1 in cancer cells. Oncotarget. 2016;7(8):8931-8943
  153. 153. Lykke-Andersen J, Wagner E. Recruitment and activation of mRNA decay enzymes by two ARE-mediated decay activation domains in the proteins TTP and BRF-1. Genes & Development. 2005;19(3):351-361
  154. 154. Filippova N, Yang X, Ananthan S, Sorochinsky A, Hackney JR, Gentry Z, et al. Hu antigen R (HuR) multimerization contributes to glioma disease progression. The Journal of Biological Chemistry. 2017
  155. 155. Gauchotte G, Hergalant S, Vigouroux C, Casse JM, Houlgatte R, Kaoma T, et al. Cytoplasmic overexpression of RNA-binding protein HuR is a marker of poor prognosis in meningioma, and HuR knockdown decreases meningioma cell growth and resistance to hypoxia. The Journal of Pathology. 2017;242(4):421-434
  156. 156. Lang M, Berry D, Passecker K, Mesteri I, Bhuju S, Ebner F, et al. HuR small-molecule inhibitor elicits differential effects in adenomatosis polyposis and colorectal carcinogenesis. Cancer Research. 2017;77(9):2424-2438
  157. 157. Xu X, Song C, Chen Z, Yu C, Wang Y, Tang Y, et al. Downregulation of HuR inhibits the progression of esophageal cancer through Interleukin-18. Cancer Research and Treatment. 2017
  158. 158. Tan S, Ding K, Chong QY, Zhao J, Liu Y, Shao Y, et al. Post-transcriptional regulation of ERBB2 by miR26a/b and HuR confers resistance to tamoxifen in estrogen receptor-positive breast cancer cells. The Journal of Biological Chemistry. 2017;292(33):13551-13564
  159. 159. Braun JE, Huntzinger E, Fauser M, Izaurralde E. GW182 proteins directly recruit cytoplasmic deadenylase complexes to miRNA targets. Molecular Cell. 2011;44(1):120-133
  160. 160. Ryu J, Yoon NA, Seong H, Jeong JY, Kang S, Park N, et al. Resveratrol induces glioma cell apoptosis through activation of tristetraprolin. Molecules and Cells. 2015;38(11):991-997
  161. 161. Lykke-Andersen S, Jensen TH. Nonsense-mediated mRNA decay: An intricate machinery that shapes transcriptomes. Nature Reviews. Molecular Cell Biology. 2015;16(11):665-677
  162. 162. Lejeune F, Li X, Maquat LE. Nonsense-mediated mRNA decay in mammalian cells involves decapping, deadenylating, and exonucleolytic activities. Molecular Cell. 2003;12(3):675-687
  163. 163. Cao L, Qi L, Zhang L, Song W, Yu Y, Xu C, et al. Human nonsense-mediated RNA decay regulates EMT by targeting the TGF-ß signaling pathway in lung adenocarcinoma. Cancer Letters. 2017;403:246-259
  164. 164. Liu C, Karam R, Zhou Y, Su F, Ji Y, Li G, et al. The UPF1 RNA surveillance gene is commonly mutated in pancreatic adenosquamous carcinoma. Nature Medicine. 2014;20(6):596-598
  165. 165. Gardner LB. Nonsense-mediated RNA decay regulation by cellular stress: Implications for tumorigenesis. Molecular Cancer Research. 2010;8(3):295-308
  166. 166. Martin L, Gardner LB. Stress-induced inhibition of nonsense-mediated RNA decay regulates intracellular cystine transport and intracellular glutathione through regulation of the cystine/glutamate exchanger SLC7A11. Oncogene. 2015;34(32):4211-4218
  167. 167. Matos ML, Lapyckyj L, Rosso M, Besso MJ, Mencucci MV, Briggiler CI, et al. Identification of a novel human E-cadherin splice variant and assessment of its effects upon EMT-related events. Journal of Cellular Physiology. 2017;232(6):1368-1386
  168. 168. Bezzi M, Teo SX, Muller J, Mok WC, Sahu SK, Vardy LA, et al. Regulation of constitutive and alternative splicing by PRMT5 reveals a role for Mdm4 pre-mRNA in sensing defects in the spliceosomal machinery. Genes & Development. 2013;27(17):1903-1916
  169. 169. Dewaele M, Tabaglio T, Willekens K, Bezzi M, Teo SX, Low DH, et al. Antisense oligonucleotide-mediated MDM4 exon 6 skipping impairs tumor growth. The Journal of Clinical Investigation. 2016;126(1):68-84
  170. 170. Voorhoeve PM, le Sage C, Schrier M, Gillis AJ, Stoop H, Nagel R, et al. A genetic screen implicates miRNA-372 and miRNA-373 as oncogenes in testicular germ cell tumors. Cell. 2006;124(6):1169-1181
  171. 171. Ota A, Tagawa H, Karnan S, Tsuzuki S, Karpas A, Kira S, et al. Identification and characterization of a novel gene, C13orf25, as a target for 13q31-q32 amplification in malignant lymphoma. Cancer Research. 2004;64(9):3087-3095
  172. 172. Ikeda S, Kitadate A, Abe F, Saitoh H, Michishita Y, Hatano Y, et al. Hypoxia-inducible microRNA-210 regulates the DIMT1-IRF4 oncogenic axis in multiple myeloma. Cancer Science. 2017;108(4):641-652
  173. 173. Li Z, Lei H, Luo M, Wang Y, Dong L, Ma Y, et al. DNA methylation downregulated mir-10b acts as a tumor suppressor in gastric cancer. Gastric Cancer. 2015;18(1):43-54
  174. 174. Wang CZ, Yuan P, Li Y. MiR-126 regulated breast cancer cell invasion by targeting ADAM9. International Journal of Clinical and Experimental Pathology. 2015;8(6):6547-6553
  175. 175. Heyn H, Engelmann M, Schreek S, Ahrens P, Lehmann U, Kreipe H, et al. MicroRNA miR-335 is crucial for the BRCA1 regulatory cascade in breast cancer development. International Journal of Cancer. 2011;129(12):2797-2806
  176. 176. Bhan A, Mandal SS. LncRNA HOTAIR: A master regulator of chromatin dynamics and cancer. Biochimica et Biophysica Acta. 2015;1856(1):151-164
  177. 177. Zhang J, Zhang P, Wang L, Piao HL, Ma L. Long non-coding RNA HOTAIR in carcinogenesis and metastasis. Acta Biochimica et Biophysica Sinica. 2014;46(1):1-5
  178. 178. Pickard MR, Williams GT. The hormone response element mimic sequence of GAS5 lncRNA is sufficient to induce apoptosis in breast cancer cells. Oncotarget. 2016;7(9):10104-10116
  179. 179. Mourtada-Maarabouni M, Pickard MR, Hedge VL, Farzaneh F, Williams GT. GAS5, a non-protein-coding RNA, controls apoptosis and is downregulated in breast cancer. Oncogene. 2009;28(2):195-208
  180. 180. Berteaux N, Aptel N, Cathala G, Genton C, Coll J, Daccache A, et al. A novel H19 antisense RNA overexpressed in breast cancer contributes to paternal IGF2 expression. Molecular and Cellular Biology. 2008;28(22):6731-6745
  181. 181. Jadaliha M, Zong X, Malakar P, Ray T, Singh DK, Freier SM, et al. Functional and prognostic significance of long non-coding RNA MALAT1 as a metastasis driver in ER negative lymph node negative breast cancer. Oncotarget. 2016
  182. 182. Gutschner T, Hammerle M, Diederichs S. MALAT1—A paradigm for long noncoding RNA function in cancer. Journal of Molecular Medicine. 2013;91(7):791-801
  183. 183. Sun L, Li Y, Yang B. Downregulated long non-coding RNA MEG3 in breast cancer regulates proliferation, migration and invasion by depending on p53’s transcriptional activity. Biochemical and Biophysical Research Communications. 2016
  184. 184. Zhou Y, Zhang X, Klibanski A. MEG3 noncoding RNA: A tumor suppressor. Journal of Molecular Endocrinology. 2012;48(3):R45-R53
  185. 185. Poliseno L, Marranci A, Pandolfi PP. Pseudogenes in human cancer. Frontiers in Medicine. 2015;2:68
  186. 186. Poliseno L, Salmena L, Zhang J, Carver B, Haveman WJ, Pandolfi PP. A coding-independent function of gene and pseudogene mRNAs regulates tumour biology. Nature. 2010;465(7301):1033-1038
  187. 187. Poliseno L, Haimovic A, Christos PJ, Vega YSMEC, Shapiro R, Pavlick A, et al. Deletion of PTENP1 pseudogene in human melanoma. The Journal of Investigative Dermatology. 2011;131(12):2497-2500
  188. 188. Liu F, Gao H, Li S, Ni X, Zhu Z. Long non-coding RNA ZFAS1 correlates with clinical progression and prognosis in cancer patients. Oncotarget. 2017;8(37):61561-61569

Written By

Carlos DeOcesano-Pereira, Fernando Janczur Velloso, Ana Claudia Oliveira Carreira, Carolina Simões Pires Ribeiro, Sheila Maria Brochado Winnischofer, Mari Cleide Sogayar and Marina Trombetta-Lima

Submitted: 19 July 2017 Reviewed: 23 October 2017 Published: 21 February 2018