Open access peer-reviewed chapter

Mechanisms of Oncogene Activation

Written By

Anca Botezatu, Iulia V. Iancu, Oana Popa, Adriana Plesa, Dana Manda, Irina Huica, Suzana Vladoiu, Gabriela Anton and Corin Badiu

Submitted: 03 December 2014 Reviewed: 22 July 2015 Published: 02 March 2016

DOI: 10.5772/61249

From the Edited Volume

New Aspects in Molecular and Cellular Mechanisms of Human Carcinogenesis

Edited by Dmitry Bulgin

Chapter metrics overview

3,927 Chapter Downloads

View Full Metrics

Abstract

The main modifications that characterize cancer are represented by alterations in oncogenes, tumor-suppressor genes, and non-coding RNA genes. Most of these alterations are somatic and the process is a multistep one. Tumors often arise from an initial transformed cell, and after subsequent genetic alterations different cytogenetically clones lead to tumor heterogeneity.

Keywords

  • Oncogenes
  • genomic instability
  • epigenetic modification

1. Introduction

The main modifications that characterize cancer are represented by alterations in oncogenes, tumor-suppressor genes, and non-coding RNA genes. Most of these alterations are somatic and the process is a multistep one, although germ-line mutations can predispose a person to heritable or familial cancer.

Tumors often arise from an initial transformed cell, and after subsequent genetic alterations different cytogenetically clones lead to tumor heterogeneity. Tumor heterogeneity determines different clinical phenotypes, leading to an individual response to treatment for tumors with the same diagnostic type.

Oncogenes encode proteins that control cell processes such as proliferation and apoptosis. Among these proteins are transcription factors, chromatin remodelers, growth factors, growth factor receptors, signal transducers, and apoptosis regulators. Activation of oncogenes by structural alterations (chromosomal rearrangement, gene fusion, mutation, and gene amplification) or epigenetic modification (gene promoter hypomethylation) confers an increased or a deregulated expression. Therefore, cells with such alterations possess a growth advantage or an increased survival rate. Translocations and mutations occur early on in tumor progression, whereas amplification usually occurs during late tumor stages.

A proto-oncogene is a normal gene that presents a potential to become an oncogene after a genetic alteration (mutation), leading to an increased expression. Usually, proto-oncogenes code for proteins that control cell growth and differentiation through signal transduction and execution of mitogenic signals. Upon activation, a proto-oncogene (or its product onco-protein) becomes a tumor-inducing agent. Most known examples of proto-oncogenes include RAS, WNT, MYC, ERK, and TRK. Another oncogene is the BCR-ABL gene found on the Philadelphia chromosome, a piece of genetic material seen in chronic myelogenous leukemia caused by the translocation of pieces from chromosomes 9 and 22 t (9; 22).

Oncogene products can comprise a variety of molecules such as transcription factors, chromatin remodelers, growth factors, growth factor receptors, signal transducers, and apoptosis regulators, each playing an important role in neoplastic transformation. For example, studies have shown that in prostate carcinomas the fusion between the TMPR552 gene and two transcription factors ERG1 or ETV1 creates a fusion protein that increases proliferation and inhibits apoptosis of cells in the prostate gland, thereby facilitating their transformation into cancer cells [1]. Another example is represented by chromatin remodeler factors such as the MLL gene that plays a critical role in acute lymphocytic leukemia and acute myelogenous leukemia [2]. Also, an essential role in cancer development is played by apoptosis regulators such as the BCL2 gene, which is involved in the initiation of almost all follicular lymphomas and some diffuse large B-cell lymphomas [3].

Advertisement

2. Mutations

Mutations in an oncogene may lead to a change in the structure of encoded protein, enhancing its transforming activity. Oncogenes are activated by point mutations (substitutions) and may either enhance or degrade the function of a protein. Table 1 shows the occurrences of mutations in each oncogene among some tissues [4].

CANCER TYPE ONCOGENES TYPE
T cell lymphoma KIT KRAS Co-occurred
Acute lymphoblastic leukemia KRAS NRAS Co-occurred
Acute myeloid leukemia KRAS NRAS Co-occurred
FLT3 KIT Mutually exclusive
Lung adenocarcinoma BRAF KRAS Mutually exclusive
KRAS NRAS Mutually exclusive
EGFR3 KRAS Mutually exclusive
Pancreatic cancer B-CATENIN KRAS Mutually exclusive
Biliary tract cancer BRAF KRAS Mutually exclusive
Colorectal cancer BRAF KRAS Mutually exclusive
KRAS PIK3CA Co-occurred
B-CATENIN KRAS Co-occurred
KRAS NRAS Mutually exclusive
Thyroid cancer BRAF NRAS Mutually exclusive
HRAS NRAS Mutually exclusive
BRAF RET Mutually exclusive
NRAS RET Mutually exclusive
Melanoma BRAF NRAS Mutually exclusive
BRAF HRAS Mutually exclusive
BRAF KRAS Mutually exclusive
KRAS BRAS Co-occurred
Prostate cancer HRAS NRAS Mutually exclusive
KRAS NRAS Mutually exclusive
Kidney cancer B-CATENIN WTI Co-occurred
Cervical cancer PIK3CA KRAS Mutually exclusive
PIK3CA BRAF Mutually exclusive

Table 1.

Frequently mutated oncogenes in various type of cancers

In cancer, mutations occur in many oncogenes, most notable being RAS and BRAF. The RAS family represents the upstream component of the RAS/RAF/MAPK pathway and mutations in RAS are one of the most common activating events in most of cancers. Mutated RAS oncogene (KRAS, HRAS, and NRAS) encodes for a protein that remains in the active state and transduces signals for continuous cell growth. KRAS mutations are common in carcinomas of the lung, colon, and pancreas [5], whereas mutations of NRAS occur in acute myelogenous leukemia and the myelodysplastic syndrome [6].

BRAF is a protein member of the RAF family (RAF1, BRAF, ARAF), also regulated by RAS binding. Mutated BRAF oncogene encodes for a protein with a modified kinase domain, resulting in a constitutively active protein that uncontrollably stimulates the MAP kinase cascade [7].

2.1. Melanoma

BRAF mutations are the most common somatic mutations in cutaneous melanoma and are extremely rare in mucosal melanoma. There are found in 48% of metastatic biopsy specimens and can precede neoplastic transformation [8, 9]. Over 90% of the identified mutations in BRAF are in codon 600. The most common is BRAFV600E, resulting in substitution of glutamic acid for valine (BRAFV600E: nucleotide 1799 T>A; codon GTG>GAG). The second most common mutation is BRAFV600K (5–6%) substituting lysine for valine, (GTG>AAG), followed by BRAFV600R (GTG>AGG), BRAFV600′E2′ (GTG>GAA). Less common BRAF mutations found in cutaneous melanoma are BRAF V600D (GTG>GAT) and L597R [1012].

There were identified mutations in hotspot codons (12, 13, and 61) of different RAS genes (HRAS, NRAS, or KRAS), but the most prevalent were HRAS substitutions that occurred preponderent at codon 61 (HRAS Q61L mutation), with fewer mutations at codon 12 and codon 13 [13]. Mutations in N-RAS appear to be significant in melanoma even earlier than the discovery of BRAF mutations [14]. The base change at position 61 seems to be important in the activation of N-RAS genes, transforming activity being detected only when mutant codon 61 was present. BRAFV600E mutations are more common in younger persons and in tumors arising from intermittently sun-exposed skin, exclusive with N-RAS [15].

C-KIT gene encodes a receptor tyrosine kinase (KIT). All the mutations were founded in exon 11, 13, and 17. The most common is V559A mutation that results in an amino acid substitution at position 559 in KIT, from a valine (V) to an alanine (A) [16]. While BRAF and NRAS mutations are common and significant in cutaneous melanomas, C-KIT mutations were detected in acral melanomas, mucosal melanomas, conjunctival melanomas, and cutaneous melanomas [17]

2.2. Colorectal cancer

The development of colorectal cancer (CRC) is a multistep process that occurs due to the accumulation of several genetic alterations, which are associated with oncogenes and tumor suppressor genes, as well as genes involved in DNA damage recognition and repair.

Most of the BRAF mutations associated with CRC are located in exons 11 and 15, coding for the kinase domain. The hotspot mutation at 1796 nucleotide is the T-to-A transversion that corresponds to the V600E mutation (7–15%) [18].

In colorectal cancer, RAS gene mutations have been reported in 40–50% and the frequency of KRAS mutations varies between 24–50%. KRAS mutation occurs most commonly in codon 12 and 13 rather than in codon 61, with the most frequent mutations: G12D, G12V, G12C, G13D, Q61H [19]. KRAS mutations exist in the presence of a vast majority of wild-type KRAS cells, which is why they not are detected in initial disease. Thirty-eight percent of patients whose tumors were initially KRAS wild-type developed KRAS mutations that were detectable in their sera after 5–6 months of treatment [20]. KRAS is mutated much more frequently than NRAS. KRAS mutations were studied to determine their role in the predictability of response to chemotherapy treatment.

2.3. Thyroid cancer

RAS mutations in thyroid cancer occur in both benign tumors and thyroid cancers (both papillary thyroid carcinoma (PTC) and follicular thyroid cancer (FTC)), with variable frequency in anaplastic thyroid cancers. PTCs with RAS mutations tend to display a lower rate of lymph node metastasis [21]. PTCs with RAS mutations often present a follicular architecture and a follicular variant of papillary thyroid carcinoma (FVPTC). There are two most common RAS mutations associated with thyroid cancer: mutations of H-RAS codon 61 and N-RAS codon 61 [22].

BRAF mutations were first detected missense mutations in thyroid cancer [23], which occurs in exon 15, due to the substitution of the amino acid valine for glutamic acid at residue position 600 [24]. This mutation is the most frequent genetic change in PTC [25], being found in 36–69% of PTC cases. BRAF mutation is responsible for the suppression of the sodium/iodide symporter (NIS), which is involved in iodine metabolism [26]. The V600E mutation comprises more than 90% of observed BRAF mutations, with the highest rate (77%) in the tall cell variant of papillary cancer, and the lowest percentage (12%) in the FVPTC. In PTC, BRAF mutation is more frequent in older patients, associated with extrathyroidal invasion [27]. By contrast, others have found that the BRAF mutation is not associated with age, gender, multicentricity, recurrence rate, lymphovascular invasion, or distant metastasis [28].

Activating point mutations of RET oncogenes are associated with hereditary cancer syndrome (multiple endocrine neoplasia type 2 -MEN 2). RET mutations are mostly missense and located in exons 10, 11 (extracellular domain of RET), 13, 14, 15, and 16 (in the TK domain) [29–31]. Mutation of the extracellular cysteine in codon 634 in exon 11 of RET causes ligand-independent dimerization of receptor molecules and enhances phosphorylation of intracellular substrates and cell transformation. Mutation of the intracellular TK (codon 918) results in cellular transformation [32].

There is a high correlation between the position of the point mutation and the phenotype of the disease. Three subtypes based on clinical presentation are defined: MEN 2A, MEN 2B, and FMTC. RET mutations are observed in 98% of MEN2A, 95% of MEN 2B, and 88% of familial medullary thyroid carcinoma (FMTC) [33]. Activating mutations of RET involving exons 10, 11, 13, 14, and 15 (encoding the highly conserved cysteine-rich domain) have been proven to cause MEN2A. [34]. The mutations for MEN2A are mostly located in exon 10 (10–15%), including codons 609, 611, 618, and 620, and exon 11 (80–85%), as well as codons 630 and 634 [35]. The mutations characteristic of FMTC occur in exons 10 and 11. However, non-cysteine point mutations also have been found in exon 8 (codons 532 and 533), exon 13 (codons 768, 790, and 791), exon 14 (codons 804 and 844), exon 15 (codon 891), and exon 16 (codon 912) [3537]. About 95% of MEN2B patients carry a M918T mutation within exon 16 and 5% have an A883F mutation in exon 15. Mutation in codon 918 gives a more aggressive phenotype [38].

2.4. Hepatocellular Carcinomas (HCC)

In HCC, only one mutation (KRAS codon 13; Gly to Asp) was detected among patients and no mutations were found in codons 12 and 61 of KRAS or codons 12, 13, and 61 of the NRAS and HRAS genes. So, the activation of RAS oncogenes by point mutations does not play a major role in hepatocellular carcinogenesis [39]. Activating mutations in the BRAF oncogene have been found in a small fraction of hepatocellular carcinomas. KRAS and BRAF mutations are rare events in HCC and therefore not a key event in hepatocarcinogenesis [40].

2.5. Pancreatic cancer

The highest incidence of KRAS mutations are found in adenocarcinomas of the pancreas (90%), with activating point mutations in codon 12 of the KRAS protein, leading to a glycine (G) to aspartic acid (D) or valine (V) substitution [41]. Single amino acid substitutions at G12, 13, or Q61 lead to the formation of mutated KRAS that are insensitive to GAP stimulation. This leads to the accumulation of persistently GTP-bound and active KRAS, which leads to pancreatic cancer formation [42].

2.6. Cervical cancer

Cervical cancer harbors high rates of potentially targetable oncogenic mutations. KRAS mutations were identified in low percentage (17%) exclusively in cervical adenocarcinomas. Most mutations were missense mutations of codon G12, well-described activating mutations, which have been associated with a worse prognosis in the metastatic process [43].

EGFR mutations were identified in 7.5% of cervical squamous cell carcinomas; a missense mutation in exon 15 of the EGFR gene produces an alternate spliced transcript (isoform D). Its presence in both tumor and adjacent normal tissue suggests that EGFR S703F may be a germline mutation [44, 45].

PIK3CA mutations are present in both squamous cell carcinomas and adenocarcinomas (31%). The PIK3CA mutations were located in the exon 9 helical domain in two hotspot mutations (E545K and E542K), which result in the constitutive activation of cellular signaling [46]. PI3KCA mutations may impart a more aggressive and treatment-resistant phenotype and decreased survival among patients with these mutations in early stage cancers [47].

Advertisement

3. Gene amplification and chromosomal translocations

The interest regarding the role of genomic context in promoting amplification was intensely investigated, but is still under debate. An important interest remains to establish the tendency of some genomic region to be subject to amplification. Past researches showed that different regions of the genome were more subjected to be amplified than others, but the molecular substrate was unknown [48]. At present, several mechanisms and models have been proposed to explain gene amplification in oncogenesis.

3.1. Gene amplification

Besides point mutations resulting in amino acid substitutions, a proto-oncogene may be activated by chromosomal alterations. Among the most important cromosomal abnormalities is gene amplification, which is an increase of the copy number for a specific chromosomal region. The consequence of chromosome fragment amplification is associated with overexpression of the amplified gene(s) and is a characteristic of cancer [49]. Amplified genes (hundreds of copies of normally diploid genes) may be organized as extrachromosomal elements (double minute chromosomes) as repeated units at a single locus or scattered throughout the genome.

At this moment, the relationships between the two forms of gene amplification found in tumors, the intrachromosomal homogeneously staining regions (HSRs) and the extrachromosomal DNA molecules, double minutes (dmins), are not well understood [50].

Several models for initiation of amplification have been described involving defects in DNA replication or telomere dysfunction and chromosomal fragile sites. Regarding the DNA replication initial proposals, based on extra rounds of replication due to replication origins misfiring appear to be incorrect modification of models invoking replication of extrachromosomal DNA [51]. Another theory involves the double-strand DNA breaks (frequent in replicating cells) generated by the collapse of replication forks that are unable to progress due to DNA structure lesions, therefore providing an opportunity to initiate the amplification process [49].

Telomeres are repetitive nucleotide sequences, with the role to prevent the loss of DNA sequences, resulted as a consequence of the incomplete DNA replication at the chromosome ends. Telomere shortening can block cell division; this mechanism appears to prevent genomic instability and development of cancer in aged cells by limiting the number of cell divisions [52]. Telomerase is responsible for telomere replication and is inactive in most somatic cells. With every cell division, the DNA telomere sequence is shortened by 40–50 bp. Telomere shortening in humans can induce replicative senescence, which blocks cell division. When telomeres are short to a critical length (replicative limit), cellular senescence is induced and normal cells cease to proliferate. This mechanism appears to prevent genomic instability and development of cancer in aged cells by limiting the number of cell divisions. In cancer, tumor cells escape replicative limit and acquire the capability to maintain telomere length through cell divisions by telomerase reactivation, or by using a recombination-based mechanism and alternate lengthening of telomeres (ALT) [53, 54]. An experimental murine model (lacking the RNA component of telomerase-TercK/K mice) for telomere dysfunction demonstrated the promotion of gene amplification. Tumor genomes arising in mutant mice contain chromosomal rearrangements, amplifications, and deletions commonly associated with human tumors [55, 56]. Despite the established correlation between telomerase reactivation and telomeres lengthening in cancer, recent literature review and analysis [52] suggest this is unlikely, because shorter telomeres and telomerase inactivation is more often associated with increased cancer rates, and the mortality from cancer occurs late in life.

In humans, shorter telomeres were associated with poorer health and aging and were also observed in preneoplastic stages, supporting a role for this mechanism in generating genomic aberrations in oncogenesis [5760]. The model for gene amplification due to telomere abnormalities and the break at fragile sites (discussed below) was first described in maize and results from the breakage/fusion/bridge (B/F/B) cycles [61]. B/F/B cycles are initiated when broken ends of chromosomes fuse, resulting in a dicentric chromosome. During anaphase, the two centromeres are pulled in opposite directions and the dicentric chromosome generates a chromosome with an inverted duplication of terminal sequences to break. The B/F/B cycle continues in the next cell cycle because this chromosome also has broken ends. The B/F/B cycles were observed like primary mechanism for gene amplification in hamster cells [62].

In human cancer, evidence of B/F/B cycles was provided by the high frequency of anaphase bridges in early passage tumor cells and tumors [63, 64]. On the other hand, it was proven that human tumor cells in culture presenting gene amplification contain DM chromosomes, and the clones with low-copy amplification contained structures related to B/F/B cycles [65, 66]. There are evidences that B/F/B cycles may generate amplicons. These results were obtained by cytogenetic analyses of HSRs in tumor cell lines and in model systems with amplifications following drug treatments [67, 68]. The model explains that loss of the DNA sequences distal to the gene under selection or their translocation to another chromosome is also possible.

HSR may arise from the integration or fusion of double minute with a chromosome [50]. Currently, the data available suggests that fusion and reintegration constitute a pathway for the evolution of extrachromosomal elements, but the site of HSR insertion has never been characterized at a nucleotide resolution [50].

3.2. Fragile sites

Fragile sites are part of normal chromosome structures existing in each individual and represent chromosome regions that are late in replicating and prone to breakage under conditions of replication stress. Fragile sites occur after partial inhibition of DNA synthesis and are constituted in regions presenting site-specific gaps and breaks on metaphase chromosomes. Common fragile sites are normally stable in somatic cells, but it was observed that following treatment of cultured cells with replication inhibitors, fragile sites display gaps, breaks, rearrangements [69, 70]. Fragile sites extend over large regions of high DNA flexibility and are associated with genes.

The molecular nature and mechanisms involved in fragile site instability was unknown till recently. In many cancer cells, fragile sites and associated genes suffer frequent deletions and/or rearrangement, demonstrating their role in genome instability during the oncogenesis process. As a group, fragile sites are heterogeneous and seem to extend over broad regions 0.3–9-Mb long. The regions comprising fragile site are particularly associated with a high frequency of recombinogenic events, including co-localization with chromosome aberrations sites related to various cancers [69].

Accordingly to several studies, there are around 127 known fragile sites in the human genome, defined as "common" or "rare" based on their frequency [71, 72]. Common fragile sites (CFSs) are a normal part of the human genome and are typically replicative stable [73]. CFSs are not the result of nucleotide repeat expansion mutations. The majority of breakages at CFSs are further distinguished depending on their sensitivity to the drugs used to induce their expression (e.g., low doses of the antibiotic aphidocilin (APH)) [74].

The breakage effect of APH may be reduced by using a co-treatment with low concentrations of the topoisomerase I inhibitor and camptothecin (CPT) [75]. CFS regions are highly conserved in vertebrate species, including mouse and primates [7678]. CFSs initiate proper replication but slow to complete it, introducing breaks from unreplicated regions of DNA [79]. The mechanism proposed for CFS instability resides in this late replication. Late replication may occur due to formation of non-B DNA structures such as hairpins and toroids that block the replication fork in AT rich regions [80].

Rare fragile sites (RFSs) are classified into two sub-groups based on the compounds that induce breakage, folate-sensitive groups and nonfolate-sensitive groups, which are sensitive at bromodeoxyuridine (BrdU) or distamycin A, an antibiotic that binds to AT-pairs of the DNA sequence. The folate-sensitive group is characterized by an expansion of CGG repeats, while the nonfolate-sensitive group contains many AT-rich minisatellite repeats [8183]. The genome instability mechanism of CGG and AT-rich repeats characteristic for RFSs can form DNA structure (hairpins and other non-B DNA) replication forks, leading to breakage [84, 85]. On the other hand, it was demonstrated that DNA polymerase stops at CTG and CGG triplet repeat sequences, which can result in continuous DNA synthesis via slippage [79].

Fragile site regions are stable in normal cells and become unstable in tumor cells. The breakage of the fragile sites may be caused by mutations leading to a blockage of replication, or by a cell cycle perturbation and gene involved in the DNA repair process deregulation [86]. Several reports developed the concepts that underlie the mechanisms leading to fragile site expression and chromosomal rearrangements at fragile sites in tumors. The analysis of DNA damage response in various tumor types, including bladder, breast, colorectal, and lung tumors, found that early stages of cancer development are associated with an active DNA damage response, including phosphorylated ATR (ataxia telangiectasia and Rad3-related protein), ATM (ataxia telangiectasia mutated), CHK1 (checkpoint kinases), CHK2 kinases, phosphorylated histone H2AX, and p53 [87-88].

These events are linked to a high frequency of LOH (loss of heterozygoty) at known fragile site regions. The explained mechanisms sustained that in precancerous lesions, the blockage or collapsed replication leads to ATR activation and with subsequent DNA double strand breaks. Tumor cells that escape apoptosis or cell cycle arrest will exhibit allelic imbalances, especially at target fragile sites because of replication sensitivity. Further, the model sustains the necessity of p53 mutation and/or other genes involved in checkpoints control, leading therefore to cancer progression. Lesions at common fragile sites are indicators of replication stress during early stages of tumorigenesis [70].

Fragile sites regions are targets for the initiation of the amplification process due to breakage. Several studies showed that boundaries of some amplicons generated through the amplification process mapped to common chromosomal fragile sites in hamster cells [8990]. Evidences of the role of fragile sites in human cancer regarding gene amplification are scarce. One example of cell line model is for the MET amplicons in the esophageal adenocarcinoma map within the fragile site FRA7G [91].

Aphidicolin-sensitive fragile sites FRA5D, FRA5F, and FRA5C, which map distal to dihydrofolate reductase gene (DHFR) on 5q, are infrequently expressed and are less likely to contribute to the amplification process. In order for the gene amplification process to take effect the target gene must be in the close proximity of fragile sites, similar to the MET amplicon. The breakage at specific genomic sites may not contribute to the amplification process, and no evidence of recurrent amplicon boundaries was found using array CGH in a human cell culture system [92].

3.3. Amplified genes in cancer

The amplification process is important for deciphering oncogenesis molecular biology, prognosis, and targeted therapies. A good example of gene amplification is dihydrofolate reductase gene (DHFR), which usually occurs during progression of methotrexate-resistant acute lymphoblastic leukemia [93]. In cancer, the most amplified genes are members of four different oncogene families: MYC, cyclin D1 (or CCND1), EGFR, and RAS. The amplified DNA segment usually involves several hundred kilobases and can contain many genes.

In breast cancer, MYC, ERBB2, CCND1, EGFR, or MDM2 were found to be amplified concomitantly [94]. Moreover, it has been reported that there is a direct correlation between the number of amplifications and an advanced breast cancer and poor survival [95]. MYC oncogene is amplified in many types of cancer such as small-cell lung cancer, breast cancer, esophageal cancer, cervical cancer, ovarian cancer, and head and neck cancer [96].

Among the best-known oncogenes that are amplified in cancer cells is N-MYC. This gene codes for a transcription factor that plays a physiologic role in stimulating cellular proliferation and is commonly amplified in neuroblastoma where patients have poor clinical prognosis. Amplification of N-MYC in neuroblastoma has a valuable prognostic significance, and is correlated with an advanced tumor stage [97], along with MYC and ERBB2 in breast cancer [94].

In malignant thyroid tumors, C-MYC gene overexpression and amplification has also been correlated with tumor aggressiveness [98]. Overexpression of cyclins is also an important element in thyroid oncogenesis, playing a crucial role in PTC pathogenesis [98]. CCND1, a cell cycle key regulator of G1/S transition, is a frequent target of mutagenesis in many tumors; amplification and rearrangement of its gene can lead to the over-production of this cell cycle regulatory protein. CCND1 amplification also occurs in breast, esophageal, hepatocellular, and head and neck cancer [99].

EGFR (ERBB1) is amplified in glioblastoma and head and neck cancer. Amplification of ERBB2 (also called HER2/neu) in breast cancer correlates with a poor prognosis. A monoclonal antibody against the product of this oncogene (trastuzumab) is effective in breast cancers that overexpress HER2/neu.

New data were acquired using array-CGH technique, bringing to knowledge the complex aspect of oncogene amplification in cancer. Using array-CGH analysis on identification of an 8p12 amplicon in anaplastic thyroid carcinoma (ATC) cell lines, six genes were found to be amplified, DUSP26,MET, MYC, PVT1, YAP1 and CIAP1 [100]. DUSP26 effectively dephosphorylates p38 and formed a physical complex with p38, promoting the survival of ATC cells by inhibiting p38-mediated apoptosis.

The AIB1 oncogene is located on chromosome 20q, a region frequently amplified and overexpressed in breast cancer [101102]. High levels of AIB1 mRNA or protein predict significantly worse prognosis and overall survival in breast cancer patients [103]. AIB1 is a transcriptional co-activator that promotes the transcriptional activity of multiple nuclear receptors such as the estrogen and progesterone receptors [104].

In cervical cancer, the array-CGH technique revealed that the 3q26.3 amplification was the most consistent chromosomal aberration in primary tissues of cervical carcinoma, and an increased copy number of PIK3CA gene was identified [105]. PIK3CA is known to be involved in the PI 3-kinase/AKT signaling pathway, which plays an important role in regulating cell growth and apoptosis.

In pancreatic cancer, chromosome 19q13 was found amplified containing PAK4 gene [106]. PAK proteins are critical effectors that link Rho GTPases to cytoskeleton reorganization and nuclear signaling. PAK4 interacts specifically with the GTP-bound form of Cdc42Hs and weakly activates the JNK family of MAP kinases. PAK4 gene is not in a mutated oncogenic form but the activation of the PAK4 gene promotes KRAS2 gene mutation, a very frequent event in pancreatic cancer [106].

DNA amplification represents an important mechanism during human multistep hepatocarcinogenesis. Several genes were found to be amplified within 1q21 amplicon in hepatocellular carcinoma: CREB3L4 (cyclic AMP responsive element binding protein3-like 4); JTB (Jumping Translocation Breakpoint) is a transmembrane protein that suffers an unbalanced translocation in various types of cancers [107]; INTS3 and SNAPAP, whose role in oncogenesis remains to be defined; SHC1 is involved in signal transduction from receptor tyrosine kinases to downstream signal to RAS [108109]; CKS1B (CDC28 protein kinase regulatory subunit 1B), Cks1 expression was closely associated with poor differentiation and also negatively associated with p27kip1 in hepatocellular carcinoma [110]; CHD1L (Chromodomain Helicase/ATPase DNA Binding Protein 1-Like, also known as Amplified in Liver Cancer 1, ALC1) whose increased expression was associated with clinicopathological features such as microsatellite tumor formation, venous infiltration, and advanced tumor stage, overall survival time, and the disease-free survival rate [111]. Moreover, Glyoxalase 1 (Glo1) gene aberrations are associated with tumorigenesis and progression in numerous cancers. Hepatocarcinoma cells with genetic amplified Glo1 gene express higher levels of Glo1 and are more sensitive to cell killing effects if Glo1 expression is down-regulated [112]. The study supports the potential of Glo1 as therapeutic target in patients with hepatocellular carcinoma and genetic Glo1 amplification.

Human oncogene JUN gene amplification/overexpression was found in highly aggressive sarcomas and in hepatocellular carcinomas, along with amplification/overexpression of MAP3K5. JUN overexpression could interfere with adipocytic differentiation and promote angiogenesis [113, 114].

Amplification of the FGFR2 gene was identified in a subset of Chinese and Caucasian patients with gastric cancer. Fibroblast growth factor receptor family members (FGFR1–4) belong to the RTK superfamily. Through interaction with FGF ligands, the receptors are involved in diverse cellular functions including regulation of development processes, mediation of cell proliferation, and differentiation, as well as angiogenesis and tissue regeneration [115, 116]. FGF ligand binding leads to kinase activation and downstream signaling to phosphoinositide 3-kinase (PI3K)-AKT and mitogen-activated protein kinase–extracellular signal–regulated kinase (MAPK-ERK) pathways [117]. Genetic modifications or overexpression of FGFRs have been associated with tumorigenesis and progression in breast, prostate, stomach, and hematologic malignancies [118, 119]. FGFR2 amplification leads to constitutive activation of the FGFR2 signaling pathway in gastric cancer, and furthermore inhibition of this pathway using a well-tolerated, potent, and selective inhibitor can lead to rapid and durable tumor regressions in FGFR2-amplified gastric cancer xenograft models, representing an important treatment target [120].

3.4. Chromosomal translocations

Chromosomal translocations (CTs) are very common in human cancer, and the molecular mechanisms involved are complex and poorly understood. CTs are involved in several types of cancer, particularly in hematopoietic and lymphoid tumors [121]. This type of chromosomal abnormality seems to provide a selective growth advantage for some stem or progenitor cells, which may further initiate the development of some malignant tumors. In case of oncogenes, CTs may change the original locations of proto-oncogenes, generating effects on the gene products through two major ways [122, 123]. One is to generate oncogenic fusion proteins and the other way is that proto-oncogenes are brought into proximity with regulatory elements, causing the overexpression of proto-oncogene.

The first specific chromosomal translocation identified in human cancer was the Philadelphia chromosome [t(9;22)], which underlies chronic myeloid leukemia (CML). The fusion of chromosomes 9 and 22 leads to the joining of two unrelated genes, the C-ABL gene, which encodes a tyrosine kinase and is located on chromosome 9, and the gene BCR (for breakpoint recombination) located on chromosome 22.10 [124]. A chimeric protein (BCR-ABL) with novel transforming properties is formed from this specific chromosomal rearrangement. BCR-ABL oncoprotein has an abnormal tyrosine kinase activity and is associated with the tumorigenesis of CML and acute lymphoblastic leukemia (ALL) [124]. Duplication of the Philadelphia chromosome leads to accelerated CML blast phase, suggesting that increased copies of this aberrant gene confer a dose-dependent transforming effect [125]. Similar to t(9;22) in acute promyelocytic leukemia (APL), a chromosomal rearrangement joins a novel gene t(15;17), resulting in the formation of promyelocytic leukemia-retinoic acid receptor α (PML-RARα) fusion oncoprotein [126].

PML-RAR function is unknown, but this translocation underlies the response of this leukemia type to treatment with trans-retinoic acid. Another intergenic, CT t(12;21), leads to a novel chromosomal translocation product, TEL-AML1, which requires a specific treatment for pediatric acute lymphoblastic leukemia [127].

A classic example is the overexpression of proto-oncogene c-MYC in Burkitt lymphoma due to t(8;14) that results in c-MYC gene juxtaposed to immunoglobulin heavy chain (IGH) regulatory elements [128, 129]. Further expression of the gene is directed by the strong immunoglobulin heavy-chain enhancer, which is constitutively active in B lymphocytes. Thus, c-MYC overexpression is a potent force driving cellular proliferation.

The t(11;14) translocation juxtaposes CCND1 and immunoglobulin enhancer elements and is characteristic of mantle-cell lymphoma. The t(11;14) translocation juxtaposes CCND1 and immunoglobulin enhancer elements [130].

The ability to grow leukemic cells in culture long enough to allow cytogenetic analysis has facilitated the characterization of chromosomal translocations in leukemia. However, specific chromosomal translocations have also been observed in solid tumors. Aside from interchromosomal translocations, intrachromosomal translocations are also associated with cancer. Around 60–70% of PTCs have a characteristic inv(10)(q11.2q21). The breakpoint is represented by RET gene locus (10q11.2), which is relegated to the opposite breakpoint of the H4 (D10S170) or NCOA4 (ELE1) gene (10q21) in the same chromosome [131]. The H4 protein is widely expressed in the nucleus and cytoplasm and its function is unknown [132]. In PTC, many types of rearrangement loci (11 rearranged forms) were noted and PTC1(H4, CCDC6)-RET and PTC3(NCOA4)-RET are the most common [133]. PTC2-RET is a less common type of PTC-RET [134]. These rearrangements can lead to constitutively ligand-independent RET activity involved in thyroid carcinogenesis. The hypothesis sustain that the distances between RET and H4 loci are 18 Mb, therefore chromosome folding may close the two loci to each other in thyroid cells, increasing the probability of recombination between them in the interphase nuclei. This chromosomal folding is specific for thyroid cells, and this may explain why inv(10) (q11.2q21) is frequently seen in PTC [135].

It has been shown that in prostate carcinomas, the fusion between TMPR552 gene and two transcription factors ERG1 or ETV1 creates a fusion protein that increases proliferation and inhibits apoptosis of cells in the prostate gland, thereby facilitating their transformation into cancer cells [1].

The translocations of ETS are often found in human cancer, such as Ewing sarcoma [136137], leukemia [138139], prostate cancer [140], and breast cancer [141]. These once disparate tumors are now defined by a chromosomal translocation fusing the EWS gene to a number of transcription factors of the ETS gene family (the most common chimeric protein is EWS-FLI1) [142]. This chimeric product presumably acts directly on target promoters to direct the expression of genes that induce cellular proliferation. Identification of EWS translocations allowed the molecular grouping of a class of tumors whose proliferation is driven by similar genetic alterations and that respond to similar chemotherapeutic regimens.

Advertisement

4. Oncogene hypomethylation

The first epigenetic modification observed in human cancer was the loss of DNA methylation at the 5’citosine level (m5C residues replaced by unmethylated C residues), reported in 1983 [143]. This discovery was often regarded as an unwelcome complication, and all of the attention was focused on the opposite effect hypermethylation of promoters of genes that are silenced in cancers (e.g., tumor-suppressor genes). Global hypomethylation of DNA in cancer was found associated especially with repeated DNA elements; this modification did not represent a research direction for many years [144]. However, changes in the pattern of DNA methylation have been a consistent modification in cancer cells. Both hypo- and hypermethylation were observed at various loci, but at this moment it is clear that DNA methylation plays an important role in carcinogenesis.

New deep sequencing methylome analyses have shown much more cancer-linked hypomethylation of unique gene sequences and hypermethylation of repeated sequences than previously found [145148]. Targeting DNA repetitive sequence, DNA hypomethylation may induce genomic instability and mutation events in cancer genomes [149–152] by altering the intranuclear positioning of chromatin enhancing recombination [153155] and activating retroviral elements [156]. Promoter hypomethylation of some genes may be associated with the development of cancer by regulating the activity of genes [157].

4.1. Genomic hypomethylation profiles in cancer

DNA methylation principally occurs at 5’ cytosine from dinucleotide CpG sites [158, 159]. CpG dinucleotides are found in C+G-rich regions in the genome termed CpG islands, localized frequently at promoter or gene regulatory level. However, the vast majority of CpG dinucleotides are localized within the intergenic and intronic regions of the DNA, particularly within repeat sequences and transposable elements. Unmethylated CpG islands at gene level are associated with gene transcription. In normal somatic cells, between 70% and 90% of CpG dinucleotides are methylated, which constitute approximately 0.75–1% of the total number of bases in the genome, while most CpG islands are unmethylated [160]. A part of genes promoter region are methylated as part of normal developmental processes or tissue specific (e.g., germ-line specific genes-MAGE genes) [161]. In X chromosomes in female dosage compensation (imprinted genes of X chromosomes in females), where only one of two copies is active, methylation of regulatory regions is involved in the repression of the expression of the silent loci [162].

Recently, high-resolution genome-wide analyses of DNA methylation changed the idea that considers oncogenesis being characterized predominantly of hypomethylated DNA repeats and hypermethylated gene regions [163164]. The hallmark for cancer is represented by global losses of DNA methylation with local hypermethylation and hypomethylation of specific genes [165167].

Evaluation of the majority of cancers showed that a major contributor to global DNA hypomethylation is hypomethylation of tandem and interspersed DNA repeats [165, 168]. Several studies using CpG methylation-sensitive restriction endonucleases or sodium bisulfite reported that hypomethylation was often found at gene sequence level (including metastasis-associated genes) [157, 165].

4.2. Hypomethylation of DNA repeats sequence in cancer

Repeat sequences are represented by transposable elements found interspersed throughout the genome, or simple repeat sequences, such as DNA satellites, found in pericentromeric or subtelomeric region of chromosomes. These are normally methylated within the healthy genome [169].

In cancer, hypomethylation of DNA repeats is a result of the demethylation process rather than the preexisting hypomethylation in a cancer stem cell [170]. The frequency of cancer-associated hypomethylation of DNA repeats is dependent with disease progression (tumor grade, stage) [171, 172]. Hypomethylation is also seen in tumor adjacent tissues and in benign tumors (breast fibroadenomas and ovarian cystadenomas), but at a lower level than cancer [145, 165, 173, 174].

Hypomethylation may affect transcription and hypomethylation of interspersed DNA repeats within promoter modifies the chromatin boundaries resulting in transcription activation of nearby genes [175, 176]. Along with the effects upon transcription, hypomethylation can affect alternative splicing and hypomethylation of a minor portion of interspersed DNA repeats may occasionally cause induction of retroviral element transcription [156]. Several studies reported numerous evidences for the causal relationships between DNA hypomethylation and increased transcription as well as hypomethylation and cancer [177179].

Regions of cancer-associated changes in DNA methylation are found in short interspersed or clustered regions, as well as in long blocks [180182]. Dante et al. described hypomethylation of LINE-1 (a highly repeated interspersed repeat) in mononuclear cells from patients with chronic lymphocytic leukemia [183]. Along with hypomethylation of LINE-1, Alu repeats were also subsequently observed hypomethylated in many other types of cancers [183186]. In breast adenocarcinomas, ovarian epithelial cancers, and Wilms tumors, a hypomethylation of centromeric and juxtacentromeric satellite DNA was noted [173, 174, 187]. Moreover, another classes of tandem repeats (macrosatellite DNAs) and segmental duplications were found hypomethylated in various cancers [188190]. The loss of DNA methylation in cancer varies according to the tumor type and subclasses of DNA repeat [191193].

Gathering the result of the presented studies, we may conclude that in many types of cancer, hypomethylation of DNA repeats represents a highly informative prognostic marker and/or predictor of survival [194197].

4.3. Hypomethylation of DNA gene enhancer sequence in cancer

Gene expression levels may be further modulated by DNA methylation levels at upstream enhancer sites [198], which can affect the binding of transcription factors at (CpG) islands [199]. In normal cells, DNA demethylation at enhancer’s level is correlated with upregulation of expression of the associated gene. It was shown that the binding of FoxA1/FOXA1 transcriptional factors to enhancers is inhibited by DNA methylation at the respective binding site [200]. In this case, modification of DNA methylation status (demethylation) at the enhancer level may lead to an open chromatin state allowing the access of transcription factors at the active enhancer [201202]. Following DNA demethylation, FoxD3 transcription factor binds at the enhancer level, allowing the recruitment of FoxA1 and conversion of the enhancer to a state that is set for activity. Local DNA demethylation leads also to changes in histone H3K27 or H3K9 methylation [200]. FOXA1 is an important factor for oncogenesis being involved in various types of cancer [203]. Thus, DNA hypomethylation from transcription regulatory regions may cause changes in expression [204].

4.4. Genomic hypomethylation in promoters and within gene bodies

Hypomethylation of transcription regulatory regions is less frequent than hypermethylation of CpG island promoters in cancer. Some of the gene regions (including transcription control sequences) were associated with loss of DNA methylation. Currently, there are data that sustain that promoter hypomethylation of some genes may be associated with the development of cancer, regulating the activity of genes [157]. For example, promoter hypomethylation of specific immunity-related genes (e.g., cytokine IL-10) may activate the specific gene expression to inhibit the immune response in breast cancer [205], and the promoter hypomethylation of SPAN-Xb, an immunogenic antigen, can induce de novo B-cell response in myeloma cells [206]. However, the biological significance of promoter hypomethylation in cancer is still poorly understood [144]. Hypomethylation of gene promoters must cooperate with other key activators such as transcriptional factors to control gene expression [207, 208].

Promoters may overlap tissue-specific (T-DMR) or cancer-specific (C-DMR) differentially methylated DNA regions [209]. Most of the non-imprinted, autosomal T-DMR promoters are not the main type of vertebrate DNA promoters, and the genes presenting T-DMR promoters become activated after experimentally induced demethylation 5-deoxyazacytidine [209].

Intragenic epigenetic marks have been also involved in normal gene expression regulation and inverse relationships between imprinted gene expression and DNA methylation level was observed [210]. T-DMR regions were found not only inside many genes, but also in downstream promoters, flanking certain subsets of genes [211, 212]. Moreover, besides first exon, T-DMRs are also present at exonic and intronic sequences, insulators, intragenic ncRNA genes, and 3 ‘terminal regions [213, 214].

The role of these regions is to connect DNA and chromatin, inducing tissue-specific chromatin epigenetic marks inside genes [215, 216]. This relationship between DNA and chromatin modification at gene level may help determine alternative promoter usage, modulate the rate of transcription initiation or elongation, and direct the choice of alternative splice sites [217, 218]. For moderately expressed genes, DNA methylation level in the middle of the gene is correlated with higher transcription rates, being related to nucleosome positioning [219]. In genes with CpG-poor promoters, methylated sequences located downstream binds Polycomb repressor complexes [212], which are being associated with repression of promoters [220].

On the other hand, certain histone modifications may direct the choice of splice junction through direct interactions with proteins that mark exon–intron junctions, altering rates of transcription and nucleosome positioning [221, 222]. As we mentioned before, DNA methylation may also be involved in regulating alternative splicing, intron–exon junctions being enriched in sharp transitions in DNA methylation levels [223] (e.g., malignant prostate cancer cells have enrichment of DNA hypermethylation at exon–intron junctions [224]). Therefore, these findings highlight the involvement of DNA methylation levels in determining alternative splicing in tumor cells, suggesting that cancer-associated DNA hypomethylation in intronic and exonic sequences can modulate the amount and type of gene products and thereby contribute to tumor formation or progression.

TGFB2 gene contains an intronic Alu repeat that was found hypomethylated in some cancer cell lines. Their hypomethylation at this site might be related to the significant upregulation of TGFB2 gene, being an example of cancer-associated hypomethylation and a target chromatin associated epigenetic changes [225]. PRDM16 presents gene-body hypomethylation (overlapping an exon) in some of the cancer cell lines, whereas NOTCH2 also showed gene-body hypomethylation (in a subregion of repetitive DNA).

Gene encoding the protease urokinase (PLAU/uPA) is overexpressed and was found hypomethylated along with tumor progression in breast cancers and prostate cancers [157]. Also, other genes were observed to display hypomethylation and transcriptional activation in cancer, S100A4, mesothelin, claudin4, trefoil factor 2, maspin, PGP9.5, POMC, and the heparinase gene [144].

DNA hypomethylation is closely associated with morphological dedifferentiation in thyroid cancers. Four oncogenes (INSL4, DPPA2, TCL1B and NOTCH4) were frequently regulated by hypomethylation in anaplastic and medulary carcinoma [226].

Hematopoietin, TNF, IL1, IL10, and IL17 families of cytokines had a significant tendency to be hypomethylated in five cancer types (colon, kidney, stomach, lung, and breast) [227].

Hypomethylation and increased expression in cancer has been shown for R-RAS [228]. A strong association of CDH3 promoter demethylation and P-cadherin expression evident with histological grade and invasiveness in breast cancer was observed [229]. In Stage III and IV gastric cancer cyclin D2 activation is associated with promoter demethylation, activation of synuclein γ is associated with progression and metastatic potential in a range of solid tumors, and maspin expression in colorectal cancer is associated with microsatellite unstable tumors [230232].

Hypomethylation and overexpression of some imprinted genes, including the IGF-IIand H19 genes, are implicated in carcinogenesis [233235].

The putative oncogene, ELMO3, is overexpressed in non-small cell lung cancer in combination with hypomethylation of its promoter and these cancer-specific events are associated with the formation of metastases [236].

Aberrant hypomethylation and overexpression of WNT5A may be functionally important in the progression of prostate cancer. Along with WNT5A, S100P, and CRIP1, which have been previously implicated in cancer progression, are also regulated at the transcriptional level in prostate cancer by hypomethylation [237].

Evidence is accumulating for the biological significance and clinical relevance of DNA hypomethylation in cancer and for cancer-linked demethylation, and those seem to be highly dynamic processes.

Advertisement

5. MicroRNA genes

At present, a special consideration is given to small non-coding RNA molecules (microRNA) to their functions and involvement in human diseases. There are an extensive number of studies that link microRNA alterations to cancer pathogenesis. MicroRNA genes encode for a single RNA strand of about 21 to 23 nucleotides, which regulate gene expression by specifically targeting certain mRNAs in order to prevent them from coding for a specific protein. Some microRNA genes are mapped in chromosomal regions that undergo rearrangements, deletions, and amplifications in cancer. A growing amount of data demonstrates that microRNA genes display a different pattern of expression in various malignancies; they are found up-regulated or down-regulated and therefore can function either as oncogenes activating the malignant transformation (by down-regulating tumor-suppressor genes), or as tumor-suppressor genes blocking the malignant transformation (by down-regulating oncogenes). In numerous types of cancers, many different microRNA have been shown to act as oncogenes, their expression profiling presenting specific signatures associated with malignant transformation. Cancer-associated microRNA molecules are also called oncomir (oncomiR).

The first microRNA that has been proven to act as oncogene in human cancer was miR-17/92 polycistronic cluster known as OncomiR-1, which comprises six microRNAs: miR-17, miR-18a, miR-19a, miR-20a, miR-19b-1, and miR-92a-1 [238]. The miR-17/92 cluster is located in the locus of intron 3 of C13orf25 gene at 13q31.3, in a region frequently amplified in several types of lymphomas and solid tumor. It has been shown that the locus is amplified and overexpressed in human B cell lymphomas, malignant lymphoma cell lines and in lung cancers especially with small-cell lung cancer histology [239, 240]. Insertional mutagenesis studies using retroviruses indicates that miR-17/92 acts as an oncogene in T cell lymphomas; it was shown that soon after SL3-3 murine leukemia virus infection, mice developed tumors if provirus integrates into the proximity of the gene encoding miR-17/92 cistron [241]. Moreover, other studies uncovered that C-MYC and E2F3 gene products may induce miR-17/92 polycistronic expression through direct binding to the cluster promoter. Two microRNAs belonging to the cluster, miR-17-5p and miR-20a negatively regulate E2F1 activity, which confirms that the miR-17/92 can promote cell proliferation through the exchange of E2F1 to E2F3 pro-apoptotic proliferative [242]. Thus, miR-17/92 represents an anti-apoptotic oncogene and miR-20a inhibition using antisense oligonucleotides can induce apoptosis after treatment with doxorubicin [243].

Along with the miR-17/92 cluster from the miR-17 family, two other paralogue miRNA gene clusters are produced, miR-106b/25 and miR-106a/363, which possess oncogenic potential and are known to be involved in wide types of cancers. The miR-106b/25 cluster is located in intron 13 of the minichromosome maintenance complex component 7 (MCM7) oncogene at 7q22.1 and it contains the following three miRNAs: miR-106b, miR-93, and miR-25. Recently, findings sustain the oncogenic potential of this cluster and reports correlate miR-106b/25 member expression levels with processes such as tumor growth, cell survival, and angiogenesis [244, 245]. The oncogenic potential of the miR-106b/25 cluster in malignant transformation is achieved by targeting and down-regulation of several tumor-suppressor genes such as p21, E2F1, and PTEN [246248]. Furthermore, other recent work suggests that in breast cancer cells, miR-106b/25 cluster overexpression leads to overcoming doxorubicin-induced senescence and cells become drug resistant through a mechanism that involves targeting E-cadherin transcriptional activators EP300 [249]. The second cluster, miR-106a/363, is located on chromosome X (Xq26.2) and comprises of six miRNAs: miR-106a, miR-18b, miR-20b, miR-19b-2, miR-92a-2, and miR-363. A series of reports indicates an oncogenic potential for members of the cluster, for example miR-106a and miR-92-2 were found overexpressed in colon and prostate cancer and also in leukemia and Ewing Sarcoma [250252].

Another important oncomir is represented by miR-155 found overexpressed in several malignancies: chronic lymphocytic leukemia (CLL), B cell lymphoma, Hodgkin's lymphoma or Burkitt's type, and breast cancer. Some reports have shown that clinical isolates from B cell lymphomas, including those with large cells, contain a number of copies of miR-155, about 30 times higher than normal B cells [253256]. Also, results suggest that the pancreatic ductal adenocarcinomas overexpression of miR-155 determined decreased levels of TP53INP1 leading to apoptosis elusion and cell growth development [257].

A promising oncomir is also miR-21, one of the most common miRNA associated with human cancers. MicroRNA-21 high expression has been found in a variety of cancers including breast cancer, brain malignant tumors, glioblastomas, pancreatic, colorectal, liver, gastric, lung, skin, thyroid, ovarian, esophagus, prostate, cervical, and different lymphatic and hematopoietic cancers [250, 258263]. Elevated miR-21 levels have been linked to cell proliferation, apoptosis reduction, and cell migration in neoplastic transformation; it has been found that this oncomir targets and down-regulates a number of tumor-suppressor genes including PTEN, PDCD4, BCL2, RECK, JAG1, HNRPK, BTG2, TGFBRII, and thus sustaining cancer's invasion and metastasis [264, 265]. Moreover, experiments using transfection of MCF-7 cell lines with anti-miR-21 oligonucleotide conducted to cell growth suppression in vitro and tumor growth in vivo had an increase of programmed cell death rate [266].

Altogether, these studies illustrate a major role for microRNA genes in cancer pathogenesis (Table 2); many of them have oncogenic activity and could represent valuable biomarkers very useful for cancer screening or assessment of the therapeutic effects of anti-cancer treatments.

MicroRNA Cancer type miRNA function Potential targets Reference
miR-21 Glioblastoma Increase cell growth
Inhibit apoptosis
Promote cell cycle
HNRNPK|TP63 [267]
Breast - RECK|TIMP3 [268]
Increase cell growth
Increase metastasis
PTEN [269, 270]
Promote cell growth
Promote cell migration
Promote tumor growth
ANKRD46 [271]
- PDCD4
PTEN
BCL2
[272]
Promote cell invasion
Promote metastasis
TPM1
PDCD4
SERPINB5
[265]
Cervix Increase cell Proliferation
Inhibit apoptosis
Increase cell migration
CCL20 [273]
Promote cell proliferation PDCD4 [274]
Prostate Promote epithelial-mesenchymal transition BTG2 [275]
Induce tumor angiogenesis
Activate AKT/ERK signaling
[276]
Increase docetaxel resistance PDCD4 [277]
Promote cell proliferation
Promote tumor growth
[278]
Inhibit apoptosis
Promote cell motility
Promote cell invasion
MARCKS [279]
Blood Promote cell migration
Reduce apoptosis
Promote cell growth
PDCD4 [280, 281]
miR-221 and 222 Prostate Promote cell cycle G1/S transition
Increase colony formation
CDKN1B [282]
Breast Increase cell migration
Increase cell invasion
epithelial-mesenchymal transition
TRPS1 [283]
Increase tamoxifen resistance CDKN1B [284]
Glioblastoma Increase cell invasion TIMP3 [285]
Promote cell proliferation
Promote cell invasion
inhibit apoptosis
GJA1 [286]
Thyroid Promote cell transition to cell cycle S phase CDKN1B [287]
Lung Induce TRAIL resistance
Increase cell migration
PTEN
TIMP3
[288]
miR-155 Blood Increase cell proliferation
Increase clonogenity
Inhibit apoptosis
HDAC4 [289]
Promote PI3K-AKT signaling PIK3R1 [290]
Increase cell proliferation BCL2 [291]
Reduce cell cycle arrest SMAD5 [292]
Breast Promote cell survival
Induce chemoresistance
Inhibit apoptosis
FOXO3 [293]
Promote cell proliferation
Promote tumor development
SOCS1 [294]
miR-17/92
cluster
Blood Increase cell growth
Increase cell cycle G1/S transition
CDKN1A [295]
Increase cell proliferation
Increase imatinib -induced cell death
[296]
Suppress apoptosis PTEN [297, 298]
Retinoblastoma Promote cell proliferation p21 and p57 [299]
Colon Promote angiogenesis TSP-1
CTGF
[300]
Breast ERa [301]
Lung Promote cell growth
Reduce reactive oxygen species (ROS) generation
HIF-1a [302]
HIF-1a, PTEN, BCL2L11, CDKNA and TSP-1. [303,304]
miR-106b-25 cluster Breast Activate TGF-beta signaling
Induce epithelial mesenchymal transition
Induce a tumor initiating cell phenotype
SMAD7 [305]
Gastric E2F1 [306]
miR-191 Liver Promote epithelial-mesenchymal transition
Increase cell migration
Increase cell invasion
TIMP3 [307]
Promote cell proliferation
Inhibit apoptosis
Promote tumor growth
[308]
Gastric NDST1 [309]

Table 2.

OncomiRs in human cancer

(Source: OncomiRDB: http://bioinfo.au.tsinghua.edu.cn/member/jgu/oncomirdb/) [310].


Advertisement

6. Concluding remarks

Oncogene activation by structural alteration (chromosomal rearrangement, gene fusion, mutation, and gene amplification) or epigenetic modification (gene promoter hypomethylation, microRNA expression pattern) confers an increased or a deregulated expression. Therefore, cells with such alterations possess a growth advantage or an increased survival rate. Given the fact that expression profiling of these alterations determines specific signatures associated with tumor classification, diagnosis, staging, prognosis and response to treatment, it highlights the importance of studying oncogenes activation mechanisms and the great potential that they hold as therapeutic tools in the near future.

Advertisement

Acknowledgments

This work was supported by Romanian Research Grant PCCA135/2012 and POSDRU/159/1.5 /S/135760

References

  1. 1. Tomlins SA, Rhodes DR, Perner S, Dhanasekaran SM, Mehra R, Sun XW, Varambally S, Cao X, Tchinda J, Kuefer R, Lee C, Montie JE, Shah RB, Pienta KJ, Rubin MA, Chinnaiyan AM. Recurrent fusion of TMPRSS2 and ETS transcription factor genes in prostate cancer. Science. 2005;310:644–8. DOI: 10.1126/science.1117679.
  2. 2. Nakamura T, Mori T, Tada S, Krajewski W, Rozovskaia T, Wassell R, Dubois G, Mazo A, Croce CM, Canaani E. ALL-1 is a histone methyltransferase that assembles a supercomplex of proteins involved in transcriptional regulation. Mol Cell. 2002;10:1119–28. http://dx.doi.org/10.1016/S1097-2765(02)00740-2.
  3. 3. Tsujimoto Y, Cossman J, Jaffe E, Croce CM. Involvement of the bcl-2 gene in human follicular lymphoma. Science. 1985;228:1440–3.
  4. 4. Yeang CH, McCormick F, Levine A. Combinatorial patterns of somatic gene mutations in cancer. FASEB J. 2008;22(8):2605–22. DOI: 10.1096/fj.08-108985
  5. 5. Rodenhuis S. RAS and human tumors. Semin Cancer Biol. 1992;3:241–7.
  6. 6. Beaupre DM, Kurzrock R. RAS and leukemia: From basic mechanisms to gene-directed therapy. J Clin Oncol. 1999;17:1071–9.
  7. 7. Kolch W. Meaningful relationships: The regulation of the Ras/Raf/MEK/ERK pathway by protein interactions. Biochemical Journal. 2000;351(2):289–305.
  8. 8. Maldonado JL, Fridlyand J, Patel H, Jain AN, Busam K, Kageshita T, Ono T, Albertson DG, Pinkel D, Bastian BC. Determinants of BRAF mutations in primary melanomas. J Natl Cancer Inst. 2003;95:1878–90. DOI: 10.1093/jnci/djg123.
  9. 9. Turri-Zanoni M, Medicina D, Lombardi D, Ungari M, Balzarini P, Rossini C, Pellegrini W, Battaglia P, Capella C, Castelnuovo P, Palmedo G, Facchetti F, Kutzner H, Nicolai P, Vermi W. Sinonasal mucosal melanoma: Molecular profile and therapeutic implications from a series of 32 cases. Head Neck. 2013;35:1066–77. DOI: 10.1002/hed.23079.
  10. 10. Colombino M, Capone M, Lissia A, Cossu A, Rubino C, De Giorgi V, Massi D, Fonsatti E, Staibano S, Nappi O, Pagani E, Casula M, Manca A, Sini M, Franco R, Botti G, Caracò C, Mozzillo N, Ascierto PA, Palmieri G. BRAF/NRAS mutation frequencies among primary tumors and metastases in patients with melanoma. J Clin Oncol. 2012;30(20):2522–2529. DOI: 10.1200/JCO.2011.41.2452.
  11. 11. Long GV, Menzies AM, Nagrial AM, Haydu LE, Hamilton AL, Mann GJ, Hughes TM, Thompson JF, Scolyer RA, Kefford RF. Prognostic and clinicopathologic associations of oncogenic BRAF in metastatic melanoma. J Clin Oncol. 2011;29(10):1239–1246. DOI: 10.1200/JCO.2010.32.4327.
  12. 12. Menzies AM, Haydu LE, Visintin L, Carlino MS, Howle JR, Thompson JF, Kefford RF, Scolyer RA, Long GV. Distinguishing clinicopathologic features of patients with V600E and V600K BRAF-mutant metastatic melanoma. Clin Cancer Res. 2012;18(12):3242–3249. DOI: 10.1158/1078-0432.CCR-12-0052.
  13. 13. Burns PA, Jack A, Neilson F, Haddow S, Balmain A. Transformation of mouse skin endothelial cells in vivo by direct application of plasmid DNA encoding the human T24 H-ras oncogene. Oncogene. 1991;6:1973–8.
  14. 14. Padua RA, Barrass NC, Currie GA. Activation of N-ras in a human melanoma cell line. Mol Cell Biol. 1985;5:582–5.
  15. 15. Bucheit AD, Syklawer E, Jakob JA, Bassett RL Jr, Curry JL, Gershenwald JE, Kim KB, Hwu P, Lazar AJ, Davies MA. Clinical characteristics and outcomes with specific BRAF and NRAS mutations in patients with metastatic melanoma. Cancer. 2013;119(21):3821–3829. DOI: 10.1002/cncr.28306.
  16. 16. Terheyden P, Houben R, Pajouh P, Thorns C, Zillikens D, Becker JC. Response to Imatinib Mesylate Depends on the Presence of the V559A-Mutated KIT Oncogene Journal of Investigative Dermatology. 2010;130:314–316. DOI: 10.1038/jid.2009.197.
  17. 17. Beadling C, Jacobson-Dunlop E, Hodi FS, Le C, Warrick A, Patterson J, Town A, Harlow A, Cruz F 3rd, Azar S, Rubin BP, Muller S, West R, Heinrich MC, Corless CL. KIT gene mutations and copy number in melanoma subtypes. Clin Cancer Res. 2008;14:6821–8. DOI: 10.1158/1078-0432.CCR-08-0575.
  18. 18. Minoo P, Moyer MP, Jass JR. Role of BRAF-V600E in the serrated pathway of colorectal tumourigenesis. Journal of Pathology. 2007;212(2):124–133. DOI: 10.1002/path.2160.
  19. 19. Custodio A, Feliu J. Prognostic and predictive biomarkers for epidermal growth factor receptor-targeted therapy in colorectal cancer: Beyond KRAS mutations. Critical Reviews in Oncology/Hematology. 2013;85(1):45–81. DOI: 10.1016/j.critrevonc.2012.05.001.
  20. 20. Diaz LA Jr, Williams RT, Wu J, Kinde I, Hecht JR, Berlin J, Allen B, Bozic I, Reiter JG, Nowak MA, Kinzler KW, Oliner KS, Vogelstein B. The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers. Nature. 2012;486(7404):537–540. DOI: 10.1038/nature11219.
  21. 21. Zhu Z, Gandhi M, Nikiforova MN, Fischer AH, Nikiforov YE. Molecular profile and clinical-pathologic features of the follicular variant of papillary thyroid carcinoma. An unusually high prevalence of ras mutations. Am J Clin Pathol. 2003;120:71–7. DOI: 10.1309/ND8D9LAJTRCTG6QD.
  22. 22. Vasko V, Ferrand M, Di Cristofaro J, Carayon P, Henry JF, de Micco C. Specific pattern of RAS oncogene mutations in follicular thyroid tumors. J Clin Endocrinol Metab. 2003;88:2745–52. DOI: http://dx.doi.org/10.1210/jc.2002-021186.
  23. 23. Cohen Y, Xing M, Mambo E, Guo Z, Wu G, Trink B, Beller U, Westra WH, Ladenson PW, Sidransky D. BRAF mutation in papillary thyroid carcinoma. J Natl Cancer Inst. 2003;95:625–7. DOI: 10.1093/jnci/95.8.625.
  24. 24. Davies H, Bignell GR, Cox C, Stephens P, Edkins S, Clegg S, Teague J, Woffendin H, Garnett MJ, Bottomley W, Davis N, Dicks E, Ewing R, Floyd Y, Gray K, Hall S, Hawes R, Hughes J, Kosmidou V, Menzies A, Paterson H, Marais R, Marshall CJ, Wooster R, Stratton MR, Futreal PA. Mutations of the BRAF gene in human cancer. Nature. 2002;417:949–54. DOI: 10.1038/nature00766.
  25. 25. Xing M, Vasko V, Tallini G, Larin A, Wu G, Udelsman R, Ringel MD, Ladenson PW, Sidransky D. BRAF T1796A transversion mutation in various thyroid neoplasms. J Clin Endocrinol Metab. 2004;89:1365–8. DOI: http://dx.doi.org/10.1210/jc.2003-031488.
  26. 26. Durante C, Puxeddu E, Ferretti E, Morisi R, Moretti S, Bruno R, Barbi F, Avenia N, Scipioni A, Verrienti A, Tosi E, Cavaliere A, Gulino A, Filetti S, Russo D. BRAF mutations in papillary thyroid carcinomas inhibit genes involved in iodine metabolism. J Clin Endocrinol Metab. 2007;92:2840–3. DOI: http://dx.doi.org/10.1210/jc.2006-2707.
  27. 27. Cantwell-Dorris ER, O’Leary JJ, Sheils OM. BRAFV600E: Implications for carcinogenesis and molecular therapy. Mol Cancer Ther. 2011;10:385–94. DOI: 10.1158/1535-7163.MCT-10-0799.
  28. 28. Gouveia C, Can NT, Bostrom A, Grenert JP, van Zante A, Orloff LA. Lack of association of BRAF mutation with negative prognostic indicators in papillary thyroid carcinoma: The University of California, San Francisco, experience. JAMA Otolaryngol Head Neck Surg. 2013;139:1164–70. DOI: 10.1001/jamaoto.2013.4501.
  29. 29. Segouffin-Cariou C, Billaud M. Transforming ability of MEN2A-RET requires activation of the phosphatidylinositol 3-kinase/AKT signaling pathway. J Biol Chem. 2000;275:3568–7. DOI: 10.1074/jbc.275.5.3568.
  30. 30. Raue F, Frank-Raue K. Update multiple endocrine neoplasia type 2. Fam Cancer. 2010;9:449–57. DOI: 10.1007/s10689-010-9320-2.
  31. 31. Hedayati M, Zarif Yeganeh M, Sheikhol Eslami S, Rezghi Barez S, Hoghooghi Rad L, Azizi F. Predominant RET germline mutations in exons 10, 11, and 16 in Iranian patients with hereditary medullary thyroid carcinoma. J Thyroid Res. 2011;2011:264248. DOI: 10.4061/2011/264248.
  32. 32. Machens A, Niccoli-Sire P, Hoegel J, Frank-Raue K, van Vroonhoven TJ, Roeher HD, Wahl RA, Lamesch P, Raue F, Conte-Devolx B, Dralle H. European Multiple Endocrine Neoplasia (EUROMEN) Study Group. Early malignant progression of hereditary medullary thyroid cancer.N Engl J Med. 2003;349:1517–2. DOI: 10.1056/NEJMoa012915.
  33. 33. Figlioli G, Landi S, Romei C, Elisei R, Gemignani F. Medullary thyroid carcinoma (MTC) and RET proto-oncogene: Mutation spectrum in the familial cases and a meta-analysis of studies on the sporadic form. Mutat Res Rev Mutat Res. 2013;752:36–44. DOI: 10.1016/j.mrrev.2012.09.002.
  34. 34. Marx SJ, Stratakis CA. Multiple endocrine neoplasia--introduction. J Intern Med. 2005;257:2–5. DOI: 10.1111/j.1365-2796.2004.01419.x.
  35. 35. Eng C, Clayton D, Schuffenecker I, Lenoir G, Cote G, Gagel RF, van Amstel HK, Lips CJ, Nishisho I, Takai SI, Marsh DJ, Robinson BG, Frank-Raue K, Raue F, Xue F, Noll WW, Romei C, Pacini F, Fink M, Niederle B, Zedenius J, Nordenskjöld M, Komminoth P, Hendy GN, Mulligan LM. The relationship between specific RET proto-oncogene mutations and disease phenotype in multiple endocrine neoplasia type 2. International RET mutation consortium analysis. JAMA. 1996;276:1575–9. DOI: 10.1001/jama.1996.03540190047028.
  36. 36. Mulligan LM, Eng C, Healey CS, Clayton D, Kwok JB, Gardner E, Ponder MA, Frilling A, Jackson CE, Lehnert H. Specific mutations of the RET proto-oncogene are related to disease phenotype in MEN 2A and FMTC. Nat Genet. 1994;6:70–74. DOI: 10.1038/ng0194-70.
  37. 37. Kouvaraki MA, Shapiro SE, Perrier ND, Cote GJ, Gagel RF, Hoff AO, Sherman SI, Lee JE, Evans DB. RET proto-oncogene: A review and update of genotype-phenotype correlations in hereditary medullary thyroid cancer and associated endocrine tumors. Thyroid. 2005;15:531–44. DOI: 10.1089/thy.2005.15.531.
  38. 38. Miyauchi A, Futami H, Hai N, Yokozawa T, Kuma K, Aoki N, Kosugi S, Sugano K, Yamaguchi K. Two germline missense mutations at codons 804 and 806 of the RET proto-oncogene in the same allele in a patient with multiple endocrine neoplasia type 2B without codon 918 mutation. Jpn J Cancer Res. 1999;90:15.
  39. 39. Taketomi A1, Shirabe K, Muto J, Yoshiya S, Motomura T, Mano Y, Ikegami T, Yoshizumi T, Sugio K, Maehara Y. A rare point mutation in the Ras oncogene in hepatocellular carcinoma. Surg Today. 2013;43(3):289–92. DOI: 10.1007/s00595-012-0462-8.
  40. 40. Rajagopalan H, Bardelli A, Lengauer C, Kinzler KW, Vogelstein B, Velculescu VE. RAF/RAS oncogenes and mismatch-repair status. Nature. 2002;418:934. DOI: 10.1038/418934a.
  41. 41. Malats N, Porta M, Corominas JM, Piñol JL, Rifà J, Real FX. Ki-ras mutations in exocrine pancreatic cancer: Associations with clinicopathological characteristics and with tobacco and alcohol consumption. International Journal of Cancer. 1997;70:661–667. DOI: 10.1002/(SICI)1097-0215(19970317)70:6<661::AID-IJC6>3.0.CO;2-T.
  42. 42. Agarwal A1, Saif MW. KRAS in pancreatic cancer. JOP. 2014;15(4):303–5. DOI: 10.6092/1590-8577/2660.
  43. 43. Corcoran RB, Cheng KA, Hata AN, Faber AC, Ebi H, Coffee EM, Greninger P, Brown RD, Godfrey JT, Cohoon TJ, Song Y, Lifshits E, Hung KE, Shioda T, Dias-Santagata D, Singh A, Settleman J, Benes CH, Mino-Kenudson M, Wong KK, Engelman JA. Synthetic lethal interaction of combined BCL-XL and MEK inhibition promotes tumor regressions in KRAS mutant cancer models. Cancer Cell. 2013;23(1):121–128. DOI: 10.1016/j.ccr.2012.11.007.
  44. 44. Wright AA, Howitt BE, Myers AP, Dahlberg SE, Palescandolo E, Van Hummelen P, MacConaill LE, Shoni M, Wagle N, Jones RT, Quick CM, Laury A, Katz IT, Hahn WC, Matulonis UA, Hirsch MS. Oncogenic mutations in cervical cancer: Genomic differences between adenocarcinomas and squamous cell carcinomas of the cervix. Cancer. 2013;119(21):3776–83. DOI: 10.1002/cncr.28288.
  45. 45. Iida K, Nakayama K, Rahman MT, Rahman M, Ishikawa M, Katagiri A, Yeasmin S, Otsuki Y, Kobayashi H, Nakayama S, Miyazaki K. EGFR gene amplification is related to adverse clinical outcomes in cervical squamous cell carcinoma, making the EGFR pathway a novel therapeutic target. Br J Cancer. 2011;105(3):420–427. DOI: 10.1038/bjc.2011.22.
  46. 46. Courtney KD, Corcoran RB, Engelman JA. The PI3K pathway as drug target in human cancer. J Clin Oncol. 2010;28(6):1075–1083. DOI: 10.1200/JCO.2009.25.3641.
  47. 47. Schwarz JK, Payton JE, Rashmi R, Xiang T, Jia Y, Huettner P, Rogers BE, Yang Q, Watson M, Rader JS, Grigsby PW. Pathway-specific analysis of gene expression data identifies the PI3K/Akt pathway as a novel therapeutic target in cervical cancer. Clin Cancer Res. 2012;18(5):1464–1471. DOI: 10.1158/1078-0432.CCR-11-2485.
  48. 48. Wahl GM, Robert de Saint Vincent B, DeRose ML. Effect of chromosomal position on amplification of transfected genes in animal cells. Nature. 1984;307:516–520.
  49. 49. Albertson DG. Gene amplification in cancer. Trends Genet. 2006;22(8):447–55. DOI: http:// dx.doi.org /10.1016/j.tig.2006.06.007.
  50. 50. Gibaud A, Vogt N, Brison O, Debatisse M, Malfoy B. Characterization at nucleotide resolution of the homogeneously staining region sites of insertion in two cancer cell lines. Nucleic Acids Res. 2013;41(17):8210–9. DOI: 10.1093/nar/gkt566.
  51. 51. Windle BE, Wahl GM. Molecular dissection of mammalian gene amplification: New mechanistic insights revealed by analyses of very early events. Mutat. Res. 1992;276:199–224. DOI: 10.1016/0165-1110(92)90009-X.
  52. 52. Eisenberg DTA. An evolutionary review of human telomere biology: The thrifty telomere hypothesis and notes on potential adaptive paternal effects. American Journal of Human Biology. 2011;23(2):149–167. DOI: 10.1002/ajhb.21127.
  53. 53. Londono-Vallejo JA. Telomere length heterogeneity and chromosome instability. Cancer Lett. 2004;212(2):135–144. DOI: http://dx.doi.org/10.1016/j.canlet.2004.05.008.
  54. 54. Sabatier L, Ricoul M, Pottier G, Murnane JP. The loss of a single telomere can result in instability of multiple chromosomes in a human tumor cell line. Mol. Cancer Res. 2005;3:139–150. DOI: 10.1158/1541-7786.MCR-04-0194.
  55. 55. O'Hagan RC, Chang S, Maser RS, Mohan R, Artandi SE, Chin L, DePinho RA. Telomere dysfunction provokes regionalamplification and deletion in cancer genomes. Cancer Cell. 2002;2:149–155. DOI: http://dx.doi.org/10.1016/S1535-6108(02)00094-6.
  56. 56. Chang S, Khoo C, DePinho RA. Modeling chromosomal instability and epithelial carcinogenesis in the telomerase-deficient mouse. Semin. Cancer Biol. 2001;11:227–239. DOI: 10.1006/scbi.2000.0374.
  57. 57. Gisselsson D, Lv M, Tsao SW, Man C, Jin C, Höglund M, Kwong YL, Jin Y. Telomere-mediated mitotic disturbances in immortalized ovarian epithelial cells reproduce chromosomal losses and breakpoints from ovarian carcinoma. Genes Chromosomes Cancer. 2005;42:22–33. DOI: 10.1002/gcc.20094.
  58. 58. Gisselsson D, Höglund M. Connecting mitotic instability and chromosome aberrations in cancer—can telomeres bridge the gap? Semin. Cancer Biol. 2005;15:13–23. DOI: 10.1016/j.semcancer.2004.09.002.
  59. 59. Shen J, Gammon MD, Terry MB, Wang Q, Bradshaw P, Teitelbaum SL, Neugut AI, Santella RM. Telomere length, oxidative damage, antioxidants and breast cancer risk. Int J Cancer. 2009;124 (7):1637–43. DOI: 10.1002/ijc.24105.
  60. 60. Cawthon RM, Smith KR, O'Brien E, Sivatchenko A, Kerber RA. Association between telomere length in blood and mortality in people aged 60 years or older. Lancet. 2003;361(9355):393–395. DOI: 10.1016/s0140-6736(03)12384-7.
  61. 61. McClintock B. The stability of broken ends of chromosomes in Zea mays. Genetics. 1941;41:234–82.
  62. 62. Toledo F, Buttin G, Debatisse M. The origin of chromosome rearrangements at early stages of AMPD2 gene amplification in Chinese hamster cells. Curr Biol. 1993;3:255–64. DOI: http://dx.doi.org/10.1016/0960-9822(93)90175-N.
  63. 63. Londono-Vallejo JA. Telomere length heterogeneity and chromosome instability. Cancer Lett. 2004;212:135–144 41. DOI: http://dx.doi.org/10.1016/j.canlet.2004.05.008.
  64. 64. Snijders AM, Fridlyand J, Mans DA, Segraves R, Jain AN, Pinkel D, Albertson DG. Shaping of tumor and drug-resistant genomes by instability and selection. Oncogene. 2003;22:4370–4379. DOI: 10.1038/sj.onc.1206482.
  65. 65. Singer MJ, Mesner LD, Friedman CL, Trask BJ, Hamlin JL. Amplification of the human dihydrofolate reductase gene via double minutes is initiated by chromosome breaks. Proc Natl Acad Sci USA. 2000;97:7921–26. DOI: 10.1073/pnas.130194897.
  66. 66. Lo AW, Sabatier L, Fouladi B, Pottier G, Ricoul M, Murnane JP. DNA amplification by breakage/fusion/bridge cycles initiated by spontaneous telomere loss in a human cancer cell line. Neoplasia. 2002;4(6):531–8. DOI: 10.1038/sj.neo.7900267.
  67. 67. Ma K, Qiu L, Mrasek K, Zhang J, Liehr T, Quintana LG, Li Z. Common Fragile Sites: Genomic Hotspots of DNA Damage and Carcinogenesis. Int. J. Mol. Sci. 2012;13:11974–11999. DOI: 10.3390/ijms130911974.
  68. 68. Murnane JP, Sabatier L. Chromosome rearrangements resulting from telomere dysfunction and their role in cancer. Bioessays. 2004;26:1164–1174. DOI: 10.1002/bies.20125.
  69. 69. Glover TW, Arlt MF, Casper AM, Durkin SG. Mechanisms of common fragile site instability. Hum. Mol. Genet. 2005;2:R197–R205. DOI: 10.1093/hmg/ddi265.
  70. 70. Glover TW. Common fragile sites. Cancer Lett. 2006;232(1):4–12. DOI: http://dx.doi.org/ 10.1016/ j.canlet. 2005.08.032.
  71. 71. Lukusa T, Fryns JP. Human chromosome fragility. Biochimica et Biophysica Acta. 2008;1779(1):3–16. DOI: 10.1016/j.bbagrm.2007.10.005.
  72. 72. Durkin SG, Glover TW. Chromosome fragile sites. Annual Review of Genetics. 2007;41:169–192. DOI: 10.1146/annurev.genet.41.042007.165900.
  73. 73. Smith DI, Huang H, Wang L. Common fragile sites and cancer (review). International journal of oncology. 1998;12(1):187–96. DOI: 10.3892/ijo.12.1.187.
  74. 74. Glover TW, Berger C, Coyle J, Echo B. DNA polymerase alpha inhibition by aphidicolin induces gaps and breaks at common fragile sites in human chromosomes. Human Genetics. 1984;67(2):136–42. DOI: 10.1007/bf00272988.
  75. 75. Arlt MF, Glover TW. Inhibition of topoisomerase I prevents chromosome breakage at common fragile sites. DNA repair. 2010;9(6):678–89. DOI: 10.1016/j.dnarep.2010.03.005.
  76. 76. Shiraishi T, Druck T, Mimori K, Flomenberg J, Berk L, Alder H, Miller W, Huebner K, Croce CM. Sequence conservation at human and mouse orthologous common fragile regions, FRA3B/FHIT and Fra14A2/Fhit. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(10):5722–7. DOI: 10.1073/pnas.091095898.
  77. 77. Krummel KA, Denison SR, Calhoun E, Phillips LA, Smith DI. The common fragile site FRA16D and its associated gene WWOX are highly conserved in the mouse at Fra8E1. Genes, chromosomes & cancer. 2002;34(2):154–67. DOI: 10.1002/gcc.10047.
  78. 78. Schmid M, Ott G, Haaf T, Scheres JM. Evolutionary conservation of fragile sites induced by 5-azacytidine and 5-azadeoxycytidine in man, gorilla, and chimpanzee. Human Genetics. 1985;71(4):342–50. DOI: 10.1007/bf00388461.
  79. 79. Kang S, Ohshima K, Shimizu M, Amirhaeri S, Wells RD. Pausing of DNA Synthesis in Vitro at Specific Loci in CTG and CGG Triplet Repeats from Human Hereditary Disease Genes. Journal of Biological Chemistry. 1995;270(45):27014–27021. DOI: 10.1074/jbc.270.45.27014.
  80. 80. Zlotorynski E, Rahat A, Skaug J, Ben-Porat N, Ozeri E, Hershberg R, Levi A, Scherer SW, Margalit H, Kerem B. Molecular basis for expression of common and rare fragile sites. Molecular and Cellular Biology. 2003;23(20):7143–51. DOI: 10.1128/mcb.23.20.7143-7151.2003.
  81. 81. Luck G, Zimmer C, Reinert KE, Arcamone F. Specific interactions of distamycin A and its analogs with (A-T) rich and (G-C) rich duplex regions of DNA and deoxypolynucleotides. Nucleic Acids Research. 1977;4(8):2655–70. DOI: 10.1093/nar/4.8.2655.
  82. 82. Balakumaran BS, Freudenreich CH, Zakian VA. CGG/CCG repeats exhibit orientation-dependent instability and orientation-independent fragility in Saccharomyces cerevisiae. Human Molecular Genetics. 2000;9(1):93–100. DOI: 10.1093/hmg/9.1.93.
  83. 83. Yu S, Mangelsdorf M, Hewett D, Hobson L, Baker E, Eyre HJ, Lapsys N, Le Paslier D, Doggett NA, Sutherland GR, Richards RI. Human chromosomal fragile site FRA16B is an amplified AT-rich minisatellite repeat. Cell. 1997;88(3):367–74. DOI: 10.1016/S0092-8674(00)81875-9.
  84. 84. Zhang H, Freudenreich CH. An AT-Rich Sequence in Human Common Fragile Site FRA16D Causes Fork Stalling and Chromosome Breakage in S. cerevisiae. Molecular Cell. 2007;27(3):367–379. DOI: 10.1016/j.molcel.2007.06.012.
  85. 85. Aguilera A, Gómez-González B. Genome instability: A mechanistic view of its causes and consequences. Nature reviews. Genetics. 2008;9(3):204–17. DOI: 10.1038/nrg2268.
  86. 86. Casper AM, Nghiem P, Arlt MF, Glover TW. ATR regulates fragile site stability. Cell. 2002;111:779–789. DOI: http://dx.doi.org/10.1016/S0092-8674(02)01113-3.
  87. 87. Gorgoulis VG, Vassiliou LV, Karakaidos P, Zacharatos P, Kotsinas A, Liloglou T, Venere M, Ditullio RA Jr, Kastrinakis NG, Levy B, Kletsas D, Yoneta A, Herlyn M, Kittas C, Halazonetis TD. Activation of the DNA damage checkpoint and genomic instability in human precancerous lesions. Nature. 2005;434:907–913. DOI: 10.1038/nature03485.
  88. 88. Bartkova J, Horejsí Z, Koed K, Krämer A, Tort F, Zieger K, Guldberg P, Sehested M, Nesland JM, Lukas C, Ørntoft T, Lukas J, Bartek J. DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis. Nature. 2005;434:864–870.
  89. 89. Coquelle A, Pipiras E, Toledo F, Buttin G, Debatisse M. Expression of fragile sites triggers intrachromosomal mammalian gene amplification and sets boundaries to early amplicons. Cell. 1997;89:215–225. DOI: http://dx.doi.org/10.1016/S0092-8674(00)80201-9.
  90. 90. Kuo MT, Vyas RC, Jiang LX, Hittelman WN. Chromosome breakage at a major fragile site associated with P-glycoprotein gene amplification in multidrugresistant CHO cells. Mol. Cell. Biol. 1994;14:5202–5211. DOI: 10.1128/MCB.14.8.5202.
  91. 91. Miller CT, Lin L, Casper AM, Lim J, Thomas DG, Orringer MB, Chang AC, Chambers AF, Giordano TJ, Glover TW, Beer DG. Genomic amplification of MET with boundaries within fragile site FRA7G and upregulation of MET pathways in esophageal adenocarcinoma. Oncogene. 2006;25:409–418. DOI: 10.1038/sj.onc.1209057.
  92. 92. Wang ND, Testa JR, Smith DI. Determination of the specificity of aphidicolin-induced breakage of the human 3pl4.2 fragile site. Genomics. 1993;17:341–347. DOI: 10.1006/geno.1993.1330.
  93. 93. Göker E, Waltham M, Kheradpour A, Trippett T, Mazumdar M, Elisseyeff Y, Schnieders B, Steinherz P, Tan C, Berman E. Amplification of the dihydrofolate reductase gene is a mechanism of acquired resistance to methotrexate in patients with acute lymphoblastic leukemia and is correlated with p53 gene mutations. Blood. 1995;86(2):677–84.
  94. 94. Al-Kuraya K, Schraml P, Torhorst J, Tapia C, Zaharieva B, Novotny H, Spichtin H, Maurer R, Mirlacher M, Köchli O, Zuber M, Dieterich H, Mross F, Wilber K, Simon R, Sauter G. Prognostic relevance of gene amplifications and coamplifications in breast cancer. Cancer Res. 2004;64:8534–8540. DOI: 10.1158/0008-5472.CAN-04-1945.
  95. 95. Blegen H, Will JS, Ghadimi BM, Nash HP, Zetterberg A, Auer G, Ried T. DNA amplifications and aneuploidy, high proliferative activity and impaired cell cycle control characterize breast carcinomas with poor prognosis. Anal. Cell. Pathol. 2003;25:103–114.
  96. 96. Schwab M, Alitalo K, Klempnauer KH, Varmus HE, Bishop JM, Gilbert F, Brodeur G, Goldstein M, Trent J. Amplified DNA with limited homology to myc cellular oncogene is shared by human neuroblastoma cell lines and a neuroblastoma tumour. Nature. 1983;305:245–8.
  97. 97. Savelyeva L, Schwab M. Amplification of oncogenes revisited: From expression profiling to clinical application. Cancer Lett. 2001;167:115–123. DOI: http://dx.doi.org/10.1016/S0304-3835(01)00472-4.
  98. 98. Várkondi E, Gyory F, Nagy A, Kiss I, Ember I, Kozma L. Oncogene amplification and overexpression of oncoproteins in thyroid papillary cancer. In Vivo. 2005;19(2):465–70.
  99. 99. Croce CM. Oncogenes and Cancer. N Engl J Med. 2008;358:502–11. DOI: 10.1056/NEJMra072367.
  100. 100. Yu W, Imoto I, Inoue J, Onda M, Emi M, Inazawa J. A novel amplification target, DUSP26, promotes anaplastic thyroid cancer cell growth by inhibiting p38 MAPK activity. Oncogene. 2007;26(8):1178–87. DOI: 10.1038/sj.onc.1209899.
  101. 101. Anzick SL, Kononen J, Walker RL, Azorsa DO, Tanner MM, Guan XY, Sauter G, Kallioniemi OP, Trent JM, Meltzer PS. AIB1, a steroid receptor coactivator amplified in breast and ovarian cancer. Science. 1997;277:965–968. DOI: 10.1126/science.277.5328.965.
  102. 102. Lahusen T, Henke RT, Kagan BL, Wellstein A, Riegel AT. The role and regulation of the nuclear receptor co-activator AIB1 in breast cancer. Breast Cancer Res Treat. 2009;116:225–237. DOI: 10.1007/s10549-009-0405-2.
  103. 103. Ory V, Tassi E, Cavalli LR, Sharif GM, Saenz F, Baker T, Schmidt MO, Mueller SC, Furth PA, Wellstein A, Riegel AT. The nuclear coactivator amplified in breast cancer 1 maintains tumor-initiating cells during development of ductal carcinoma in situ. Oncogene. 2014;33(23):3033–42. DOI: 10.1038/onc.2013.263.
  104. 104. York B, O'Malley BW. Steroid receptor coactivator (SRC) family: Masters of systems biology. J Biol Chem. 2010;285:38743–38750. DOI: 10.1074/jbc.R110.193367.
  105. 105. Ma YY, Wei SJ, Lin YC, Lung JC, Chang TC, Whang-Peng J, Liu JM, Yang DM, Yang WK, Shen CY. PIK3CA as an oncogene in cervical cancer. Oncogene. 2000;19(23):2739–44.
  106. 106. Chen S, Auletta T, Dovirak O, Hutter C, Kuntz K, El-ftesi S, Kendall J, Han H, Von Hoff DD, Ashfaq R, Maitra A, Iacobuzio-Donahue CA, Hruban RH, Lucito R. Copy number alterations in pancreatic cancer identify recurrent PAK4 amplification. Cancer Biol Ther. 2008;7(11):1793–802. DOI: 10.4161/cbt.7.11.6840.
  107. 107. Pan JS, Cai JY, Xie CX, Zhou F, Zhang ZP, Dong J, Xu HZ, Shi HX, Ren JL. Interacting with HBsAg compromises resistance of jumping translocation breakpoint protein to ultraviolet radiation-induced apoptosis in 293FT cells. Cancer Lett. 2009;285:151–6. DOI: 10.1016/j.canlet.2009.05.009.
  108. 108. Huebner K, Kastury K, Druck T, Salcini AE, Lanfrancone L, Pelicci G, Lowenstein E, Li W, Park SH, Cannizzaro L. Chromosome locations of genes encoding human signal transduction adapter proteins, Nck (NCK), Shc (SHC1), and Grb2 (GRB2). Genomics. 1994;22:281–7. DOI: 10.1006/geno.1994.1385.
  109. 109. Kraut-Cohen J, Muller WJ, Elson A. Protein-tyrosine phosphatase epsilon regulates Shc signaling in a kinase-specific manner: Increasing coherence in tyrosine phosphatase signaling. J Biol Chem. 2008;283:4612–21. DOI: 10.1074/jbc.M708822200.
  110. 110. Shen DY, Fang ZX, You P, Liu PG, Wang F, Huang CL, Yao XB, Chen ZX, Zhang ZY. Clinical significance and expression of cyclin kinase subunits 1 and 2 in hepatocellular carcinoma. Liver Int. 2009;30:19–25. DOI: 10.1111/j.1478-3231.2009.02106.x.
  111. 111. Chen L, Chan TH, Guan XY. Chromosome 1q21 amplification and oncogenes in hepatocellular carcinoma. Acta Pharmacol Sin. 2010;31(9):1165-71. DOI: 10.1038/aps.2010.94.
  112. 112. Zhang S, Liang X, Zheng X, Huang H, Chen X, Wu K, Wang B, Ma S, Glo1 genetic amplification as a potential therapeutic target in hepatocellular carcinoma. Int J Clin Exp Pathol. 2014;7(5):2079–90.
  113. 113. Mariani O, Brennetot C, Coindre JM, Gruel N, Ganem C, Delattre O, Stern MH, Aurias A. JUN oncogene amplification and overexpression block adipocytic differentiation in highly aggressive sarcomas. Cancer Cell. 2007;11(4):361–74. DOI: http://dx.doi.org/10.1016/j.ccr.2007.02.007.
  114. 114. Endo M, Yasui K, Nakajima T, Gen Y, Tsuji K, Dohi O, Zen K, Mitsuyoshi H, Minami M, Itoh Y, Taniwaki M, Tanaka S, Arii S, Okanoue T, Yoshikawa T. Infrequent amplification of JUN in hepatocellular carcinoma. Anticancer Res. 2009;29(12):4989–94.
  115. 115. Eswarakumar VP, Lax I, Schlessinger J. Cellular signaling by fibroblast growth factor receptors. Cytokine Growth Factor Rev. 2005;16:139–49. DOI: http://dx.doi.org/10.1016/j.cytogfr.2005.01.001.
  116. 116. Brooks AN, Kilgour E, Smith PD. Molecular pathways: Fibroblast growth factor signaling: a new therapeutic opportunity in cancer. Clin Cancer Res. 2012;18:1855–62. DOI: 10.1158/1078-0432.CCR-11-0699.
  117. 117. Katoh M, Katoh M. FGF signaling network in the gastrointestinal tract (review). Int J Oncol. 2006;29:163–8. DOI: 10.3892/ijo.29.1.163.
  118. 118. Katoh M. Genetic alterations of FGF receptors: An emerging field in clinical cancer diagnostics and therapeutics. Expert Rev Anticancer Ther. 2010;10:1375–9. DOI: 10.1586/era.10.128.
  119. 119. Grose R, Dickson C. Fibroblast growth factor signaling in tumorigenesis. Cytokine Growth Factor Rev. 2005;16:179–86. DOI: http://dx.doi.org/10.1016/j.cytogfr.2005.01.003.
  120. 120. Xie L, Su X, Zhang L, Yin X, Tang L, Zhang X, Xu Y, Gao Z, Liu K, Zhou M, Gao B, Shen D, Zhang L, Ji J, Gavine PR, Zhang J, Kilgour E, Zhang X, Ji Q. FGFR2 gene amplification in gastric cancer predicts sensitivity to the selective FGFR inhibitor AZD4547. Clin Cancer Res. 2013;19(9):2572–83. DOI: 10.1158/1078-0432.CCR-12-3898.
  121. 121. Mitelman F, Johansson B and Mertens F. The impact of translocations and gene fusions on cancer causation. Nat Rev Cancer. 2007;7:233–245. DOI: 10.1038/nrc2091.
  122. 122. Nambiar M, Kari V and Raghavan SC. Chromosomal translocations in cancer. Biochim Biophys Acta. 2008;1786:139–152. DOI: 10.1016/j.bbcan.2008.07.005.
  123. 123. Fröhling S and Döhner H. Chromosomal abnormalities in cancer. N Engl J Med. 2008;359:722–734. DOI: 10.1056/NEJMra0803109.
  124. 124. Pui CH, Relling MV and Downing JR. Acute lymphoblastic leukemia. N Engl J Med. 2004;350:1535–1548.
  125. 125. Soupir CP, Dal Cin P, Muzikansky A, Kantarjian H, Jones D, Hasserjian RP. Philadelphia chromosome-positive acute myeloid leukemia: A rare aggressive leukemia with clinicopathologic features distinct from chronic myeloid leukemia in myeloid blast crisis. Am J Clin Pathol. 2007;127(4):642–50. DOI: 10.1016/S0140-6736(08)60457-2.
  126. 126. Collins SJ. Retinoic acid receptors, hematopoiesis and leukemogenesis. Curr Opin Hematol. 2008;15:346–351. DOI: 10.1097/MOH.0b013e3283007edf.
  127. 127. Stams WA, den Boer ML, Beverloo HB, Meijerink JP, van Wering ER, Janka-Schaub GE, Pieters R. 2005; Expression levels of TEL, AML1, and the fusion products TEL-AML1 and AML1-TEL versus drug sensitivity and clinical outcome in t(12;21)-positive pediatric acute lymphoblastic leukemia. Clin. Cancer Res. 2005;11(8):2974–80. DOI: 10.1158/1078-0432.CCR-04-1829.
  128. 128. Roix JJ, McQueen PG, Munson PJ, Parada LA and Misteli T. Spatial proximity of translocation-prone gene loci in human lymphomas. Nat Genet. 2011;34:287–291. DOI: 10.1038/ng1177.
  129. 129. Misteli T. The inner life of the genome. Sci Am. 2003;304:66–73.
  130. 130. Osborne CS, Chakalova L, Mitchell JA, Horton A, Wood AL, Bolland DJ, Corcoran AE, Fraser P. Myc dynamically and preferentially relocates to a transcription factory occupied by Igh. PLoS Biol. 2007;5:e1922007. DOI: 10.1371/journal.pbio.0050192.
  131. 131. Gandhi M, Evdokimova V and Nikiforov YE. Mechanisms of chromosomal rearrangements in solid tumors: The model of papillary thyroid carcinoma. Mol Cell Endocrinol. 2010;321:36–43. DOI: 10.1016/j.mce.2009.09.013.
  132. 132. Merolla F, Pentimalli F, Pacelli R, Vecchio G, Fusco A, Grieco M, Celetti A. Involvement of H4(D10S170) protein in ATM-dependent response to DNA damage. Oncogene. 2007;26:6167–6175. DOI: 10.1038/sj.onc.1210446.
  133. 133. Nikiforov YE. Thyroid carcinoma: Molecular pathways and therapeutic targets. Mod Pathol. 2008;21(2):S37–S43. DOI: 10.1038/modpathol.2008.10.
  134. 134. Ciampi R, Giordano TJ, Wikenheiser-Brokamp K, Koenig RJ, Nikiforov YE. HOOK3-RET: A novel type of RET/PTC rearrangement in papillary thyroid carcinoma. Endocr Relat Cancer. 2007;14:445–452. DOI: 10.1677/ERC-07-0039.
  135. 135. Gandhi M, Medvedovic M, Stringer JR, Nikiforov YE. Interphase chromosome folding determines spatial proximity of genes participating in carcinogenic RET/PTC rearrangements. Oncogene. 2006;25:2360–2366. DOI: 10.1038/sj.onc.1209268.
  136. 136. Sankar S, Lessnick SL. Promiscuous partnerships in Ewing’s sarcoma. Cancer Genet. 2011;204:351–365. DOI: 10.1016/j.cancergen.2011.07.008.
  137. 137. Patel M, Simon JM, Iglesia MD, Wu SB, McFadden AW, Lieb JD, Davis IJ. Tumor-specific retargeting of an oncogenic transcription factor chimera results in dysregulation of chromatin and transcription. Genome Res. 2012;22:259–270. DOI: 10.1101/gr.125666.111.
  138. 138. Zelent A, Greaves M, Enver T. Role of the TEL-AML1 fusion gene in the molecular pathogenesis of childhood acute lymphoblastic leukaemia. Oncogene. 2004;23:4275–4283. DOI: 10.1038/sj.onc.1207672.
  139. 139. Bohlander SK. ETV6: A versatile player in leukemogenesis. Semin Cancer Biol. 2005;15:162–174. DOI: 10.1016/j.semcancer.2005.01.008.
  140. 140. Lin C, Yang L, Tanasa B, Hutt K, Ju BG, Ohgi K, Zhang J, Rose DW, Fu XD, Glass CK, Rosenfeld MG. Nuclear receptor-induced chromosomal proximity and DNA breaks underlie specific translocations in cancer. Cell. 2009;139:1069–1083. DOI: 10.1016/j.cell.2009.11.030.
  141. 141. Li Z, Tognon CE, Godinho FJ, Yasaitis L, Hock H, Herschkowitz JI, Lannon CL, Cho E, Kim SJ, Bronson RT, Perou CM, Sorensen PH, Orkin SH. ETV6-NTRK3 fusion oncogene initiates breast cancer from committed mammary progenitors via activation of AP1 complex. Cancer Cell. 2007;12:542–558. DOI: http://dx.doi.org/10.1016/j.ccr.2007.11.012.
  142. 142. Seth A, Watson DK. ETS transcription factors and their emerging roles in human cancer. Eur J Cancer. 2005;41:2462–2478.
  143. 143. Gama-Sosa MA, Slagel VA, Trewyn RW, Oxenhandler R, Kuo KC, Gehrke CW, Ehrlich M. The 5-methylcytosine content of DNA from human tumors. Nucleic Acids Res. 1983;11:6883–6894. DOI: 10.1093/nar/11.19.6883.
  144. 144. Ehrlich M. DNA hypomethylation in cancer cells Epigenomics. 2009;1(2):239–259. DOI: 10.2217/epi.09.33.
  145. 145. Alvarez H, Opalinska J, Zhou L, Sohal D, Fazzari MJ, Yu Y, Montagna C, Montgomery EA, Canto M, Dunbar KB, Wang J, Roa JC, Mo Y, Bhagat T, Ramesh KH, Cannizzaro L, Mollenhauer J, Thompson RF, Suzuki M, Meltzer SJ, Melnick A, Greally JM, Maitra A,Verma A. Widespread hypomethylation occurs early and synergizes with gene ampli fi cation during esophageal carcinogenesis. PLoS Genet. 2011;7(3):e1001356. DOI: 10.1371/journal.pgen.1001356.
  146. 146. Sandoval J, Heyn HA, Moran S, Serra-Musach J, Pujana MA, Bibikova M, Esteller M. Validation of a DNA methylation microarray for 450,000 CpG sites in the human genome.Epigenetics. 2011;6(6):692–702. DOI: 10.4161/epi.6.6.16196.
  147. 147. Ruike Y, Imanaka Y, Sato F, Shimizu K, Tsujimoto G. Genome-wide analysis of aberrant methylation in human breast cancer cells using methyl-DNA immunoprecipitation combined with high-throughput sequencing. BMC Genomics. 2010;11:137. DOI: 10.1186/1471-2164-11-137.
  148. 148. Hansen KD, Timp W, Bravo HC, Sabunciyan S, Langmead B, McDonald OG, Wen B, Wu H, Liu Y, Diep D, Briem E, Zhang K, Irizarry RA, Feinberg AP. Increased methylationvariation in epigenetic domains across cancer types. Nat Genet. 2011;43(8):768–775. DOI: 10.1038/ng.865.
  149. 149. Ferguson AT, Vertino PM, Spitzner JR, Baylin SB, Muller MT, Davidson NE. Role of estrogen receptor gene demethylation and DNA methyltransferase. DNA adduct formation in 5-aza-2’deoxycytidine-induced cytotoxicity in human breast cancer cells. J Biol Chem. 1997;272:32260–32266. DOI: 10.1074/jbc.272.51.32260.
  150. 150. Ehrlich M. DNA hypomethylation, cancer, the immunodeficiency, centromeric region instability, facial anomalies syndrome and chromosomal rearrangements. J Nutr. 2002;132:2424S–2429S.
  151. 151. Ehrlich M, Woods CB, Yu MC, Dubeau L, Yang F, Campan M, Weisenberger DJ, Long T, Youn B, Fiala ES, Laird PW. Quantitative analysis of associations between DNA hypermethylation, hypomethylation, and DNMT RNA levels in ovarian tumors. Oncogene. 2006;25:2636–2645. DOI: 10.1038/sj.onc.1209145.
  152. 152. Cash HL, Tao L, Yuan JM, Marsit CJ, Houseman EA, Xiang YB, Gao YT, Nelson HH, Kelsey KT. LINE-1hypomethylation is associated with bladder cancer risk among nonsmoking Chinese. Int J Cancer. 2012;130(5):1151–9. DOI: 10.1002/ijc.26098.
  153. 153. Eden A, Gaudet F, Waghmare A, Jaenisch R. Chromosomal instability and tumors promoted by DNA hypomethylation. Science. 2003;300(5618):455. DOI: 10.1126/science.1083557.
  154. 154. Yamada Y, Jackson-Grusby L, Linhart H, Meissner A, Eden A, Lin H, Jaenisch R. Opposing effects of DNA hypomethylation on intestinal and liver carcinogenesis. Proc NatlAcad Sci USA. 2005;102(38):13580–13585. DOI: 10.1073/pnas.0506612102.
  155. 155. Cadieux B, Ching TT, Vandenberg SR, Costello JF. Genome-wide hypomethylation in human glioblastomas associated with specific copy number alteration, methylenetetrahydrofolatereductase allele status, and increased proliferation. Cancer Res. 2006;66(17):8469–8476. DOI: 10.1158/0008-5472.CAN-06-1547.
  156. 156. Goering W, Ribarska T, Schulz WA. Selective changes of retroelement expression in human prostate cancer. Carcinogenesis. 2011;32(10):1484–92. DOI: 10.1093/carcin/bgr181.
  157. 157. Pulukuri SM, Estes N, Patel J, Rao JS. Demethylation-linked activation of urokinase plasminogen activator is involved in progression of prostate cancer. Cancer Res. 2007;67:930–939. DOI: 10.1158/0008-5472.CAN-06-2892.
  158. 158. Bird A. DNA methylation patterns and epigenetic memory. Genes Dev. 2002;16:6–21. DOI: 10.1101/gad.947102.
  159. 159. Fazzari MJ, Greally JM. Epigenomics: Beyond CpG islands. Nat. Rev.Genet. 2004;5:446–455. DOI: 10.1038/nrg1349.
  160. 160. Tuck-Muller CM, Narayan A, Tsien F, Smeets DF, Sawyer J, Fiala ES, Sohn OS, Ehrlich M. DNA hypomethylation and unusual chromosome instability in cell lines from ICF syndrome patients, Cytogenet. Cell Genet. 2000;89:121–128. DOI: 10.1159/000015590.
  161. 161. Reik W, Dean W, Walter J. Epigenetic reprogramming in mammalian development. Science. 2001;293:1089–1093. DOI: 10.1126/science.1063443.
  162. 162. Avner P, Heard E. X-chromosome inactivation: Counting, choice and initiation. Nat. Rev. Genet. 2001;2:59–67. DOI: 10.1038/35047580.
  163. 163. Pfeifer GP, Rauch TA. DNA methylation patterns in lung carcinomas. Semin Cancer Biol. 2009;19(3):181–187. DOI: 10.1016/j.semcancer.2009.02.00.
  164. 164. Park SY, Yoo EJ, Cho NY, Kim N, Kang GH. Comparison of CpG island hypermethylation and repetitive DNA hypomethylation in premalignant stages of gastric cancer, stratified for Helicobacter pylori infection. J Pathol. 2009;219(4):410–6. DOI: 10.1002/path.2596.
  165. 165. Ehrlich M. Cancer-linked DNA hypomethylation and its relationship to hypermethylation. Curr Top Microbiol Immunol. 2006;310:251–274.
  166. 166. Ehrlich M. DNA methylation in cancer: Too much, but also too little. Oncogene. 2002;21(35):5400–5413. DOI: 10.1038/sj.onc.1205651.
  167. 167. Feber A, Wilson GA, Zhang L, Presneau N, Idowu B, Down TA, Rakyan VK, Noon LA, Lloyd AC, Stupka E, Schiza V, Teschendorff AE, Schroth GP, Flanagan A, Beck S. Comparative methylome analysis of benign and malignant peripheral nerve sheath tumors. Genome Res. 2011;21(4):515–524. DOI: 10.1101/gr.109678.110.
  168. 168. Rodriguez J, Vives L, Jorda M, Morales C, Munoz M, Vendrell E, Peinado MA. Genome-wide tracking of unmethylated DNA Alu repeats in normal and cancer cells. Nucleic Acids Res. 2008;36(3):770–784. DOI: 10.1093/nar/gkm1105.
  169. 169. Wilson AS, Power BE, Molloy PL. DNA hypomethylation and human diseases Biochimica et Biophysica Acta. 2007;1775(1):138–162. DOI: 10.1016/j.bbcan.2006.08.007.
  170. 170. Feinberg AP, Tycko B. The history of cancer epigenetics. Nat. Rev. Cancer. 2004;4:143–153. DOI: 10.1038/nrc1279.
  171. 171. Widschwendter M, Jiang G, Woods C, Muller HM, Fiegl H, Goebel G, Marth C, Holzner EM, Zeimet AG, Laird PW, Ehrlich M. DNA hypomethylation and ovarian cancer biology. Cancer Res. 2004;64(13):4472–4480. DOI: 10.1158/0008-5472.CAN-04-0238.
  172. 172. Ehrlich M, Hopkins N, Jiang G, Dome JS, Yu MS, Woods CB, Tomlinson GE, Chintagumpala M, Champagne M, Diller L, Parham DM, Sawyer J. Satellite hypomethylation in karyotyped Wilms tumors. Cancer Genet Cytogenet. 2003;141:97–105. DOI: http://dx.doi.org/10.1016/S0165-4608(02)00668-4.
  173. 173. Narayan A, Ji W, Zhang X-Y, Marrogi A, Graff JR, Baylin SB, Ehrlich M. Hypomethylation of pericentromeric DNA in breast adenocarcinomas. Int J Cancer. 1998;77:833–838. DOI: 10.1002/(SICI)1097-0215(19980911)77:6<833::AID-IJC6>3.0.CO;2-V.
  174. 174. Qu G, Dubeau L, Narayan A, Yu M, Ehrlich M. Satellite DNA hypomethylation vs. overall genomic hypomethylation in ovarian epithelial tumors of different malignant potential. Mut Res. 1999;423:91–101. DOI: 10.1016/S0027-5107(98)00229-2.
  175. 175. Roman AC, Gonzalez-Rico FJ, Molto E, Hernando H, Neto A, Vicente-Garcia C, Ballestar E, Gomez-Skarmeta JL, Vavrova-Anderson J, White RJ, Montoliu L, Fernandez-Salguero PM. Dioxin receptor and SLUG transcription factors regulate the insulator activity of B1 SINE retrotransposons via an RNA polymerase switch. Genome Res. 2011;21(3):422–432. DOI: 10.1101/gr.111203.110.
  176. 176. Wang J, Lunyak VV, Jordan IK. Genome-wide prediction and analysis of human chromatin boundary elements. Nucleic Acids Res. 2011;40(2):511–29. DOI: 10.1093/nar/gkr750.
  177. 177. Hansen KD, Timp W, Bravo HC, Sabunciyan S, Langmead B, McDonald OG, Wen B, Wu H, Liu Y, Diep D, Briem E, Zhang K, Irizarry RA, Feinberg AP. Increased methylation variation in epigenetic domains across cancer types. Nat Genet. 2011;43(8):768–775. DOI: 10.1038/ng.865.
  178. 178. Irizarry RA, Ladd-Acosta C, Wen B, Wu Z, Montano C, Onyango P, Cui H, Gabo K, Rongione, M, Webster M, Ji H, Potash JB, Sabunciyan S, Feinberg AP. The human colon cancer methylome shows similar hypo- and hypermethylation at conserved tissue-speci fi c CpG island shores. Nat Genet. 2009;41(2):178–186. DOI: 10.1038/ng.298.
  179. 179. Brenet F, Moh M, Funk P, Feierstein E, Viale AJ, Socci ND, Scandura JM. DNA methylation of the fi rst exon is tightly linked to transcriptional silencing. PLoS One. 2011;6(1):e14524. DOI: 10.1371/journal.pone.0014524.
  180. 180. Andrews J, Kennette W, Pilon J, Hodgson A, Tuck AB, Chambers AF, Rodenhiser DI. Multi-platform whole-genome microarray analyses refine the epigenetic signature of breast cancer metastasis with gene expression and copy number. PLoS One. 2010;5(1):e8665. DOI: 10.1371/journal.pone.0008665.
  181. 181. Coolen MW, Stirzaker C, Song JZ, Statham AL, Kassir Z, Moreno CS, Young AN, Varma V, Speed TP, Cowley M, Lacaze P, Kaplan W, Robinson MD, Clark SJ. Consolidation of the cancer genome into domains of repressive chromatin by long-range epigenetic silencing (LRES) reduces transcriptional plasticity. Nat Cell Biol. 2010;12(3):235–246. DOI: 10.1038/ncb2023.
  182. 182. Widschwendter M, Jiang G, Woods C, Muller HM, Fiegl H, Goebel G, Marth C, Holzner EM, Zeimet AG, Laird PW, Ehrlich M. DNA hypomethylation and ovarian cancer biology. Cancer Res. 2004;64(13):4472–4480. DOI: 10.1158/0008-5472.CAN-04-0238.
  183. 183. Dante R, Dante-Paire J, Rigal D, Roizes G. Methylation patterns of long interspersed repeated DNA and alphoid repetitive DNA from human cell lines and tumors. Anticancer Res. 1992;12(2):559–563.
  184. 184. Weisenberger DJ, Campan M, Long TI, Kim M, Woods C, Fiala E, Ehrlich M, Laird PW. Analysis of repetitive element DNA methylation by MethyLight. Nucleic Acids Res. 2005;33(21):6823–6836. DOI: 10.1093/nar/gki987.
  185. 185. Schulz WA, Steinhoff C, Florl AR. Methylation of endogenous human retroelements in health and disease. Curr Top Microbiol Immunol. 2006;310:211–250.
  186. 186. Chan SW, Zilberman D, Xie Z, Johansen LK, Carrington JC, Jacobsen SE. RNA silencing genes control de novo DNA methylation. Science. 2004;303(5662):1336. DOI: DOI: 10.1126/science.1095989.
  187. 187. Qu G (a), Grundy PE, Narayan A, Ehrlich M. Frequent hypomethylation in Wilms tumors of pericentromeric DNA in chromosomes 1 and 16. Cancer Genet Cytogenet. 1999;109:34–39. DOI: http://dx.doi.org/10.1016/S0165-4608(98)00143-5.
  188. 188. Thoraval D, Asakawa J, Wimmer K, Kuick R, Lamb B, Richardson B, Ambros P, Glover T, Hanash S. Demethylation of repetitive DNA sequences in neuroblastoma. Genes Chromosomes Cancer. 1996;17(4):234–244. DOI: 10.1016/S0027-5107(98)00229-2.
  189. 189. Nagai H, Baba M, Konishi N, Kim YS, Nogami M, Okumura K, Emi M, Matsubara K. Isolation of NotI clusters hypomethylated in HBV-integrated hepatocellular carcinomas by two-dimensional electrophoresis. DNA Res. 1999;6(4):219–225. DOI: 10.1093/dnares/6.4.219.
  190. 190. Itano O, Ueda M, Kikuchi K, Hashimoto O, Hayatsu S, Kawaguchi M, Seki H, Aiura K, Kitajima M. Correlation of postoperative recurrence in hepatocellular carcinoma with demethylation of repetitive sequences. Oncogene. 2002;21(5):789–797. DOI: 10.1038/sj/onc/1205124.
  191. 191. Katargin AN, Pavlova LS, Kisseljov FL, Kisseljova NP. Hypermethylation of genomic 3.3-kb repeats is frequent event in HPV-positive cervical cancer. BMC Med Genomics. 2009;2:30. DOI: 10.1186/1755-8794-2-30.
  192. 192. Szpakowski S, Sun X, Lage JM, Dyer A, Rubinstein J, Kowalski D, Sasaki C, Costa J, Lizardi PM. Loss of epigenetic silencing in tumors preferentially affects primate-speci fi c retroelements. Gene. 2009;48(2):151–167. DOI: 10.1016/j.gene.2009.08.006.
  193. 193. Choi SH, Worswick S, Byun HM, Shear T, Soussa JC, Wolff EM, Douer D, Garcia-Manero, G, Liang G, Yang AS. Changes in DNA methylation of tandem DNA repeats are different from interspersed repeats in cancer. Int J Cancer. 2009;125(3):723–729. DOI: 10.1002/ijc.24384.
  194. 194. Itano O, Ueda M, Kikuchi K, Shimazu M, Kitagawa Y, Aiura K, Kitajima M. A new predictive factor for hepatocellular carcinoma based on two- dimensional electrophoresis of genomic DNA. Oncogene. 2000;19(13):1676–1683.
  195. 195. Grunau C, Sanchez C, Ehrlich M, van der Bruggen P, Hindermann W, Rodriguez C, Krieger S, De Sario A. Frequent DNA hypomethylation in the human juxtacentromeric BAGE loci in cancer. Genes Chrom Cancer. 2005;43(1):11–24. DOI: 10.1002/gcc.20155.
  196. 196. Roman-Gomez J, Jimenez-Velasco A, Agirre X, Castillejo JA, Navarro G, San Jose-Eneriz E, Garate L, Cordeu L, Cervantes F, Prosper F, Heiniger A, Torres A. Repetitive DNA hypomethylation in the advanced phase of chronic myeloid leukemia. Leuk Res. 2008;32(3):487–490. DOI: http://dx.doi.org/10.1016/j.leukres.2007.07.021.
  197. 197. Yegnasubramanian S, Haffner MC, Zhang Y, Gurel B, Cornish TC, Wu Z, Irizarry RA, Morgan J, Hicks J, DeWeese TL, Isaacs WB, Bova GS, De Marzo AM, Nelson WG. DNA hypomethylation arises later in prostate cancer progression than CpG island hypermethylation and contributes to metastatic tumor heterogeneity. Cancer Res. 2008;68(21):8954–8967. DOI: 10.1158/0008-5472.CAN-07-6088.
  198. 198. Horiuchi A, Hayashi T, Kikuchi N, Hayashi A, Fuseya C, Shiozawa T, Konishi I. Hypoxia upregulates ovarian cancer invasiveness via the binding of HIF-1alpha to a hypoxia-induced, methylation-free hypoxia response element of S100A4 gene. Int J Cancer. 2012;131:1755–1767. DOI: 10.1002/ijc.27448.
  199. 199. Irizarry RA, Ladd-Acosta C, Wen B, Wu Z, Montano C, Onyango P, Cui H, Gabo K, Rongione M, Webster M, Ji H, Potash JB, Sabunciyan S, Feinberg AP.The human colon cancer methylome shows similar hypo- and hypermethylation at conserved tissue-specific CpG island shores. Nat Genet. 2009;41:178–186. DOI: 10.1038/ng.298.
  200. 200. Smale ST. Pioneer factors in embryonic stem cells and differentiation. Curr Opin Genet Dev. 2010;20(5):519–526. DOI: 10.1016/j.gde.2010.06.010.
  201. 201. Serandour AA, Avner S, Percevault F, Demay F, Bizot M, Lucchetti-Miganeh C, Barloy-Hubler F, Brown M, Lupien M, Metivier R, Salbert G, Eeckhoute J. Epigenetic switch involved in activation of pioneer factor FOXA1-dependent enhancers. Genome Res. 2011;21(4):555–565. DOI: 10.1101/gr.111534.110.
  202. 202. Taube JH, Allton K, Duncan SA, Shen L, Barton MC. Foxa1 functions as a pioneer transcription factor at transposable elements to activate Afp during differentiation of embryonic stem cells. J Biol Chem. 2010;285(21):16135–16144. DOI: 10.1074/jbc.M109.088096.
  203. 203. Magnani L, Eeckhoute J, Lupien M. Pioneer factors: Directing transcriptional regulators within the chromatin environment. Trends Genet. 2011;27(11):465–74. DOI: 10.1016/j.tig.2011.07.002.
  204. 204. Hatada I, Namihira M, Morita S, Kimura M, Horii T, Nakashima K. Astrocyte-specific genes are generally demethylated in neural precursor cells prior to astrocytic differentiation. PLoS One. 2008;3(9):e3189. DOI: 10.1371/journal.pone.0003189.
  205. 205. Son KS, Kang HS, Kim SJ, Jung SY, Min SY, Lee SY, Kim SW, Kwon Y, Lee KS, Shin KH, Ro J. Hypomethylation of the interleukin-10 gene in breast cancer tissues. Breast. 2010;19:484–488. DOI: 10.1016/j.breast.2010.05.011.
  206. 206. Wang Z, Zhang J, Zhang Y, Lim SH. SPAN-Xb expression in myeloma cells is dependent on promoter hypomethylation and can be upregulated pharmacologically. Int J Cancer. 2006;118:1436–1444. DOI: 10.1002/ijc.21499.
  207. 207. Noushmehr H, Weisenberger DJ, Diefes K, Phillips HS, Pujara K, Berman BP, Pan F, Pelloski CE, Sulman EP, Bhat KP, Verhaak RG, Hoadley KA, Hayes DN, Perou CM, Schmidt HK, Ding L, Wilson RK, Van Den Berg D, Shen H, Bengtsson H, Neuvial P, Cope LM, Buckley J, Herman JG, Baylin SB, Laird PW, Aldape K; Cancer Genome Atlas Research Network. Identification of a CpG island methylator phenotype that defines a distinct subgroup of glioma. Cancer Cell. 2010;17:510–522. DOI: 10.1016/j.ccr.2010.03.017.
  208. 208. Hoffmann MJ, Schulz WA. Causes and consequences of DNA hypomethylation in human cancer. Biochem Cell Biol. 2005;83:296–321. DOI: 10.1139/o05-036.
  209. 209. Eckhardt F, Lewin J, Cortese R, Rakyan VK, Attwood J, Burger M, Burton J, Cox TV, Davies R, Down TA, Hae fl iger C, Horton R, Howe K, Jackson DK, Kunde J, Koenig C, Liddle J, Niblett D, Otto T, Pettett R, Seemann S, Thompson C, West T, Rogers J, Olek A, Berlin K, Beck S. DNA methylation pro fi ling of human chromosomes 6, 20 and 22. Nat Genet. 2006;38(12):1378–1385. DOI: 10.1038/ng1909.
  210. 210. Baba Y, Nosho K, Shima K, Huttenhower C, Tanaka N, Hazra A, Giovannucci EL, Fuchs CS, Ogino S. Hypomethylation of the IGF2 DMR in colorectal tumors, detected by bisulfite pyrosequencing, is associated with poor prognosis. Gastroenterology. 2010;139(6):1855–1864. DOI: 10.1053/j.gastro.2010.07.050.
  211. 211. Ball MP, Li JB, Gao Y, Lee JH, LeProust EM, Park IH, Xie B, Daley GQ, Church GM. Targeted and genome-scale strategies reveal gene-body methylation signatures in human cells. Nat Biotechnol. 2009;27(4):361–368. DOI: 10.1038/nbt.1533.
  212. 212. Wu H, Coskun V, Tao J, Xie W, Ge W, Yoshikawa K, Li E, Zhang Y, Sun YE. Dnmt3a-dependent nonpromoter DNA methylation facilitates transcription of neurogenic genes. Science. 2010;329(5990):444–448. DOI: 10.1126/science.
  213. 213. Lister R, Pelizzola M, Dowen RH, Hawkins RD, Hon G, Tonti-Filippini J, Nery JR, Lee L, Ye Z, Ngo QM, Edsall L, Antosiewicz-Bourget J, Stewart R, Ruotti V, Millar AH, Thomson JA, Ren B, Ecker JR. Human DNA methylomes at base resolution show widespread epigenomic differences. Nature. 2009;462(7271):315–322. DOI: 10.1038/nature08514.
  214. 214. De Bustos C, Ramos E, Young JM, Tran RK, Menzel U, Langford CF, Eichler EE, Hsu L, Henikoff S, Dumanski JP, Trask BJ. Tissue-speci fi c variation in DNA methylation levels along human chromosome 1. Epigenetics Chromatin. 2009;2(1):7. DOI: 10.1186/1756-8935-2-7.
  215. 215. Ke XS, Qu Y, Cheng Y, Li WC, Rotter V, Oyan AM, Kalland KH. Global pro fi ling of histone and DNA methylation reveals epigenetic-based regulation of gene expression during epithelial to mesenchymal transition in prostate cells. BMC Genomics. 2010;11:669. DOI: 10.1186/1471-2164-11-669.
  216. 216. Cheng X, Blumenthal RM. Coordinated chromatin control: Structural and functional linkage of DNA and histone methylation. Biochemistry. 2010;49(14):2999–3008. DOI: 10.1021/bi100213t.
  217. 217. Stengel S, Fiebig U, Kurth R, Denner J. Regulation of human endogenous retrovirus-K expression in melanomas by CpG methylation. Genes Chromosomes Cancer. 2010;49(5):401–411. DOI: 10.1002/gcc.20751.
  218. 218. Appanah R, Dickerson DR, Goyal P, Groudine M, Lorincz MC. An unmethylated 3’ promoter-proximal region is required for ef fi cient transcription initiation. PLoS Genet. 2007;3(2):e27. DOI: 10.1371/journal.pgen.0030027.
  219. 219. Chodavarapu RK, Feng S, Bernatavichute YV, Chen PY, Stroud H, Yu Y, Hetzel JA, Kuo F, Kim J, Cokus SJ, Casero D, Bernal M, Huijser P, Clark AT, Kramer U, Merchant SS, Zhang X, Jacobsen SE, Pellegrini M. Relationship between nucleosome positioning and DNA methylation. Nature. 2010;466(7304):388–392. DOI: 10.1038/nature09147.
  220. 220. Maunakea AK, Nagarajan RP, Bilenky M, Ballinger TJ, D’Souza C, Fouse SD, Johnson BE, Hong C, Nielsen C, Zhao Y, Turecki G, Delaney A, Varhol R, Thiessen N, Shchors K, Heine VM, Rowitch DH, Xing X, Fiore C, Schillebeeckx M, Jones SJ, Haussler D, Marra MA, Hirst M, Wang T, Costello JF. Conserved role of intragenic DNA methylation in regulating alternative promoters. Nature. 2010;466(7303):253–257. DOI: 10.1038/nature09165.
  221. 221. Hodges E, Smith AD, Kendall J, Xuan Z, Ravi K, Rooks M, Zhang MQ, Ye K, Bhattacharjee A, Brizuela L, McCombie WR, Wigler M, Hannon GJ, Hicks JB. High definition profiling of mammalian DNA methylation by array capture and single molecule bisulfite sequencing. Genome Res. 2009;19(9):1593–1605. DOI: 10.1101/gr.095190.109.
  222. 222. Luco RF, Allo M, Schor IE, Kornblihtt AR, Misteli. Epigenetics in alternative premRNA splicing. Cell. 2011;144(1):16–26. DOI: 10.1016/j.cell.2010.11.056.
  223. 223. Laurent L, Wong E, Li G, Huynh T, Tsirigos A, Ong CT, Low HM, Kin Sung KW, Rigoutsos I, Loring J, Wei CL. Dynamic changes in the human methylome during differentiation. Genome Res. 2010;20(3):320–331. DOI: 10.1101/gr.101907.109.
  224. 224. Yegnasubramanian S, Wu Z, Haffner MC, Esopi D, Aryee MJ, Badrinath R, He TL, Morgan JD, Carvalho B, Zheng Q, De Marzo AM, Irizarry RA, Nelson WG. Chromosome wide mapping of DNA methylation patterns in normal and malignant prostate cells reveals pervasive methylation of gene-associated and conserved intergenic sequences. BMC Genomics. 2011;12:313. DOI: 10.1186/1471-2164-12-313.
  225. 225. Nagai H, Kim YS, Yasuda T, Ohmachi Y, Yokouchi H, Monden M, Emi M, Konishi N, Nogami M, Okumura K, Matsubara K. A novel sperm-speci fi c hypomethylation sequence is a demethylation hotspot in human hepatocellular carcinomas. Gene. 1999;237(1):15–20. DOI: 10.1016/S0378-1119(99)00322-4.
  226. 226. Faam B, Ghaffari MA, Ghadiri A, Azizi F. Epigenetic modifications in human thyroid cancer. Biomed Rep. 2015;3(1):3-8. DOI: 10.3892/br.2014.375.
  227. 227. Shen X, He Z, Li H, Yao C, Zhang Y, He L, Li S, Huang J, Guo Z. Distinct functional patterns of gene promoter hypomethylation and hypermethylation in cancer genomes. PLoS One. 2012;7(9):e44822. DOI: 10.1371/journal.pone.0044822.
  228. 228. Nishigaki M, Aoyagi K, Danjoh I, Fukaya M, Yanagihara K, Sakamoto H, Yoshida T, Sasaki H. Discovery of aberrant expression of R-RAS by cancer-linked DNA hypomethylation in gastric cancer using microarrays, Cancer Res. 2005;65:2115–2124. DOI: 10.1158/0008-5472.CAN-04-3340.
  229. 229. Paredes J, Albergaria A, Oliveira JT, Jerónimo C, Milanezi F, Schmitt FC. P-cadherin overexpression is an indicator of clinical outcome in invasive breast carcinomas and is associated with CDH3 promoter hypomethylation. Clin. Cancer Res. 2005;11:5869–5877. DOI: 10.1158/1078-0432.CCR-05-0059.
  230. 230. Oshimo Y, Nakayama H, Ito R, Kitadai Y, Yoshida K, Chayama K, Yasui W. Promoter methylation of cyclin D2 gene in gastric carcinoma, Int. J. Oncol. 2003;23:1663–1670. DOI: 10.3892/ijo.23.6.1663.
  231. 231. Liu H, Liu W, Wu Y, Zhou Y, Xue R, Luo C, Wang L, Zhao W, Jiang JD, Liu J. Loss of epigenetic control of synuclein-gamma gene as a molecular indicator of metastasis in a wide range of human cancers, Cancer Res. 2005;65:7635–7643. DOI: 10.1158/0008-5472.CAN-05-1089.
  232. 232. Bettstetter M, Woenckhaus M, Wild PJ, Rümmele P, Blaszyk H, Hartmann A, Hofstädter F, Dietmaier W. Elevated nuclear maspin expression is associated with microsatellite instability and high tumour grade in colorectal cancer, J. Pathol. 2005;205:606–614. DOI: 10.1002/path.1732.
  233. 233. Holm TM, Jackson-Grusby L, Brambrink T, Yamada Y, Rideout WM 3rd, Jaenisch R. Global loss of imprinting leads to widespread tumorigenesis in adult mice. Cancer Cell. 2005;8:275–285. DOI: http://dx.doi.org/10.1016/j.ccr.2005.09.007.
  234. 234. Pogribny IP, Beland FA. DNA hypomethylation in the origin and pathogenesis of human diseases. Cell Mol Life Sci. 2009;66:2249–2261. DOI: 10.1007/s00018-009-0015-5.
  235. 235. Tycko, B. Genomic imprinting and human neoplasia. In: Ehrlich, M., editor. DNA and Alterations in Cancer: Genetic and Epigenetic Alterations. Eaton Publishing; MA, USA: 2000. p. 333-349.
  236. 236. Søes S, Daugaard IL, Sørensen BS, Carus A, Mattheisen M, Alsner J, Overgaard J, Hager H, Hansen LL, Kristensen LS. Hypomethylation and increased expression of the putative oncogene ELMO3 are associated with lung cancer development and metastases formation. Oncoscience. 2014;1(5):367–74. eCollection 2014.
  237. 237. Wang Q, Williamson M, Bott S, Brookman-Amissah N, Freeman A, Nariculam J, Hubank MJ, Ahmed A, Masters JR. Hypomethylation of WNT5A, CRIP1 and S100P in prostate cancer. Oncogene. 2007;4;26(45):6560–5. DOI: 10.1038/sj.onc.1210472.
  238. 238. Ota A, Tagawa H, Karnan S, Tsuzuki S, Karpas A, Kira S, Yoshida Y, Seto M. Identification and characterization of a novel gene, C13orf25, as a target for 13q31-q32 amplification in malignant lymphoma. Cancer Res. 2004;64(9):3087–95. DOI: 10.1158/0008-5472.CAN-03-3773.
  239. 239. Tagawa H, Seto M. A microRNA cluster as a target of genomic amplification in malignant lymphoma.Leukemia. 2005;19(11):2013–6. DOI: 10.1038/sj.leu.2403942.
  240. 240. Hayashita Y, Osada H, Tatematsu Y, Yamada H, Yanagisawa K, Tomida S, Yatabe Y, Kawahara K, Sekido Y, Takahashi T. A polycistronic microRNA cluster, miR-17-92, is overexpressed in human lung cancers and enhances cell proliferation. Cancer Res. 2005;65(21):9628–32. DOI: 10.1158/0008-5472.CAN-05-2352.
  241. 241. Wang CL, Wang BB, Bartha G, Li L, Channa N, Klinger M, Killeen N, Wabl M. Activation of an oncogenic microRNA cistron by provirus integration.Proc Natl Acad Sci U S A. 2006;103(49):18680–4. DOI: 10.1073/pnas.0609030103.
  242. 242. Woods K, Thomson JM, Hammond SM. Direct regulation of an oncogenic micro-RNA cluster by E2F transcription factors. J Biol Chem. 2007;282(4):2130–4. DOI: 10.1074/jbc.C600252200.
  243. 243. Sylvestre Y, De Guire V, Querido E, Mukhopadhyay UK, Bourdeau V, Major F, Ferbeyre G, Chartrand P. An E2F/miR-20a autoregulatory feedback loop. J Biol Chem. 2007;282(4):2135−43. DOI: 10.1074/jbc.M608939200.
  244. 244. Fang L, Deng Z, Shatseva T, Yang J, Peng C, Du WW, Yee AJ, Ang LC, He C, Shan SW, Yang BB. MicroRNA miR-93 promotes tumor growth and angiogenesis by targeting integrin-beta8. Oncogene. 2011;30:806–821. DOI: 10.1038/onc.2010.465.
  245. 245. Ivanovska I, Ball AS, Diaz RL, Magnus JF, Kibukawa M, Schelter JM, Kobayashi SV, Lim L, Burchard J, Jackson AL, Linsley PS, Cleary MA. MicroRNAs in the miR-106b family regulate p21/CDKN1A and promote cell cycle progression. Mol Cell Biol. 2008;28:2167–2174. DOI: 10.1128/MCB.01977-07.
  246. 246. Poliseno L, Salmena L, Riccardi L, Fornari A, Song MS, Hobbs RM, Sportoletti P, Varmeh S, Egia A, Fedele G, Rameh L, Loda M, Pandolfi PP. Identification of the miR-106b~25 microRNA cluster as a proto-oncogenic PTEN-targeting intron that cooperates with its host gene MCM7 in transformation. Sci Signal. 2010;3(117):ra29. DOI: 10.1126/scisignal.2000594.
  247. 247. Petrocca F, Visone R, Onelli MR, Shah MH, Nicoloso MS, de Martino I, Iliopoulos D, Pilozzi E, Liu CG, Negrini M, Cavazzini L, Volinia S, Alder H, Ruco LP, Baldassarre G, Croce CM, Vecchione A. E2F1-regulated microRNAs impair TGF[beta]-dependent cell-cycle arrest and apoptosis in gastric cancer. Cancer Cell. 2008;13:272–286. DOI: 10.1016/j.ccr.2008.02.013.
  248. 248. Smith AL, Iwanaga R, Drasin DJ, Micalizzi DS, Vartuli RL, Tan AC, Ford HL. The miR-106b-25 cluster targets Smad7, activates TGF-beta signaling, and induces EMT and tumor initiating cell characteristics downstream of Six1 in human breast cancer. Oncogene. 2012;31:5162–5171. DOI: 10.1038/onc.2012.11.
  249. 249. Zhou Y, Hu Y, Yang M, Jat P, Li K, Lombardo Y, Xiong D, Coombes RC, Raguz S, Yagüe E. The miR-106b~25 cluster promotes bypass of doxorubicin-induced senescence and increase in motility and invasion by targeting the E-cadherin transcriptional activator EP300. Cell Death Differ. 2014;21(3):462–74. DOI: 10.1038/cdd.2013.167.
  250. 250. Volinia S, Calin GA, Liu CG, Ambs S, Cimmino A, Petrocca F, Visone R, Iorio M, Roldo C, Ferracin M, Prueitt RL, Yanaihara N, Lanza G, Scarpa A, Vecchione A, Negrini M, Harris CC, Croce CM. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc Natl Acad Sci U S A. 2006;103:2257–61. DOI: 10.1073/pnas.0510565103.
  251. 251. Yu J, Wang F, Yang GH, Wang FL, Ma YN, Du ZW, Zhang JW. Human microRNA clusters: Genomic organization and expression profile in leukemia cell lines. Biochem Biophys Res Commun. 2006;349:59–68. DOI: 10.1016/j.bbrc.2006.07.207.
  252. 252. Dylla L, Jedlicka P. Growth-promoting role of the miR-106a~363 cluster in Ewing sarcoma. PLoS One. 2013;8(4):e63032. DOI: 10.1371/journal.pone.0063032.
  253. 253. Fulci V, Chiaretti S, Goldoni M, Azzalin G, Carucci N, Tavolaro S, Castellano L, Magrelli A, Citarella F, Messina M, Maggio R, Peragine N, Santangelo S, Mauro FR, Landgraf P, Tuschl T, Weir DB, Chien M, Russo JJ, Ju J, Sheridan R, Sander C, Zavolan M, Guarini A, Foà R, Macino G. Quantitative technologies establish a novel microRNA profile of chronic lymphocytic leukemia. Blood. 2007;109(11):4944–51. DOI: http://dx.doi.org/10.1182/blood-2006-12-062398.
  254. 254. Eis PS, Tam W, Sun L, Chadburn A, Li Z, Gomez MF, Lund E, Dahlberg JE. Accumulation of miR-155 and BIC RNA in human B cell lymphomas. Proc Natl Acad Sci U S A. 20058;102(10):3627–32. DOI: 10.1073/pnas.0500613102.
  255. 255. Metzler M, Wilda M, Busch K, Viehmann S, Borkhardt A. High expression of precursor microRNA-155/BIC RNA in children with Burkitt lymphoma. Genes Chromosomes Cancer. 2004;39(2):167–9. DOI: 10.1002/gcc.10316.
  256. 256. Iorio MV, Ferracin M, Liu CG, Veronese A, Spizzo R, Sabbioni S, Magri E, Pedriali M, Fabbri M, Campiglio M, Ménard S, Palazzo JP, Rosenberg A, Musiani P, Volinia S, Nenci I, Calin GA, Querzoli P, Negrini M, Croce CM. MicroRNA gene expression deregulation in human breast cancer. Cancer Res. 2005;65(16):7065–70. DOI: 10.1158/0008-5472.CAN-05-1783.
  257. 257. Gironella M, Seux M, Xie MJ, Cano C, Tomasini R, Gommeaux J, Garcia S, Nowak J, Yeung ML, Jeang KT, Chaix A, Fazli L, Motoo Y, Wang Q, Rocchi P, Russo A, Gleave M, Dagorn JC, Iovanna JL, Carrier A, Pébusque MJ, Dusetti NJ. Tumor protein 53-induced nuclear protein 1 expression is repressed by miR-155, and its restoration inhibits pancreatic tumor development. Proc Natl Acad Sci U S A. 2007;104(41):16170–5. DOI: 10.1073/pnas.0703942104.
  258. 258. Krichevsky AM, Gabriely G. miR-21: A small multi-faceted RNA. J Cell Mol Med. 2009;13(1):39–53. DOI: 10.1111/j.1582-4934.2008.00556.
  259. 259. Chan JA, Krichevsky AM, Kosik KS: MicroRNA-21 is an antiapoptotic factor in human glioblastoma cells. Cancer Res. 2005;65:6029–6033. DOI: 10.1158/0008-5472.CAN-05-0137.
  260. 260. Iorio MV, Visone R, Di Leva G, Donati V, Petrocca F, Casalini P, Taccioli C, Volinia S, Liu CG, Alder H, Calin GA, Ménard S, Croce CM. MicroRNA signatures in human ovarian cancer. Cancer Res. 2007;67(18):8699–707. DOI: 10.1158/0008-5472.
  261. 261. Lui WO, Pourmand N, Patterson BK, Fire A. Patterns of known and novel small RNAs in human cervical cancer. Cancer Res.2007;67(13):6031–43. DOI: 10.1158/0008-5472.
  262. 262. Hu Y, Correa AM, Hoque A, Guan B, Ye F, Huang J, Swisher SG, Wu TT, Ajani JA, Xu XC. Prognostic significance of differentially expressed miRNAs in esophageal cancer". Int. J. Cancer. 2011;128(1):132–43. DOI: 10.1002/ijc.25330.
  263. 263. Tetzlaff MT, Liu A, Xu X, Master SR, Baldwin DA, Tobias JW, Livolsi VA, Baloch ZW. Differential expression of miRNAs in papillary thyroid carcinoma compared to multinodular goiter using formalin fixed paraffin embedded tissues. Endocr. Pathol. 2007;18(3):163–73. DOI: 10.1007/s12022-007-0023-7.
  264. 264. Asangani IA, Rasheed SA, Nikolova DA, Leupold JH, Colburn NH, Post S, Allgayer H. MicroRNA-21 (miR-21) post-transcriptionally downregulates tumor suppressor Pdcd4 and stimulates invasion, intravasation and metastasis in colorectal cancer. Oncogene. 2008;27(15):2128–36. DOI: 10.1038/sj.onc.1210856.
  265. 265. Zhu S, Wu H, Wu G, Nie D, Sheng S, Mo YY. MicroRNA-21 targets tumor suppressor genes in invasion and metastasis. Cell Res. 2008;18:350–359. DOI: 10.1038/cr.2008.24.
  266. 266. Cheng AM, Byrom MW, Shelton J, Ford LP. Antisense inhibition of human miRNAs and indications for an involvement of miRNA in cell growth and apoptosis. Nucleic Acids Res. 2005;33(4):1290–7. DOI: 10.1093/nar/gki200.
  267. 267. Papagiannakopoulos T, Shapiro A, Kosik KS. MicroRNA-21 targets a network of key tumor-suppressive pathways in glioblastoma cells. Cancer Res. 2008;68(19):8164–72. DOI: 10.1158/0008-5472.
  268. 268. Gabriely G, Wurdinger T, Kesari S, Esau CC, Burchard J, Linsley PS, Krichevsky AM. MicroRNA 21 promotes glioma invasion by targeting matrix metalloproteinase regulators. 2. Mol Cell Biol. 2008;28(17):5369–80. DOI: 10.1128/MCB.00479-08.
  269. 269. Mandal CC, Ghosh-Choudhury T, Dey N, Choudhury GG, Ghosh-Choudhury N. miR-21 is targeted by omega-3 polyunsaturated fatty acid to regulate breast tumor CSF-1 expression. Carcinogenesis. 2012;33(10):1897–908. DOI: 10.1093/carcin/bgs198.
  270. 270. Wang ZX, Lu BB, Wang H, Cheng ZX, Yin YM. MicroRNA-21 modulates chemosensitivity of breast cancer cells to doxorubicin by targeting PTEN. Arch Med Res. 2011;42(4):281–90. DOI: 10.1016/j.arcmed.2011.06.008.
  271. 271. Yan LX, Wu QN, Zhang Y, Li YY, Liao DZ, Hou JH, Fu J, Zeng MS, Yun JP, Wu QL, Zeng YX, Shao JY. Knockdown of miR-21 in human breast cancer cell lines inhibits proliferation, in vitro migration and in vivo tumor growth. Breast Cancer Res. 2011;13(1):R2. DOI: 10.1186/bcr2803.
  272. 272. Wickramasinghe NS, Manavalan TT, Dougherty SM, Riggs KA, Li Y, Klinge CM. Estradiol downregulates miR-21 expression and increases miR-21 target gene expression in MCF-7 breast cancer cells. Nucleic Acids Res. 2009;37(8):2584–95. DOI: 10.1093/nar/gkp117.
  273. 273. Yao T, Lin Z. MiR-21 is involved in cervical squamous cell tumorigenesis and regulates CCL20.Biochim Biophys Acta. 2012;1822(2):248–60. DOI: 10.1016/j.bbadis.2011.09.018.
  274. 274. Yao Q, Xu H, Zhang QQ, Zhou H, Qu LH. MicroRNA-21 promotes cell proliferation and down-regulates the expression of programmed cell death 4 (PDCD4) in HeLa cervical carcinoma cells. Biochem Biophys Res Commun. 2009;388(3):539–42. DOI: 10.1016/j.bbrc.2009.08.044.
  275. 275. Coppola V, Musumeci M, Patrizii M, Cannistraci A, Addario A, Maugeri-Saccà M, Biffoni M, Francescangeli F, Cordenonsi M, Piccolo S, Memeo L, Pagliuca A, Muto G, Zeuner A, De Maria R, Bonci D. BTG2 loss and miR-21 upregulation contribute to prostate cell transformation by inducing luminal markers expression and epithelial-mesenchymal transition. Oncogene. 2013;32(14):1843–53. DOI: 10.1038/onc.2012.194.
  276. 276. Liu LZ, Li C, Chen Q, Jing Y, Carpenter R, Jiang Y, Kung HF, Lai L, Jiang BH. MiR-21-induced angiogenesis through AKT and ERK activation and HIF-1α expression. PLoS One. 2011;6(4):e19139. DOI: 10.1371/journal.pone.0019139.
  277. 277. Shi GH, Ye DW, Yao XD, Zhang SL, Dai B, Zhang HL, Shen YJ, Zhu Y, Zhu YP, Xiao WJ, Ma CG. Involvement of microRNA-21 in mediating chemo-resistance to docetaxel in androgen-independent prostate cancer PC3 cells. Acta Pharmacol Sin. 2010;31(7):867–73. DOI: 10.1038/aps.2010.48.
  278. 278. Ribas J, Ni X, Haffner M, Wentzel EA, Salmasi AH, Chowdhury WH, Kudrolli TA, Yegnasubramanian S, Luo J, Rodriguez R, Mendell JT, Lupold SE. miR-21: An androgen receptor-regulated microRNA that promotes hormone-dependent and hormone-independent prostate cancer growth. Cancer Res. 2009;69(18):7165–9. DOI: 10.1158/0008-5472.CAN-09-1448.
  279. 279. Li T, Li D, Sha J, Sun P, Huang Y. MicroRNA-21 directly targets MARCKS and promotes apoptosis resistance and invasion in prostate cancer cells. Biochem Biophys Res Commun. 2009;383(3):280–5. DOI: 10.1016/j.bbrc.2009.03.077.
  280. 280. Hu H, Li Y, Gu J, Zhu X, Dong D, Yao J, Lin C, Fei J. Antisense oligonucleotide against miR-21 inhibits migration and induces apoptosis in leukemic K562 cells. Leuk Lymphoma. 2010;51(4):694–701. DOI: 10.3109/10428191003596835.
  281. 281. Gu J, Zhu X, Li Y, Dong D, Yao J, Lin C, Huang K, Hu H, Fei J. miRNA-21 regulates arsenic-induced anti-leukemia activity in myelogenous cell lines. Med Oncol. 2011;28(1):211–8. DOI: 10.1007/s12032-009-9413-7.
  282. 282. Galardi S, Mercatelli N, Giorda E, Massalini S, Frajese GV, Ciafrè SA, Farace MG. miR-221 and miR-222 expression affects the proliferation potential of human prostate carcinoma cell lines by targeting p27Kip1. J Biol Chem. 2007;282(32):23716–24. DOI: 10.1074/jbc.M701805200.
  283. 283. Stinson S, Lackner MR, Adai AT, Yu N, Kim HJ, O'Brien C, Spoerke J, Jhunjhunwala S, Boyd Z, Januario T, Newman RJ, Yue P, Bourgon R, Modrusan Z, Stern HM, Warming S, de Sauvage FJ, Amler L, Yeh RF, Dornan D. miR-221/222 targeting of trichorhinophalangeal 1 (TRPS1) promotes epithelial-to-mesenchymal transition in breast cancer. Sci Signal. 2011;4(186):pt5. DOI: 10.1126/scisignal.2002258.
  284. 284. Miller TE, Ghoshal K, Ramaswamy B, Roy S, Datta J, Shapiro CL, Jacob S, Majumder S. MicroRNA-221/222 confers tamoxifen resistance in breast cancer by targeting p27Kip1. J Biol Chem. 2008;283(44):29897–903. DOI: 10.1074/jbc.M804612200.
  285. 285. Zhang C, Zhang J, Hao J, Shi Z, Wang Y, Han L, Yu S, You Y, Jiang T, Wang J, Liu M, Pu P, Kang C. High level of miR-221/222 confers increased cell invasion and poor prognosis in glioma. J Transl Med. 2012;10:119. DOI: 10.1186/1479-5876-10-119.
  286. 286. Hao J, Zhang C, Zhang A, Wang K, Jia Z, Wang G, Han L, Kang C, Pu P. miR-221/222 is the regulator of Cx43 expression in human glioblastoma cells. Oncol Rep. 2012;27(5):1504–10. DOI: 10.3892/or.2012.1652.
  287. 287. Visone R, Russo L, Pallante P, De Martino I, Ferraro A, Leone V, Borbone E, Petrocca F, Alder H, Croce CM, Fusco A. MicroRNAs (miR)-221 and miR-222, both overexpressed in human thyroid papillary carcinomas, regulate p27Kip1 protein levels and cell cycle. Endocr Relat Cancer. 2007;14(3):791–8. DOI: 10.1677/ERC-07-0129.
  288. 288. Garofalo M, Di Leva G, Romano G, Nuovo G, Suh SS, Ngankeu A, Taccioli C, Pichiorri F, Alder H, Secchiero P, Gasparini P, Gonelli A, Costinean S, Acunzo M, Condorelli G, Croce CM. miR-221&222 regulate TRAIL resistance and enhance tumorigenicity through PTEN and TIMP3 downregulation. Cancer Cell. 2009;16(6):498–509. DOI: 10.1016/j.ccr.2009.10.014.
  289. 289. Sandhu SK, Volinia S, Costinean S, Galasso M, Neinast R, Santhanam R, Parthun MR, Perrotti D, Marcucci G, Garzon R, Croce CM. miR-155 targets histone deacetylase 4 (HDAC4) and impairs transcriptional activity of B-cell lymphoma 6 (BCL6) in the Eμ-miR-155 transgenic mouse model. Proc Natl Acad Sci U S A. 2012;109(49):20047–52. DOI: 10.1073/pnas.1213764109.
  290. 290. Huang X, Shen Y, Liu M, Bi C, Jiang C, Iqbal J, McKeithan TW, Chan WC, Ding SJ, Fu K. Quantitative proteomics reveals that miR-155 regulates the PI3K-AKT pathway in diffuse large B-cell lymphoma. Am J Pathol. 2012;181(1):26–33. DOI: 10.1016/j.ajpath.2012.03.013.
  291. 291. Willimott S, Wagner SD. miR-125b and miR-155 contribute to BCL2 repression and proliferation in response to CD40 ligand (CD154) in human leukemic B-cells. J Biol Chem. 2012 Jan 20;287(4):2608–17. doi: 10.1074/jbc.M111.285718.
  292. 292. Rai D, Kim SW, McKeller MR, Dahia PL, Aguiar RC. Targeting of SMAD5 links microRNA-155 to the TGF-beta pathway and lymphomagenesis. Proc Natl Acad Sci U S A. 2010;107(7):3111–6. DOI: 10.1073/pnas.0910667107.
  293. 293. Kong W, He L, Coppola M, Guo J, Esposito NN, Coppola D, Cheng JQ. MicroRNA-155 regulates cell survival, growth, and chemosensitivity by targeting FOXO3a in breast cancer. J Biol Chem. 2010;285(23):17869–79. DOI: 10.1074/jbc.M110.101055.
  294. 294. Jiang S, Zhang HW, Lu MH, He XH, Li Y, Gu H, Liu MF, Wang ED. MicroRNA-155 functions as an OncomiR in breast cancer by targeting the suppressor of cytokine signaling 1 gene. Cancer Res. 2010;70(8):3119–27. DOI: 10.1158/0008-5472.CAN-09-4250.
  295. 295. Inomata M, Tagawa H, Guo YM, Kameoka Y, Takahashi N, Sawada K. MicroRNA-17-92 down-regulates expression of distinct targets in different B-cell lymphoma subtypes. Blood. 2009;113(2):396–402. DOI: 10.1182/blood-2008-07-163907.
  296. 296. Venturini L, Battmer K, Castoldi M, Schultheis B, Hochhaus A, Muckenthaler MU, Ganser A, Eder M, Scherr M. Expression of the miR-17-92 polycistron in chronic myeloid leukemia (CML) CD34+ cells. Blood. 2007;109(10):4399–405. DOI: http://dx.doi.org/10.1182/blood-2006-09-045104.
  297. 297. He L, Thomson JM, Hemann MT, Hernando-Monge E, Mu D, Goodson S, Powers S, Cordon-Cardo C, Lowe SW, Hannon GJ, Hammond SM. A microrna polycistron as a potential human oncogene. Nature. 2005;435:828–833. DOI: 10.1038/nature03552.
  298. 298. Mu P, Han YC, Betel D, Yao E, Squatrito M, Ogrodowski P, de Stanchina E, D'Andrea A, Sander C, Ventura A. Genetic dissection of the mir-17B92 cluster of micrornas in myc-induced b-cell lymphomas. Genes Dev. 2009;23:2806–2811. DOI: 10.1101/gad.1872909.
  299. 299. Conkrite K, Sundby M, Mukai S, Thomson JM, Mu D, Hammond SM, MacPherson D. Mir-17B92 cooperates with rb pathway mutations to promote retinoblastoma. Genes Dev. 2011;25:1734–1745. DOI: 10.1101/gad.17027411.
  300. 300. Dews M, Homayouni A, Yu D, Murphy D, Sevignani C, Wentzel E, Furth EE, Lee WM, Enders GH, Mendell JT, Thomas-Tikhonenko A. Augmentation of tumor angiogenesis by a myc-activated microrna cluster. Nat Genet. 2006;38:1060–1065. DOI: 10.1038/ng1855.
  301. 301. Leivonen SK, Mäkelä R, Ostling P, Kohonen P, Haapa-Paananen S, Kleivi K, Enerly E, Aakula A, Hellström K, Sahlberg N, Kristensen VN, Børresen-Dale AL, Saviranta P, Perälä M, Kallioniemi O. Protein lysate microarray analysis to identify micrornas regulating estrogen receptor signaling in breast cancer cell lines. Oncogene. 2009;28:3926–3936. DOI: 10.1038/onc.2009.241.
  302. 302. Taguchi A, Yanagisawa K, Tanaka M, Cao K, Matsuyama Y, Goto H, Takahashi T. Identification of hypoxia-inducible factor-1 alpha as a novel target for miR-17-92 microRNA cluster. Cancer Res. 2008;68(14):5540–5. DOI: 10.1158/0008-5472.CAN-07-6460.
  303. 303. Ebi H, Sato T, Sugito N, Hosono Y, Yatabe Y, Matsuyama Y, Yamaguchi T, Osada H, Suzuki M, Takahashi T. Counterbalance between RB inactivation and miR-17-92 overexpression in reactive oxygen species and DNA damage induction in lung cancers. Oncogene. 2009;28(38):3371–9. DOI: 10.1038/onc.2009.201.
  304. 304. Osada H, Takahashi T. Let-7 and mir-17-92: Small-sized major players in lung cancer development. Cancer Sci. 2011;102:9–17. DOI: 10.1111/j.1349-7006.2010.01707.x.
  305. 305. Smith AL, Iwanaga R, Drasin DJ, Micalizzi DS, Vartuli RL, Tan AC, Ford HL. The miR-106b-25 cluster targets Smad7, activates TGF-β signaling, and induces EMT and tumor initiating cell characteristics downstream of Six1 in human breast cancer. Oncogene. 2012;31(50):5162–71. DOI: 10.1038/onc.2012.11.
  306. 306. Petrocca F, Visone R, Onelli MR, Shah MH, Nicoloso MS, de Martino I, Iliopoulos D, Pilozzi E, Liu CG, Negrini M, Cavazzini L, Volinia S, Alder H, Ruco LP, Baldassarre G, Croce CM, Vecchione A.E2F1-regulated microRNAs impair TGFbeta-dependent cell-cycle arrest and apoptosis in gastric cancer. Cancer Cell. 2008;13(3):272–86. DOI: 10.1016/j.ccr.2008.02.013.
  307. 307. He Y, Cui Y, Wang W, Gu J, Guo S, Ma K, Luo X. Hypomethylation of the hsa-miR-191 locus causes high expression of hsa-mir-191 and promotes the epithelial-to-mesenchymal transition in hepatocellular carcinoma. Neoplasia. 2011;13(9):841–53. DOI: 10.1593/neo.11698.
  308. 308. Elyakim E, Sitbon E, Faerman A, Tabak S, Montia E, Belanis L, Dov A, Marcusson EG, Bennett CF, Chajut A, Cohen D, Yerushalmi N.hsa-miR-191 is a candidate oncogene target for hepatocellular carcinoma therapy. Cancer Res. 2010;70(20):8077–87. DOI: 10.1158/0008-5472.CAN-10-1313.
  309. 309. Shi X, Su S, Long J, Mei B, Chen Y. MicroRNA-191 targets N-deacetylase/N-sulfotransferase 1 and promotes cell growth in human gastric carcinoma cell line MGC803. Acta Biochim Biophys Sin (Shanghai). 2011;43(11):849-56. DOI: 10.1093/abbs/gmr084.
  310. 310. Dongfang Wang, Jin Gu, Ting Wang, Zijian Ding. OncomiRDB: A database for the experimentally verified oncogenic and tumor-suppressive microRNAs. Bioinformatics. 2014:30(15):2237–2238. DOI: 10.1093/bioinformatics/btu155.

Written By

Anca Botezatu, Iulia V. Iancu, Oana Popa, Adriana Plesa, Dana Manda, Irina Huica, Suzana Vladoiu, Gabriela Anton and Corin Badiu

Submitted: 03 December 2014 Reviewed: 22 July 2015 Published: 02 March 2016