Open access peer-reviewed chapter

The Involvement of Epigenetic Mechanisms in HPV‐Induced Cervical Cancer

Written By

Adriana Plesa, Iulia V. Iancu, Anca Botezatu, Irina Huica, Mihai Stoian and Gabriela Anton

Submitted: 29 October 2015 Reviewed: 03 March 2016 Published: 13 July 2016

DOI: 10.5772/62833

From the Edited Volume

Human Papillomavirus - Research in a Global Perspective

Edited by Rajamanickam Rajkumar

Chapter metrics overview

1,840 Chapter Downloads

View Full Metrics

Abstract

High‐risk human papillomavirus (HPV) genotypes infection associates with cervical dysplasia and carcinogenesis. hr‐HPV transforming potential is based on E6 and E7 viral oncoproteins actions on cellular proteins. A persistent infection with hr‐HPV leads to progression from precursor lesions to invasive cervical cancer inducing changes in host genome and epigenome. Pathogenesis and development of cancer associated with both genetic and epigenetic defects alter transcriptional program. An important role for malignant transformation in HPV‐induced cervical cancer is played by epigenetic changes that occur in both viral and host genome. Furthermore, there are observations demonstrating that oncogenic viruses, once they integrated into host genome, become susceptible to epigenetic alterations made by host machinery. Epigenetic regulation of viral gene expression is an important factor in HPV‐associated disease. Gene expression control is complex and involves epigenetic changes: DNA methylation, histone modification, and non‐coding RNAs activity. Persistent infection with hr‐HPV can cause viral DNA integration into host genome attracting defense mechanisms such as methylation machinery. In this chapter, we aim to review HPV infection role in chromatin modification/remodeling and the impact of HPV infection on non‐coding RNAs in cervix oncogenesis. The reversible nature of epigenetic alterations provides new opportunities in the development of therapeutic agents targeting epigenetic modification in oncogenesis.

Keywords

  • HPV
  • epigenetic regulation
  • DNA methylation
  • histone modification
  • ncRNAs

1. Introduction

Cervical cancer accounts for almost 12% of all cancers in women, representing the second most frequent gynecological malignancy in the world, human papillomavirus (HPV) being considered as etiologic agent of this malignancy [1, 2]. HPVs exhibit tropism for skin or mucosal epithelium where they cause warts, benign lesions that usually regress. HPV prevalence is a combination of incidental and persistent infections that have accumulated over time, due to lack of clearance. Infection with a high‐risk HPV (hr‐HPV) type is considered necessary for the development of cervical cancer, but by itself, it is not sufficient to cause cancer [3, 4].

The persistent infection with hr‐HPVs that have tropism for mucosal epithelia has been defined as the cause of more than 98% of cervical carcinomas as well as a high proportion of other cancers of the anogenital region (vulvar, vaginal, and penial) and oropharyngeal region [5]. It is known that persistent infection with hr‐HPV genotypes is necessary but not sufficient for the development of high‐grade cervical lesions and progression to malignancy. Persistent infection is characterized by continuous detection of the virus or its intermittent detection, due to latency, although the mechanism of latency has not yet been established but it is clear that the differences between active and latent cervical infection are qualitative and/or quantitative. The high prevalence of HPV infection in precancerous and cancerous cervical lesions confirms its oncogenic potential, different genotypes seem to be responsible for invasive cancer development. Approximately, 40 HPV genotypes were found to be associated with anogenital infections and are generally classified according to their oncogenic potential into low‐, high‐, and intermediate‐risk types. High‐risk or oncogenic types such as HPV16, 18, 31, 33, 35, 39, 45, 51, 52, 56, 58, 59, 68, 73, and 82 are considered so due to their presence in high‐grade squamous intraepithelial lesions (HSIL) or cervical cancer [6]. hr‐HPV genotypes 16 and 18 are causing more than 70% cervical cancer cases and most of the anogenital cancer as well as oropharyngeal tumors in men and women.

The molecular mechanism of cellular transformation induction involves epigenetic abnormalities along with genetic alterations. HPV disrupts normal cell‐cycle control, promoting uncontrolled cell division, and the accumulation of genetic damage. The transforming properties of hr‐HPV E6 and E7 oncoproteins are in interaction with many host cell proteins resulting in the maintenance and the reentering into cell cycle, permitting the virus to replicate, as it is dependent on the host cell DNA replication machinery.

Both E6 and E7 oncoproteins are able to interfere with key cellular processes (cell cycle, senescence, apoptosis and telomere shortening, differentiation). Furthermore, because of the frequent integration of the hr‐HPV genome into a host cell chromosome, those two proteins are the only viral proteins known to be consistently expressed in HPV‐associated cancers [7, 8]. Persistent infection with hr‐HPV genotypes determines progression from precursor lesions to invasive cervical cancer by inducing changes in the host genome and epigenome. Transcriptional modification program through genetic and epigenetic alterations leads to cancer development. Gene expression control is complex and involves epigenetic changes.

hr‐E6 protein is one of the most studied HPV proteins due to its many functions and is found to be interacting with many host cell proteins. Although E6 protein leads to p53 protein loss, an important element of cell transformation [9], many studies have identified a number of additional cellular targets that may play an important role. HPV E6 interferes with different apoptosis pathways by additional interactions with key mediators (TNFR‐1, FADD, CASP8, BAK, DFF40, GADD34/PP1, and TIP60) [1015]. E6 protein is found to interact with proteins involved in cell cycle, cell-cell contact and polarity (MUPPI, E6BP, MAGI 1/3, DLG, PAT, paxillin, interacts with many proteins directly involved in DNA repair such as BRCA1, XRCC1, and MGMT [1624] and targets other cell proteins involved in chromosomal and DNA stability for instance NFX1, hTERT, MCM7 [25, 26]. The E6 protein appears to have role in immune evasion as it interacts with tyk‐2 and IRF‐3 proteins both of which are involved in interferon signaling [27, 28].

E7 protein key activity is to overcome tumor suppressor block controlled by the pRb family proteins (RB, p107, p130) through disruption of pRb–E2F complexes thereby initiating the E2F mediated transcription [29]. E7‐pRB complex leads to functional inactivation and disruption of cell‐cycle progression in S phase. Another E7 hr‐HPV function is as cell‐cycle regulator, in doing so, the oncoprotein binds to p21/p27 and subsequent inactivates the CDK inhibitors [29, 30] and also to cyclins A, E in order to regulate cell cycle through pRb, p107 binding [31, 32]. On the other hand, E7 is known to be involved in transcription modulation by targeting host cell proteins AP1, TBP, MPP2, E2F6, and Skip [3337]. In addition, E7 hr‐HPV protein binds to histone deacetylases (HDACs) in a pRb‐independent manner, which promotes cell growth. The E7 protein can also associate, directly or indirectly, with histone acetyl transferases (HATs) (p300, pCAF, and SRC1) and abrogates SRC1 associated HAT activity [38].

Following persistent infections with hr‐HPVs, E6, and E7 oncoproteins acts on the DNA and cause epigenetic changes. The cooperation between genetic and epigenetic alterations leads to the malignant phenotype and cancer progression. In contradiction to genetic alterations, the epigenetic changes are reversible, making them therapeutic targets in various conditions, and do not affect DNA sequence of the genes, but determine the gene expression regulation acting on the genome. It is well supported that cancers are epigenetically deregulated. Disruption of epigenetic processes determines altered gene function leading to imprinting disorders, developmental abnormalities and cancer. Epigenetic regulation of viral gene expression is an important factor in HPV associated diseases, due to processes that arise independently of changes in the underlying DNA sequence. Gene expression control is complex and involves epigenetic changes such as DNA methylation [39], histone modifications, and chromatin‐remodeling proteins [40] and DNA silencing by non‐coding RNAs (ncRNAs) [41].

Taking into account that molecular mechanism induction of cellular transformation involves epigenetic abnormalities along with genetic alterations, in this chapter, we aim to review: (1) DNA methylation and cervical cancer; (2) the role of HPV infection in chromatin modification/remodeling; (3) the impact of HPV infection on ncRNA in cervix oncogenesis; (4) epigenetic changes involved in viral gene expression; (5) potential epigenetic biomarkers in cervical cancer.

Advertisement

2. DNA methylation and cervical cancer

The most studied epigenetic mechanism is DNA methylation. DNA methylation is a general term for processes of DNA bases (adenine, cytosine, and guanine) change by addition of a methyl group. Methylation of DNA bases can be achieved either under physiological conditions after a specific endogenous enzyme reaction by transferring the methyl group from a donor (biological methylation), or non‐physiological conditions through the action of chemical compounds: alkylating agents. DNA methylation plays an important role in various cellular processes including gene expression, silencing of transposable elements, as well as in the defense mechanism against viral infection.

In several types of cancer, many genes have been reported to be hypermethylated. DNA hypermethylation results in blocking of affected gene transcription, causing silencing them. In cancer, hypermethylation is considered one of the most important mechanisms for tumor suppressor gene silencing, responsible for the control of the normal cellular differentiation and/or inhibition of cell growth. The main chemical DNA modification is methylation of cytosine, commonly found in areas with CpG dinucleotides islands. Almost 60% of promoters of genes encoding proteins in the human genome contain CpG islands, and the majority are methylated in varying degrees, depending on the tissue [42].

Cytosine methylation is a stable inherited and reversible hallmark and is generally associated with transcriptional repression. The methylation inhibits transcription factors that bind to recognized DNA sequences by the recruitment of methyl cytosine binding protein (MECP and MBD) with corepressor molecules. The way that 5‐methylcytosine (5‐mC) repress transcription at the promoter level is by the recruitment of methyl binding proteins (MeCP2, MBD1, MBD2, MBD3, MBD4), which subsequently interacts with another protein to repress DNA transcription as well as HDAC and other chromatin remodeling enzymes [43].

DNA methylation is controlled by DNA methyltransferase (DNMT), which catalyses the transfer of the methyl group from S‐adenosyl methionine donor (SAM). Three active catalytic DNA methyltransferases were identified as follows: DNMT1, DNMT3A, and DNMT3B.

First tumor suppressor gene identified as hypermethylated was pRB, and then was followed by multiple publications describing similar phenomena for a variety of tumor suppressor genes such as p16, MLH1, VHL, and E‐cadherin [44, 45]. It remains controversial whether tumor suppressor gene hypermethylation is a cause or a consequence of silencing them. DNA methylation is reversible and various chemical compounds are known that can reactivate gene expression [46].

On the other hand, DNA hypermethylation may be a secondary process, due to changes in chromatin role in maintaining the status repression of gene expression. More evidence of this hypothesis, resulting from experiments showing that when DNA methyltransferase expression was blocked in vitro, histone H3K9 methylation determined silencing of p16 gene in the absence of promoter DNA methylation [47, 48]. It was shown in cervical carcinoma that tumor suppressor genes are silent or abnormal diminished expressed due promoter hypermethylation (Table 1).

Gene Methylation percentage Activity
DcR1/DcR2 100% Apoptosis [5053]
hTERT 57%
p16 8–42% Cell cycle control [54, 55]
p73 39%
PTEN 58% WNT pathway [56, 57]
E‐cadherin 28–80.5%
APC 11–94%
MGMT 5–81% DNA repair [58]
FANCF 30% FA‐BRAC pathway [59, 60]
BRACI 6.1%
hMLH1 5% DNA mismatch repair [61]
RASSF1A 0–45% Ras negative effector [62]
DAPK 45–100% Cell death/metastasis [63]
TSCLC1 58–65% Tumor suppressor [64]
FHIT 11–88% Cell death/repair [65, 66]
HIC1 18–45% Transcriptional factor [55, 67]
RARβ 33–66% Cell differentiation [68, 69]
TIMP2/TIMP3 47% Metalloprotease inhibitors [70, 71]
Calveolina1 6% Calveole membrane [72, 73]
ERα 25% Estrogen receptor alpha [74]
miR124 59% Tumor suppressor [75]
miR‐34b 48%
miR‐203 57%

Table 1.

Hypermethylated tumor suppressor genes in invasive[R1] cervical cancer (Adapted with permission from Dueñas‐González et al. [49]).

On the other hand, microRNA genes undergo methylation‐mediated transcriptional repression in cervical cancer miR‐149, miR‐375, miR‐432, miR‐1286, miR‐641, miR‐1290, miR‐1287, and miR‐95 [7577].

CADM1, MAL, PAX1, and ADCYAP1 genes promoter hypermethylation were found to be involved in HPV‐mediated transformation and may be significantly associated with the development of cervical cancer [78].

It was shown that hTERT, mir124‐2, and PRDM14 were the first genes that became methylated during experimental immortalization. Following immortalization, ROBO3 methylation and CYGB was methylated, followed by CADM1, FAM19A4, MAL, PHACTR3, and SFRP2 [79].

Advertisement

3. The role of HPV infection in chromatin modification/remodeling

Nucleosomes are the basic repetitive units of chromatin and are intended to pack huge eukaryotic genome in the nucleus (mammalian cells contain approximately 2 m linear DNA packed into a nucleus‐sized 10 mm diameter). Nucleosomes are further compacted to form chromosomes. These structures confer DNA compaction, but also create a base for the gene expression regulation. Nucleosome core particle is approximately 147 base pairs wrapped around a histone octamer made up of two copies of the histones H2A, H2B, H3, and H4. Histone H1 (linker histone) and its isoforms are involved in chromatin compaction and underlying nucleosomes condensation. Compaction of chromatin in the cell nucleus is necessary but is not fully understood. In nucleosomes compacting DNA linker—10 nm has an important role. A chain of nucleosomes can be arranged in 30 nm chromatin fibers whose formation is dependent of histone H1. A 30‐nm chromatin fiber is arranged as a loop around a central protein scaffold to generate the active form of the transcription‐euchromatin. Compacting the fibers can lead to transcriptionally inactive form—heterochromatin [80].

3.1. Histone modifications

Covalent modifications of histones (epigenetic changes) are regulatory elements important in many biological processes. They affect chromatin interactions by structural changes in the histones or by modifying the electrostatic interactions and non‐histone proteins recruit [81]. Histones can undergo a variety of post‐translational modifications at the N‐terminus, which is represented by acetylation, methylation, phosphorylation, sumoylation, ADP–ribosylation, and ubiquitination. They can alter the DNA histone interaction, with a major impact on chromatin structure. Some covalent modifications of histones are involved in transcription and are associated with DNA repair process. On the other hand, the phosphorylation of histone H2AX appears to be unique modification in DNA repair. Histone modifications may influence both among themselves and in interaction with methylated DNA, and the presence of numerous changes in the combined space‐time context creates a program of genome expression profile specific to each cell to keep identity.

3.2. Histone acetylation

Histone acetylation is a modification of the lysine aminoacid that neutralizes the positive charges occurring at specific targets of nucleosome core. It has been speculated that histone acetylation can alter DNA interaction, helping to create more open chromatin architecture. Acetylation of lysine residues is catalyzed by HATs by transfer of an acetyl group from acetyl‐coenzyme A to ε nitrogen of the lysine aminoacid. With few exceptions, these changes tend to create a relaxed form of chromatin, open for transcription, while deacetylation performed by HDACs is associated with transcriptional repression. Many transcriptional activators have been identified possessing intrinsic activity acetyl transferase: Gcn5/PCAF, CBP/p300, and SRC‐1. Similar to these co‐activators that exhibit HAT activity, there are co‐repressors with HDAC activity, such as mSin3a, NcoR/SMRT and NURD/Mi‐2 [82]. Rpd3 complex is an exception, being a complex with HDAC activity, associated with the active form of RNA polymerase II. Through this association is Rpd3 complex transcriptional repressor of initiation [83].

3.3. Histone methylation

Methylation of H3 and H4 histone at lysine and arginine residues arises in mono‐, di‐, or tri‐methylated form and is conducted using a specific histone methyltransferase (HMT) that acts at the level of these residues. HMTs can catalyze the addition of up to three methyl groups on the ε nitrogen of lysine [84]. Co‐activators, such as arginine methyltransferase (CARM1) and protein arginine methyltransferase (PRMT1), are essential for histone H3 and H4 arginine methylation [85]. Most of HMTs contain catalytic domains, named SET conserved domains named after D. melanogaster Su (var)3‐9 Enhancer of Z‐(E(z)) and trithorax (TRX), although there are some exceptions of HMTs without SET domain [8588]. H3K4 and H3K36 methylation is carried out by several methyltransferases. This redundancy makes the study of histone methylation more complex. Mammalian H3K79 methyltransferases include Suv39h1,2, G9a, and ESET [89]. EZH2 catalyses methylation of histone H3K27 and PR‐Set2 (also known SET 8) with Suv4‐20h1,2 catalyzes histone H4K20 methylation [90]. Methylation of H3K9, H3K27, and H4K20 is generally linked to the formation of heterochromatin in the presence of a transcriptional repressor HP1, while methylation of H3K4 and H3K36 is associated with transcriptionally active regions [91]. Methylation of histones lysine is reversible, being made by two enzymes families: aminooxidases (LSD1) and hydroxylases (JmjC family members), which may also demethylate trimethylated lysine [92, 93]. It was identified a series of proteins that bind to post‐translationally modified histones. For example, methylated lysine residues can bind to protein with conserved regions, like plant‐homeodomain (PHD) and chromodomain (CHD), whereas acetylated lysine residues bind to proteins with bromodomain [94]. These recruitment and recognition events can serve as a regulatory mechanism for other mechanisms that lead to other modifications of the histones. Two main complexes were identified, accompanying epigenetic changes, and containing members of trithorax (TrxG) and Polycomb (PCG) group. Some components of complex PCG and TrxG exhibit histone–methyltransferase activity, while other members interpret histones modifications playing a central role in gene regulation, coordinating such DNA availability for development and to establish cell faith. This are accomplished by pausing the state of balance between silent transcriptionally heterochromatin (PCG) and competent transcriptionally euchromatin (TrxG) [95].

3.4. Histone phosphorylation

Histones are phosphorylated at specific sites (serine residues) during cell division [96]. Phosphorylation process requires certain kinases. All four histone suffer phosphorylation; their biological meanings depend on the context. For example, histone H4S1 is evolutionary conserved role in chromatin compaction during the late stages of gametogenesis [97]. Phosphorylation of histone H2A (human—Ser14 in yeast—Ser10) is correlated with meiotic chromosome condensation, but disappears during meiotic division [98]. Chromatin condensation in apoptosis has been linked to the phosphorylation of histone H2B, in both humans and yeast. Histone phosphorylation seems to have a role in transcription. It was shown that phosphorylation of histone H3 determines the competence of transcriptional response for JUN and FOS genes immediately. These changes occur due to activation of Ras‐MAP kinase pathway by growth factors.

3.5. Histone ubiquitination

Ubiquitin is a polypeptide that is attached covalently to other proteins as a result of a steps series involving activation and conjugation enzymes of ubiquitin E1 (ubiquitin activating enzyme), E2 (ubiquitin conjugating enzyme) and ubiquitin ligase (E3) [99]. Polyubiquitination or more ubiquitin molecules addition to a protein is a classic signal for degradation via the proteasome. Histone H2A was first protein identified to serve as ubiquitin substrate [100]. Histone ubiquitination may be reversible using deubiquitinases. Like histone acetylation, ubiquitination is important in regulating gene expression. Highly ubiquitinated histones H2A and H2B have been associated with transcriptionally active sequences. Removing the ubiquitin residues on histone H2A leads to transcriptional repression [101].

3.6. ADP‐ribosylation

ADP‐ribosylation is a post‐translational modification of proteins, including histones, which involves the addition of one or more residues of ADP and ribose. Mono or poly ADP-ribosylation is mediated by MARTs (Mono-ADP-ribosyltransferases) or PARPs (poly-ADP-ribose polymerases) enzymes [102]. ADP‐ribosylation of histones is carried out in a single‐site H2BE2ar1 [103]. Recently it was demonstrated the role of PARP‐1 in transcriptional activity, but only if that DNA repair process was induced [104].

3.7. Crosstalk between DNA methylation and histone modifications

Several studies have shown that the relationship between DNA methylation and histone modifications is mediated by a group of proteins whose function is the binding to methyl groups from DNA, including proteins which bind to CpG methylated islands (MeCP2), proteins with binding domain to CpG methylated islands (methyl‐CpG binding domain protein 1, MBD1), and Kaiso protein‐also known as ZBTB 33 (Zinc finger and BTB domain containing protein 33). These proteins are localized to the methylated promoters and recruit a protein complex containing HDACs and HMTs [105107]. These studies suggest that DNA methylation may induce structural changes to the chromatin by altering the histone modifications. It is also known that DNA methylation inhibits methylation of histone H3K4me [108, 109]. In embryonic stem cells gene, Oct3/4 is inactivated after the fate is determined to a particular cell type. The silencing process is realized by recruiting a co‐repressor complex consisting of G9a methyltransferase and with HDAC enzyme activity. DNMT3A and DNMT3B DNA methyltransferases are subsequently recruited, catalysing the de novo DNA methylation at the gene promoter level [110]. Interaction between G9a protein and DNA methyltransferases (DNMT3A and DNMT3B) depends on the ankyrin motif of G9a protein [111]. In exchange, the SET domain responsible for methyl transferase activity of the G9a protein does not interact with DNA methyltransferases [112, 113]. These data suggest that DNA methylation at the promoter level depends on recruiting especially G9a protein and less of its methyltransferase activity. Interaction between histone H3K9 methylation and DNA methylation represents a model in which these two changes determine a strong silencing loop or bidirectional interference.

Recently, it was established with the aid ChIP and bioinformatics a link between methylation‐mediated by PCG on histone H3K27 and de novo DNA methylation in cancers, which claims that the signal required by PRC2 during development predisposing certain genes to de novo methylation later [113115]. In tumor cells were observed interactions between DNA methylation and histone H3K9 methylated, thereby contributing to a stable silencing mechanism. PCG and EZH2 proteins are members of Polycomb repressor complex 2 (PRC2) which has methyltransferase activity with substrate specificity for histone H3K27. Histone H3K27me3 serves as specific binding signal to a chromodomain of another Polycomb repressor complex (PRC1). PRC1 blocks transcriptional factors recruitment, therefore the presence of PRC1 stops transcription initiation.

Biochemical studies have also shown that DNA methyltransferase binds EZH2 to certain conditions [115117]. Histone H3K9 and H3K27 methylation presence does not always lead to the de novo DNA methylation. A subset of target genes for complex PCG can be methylated in cancer. Additional factors are required for DNA methylation in genes showing changes in the histones. Recently, it has been shown that the histone H3K27 trimethylation is PCG mediated and is a mechanism that determines tumor suppressor gene silencing in cancer, which is independent of promoter methylation [118, 119]. This lack of dependence between DNA methylation and histone modifications of these studies demonstrated conflicting results of previous studies. It should be noted that most of the genes presenting at their level histone H3K27me3 in prostate cancer do not show islands CpG motifs in the promoter, instead gene targeted by PCG complexes show generally the CpG promoters islands in embryonic stem cells ES [120]. This indicates that the histone H3K27 methylation processes mediated by the PCG complex in ES, normal and tumor cells are different because the tumor cells by removing the functional path of histone H3K27me3 usurps silencing mechanisms. Therefore, it was established the existence of 3 directions involved in silencing machinery associated H3K27 methylated histone mediated by PCG complex. The first relates to de novo repressed genes by methylation of histone H3K27, PCG mediated, and targets certain gene in particular which do not present CpG islands at promoter level. The second direction supports that during oncogenesis an early gene subset became methylated and CpG islands of the promoters are initially marked by PCG complex. This includes also those genes which undergo epigenetic reprogramming and are silent initially by the PCG and then suffer DNA methylation process like an alternated silencing mechanism. This epigenetic silencing switch through DNA methylation reduces epigenetic plasticity, blocking key regulators and contributes to tumourigenesis [121]. Third mechanism supports the fact that DNA methylation and histone H3K27me3 co‐exist at the same promoter and methylation H3K27me3 histone by PCG silencing machinery is dominant. The silencing machinery may contribute to oncogenesis process in various forms, which can constitute in a repressor mechanism from flexible to plastic up to stable inactivation maintained by DNA methylation.

3.8. Chromatin and cancer

The involvement of DNA methylation process and chromatin changes in oncogenesis is indisputable, but separation of the genetic from epigenetic events is artificial. New evidence has shown that primary genetic defects (mutations in the genes coding for the receptors of growth factors, adhesion molecules, the gene that affects the DNA methylation and histone modifications as DNMT, HAT, or HDAC) lead to altered DNA methylation and changes in chromatin pattern. Both the endogenous and exogenous carcinogens do not cause genetic mutations but first epigenetic alterations, which highlights that epigenetic alteration is a step in oncogenesis.

All classical genetic alterations as mutations in tumor suppressor genes and in oncogenes can affect gene transcription (e.g., mutations in Ras gene, HER2 gene amplification). It is not surprising that the control of gene transcription machinery can be directly involved in oncogenesis. Although the complex nature of transcriptional regulation is uncertain, balance disruption of enzymatic activity responsible for maintaining acetylated histones status is expected to occur in cancer. p300/CBP histone acethyltransferase gene exhibits mutations in various cancer type (lung tumors, esophageal, ovarian, and gastric) [122125]. Chromosomal translocations that targeted CBP/p300 gene locus affects transcription by their merger the translocated fragment with genes located in the chromatid area where they were joined (event met in hematological cancers such as acute myeloid leukemia) [126, 127].

Limited data regarding the global profile of histone modifications in oncogenesis can be found, but as a highlight is the overall loss of H4K16 monoacetylation and H4K20 trimethylation [128]. It has also been found that an important role in tumourigenesis is represented by changes of histone from promoters of tumor suppressor genes that determine their silencing. Such modifications are the loss of histone H3K9 acetylation and di/trimethylation of H3K4, H3K9 dimethylation, or trimethylation of H3K27 [129]. Several studies have reported a high level of EZH2 expression, which promotes tumor growth in both in vitro and in vivo, as identified in a number of human cancers such as melanoma, leukemia, prostate, and breast cancer [130, 131]. It has been shown that EZH2 could be a potential biomarker, and its expression was correlated with aberrant H3K27 trimethylation and silencing of tumor‐suppressor genes [119, 132].

Another frequent mechanism in cancer is the inactivation H3K27 demethylase‐UTX/KDM6A (lysine (K)‐specific demethylase 6A). KDM6A gene mutations have been reported in many types of tumors: multiple myeloma, esophageal squamous cell carcinoma, and renal cell carcinoma [133].

3.9. Alteration of histones changes in cervical cancer

Histone modifications and alterations have recently begun to be studied in the cervical cancer. Analysis of histone modifications in the progression of cervical lesions is relatively at the beginning, there are few studies which indicate an association between alterations of histones and cervical cancer development. There are some data supporting that chromatin pattern in cervical samples may help in cervical neoplasia diagnosis, particularly for glandular lesions. The molecular basis of chromatin modifications is not fully determined [134]. E6 and E7 viral oncogenes expression is essential but not sufficient for neoplastic transformation, many studies highlight the important role of epigenetic changes in cervical carcinogenesis. Recently, it has been shown that E6 and E7 oncoproteins interacts with histone‐modulating enzyme, which regulates transcription via the host cell chromatin [84]. A recent report showed that in tumourigenesis, tumor cells lose monoacethylated and trimethylated histone H4 (acetylated Lys16 and trimethylated Lys20) form, that being associated with hypomethylation of repetitive DNA sequences [135]. Huang et al. [136] showed that the expression levels of HDACs were found to be increased in cervical dysplasia and invasive carcinoma.

It was reported that MGMT a DNA repair protein silencing seems to be associated with a reduction in acetylated histones [137]. Moreover, the activation of Wnt signaling pathway may be realized by a transcriptionally repressed Wnt antagonist DICKKOPF‐1 (DKK‐1), by histone deacetylation in HPV‐infected cervical cells [138]. HDAC function is necessary for HIF‐1 (hypoxia inducible factor‐1) activity, and it was found that E7HPV protein can block the interaction of HDACs with HIF‐1α, activating HIF‐1‐dependent transcription for a range of pro‐angiogenic factors [139, 140]. Silencing of proliferation repressor protein osteo‐protegerin (OPG) and retinoic acid receptor β2 (RAR‐β2) was found to occur through histone modification as well as DNA methylation [141, 142].

It has been shown that phosphorylated and acetylated forms of histone H3 in cervical swabs are associated with progression from CIN I to CIN II and CIN III [143]. The balance between HDACs and HATs activity has a key role in regulating gene transcription [144]. This balance must be maintained in normal cells, to prevent an uncontrolled proliferation and cell death. E6 and E7 HPV target numerous cellular proteins to disrupt cell growth and proliferation, including HDACs and HATs. E7 hr‐HPV protein binds to HDACs, this interaction being performed by Mi2β, a member of the nucleosomes remodeling complex and acetylation of histones (NuRD), which possess the ability to modify chromatin structure by both the deacetylation of histones and by the repositioning ATP‐dependent nucleosomes [145]. The interaction the E7‐HDACs is independent of binding to Rb protein and E7 gene mutations abolish its ability to target the HDACs and to transform mouse fibroblasts [84]. E6 hr‐HPV protein shares with other DNA tumorigenic viruses’ ability to target CBP/p300. The interaction involves C‐terminus zinc finger of E6 protein and 1808–1826 residues of CBP; as a result, the p53 transcriptional activity is reduced, independently of p53 protein removal through the proteasome degradation pathway [146]. E7, E6 protein binds to the transcriptional co‐activator p300/CBP, being a crucial step in cellular transformation [147].

Histone methylation is acknowledged to be a dynamically process controlled by two types of enzymes that work together to maintain global histone methylation patterns: HMTs and histone lysine demethylases (KDMs) [42, 95]. Histone methylation can occur at different lysine residues. The interaction between HMTs and KDMs locally adjusts the degree of methylation which results in the activation or repression of gene expression, depending on the specific target lysine residue [95]. Thus, the degree of methylation and the position of methylated lysine have different consequences: overall methylation of H3K9 (histone 3 lysine at position 9), H3K27 (histone 3 lysine at position 27), and H4K20 (histone 4 lysine at position 20) is linked to the heterochromatin formation in the presence of a transcriptional repressor associated with HP1, while the methylation of H3K4 (histone 3 lysine in position 4) and H3K36 (histone 3 lysine in position 4) is associated with transcriptionally active regions [148152]. Generally, methylated H3K4, H3K36, and H3K79 are considered activating marks, whereas methylation of H3K9, H3K27, and H4K20 are often associated with gene silencing [150154].

E6, E7 oncoproteins can associate with enzymes that modulate histone acetylation, and thus, regulate the transcriptional capacity of host cell chromatin [151, 152, 155, 156]. Especially, KDMs expression was found to be deregulated and associated with cancer aggressiveness. KDMs were further proposed as potential tumor biomarkers and could play distinct role in cancer progression acting either as putative oncogene or tumor suppressor based on different transcriptional role (gene activation/repression) [157, 158].

McLaughlin‐Drubin et al. [159] sustain that E7 HPV16 can induce epigenetic and transcriptional alterations by transcriptional induction of the KDM6A and KDM6B histone 3 lysine 27 (H3K27)‐specific demethylases.

KDM5C demethylase role in the pathogenesis of HPV‐induced has been described in the literature. KDM5C is recruited by the E2 viral protein for E6 and E7 oncogenes transcriptional repression through the LCR region of HPV. The results obtained indicate KDM5C as a good marker for severe lesions and SCC [160].

Another recent study showed that KDM4C, KDM5C, KDM6A, and KDM6B genes expression significantly increase in high‐grade lesions (CIN 2+) and SCC presenting a positive correlation with HPV infection. A significantly increased of KDM4C expression levels in SCC samples compared with precancerous lesions propose it as a suitable tumor marker. KDM4C/GASC1/JMJD2C/ is a histone demethylase that is mainly regarded as oncogene due to its role in demethylating heterochromatic H3K9me3/2 [161]. Another good marker for high‐risk lesions and SCC seems to be KDM5C whose expression levels were found increased in CIN2+ lesions and significantly increased in SCC cases [160, 161].

p16 gene expression in normal cells is generally low due to gene silencing by H3K27 trimethylation and PRC complex action. It was observed that the E7 oncogene expression may reduce residues H3K27 required for repression of PRC1 complex, leading to transcriptional activation of histone H3K27, histone demethylases KDM6A, and KDM6B through an unknown mechanism. In response to the stimulation of the RAS/RAF transcriptional activation of KDM6B occurs possible via AP1, leading to the removal of H3K27me3 (histone 3 lysine at position 27 trimethylated) residues and increasing expression of p16INK4a [162]

The literature data suggest an important role as biomarker for p16INK4a tumor‐suppressor gene in HPV‐induced lesions and cervical cancers. The mechanism of induction of p16 expression by the E7 viral oncogene is believed to be achieved by the activation of E2F transcription factor [163]. Later it was observed that from the p16 promoter are missing response elements to E2F and E7 HPV16 mutated variants that are defective in binding/degradation of pRb and E2F transcription are not activated; p16 expression can be induced by the wild‐type and variants [159]. p16INK4 expression is induced by demethylation of H3K27 residues KDM6B mediated, that underpins the induction of senescence by oncogenes (Oncogene induced senescence—OIS), an intrinsic cellular innate tumor suppressor mechanism triggered by oncogenes such as RAS [164]. E7 oncogene causes degradation of pRB, the main mediator of halting cell growth and senescence induced by p16, repealing the mechanism of induction of senescence by oncogenes. The mechanism of inactivation of pRB by E7 can be explained by the necessity to avoid eliminating E7 HPV positive cells targeted by OIS. Such high levels of p16 observed in this study correlated with an increased expression of KDM6B histone demethylase due to E7 oncogene activity on H3K27 modulators.

Advertisement

4. The impact of HPV infection on ncRNA in cervix oncogenesis

In the latest years, thanks to a growing number of studies focusing on high‐throughput next generation sequencing (NGS), large‐scale genome, and genome‐wide transcriptome methods, a new world of RNA molecules: ncRNAs have emerged [165].

More recently, through deep sequencing data obtained by transcriptome projects such as ENCODE (Encyclopedia of DNA Elements Consortium), it has been revealed that around 90% of genomic DNA in eukaryotes is transcribed with just 1–2% of the transcript encoding for proteins, the vast majority being transcribed as ncRNAs [166].

Some of the ncRNAs molecules appear to be important players in genome functioning acting as “regulatory RNAs”. Experimentally data gathered so far sustain the ncRNAs involvement in many biological processes; they seem to have important roles in genes transcriptional and posttranscriptional regulation, RNA splicing, translation and turnover, also in epigenetic modifications [167, 168].

Furthermore, given the regulatory role that these non‐coding molecules possess in normal biological processes, it has been presumed that they might play a significant role also in different types of pathologies. There are accumulating evidence highlighting a major role for these molecules in various diseases where they appear to have aberrant expression and contributes to disease development and progression.

Several studies showed ncRNAs involvement in diseases such as neurodegenerative, cardiovascular, immune diseases and in neoplastic transformation [169]. Regulatory ncRNAs could be classified according to their length in three categories [170]:

  • Small ncRNAs (approximately 18–31 nucleotides) which comprises: small interfering RNAs (siRNAs), microRNAs (miRNAs), PIWI‐interacting RNAs (piRNAs).

  • Medium ncRNAs (31–200 nucleotides): promoter‐associated small RNAs (PASRs), terminal‐associated small RNAs (TASRs), transcription initiation (tiRNAs).

  • Long ncRNAs (lncRNAs) (>200 nucleotides in length) that includes: long intergenic ncRNAs (lincRNAs), pseudogenes, sense and antisense RNA, enhancer RNAs (eRNAs).

miRNAs molecules are approximately 18–24 nucleotides in length, ncRNAs that regulate genes expression in eukaryotic organisms. These RNA molecules are known to be a part of RISC complex (RNA‐induced silencing complex) and are involved in gene silencing by pairing with complementary sequences at 3' UTR (untranslated regions) or coding region of a target messenger RNAs (mRNAs) that leads to mRNA degradation and blocking protein synthesis [171173].

Through this interaction miRNAs molecules play an important role in specific cellular processes including cellular development, proliferation, differentiation, apoptosis, and thereby controlling the expression level of hundreds important genes involved in these processes [174, 175].

Numerous studies have reported miRNAs aberrant expression profiles in different types of cancer. Until recently, the most extensively studied ncRNAs in oncogenesis, miRNAs appear to have a dual nature in neoplastic transformation acting either as tumor suppressors and/or as oncogenes depending on the cellular context [176].

miRNAs known to have oncogenic functions also called “oncomiRs” have frequently been demonstrated to control processes such as cell differentiation, apoptosis, and tumor development through tumor suppressor genes inhibition. Several examples of well‐known oncomiRs linked to malignant transformation are miR‐15, miR‐16 found upregulated in many types of leukemia's and lymphomas [177, 178]; miR‐155 overexpressed in chronic lymphocytic leukemia (CLL), B‐cell lymphoma, anaplastic large cell lymphoma (ALCL), Hodgkin's or Burkitt's lymphoma and in breast tumors [179182]; miRNA‐17‐92 cluster (oncomiR‐1) deregulated in multiple types of cancer: lung (particularly in small‐cell lung cancer and aggressive forms), pancreatic, hepatocellular, colorectal, breast, ovarian, and hematopoietic cancers [183, 184]; meanwhile, miR‐106a has an oncogenic role in pancreatic, colon cancer, and T‐cell lymphoma [185187]; another promising oncomiR is miR‐21 found overexpressed in various cancers including breast cancer, lung cancer, colorectal cancer, hepatocellular carcinoma, glioblastoma [188196].

On the other hand, in cancer, it was shown that some miRNAs are consistently downregulated and act as tumor suppressor with example such as: miR‐15a and miR‐16 cluster that is often deleted or downregulated in tumor cells [197200] or miR‐34 family members that were identified as potential tumor suppressor in many cancers [201203], also miR‐124 was found significantly downregulated in several types of human cancers [203206]; miR‐122 demonstrated to regulate intrahepatic metastasis in hepatocellular carcinoma and thus acting as a tumor suppressor for this pathology [207] and mir‐203 was shown to suppress cell proliferation and migration in various types of cancer [208210].

A malignancy where miRNAs role has been extensively investigated is cervical cancer. There are many reports that emphasize a substantial role for these non‐coding molecules in cervical oncogenesis.

An interesting research direction in miRNAs field is their relationships with viral infections. Various studies support the fact that some cellular miRNAs expression can be regulated by virus infection and these observations are not surprising given the host defense mechanisms against pathogen agents such as bacteria and viruses.

Researchers also have identified several cellular miRNAs whose levels can be modulated by HPV infection, respectively, by viral E6 or E7 oncoprotein of high‐risk genotypes [211].

Studies based on miRNAs expression profiles revealed a differentially pattern of expression in cervical tumors tissue compared with normal tissue, still due to different detection methods and experimental systems use in some cases the observations are contradictory (Table 2).

miRNA Expression pattern miRNA detection method References
‐21 High Microarray, Northern blot, qRT‐PCR [212214]
‐27a High qRT‐PCR [214]
‐34a High qRT‐PCR [214, 215]
‐155 High Microarray, Northern blot, qRT‐PCR [214, 216, 217]
‐146a High Microarray, Northern blot [216]
‐125 High qRT‐PCR [215]
‐196a High qRT‐PCR [214, 218, 219]
‐203 High qRT‐PCR [214, 220, 221]
‐138 High qRT‐PCR [222]
‐7 High qRT‐PCR [221]
‐20a High qRT‐PCR [220, 223]
‐221 High qRT‐PCR [214, 224]
‐200 High qRT‐PCR [225]
‐93 High qRT‐PCR [225]
‐124 Low qRT‐PCR [226, 227]
‐143 Low Microarray, Northern blot, qRT‐PCR [216, 228]
‐145 Low Microarray, Northern blot [216, 229]
‐149 Low [212]
‐195 Low Microarray [230]
‐34a Low qRT‐PCR, Northern blot [231, 232]
‐214 Low Northern blot, qRT‐PCR [233]
‐23b Low Microarray
qRT‐PCR [212, 234]
‐519 Low Northern blot, qRT‐PCR [235]
‐218 Low Northern blot, qRT‐PCR [236]
‐372 Low qRT‐PCR [237]

Table 2.

Examples of miRNAs aberrant expressed in cervical cancer.

Recently based on the observation that some viruses could express their own set of miRNAs, there is an ongoing effort to identifying these miRNAs and to establish their role during viral infection. Reports revealed that miRNAs encoded by viruses target host genes involved cell proliferation, apoptosis, host immunity regulation, in order to maintain their survival and to escape from immune system response.

Over 200 miRNAs encoded by several virus families have been identified to date, many of them being found for herpes viruses and Epstein–Barr virus (EBV) [237]. For instance, it was found that EBV encodes more than 40 miRNAs that presents different expression levels during viral infection and some are involved in maintaining viral latency [238, 239]. From our knowledge to date, there are no reports on the existence of HPV‐encoded microRNAs.

Although researcher's attention in latest years was mainly focus on short ncRNAs molecules and their functions in normal/pathological conditions, at present great efforts are put into investigating the major part of the non‐coding transcriptome namely lncRNAs transcripts.

It has been revealed that certain lncRNAs can control gene expression through a range of different mechanism including transcriptional, splicing, and post‐transcriptional regulation or at epigenetic levels by chromatin remodeling and histone modification regulation [240242].

Even though few lncRNAs have been well characterized, from the knowledge accumulated so far it is clear that they represent significant gene regulators and play critical roles in many cellular and development processes. Therefore, taken into account, the wide functions that lncRNAs hold, it is not surprising that their alterations are associated with an extensive range of disease.

Several studies have reported lncRNAs involvement in cardiovascular diseases, neurological disorders, immune disease, and also in cancer, data indicates a differential lncRNAs expression in many types of malignancy including, breast cancer, colon cancer, prostate cancer, hepatocellular carcinoma, pancreatic cancer, lymphomas [243].

Currently, the expression profile of various ncRNAs has become an important feature of oncogenesis process. There are numerous publications indicating an association between lncRNAs expression and malignant transformation and the number is still rising. Despite the keen interest shown by these molecules for many of them, the functional role in normal/pathological condition is still unclear, additional studies are needed. Recently with the help of the latest NGS techniques, new information is brought to light for better understanding lncRNAs role, mechanisms of action, and also the potential use of them in various cancer therapies.

Among the best well‐characterized lncRNAs are XIST (X inactive specific transcript) a 17‐kb‐long transcript known for its role in dosage compensation involving X chromosome inactivation and H19 transcript 2.5‐kb‐long that plays an important role in imprinting [244, 245].

Experimentally data sustain a potential oncogenic role for H19, an aberrant expression have been identified in a variety of cancers: breast, ovarian, hepatocellular, gastric, lung, colon, esophagus [246251]. It has been shown that H19 oncogenic role is also due to the fact that the transcript acts as a precursor for miARN‐675 leading to pRB gene expression decrease [252].

Another lncRNA having oncogenic potential is MALAT1 (metastasis‐associated lung adenocarcinoma transcript 1) who it was suggested in many studies that can promote cell proliferation apoptosis, invasion, and metastasis. Significantly high levels of MALAT1 expression were detected in lung cancer, prostate cancer, colorectal cancer, hepatocellular carcinomas, gynecologic (endometrial, cervical) cancer, osteosarcoma [253262].

There are identified lncRNAs that appear to have tumor suppressor function in carcinogenesis. For GAS5 (growth arrest specific 5) lncRNA, it was found to play an important role in apoptosis induction; studies have reported a significantly reduced GAS5 levels of expression in breast cancer and prostate cancer [263, 264]. Another report shows that GAS5 low level of expression is associated with poor prognosis in hepatocellular carcinoma [265].

MEG3 (maternally expressed 3) is lncRNAs that presents a reduced expression in several types of cancer. Several experimental evidences demonstrate that MEG3 interacts with p53 tumor suppressor gene and regulates p53 target gene expression, therefore, inhibits tumor cell proliferation and cancer progression. Aberrant levels of MEG3 expression have been identified in glioblastoma, ovarian, colon, cervical, lung cancer [266271].

In literature, there are a relatively small number of experimental data showing an association regarding lncRNAs involvement in cervical carcinogenesis, but due to high interest shown toward these molecules the number is growing fast. The data collected so far on lncRNAs involvement in cervical cancer are presented in Table 3.

LncRNA name Functions Expression pattern References
H19 (imprinted maternally expressed transcript) Located in an imprinted region near the insulin‐like growth factor 2 (IGF2) gene. Potential dual role in oncogenesis (tumor suppressor/oncogene) Upregulated [275]
HOTAIR (HOX antisense intergenic RNA) Located within the Homeobox C (HOXC) gene locus is co‐expressed with HOXC genes. Involved in HOXD genes transcription repression via histones methylation using PCR2 and LSD1 complex Upregulated [276279]
XIST (X inactive specific transcript) Role in X chromosome inactivation Upregulated [280]
MALAT1 (metastasis‐associated lung adenocarcinoma transcript 1) Regulates transcription of genes involved in cancer metastasis cell migration, proliferation and cell cycle regulation Upregulated [281, 282]
ANRIL/CDKN2B‐AS1 (Antisense non‐coding RNA in the INK4 Locus; CDKN2B antisense RNA 1) Located within the CDKN2B‐CDKN2A gene cluster; interacts with PRC1 and PRC2 complexes for epigenetic silencing of genes in this cluster Upregulated [283, 284]
TUSC8 (tumour suppressor candidate 8; XLOC_010588) Unknown Downregulated [285]
BC200/BCYRN1 (brain cytoplasmic RNA 1) Encodes a neural small non‐messenger RNA Upregulated [286]
lncRNA‐EBIC (TMPOP2—thymopoietin pseudogene 2) Unknown Upregulated [287]
GAS5 (growth arrest specific 5) Promotes cellular growth arrest and apoptosis Downregulated [288]
MEG3 (maternally expressed gene 3) Inhibits tumor cell proliferation; a potential tumor suppressor role due to p53 activation Downregulated [269]
lincRNA‐p21 (TP53COR1—tumor protein p53 pathway corepressor 1) Repress genes transcriptionally regulated by p53 Upregulated [283]
lncRNA‐LET (NPTN‐IT1—NPTN intronic transcript 1) Stabilizes nuclear factor 90 protein, promotes hypoxia Downregulated [289]
CCHE1 (CCEPR—cervical carcinoma expressed PCNA regulatory lncRNA)
CCND1 ncRNA (cyclin D1 non‐coding RNA)
Regulates cyclin D1 gene expression Upregulated
Upregulated
[290]
[283]
SncmtRNA (Sense mitochondrial ncRNA)
ASncmtRNA (Antisense mitochondrial ncRNA transcript)
Unknown ASncmtRNA
Downregulated
SncmtRNA‐2 (Upregulated)
[291]

Table 3.

List of lncRNAs expressed in cervical cancer [272274].

Advertisement

5. Epigenetic changes involved in viral gene expression

Methylation status of integrated HPV depends of viral life cycle as well as of neoplastic transformation, this making HPV methylome a potential tool in cancer diagnostic. HPV genome methylation status depends on the viral life cycle and is associated with neoplastic progression.

According to Johanssen and Lambert study, viral genome is subjected to de novo methylation by host DNMTs. Methylation of the viral genome may be a part of a mechanism involved in innate response to pathogens by which the host attempts to suppress viral gene expression. The authors note that in HeLa cells, HPV18 genome chromatin histone modification status correlates with the occupancy of host transcriptional machinery specifically within the LCR [292]. E7 and E6 oncoproteins of hr‐HPVs appear to modulate host epigenetic machinery through their interplay with both DNA methylation enzymes as well as chromatin remodeling enzymes [159].

Mirabello et al. [293] reported in 2013 that they found 3‐region in L1 strongly methylated in cancers and only in a small percentage in CIN I and CIN II lesions. In addition, the authors shown that methylation at certain CpG sites can indicate an evolution toward CIN II+ years before it happens.

Evaluation of cervical samples from HPV positive women, presenting precancerous lesions or invasive ones, showing that the hypomethylation degree in LCR and E6 gene region increase with the increasing of lesion severity. These data convinced the authors to conclude that neoplastic transformation could be suppressed by hypermethylation, while hypomethylation accompanies or leads to progression toward cancer [294].

Using laser microdissection on different layers from samples with HPV‐infected lesions, Vinokurova and von Knebel Doeberitz [295] found dynamic changes in HPV16 LCR methylation in the context of the viral life cycle. A decrease in methylation in the transcriptional enhancer region within the LCR was observed in terminally differentiated epithelial compartment and meanwhile an increase in methylation within the region of the LCR containing the early promoter was noted [295].

Another study highlighted heterogeneity of methylation status among patients, even in samples from the same patient. Methylation frequency was found to be approximately 30% in L1 region, less than in CpG islands around enhancer and promoter of HPV16. In most of the HPV genome sites, hypermethylation is associated better with carcinoma than with dysplastic lesions [296].

On the other hand, a study regarding methylation status in HPV18 immortalized cell lines (HeLa and C4‐1) and in samples from patients, determined a clonally heterogeneity of methylation status in different regions of viral genome. The clinical samples showed partial or total methylation in HPV enhancer region, while in asymptomatic patient's, samples were fully unmethylated. Viral promoter was reported to be methylated in tumor samples and in cervical smears [297].

These studies indicate that methylation status of viral oncogenes in cervical lesions could be the result of transcriptional activity level and not an event that leads toward neoplastic progression. Further studies regarding the influence of DNA methylation on viral life cycle focused on E2 (early gene involved in viral transcription and replication) gene methylation. In vitro studies revealed that HPV16 URR (upstream regulatory region) methylation inhibit E2 protein capacity to bind DNA [298]. By looking at methylation status of E2BS (E2 binding sites) in immortalized epithelial cells from a HPV16 positive patient, Kim et al. found this region to be selectively hypomethylated in highly differentiated cell population, while heavily methylated in basal‐like differentiated cells. The conclusion was that methylation status of E2BS may vary during the viral life cycle, this giving an insight on E2 modulation function during progression of infection [298]. E2BS is more frequently found in a hypermethylated state in cervical lesions with extrachromosomal state of viral genome, while upon integration in the host genome, it was found to be hypomethylated, except the cases in which viral genome integrates as a concatemer, when only a small proportion are found hypomethylated and most of them hypermethylated [292].

All this experimentally observations conclude that HPV genome methylation status could hold a prognostic and progression value for cervical lesions.

Advertisement

6. Potential epigenetic biomarkers in cervical cancer

Cancer epigenome is currently in the researchers spotlight due to the fact that all the epigenetic changes that accompany cervical carcinogenesis can be exploited as biomarkers. Thus, once deciphered, the epigenetic peculiarities of cervical cancer might be used in the development of new alternatives for screening or for the assessment of prognostic [299]. On the other hand, the reversible nature of epigenetic alterations makes them attractive targets for new therapeutic approaches. Some of these discoveries have been proposed as investigation methods or resulted in new treatment approaches and commercial tests [300]. By far, the most studied epigenetic changes are the methylation patterns, especially the methylation markers of the host. Abnormal methylation of promoters of tumor suppressor genes is common in different type of cancers with the prospect of becoming a biomarker in oncology [301]. As the methylation profile of these genes increases with the severity of cervical lesions, their status might be used as potential biomarker for early detection of cervical cancer disease [302]. For a better stratification of cervical cancer and precursor lesions, different specific methylation panels have been suggested [303]. Using DMH (differential methylation hybridization) technique and qPCR, Lai et al. [304] found in scrapings isolated from CIN3 lesions, a higher frequency of methylation for SOX1, NKX6‐1, PAX1, WT1, and LMX1A genes. Siegel et al. [305] demonstrated that aberrant methylation levels of DAPK1, RARB, WIF1, and SLIT2 might increase specificity to identify cervical cancer compared to viral testing alone. Also, the methylation patterns of GGTLA4 (183 bp) and ZNF516 (241 bp) genes were proposed in a patent as biomarkers for diagnosis of premalignant cervical lesions [306], while aberrant methylation of PAX1, PTPRR, SOX1, and ZNF582 promoters were suggested as markers for AC screening [307]. The studies that associate the methylation profile with cervical lesion severity have resulted in a commercial test (GynTect) [308]. GynTect assay is based on methylation‐specific PCR (MS‐PCR) and, if positive, detects specific methylated DNA sites in cervical smears. Manufacturer recommend the test for cervical cancer screening, allowing the triage of women over 30 years who tested positive for HPV. Moreover, GynTect may be performed using residual material from the HPV test. So, this methylation assays might be use as a secondary marker after HPV DNA testing in order to guide the subsequent clinical approach (referral to colposcopy or initiating a certain therapy) [309].

Other authors correlated changes in host DNA methylation with the development of drug resistance. Chen et al. [310] identified both genome‐wide and within individual loci changes in an oxaliplatin‐resistant cervical cancer cell line derived from SiHa cell line. The methylation of Casp8AP2 gene resulted in increased drug resistance in different cells.

Masuda et al. [311] reported that aberrant methylation of Werner (WRN) gene that encode for a DNA helicase, increased the sensitivity to CPT‐11 (an inhibitor of DNA topoisomerase I). Iida et al. [312] reported aberrant hypermethylation of CHFR (checkpoint with forkhead and ring finger) in adenocarcinoma and HeLa cell line (immortalized with HPV18) and correlated this profile with lower sensitivity to anticancer therapy when compared to SSC, proposing this pattern in adenocarcinoma as a potential biomarker for sensitivity to paclitaxel. Therefore, the identification of methylation patterns associated with drug‐resistance might become a valuable tool in cervical treatment with demethylation agents that can revert this epigenetic change.

Regarding the methylation of viral DNA, data are still under debate. While some authors have suggested that it is a defense mechanism of the host cell, others considered it is a way by which the virus contributes to persistent infection. [313]. Other researchers considered that neoplastic transformation may be suppressed by HPV CpG methylation, while demethylation occurs as the cause of or concomitant with neoplastic progression [314]. Several authors proposed HPV16 L1 ORF methylation as a predictive marker for CIN3+ [315] and elevated levels of CpG 6367 L1HPV16 methylation as marker to predict future CIN2+ in women older than 28 years [293]. Also, Mirabello et al. [316] correlated elevated levels of CpG methylation in the L1, L2, E2/E4 with CIN3 or worse and data were confirmed by other papers [317]. Moreover, Wentzensen et al. [317] found differential methylation patterns in CIN3 patients with multiple infections thus suggesting a possible way to identify the causal type of HPV.

Cervical carcinogenesis is accompanied also by altered expression of methyltransferases. For therapeutic purpose, Hamamoto et al. [318] had synthesized double‐stranded molecules that inhibit the expression of SUV39H2 (suppressor of variegation 3–9 homolog 2) gene. This gene encodes a HMT that methylate the H3K9 lysine residue and its hyperexpression correlates with carcinogenesis. The silencing of CHFR through its promoter hypermethylation leads also to the activation of DNA methyltransferases (including DNMT1). Different patterns of demethylation obtained by silencing DNMT1 in experimental model (HeLa and SiHa cell lines) indicate the inhibition of DNMT1 as a target for the treatment of cervical cancer with HPV18 infection [312]. These results showed that infection with different HPV genotypes differently interfere with epigenetic mechanisms.

Molecular investigations of cervical tumors and cell lines immortalized with HPV have shown that, from all ncRNA molecules, miRNAs profile is significantly changed when compared to normal tissue, even in early stages of carcinogenesis [309]. Zheng et al. [319] provided data that viral E6 and E7 oncoproteins deregulate the expression of several miRNAs via the E6‐p53 and E7‐pRb pathways. In turn, miRNAs may influence the expression of HPV genes by targeting viral RNA transcripts, these recommended miRNAs as new biomarkers in cervical screening. The panel of four circulating miRNAs (miR‐16‐2*, miR‐195, miR‐2861, miR‐49) [320] are suggested as predictive biomarkers for the prognosis of cervical cancer patients, upregulate expression of serum miR‐205 [321] and serum pattern of miR‐29a and miR‐200a may indicate tumor histological grade and progression stage [322]. Li et al. [323] found lower levels of miR‐218 levels in patients with high‐risk HPV comparing with control or those with low‐risk or intermediate‐risk HPV. In Chinese population, Zhou et al. [324] reported a good correlation between a miR‐218 polymorphism and its target laminin 5B3 in cervical cancer invasiveness. Epigenetic changes through methylation of miRNAs might correlate with cervical disease. A panel of three miRs (miR‐149, miR‐203, and miR‐375) was found hypermethylated in HPV‐positive cell lines [76] and miR‐203 and miR‐375 hypermethylation correlated with uterine precancerous lesions [325].

miRNAs might be also used for cervical cancer therapy. On animal model, Liu et al. [227] who noticed an inverse correlation between the expression of miR‐143 and Bcl2 suggested the possibility of a therapeutic approach by targeting this pathway. Also, miRNAs might modulate the sensitivity to chemotherapy. For example, miR‐375 might be a therapeutic target in paclitaxel‐resistance of cervical cancer cells [326], while miR‐155 and miR‐281 increase sensitivity to cisplatin [325]. Therefore, miRNA deregulation may become a target of the investigations for evaluating the effectiveness of treatments in cervical cancer [327].

Advertisement

7. Conclusions

All these data underline the importance of epigenetic modification in tumor development and cervical cancer risk assessment. Epigenetic alterations could be used as biomarkers for the prognosis and evolution of the disease and for therapy response prediction. New techniques in epigenetic investigations may yield better detection systems in order to identify new and sensitive biomarkers that might contribute to improved screening assays, new therapeutic approaches, and prediction biomarkers. The reversible nature of epigenetic alterations provides new opportunities in the development of therapeutic agents targeting epigenetic modification in oncogenesis.

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Acknowledgments

POSCCE SMIS 14049

This work is supported by The Executive Agency for Higher Education, Research, Development and Innovation Funding (UEFISCDI) under grant: PNII‐RU‐TE‐2014‐4‐2502 (OncoNuRD).

References

  1. 1. Pisani P, Bray F, Parkin DM. Estimates of the world‐wide prevalence of cancer for 25 sites in the adult population. Int J Cancer. 2002;97:72–81. doi:10.1002/ijc.1571
  2. 2. Gillison ML, Shah KV. Human papillomavirus‐associated head and neck squamous cell carcinoma: mounting evidence for an etiologic role for human papillomavirus in a subset of head and neck cancers. Curr Opin Oncol. 2001;13:183–188.
  3. 3. Munoz N, Mendez F, Posso H. Incidence, duration, and determinants of cervical human papillomavirus infection in a cohort of Colombian women with normal cytological results. J Infect Dis. 2004;190:2077–2087. doi:10.1086/425907
  4. 4. Castle PE, Rodriguez AC, Burk RD. Short term persistence of human papillomavirus and risk of cervical precancer and cancer: population based cohort study. BMJ. 2009;339:b2569. doi:10.1136/bm
  5. 5. Pim D, Banks L. Interaction of viral oncoproteins with cellular target molecules: infection with high‐risk vs low‐risk human papillomaviruses. APMIS. 2010;118:471–493. doi:10.1111/j.1600‐0463.2010.02618.x
  6. 6. Alsbeih G, Ahmed R, Al‐Harbi N, Venturina LA, Tulbah A, Balaraj K. Prevalence and genotypes’ distribution of human papillomavirus in invasive cervical cancer in Saudi Arabia. Gynecol Oncol. 2011;121:522–526. doi:10.1016/j.ygyno.2011.01.033
  7. 7. Grce M, Sabol I, Milutin‐Gasperov N. Burden and prevention of HPV related diseases: Situation in Croatia. Period Biol. 2012;114(4):175–186. doi:10.18054.
  8. 8. Grce M, Sabol I, Milutin‐Gasperov N. The transforming properties of human papillomavirus oncoproteins. Period Biol. 2012;114(4):479–487. doi:10.18054
  9. 9. Scheffner M, Werness BA, Huibregtse JM, Levine AJ, Howley PM. The E6 oncoprotein encoded by human papillomavirus types 16 and 18 promotes the degradation of p53. Cell. 1990;63:1129–1136. doi:10.1016/0092‐8674(90)90409‐8
  10. 10. Filippova M, Song H, Connolly JL, Dermody TS, Duerksen‐Hughes PJ. The human papillomavirus 16 E6 protein binds to tumor necrosis factor (TNF) R1 and protects cells from TNF‐induced apoptosis. J Biol Chem. 2002;277:21730–21739. doi:10.1074/jbc.M200113200
  11. 11. Filippova M, Parkhurst L, Duerksen‐Hughes PJ. The human papillomavirus 16 E6 protein binds to Fas‐associated death domain and protects cells from Fas‐triggered apoptosis. J Biol Chem. 2004;279:25729–25744. doi:10.1074/jbc.M401172200
  12. 12. Filippova M, Johnson MM, Bautista M, Filippov V, Fodor N, Tungteakkhun SS, Williams K, Duerksen‐Hughes PJ. The large and small isoforms of human papillomavirus type 16 E6 bind to and differentially affect procaspase 8 stability and activity. J Virol. 2007;81:4116–4129. doi:10.1128/JVI.01924‐06
  13. 13. Thomas M, Banks L. Inhibition of Bak‐induced apoptosis by HPV‐18 E6. Oncogene. 1998;17:2943–2954.
  14. 14. Jong JE, Jeong KW, Shin H, Hwang LR, Lee D, Seo T. Human papillomavirus type 16 E6 protein inhibits DNA fragmentation via interaction with DNA fragmentation factor 40. Cancer Lett. 2012;324:109–117. doi:10.1016/j.canlet.2012.05.010
  15. 15. Jha S, Vande Pol S, Banerjee NS, Dutta AB, Chow LT, Dutta A. Destabilization of TIP60 by human papillomavirus E6 results in attenuation of TIP60‐dependent transcriptional regulation and apoptotic pathway. Mol Cell. 2010;38:700–711. doi:10.1016/j.molcel.2010.05.020
  16. 16. Lee SS, Glaunsinger B, Mantovani F, Banks L, Javier RT. Multi‐PDZ domain ProteinMUPP1 is a cellular target for both adenovirus E4‐ORF1 and high‐risk papillomavirus type 18 E6 oncoproteins. J Virol. 2000;74:9680–9693. doi:10.1128/JVI.74.20.9680‐9693.2000
  17. 17. Chen JJ, Reid CE, Band V, Androphy EJ. Interaction of papillomavirus E6 oncoproteins with a putative calcium‐binding protein. Science. 1995;269:529–531
  18. 18. Glaunsinger BA, Lee SS, Thomas M, Banks L, Javier R. Interactions of the PDZ‐protein MAGI‐1 with adenovirus E4‐ORF1 and high‐risk papillomavirus E6 oncoproteins. Oncogene. 2000;19:5270–5280. doi:10.1038/sj.onc.1203906
  19. 19. Kiyono T, Hiraiwa A, Fujita M, Hayashi Y, Akiyama T, Ishibashi M. Binding of high‐risk human papillomavirus E6 oncoproteins to the human homologue of the Drosophila discs large tumor suppressor protein. Proc Natl Acad Sci USA. 1997;94:11612–11616
  20. 20. Latorre IJ, Roh MH, Frese KK, Weiss RS, Margolis B, Javier RT. Viral oncoprotein‐induced mislocalization of select PDZ proteins disrupts tight junctions and causes polarity defects in epithelial cells. J Cell Sci. 2005;118:4283–4293. doi:10.1242/jcs.02560
  21. 21. Tong X, Howley PM. The bovine papillomavirus E6 oncoprotein interacts with paxillin and disrupts the actin cytoskeleton. Proc Natl Acad Sci USA.1997;94:4412–4417.
  22. 22. Zhang Y, Fan S, Meng Q, Ma Y, Katiyar P, Schlegel R, Rosen EM. BRCA1 interaction with human papillomavirus oncoproteins. J Biol Chem. 2005;280:33165–33177. doi:10.1074/jbc.M505124200
  23. 23. Iftner T, Elbel M, Schopp B, Hiller T, Loizou JI, Caldecott KW, Stubenrauch F. Interference of papillomavirus E6 protein with single‐strand break repair by interaction with XRCC1. EMBO J. 2002;21:4741–4748. doi:10.1093/emboj/cdf443
  24. 24. Srivenugopal KS, Ali‐Osman F. The DNA repair protein, O(6)‐methylguanine‐DNA methyltransferase is a proteolytic target for the E6 human papillomavirus oncoprotein. Oncogene. 2002;21:5940–5945. doi:10.1038/sj.onc.1205762
  25. 25. Gewin L, Myers H, Kiyono T, Galloway DA. Identification of a novel telomerase repressor that interacts with the human papillomavirus type‐16 E6/E6‐AP complex. Genes Dev. 2004;18:2269–2282. doi:10.1101/gad.1214704
  26. 26. Liu X, Dakic A, Zhang Y, Dai Y, Chen R, Schlegel R. HPV E6 protein interacts physically and functionally with the cellular telomerase complex. Proc Natl Acad Sci USA. 2009;106:18780–18785. doi:10.1073/pnas.0906357106
  27. 27. Li S, Labrecque S, Gauzzi MC, Cuddihy AR, Wong AH, Pellegrini S, Matlashewski GJ, Koromilas AE. The human papilloma virus (HPV)‐18 E6 oncoprotein physically associates with Tyk2 and impairs Jak‐STAT activation by interferon‐alpha. Oncogene. 1999;18:5727–5737
  28. 28. Ronco LV, Karpova AY, Vidal M, Howley PM. Human papillomavirus 16 E6 oncoprotein binds to interferon regulatory factor‐3 and inhibits its transcriptional activity. Genes Dev. 1998;12:2061–2072. doi:10.1101/gad.12.13.2061
  29. 29. Ghittoni R, Accardi R, Hasan U, Gheit T, Sylla B, Tommasino M. The biological properties of E6 and E7 oncoproteins from human papillomaviruses. Virus Genes, 2010;40:1–13. doi:10.1007/s11262‐009‐0412‐8
  30. 30. Zerfass‐Thome K, Zwerschke W, Mannhardt B, Tindle R, Botz JW, Jansen‐Dürr P. Inactivation of the cdk inhibitor p27KIP1 by the human papillomavirus type 16 E7 oncoprotein. Oncogene. 1996;13:2323–2330
  31. 31. Arroyo M, Bagchi S, Raychaudhuri P. Association of the human papillomavirus type 16 E7 protein with the S‐phase—specific E2F‐cyclin A complex. Mol Cell Biol. 1993;13:6537–6546. doi:10.1128/MCB.13.10.6537
  32. 32. McIntyre MC, Ruesch MN, Laimins LA. Human papillomavirus E7 oncoproteins bind a single form of cyclin E in a complex with cdk2 and p107. Virology. 1996;215:73–82. doi:10.1006/viro.1996.0008
  33. 33. Antinore MJ, Birrer MJ, Patel D, Nader L, McCance DJ. The human papillomavirus type 16 E7 gene product interacts with and trans‐activates the AP1 family of transcription factors. EMBO J.1996;15:1950–1960
  34. 34. Phillips AC, Vousden KH. Analysis of the interaction between human papillomavirus type 16 E7 and the TATA‐binding protein, TBP. J Gen Virol. 1997;78(Pt4):905–909. doi:10.1099/0022‐1317‐78‐4‐905
  35. 35. Lüscher‐Firzlaff JM, Westendorf JM, Zwicker J, Burkhardt H, Henriksson M, Müller R, Pirollet F, Lüscher B. Interaction of the fork head domain transcription factor MPP2 with the human papilloma virus 16 E7 protein: enhancement of transformation and transactivation. Oncogene. 1999;18:5620–5630
  36. 36. McLaughlin‐Drubin ME, Huh KW, Münger K. Human papillomavirus type 16E7 oncoprotein associateswith E2F6. J Virol. 2008;82:8695–8705. doi:10.1128/JVI.00579‐08
  37. 37. Prathapam T, Kühne C, Banks L. The HPV‐16 E7 oncoprotein binds Skip and suppresses its transcriptional activity. Oncogene. 2001;20:7677–7685. 10.1093/nar/29.17.3469
  38. 38. Baldwin A, Huh KW, Münger K. Human papillomavirus E7 oncoprotein dysregulates steroid receptor coactivator 1 localization and function. J Virol. 2006;80:6669–6677. doi:10.1128/JVI.02497‐05
  39. 39. Suzuki MM, Bird A. DNA methylation landscapes: provocative insights from epigenomics. Nat Revi Genet. 2008;9:465–476. doi:10.1038/nrg2341
  40. 40. Bannister AJ, Kouzarides T. Regulation of chromatin by histone modifications. Cell Res. 2011;21:381–395. doi:10.1038/cr.2011.22
  41. 41. Storz G. An expanding universe of noncoding RNAs. Science. 2002;296:1260–1263. doi:10.1126/science.1072249
  42. 42. Bird A. DNA methylation patterns and epigenetic memory. Genes Dev. 2002;16:6–21. doi:10.1101/gad.947102
  43. 43. Wade PA. Methyl CpG‐binding protiens and transcriptional repression. Bioessays. 2001;23:1131–1137. doi:10.1002/bies.10008
  44. 44. Greger V, Passarge E, Hopping W, Messmer E, Horsthemke B. Epigenetic changes may contribute to the formation and spontaneous regression of retinoblastoma. Hum Genet. 1989;83:155–158.
  45. 45. Santini V, Kantarjian HM, Issa JP. Changes in DNA methylation in neoplasia: pathophysiology and therapeutic implications. Ann Intern Med. 2001;134:573–586. doi:10.7326/0003‐4819‐134‐7‐200104030‐00011
  46. 46. Szyf M. Targeting DNA methylation in cancer. Ageing Res Rev. 2003;2:299–328. doi:10.1016/S1568‐1637(03)00012‐6
  47. 47. Bachman KE, Park BH, Rhee I, Rajagopalan H, Herman JG, Baylin SB, Kinzler KW, Vogelstein B. Histone modifications and silencing prior to DNA methylation of a tumor suppressor gene. Cancer Cell. 2003;3:89–95. doi:10.1016/S1535‐6108(02)00234‐9
  48. 48. Clark SJ, Melki J. DNA methylation and gene silencing in cancer: which is the guilty part? Oncogene. 2002;21:5380–5387. doi:10.1038/sj.onc.1205598
  49. 49. Dueñas‐González A, Lizano M, Candelaria M, Cetina L, Arce C, Cervera E. Epigenetics of cervical cancer. An overview and therapeutic perspectives. Mol Cancer. 2005;4:38. doi:10.1186/1476‐4598‐4‐38
  50. 50. Soengas MS, Capodieci P, Polsky D, Mora J, Esteller M, Opitz‐Araya X, McCombie R, Herman JG, Gerald WL, Lazebnik YA, Cordon‐Cardo C, Lowe SW. Inactivation of the apoptosis effector Apaf‐1 in malignant melanoma. Nature. 2001;409:207–211. doi:10.1038/35051606
  51. 51. van Noesel MM, van Bezouw S, Salomons GS, Voute PA, Pieters R, Baylin SB, Herman JG, Versteeg R. Tumor‐specific down‐regulationof the tumor necrosis factor‐related apoptosis‐inducing ligand decoy receptors DcR1 and DcR2 is associated with dense promoter hypermethylation. Cancer Res. 2002;62:2157–2161.
  52. 52. Ashkenazi A, Dixit VM. Apoptosis control by death and decoy receptors. Curr Opin Cell Biol. 1999;11:255–260. doi:10.1016/S0955‐0674(99)80034‐9
  53. 53. Vousden KH. p53: death star. Cell. 2000;103:691–694. doi:10.1016/S0092‐8674(00)00171‐9
  54. 54. Nakashima R, Fujita M, Enomoto T, Haba T, Yoshino K, Wada H, Kurachi H, Sasaki M, Wakasa K, Inoue M, Buzard G, Murata Y. Alteration of p16 and p15 genes in human uterine tumours. Br J Cancer. 1999;80:458–467. doi:10.1038/sj.bjc.6690379
  55. 55. Narayan G, Arias‐Pulido H, Koul S, Vargas H, Zhang FF, Villella J, Schneider A, Terry MB, Mansukhani M, Murty VV. Frequent promoter methylation of CDH1, DAPK, RARB, and HIC1 genes in carcinoma of cervix uteri: its relationship to clinical outcome. Mol Cancer. 2003;2:24. doi:10.1186/1476‐4598‐2‐24
  56. 56. Cheung TH, Lo KW, Yim SF, Chan LK, Heung MS, Chan CS, Cheung AY, Chung TK, Wong YF. Epigenetic and genetic alternation of PTEN in cervical neoplasm. Gynecol Oncol. 2004;93:621–627. doi:10.1016/j.ygyno.2004.03.013
  57. 57. Widschwendter A, Ivarsson L, Blassnig A, Muller HM, Fiegl H, Wiedemair A, Muller‐Holzner E, Goebel G, Marth C, Widschwendter M. CDH1 and CDH13 methylation in serum is an independent prognostic marker in cervical cancer patients. Int J Cancer. 2004;109:163–166. doi:10.1002/ijc.11706
  58. 58. Qi J, Zhu YQ, Huang MF, Yang D. Hypermethylation of CpG island in O6‐methylguanine‐DNA methyltransferase gene was associated with K‐ras G to A mutation in colorectal tumor. World J Gastroenterol. 2005;11:2022–2025. doi:10.3748/wjg.v11.i13.2022
  59. 59. D'Andrea AD, Grompe M: The Fanconi anaemia/BRCA pathway. Nat Rev Cancer. 2003;3:23–34. doi:10.1038/nrc970
  60. 60. Taniguchi T, Tischkowitz M, Ameziane N. Disruption of the Fanconi anemia‐BRCA pathway in cisplatin‐sensitive ovarian tumors. Nat Med. 2003;9:568–574. doi:10.1038/nm852
  61. 61. Aaltonen LA, Peltomaki P, Mecklin JP, Jarvinen H, Jass JR, Green JS, Lynch HT, Watson P, Tallqvist G, Juhola M, Sistonen P, Hamilton SR,Kinzler KW, Vogelstein B, de la Chapelle A. Replication errors in benign and malignant tumors from hereditary non‐polyposis colorectal cancer patients. Cancer Res. 1994;54:1645–1648.
  62. 62. Dammann R, Schagdarsurengin U, Seidel C, Strunnikova M, Rastetter M, Baier K, Pfeifer GP: The tumor suppressor RASSF1A in human carcinogenesis: an update. Histol Histopathol. 2005;20:645–663
  63. 63. Uthoff SM, Eichenberger MR, McAuliffe TL, Hamilton CJ, Galandiuk S. Wingless‐type frizzled protein receptor signaling and its putative role in human colon cancer. Mol Carcinog. 2001;31:56–62. doi:10.1002/mc.1039
  64. 64. Watabe K, Ito A, Koma YI, Kitamura Y. IGSF4: a new intercellular adhesion molecule that is called by three names, TSLC1, SgIGSF and SynCAM, by virtue of its diverse function. Histol Histopathol. 2003;18:1321–1329.
  65. 65. Greenspan DL, Connolly DC, Wu R, Lei RY, Vogelstein JT, Kim YT,Mok JE, Muñoz N, Bosch FX, Shah K, Cho KR. Loss of FHIT expression in cervical carcinoma cell lines and primary tumors. Cancer Res. 1997;57:4692–4698.
  66. 66. Roz L, Gramegna M, Ishii H, Croce CM, Sozzi G. Restoration of fragile histidine triad (FHIT) expression induces apoptosis and suppresses tumorigenicity in lung and cervical cancer cell lines. Proc Natl Acad Sci USA. 2002;99:3615–3620. doi:10.1073/pnas.062030799
  67. 67. Virmani AK, Muller C, Rathi A, Zoechbauer‐Mueller S, Mathis M,Gazdar AF. Aberrant methylation during cervical carcinogenesis. Clin Cancer Res. 2001;7:584–589.
  68. 68. Lippman SM, Kavanagh JJ, Paredes‐Espinoza M, Delgadillo‐Madrueno F, Paredes Casillas P, Hong WK, Massimini G, Holdener EE, Kazakoff IH. 13‐cis‐Retinoic acid plus interferon‐α2a in locally advanced squamous cell carcinoma of cervix. J Natl Cancer Inst. 1993;85:499–500. doi:10.1093/jnci/85.6.499
  69. 69. Geraddts J, Chen JY, Russell EA, Yankaskas JR, Nieves L, Minna JD. Human cancer cell lines exhibit resistance to retinoic acid treatment. Cell Growth Differ. 1993;4:799–809
  70. 70. Ivanova T, Petrenko A, Gritsko T, Vinokourova S, Eshilev E, Kobzeva V, Kisseljov F, Kisseljova N. Methylation and silencing of the retinoic acid receptor‐beta 2 gene in cervical cancer. BMC Cancer. 2002;2:4. doi:10.1016/j.canlet.2006.05.013
  71. 71. Feng Q, Balasubramanian A, Hawes SE, Toure P, Sow PS, Dem A, Dembele B, Critchlow CW, Xi L, Lu H, McIntosh MW, Young AM, Kiviat NB. Detection of hypermethylated genes in women with and without cervical neoplasia. J Natl Cancer Inst. 2005;97:273–282. doi:10.1093/jnci/dji041
  72. 72. Widschwendter A, Muller HM, Fiegl H, Ivarsson L, Wiedemair A, Muller‐Holzner E, Goebel G, Marth C, Widschwendter M. DNA methylation in serum and tumors of cervical cancer patients. Clin Cancer Res. 2004;10:565–571. doi:10.1158/1078‐0432.CCR‐0825‐03
  73. 73. Razani B, Altschuler Y, Zhu L, Pestell RG, Mostov KE, Lisanti MP. Caveolin‐1 expression is down‐regulated in cells transformed by the human papilloma virus in a p53‐dependent manner. Replacement of caveolin‐1 expression suppresses HPV‐mediated cell transformation. Biochemistry. 2000;39:13916–13924. doi:10.1021/bi001489b
  74. 74. Kirn V, Zaharieva I, Heublein S, Thangarajah F, Friese K, Mayr D, Jeschke U. ESR1 promoter methylation in squamous cell cervical cancer. Anticancer Res. 2014;34(2):723–727.
  75. 75. Botezatu A, Goia‐Rusanu CD, Iancu IV, Huica I, Plesa A, Socolov D, Ungureanu C, Anton G. Quantitative analysis of the relationship between microRNA-124a, ‐34b and ‐203 gene methylation and cervical oncogenesis. Mol Med Rep. 2011;4(1):121–128. doi:10.3892/mmr.2010.394
  76. 76. Wilting SM, Verlaat W, Jaspers A, Makazaji NA, Agami R, Meijer CJ, et al. Methylation‐mediated transcriptional repression of microRNAs during cervical carcinogenesis. Epigenetics. 2013;8:220–228.
  77. 77. Yao T, Rao Q, Liu L, Zheng C, Xie Q, Liang J. Exploration of tumor‐suppressive microRNAs silenced by DNA hypermethylation in cervical cancer. Virol J. 2013;10:175.
  78. 78. Ki EY, Lee KH, Hur SY, Rhee JE, Kee MK, Kang C, Park JS. Methylation of cervical neoplastic cells infected with human papillomavirus 16. Int J Gynecol Cancer. 2016;26(1):176–183. doi:10.1097/IGC.0000000000000582
  79. 79. Schütze DM, Kooter JM, Wilting SM, Meijer CJ, Quint W, Snijders PJ, Steenbergen RD. Longitudinal assessment of DNA methylation changes during HPVE6E7‐induced immortalization of primary keratinocytes. Epigenetics. 2015;10(1):73–81. doi:10.4161/15592294.2014.990787
  80. 80. Segal E, Fondufe‐Mittendorf Y, Chen L, Thåström A, Field Y, Moore IK, Wang JP, Widom J. A genomic code for nucleosome positioning. Nature. 2006;442(7104):772–778. doi:10.1038/nature04979
  81. 81. Endo M, Majima T. Control of a double helix DNA assembly by use of cross‐linked oligonucleotides. J Am Chem Soc. 2003;125(45):13654–13655. doi:10.1021/ja036752l
  82. 82. Kouzarides T. Chromatin modifications and their function. Cell. 2007;128:693–705. doi:10.1016/j.cell.2007.02.005
  83. 83. Shahbazian MD, Grunstein M. Functions of site‐specific histone acetylation and deacetylation. Annu Rev Biochem. 2007;76:75–100. doi:10.1146/annurev.biochem.76.052705.162114
  84. 84. Shogren‐Knaak M, Ishii H, Sun JM, Pazin MJ, Davie JR, Peterson CL. Histone H4‐K16 acetylation controls chromatin structure and protein interactions. Science. 2006;311(5762):844–847. doi:10.1126/science.1124000
  85. 85. Braunstein M, Sobel RE, Allis CD, Turner BM, Broach JR. Efficient transcriptional silencing in Saccharomyces cerevisiae requires a heterochromatin histone acetylation pattern. Mol Cell Biol. 1996;16(8):4349–4356. doi:10.1128/MCB.16.8.4349
  86. 86. Turner BM, Birley AJ, Lavender J. Histone H4 isoforms acetylated at specific lysine residues define individual chromosomes and chromatin domains in Drosophila polytene nuclei. Cell. 1992;69(2):375–384. doi:10.1016/0092‐8674(92)90417‐B
  87. 87. Zhang Y, Reinberg D. Transcription regulation by histone methylation: interplay between different covalent modifications of the core histone tails. Genes Dev. 2001;15:2343–2360. doi:10.1101/gad.927301
  88. 88. Shilatifard A. Chromatin modifications by methylation and ubiquitination: implications in the regulation of gene expression. Annu Rev Biochem. 2006;75:243–269. doi:10.1146/annurev.biochem.75.103004.142422
  89. 89. Krogan NJ, Dover J, Khorrami S, Greenblatt JF, Schneider J, Johnston M, Shilatifard A. COMPASS, a histone H3 (Lysine 4) methyltransferase required for telomeric silencing of gene expression. J Biol Chem. 2002;277:10753–10755. doi:10.1074/jbc.C200023200
  90. 90. Lee DY, Teyssier C, Strahl BD, Stallcup MR. Role of protein methylation in regulation of transcription. Endocr Rev. 2005;26:147–170. doi:10.1210/er.2004‐0008.
  91. 91. Feng Q, Wang H, Ng HH, Erdjument‐Bromage H, Tempst P, Struhl K, Zhang Y Methylation of H3‐lysine 79 is mediated by a new family of HMTases without a SET domain. Curr Biol. 2002;12(12):1052–1058. doi:10.1016/S0960‐9822(02)00901‐6
  92. 92. Lacoste N, Utley RT, Hunter JM, Poirier GG, Côte J. Disruptor of telomeric silencing‐1 is a chromatin‐specific histone H3 methyltransferase. J Biol Chem. 2002;277:30421–30424. doi:10.1074/jbc.C200366200
  93. 93. Sims RJ III, Reinberg D. Histone H3 Lys 4 methylation: caught in a bind? Genes. 2006;20:2779–2786. doi:10.1101/gad.1468206
  94. 94. Lee N, Zhang J, Klose RJ, Erdjument‐Bromage H, Tempst P, Jones RS, Zhang Y. The trithorax‐group protein Lid is a histone H3 trimethyl‐Lys4 demethylase. Nat Struct Mol Biol. 2007;14(4):252–254. doi:10.1038/nsmb1216
  95. 95. Martin C, Zhang Y. The diverse functions of histone lysine methylation. Nat Rev Mol Cell Biol. 2005;6:838–849. doi:10.1038/nrm1761
  96. 96. Shi X, Hong T, Walter KL, Ewalt M, Michishita E, Hung T, Carney D, Peña P, Lan F, Kaadige MR, Lacoste N, Cayrou C, Davrazou F, Saha A, Cairns BR, Ayer DE, Kutateladze TG, Shi Y, Côté J, Chua KF, Gozani O. ING2 PHD domain links histone H3 lysine 4 methylation to active gene repression. Nature. 2006;442(7098):96–99. doi:10.1038/nature04835
  97. 97. Barber CM, Turner FB, Wang Y, Hagstrom K, Taverna SD, Mollah S, Ueberheide B, Meyer BJ, Hunt DF, Cheung P, Allis CD. The enhancement of histone H4 and H2A serine 1 phosphorylation during mitosis and S‐phase is evolutionarily conserved. Chromosoma. 2004;112(7):360–371. doi:10.1007/s00412‐004‐0281‐9
  98. 98. Bradbury EM. Reversible histone modifications and the chromosome cell cycle. Bioessay.1992;14(1):9–16. doi:10.1002/bies.950140103
  99. 99. Wendt KD, Shilatifard A. Packing for the germy: the role of histone H4 Ser1 phosphorylation in chromatin compaction and germ cell development. Genes Dev. 2006;20:2487–2491. doi:10.1101/gad.1477706
  100. 100. Downs JA, Lowndes NF, Jackson SP. A role for Saccharomyces cerevisiae histone H2A in DNA repair. Nature. 2000;408(6815):1001–1004. doi:10.1038/35050000
  101. 101. Osley MA. Regulation of histone H2A and H2B ubiquitylation. Brief Funct Genom Proteom. 2006;5:179–189. doi:10.1093/bfgp/ell022
  102. 102. Henry KW, Wyce A, Lo WS, Duggan LJ, Emre NC, Kao CF, Pillus L, Shilatifard A, Osley MA, Berger SL. Transcriptional activation via sequential histone H2B ubiquitylation and deubiquitylation, mediated by SAGA‐associated Ubp8. Genes Dev. 2003;17:2648–2663. doi:10.1101/gad.1144003
  103. 103. Zhu P, Zhou W, Wang J, Puc J, Ohgi KA, Erdjument‐Bromage H, Tempst P, Glass CK, Rosenfeld MG. Histone H2A deubiquitinase complex coordinating histone acetylation and H1 dissociation. Mol Cell. 2007;27:609–621. doi:10.1016/j.molcel. 2007.07.024
  104. 104. Hassa PO, Haenni SS, Elser M, Hottiger MO. Nuclear ADP‐ribosylation reactions in mammalian cells: where are we today and where are we going? Microbiol Mol Biol Rev. 2006;70:789–829. doi:10.1128/MMBR.00040‐05
  105. 105. Zhu B, Zheng Y, Pham AD, Mandal SS, Erdjument‐Bromage H, Tempst P, Reinberg D. Monoubiquitination of human histone H2B: the factors involved and their roles in HOX gene regulation. Mol Cell. 2005;20:601–611. doi:10.1016/j.molcel.2005.09.025
  106. 106. Nakagawa T, Kajitani T, Togo S, Masuko N, Ohdan H, Hishikawa Y, Koji T, Matsuyama T, Ikura T, Muramatsu M, Ito T. Deubiquitylation of histone H2A activates transcriptional initiation via trans‐histone cross‐talk with H3K4 di‐ and trimethylation. Genes Dev. 2008;22:37–49. doi:10.1101/gad.1609708
  107. 107. Nan X, Ng HH, Johnson CA, Laherty CD, Turner BM, Eisenman RN, Bird A. Transcriptional repression by the methyl‐CpG‐binding protein MeCP2 involves a histone deacetylase complex. Nature.1998;393:386–389. doi:10.1038/30764
  108. 108. Hendrich B, Bird A. Identification and characterization of a family of mammalian methyl‐CpG binding proteins. Mol Cell Biol. 1998;18:6538–6547. doi:10.1128/MCB.18.11.6538
  109. 109. Okitsu CY, Hsieh CL. DNA methylation dictates histone H3K4 methylation. Mol Cell Biol. 2007;27:2746–2757. doi:10.1128/MCB.02291‐06
  110. 110. Feldman N, Gerson A, Fang J, Li E, Zhang Y, Shinkai Y, Cedar H, Bergman Y. G9a‐mediated irreversible epigenetic inactivation of Oct‐3/4during early embryogenesis. Nat Cell Biol. 2006;8:188–194. doi:10.1038/ncb1353
  111. 111. Epsztejn‐Litman S, Feldman N, Abu‐Remaileh M, Shufaro Y, Gerson A, Ueda J, Deplus R, Fuks F, Shinkai Y, Cedar H, Bergman Y. De novo DNA methylation promoted byG9a prevents reprogramming of embryonically silenced genes. Nat Struct Mol Biol. 2008;15:1176–1183. doi:10.1038/nsmb.1476
  112. 112. Dong KB, Maksakova IA, Mohn F, Leung D, Appanah R, Lee S, Yang HW, Lam LL, Mager DL, Schübeler D, Tachibana M, Shinkai Y, Lorincz MC. DNA methylation in ES cells requires the lysine methyltransferase G9a but not its catalytic activity. EMBO J. 2008;27:2691–2701. doi:10.1038/emboj.2008
  113. 113. Tachibana M, Matsumura Y, Fukuda M, Kimura H, Shinkai Y. G9a/GLP complexes independently mediate H3K9 and DNA methylation to silence transcription. EMBO J. 2008;27:2681–2690. doi:10.1038/emboj.2008.192
  114. 114. Ohm JE, McGarvey KM, Yu X, Cheng L, Schuebel KE, Cope L, Mohammad HP, Chen W, Daniel VC, Yu W, Berman DM, Jenuwein T, Pruitt K, Sharkis SJ, Watkins DN, Herman JG, Baylin SB. A stem cell‐like chromatin pattern may predispose tumor suppressor genes to DNA hypermethylation and heritable silencing. Nat Genet. 2007;39:237–242. doi:10.1038/ng1972
  115. 115. Widschwendter M, Fiegl H, Egle D, Mueller‐Holzner E, Spizzo G, Marth C, Weisenberger DJ, Campan M, Young J, Jacobs I, Laird PW. Epigenetic stem cell signature in cancer. Nat Genet. 2007;39:157–158. doi:10.1038/ng1941
  116. 116. Schlesinger Y, Straussman R, Keshet I, Farkash S, Hecht M, Zimmerman J, Eden E, Yakhini Z, Ben‐Shushan E, Reubinoff BE, Bergman Y, Simon I, Cedar H. Polycomb‐mediated methylation on Lys27of histone H3 pre‐marks genes for de novo methylation in cancer. Nat Genet. 2007;39:232–236. doi:10.1038/ng1950
  117. 117. Viré E, Brenner C, Deplus R, Blanchon L, Fraga M, Didelot C, Morey L, Van Eynde A, Bernard D, Vanderwinden JM, Bollen M, Esteller M, Di Croce L, de Launoit Y, Fuks F. The Polycomb group protein EZH2 directly controls DNA methylation. Nature. 2006;439:871–874. doi:10.1038/nature04431
  118. 118. Viré E, Brenner C, Deplus R, Blanchon L, Fraga M, Didelot C, Morey L, Van Eynde A, Bernard D, Vanderwinden JM, Bollen M, Esteller M, Di Croce L, de Launoit Y, Fuks F. Promoter CpG methylation contributes to ES cell gene regulation in parallel with Oct4/Nanog, PcG complex, and histone H3 K4/K27 trimethylation. Cell Stem Cell. 2008;2:160–169. doi:10.1038/nature04431
  119. 119. Kondo Y, Shen L, Cheng AS, Ahmed S, Boumber Y, Charo C, Yamochi T, Urano T, Furukawa K, Kwabi‐Addo B, Gold DL, Sekido Y, Huang TH, Issa JP. Gene silencing in cancer by histone H3 lysine 27 trimethylation independent of promoter DNA methylation. Nat Genet. 2008;40:741–750. doi:10.1038/ng.159
  120. 120. Gal‐Yam EN, Egger G, Iniguez L, Holster H, Einarsson S, Zhang X, Lin JC, Liang G, Jones PA, Tanay A. Frequent switching of Polycomb repressive marks and DNA hypermethylation in the PC3 prostate cancer cell line. Proc Natl Acad Sci USA. 2008;105:12979–12984. doi:10.1073/pnas.0806437105
  121. 121. Mikkelsen TS, Hanna J, Zhang X, Ku M, Wernig M, Schorderet P, Bernstein BE, Jaenisch R, Lander ES, Meissner A. Dissecting direct reprogramming through integrative genomic analysis. Nature. 2008;454:49–55. doi:10.1038/nature07056
  122. 122. Ward R, Johnson M, Shridhar V, van Deursen J, Couch FJ. CBP truncating mutations in ovarian cancer. J Med Genet. 2005;42:514–518. doi:10.1136/jmg.2004.025080
  123. 123. So CK, Nie Y, Song Y, Yang GY, Chen S, Wei C, Wang LD, Doggett NA, Yang CS. Loss of heterozygosity and internal tandem duplication mutations of the CBP gene are frequent events in human esophageal squamous cell carcinoma. Clin Cancer Res. 2004;10:19–27.
  124. 124. Kishimoto M, Kohno T, Okudela K, Otsuka A, Sasaki H, Tanabe C, Sakiyama T, Hirama C, Kitabayashi I, Minna JD, Takenoshita S, Yokota J. Mutations and deletions of the CBP gene in human lung cancer. Clin Cancer Res. 2005;11:512–519.
  125. 125. Sobulo OM, Borrow J, Tomek R, Reshmi S, Harden A, Schlegelberger B, Housman D, Doggett NA, Rowley JD, Zeleznik‐Le NJ. MLL is fused to CBP, a histone acetyltransferase, in therapy‐related acute myeloid leukemia with a t(11;16)(q23;p13.3). Proc Natl Acad Sci USA. 1997;94:8732–8777
  126. 126. Chaffanet M, Gressin L, Preudhomme C, Soenen‐Cornu V, Birnbaum D, Pebusque MJ. MOZ is fused to p300 in an acute monocytic leukemia with t(8;22). Genes Chromosomes Cancer. 2000;28:138–144. doi:10.1002/(SICI)1098‐2264(200006)28:2<138::AID‐GCC2>3.0.CO
  127. 127. Lin RJ, Nagy L, Inoue S, Shao W, Miller WH Jr, Evans RM. Role of the histone deacetylase complex in acute promyelocytic leukaemia. Nature. 1998;391:811–814. doi:10.1038/35895
  128. 128. Esteller M. Epigenetics in cancer. N Engl J Med. 2008;358:1148–1159. doi:10.1056/NEJMra072067
  129. 129. Esteller M. Cancer epigenomics: DNA methylomes and histone‐modification maps. Nat Rev Genet. 2007;8(4):286–298. doi:10.1038/nrg2005
  130. 130. Kleer CG, Cao Q, Varambally S, Shen R, Ota I, Tomlins SA, Ghosh D, Sewalt RG, Otte AP, Hayes DF, Sabel MS, Livant D, Weiss SJ, Rubin MA, Chinnaiyan AM. EZH2 is a marker of aggressive breast cancer and promotes neoplastic transformation of breast epithelial cells. Proc Natl Acad Sci USA. 2003;100:11606–11611. doi:10.1073/pnas.1933744100
  131. 131. Martinez‐Garcia E, Licht JD. Deregulation of H3K27 methylation in cancer. Nat Genet.2010;42:100–101. doi:10.1038/ng0210‐100
  132. 132. Yu J, Yu J, Rhodes DR, Tomlins SA, Cao X, Chen G, Mehra R, Wang X, Ghosh D, Shah RB, Varambally S, Pienta KJ, Chinnaiyan AM. A polycomb repression signature in metastatic prostate cancer predicts cancer outcome. Cancer Res. 2007;67(22):10657–10663. doi:10.1158/0008‐5472.CAN‐07‐2498
  133. 133. van Haaften G, Dalgliesh GL, Davies H, Chen L, Bignell G, Greenman C, Edkins S, Hardy C, O'Meara S, Teague J, Butler A, Hinton J, Latimer C, Andrews J, Barthorpe S, Beare D, Buck G, Campbell PJ, Cole J, Forbes S, Jia M, Jones D, Kok CY, Leroy C, Lin ML, McBride DJ, Maddison M, Maquire S, McLay K, Menzies A, Mironenko T, Mulderrig L, Mudie L, Pleasance E, Shepherd R, Smith R, Stebbings L, Stephens P, Tang G, Tarpey PS, Turner R, Turrell K, Varian J, West S, Widaa S, Wray P, Collins VP, Ichimura K, Law S, Wong J, Yuen ST, Leung SY, Tonon G, DePinho RA, Tai YT, Anderson KC, Kahnoski RJ, Massie A, Khoo SK, Teh BT, Stratton MR, Futreal PA. Somatic mutations of the histone H3K27 demethylase gene UTX in human cancer. Nat Genet. 2009;41(5):521–523. doi:10.1038/ng.349
  134. 134. de Capoa A, Musolino A, Della Rosa S, Caiafa P, Mariani L, Del Nonno F, Vocaturo A, Donnorso RP, Niveleau A, Grappelli C. DNA demethylation is directly related to tumour progression: evidence in normal, pre‐malignant and malignant cells from uterine cervix samples. Oncol Rep. 2003;10:545–549. doi:10.3892/or.10.3.545
  135. 135. Bernstein BE, Humphrey EL, Erlich RL, Schneider R, Bouman P, Liu JS, Kouzarides T, Schreiber SL. Methylation of histone H3 Lys 4 in coding regions of active genes. Proc Natl Acad Sci USA. 2002;99:8695–8700. doi:10.1073/pnas.082249499
  136. 136. Huang BH, Laban M, Leung CH, Lee L, Lee CK, Salto‐Tellez M, Raju GC, Hooi SC. Inhibition of histone deacetylase 2 increases apoptosis and p21Cip1/WAF1 expression, independent of histone deacetylase 1. Cell Death Differ. 2005;12:395–404. doi:10.1038/sj.cdd.4401567
  137. 137. Danam RP, Howell SR, Brent TP, Harris LC. Epigenetic regulationof O6‐methylguanine‐DNA methyltransferase gene expression byhistone acetylation and methyl‐CpG binding proteins. Mol Cancer Ther. 2005;4:61–69.
  138. 138. Jaehyouk L, Young Soo Y, Jae Hoon C. Epigenetic silencing of the WNT antagonist DICKKOPF‐1 in cervical cancer cell lines. Gynecol Oncol. 2008;109:270–274. doi:10.1016/j.ygyno.2008.01.034
  139. 139. Bardos JI, Ashcroft M. Negative and positive regulation of HIF‐1: acomplex network. Biochim Biophys Acta. 2005;1755:107–120. doi:10.1016/j.bbcan.2005.05.001
  140. 140. Bodily JM, Mehta KP, Laimins LA. Human papillomavirus e7enhances hypoxia‐inducible factor 1 mediated transcription byinhibiting binding of histone deacetylases. Cancer Res. 2011;71:1187–1195. doi:10.1158/0008‐5472.CAN‐10‐2626
  141. 141. Lu TY, Kao CF, Lin CT, Huang DY, Chiu CY, Huang YS, Wu HC. DNA methylation and histone modification regulate silencing of OPG during tumor progression. J Cell Biochem. 2009;108:315–325. doi:10.1002/jcb.22256
  142. 142. Zhang Z, Joh K, Yatsuki H, Zhao W, Soejima H, Higashimoto K, Noguchi M, Yokoyama M, Iwasaka T, Mukai T. Retinoic acid receptor beta2 is epigenetically silenced either by DNA methylation or repressive histone modifications at the promoter in cervical cancer cells. Cancer Lett. 2007;247:318–327. doi:10.1016/j.canlet.2006.05.013
  143. 143. Seligson DB, Horvath S, Shi T, Yu H, Tze S, Grunstein M, Kurdistani SK. Global histone modification patterns predict risk of prostate cancer recurrence. Nature. 2005;435:1262–1266. doi:10.1038/nature03672
  144. 144. Anton M, Horký M, Kuchtícková S, Vojtĕsek B, Bláha O. Immunohistochemical detection of acetylation and phosphorylation of histone H3 in cervical smears. Ceska Gynekol. 2004;69:3–6.
  145. 145. Vigushin DM, Coombes RC. Targeted histone deacetylase inhibition for cancer therapy. Curr Cancer Drug Targets. 2004;4:205–218. doi:10.2174/1568009043481560
  146. 146. Brehm A, Nielsen SJ, Miska EA, McCance DJ, Reid JL, Bannister AJ, Kouzarides T. The E7 oncoprotein associates with Mi2 and histone deacetylase activity to promote cell growth. EMBO J. 1999;18:2449–2458. doi:10.1093/emboj/18.9.2449
  147. 147. Zimmermann H, Degenkolbe R, Bernard HU, O'Connor MJ. The human papillomavirus type 16 E6 oncoprotein can downregulate p53 activity by targeting the transcriptional coactivator CBP/p300. J Virol. 1999;73:6209–6219.
  148. 148. Mosammaparast N, Shi Y. Reversal of histone methylation: biochemical and molecular mechanisms of histone demethylases. Annu Rev Biochem. 2010;79:155–179. doi:10.1146/annurev.biochem.78.070907.103946
  149. 149. Van Rechem C, Whetstine JR. Examining the impact of gene variants on histone lysine methylation. Biochim Biophys Acta. 2014;1839(12):1463–1476. doi:10.1016/j.bbagrm.2014.05.014
  150. 150. Fischle W, Wang Y, Allis CD. Histone and chromatin cross talk. Curr Opin Cell Biol. 2003;15:172–183. doi:10.1016/S0955‐0674(03)00013‐9
  151. 151. Agger K, Christensen J, Cloos PA, Helin K. The emerging functions of histone demethylases. Curr Opin Genet Dev. 2008;18:159–168. doi:10.1016/j.gde.2007.12.003
  152. 152. Bernat A, Avvakumov N, Mymryk JS, Banks L. Interaction between the HPV E7 oncoprotein and the transcriptional coactivator p300. Oncogene. 2003;22:7871–7881. doi:10.1038/sj.onc.1206896
  153. 153. Vaquero A, Loyola A, Reinberg D. The constantly changing face of chromatin. Sci Aging Knowledge Environ. 2003;14:RE4. doi:10.1126/sageke.2003.14.re4
  154. 154. Peterson CL, Laniel MA. Histones and histone modifications. Curr Biol.2004;14:546–551. doi:10.1016/j.cub.2004.07.007
  155. 155. Longworth MS, Laimins LA. The binding of histone deacetylases and the integrity of zinc finger‐like motifs of the E7 protein are essential for the life cycle ofhuman papillomavirus type 31. J Virol. 2004;78:3533–3541. doi:10.1128/JVI.78.7.3533‐3541.2004
  156. 156. Avvakumov N, Torchia J, Mymryk JS. Interaction of the HPV E7 proteins with the pCAF acetyltransferase, Oncogene. 2003;22:3833–3841. doi:10.1038/sj.onc.1206562
  157. 157. Varier RA, Timmers HT. Histone lysine methylation and demethylation pathways in cancer. Biochim Biophys Acta. 2010;1815:75–89. doi:10.1016/j.bbcan.2010.10.002
  158. 158. Kampranis SC, Tsichlis PN. Histone demethylases and cancer. Adv Can Res. 2009;102:103–169. doi:10.1016/S0065‐230X(09)02004‐1
  159. 159. McLaughlin‐Drubin ME, Crum CP, Münger K. Human papillomavirus E7 oncoprotein induces KDM6A and KDM6B histone demethylase expression and causes epigenetic reprogramming. Proc Natl Acad Sci USA. 2011;108(5):2130–2135. doi:10.1073/pnas.1009933108
  160. 160. Smith JA, White EA, Sowa ME, Powell ML, Ottinger M, Harper JW, Howley PM. Genome‐wide siRNA screen identifies SMCX, EP400, and Brd4 as E2‐dependent regulators of human papillomavirus oncogene expression. Proc Natl Acad Sci USA. 2010;107(8), 3752–3757. doi:10.1073/pnas.0914818107
  161. 161. Iancu IV, Botezatu A, Plesa A, Huica I, Socolov D, Anton G. Histone lysine demethylases as epigenetic modifiers in HPV‐induced cervical neoplasia. Roman Biotechnol Lett. 2015;20(2):10236–10244
  162. 162. McLaughlin‐Drubin ME, Munger K. Biochemical and functional interactions of human papillomavirus proteins with polycomb group proteins. Viruses. 2013;5:1231–1249. doi:10.3390/v5051231
  163. 163. Khleif SN, DeGregori J, Yee CL, Otterson GA, Kaye FJ, Nevins JR, Howley PM. Inhibition of cyclin D‐CDK4/CDK6 activity is associated with an E2F‐mediated induction of cyclin kinase inhibitor activity. Proc Natl Acad Sci USA. 1996;93:4350–4354.
  164. 164. Barradas M, Anderton E, Acosta JC, Li S, Banito A, Rodriguez‐Niedenführ M, Maertens G, Banck M, Zhou MM, Walsh MJ, Peters G, Gil J. Histone demethylase JMJD3 contributes to epigenetic control of INK4a/ARF by oncogenic RAS. Genes Dev. 2009;23:1177–1182. doi:10.1101/gad.511109
  165. 165. Lao KQ, Tang F, Barbacioru C, Wang Y, Nordman E, Lee C, Xu N, Wang X, Tuch B, Bodeau J, Siddiqui A, Surani MA. mRNA‐sequencing whole transcriptome analysis of a single cell on the SOLiD system. J Biomol Tech. 2009;20(5):266–271.
  166. 166. Birney E, Stamatoyannopoulos JA, Dutta A, Guigó R, Gingeras TR, Margulies EH, Weng Z, Snyder M,Dermitzakis ET, Thurman RE, Kuehn MS, Taylor CM, Neph S, Koch CM, Asthana S, Malhotra A, Adzhubei I, Greenbaum JA,Andrews RM, Flicek P, Boyle PJ, Cao H “Identification and analysis of functional elements in 1% of the human genome by the ENCODE pilot project.” Nature. 2007;447(7146):799–816. doi:10.1038/nature05874
  167. 167. Dinger ME, Amaral PP, Mercer TR, Pang KC, Bruce SJ, Gardiner BB, Askarian‐Amiri ME, Ru K, Soldà G, Simons C, Sunkin SM,Crowe ML, Grimmond SM, Perkins AC, Mattick JS. Long noncoding RNAs in mouse embryonic stem cell pluripotency and differentiation. Genome Res. 2008;18:1433–1445. doi:10.1101/gr.078378.108
  168. 168. Kaikkonen MU, Lam MTY, Glass CK. Non‐coding RNAs as regulators of gene expression and epigenetics. Cardiovasc Res. 2011;90:430–440. doi:10.1093/cvr/cvr097
  169. 169. Amaral PP, Dinger ME, Mattick JS. Non‐coding RNAs in homeostasis, disease and stress responses: an evolutionary perspective. Brief Funct Genom. 201;12(3):254–278. doi:10.1093/bfgp/elt016
  170. 170. Nagano T, Fraser P. No‐nonsense functions for long noncoding RNAs. Cell. 2011;145(2):178–181. doi:10.1016/j.cell.2011.03.014
  171. 171. Lai EC. miRNAs: whys and wherefores of miRNA‐mediated regulation. Curr Biol. 2005;15(12):R458–R460. doi:10.1016/j.cub.2005.06.015
  172. 172. Gregory R, Chendrimada T, Shiekhattar R. “MicroRNA biogenesis: isolation and characterization of the microprocessor complex.” Methods Mol Biol. 2006;342:33–47. doi:10.1385/1‐59745‐123‐1:33
  173. 173. Pillai RS, Bhattacharyya SN, Filipowicz W. “Repression of protein synthesis by miRNAs: how many mechanisms?” Trends Cell Biol. 2007;17(3):118. doi:10.1016/j.tcb.2006.12.007
  174. 174. Calin GA, Croce CM. “MicroRNA signatures in human cancers.” Nat Rev Cancer. 2006;6(11):857–866. doi:10.1038/nrc1997
  175. 175. Bartel DP. “MicroRNAs: genomics, biogenesis, mechanism, and function.” Cell. 2004;116(2):281–297. doi:10.1016/S0092‐8674(04)00045‐5
  176. 176. Kent OA, Mendell JT. A small piece in the cancer puzzle: microRNAs as tumor suppressors and oncogenes. Oncogene. 2006;25:6188–6196. doi:10.1038/sj.onc.1209913
  177. 177. Cimmino A Calin GA, Fabbri M, Iorio MV, Ferracin M, Shimizu M, Wojcik SE, Aqeilan RI, Zupo S, Dono M, Rassenti L, Alder H,Volinia S, Liu CG, Kipps TJ, Negrini M, Croce CM. miR‐15 and miR‐16 induce apoptosis by targeting BCL2. Proc Natl Acad Sci USA. 2005;102:13944–13949. doi:10.1073/pnas.0506654102
  178. 178. Sanchez‐Beato M, Sanchez‐Aguilera A, Piris MA. Cell cycle deregulation in B‐cell lymphomas. Blood. 2003;101:1220–1235. doi:10.1182/blood‐2002‐07‐2009
  179. 179. 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 USA. 2005;102(10):3627–3632. doi:10.1073/pnas.0500613102
  180. 180. 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):4944–4951. doi:10.1182/blood‐2006‐12‐062398
  181. 181. Merkel O, Hamacher F, Griessl R, Grabner L, Schiefer AI, Prutsch N, Baer C, Egger G, Schlederer M, Krenn PW, Hartmann TN, Simonitsch‐Klupp I, Plass C, Staber PB, Moriggl R, Turner SD, Greil R, Kenner L. Oncogenic role of miR‐155 in anaplastic large cell lymphoma lacking the t(2;5) translocation. J Pathol. 2015;236(4):445–456. doi:10.1002/path.4539
  182. 182. 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):7065–7070. doi:10.1158/0008‐5472.CAN‐05‐1783
  183. 183. Concepcion CP, Bonetti C, Ventura A. The microrna‐17‐92 family of microrna clusters in development and disease. Cancer J. 2012;18:262–267. doi:10.1097/PPO.0b013e318258b60a
  184. 184. 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;65:9628–9632. doi:10.1158/0008‐5472.CAN‐05‐2352
  185. 185. Li P, Xu Q, Zhang D, Li X, Han L, Lei J, Duan W, Ma Q, Wu Z, Wang Z. Upregulated miR‐106a plays an oncogenic role in pancreatic cancer. FEBS Lett. 2014;588(5):705–712. doi:10.1016/j.febslet.2014.01.007
  186. 186. Díaz R, Silva J, García JM, Lorenzo Y, García V, Peña C, Rodríguez R, Muñoz C, García F, Bonilla F, Domínguez G. Deregulated expression of miR‐106a predicts survival in human colon cancer patients. Genes Chromosomes Cancer. 2008;47(9):794–802. doi:10.1002/gcc.20580
  187. 187. Lum AM, Wang BB, Li L, Channa N, Bartha G, Wabl M. Retroviral activation of the mir‐106a microRNA cistron in T lymphoma. Retrovirology. 2007;4:5. doi:10.1186/1742‐4690‐4‐5
  188. 188. Huang GL, Zhang XH, Guo GL, Huang KT, Yang KY, Shen X, You J, Hu XQ. Clinical significance of miR‐21 expression in breast cancer: SYBR‐Green I‐based real‐time RT‐PCR study of invasive ductal carcinoma. Oncol Rep. 2009;21(3):673–679. doi:10.3892/or_00000270
  189. 189. Yan LX, Huang XF, Shao Q, Huang MY, Deng L, Wu QL, Zeng YX, Shao JY. MicroRNA miR‐21 overexpression in human breast cancer is associated with advanced clinical stage, lymph node metastasis and patient poor prognosis. RNA. 2008;14(11):2348–2360. doi:10.1261/rna.1034808
  190. 190. Frezzetti D, De Menna M, Zoppoli P, Guerra C, Ferraro A, Bello AM, De Luca P, Calabrese C, Fusco A, Ceccarelli M, Zollo M,Barbacid M, Di Lauro R, De Vita G. Upregulation of mir‐21 by ras in vivo and its role in tumor growth. Oncogene. 2011;30:75–286. doi:10.1038/onc.2010.416
  191. 191. Capodanno A, Boldrini L, Proietti A, Alì G, Pelliccioni S, Niccoli C, D'Incecco A, Cappuzzo F, Chella A, Lucchi M, Mussi A, Fontanini G. Let‐7g and miR‐21 expression in non‐small cell lung cancer: correlation with clinicopathological and molecular features. Int J Oncol. 2013;43(3):765–774. doi:10.3892/ijo.2013.2003
  192. 192. 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):2128–2136. doi:10.1038/sj.onc.1210856
  193. 193. Meng F, Henson R, Wehbe‐Janek H, Ghoshal K, Jacob ST, Patel T. “MicroRNA‐21 regulates expression of the PTEN tumor suppressor gene in human hepatocellular cancer”. Gastroenterology. 2007;133(2):647–658. doi:10.1053/j.gastro.2007.05.022
  194. 194. Huang CS, Yu W, Cui H, Wang YJ, Zhang L, Han F, Huang T. Increased expression of miR‐21 predicts poor prognosis in patients with hepatocellular carcinoma. Int J Clin Exp Pathol. 2015;8(6):7234–7238. eCollection 2015.
  195. 195. Yang CH, Yue J, Pfeffer SR, Fan M, Paulus E, Hosni‐Ahmed A, Sims M, Qayyum S, Davidoff AM, Handorf CR, Pfeffer LM. MicroRNA‐21 promotes glioblastoma tumorigenesis by down‐regulating insulin‐like growth factor‐binding protein‐3 (IGFBP3). J Biol Chem. 2014;289(36):25079–25087. doi:10.1074/jbc.M114.593863. Epub 2014 July 24.
  196. 196. Gabriely G, Wurdinger T, Kesari S, Esau CC, Burchard J, Linsley PS, Krichevsky AM. “MicroRNA 21 promotes glioma invasion by targeting matrix metalloproteinase regulators.” Mol Cell Biol. 2008;28(17):5369–5380. doi:10.1128/MCB.00479‐08
  197. 197. Rivas MA, Venturutti L, Huang YW, Schillaci R, Huang TH, Elizalde PV. Downregulation of the tumor‐suppressor miR‐16 via progestin‐mediated oncogenic signaling contributes to breast cancer development. Breast Cancer Res. 2012;14(3):R77. doi:10.1186/bcr3187
  198. 198. Bottoni A, Piccin D, Tagliati F, Luchin A, Zatelli MC, degli Uberti EC. miR‐15a and miR‐16‐1 down‐regulation in pituitary adenomas. J Cell Physiol. 2005;204:280–285. doi:10.1002/jcp.20282
  199. 199. Nicoloso MS, Kipps TJ, Croce CM, Calin GA. MicroRNAs in the pathogeny of chronic lymphocytic leukaemia. Br J Haematol. 2007;139:709–716. doi:10.1111/j.1365‐2141.2007.06868.x
  200. 200. Bonci D, Coppola V, Musumeci M, Addario A, Giuffrida R, Memeo L, D'Urso L, Pagliuca A, Biffoni M, Labbaye C, Bartucci M, Muto G,Peschle C, De Maria R. The miR‐15a‐miR‐16‐1 cluster controls prostate cancer by targeting multiple oncogenic activities. Nat Med. 2008;14:1271–1277. doi:10.1038/nm.1880
  201. 201. Chang TC, Wentzel EA, Kent OA, Ramachandran K, Mullendore M, Lee KH, Feldmann G, Yamakuchi M, Ferlito M, Lowenstein CJ. Transactivation of miR‐34a by p53 broadly influences gene expression and promotes apoptosis. Mol Cell. 2007;26(5):745–752. doi:10.1016/j.molcel.2007.05.010
  202. 202. Tazawa H, Tsuchiya N, Izumiya M, Nakagama H. Tumor‐suppressive miR‐34a induces senescence‐like growth arrest through modulation of the E2F pathway in human colon cancer cells. Proc Natl Acad Sci USA. 2007;104(39):15472–15477. doi:10.1073/pnas.0707351104
  203. 203. Li R, Shi X, Ling F, Wang C, Liu J, Wang W, Li M. MiR‐34a suppresses ovarian cancer proliferation and motility by targeting AXL. Tumour Biol. 2015;36(9):7277–7283. doi:10.1007/s13277‐015‐3445‐8
  204. 204. Furuta M, Kozaki KI, Tanaka S, Arii S, Imoto I, Inazawa J. miR‐124 and miR‐203 are epigenetically silenced tumor‐suppressive microRNAs in hepatocellular carcinoma. Carcinogenesis. 2010;31:766–776. doi:10.1093/carcin/bgp250
  205. 205. Agirre X, Vilas‐Zornoza A, Jiménez‐Velasco A, Martin‐Subero JI, Cordeu L, Gárate L, San José‐Eneriz E, Abizanda G, Rodríguez‐Otero P, Fortes P, Rifón J, Bandrés E, Calasanz MJ, Martín V, Heiniger A, Torres A, Siebert R, Román‐Gomez J, Prósper F. Epigenetic silencing of the tumor suppressor microRNA Hsa‐miR‐124a regulates CDK6 expression and confers a poor prognosis in acute lymphoblastic leukemia. Cancer Res. 2009;69(10):4443–4453. doi:10.1158/0008‐5472.CAN‐08‐4025
  206. 206. Baroukh N, Ravier MA, Loder MK, Hill EV, Bounacer A, Scharfmann R, Rutter GA, Van Obberghen E. MicroRNA‐124a regulates Foxa2 expression and intracellular signaling in pancreatic beta‐cell lines. J Biol Chem. 2007;282:19575–19588. doi:10.1074/jbc.M611841200
  207. 207. Tsai WC, Hsu PW, Lai TC, Chau GY, Lin CW, Chen CM, Lin CD, Liao YL, Wang JL, Chau YP, Hsu MT, Hsiao M, Huang HD, Tsou AP. MicroRNA‐122, a tumor suppressor microRNA that regulates intrahepatic metastasis of hepatocellular carcinoma. Hepatology. 2009;49(5):1571–1582. doi:10.1002/hep.22806
  208. 208. Bueno MJ, Pérez de Castro I, Gómez de Cedrón M, Santos J, Calin GA, Cigudosa JC, Croce CM, Fernández‐Piqueras J, Malumbres M. Genetic and epigenetic silencing of microRNA‐203 enhances ABL1 and BCR‐ABL1 oncogene expression. Cancer Cell. 2008;13(6):496–506. doi:10.1016/j.ccr.2008.04.018
  209. 209. Wang C, Zheng X, Shen C, Shi Y. MicroRNA‐203 suppresses cell proliferation and migration by targeting BIRC5 and LASP1 in human triple‐negative breast cancer cells. J Exp Clin Cancer Res. 2012;31:58. doi:10.1186/1756‐9966‐31‐58
  210. 210. Viticchiè G, Lena AM, Latina A, Formosa A, Gregersen LH, Lund AH, Bernardini S, Mauriello A, Miano R, Spagnoli LG, Knight RA,Candi E, Melino G. MiR‐203 controls proliferation,migration and invasive potential of prostate cancer cell lines. Cell Cycle. 2011;10:1121–1131.
  211. 211. Wang X, Wang HK, Li Y, Hafner M, Banerjee NS, Tang S, Briskin D, Meyers C, Chow LT, Xie X, Tuschl T, Zheng ZM. microRNAs are biomarkers of oncogenic human papillomavirus infections. Proc Natl Acad Sci USA. 2014;111(11):4262–4267. doi:10.1073/pnas.1401430111
  212. 212. Lui WO, Pourmand N, Patterson BK, Fire A.Patterns of known and novel small RNAs in human cervical cancer. PMID:17616659 Cancer Res. 2007;67(13):6031–6043. doi:10.1158/0008‐5472.CAN‐06‐0561
  213. 213. 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):539–542. doi:10.1016/j.bbrc.2009.08.044
  214. 214. Gocze K, Gombos K, Juhasz K, Kovacs K, Kajtar B, Benczik M, Gocze P, Patczai B, Arany I, Ember I. “Unique microRNA expression profiles in cervical cancer.” Anticancer Res. 2013;33(6):2561–2567
  215. 215. Ribeiro J, Marinho‐Dias J, Monteiro P, Loureiro J, Baldaque I, Medeiros R, Sousa H. miR‐34a and miR‐125b expression in HPV infection and cervical cancer development. Biomed Res Int. 2015;2015:304584. doi:10.1155/2015/304584
  216. 216. Wang X, Tang S, Le SY, Lu R, Rader JS, Meyers C, Zheng ZM. Aberrant expression of oncogenic and tumor‐suppressive microRNAs in cervical cancer is required for cancer cell growth. Plos One. 2008;3(7):e2557. doi:10.1371/journal.pone.0002557
  217. 217. Lao G, Liu P, Wu Q, Zhang W, Liu Y, Yang L, Ma C. Mir‐155 promotes cervical cancer cell proliferation through suppression of its target gene LKB1. Tumour Biol. 2014;35(12):11933–11938. doi:10.1007/s13277‐014‐2479‐7
  218. 218. Liu P, Xin F, Ma CF. Clinical significance of serum miR‐196a in cervical intraepithelial neoplasia and cervical cancer. Genet Mol Res. 2015;14(4):17995–18002. doi:10.4238/2015.December.22.25.
  219. 219. Hou T, Ou J, Zhao X, Huang X, Huang Y, Zhang Y. MicroRNA‐196a promotes cervical cancer proliferation through the regulation of FOXO1 and p27Kip1. Br J Cancer. 2014;110(5):1260–1268. doi:10.1038/bjc.2013.829
  220. 220. Zhao S, Yao DS, Chen JY, Ding N. Aberrant expression of miR‐20a and miR‐203 in cervical cancer. Asian Pac J Cancer Prev. 2013;14(4):2289–2293.
  221. 221. Seifoleslami M, Khameneie MK, Mashayekhi F, Sedaghati F, Ziari K, Mansouri K, Safari A. Identification of microRNAs (miR‐203/miR‐7) as potential markers for the early detection of lymph node metastases in patients with cervical cancer. Tumour Biol. 2015;22. [Epub ahead of print].
  222. 222. Chakrabarti M, Banik, NL, Ray SK. “miR‐138 overexpression is more than hTERT knockdown to potentiate apigenin for apoptosis in neuroblastoma in vitro and in vivo.” Exp Cell Res. 2013;319(10);1575–1585. doi:10.1016/j.yexcr.2013.02.025
  223. 223. Zhao S, Yao D, Chen J, Ding N. Genetic testing and molecular biomarkers. 2013;17(8):631–636. doi:10.1089/gtmb.2013.0085
  224. 224. Du J, Wang L, Li C, Yang H, Li Y, Hu H, Li H, Zhang Z. MicroRNA‐221 targets PTEN to reduce the sensitivity of cervical cancer cells to gefitinib through the PI3K/Akt signaling pathway. Tumour Biol. 2015. [Epub ahead of print].
  225. 225. Wang L, Wang Q, Li HL, Han LY. Expression of miR‐200a, miR‐93, metastasis‐related gene RECK and MMP2/MMP9 in human cervical carcinoma—relationship with prognosis. Asian Pac J Cancer Prev. 2013;14:2113–2118. doi:10.7314/APJCP.2013.14.3.2113
  226. 226. Wilting SM, van Boerdonk RA, Henken FE, Meijer CJ, Diosdado B, Meijer GA, le Sage C, Agami R, Snijders PJ, Steenbergen RD. Methylation‐mediated silencing and tumour suppressive function of hsa‐miR‐124 in cervical cancer. Mol Cancer. 2010;9:167–181. doi:10.1186/1476‐4598‐9‐167
  227. 227. Liu L, Yu X, Guo X, Tian Z, Su M, Long Y, Huang C, Zhou F, Liu M, Wu X, Wang X. miR‐143 is downregulated in cervical cancer and promotes apoptosis and inhibits tumor formation by targeting Bcl‐2. Mol Med Rep. 2012;5(3):753–760. doi:10.3892/mmr.2011.696
  228. 228. Wang Q, Qin J, Chen A, Zhou J, Liu J, Cheng J, Qiu J, Zhang J. Downregulation of microRNA‐145 is associated with aggressive progression and poor prognosis in human cervical cancer. Tumour Biol. 2015;36(5):3703–3708. doi:10.1007/s13277‐014‐3009‐3
  229. 229. Rao Q, Shen Q, Zhou H, Peng Y, Li J, Lin Z. Aberrant microRNA expression in human cervical carcinomas. Med Oncol. 2012;29:1242–1248. doi:10.1007/s12032‐011‐9830‐2
  230. 230. Geng D, Song X, Ning F, Song Q, Yin H. MiR‐34a Inhibits viability and invasion of human papillomavirus‐positive cervical cancer cells by targeting E2F3 and regulating survivin. Int J Gynecol Cancer. 2015;25(4):707–713. doi:10.1097/IGC.0000000000000399
  231. 231. Wang X, Wang HK, McCoy JP, Banerjee NS, Rader JS, Broker TR, Meyers C, Chow LT, Zheng ZM. Oncogenic HPV infection interrupts the expression of tumor‐suppressive miR‐34a through viral oncoprotein E6. RNA. 2009;15(4):637–647. doi:10.1261/rna.1442309
  232. 232. Frasco MA, Ayhan A, Zikan M, Cibula D, Iyibozkurt CA, Yavuz E, Hauser‐Kronberger C, Dubeau L, Menon U, Jacobs IJ. HOXA methylation in normal endometrium from premenopausal women is associated with the presence of ovarian cancer: a proof of principle study. Int J Cancer.2009;125(9):2214–2218. doi:10.1002/ijc.24599
  233. 233. Campos‐Viguri GE, Jiménez‐Wences H, Peralta‐Zaragoza O, Torres‐Altamirano G, Soto‐Flores DG, Hernández‐Sotelo D, Alarcón‐Romero Ldel C, Jiménez‐López MA, Illades‐Aguiar B, Fernández‐Tilapa G. miR‐23b as a potential tumor suppressor and its regulation by DNA methylation in cervical cancer. Infect Agent Cancer. 2015;10:42. doi:10.118/s13027‐015‐0037‐6. eCollection 2015
  234. 234. Abdelmohsen K, Kim MM, Srikantan S, Mercken EM, Brennan SE, Wilson GM, de Cabo R, Gorospe M. miR‐519 suppresses tumor growth by reducing HuR levels. Cell Cycle. 2010;9(7):1354–1359. doi:10.4161/cc.9.7.11164
  235. 235. Martinez I, Gardiner AS, Board KF, Monzon FA, Edwards RP, Khan SA. Human papillomavirus type 16 reduces the expression of microRNA‐218 in cervical carcinoma cells. Oncogene. 2008;27(18):2575–2582. doi:10.1038/sj.onc.1210919
  236. 236. Tian RQ, Wang XH, Hou LJ, Jia WH, Yang Q, Li YX, Liu M, Li X, Tang H. MicroRNA‐372 is down‐regulated and targets cyclin‐dependent kinase 2 (CDK2) and cyclin A1 in human cervical cancer, which may contribute to tumorigenesis. J Biol Chem. 2011;286(29):25556–25563. doi:10.1074/jbc.M111.221564
  237. 237. Boss IW, Renne R. Viral miRNAs and immune evasion. Biochim Biophys Acta. 2011;1809(11–12):708–714. doi:10.1016/j.bbagrm.2011.06.012
  238. 238. Cai X, Schäfer A, Lu S, Bilello JP, Desrosiers RC, Edwards R, Raab‐Traub N, Cullen BR. Epstein‐Barr Virus microRNAs are evolutionarily conserved and differentially expressed. Plos Pathog. 2006;2:e23. doi:10.1371/journal.ppat.0020023
  239. 239. Qiu J, Thorley‐Lawson DA. EBV microRNA BART 18‐5p targets MAP3K2 to facilitate persistence in vivo by inhibiting viral replication in B cells. Proc Natl Acad Sci USA. 2014;111(30):11157–11162. doi:10.1073/pnas.1406136111
  240. 240. Wang KC, Chang HY. Molecular mechanisms of long noncoding RNAs. Mol Cell. 2011;43:904–914. doi:10.1016/j.molcel.2011.08.018
  241. 241. Mercer TR, Dinger ME, Mattick JS. Long non‐coding RNAs: insights into functions. Nat Rev Genet. 2009;10(3):155–159. doi:10.1038/nrg2521
  242. 242. Khalil AM, Guttman M, Huarte M, Garber M, Raj A, Rivea Morales D, Thomas K, Presser A, Bernstein BE, van Oudenaarden A,Regev A, Lander ES, Rinn JL. Many human large intergenic noncoding RNAs associate with chromatin‐modifying complexes and affect gene expression. Proc Natl Acad Sci USA. 2009;106:11667–11672. doi:10.1073/pnas.0904715106
  243. 243. Gibb EA, Brown CJ, Lam WL. The functional role of long non‐coding RNA in human carcinomas. Mol Cancer. 2011;10:38. doi:10.1186/1476‐4598‐10‐38
  244. 244. Agrelo R, Wutz A. ConteXt of change—X inactivation and disease. EMBO Mol Med. 2010;2(1):6–15. doi:10.1002/emmm.200900053
  245. 245. DeChiara TM, Robertson EJ, Efstratiadis A. Parental imprinting of the mouse insulin‐like growth factor II gene. Cell. 1991;64(4):849–859. doi:10.1016/0092‐8674(91)90513‐X
  246. 246. Berteaux N, Lottin S, Monte D, Pinte S, Quatannens B, Coll J, Hondermarck H, Curgy JJ, Dugimont T, Adriaenssens E. H19 mRNA‐like noncoding RNA promotes breast cancer cell proliferation through positive control by E2F1. J Biol Chem. 2005;280(33):29625–29636. doi:10.1074/jbc.M504033200
  247. 247. Berteaux N, Lottin S, Monté D, Pinte S, Quatannens B, Coll J, Hondermarck H, Curgy JJ, Dugimont T, Adriaenssens E. The H19 noncoding RNA is essential for human tumor growth. Plos One. 2007;2:e845. doi:10.1074/jbc.M504033200
  248. 248. Yang F, Bi J, Xue X, Zheng L, Zhi K, Hua J, Fang G. Up‐regulated long non‐coding RNA H19 contributes to proliferation of gastric cancer cells. FEBS J. 2012;279:3159–3165. doi:10.1111/j.1742‐4658.2012.08694.x
  249. 249. Kondo M, Suzuki H, Ueda R, Osada H, Takagi K, Takahashi T, Takahashi T. Frequent loss of imprinting of the H19 gene is often associated with its overexpression in human lung cancers. Oncogene. 1995;10(6):1193–1198.
  250. 250. Tanos V, Prus D, Ayesh S, Weinstein D, Tykocinski ML, De‐Groot N, Hochberg A, Ariel I. Expression of the imprinted H19 oncofetal RNA in epithelial ovarian cancer. Eur J Obstet Gynecol Reprod Biol. 1999;85(1):7–11. doi:10.1016/S0301‐2115(98)00275‐9
  251. 251. Ariel I, Miao HQ, Ji XR, Schneider T, Roll D, de Groot N, Hochberg A, Ayesh S. Imprinted H19 oncofetal RNA is a candidate tumour marker for hepatocellular carcinoma. Mol Pathol. 1998;51(1):21–25.
  252. 252. Tsang WP, Ng EK, Ng SS, Jin H, Yu J, Sung JJ, Kwok TT. Oncofetal H19‐derived miR‐675 regulates tumor suppressor RB in human colorectal cancer. Carcinogenesis. 2010;31(3):350–358. doi:10.1093/carcin/bgp181
  253. 253. Ji P, Diederichs S, Wang W, Böing S, Metzger R, Schneider PM, Tidow N, Brandt B, Buerger H, Bulk E, Thomas M, Berdel WE,Serve H, Müller‐Tidow C. MALAT‐1, anovel noncoding RNA, and thymosin beta4 predict metastasis and survival inearly‐stage non‐small cell lung cancer. Oncogene. 2003;22:8031–8041. doi:10.1038/sj.onc.1206928
  254. 254. Lin R., Maeda S, Liu C, Karin M, Edgington TS. A large noncoding RNA is a marker for murine hepatocellular carcinomas and a spectrum of human carcinomas. Oncogene. 2007;26:851–858. doi:10.1038/sj.onc.1209846
  255. 255. Luo JH, Ren B, Keryanov S, Tseng GC, Rao UN, Monga SP, Strom S, Demetris AJ, Nalesnik M, Yu YP, Ranganathan S,Michalopoulos GK. Transcriptomic and genomic analysis of human hepatocellular carcinomas and hepatoblastomas. Hepatology. 2006;44:1012–1024. doi:10.1002/hep.21328
  256. 256. Yamada K, Kano J, Tsunoda H, Yoshikawa H, Okubo C, Ishiyama T, Noguchi M. Phenotypic characterization of endometrial stromal sarcoma of the uterus. Cancer Sci. 2006;97:106–112. doi:10.1111/j.1349‐7006.2006.00147.x
  257. 257. Ren S, Liu Y, Xu W, Sun Y, Lu J, Wang F, Wei M, Shen J, Hou J, Gao X, Xu C, Huang J, Zhao Y, Sun Y. Long noncoding RNA MALAT‐1 is a new potential therapeutic target for castration resistant prostate cancer. J Urol. 2013;190(6):2278–2287. doi:10.1016/j.juro.2013.07.001
  258. 258. Ji Q, Zhang L, Liu X, Zhou L, Wang W, Han Z, Sui H, Tang Y, Wang Y, Liu N, Ren J, Hou F, Li Q. Long non‐coding RNA MALAT1 promotes tumour growth and metastasis in colorectal cancer through binding to SFPQ and releasing oncogene PTBP2 from SFPQ/PTBP2 complex. Br J Cancer. 2014;111(4):736–748. doi:10.1038/bjc.2014.383
  259. 259. Fellenberg J, Bernd L, Delling G, Witte D, Zahlten‐Hinguranage A. Prognostic significance of drug‐regulated genes in high‐gradeosteosarcoma. Mod Pathol.2007;20:1085–1094.
  260. 260. Lai MC, Yang Z, Zhou L, Zhu QQ, Xie HY, Zhang F, Wu LM, Chen LM, Zheng SS. Long non‐coding RNA MALAT‐1 overexpression predicts tumor recurrence of hepatocellular carcinoma after liver transplantation. Med Oncol. 2012;29:1810–1816. doi:10.1007/s12032‐011‐0004‐z
  261. 261. Schmidt LH, Spieker T, Koschmieder S, Schäffers S, Humberg J, Jungen D. The long noncoding MALAT‐1 RNA indicates a poor prognosis in non‐small cell lung cancer and induces migration and tumor growth. J Thorac Oncol. 2011;6(12):1984–1992. doi:10.1097/JTO.0b013e3182307eac
  262. 262. Müller‐Tidow C, Diederichs S, Thomas M, Serve H. Genome‐wide screening for prognosis‐predicting genes in early‐stage non‐small‐cell lung cancer. Lung Cancer. 2004;45:S145–S150. doi:10.1016/j.lungcan.2004.07.979
  263. 263. Mourtada‐Maarabouni M, Pickard MR, Hedge VL, Farzaneh F,Williams GT. GAS5, a non‐protein‐coding RNA, controls apoptosis and is downregulated in breast cancer. Oncogene. 2009;28:195–208. doi:10.1038/onc.2008.373
  264. 264. Pickard MR, Mourtada‐Maarabouni M, Williams GT. Long non‐coding RNA GAS5 regulates apoptosis in prostate cancer cell lines. Biochim Biophys Acta. 2013;1832(10):613–1623. doi:10.1016/j.bbadis.2013.05.005
  265. 265. Tu ZQ, Li RJ, Mei JZ, Li XH. Down‐regulation of long non‐coding RNA GAS5 is associated with the prognosis of hepatocellular carcinoma. Int J Clin Exp Pathol. 2014;7:4303–4309.
  266. 266. Zhou Y, Zhang X, Klibanski A. MEG3 noncoding RNA: a tumor suppressor. J Mol Endocrinol. 2012;48:R45–R53. doi:10.1530/JME‐12‐0008
  267. 267. Wang P, Ren Z, Sun P. Overexpression of the long non‐coding RNA MEG3 impairs in vitro glioma cell proliferation. J Cell Biochem. 2012;113:1868–1874. doi:10.1002/jcb.24055
  268. 268. Sheng X, Li J, Yang L, Chen Z, Zhao Q, Tan L. Promoter hypermethylation influences the suppressive role of maternally expressed 3, a long non‐coding RNA, in the development of epithelial ovarian cancer. Oncol Rep. 2014;32:277–285. doi:10.3892/or.2014.3208
  269. 269. Qin R, Chen Z, Ding Y, Hao J, Hu J, Guo F. Long non‐coding RNA MEG3 inhibits the proliferation of cervical carcinoma cells through the induction of cell cycle arrest and apoptosis. Neoplasma. 2013;60(5)486–492. doi:10.4149/neo_2013_063
  270. 270. Yin DD, Liu ZJ, Zhang E, Kong R, Zhang ZH, Guo RH. Decreased expression of long noncoding RNA MEG3 affects cell proliferation and predicts a poor prognosis in patients with colorectal cancer. Tumour Biol. 2015;36(6):4851–4859. doi:10.1007/s13277‐015‐3139‐2
  271. 271. Lu KH, Li W, Liu XH, Sun M, Zhang ML, Wu WQ, Xie WP, Hou YY. Long non‐coding RNA MEG3 inhibits NSCLC cells proliferation and induces apoptosis by affecting p53 expression. BMC Cancer. 2013;13:461. doi:10.1186/1471‐2407‐13‐461
  272. 272. Ning S, Zhang J, Wang P, Zhi H, Wang J, Liu Y, Gao Y, Guo M, Yue M, Wang L, Li X. Lnc2Cancer: a manually curated database of experimentally supported lncRNAs associated with various human cancers. Nucleic Acids Res. 2016;4;44(D1):D980–D985. doi:10.1093/nar/gkv1094
  273. 273. LncRNADisease database (http://www.cuilab.cn/lncrnadisease)
  274. 274. Quek XC, Thomson DW, Maag JL, Bartonicek N, Signal B, Clark MB, Gloss BS, Dinger ME. lncRNAdb v2.0: expanding the reference database for functional long noncoding RNAs. Nucleic Acid Res. 2014;43:D168–D173. doi:10.1093/nar/gku988
  275. 275. Feigenberg T, Gofrit ON, Pizov G, Hochberg A, Benshushan A. Expression of the h19 oncofetal gene in premalignant lesions of cervical cancer: a potential targeting approach for development of nonsurgical treatment of high‐risk lesions. ISRN Obstet Gynecol. 2013;137509. doi:10.1155/2013/137509
  276. 276. Huang L, Liao LM, Liu AW, Wu JB, Cheng XL, Lin JX, Zheng M. Overexpression of long noncoding RNA HOTAIR predicts a poor prognosis in patients with cervical cancer. Arch Gynecol Obstet. 2014;290(4):717–723. doi:10.1007/s00404‐014‐3236‐2
  277. 277. Sharma S, Mandal P, Sadhukhan T, Roy Chowdhury R, Ranjan Mondal N, Chakravarty B, Chatterjee T, Roy S, Sengupta S. Bridging links between long noncoding RNA HOTAIR and HPV oncoprotein E7 in cervical cancer pathogenesis. Sci Rep. 2015;5:11724. doi:10.1038/srep11724
  278. 278. Kim HJ, Lee DW, Yim GW, Nam EJ, Kim S, Kim SW, Kim YT. Long non‐coding RNA HOTAIR is associated with human cervical cancer progression. Int J Oncol. 2014;46, 521–530. doi:10.3892/ijo.2014.2758
  279. 279. Li J, Wang Y, Yu J, Dong R, Qiu H. A high level of circulating HOTAIR is associated with progression and poor prognosis of cervical cancer. Tumor Biol. 2015;36(3):1661–1665. doi:10.1007/s13277‐014‐2765‐4
  280. 280. Gibb EA, Becker‐Santos DD, Enfield KS, Guillaud M, Niekerk Dv, Matisic JP, Macaulay CE, Lam WL. Aberrant expression of long noncoding RNAs in cervical intraepithelial neoplasia. Int J Gynecol Cancer. 2012;22(9):1557–1563. doi:10.1097/IGC.0b013e318272f2c9
  281. 281. Guo F, Li Y, Liu Y, Wang J, Li Y, Li G. Inhibition of metastasis‐associated lung adenocarcinoma transcript 1 in CaSki human cervical cancer cells suppresses cell proliferation and invasion. Acta Biochim Biophys Sin (Shanghai). 2010;42:224–229. doi:10.1093/abbs/gmq008
  282. 282. Jiang Y, Li Y, Fang S, Jiang B, Qin C, Xie P, Zhou G, Li G. The role of MALAT1 correlates with HPV in cervical cancer. Oncol Lett. 2014;7(6):2135–2141.doi:10.3892/ol.2014.1996
  283. 283. Özgür E, Mert U, Isin M, Okutan M, Dalay N, Gezer U. Differential expression of long non‐coding RNAs during genotoxic stress‐induced apoptosis in HeLa and MCF‐7 cells. Clin Exp Med. 2013;13(2):119–126. doi:10.1007/s10238‐012‐0181‐x
  284. 284. Naemura M, Murasaki C, Inoue Y, Okamoto H, Kotake Y. Long Noncoding RNA ANRIL regulates proliferation of non‐small cell lung cancer and cervical cancer cells. Anticancer Res. 2015;35(10):5377–5382.
  285. 285. Liao LM, Sun XY, Liu AW, Wu JB, Cheng XL, Lin JX, Zheng M, Huang L. Low expression of long noncoding XLOC_010588 indicates a poor prognosis and promotes proliferation through upregulation of c‐Myc in cervical cancer. Gynecol Oncol. 2014;133(3):616–623. doi:10.1016/j.ygyno.2014.03.555
  286. 286. Chen W, Böcker W, Brosius J, Tiedge H. Expression of neural BC200 RNA in human tumours. J Pathol. 1997;183(3):345–351. doi:10.1002/(SICI)1096‐9896(199711)183:3<345::AID‐PATH930>3.0.CO;2‐8
  287. 287. Sun NX, Ye C, Zhao Q, Zhang Q, Xu C, Wang SB, Jin ZJ, Sun SH, Wang F, Li W. Long noncoding RNA‐EBIC promotes tumor cell invasion by binding to EZH2 and repressing E‐cadherin in cervical cancer. Plos One. 2014;9(7):e100340. doi:10.1371/journal.pone.0100340. eCollection 2014.
  288. 288. Cao S, Liu W, Li F, Zhao W, Qin C. Decreased expression of lncRNA GAS5 predicts a poor prognosis in cervical cancer. Int J Clin Exp Pathol. 2014;7:6776–6783.
  289. 289. Jiang S, Wang HL, Yang J. Low expression of long non‐coding RNA LET inhibits carcinogenesis of cervical cancer. Int J Clin Exp Pathol. 2015;8(1):806–811.
  290. 290. Yang M, Zhai X, Xia B, Wang Y, Lou G. Long noncoding RNA CCHE1 promotes cervical cancer cell proliferation via upregulating PCNA. Tumour Biol. 2015;36(10):7615–7622 doi:10.1007/s13277‐015‐3465‐4
  291. 291. Villota C, Campos A, Vidaurre S, Oliveira‐Cruz L, Boccardo E, Burzio VA, Varas M, Villegas J, Villa LL, Valenzuela P, Socias M, Roberts S, Burzio LO. Expression of mitochondrial ncRNAs is modulated by high risk HPV oncogenes. J Biol Chem. 2012;287(25):21303–21315. doi:10.1074/jbc.M111.326694
  292. 292. Johannsen E, Lambert PF. Epigenetics of human papillomaviruses. Virology. 2013;445(1–2):205–212. doi:10.1016/j.virol.2013.07.016
  293. 293. Mirabello L, Schiffman M, Ghosh A, Rodriguez AC, Vasiljevic N, Wentzensen N, Herrero R, Hildesheim A, Wacholder S, Scibior‐Bentkowska D, Burk RD, Lorincz AT. 2013. Elevated methylation of HPV16 DNA is associated with thedevelopment of high grade cervical intraepithelial neoplasia. International journal of cancer. Int J Cancer. 2013;132:1412–1422. doi:10.1002/ijc.27750
  294. 294. Badal V, Chuang LS, Tan EH, Badal S, Villa LL, Wheeler CM, Li BF, Bernard HU. CpG methylation of human papillomavirus type 16 DNA in cervical cancer cell lines and in clinical specimens: genomic hypomethylation correlates with carcinogenic progression. Virology. 2003;77:6227–6234. doi:10.1128/JVI.77.11.6227‐6234.2003
  295. 295. Vinokurova S, von Knebel Doeberitz M. Differential methylation of the HPV 16 upstream regulatory region during epithelial differentiation and neoplastic transformation. Plos One. 2011;6:e24451
  296. 296. Kalantari M, Calleja‐Macías IE, Tewari D, Hagmar B, Lie K, Barrera‐Saldaña HA, Wiley DJ, Bernard HU. Conserved methylation patterns of human papillomavirus type 16 DNA in asymptomatic infection and cervical neoplasia. J Virol. 2004;78:12762–12772. doi:10.1128/JVI.78.23.12762‐12772.2004
  297. 297. Badal S, Badal V, Calleja‐Macías IE, Kalantari M, Chuang LS, Li BF, Bernard HU. The human papillomavirus‐18 genome is efficiently targeted by cellular DNA methylation. Virology. 2004;324:483–492. doi:10.1016/j.virol.2004.04.002
  298. 298. Thain A, Jenkins O, Clarke AR, Gaston K. CpG methylation directly inhibits binding of the human papillomavirus type 16 E2 protein to specific DNA sequences. J Virol. 1996;70:7233–7235.
  299. 299. Saavedra K, Brebi P, Roa JC. Epigenetic alterations in preneoplastic and neoplastic lesions of the cervix. Clin Epigenetics. 2012;4:13, doi:10.1186/1868‐7083‐4‐13
  300. 300. Tharappel LJP, Patel P, Kaur G, Addepalli V. Cervical cancer: a perspective on recent patents. Int J Cancer Res. 2015;11(4):159–163.
  301. 301. Sharma S, Kelly TK, Jones PA. Epigenetics in cancer. Carcinogenesis. 2010;31:27–36. doi:10.1093/carcin/bgp220
  302. 302. Sahasrabuddhe VV, Luhn P, Wentzensen N. Human papillomavirus and cervical cancer: biomarkers for improved prevention efforts. Future Microbiol. 2011;6(9):1083–1098. doi:10.2217/fmb.11.87
  303. 303. Sova P, Feng Q, Geiss G, Wood T, Strauss R, Rudolf V, Lieber A, Kiviat N. Discovery of novel methylation biomarkers in cervical carcinoma by global demethylation and microarray analysis. Cancer Epidemiol Biomark Prev. 2006;15(1):114–123. doi:10.1158/1055‐9965. EPI‐05‐0323
  304. 304. Lai HC, Lin YW, Huang TH, Yan P, Huang RL, Wang HC, Liu J, Chan MW, Chu TY, Sun CA, Chang CC, Yu MH. Identification of novel DNA methylation markers in cervical cancer. Int J Cancer. 2008;123:161–167. doi:10.1002/ijc.23519
  305. 305. Siegel EM, Riggs BM, Delmas AL, KochA, Hakam A, Brown KD. Quantitative DNA methylation analysis of candidate genes in cervical cancer. Plos One. 2015;10(3):e0122495. doi:10.1371/journal.pone.0122495
  306. 306. Guerrero‐Preston RE, Sidransky D, Brebi‐Mieville P. Hypermethylation biomarkers for detection of cervical cancer. Publication No. US 20130109584 B2. 2014. http://www.google.com/patents/US8859468.
  307. 307. Chang CC, Huang RL, Wang HC, Liao YP, Yu MH, Lai HC. High methylation rate of LMX1A, NKX6‐1, PAX1, PTPRR, SOX1, and ZNF582 genes in cervical adenocarcinoma. Int J Gynecol Cancer. 2014;24(2):201–209. doi:10.1097/IGC.0000000000000054
  308. 308. https://www.gbo.com/fileadmin/user_upload/Downloads/Brochures/Brochu
  309. 309. Sahasrabuddhe VV, Luhn P, Wentzensen N. Human papillomavirus and cervical cancer: biomarkers for improved prevention efforts. Future Microbiol. 2011;6(9). doi:10.2217/fmb.11.87
  310. 310. Chen CC, Lee KD, Pai MY, Chu PY, Hsu CC, Chiu CC, Chen LT, Chang JY, Hsia SH, Leu YW. Changes in DNA methylation are associated with the development of drug resistance in cervical cancer cells. Cancer Cell Int. 2015;15:98. doi:10.1186/s12935‐015‐0248‐3
  311. 311. Masuda K, Banno K, Yanokura M, Tsuji K, Kobayashi T, Kisu I, Ueki A, Yamagami W, Nomura H, Tominaga E, Susumu N, Aoki D. Association of epigenetic inactivation of the WRN gene with anticancer drug sensitivity in cervical cancer cells. Oncol Rep. 2012;28:1146–1152. doi:10.3892/or.2012.1912
  312. 312. Iida M, Banno K, Yanokura M, Nakamur K, Adachi Mm Nogami Y, Umene K, Masuda K, Kisu I, Iwata T, Tanaka K, Aoki D. Candidate biomarkers for cervical cancer treatment: potential for clinical practice (Review). Mol Clin Oncol. 2014;2:647–655. doi:10.3892/mco.2014.324
  313. 313. Fernandez AF, Esteller M. Viral epigenomes in human tumorigenesis. Oncogene. 2010;29(10):1405–1420. doi:10.1038/onc.2009.517
  314. 314. Badal V, Chuang LS, Tan EH, Badal S, Villa LL, Wheeler CM, Li BF, Bernard HU. CpG methylation of human papillomavirus type 16 DNA in cervical cancer cell lines and in clinical specimens: genomic hypomethylation correlates with carcinogenic progression. J Virol. 2003;77(11):6227–6234. doi:10.1128/JVI.77.11.6227‐6234.2003
  315. 315. Sun C, Reimers LL, Burk RD. Methylation of HPV16 genome CpG sites is associated with cervix precancer and cancer. Gynecol Oncol. 2011;121(1):59–63. doi:10.1016/j.ygyno.2011.01.013
  316. 316. Mirabello L, Sun C, Ghosh A, Rodriguez AC, Schiffman M, Wentzensen N, Hildesheim A, Herrero R, Wacholder S, Lorincz A, Burk RD. Methylation of human papillomavirus type 16 genome and risk of cervical precancer in a costa rican population.” J Natl Cancer Inst. 2012;104(7):556–565. doi:10.1093/jnci/djs135
  317. 317. Wentzensen N, Sun C, Ghosh A, Kinney W, Mirabello L, Wacholder S, Shaber R, LaMere B, Clarke M, Lorincz AT, Castle PE,Schiffman M, Burk RD. Methylation of HPV18, HPV31, and HPV45 genomes and cervica lintraepithelial neoplasia grade3. JNCI. 2012;104:1738–1749. doi:10.1093/jnci/djs425
  318. 318. Hamamoto R, Nakamura Y, Tsunoda T. Suv39h2 as a target gene for cancer therapy and diagnosis. Publication No. EP2714903 A1. 2014. http://www.google.com/patents/EP2714903A1.
  319. 319. Zheng ZM, Wang X. Regulation of cellular miRNA expression by human papillomaviruses. Biochim Biophys Acta. 2011;1809(11–12):668–677. doi:10.1016/j.bbagrm.2011.05.005
  320. 320. Zhang Y, Zhang D, Wang F, Xu D, Guo Y, Cui W. Serum miRNAs panel (miR‐16‐2*, miR‐195, miR‐2861, miR‐497) as novel non‐invasive biomarkers for detection of cervical cancer. Sci Rep. 2015;5:17942. doi:10.1038/srep17942
  321. 321. Ma Q, Wan G, Wang S, Yang W, Zhang J, Yao X. Serum microRNA‐205 as a novel biomarker for cervical cancer patients. Cancer Cell Int. 2014;14:81. doi:10.1186/s12935‐014‐0081‐0. eCollection 2014.
  322. 322. Jia W, Wu Y, Zhang Q, Gao G, Zhang C, Xiang Y. Expression profile of circulating microRNAs as a promising fingerprint for cervical cancer diagnosis and monitoring. Mol Clin Oncol. 2015;3:851–858. doi:10.3892/mco.2015.560
  323. 323. Li Y, Liu J, Yuan C, Cui B, Zou X, Qiao Y. High‐risk human papillomavirus reduces the expression of microRNA‐218 in women with cervical intraepithelial neoplasia. J Int Med Res. 2010;38:1730–1736. doi:10.1177/147323001003800518
  324. 324. Zhou X, Chen X, Hu L, Han S, Qiang F, Wu Y, Pan L, Shen H, Li Y, Hu Z: Polymorphisms involved in the miR‐218‐LAMB3 pathway and susceptibility of cervical cancer, a case–control study in Chinese women. Gynecol Oncol. 2010;117:287–290. doi:10.1016/j.ygyno.2010.01.020
  325. 325. Banno K, Iida M, Yanokura M, Kisu I, Iwata T, Tominaga E, Tanaka K, Aoki D, MicroRNA in cervical cancer: OncomiRs and tumor suppressor miRs in diagnosis and treatment. Sci World J. 2014;2014:178075. doi:10.1155/2014/178075. eCollection 2014
  326. 326. Shen Y, Wang P, Li Y, Ye F, Wang F, Wan X, Cheng X, Lu W, Xie X. miR‐375 is upregulated in acquired paclitaxel resistance in cervical cancer. Br J Cancer. 2013;109(1):92–99. doi:10.1038/bjc.2013.308
  327. 327. Gadducci, A, Guerrieri ME, Greco C. Tissue biomarkers as prognostic variables of cervical cancer. Crit Rev Oncol Hematol. 2013;86:104–129. doi:10.1016/j.critrevonc.2012.09.003

Written By

Adriana Plesa, Iulia V. Iancu, Anca Botezatu, Irina Huica, Mihai Stoian and Gabriela Anton

Submitted: 29 October 2015 Reviewed: 03 March 2016 Published: 13 July 2016