Open access peer-reviewed chapter

Endogenous Retroelements in Cancer: Molecular Roles and Clinical Approach

Written By

Kang-Hoon Lee and Je-Yoel Cho

Submitted: 11 March 2020 Reviewed: 13 July 2020 Published: 12 August 2020

DOI: 10.5772/intechopen.93370

From the Edited Volume

Methods in Molecular Medicine

Edited by Yusuf Tutar

Chapter metrics overview

549 Chapter Downloads

View Full Metrics

Abstract

Retroelements have been considered as “Junk” DNA although the encyclopedia of DNA elements (ENCODE) project has demonstrated that most of the genome is functional. Since the contribution of LINE1 (L1) and human endogenous retrovirus (HERV) has been suspected to cause human cancers, their regulations and putative molecular functions have been investigated in diverse types of cancer. Their diagnostic, prognostic, and therapeutic potentials have been incessantly proposed using cancer associated or specific properties, such as hypomethylation, increased transcripts, and reverse transcriptase, as well as cancer-associated antigens. This chapter presents the current knowledge on retroelements in various aspects during tumorigenesis and their clinical usage in many cancer studies.

Keywords

  • retrotransposons
  • repetitive elements
  • tumorigenesis
  • cancer
  • LINE
  • HERV
  • retroelement

1. Introduction

In recent decades, the development of genomic analysis technology has played an important role in the study and treatment of various diseases [1, 2]. However, these studies have been focused on genes that form proteins that account for about 1–2% of the entire genome, and the understanding of other parts remains relatively insufficient. A retroelement (RE), also called a retrotransposon, is a type I transposable element that replicates itself via RNA and reverse transcription and can be largely classified into two types based on the genome structure, including long terminal repeat sequences (LTRs). The intact endogenous retrovirus (ERVs) retains two LTRs at both ends of the genome, instead of long and short interspersed nuclear elements (LINE and SINE), which are non-LTR groups. LTRs compose ~8% of the human genome and most are known to be inactive due to accumulated mutations. Yet, interestingly, many are transcriptionally active [3]. The non-LTR groups can be divided again into autonomous LINEs and nonautonomous SINEs that need LINE’s proteins [4]. The LINE1s (L1s), known as the only active REs, makes up ~17% of the human genome. Intact L1s retain ~6 kb of the genome, which encodes two proteins, ORF1 and ORF2, which are essential for replication and reverse transcription [5]. There are about 145 full-length, functional L1 elements in the human genome. On the other hand, SINEs, which are nonautonomous retroelements, have ~300 bp genomes without coding potential. Most SINEs are of the Alu type of which there are over one million copies in the human genome [6].

The association between REs and cancer has been suggested since 1950. As the presence of a viral-oncogene was unveiled and mouse mammary tumor virus (MMTV) became the accepted etiological agent of mammary tumors in mice, the possible carcinogenesis mechanism of ERV was also revealed, raising hope for overcoming cancer [7, 8]. Many studies have reported the association of RE expression with various cancer types, including breast cancer, melanoma, and kidney cancer [9]. However, the function of RE expression in cancer as a driver or passenger remains controversial [10, 11]. It is a chicken and egg situation, since the cancer-associated RE expression can cause malignant cell transformation and malignant cell transformation leads to global DNA hypomethylation, which in turn contributes to oncogenic RE expression [12, 13, 14, 15]. In addition, the fact that most REs have lost their transposition activity due to accumulated mutations makes it difficult to evaluate the role of REs [16]. The RE sequences that occupies about half of the mammalian genome is known as “junk DNA,” and, as the name suggests, little research has been done it [17]. However, in certain areas such as in the early embryogenesis process, degenerative disease, and cancer, the expression of REs have been studied relatively well [18, 19]. In particular, several studies have been conducted to reveal the relationship among the environmental stress, RE responses, and associated diseases [20, 21]. Although no direct relationship has been revealed yet, genome instability by activated RE is known to be the main mechanism linking RE with disease [22]. However, the transposition ratio of all the REs is about 0.02 germline events per generation [23], so it is too rare to explain their various roles.

In this chapter, we focus on the functional mechanisms of REs in various cancers from development to metastasis and from diagnosis to cancer therapy.

Advertisement

2. RE regulation in normal cells and abnormal reactivation and expansion in cancer

Fortunately, except for during the reprogramming process in early stage germ cells, most REs are strongly silenced by diverse epigenomic controls and their reactivation is molecularly inhibited [24, 25].

DNA methylation is a major epigenetic mechanism that contributes to retrotransposon silencing in both normal and cancer cells [26]. In early embryogenesis, a genome-wide DNA methylation is established by the DNA methyltransferase 3 (Dnmt3) and maintained by the methyltransferase1 (Dnmt1) [27]. Parental methylation pattern is genome-wide demethylated and methylated again at imprinted loci and REs by the Dnmt3, and these patterns are maintained by Dnmt1 in somatic cells [28, 29, 30]. Association between demethylation and RE expression was demonstrated in that the inactivation of DNMT3L, which is a non-catalytic homolog of DNMT3A/3B, causes the reactivation of L1 and IAP and leads to meiotic arrest as well as male sterility in male germ cells [31, 32, 33].

In cancer cells, a genome-wide DNA hypomethylation and the reactivation of REs that may result in the loss of chromosomal stability and imprinting patterns are well known [34]. Alteration of L1 methylation has been investigated in many types of cancers, including breast, colon, lung, ovarian, and prostate cancers [35, 36, 37]. Mostly, hypomethylation of the L1 promoter is associated with genome instability, aggressive histology, poor prognosis, and some metastasis [38]. Interestingly, some abnormal features, such as chromosome 8 abnormalities, are also associated with L1 hypomethylation [39]. In addition, due to their prevalent unmethylation in cancer samples, a moderate increase of Alu was also observed in cancer samples with a hypomethylated L1 promoter [40]. Similarly, hypomethylation of HERV has also been reported in various cancer cells [9, 12, 41, 42, 43, 44]. Hypomethylation of its long terminal repeat (LTR) where the promoter is located is associated with its overexpression in cancer [45]. Numerous HERV family members were expressed in cancer cell lines and primary tumor tissues. In a head and neck cancer study, tumor-specific methylation changes were found in HERV-H, HERV-W, and HERV-K families [24, 46]. Similarly, the hypomethylated CpGs resulting in high expression of HERV-K, -W, and L1 was reported in ovarian cancer [47]. Moreover, the hypomethylation of REs has been observed in specific stages or subtypes of cancer, such as during ovarian cancer progression and in the basal subtype of invasive ductal carcinoma breast cancer [48, 49]. Remarkably, individual RE expressions associated with cancer such as HERV-K at 22q11.23 (H22q), HERV-H5, HERV-H48–1, and HERV-E4 are highlighted in various cancers [46, 50, 51]. Their transcripts or viral proteins have been detected in sera from bladder, breast, liver, lung, ovarian, and prostate cancer patients [11].

The last cellular epigenomic regulation mechanism for silencing RE expression is histone modification [52]. In normal spermatogonia, one of the repressive histone modification marks, histone 3 lysine 9 dimethylation (H3K9me2), causes transcriptional repression and is sufficient to maintain L1 silencing in the absence of DNA methylation. Thus, the loss of H3K9me2 combined with the absence of DNA methylation may be the cause of LINE1 activation [53]. On the other hand, in the study of the association of histone modification with RE expression in cancer, two repressive histone modifications, H3K9me3 and H3K27me3, were more enriched at H22q, HERVK17, and L1 sequences in PC3 than in LNCaP prostate cell lines, of which RE expression levels are high and low, respectively. By contrast, the active modification H3K4me3 was the most enriched in LNCaP at the H22q LTR [54].

The expressed RE transcripts can eventually be knocked down by the PIWI system [55]. Piwi-interacting RNA (piRNA) is a well-studied mechanism that contributes to the silencing of REs in many animal germline cells [56, 57]. The piRNA system is a ribonucleoprotein complex consisting of a piRNA, and a P-element-induced wimpy testis (PIWI) subfamily of Argonaut nucleases protein [58]. The piRNA recognizes RE sequences and the PIWI protein destroys the RE transcripts [58, 59]. The piRNA system silences RE expression both at the transcriptional and posttranscriptional levels by modifying repressive chromatin modifications and by cleaving RE transcripts, respectively [57, 60]. However, the role of piRNA in posttranscriptional regulation is not similar to that of miRNA via providing sequence specificity because most piRNA sequences are found not to be complementary to target gene transcripts, suggesting that piRNAs may be involved in epigenetic regulation rather than posttranscriptional regulation of mRNA [61]. The deficient of the piRNA pathway causes overexpression of REs, significantly compromised genome structure and, invariably, germ cell death and sterility [58]. The aberrant expression of piRNAs has been reported in the development of cancer including the proliferation, apoptosis, metastasis, and invasion of cancer cells [62]. Moreover, the high expression of PIWI proteins has been documented in many cancer types, including gastric cancer, liver cancer, intestinal cancer, breast cancer, nonsmall cell lung cancer, bladder cancer, ovarian cancer, and melanoma and is furthermore associated with the aggressiveness of sarcomas, gliomas, and leukemia [61, 63]. The roles of PIWI proteins have been investigated separately in cancer invasion, migration, proliferation, division, and survival [64]. PIWIL1 has been known to induce epithelial-mesenchymal transition and confer migration and invasion of endometrial cancer cells [65]. The association of PIWIL2 via increasing the expression of CDK2 and cyclin A in cancer cells is reported in glioma and nonsmall lung cancer (NSCLC) cells [66]. PIWIL3 promotes the cancer proliferation, migration, and invasion through the JAK2/STAT3 signal pathway [67]. PIWIL4 can promote cancer cell division, migration, and survival of breast cancer by activating TGF-β, MAPK/ERK, and FGF signaling pathways [68].

The apolipoprotein B mRNA editing catalytic polypeptide 3 (APOBEC3) proteins are cytidine deaminases of which family consists of seven family members (APOBEC3-A through -H) with diverse activities against a variety of retroviruses and endogenous REs, even though the activity of L1 suppression does not correlate either with antiviral activity against Vif-deficient HIV-1 and murine leukemia virus, or with patterns of subcellular localization [69, 70]. Thus, the inhibitory effect of APOBEC3 family members, specifically APOBEC3G on L1 transposition might not be due to deaminase activity, but due to novel mechanism(s) [70].

Besides APOBEC3G, MOV10, SAMHD1, and ZAP have all been identified to be able to inhibit L1 activity through diverse mechanisms [71]. MOV10 inhibits L1 mobility through interacting with L1 RNP resulting in L1 transcript degradation [72]. SAMHD1 inhibits the L1 RT activity [73]. ZAP also restricts L1 activity through the loss of L1 transcripts and ribonucleoprotein integrity [74].

Together, it will be a universal explanation for the various epigenomic modifications that are directly associated with both genome-wide RE silencing and reactivation that is much more commonly found in diverse human cancers as frequent as 4–100 de novo insertions per tumor.

Advertisement

3. Roles of RE expressed in cancers

The genomic instability caused by de novo insertions of REs that frequently occur in cancer is the major pathophysiological role accepted by the public [75, 76]. However, this is a very limited explanation of the universal functions of REs, because most REs lose their ability to mobilize [16]. Although some retain their coding potentials, these are silenced tightly by various mechanisms and at various levels, such as epigenomic mechanisms, transcription, and posttranscription [77]. Thus, a more in-depth understanding of RE function is mandatory.

3.1 The source of genome instability

De novo insertions of REs, despite their defective form, can both directly and indirectly affect surrounding human genome sequences [78]. Some of these events occur at high enough frequency to result in vast amounts of rearrangement of the host genome sequence [16]. This does not happen only via the mechanism of transposition activity followed by reintegration but also via the homologous recombination between dispersed REs, resulting in large structural variations (SVs) including duplications, inversions, and deletions [79]. REs are also the source of small SVs such as single-nucleotide variants (SNVs) and short indels, which are caused by template switching during repair of replication errors [16]. The SVs derived from reactivation and expansion of REs via either mobilization activity or homologous recombination have been frequently found in many cancers (~50%) [80, 81]. A high enrichment was reported especially in certain types of cancers, such as esophageal cancers, colon cancers, and squamous cell lung cancers (> 90%) [82]. Although this result indicated that somatic L1 insertions are very frequently found in certain cancers, it is known that a majority of RE somatic integrations are passenger mutations with little or no effect on cancer development [83].

Nevertheless, specific SV loci derived from somatic L1 insertions have also been identified as drivers in most cancer types, including colorectal, breast, lung, and liver cancers [84, 85, 86, 87, 88]. For example, disruption of the APC gene by the insertion of L1 in colon cancer has been well studied [89]. Additionally, a recent study identified driver SV by L1 insertion in liver cancer [90]. L1 integration in the intron of the ST18 gene disrupted a cis-regulatory repressor element, resulting in increased expression of the ST18 gene [84].

Several algorithms have also been developed for the sensitive and precise detection of SVs from the whole genome sequence (WGS) and whole exome sequence (WES) data published in large international consortia such as The International Cancer Genome Consortium (ICGC) and The Cancer Genome Atlas (TCGA), and driver SV events with remarkable functional consequences have been identified [82, 91]. The most SVs were generated by L1 (99%), followed by SINE VNTR Alu (SVA) and ERV [92]. Yet, few retrotranspositions of HERVs have been reported in human cancers [84, 93].

3.2 Epigenomic regulation and reactivation of REs in cancer

Since 1993 when the methylation status of L1 in cancer cells was first measured by Thayer et al., L1 hypomethylation has been reported in many types of human cancers, including prostate, ovarian, head and neck, lung, thyroid, and breast cancer [94, 95]. However, some controversial results showed no changes in L1 methylation levels of cancers including thyroid cancer, renal cancer, lymphoma, and leukemia [96]. This discrepancy may be due to differences in the tumor histological type, because association between L1 hypomethylation and clinical outcome has been demonstrated in melanoma patients. However, the mechanism of L1 hypomethylation effects on aggressive tumor behavior has not been fully investigated [49]. The most likely mechanism is the causing of DNA instability, which has been suspected as the main role of REs [92]. A DNA methyltransferase 1 (Dnmt1) mutation showed substantial genome-wide hypomethylation in all types of tissue and also known to be associated with aggressive T cell lymphomas [97, 98]. Notably, the mutation also showed a high frequency of chromosome 15 trisomy, which suggested that the DNA hypomethylation has a causal role in cancers by promoting genome instability [98]. Another possible mechanism is a dysregulation in transcription level, which activates proto-oncogenes and REs that affect tumor aggressiveness [99]. MicroRNAs, which are closely related to the development of human cancer, can be increased by global DNA hypomethylation, contributing to the acquisition of tumor aggressiveness [100]. In addition, it is possible that the L1 methylation state itself exerts a biological effect. It is known that L1 regulates the function of multiple genes by providing an alternative promoter and contributing to noncoding RNA expression [101, 102]. Therefore, further studies are needed to explain the mechanisms in which L1 hypomethylation affects tumor behavior.

3.3 REs, the origin of cancer associated non-coding transcripts

RNA sequencing using next-generation sequencing technology has provided a large amount of gene expression data in both normal and disease conditions, such as cancer [103]. Growing evidence suggests that REs in the intergenic regions of the human genome are sources of noncoding RNAs, including micro RNAs (miRNAs) and long noncoding RNAs (lncRNAs) [104]. Notably, about 30% of human lncRNAs originate from REs, specifically HERVs. In addition, about 80% of lncRNAs contain RE-originated sequences within or nearby their transcription start sites [105]. Importantly, a recent study has reported that many lncRNAs have a crucial role in a variety of fundamental cellular processes and diseases [106]. A recent study reported that a single-nucleotide polymorphism (SNP) in an L1-containing lncRNA sequence located in an intron of SLC7A2 leads to a decrease in its expression and results in a lethal encephalopathy phenotype [107]. Alu elements, which encode no functional proteins, are also frequently found at multiple locations in lncRNA sequences [108]. Recently, many studies have suggested that Alu sequence in lncRNAs can contribute to the function of lncRNAs. For example, Alu-mediated CDKN1A/p21 transcriptional regulator (APTR) negatively regulates p21 expression by recruiting polycomb repressive proteins to the p21 promoter. The Alu sequence is crucial to the localization of APTR on the p21 promoter that regulates cell growth and proliferation [109].

Despite the limited contribution of L1 and Alu to lncRNAs, a close association between HERVs and ncRNAs was reported by Kelley and Rinn [110]. Hundreds of ncRNAs originated from HERV-H. For example, the lncRNA ROR known to promote the progression of human cancers is one of the ncRNAs promoted by a HERV-H element [111]. Moreover, the lncRNA produced by HERV-K11 directly binds to polypyrimidine tract-binding protein-associated splicing factor (PSF), of which the function is to repress proto-oncogene transcription, reversing the PSF-mediated repression of proto-oncogene transcription and subsequently driving tumorigenesis [46, 112]. Other HERV-related lncRNAs with tumor-suppressive potential have also been identified in the intronic RNAs arising from ERV-9 [45]. It has been reported that its antisense RNA at 3′-untranslated regions was found to physically bind to key transcription factors for cell proliferation such as NF-Y, p53, and sp1. This means that the HERV-related lncRNAs may have a function as decoy targets or traps for the transcription factors resulting in the growth retardation of cancer cells [113].

Another role of RE transcripts related to human disease is to form a complex with the cytoplasmic cDNA of the reactivated RE transcripts to trigger the signal of the inflammatory pathway [23]; for example, RE-derived cytosolic DNA accumulated in Aicardi-Goutières syndrome (AGS) [114]. IFNB1 expression also has an anticorrelation with L1 retrotransposition in cancer cells [115]. Moreover, the study by Ishak et al. showed that mutation of the RB1 gene causes both genome-wide upregulation of L1 expression in somatic cells as well as increased susceptibility to leukemia [116]. Gasche et al. reported that the IL-6 treatment of a cancer cell line induced genome-wide L1 promoter hypomethylation [117]. Altogether, the evidence indicates that REs modify an important aspect of human tumorigenesis.

3.4 RE proteins associated with tumorigenesis

ORF1 and ORF2 in L1 and GAG, POL, and ENV in HERV are proteins encoded by REs that are essential to complete the replication cycle, whereas Alu’s are RNA polymerase III-transcribed sequences without coding potential [118]. Most REs lose their coding potential due to accumulated mutations; however, it is well known that hundreds of L1 are still active to produce two essential proteins, ORF1 (p40, RNA binding protein) and ORF2 (p109, endonuclease and reverse transcriptase activities) [119, 120]. Additionally, although no infectious virus formed by HERVs is reported, multiple protein expressions and their functions have been studied in various HERV families [46]. Most comprehensive studies have reported on envelop proteins (ENV) and their pathogenic properties. The transcripts encoding capsid and protease (GAG) and reverse transcriptase with RNase H domain and integrase (POL) ORFs have been detected in many cells and tissues from both diseased and healthy individuals [121]. Remarkably, HERV-W encodes an ENV protein known as ERVWE1 (Syncytin1), which has been adopted by the human to functionally contribute in placenta biogenesis [122]. Similarly, Syncytin2 encoded by ERVFRD1 is known to have a key role in the implantation of human embryos [123]. Aberrant expression of HERV-W has been known to be associated with various human diseases including cancer [122, 124, 125].

In cancer, an increase in retroviral protein expression was generally detected. Overexpression of L1 ORF1 protein was detected from more than 90% of breast, ovarian, and pancreatic cancers followed by tubular gastrointestinal tract, lung, and prostate cancers (about 50%) [126, 127]. However, the high expression of L1 ORF1p expression is dependent on tumor origin, and it differs case by case even within a similar histological type of cancer. For example, L1 ORF1p is detected in lung adenocarcinoma at greatly varying levels (about 20% are very high, about 30% are moderate, and the rest are undetectable) [128]. Several antibodies targeting ORF2p have recently been produced, and thus, the overexpression of ORF2p was detected in many cancers. Although the functional effects of L1 proteins in human cancers remain unclear in most cancer contexts, this data suggests that L1 proteins are potential cancer biomarkers for the diagnosis of cancer development or the prognosis of clinical outcomes [126, 129]. On the other hand, the HERV-K ENV protein has been identified in various cancer tissues and several different mechanisms by which it associates with tumorigenesis have been proposed [130]. The melanocyte antigen HERV-K-MEL is expressed in about 85% of malignant melanocytes, whereas breast cancer, ovarian cancer, teratocarcinoma, sarcoma, and bladder cancer also express HERV-K ENV [131]. Other HERV families, HERV-E, and ERV3 have also been detected in more than 30% of ovarian cancer patients and are higher in patients with lymph-node-positive breast cancer [11, 132]. Moreover, some antibodies against HERV-K have been detected in serum samples with melanoma [133].

Despite HERVs being known to be incompetent in transposition, studies have shown that the protein-coding potentials can still promote neoplastic properties during tumorigenesis through diverse mechanisms [134]. The oncogenic role of HERV proteins is well investigated with NP9 and REC, which are accessory splice proteins of HERV-K [135]. The transcripts encoding these proteins are overexpressed in many tumors including breast cancers and both are known to interact with the promyelocytic leukemia zinc finger (PLZF) tumor suppressor, which is a transcriptional repressor and epigenetic modulator implicated in cancer. C-myc proto-oncogene is one of the major targets of PLZF. Interaction of NP9 and REC with PLZF abrogates the transcriptional repression of the c-Myc gene promoter, which results in c-Myc overproduction [136]. In addition, the abnormal cell-to-cell fusion activity of HERV-W ENV proteins has been shown to possibly contribute to tumor development and metastasis [130]. Further studies to characterize the expression and molecular functions of these HERV proteins in cancers are demanded.

Advertisement

4. Implementation of REs for cancer diagnosis and prognosis

4.1 Structural variations (SVs) associated with REs in cancer

Identification of somatic mutation hotspots associated with cancer is very important for functional analysis and diagnosis [137]. Several methods have been developed for the identification of somatic RE insertions in cancers (L1-seq, TIPseq, and ERVcaller), and many bioinformatics tools to discover somatic L1 insertions in silico using WGS or WES data have been developed [138, 139]. SVs via L1 insertion associated with cancer have been well investigated in a couple of genes, such as the APC gene that is considered to be a tumor suppressor of colorectal polyposis in colorectal cancer [89]. A potential suppressor of L1, TP53 mutation by L1 insertions, has been observed frequently in tumors. In addition, L1 insertional mutation of MOV10, which is a key L1 suppressor, decreased the expression of the MOV10 in tumors with high L1 insertions [140].

On the other hand, instead of cancer-associated SVs caused by RE insertion, genome variations that might be associated with HERVs or around gene expression in cancer have been identified. Chang et al. identified that four HERVs with mutation hotspots overlapped with exons of four human protein coding genes, which are TNN (HERV-9/LTR12), OR4K15 (HERV-IP10F/LTR10F), ZNF99 (HERV-W/HERV17/LTR17), and KIR2DL1 (MST/MaLR). They also evaluated the effect of each non-synonymous SNV on the survival of kidney cancer patients. Furthermore, they identified 788 HERVs harboring significantly increased the numbers of somatic single-nucleotide variations (SNVs) [141].

4.2 Global hypomethylation in cancer and identification of cancer associated RE methylation

Several studies have shown that global hypomethylation is very common in cancer [142]. The DNA methylation levels of L1 5′-untranslated region (UTR) in cancer have been extensively evaluated for potential use as an epigenomic marker for cancer diagnosis. The level of L1 hypomethylation increases in more advanced cancers; however, other types of REs, such as Alu and HERVs, have been lesser evaluated [143]. Since DNA methylation analysis has some benefits in handling tumor specimens, such as similar efficiency in fresh frozen and formalin-fixed paraffin-embedded tissue, many studies indeed have proposed DNA methylation as a diagnostic marker using fresh tumor biopsies or fixed tissue blocks [144]. Association between L1 hypomethylation and diagnostic and prognostic needs, such as tumor stage group, metastasis, the recurrence rate, and the survival rate, has been studied [145]. Also, L1 hypomethylation has been demonstrated to be a surrogate marker for predicting the response to cancer treatment [146]. Moreover, L1 hypomethylation is observed in very different types of specimen, including blood leukocyte DNA, serum, and oral rinse [147]. Hypomethylation of Alu was reported in several cancers, whereas hypomethylation of HERV-K and HERV-W genomes were found in urothelial cancer and ovarian cancer, respectively [47, 131, 148].

Classically, CpG methylation analyses have been performed in targeted sequence by discriminating between methylated and unmethylated DNA using bisulfite treatment followed by PCR amplification [149]. Although recent nanopore technology can separate between methylated and unmethylated DNA without any treatment, most analyses are usually based on methylation-specific PCR after bisulfite treatment (MSP) [150, 151]. Pyrosequencing-based analysis, specifically methylation-sensitive single-nucleotide-primer extension (MsSNuPE) and Methylight, is a promising method that can be used to reliably measure L1 methylation in paraffin-embedded cancer tissues with higher reproducibility [152]. Using this method, L1 hypomethylation has been tested in various human cancer patients, including gastric cancer, colon cancer, colorectal cancer, melanoma, and breast cancer, and its clinical implications have been suggested [153]. Recent studies have addressed that methylated L1 in circulating cell-free DNA (cfDNA) can be used as a potential prognostic and diagnostic target in cancers, and have promoted its potential as a minimally invasive screening technique. Lee et al. showed L1 hypomethylation in cfDNA of both human breast cancer and dog mammary tumor [154, 155].

Unfortunately, there are not many products in the marketplace that capitalize on the association between RE hypomethylation and diverse cancer types and features, even though many studies have provided evidence for it. Representatively, the only clinical test targeting methylation of L1 is used in the detection of bladder cancer in voided urine [156].

4.3 RE transcripts in cancer diagnosis

First of all, the quantitation of various HERV gene expressions was performed using a real-time PCR. The transcript expression of HERV-H, -K, -P, and -R ENV was significantly increased in the blood of lung cancer patients, and the level was generally much higher in the squamous cell carcinoma (SCC) subtype than the small-cell lung cancer (SCLC) subtype [157]. The level of HERV-K (HML-2) was found to be an independent prognostic factor for the overall survival rate of hepatocellular carcinoma patients [158]. The expression of HERV-H LTR-associating protein 2 (HHLA2) was significantly upregulated in bladder cancer, and it was suggested as a prognostic factor of tumor metastasis and poor survival of bladder cancer patients [159]. The elevated HERV-K (HML-2) was detected in both protein and transcripts level in the blood of breast cancer patients at an early stage and was further increased with developing metastasis. Thus, HERV-K (HML-2) expression will be one of a best candidate for the early detection of an increased risk for breast cancer in women [160]. The expression of HERV-E transcripts is observed in von Hippel-Lindau (VHL)-deficient renal carcinomas. Interestingly, the introduced VHL gene suppressed HERV-E expression in VHL-deficient carcinoma [11]. In addition, high blood levels of the ENV transcripts of various HERV types have been detected in breast cancer patients and that are decreased by treatment of adjuvant chemotherapy which means that alteration of blood HERV transcripts is a very good candidate for diagnosis and is a prognosis marker of breast cancer [132].

4.4 Detection of RE proteins in cancer specimens

A correlation between HERV protein expression and human cancer has been described [11]. HERV proteins, GAG, POL, and ENV, have been identified in cancer tissues, and several factors from environment and hormone response, such as UV radiation, inflammation, estrogen and smoking, have been proposed as a cause of HERV protein expression in various cancer tissues [161]. Remarkably, the envelop protein, ENV, of HERV-K has been identified in melanoma by immunohistochemistry [162]. In melanomas, the expression of HERV-K ENV is higher than that in benign lesions, especially in metastatic tumors. Moreover, it has also been found in other types of cancers, such as breast, ovarian, and bladder cancer. Antibodies targeting HERV-E, HERV-K (HML-2), and ERV3 have also been detected in more than 30% of ovarian cancer patients and are higher in patients with lymph-node-positive breast tumors. In addition, the presence of serum antibodies against HERV-K proteins has been suggested as a prognostic factor for poor survival of melanoma patients [11].

In L1 proteins, high levels of ORF1 protein was prevalent in certain cancers, including breast and ovarian cancer, whereas no or little expression was detected from other cancers such as renal, liver, and cervical cancer [36]. Rodic et al. and Ardeljan et al. separately detected ORF1 protein via IHC in ~90% of ovarian cancer and in ~90% of the breast cancer samples examined [127, 163]. Chen et al. reported that the ORF1 protein level is very high in ductal carcinoma in situ (DCIS) [164]. Moreover, the ORF1 level was the highest in high-grade ovarian carcinoma, but the expression of ORF1 in prostate cancer has not been fully confirmed [36]. Ardeljan et al. reported ORF1 positivity in ~41% of all prostate cancer tissue samples examined [163]. ORF1 levels could be clinically measured using CT scans on the blood of lung cancer patients. On the other hand, ORF2 has only been limitedly tested as a diagnostic marker for cancer when compared to ORF1 expression. However, since ORF2 encodes a reverse transcriptase that is heavily associated with L1 activity, similar to L1 hypomethylation, it may yet be a better diagnostic marker for L1-associated disease development. High expression of ORF2 in transitional colon mucosa but no expression in normal colon mucosa was detected via IHC. ORF2 was also detected in prostate intraepithelial neoplasia [36]. However, since the ORF2 expression has been reported to be much less than that of ORF1, there are challenges to measure it in clinic.

Advertisement

5. RE in cancer therapy

Aberration of RE activities in various aspects has been suggested as a potential target for cancer therapy [165]. Several studies have shown that inhibiting RT activity is a great therapeutic target for cancer. Sciamanna et al., 2005, uncovered that pharmacologic L1 inhibition by two reverse transcriptase inhibitors slows down the growth of malignant melanoma and prostatic cancer [166]. Carlini et al. similarly demonstrated the efficacy of reverse transcription inhibition of prostate cancer growth [167]. Furthermore, a clinical trial showed that Efavirenz, a non-nucleoside reverse transcriptase inhibitor (NNRTI), provides a therapeutic benefit by increasing the progression free survival in a high-stage castration-resistant prostate cancer cohort [168]. Recently, Efavirenz has been shown to suppress L1 activity and promote morphological differentiation in melanoma cells [169]. On the other hand, another class of RT inhibitor, the nucleoside reverse transcriptase inhibitor (NRTI), has also been shown to suppress L1 activity and induce anticancer activity in prostate cancer cell lines. Importantly, no significant effects were observed in normal cells [167]. Despite these successful findings, it is still unclear regarding the molecular function of RT inhibition in the gene expression regulation.

RNA interference (RNAi)-mediated downregulation of L1 generated identical effects to those observed with RT inhibitory drugs in human melanoma, which indicates that RT activity has a crucial role in L1 activity in human cancer [170]. Recently, a phase II human trial using Efavirenz on a cohort of metastatic patients with prostate cancer showed nonprogression when Efavirenz reached an optimal concentration in the blood [171]. Altogether, preclinical and clinical data provide evidence that RT inhibition is a potentially effective tool in a novel anticancer therapy against diverse human cancers with noncytotoxic effects on non-cancer cells [172].

Another approach regarding REs is an immunotherapy approach to target the pro-oncogenic effects of HERV ENV, which is possibly involved in tumor progression and in downstream metastatic spread, in a number of tissues. HERV ENVs exclusively upregulated in tumor tissues will be suitable targets to direct both passive and active immunotherapy against in cancer cells [130]. The antibodies recognizing the HERV ENVs has been developed, and currently, a monoclonal antibody against HERV-K (HML2) ENV successfully inhibits human breast cancer proliferation, with the activation of apoptosis [173]. On the other hand, various HERV-derived ENVs have been investigated as candidates of anticancer immunotherapy, either as tumor-associated or tumor-specific antigens in cancer cells [130]. ERVs were first used for antitumor immunization in the murine cancer models expressing ERV [9]. Similarly, in humans, protective immunity against the HERV-K MEL antigen in melanoma development has been investigated. This active immunotherapy is considered more advantageous with respect to passive immunization [130]. However, despite the antigenic similarity between HERV-K-MEL and yellow fever virus (YFV), no significant protective effects were shown in the 10 years post-anti-YFV vaccinations in the melanoma cohorts [174, 175]. HERV-H ENV (Xp22.3) is an another antigen significantly upregulated in a subset of gastrointestinal cancers. T cells that was sensitized to HERV-H ENV (Xp22.3) had lytic effects against colorectal cancer expressing the ENV. HERV-E ENV showed similar effects in renal carcinoma [130, 176].

In addition, demethylating drugs are commonly used as anticancer agents and are known to trigger RE reexpression [177]. Interestingly, DNA methyl transferase inhibitors are caused by immune attacks that increase the expression of HERV and thereby increase the viral dsRNA [178]. Accordingly, individual knocking down of MDA5, MAVS, or IRF7 inhibits the ability of DNA methyl transferase inhibitor to target colorectal cancers resulting in significantly reduced the anticancer activity [179]. Altogether, immunotherapy approaches targeting HERV ENV in a broad spectrum of cancers might be valuable for the expansion of target cancers and for use with other cancer therapies.

Advertisement

6. Conclusions

In this chapter, we reviewed and summarized the functions and regulatory mechanisms of retroelements in the development and progression of cancers, and further presented applications in the development of diagnosis and treatment targets using these characteristics (Table 1). We looked at the retrovirus as a functional genomic element that forms the genome, not as an ancient infected virus and its useless remnants. Reactivation of retroelements means that it affects various regulation processes of cells beyond not only controlling the functions of surrounding genes but also increasing the protein formed therefrom or its function, or prompting its reinsertion into a new position. Because of this, it is very important to analyze and understand retroelements’ functions with regard to various target substances, for example, miRNA, transcription factors, epigenetic modifiers, and so on (Figure 1).

RE type Cancer type Experimental technique References
L1 Colon, breast, lung, ovarian, prostate cancer Bisulfite-pyrosequencing [45]
MCA/CpG island microarray
L1 Intrahepatic cholangiocarcinoma Bisulfite-pyrosequencing [48]
L1 Prostate cancer Southern blot analysis [49]
Alu Colon cancer Next-generation sequencing of unmethylated Alu [50]
HERV-K Breast cancer RT-PCR, northern blot, in situ hybridization [51]
HERV-K Breast cancer TCGA RNA-seq, RPPA data anaylsis [52]
HERV-K Melanoma IHC, immunoblotting [53]
Cell fusion-dependent colony formation assay
HERV-K Kidney cancer RT-PCR, northern blot [54]
HERV-K Head and neck cancer Microarray [34]
LI Ovarian cancer Southern hybridization, RT-PCR [57]
HERV-W Ovarian cancer Southern hybridization, RT-PCR [57]
L1 Breast cancer Absolute quantitative assessment of methylated alleles (AQAMA) PCR [58]
HERV-K Urothelial carcinoma RT-PCT, bisulfite-pyrosequencing [60]
L1 Urothelial carcinoma RT-PCT, bisulfite-pyrosequencing [60]
HERV-K Pancreatic cancer RT-PCR, IHC, IF, ELISA, female immunodeficient nude [61]
HERV-K Prostate cancer RT-PCR, bisulfite-pyrosequencing, ChIP [64]
LI Colorectal, blood, brain, prostate, ovarian cancer Tea (TE analyzer) from paired-end, whole-genome sequencing [96]
L1 11 types of cancer Whole genome, exome sequencing [94]
L1 Lung, brain cancer L1-seq [97]
L1 Liver cancer Retrotransposon capture sequencing (RC-seq) [95]
L1 Colon cancer Southern blot, isolation of the fragment containing the insertion [98]
L1 Liver cancer RC-seq, whole genome sequencing [99]
L1 Encephalopathy DNA-seq, RT-PCR [116]
Alu Multiple cancer cell lines RNA immunoprecipitation, RT-PCR [118]
ERV-9 Multiple cancer cell lines RT-PCR, western blot, RNA immunoprecipitation, xenograft models [122]
L1 Multiple cancer cell lines IF, LINE-1 activation assay, RT-PCR [124]
L1 Leukemia ChIP-seq, RNA-seq [125]
L1 Oral cancer Bisulfite-pyrosequencing [126]
HERV-W Testicular cancer HERV GeneChip microarray, bisulfite sequencing PCR [133]
HERV-W Endometrial cancer RT-PCR, DNA-seq, immunoblot [134]
L1 Breast, ovarian, pancreatic, lung, prostate cancer Immunohistochemistry [136]
L1 Colon, prostate cancer Immunoblot, IF, IHC [138]
HERV Breast cancer RT-PCR [141]
HERV-K Teratocarcinoma CRISPR/Cas9, immunoblot [144]
HERV-K Breast cancer GST pull-down assay, Co-IP [145]
L1 Gastrointestinal cancer Tea (TE analyzer) from paired-end whole-genome sequencing, somatic SNV, indel call, RNA-seq for TCGA [149]
HERV Multiple cancer types SNV, DNA functional elements analysis [150]
L1 Liver cancer Bisulfite pyrosequencing [154]
L1 NSCLC Methylation-specific real-time PCR assay [155]
L1 Colon cancer Bisulfite pyrosequencing [162]
L1 Breast cancer Bisulfite sequencing, MSRED, and RT-PCR [163]
HERV Lung cancer RT-PCR [166]
HERV-K Liver cancer RT-PCR [167]
HERV-H Multiple cancer types Immunohistochemistry [168]
HERV-K Breast cancer ELISA, RT-PCR [169]
L1 Multiple cancer types Immunohistochemistry [172]
L1 Breast cancer Western blot, IHC [173]
L1 Melanoma, prostate cancer IF, Western blot, xenograft model [175]
L1 Prostate cancer RT activity assay, RT-PCR [176]
L1 Melanoma IF, RT-PCR, western blot, xenograft model [178]

Table 1.

RE expression in human cancers.

Figure 1.

Overall involvement of REs in cancer studies. RE expression was regulated by epigenomic controls such as histone modification and methylation. Reactivated RE by hypomethylation causes genome instability and the enrichment of cytoplasmic RE transcripts which may increase inflammatory signal. These may be involved in diverse biological process as a source of ncRNA including miRNAs. RE proteins are also involved in tumorigenesis process, and PIWI and APOBEC3 systems regulate RE activity in various ways.

Advertisement

Acknowledgments

This research was supported by the Bio & Medical Technology Development Program of the National Research Foundation (NRF), funded by the Ministry of Science and ICT (#2016M3A9B6026771) and by (NRF-2019R1I1A1A01060265) at least partially. J.Y.C. conceived and developed the entire study and revised the chapter, and K.H.L. mainly wrote the first draft. We thank Hyeon-Ji Hwang for data acquisition, and Johannes Schabort for English editing.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Collins FS et al. A vision for the future of genomics research. Nature. 2003;422(6934):835-847
  2. 2. Hofker MH, Fu J, Wijmenga C. The genome revolution and its role in understanding complex diseases. Biochimica et Biophysica Acta. 2014;1842(10):1889-1895
  3. 3. O'Donnell KA, Burns KH. Mobilizing diversity: Transposable element insertions in genetic variation and disease. Mobile DNA. 2010;1(1):21
  4. 4. Cordaux R, Batzer MA. The impact of retrotransposons on human genome evolution. Nature Reviews. Genetics. 2009;10(10):691-703
  5. 5. Beck CR et al. LINE-1 retrotransposition activity in human genomes. Cell. 2010;141(7):1159-1170
  6. 6. Richardson SR et al. The influence of LINE-1 and SINE retrotransposons on mammalian genomes. Microbiology Spectrum. 2015;3(2) MDNA3-0061-2014
  7. 7. Mason AL, Gilady SY, Mackey JR. Mouse mammary tumor virus in human breast cancer red herring or smoking gun? American Journal of Pathology. 2011;179(4):1588-1590
  8. 8. Kassiotis G. Endogenous retroviruses and the development of cancer. Journal of Immunology. 2014;192(4):1343-1349
  9. 9. Attermann AS et al. Human endogenous retroviruses and their implication for immunotherapeutics of cancer. Annals of Oncology. 2018;29(11):2183-2191
  10. 10. Bannert N, Kurth R. Retroelements and the human genome: New perspectives on an old relation. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(Suppl 2):14572-14579
  11. 11. Gonzalez-Cao M et al. Human endogenous retroviruses and cancer. Cancer Biology & Medicine. 2016;13(4):483-488
  12. 12. Bannert N et al. HERVs new role in cancer: From accused perpetrators to cheerful protectors. Frontiers in Microbiology. 2018;9:178
  13. 13. Yu HL, Zhao ZK, Zhu F. The role of human endogenous retroviral long terminal repeat sequences in human cancer (review). International Journal of Molecular Medicine. 2013;32(4):755-762
  14. 14. Pfeifer GP. Defining driver DNA methylation changes in human cancer. International Journal of Molecular Sciences. 2018;19(4):1166
  15. 15. Jansz N. DNA methylation dynamics at transposable elements in mammals. DNA Methylation. 2019;63(6):677-689
  16. 16. Bourque G et al. Ten things you should know about transposable elements. Genome Biology. 2018;19:199
  17. 17. Palazzo AF, Gregory TR. The case for junk DNA. PLoS Genetics. 2014;10(5):e1004351
  18. 18. Tokuyama M et al. ERVmap analysis reveals genome-wide transcription of human endogenous retroviruses. Proceedings of the National Academy of Sciences of the United States of America. 2018;115(50):12565-12572
  19. 19. Mita P, Boeke JD. How retrotransposons shape genome regulation. Current Opinion in Genetics & Development. 2016;37:90-100
  20. 20. Casacuberta E, Gonzalez J. The impact of transposable elements in environmental adaptation. Molecular Ecology. 2013;22(6):1503-1517
  21. 21. Miousse IR et al. Response of transposable elements to environmental stressors. Mutation Research, Reviews in Mutation Research. 2015;765:19-39
  22. 22. Jung YD et al. Retroelements: Molecular features and implications for disease. Genes & Genetic Systems. 2013;88(1):31-43
  23. 23. Tam OH, Ostrow LW, Gale Hammell M. Diseases of the nervous system: Retrotransposon activity in neurodegenerative disease. Mobile DNA. 2019;10:32
  24. 24. Szpakowski S et al. Loss of epigenetic silencing in tumors preferentially affects primate-specific retroelements. Gene. 2009;448(2):151-167
  25. 25. Walter M et al. An epigenetic switch ensures transposon repression upon dynamic loss of DNA methylation in embryonic stem cells. eLife. 2016;5:e11418
  26. 26. Li E, Zhang Y. DNA methylation in mammals. Cold Spring Harbor Perspectives in Biology. 2014;6(5):a019133
  27. 27. Law JA, Jacobsen SE. Establishing, maintaining and modifying DNA methylation patterns in plants and animals. Nature Reviews. Genetics. 2010;11(3):204-220
  28. 28. Adalsteinsson BT, Ferguson-Smith AC. Epigenetic control of the genome-lessons from genomic imprinting. Genes (Basel). 2014;5(3):635-655
  29. 29. Leung D et al. Regulation of DNA methylation turnover at LTR retrotransposons and imprinted loci by the histone methyltransferase Setdb1. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(18):6690-6695
  30. 30. Jones PA, Liang GN. OPINION rethinking how DNA methylation patterns are maintained. Nature Reviews Genetics. 2009;10(11):805-811
  31. 31. Rowe HM, Trono D. Dynamic control of endogenous retroviruses during development. Virology. 2011;411(2):273-287
  32. 32. Zeng Y, Chen TP. DNA methylation reprogramming during mammalian development. Genes. 2019;10(4):257
  33. 33. Zamudio N et al. DNA methylation restrains transposons from adopting a chromatin signature permissive for meiotic recombination. Genes & Development. 2015;29(12):1256-1270
  34. 34. Costello JF, Plass C. Methylation matters. Journal of Medical Genetics. 2001;38(5):285-303
  35. 35. Estecio MRH et al. LINE-1 hypomethylation in cancer is highly variable and inversely correlated with microsatellite instability. PLoS One. 2007;2(5):e399
  36. 36. Lavasanifar A et al. Long interspersed nuclear element-1 mobilization as a target in cancer diagnostics, prognostics and therapeutics. Clinica Chimica Acta. 2019;493:52-62
  37. 37. Cheung HH et al. DNA methylation of cancer genome. Birth Defects Research Part C-Embryo Today-Reviews. 2009;87(4):335-350
  38. 38. Jeong S et al. Tumoral LINE-1 hypomethylation is associated with poor survival of patients with intrahepatic cholangiocarcinoma. BMC Cancer. 2017;17:588
  39. 39. Schulz WA et al. Factor interaction analysis for chromosome 8 and DNA methylation alterations highlights innate immune response suppression and cytoskeletal changes in prostate cancer. Molecular Cancer. 2007;6:14
  40. 40. Jorda M et al. The epigenetic landscape of Alu repeats delineates the structural and functional genomic architecture of colon cancer cells. Genome Research. 2017;27(1):118-132
  41. 41. Wang-Johanning F et al. Expression of human endogenous retrovirus k envelope transcripts in human breast cancer. Clinical Cancer Research. 2001;7(6):1553-1560
  42. 42. Huang G et al. Human endogenous retroviral K element encodes fusogenic activity in melanoma cells. Journal of Carcinogenesis. 2013;12:5
  43. 43. Cherkasova E et al. Detection of an immunogenic HERV-E envelope with selective expression in clear cell kidney cancer. Cancer Research. 2016;76(8):2177-2185
  44. 44. Johanning GL et al. Expression of human endogenous retrovirus-K is strongly associated with the basal-like breast cancer phenotype. Scientific Reports. 2017;7:41960
  45. 45. Babaian A, Mager DL. Endogenous retroviral promoter exaptation in human cancer. Mobile DNA. 2016;7:24
  46. 46. Zhang M, Liang JQ , Zheng S. Expressional activation and functional roles of human endogenous retroviruses in cancers. Reviews in Medical Virology. 2019;29(2):e2025
  47. 47. Menendez L, Benigno BB, McDonald JF. L1 and HERV-W retrotransposons are hypomethylated in human ovarian carcinomas. Molecular Cancer. 2004;3:12
  48. 48. van Hoesel AQ et al. Hypomethylation of LINE-1 in primary tumor has poor prognosis in young breast cancer patients: A retrospective cohort study. Breast Cancer Research and Treatment. 2012;134(3):1103-1114
  49. 49. Miousse IR, Koturbash I. The fine LINE: Methylation drawing the cancer landscape. BioMed Research International. 2015;2015:131547
  50. 50. Kreimer U et al. HERV-K and LINE-1 DNA methylation and reexpression in urothelial carcinoma. Frontiers in Oncology. 2013;3:255
  51. 51. Li M et al. Downregulation of human endogenous retrovirus type K (HERV-K) viral env RNA in pancreatic cancer cells decreases cell proliferation and tumor growth. Clinical Cancer Research. 2017;23(19):5892-5911
  52. 52. Molaro A, Malik HS. Hide and seek: How chromatin-based pathways silence retroelements in the mammalian germline. Current Opinion in Genetics & Development. 2016;37:51-58
  53. 53. Yang F, Wang PJ. Multiple LINEs of retrotransposon silencing mechanisms in the mammalian germline. Seminars in Cell & Developmental Biology. 2016;59:118-125
  54. 54. Goering W, Ribarska T, Schulz WA. Selective changes of retroelement expression in human prostate cancer. Carcinogenesis. 2011;32(10):1484-1492
  55. 55. Sytnikova YA et al. Transposable element dynamics and PIWI regulation impacts lncRNA and gene expression diversity in Drosophila ovarian cell cultures. Genome Research. 2014;24(12):1977-1990
  56. 56. Ozata DM et al. PIWI-interacting RNAs: Small RNAs with big functions. Nature Reviews. Genetics. 2019;20(2):89-108
  57. 57. Lim AK, Tao L, Kai T. piRNAs mediate posttranscriptional retroelement silencing and localization to pi-bodies in the Drosophila germline. The Journal of Cell Biology. 2009;186(3):333-342
  58. 58. Toth KF et al. The piRNA pathway guards the germline genome against transposable elements. Advances in Experimental Medicine and Biology. 2016;886:51-77
  59. 59. Siomi MC et al. PIWI-interacting small RNAs: The vanguard of genome defence. Nature Reviews. Molecular Cell Biology. 2011;12(4):246-258
  60. 60. Inoue K et al. Switching of dominant retrotransposon silencing strategies from posttranscriptional to transcriptional mechanisms during male germ-cell development in mice. PLoS Genetics. 2017;13(7):e1006926
  61. 61. Cheng Y et al. Emerging roles of piRNAs in cancer: Challenges and prospects. Aging (Albany NY). 2019;11(21):9932-9946
  62. 62. Yu Y, Xiao J, Hann SS. The emerging roles of PIWI-interacting RNA in human cancers. Cancer Management and Research. 2019;11:5895-5909
  63. 63. Maleki Dana P, Mansournia MA, Mirhashemi SM. PIWI-interacting RNAs: New biomarkers for diagnosis and treatment of breast cancer. Cell & Bioscience. 2020;10:44
  64. 64. Liu Y et al. The emerging role of the piRNA/piwi complex in cancer. Molecular Cancer. 2019;18(1):123
  65. 65. Chen Z et al. Stem cell protein Piwil1 endowed endometrial cancer cells with stem-like properties via inducing epithelial-mesenchymal transition. BMC Cancer. 2015;15:811
  66. 66. Qu X et al. PIWIL2 promotes progression of non-small cell lung cancer by inducing CDK2 and Cyclin A expression. Journal of Translational Medicine. 2015;13:301
  67. 67. Jiang L et al. Downregulation of Piwil3 suppresses cell proliferation, migration and invasion in gastric cancer. Cancer Biomarkers. 2017;20(4):499-509
  68. 68. Wang Z et al. The role of PIWIL4, an Argonaute family protein, in breast cancer. The Journal of Biological Chemistry. 2016;291(20):10646-10658
  69. 69. Chiu YL, Greene WC. The APOBEC3 cytidine deaminases: An innate defensive network opposing exogenous retroviruses and endogenous retroelements. Annual Review of Immunology. 2008;26:317-353
  70. 70. Kinomoto M et al. All APOBEC3 family proteins differentially inhibit LINE-1 retrotransposition. Nucleic Acids Research. 2007;35(9):2955-2964
  71. 71. Liang W et al. APOBEC3DE inhibits LINE-1 Retrotransposition by interacting with ORF1p and influencing LINE reverse transcriptase activity. PLoS One. 2016;11(7):e0157220
  72. 72. Choi J, Hwang SY, Ahn K. Interplay between RNASEH2 and MOV10 controls LINE-1 retrotransposition. Nucleic Acids Research. 2018;46(4):1912-1926
  73. 73. Hu S et al. SAMHD1 inhibits LINE-1 retrotransposition by promoting stress granule formation. PLoS Genetics. 2015;11(7):e1005367
  74. 74. Goodier JL et al. The broad-spectrum antiviral protein ZAP restricts human retrotransposition. PLoS Genetics. 2015;11(5):e1005252
  75. 75. Kohnken R, Kodigepalli KM, Wu L. Regulation of deoxynucleotide metabolism in cancer: Novel mechanisms and therapeutic implications. Molecular Cancer. 2015;14:176
  76. 76. Schneider AM et al. Roles of retrotransposons in benign and malignant hematologic disease. Cell. 2009;6(2):121-145
  77. 77. Crichton J et al. Defending the genome from the enemy within: Mechanisms of retrotransposon suppression in the mouse germline. Cellular and Molecular Life Sciences. 2014;71(9):1581-1605
  78. 78. Huang CR, Burns KH, Boeke JD. Active transposition in genomes. Annual Review of Genetics. 2012;46:651-675
  79. 79. Feschotte C, Pritham EJ. DNA transposons and the evolution of eukaryotic genomes. Annual Review of Genetics. 2007;41:331-368
  80. 80. Belancio VP, Roy-Engel AM, Deininger PL. All y’all need to know 'bout retroelements in cancer. Seminars in Cancer Biology. 2010;20(4):200-210
  81. 81. Solyom S, Kazazian HH Jr. Mobile elements in the human genome: Implications for disease. Genome Medicine. 2012;4(2):12
  82. 82. Yi K, Ju YS. Patterns and mechanisms of structural variations in human cancer. Experimental & Molecular Medicine. 2018;50(8):98
  83. 83. Erwin JA et al. L1-associated genomic regions are deleted in somatic cells of the healthy human brain. Nature Neuroscience. 2016;19(12):1583-1591
  84. 84. Scott EC, Devine SE. The role of somatic L1 retrotransposition in human cancers. Viruses. 2017;9(6):131
  85. 85. Helman E et al. Somatic retrotransposition in human cancer revealed by whole-genome and exome sequencing. Genome Research. 2014;24(7):1053-1063
  86. 86. Shukla R et al. Endogenous retrotransposition activates oncogenic pathways in hepatocellular carcinoma. Cell. 2013;153(1):101-111
  87. 87. Lee E et al. Landscape of somatic retrotransposition in human cancers. Science. 2012;337(6097):967-971
  88. 88. Iskow RC et al. Natural mutagenesis of human genomes by endogenous retrotransposons. Cell. 2010;141(7):1253-1261
  89. 89. Miki Y et al. Disruption of the APC gene by a retrotransposal insertion of L1 sequence in a colon cancer. Cancer Research. 1992;52(3):643-645
  90. 90. Schauer SN et al. L1 retrotransposition is a common feature of mammalian hepatocarcinogenesis. Genome Research. 2018;28(5):639-653
  91. 91. Nakagawa H, Fujita M. Whole genome sequencing analysis for cancer genomics and precision medicine. Cancer Science. 2018;109(3):513-522
  92. 92. Hancks DC, Kazazian HH Jr. Roles for retrotransposon insertions in human disease. Mobile DNA. 2016;7:9
  93. 93. Cakmak Guner B et al. Detection of HERV-K6 and HERV-K11 transpositions in the human genome. Biomedical Reports. 2018;9(1):53-59
  94. 94. Thayer RE, Singer MF, Fanning TG. Undermethylation of specific LINE-1 sequences in human cells producing a LINE-1-encoded protein. Gene. 1993;133(2):273-277
  95. 95. Baba Y et al. Long interspersed element-1 methylation level as a prognostic biomarker in gastrointestinal cancers. Digestion. 2018;97(1):26-30
  96. 96. Piskareva O et al. The human L1 element: A potential biomarker in cancer prognosis, current status and future directions. Current Molecular Medicine. 2011;11(4):286-303
  97. 97. Zhang W, Xu J. DNA methyltransferases and their roles in tumorigenesis. Biomarker Research. 2017;5:1
  98. 98. Gaudet F et al. Induction of tumors in mice by genomic hypomethylation. Science. 2003;300(5618):489-492
  99. 99. Lamprecht B, Bonifer C, Mathas S. Repeat-element driven activation of proto-oncogenes in human malignancies. Cell Cycle. 2010;9(21):4276-4281
  100. 100. Veeck J, Esteller M. Breast cancer epigenetics: From DNA methylation to microRNAs. Journal of Mammary Gland Biology and Neoplasia. 2010;15(1):5-17
  101. 101. Chishima T, Iwakiri J, Hamada M. Identification of transposable elements contributing to tissue-specific expression of long non-coding RNAs. Genes (Basel). 2018;9(1):23
  102. 102. Tufarelli C, Cruickshanks HA, Meehan RR. LINE-1 activation and epigenetic silencing of suppressor genes in cancer: Causally related events? Mobile Genetic Elements. 2013;3(5):e26832
  103. 103. Kukurba KR, Montgomery SB. RNA sequencing and analysis. Cold Spring Harbor Protocols. 2015;2015(11):951-969
  104. 104. Zhang X et al. Non-coding RNAs and retroviruses. Retrovirology. 2018;15(1):20
  105. 105. Kapusta A et al. Transposable elements are major contributors to the origin, diversification, and regulation of vertebrate long noncoding RNAs. PLoS Genetics. 2013;9(4):e1003470
  106. 106. Sparber P et al. The role of long non-coding RNAs in the pathogenesis of hereditary diseases. BMC Medical Genomics. 2019;12(Suppl 2):42
  107. 107. Cartault F et al. Mutation in a primate-conserved retrotransposon reveals a noncoding RNA as a mediator of infantile encephalopathy. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(13):4980-4985
  108. 108. Hadjiargyrou M, Delihas N. The intertwining of transposable elements and non-coding RNAs. International Journal of Molecular Sciences. 2013;14(7):13307-13328
  109. 109. Negishi M et al. A new lncRNA, APTR, associates with and represses the CDKN1A/p21 promoter by recruiting polycomb proteins. PLoS One. 2014;9(4):e95216
  110. 110. Kelley D, Rinn J. Transposable elements reveal a stem cell-specific class of long noncoding RNAs. Genome Biology. 2012;13(11):R107
  111. 111. Pan Y et al. The emerging roles of long noncoding RNA ROR (lincRNA-ROR) and its possible mechanisms in human cancers. Cellular Physiology and Biochemistry. 2016;40(1-2):219-229
  112. 112. Wang G et al. Regulation of proto-oncogene transcription, cell proliferation, and tumorigenesis in mice by PSF protein and a VL30 noncoding RNA. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(39):16794-16798
  113. 113. Xu L et al. A novel function of RNAs arising from the long terminal repeat of human endogenous retrovirus 9 in cell cycle arrest. Journal of Virology. 2013;87(1):25-36
  114. 114. Stetson DB et al. Trex1 prevents cell-intrinsic initiation of autoimmunity. Cell. 2008;134(4):587-598
  115. 115. Yu Q et al. Type I interferon controls propagation of long interspersed element-1. The Journal of Biological Chemistry. 2015;290(16):10191-10199
  116. 116. Ishak CA et al. An RB-EZH2 complex mediates silencing of repetitive DNA sequences. Molecular Cell. 2016;64(6):1074-1087
  117. 117. Gasche JA et al. Interleukin-6 promotes tumorigenesis by altering DNA methylation in oral cancer cells. International Journal of Cancer. 2011;129(5):1053-1063
  118. 118. Finnegan DJ. Retrotransposons. Current Biology. 2012;22(11):R432-R437
  119. 119. Jiang F et al. Large-scale transcriptome analysis of retroelements in the migratory locust, Locusta migratoria. PLoS One. 2012;7(7):e40532
  120. 120. Martin SL. The ORF1 protein encoded by LINE-1: Structure and function during L1 retrotransposition. Journal of Biomedicine & Biotechnology. 2006;2006(1):45621
  121. 121. Garcia-Montojo M et al. Human endogenous retrovirus-K (HML-2): A comprehensive review. Critical Reviews in Microbiology. 2018;44(6):715-738
  122. 122. Li F, Karlsson H. Expression and regulation of human endogenous retrovirus W elements. APMIS. 2016;124(1-2):52-66
  123. 123. Soygur B, Sati L. The role of syncytins in human reproduction and reproductive organ cancers. Reproduction. 2016;152(5):R167-R178
  124. 124. Gimenez J et al. Custom human endogenous retroviruses dedicated microarray identifies self-induced HERV-W family elements reactivated in testicular cancer upon methylation control. Nucleic Acids Research. 2010;38(7):2229-2246
  125. 125. Strick R et al. Proliferation and cell-cell fusion of endometrial carcinoma are induced by the human endogenous retroviral Syncytin-1 and regulated by TGF-beta. Journal of Molecular Medicine (Berlin, Germany). 2007;85(1):23-38
  126. 126. Burns KH. Transposable elements in cancer. Nature Reviews. Cancer. 2017;17(7):415-424
  127. 127. Rodic N et al. Long interspersed element-1 protein expression is a hallmark of many human cancers. The American Journal of Pathology. 2014;184(5):1280-1286
  128. 128. Sur D et al. Detection of the LINE-1 retrotransposon RNA-binding protein ORF1p in different anatomical regions of the human brain. Mobile DNA. 2017;8:17
  129. 129. De Luca C et al. Enhanced expression of LINE-1-encoded ORF2 protein in early stages of colon and prostate transformation. Oncotarget. 2016;7(4):4048-4061
  130. 130. Grandi N, Tramontano E. HERV envelope proteins: Physiological role and pathogenic potential in cancer and autoimmunity. Frontiers in Microbiology. 2018;9:462
  131. 131. Cegolon L et al. Human endogenous retroviruses and cancer prevention: Evidence and prospects. BMC Cancer. 2013;13:4
  132. 132. Rhyu DW et al. Expression of human endogenous retrovirus env genes in the blood of breast cancer patients. International Journal of Molecular Sciences. 2014;15(6):9173-9183
  133. 133. Cherkasova E, Weisman Q, Childs RW. Endogenous retroviruses as targets for antitumor immunity in renal cell cancer and other tumors. Frontiers in Oncology. 2013;3:243
  134. 134. Xue B, He L. An expanding universe of the non-coding genome in cancer biology. Carcinogenesis. 2014;35(6):1209-1216
  135. 135. Chan SM et al. The HERV-K accessory protein Np9 controls viability and migration of teratocarcinoma cells. PLoS One. 2019;14(2):e0212970
  136. 136. Denne M et al. Physical and functional interactions of human endogenous retrovirus proteins Np9 and rec with the promyelocytic leukemia zinc finger protein. Journal of Virology. 2007;81(11):5607-5616
  137. 137. Baeissa H et al. Identification and analysis of mutational hotspots in oncogenes and tumour suppressors. Oncotarget. 2017;8(13):21290-21304
  138. 138. Steranka JP et al. Transposon insertion profiling by sequencing (TIPseq) for mapping LINE-1 insertions in the human genome. Mobile DNA. 2019;10:8
  139. 139. Ewing AD. Transposable element detection from whole genome sequence data. Mobile DNA. 2015;6:24
  140. 140. Jung H, Choi JK, Lee EA. Immune signatures correlate with L1 retrotransposition in gastrointestinal cancers. Genome Research. 2018;28(8):1136-1146
  141. 141. Chang TC et al. Investigation of somatic single nucleotide variations in human endogenous retrovirus elements and their potential association with cancer. PLoS One. 2019;14(4):e0213770
  142. 142. Ehrlich M. DNA methylation in cancer: Too much, but also too little. Oncogene. 2002;21(35):5400-5413
  143. 143. Kitkumthorn N, Mutirangura A. Long interspersed nuclear element-1 hypomethylation in cancer: Biology and clinical applications. Clinical Epigenetics. 2011;2:315-330
  144. 144. Locke WJ et al. DNA methylation cancer biomarkers: Translation to the clinic. Frontiers in Genetics. 2019;10:1150
  145. 145. Miyata T et al. Prognostic value of LINE-1 methylation level in 321 patients with primary liver cancer including hepatocellular carcinoma and intrahepatic cholangiocarcinoma. Oncotarget. 2018;9(29):20795-20806
  146. 146. Saito K et al. Long interspersed nuclear element 1 hypomethylation is a marker of poor prognosis in stage IA non-small cell lung cancer. Clinical Cancer Research. 2010;16(8):2418-2426
  147. 147. Barchitta M et al. LINE-1 hypomethylation in blood and tissue samples as an epigenetic marker for cancer risk: A systematic review and meta-analysis. PLoS One. 2014;9(10):e109478
  148. 148. Bakshi A et al. DNA methylation variation of human-specific Alu repeats. Epigenetics. 2016;11(2):163-173
  149. 149. Leontiou CA et al. Bisulfite conversion of DNA: Performance comparison of different kits and methylation quantitation of epigenetic biomarkers that have the potential to be used in non-invasive prenatal testing. PLoS One. 2015;10(8):e0135058
  150. 150. Kurdyukov S, Bullock M. DNA methylation analysis: Choosing the right method. Biology (Basel). 2016;5(1):3
  151. 151. Shim J et al. Detection and quantification of methylation in DNA using solid-state Nanopores. Scientific Reports. 2013;3:1389
  152. 152. Gonzalgo ML, Liang G. Methylation-sensitive single-nucleotide primer extension (Ms-SNuPE) for quantitative measurement of DNA methylation. Nature Protocols. 2007;2(8):1931-1936
  153. 153. Irahara N et al. Precision of pyrosequencing assay to measure LINE-1 methylation in colon cancer, normal colonic mucosa, and peripheral blood cells. The Journal of Molecular Diagnostics. 2010;12(2):177-183
  154. 154. Lee KH et al. Methylation of LINE-1 in cell-free DNA serves as a liquid biopsy biomarker for human breast cancers and dog mammary tumors. Scientific Reports. 2019;9(1):175
  155. 155. Salvi S et al. Cell-free DNA as a diagnostic marker for cancer: Current insights. Oncotargets and Therapy. 2016;9:6549-6559
  156. 156. Martinez VG et al. Epigenetics of bladder cancer: Where biomarkers and therapeutic targets meet. Frontiers in Genetics. 2019;10
  157. 157. Zare M et al. Human endogenous retrovirus env genes: Potential blood biomarkers in lung cancer. Microbial Pathogenesis. 2018;115:189-193
  158. 158. Ma WJ et al. Human endogenous retroviruses-K (HML-2) expression is correlated with prognosis and progress of hepatocellular carcinoma. BioMed Research International. 2016;2016:8201642
  159. 159. Janakiram M et al. Expression, clinical significance, and receptor identification of the newest B7 family member HHLA2 protein. Clinical Cancer Research. 2015;21(10):2359-2366
  160. 160. Wang-Johanning F et al. Human endogenous retrovirus type K antibodies and mRNA as serum biomarkers of early-stage breast cancer. International Journal of Cancer. 2014;134(3):587-595
  161. 161. Downey RF et al. Human endogenous retrovirus K and cancer: Innocent bystander or tumorigenic accomplice? International Journal of Cancer. 2015;137(6):1249-1257
  162. 162. Buscher K et al. Expression of human endogenous retrovirus K in melanomas and melanoma cell lines. Cancer Research. 2005;65(10):4172-4180
  163. 163. Ardeljan D et al. The human long interspersed element-1 retrotransposon: An emerging biomarker of neoplasia. Clinical Chemistry. 2017;63(4):816-822
  164. 164. Chen L et al. Prognostic value of LINE-1 retrotransposon expression and its subcellular localization in breast cancer. Breast Cancer Research and Treatment. 2012;136(1):129-142
  165. 165. Anwar SL, Wulaningsih W, Lehmann U. Transposable elements in human cancer: Causes and consequences of deregulation. International Journal of Molecular Sciences. 2017;18(5):974
  166. 166. Sciamanna I et al. Inhibition of endogenous reverse transcriptase antagonizes human tumor growth. Oncogene. 2005;24(24):3923-3931
  167. 167. Carlini F et al. The reverse transcription inhibitor Abacavir shows anticancer activity in prostate cancer cell lines. PLoS One. 2010;5(12):e14221
  168. 168. Hecht M et al. Efavirenz has the highest anti-proliferative effect of non-nucleoside reverse transcriptase inhibitors against pancreatic cancer cells. PLoS One. 2015;10(6):e0130277
  169. 169. Oricchio E et al. Distinct roles for LINE-1 and HERV-K retroelements in cell proliferation, differentiation and tumor progression. Oncogene. 2007;26(29):4226-4233
  170. 170. Sciamanna I, De Luca C, Spadafora C. The reverse transcriptase encoded by LINE-1 retrotransposons in the genesis, progression, and therapy of cancer. Frontiers in Chemistry. 2016;4:6
  171. 171. Houede N et al. A phase II trial evaluating the efficacy and safety of efavirenz in metastatic castration-resistant prostate cancer. The Oncologist. 2014;19(12):1227-1228
  172. 172. Sciamanna I et al. LINE-1-encoded reverse transcriptase as a target in cancer therapy. Frontiers in Bioscience-Landmark. 2018;23:1360-1369
  173. 173. Wang-Johanning F et al. Immunotherapeutic potential of anti-human endogenous retrovirus-K envelope protein antibodies in targeting breast tumors. Journal of the National Cancer Institute. 2012;104(3):189-210
  174. 174. Hodges-Vazquez M et al. The yellow fever 17D vaccine and risk of malignant melanoma in the United States military. Vaccine. 2012;30(30):4476-4479
  175. 175. Mastrangelo G et al. Does yellow fever 17D vaccine protect against melanoma? Vaccine. 2009;27(4):588-591
  176. 176. Mullins CS, Linnebacher M. Endogenous retrovirus sequences as a novel class of tumor-specific antigens: An example of HERV-H env encoding strong CTL epitopes. Cancer Immunology, Immunotherapy. 2012;61(7):1093-1100
  177. 177. Kong Y et al. Transposable element expression in tumors is associated with immune infiltration and increased antigenicity. Nature Communications. 2019;10:5228
  178. 178. Grandi N, Tramontano E. Human endogenous retroviruses are ancient acquired elements still shaping innate immune responses. Frontiers in Immunology. 2018;9:2039
  179. 179. Roulois D et al. DNA-demethylating agents target colorectal cancer cells by inducing viral mimicry by endogenous transcripts. Cell. 2015;162(5):961-973

Written By

Kang-Hoon Lee and Je-Yoel Cho

Submitted: 11 March 2020 Reviewed: 13 July 2020 Published: 12 August 2020