Next Article in Journal
Standard of Care and Promising New Agents for Triple Negative Metastatic Breast Cancer
Next Article in Special Issue
Epstein-Barr Virus in Gastric Carcinoma
Previous Article in Journal
Desmoplasia and Chemoresistance in Pancreatic Cancer
Previous Article in Special Issue
From Human Papillomavirus (HPV) Detection to Cervical Cancer Prevention in Clinical Practice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Viral Carcinogenesis: Factors Inducing DNA Damage and Virus Integration

Department of Basic Science, Loma Linda University, Loma Linda, CA 92354, USA
*
Author to whom correspondence should be addressed.
Cancers 2014, 6(4), 2155-2186; https://doi.org/10.3390/cancers6042155
Submission received: 31 July 2014 / Revised: 3 October 2014 / Accepted: 9 October 2014 / Published: 22 October 2014
(This article belongs to the Special Issue DNA Viruses in Human Cancer)

Abstract

:
Viruses are the causative agents of 10%–15% of human cancers worldwide. The most common outcome for virus-induced reprogramming is genomic instability, including accumulation of mutations, aberrations and DNA damage. Although each virus has its own specific mechanism for promoting carcinogenesis, the majority of DNA oncogenic viruses encode oncogenes that transform infected cells, frequently by targeting p53 and pRB. In addition, integration of viral DNA into the human genome can also play an important role in promoting tumor development for several viruses, including HBV and HPV. Because viral integration requires the breakage of both the viral and the host DNA, the integration rate is believed to be linked to the levels of DNA damage. DNA damage can be caused by both endogenous and exogenous factors, including inflammation induced by either the virus itself or by co-infections with other agents, environmental agents and other factors. Typically, cancer develops years to decades following the initial infection. A better understanding of virus-mediated carcinogenesis, the networking of pathways involved in transformation and the relevant risk factors, particularly in those cases where tumorigenesis proceeds by way of virus integration, will help to suggest prophylactic and therapeutic strategies to reduce the risk of virus-mediated cancer.

1. Introduction

Viruses are the causative agents of approximately 10%–15% of all cancers worldwide. Viruses that have been linked to carcinogenesis include several DNA viruses: Kaposi’s sarcoma herpesvirus (KSHV), Merkel cell polyomavirus (MCV), Epstein-Barr virus (EBV), Human papillomavirus (HPV), hepatitis B virus (HBV) and the simian virus 40 (SV40), as well as at least two RNA viruses: human T-lymphotropic virus-1 (HTLV-1) and the hepatitis C virus (HCV). These viruses and their associated cancers are shown in Table 1.
The mechanism of virus transmission varies depending on the type of virus, the type of infection, its targets and the associated life cycle. Sexual transmission has been attributed to HBV [1], HCV [2] and HPV [3]. HBV and HCV infections can be also acquired from contaminated blood transfusions [4,5] or can be transmitted through needle sharing between intravenous drug users [5,6]. Perinatal transmission of HBV [7] and HCV [8] has been observed from mother to child during pregnancy or labor, and EBV infections can be transmitted through saliva from one individual to another [9]. SV40 has the potential to be introduced into humans by the poliovirus vaccine [10].
Progression to cancer as a result of infection with an oncogenic virus is usually a rare event. For example, the overall prevalence of high-risk HPV infection is 23% [11]. Most of these infections clear up without any intervention within a few months after acquisition, and about 90% clear within two years. Only 0.3%–1.2% of these initial infections will eventually progress to invasive cervical cancer, according to 2013 World Health Organization (WHO) data [12]. Another example relates to the hepatitis B virus (HBV). The highest rate of infection was found in Sub-Saharan Africa and East Asia, where the prevalence of chronic hepatitis in the adult population is between 5%–10% according to 2014 WHO data [13]. More than 90% of those infected people will recover and be completely cleared of the virus within six months, while less than 5% of infected people will develop chronic hepatitis. Twenty percent of chronic hepatitis B cases will progress to hepatic cirrhosis, and of these, only 5% will progress to hepatocellular carcinoma (HCC). In the case of another agent causing HCC, HCV, approximately 2%–3% (130–170 million) of the worldwide population has been infected with HCV [14]. According to 2014 data from the WHO, chronic HCV infection will develop in 55%–85% of infected persons, with the risk of hepatic cirrhosis being approximately 15%–30% within 20 years [15]. The transformation rate of cirrhosis to HCC is approximately 1%–3% per year [5,16]. According to WHO serological testing data, EBV, another oncogenic virus, is present in approximately 95% of adults worldwide, including those residing in the United States. However, only in very rare cases do these infections progress to Burkitt’s lymphoma or nasopharyngeal carcinoma. Similarly, the prevalence of HTLV-I infection in some endemic areas is 15% [17], while adult T-cell leukemia (ATL) only develops in an estimated 2%–4% of these infected persons in endemic regions, and only where early childhood infection is common [18,19]. Overall, these data indicate that infection with oncogenic viruses does not mean the obligatory development of cancer, although infected individuals can be considered an at-risk population for cancer development. Epidemiological studies demonstrate that risk factors for cancer include virus load, persistence of infection and duration of infection [20,21].
Although these viruses belong to different groups, display different etiologies, carry out a variety of life cycles, target different organs and utilize variable mechanisms to induce cancer, some common features of cancer development can be noted. Overall, these common features are related to virus-mediated and exogenously-derived factors that create favorable conditions for cancer promotion and progression. In this current review, we summarize data on the factors that induce alterations in host genomes that contribute in tumorigenesis. We will also discuss the outcome of DNA damage on viral genome integration, which plays a pivotal role in cancers induced by some viruses.
Table 1. Oncogenic viruses and associated cancers. HTLV-1, human T-lymphotropic virus-1; KSHV, Kaposi’s sarcoma herpesvirus; SV40, simian virus 40.
Table 1. Oncogenic viruses and associated cancers. HTLV-1, human T-lymphotropic virus-1; KSHV, Kaposi’s sarcoma herpesvirus; SV40, simian virus 40.
VirusGenomeAssociated Cancers
HTLV-IPositive-strand, single-stranded RNA retrovirusAdult T-cell leukemia (ATL) [22]. ATL is a malignancy of CD4+ T-lymphocytes, which exhibits severe immunodeficiency and resistance to intensive chemotherapies [23].
HCVPositive-strand, single-stranded RNA flavivirusSome hepatocellular carcinoma (HCC) and possibly some lymphomas [24,25]. The risk of HCC is 11.5- to 17-fold in HCV-infected patients [26,27].
KSHVDouble-stranded DNA herpesvirusKaposi’s sarcoma, primary effusion lymphoma [28].
Kaposi’s sarcoma is the most frequent cause of malignancy among AIDS patients.
MCVDouble-stranded DNA polyomavirusMerkel cell carcinoma (MCC) [29]. MCC is a rare, but aggressive human skin cancer, and it typically affects the elderly, as well as immunosuppressed individuals.
EBVDouble-stranded DNA herpesvirusMost Burkitt’s lymphoma and nasopharyngeal carcinoma [30].
HBVPartially double-stranded DNA hepadnavirus with retroviral featuresChronic infection with HBV has been linked to the development of HCC for over 30 years [31,32].
HPVsA group of circular, double-stranded DNA viruses [33].High-risk human papillomaviruses (HPV) 16 and HPV 18 (some other α-HPV types are also carcinogenic) are associated with cervical cancer, penile cancers and some other anogenital and head and neck cancers [34,35].
SV40Polyomavirus of the rhesus macaque [10]SV40 sequences are detected in 60% of human mesothelioma, a rare tumor related to exposure to asbestos [36]. It is also detected in brain tumors [37,38,39], osteosarcoma [40] and non-Hodgkin lymphoma (NHL) [41,42].

2. Mechanisms of Viral Carcinogenesis

Cellular transformation is a multi-step process that results in the transformation of healthy cells into cancer cells. It requires a progression of changes at the cellular, genetic and epigenetic levels that ultimately lead to the cellular changes necessary for uncontrolled cell division and formation of a malignant mass. Hanahan and Weinberg, in their landmark review “Hallmarks of Cancer”, listed six essential alterations that must occur in a cell’s physiology to cause malignancy, including self-sufficiency in growth signals, insensitivity to growth-inhibitory (antigrowth) signals, evasion of programmed cell death (apoptosis), limitless replicative potential, sustained angiogenesis and tissue invasion and metastasis [43,44]. More recently, two emerging hallmarks have been added to the list: deregulating cellular energetics and avoiding immune destruction [44]. Cellular transformation induced by oncogenic viruses also adheres to Weinberg’s hallmarks; in particular, limitless replicative potential, evasion of apoptosis and genome instability [43]. Viruses, when functioning as carcinogenic agents, utilize a variety of carcinogenic mechanisms to transform human cells. One such mechanism is direct transformation, where the virus expresses viral oncogenes that can directly transform infected cells. Several viruses, including HPV, EBV, KSHV, SV40, HCV and HTLV, encode oncoproteins that employ several mechanisms to inactivate two of the major regulators of genome stability, cell viability and cell cycle; namely, the p53 and retinoblastoma proteins (pRB) [45].

2.1. Direct Transformation through Expression of Viral Genes

The tumor suppressor p53 is the product of the TP53 gene; it induces cell cycle arrest or apoptosis in response to cellular damage or insult, guards against genomic instability and plays a critical role in DNA repair [46]. Inactivation of p53 or depletion of its function in infected cells results in an accumulation of genomic mutations and DNA breaks, accumulating genomic instability, a loss of growth suppression and apoptosis, leading to promotion of cellular transformation [46,47]. Perhaps the best-studied example of viral inactivation of p53 is provided by the E6 protein from high-risk (HR) HPV. The HR HPV E6 protein induces ubiquitin-mediated degradation of p53, resulting in disabling of the normal cellular response to many insults, including the DNA damage response [48]. Another example of a virus with the ability to inactivate p53 is KSHV. The large multifunctional protein latency-associated nuclear antigen 1 (LANA1) expressed by KSHV interacts efficiently with p53, represses its transcriptional activity and inhibits p53-induced cell death [49]. Furthermore, the SV40 large T antigen (Tag) is well-known for its ability to bind and inactivate p53 [50]. In addition, it also plays a crucial role in cell-cycle derangement of human mesothelial cells [51], followed by transformation of the cells [52]. In the case of HCV, the NS5A viral oncoprotein is involved in apoptosis inhibition, signal transduction, transcription, transformation and the production of reactive oxygen species (ROS). In particular, NS5A has been shown to bind directly to p53 and to repress transcription of the tumor suppressor p21WAF1 in a p53-dependent manner [53,54,55].
In addition to p53, the tumor suppressor, pRB, is also a frequent target for oncogenic viruses. pRB regulates apoptosis during development, and its loss results in deregulation of growth and apoptosis [56]. In normal cells, pRB is hypophosphorylated in early G1, and becomes increasingly phosphorylated by cyclin D/CDK4/6 complexes as the cell moves towards S phase in response to a signal to divide. This results in the release of the E2F protein, which then activates the transcription of genes required for the S-phase transition [57]. E7, another HPV oncogene, mimics this process by binding to pRB and releasing the E2F protein [58], thereby driving quiescent, infected cells back into a proliferative state in order to enable viral genome replication [59]. In addition, the HR HPV E7 protein binds directly to E2F-1, leading to the activation of E2F-1-dependent transcription, and the affinity of E7 for E2F-1 appears to correlate with the oncogenic potential of the HPV [60]. Therefore, HR HPV E6 works together with E7 to induce cellular genomic instability and mitotic defects, known contributors to carcinogenesis [61]. Another example of pRb inactivation is provided by the large T antigen (Tag) expressed by SV40. One domain of Tag, the LxCxE motif, binds to pRB and inactivates the protein [62,63]. Despite the fact that adenoviruses do not cause cancer in humans, they can disrupt the pRB-E2F interactions via the activity of E1A, thus driving expression of viral transcription and inducing cell cycle progression [64].
In addition to these two major targets, pRB and p53, several other molecules serve as common targets for oncogenic viruses. Telomerase, which is usually found in embryonic cells and is absent in somatic cells, is one such target. In normal cells, the telomeric regions shorten with each round of division [65]; inappropriate expression of telomerase can lead to immortalization. HR HPV E6 can activate the expression of the catalytic subunit of telomerase, telomerase reverse transcriptase (hTERT) [66,67], in order to maintain telomeres through telomerase activation [68,69], and thereby contributing to immortalization. Another virus, EBV, encodes a principal oncoprotein, latent membrane protein 1 (LMP1), which is also able to activate hTERT [70]. HTLV-I expresses Tax protein, which is believed to be critical in the transformation of infected cells to adult T-cell leukemia (ATL). Tax is a 40-kDa trans-regulatory protein encoded by the tax gene located in ORF IV of the pX region [71]. Tax can repress the expression of hTERT by competition with c-Myc through a canonical c-Myc binding site within the hTERT promoter [72]. In addition to this function, it can also reprogram G1 to S progression through direct protein-protein binding, transcriptional induction/repression and post-translational modification [71]. Another target of oncogenic viruses is the tumor suppressor RASSF1A. The inactivation of this gene is correlated with the hypermethylation of its CpG-island promoter region, and the silencing of RASSF1A induces telomerase activity [73]. Tag of SV40 not only inactivates p53, but also inactivates the RASSF1A gene [74,75].
Several other signaling pathways involved in carcinogenesis are directly regulated by multifunctional viral oncoproteins. For example, the SV40 Tag can activate growth factor receptors, such as Met [76], Notch-1 [77] and IGF-1R [78], leading to the activation of the ERK-kinase and AP-1 pathways that promote cell division and contribute to SV40-induced carcinogenesis [79]. In addition, NS5A, the HCV protein that depletes p53, can also interact with Bax and prevent apoptosis in a p53-independent manner [80]. KSHV can express an interferon regulatory factor (IRF)-like signal-transduction protein, ORF K9, and this protein inhibits interferon-induced signaling pathways. This inhibition allows KSHV to overcome interferon-mediated antiviral activity and, thus, can contribute to host cell transformation [81]. The LMP1 oncoprotein expressed by EBV [82], as well as Tax expressed by HTLV-I [23], both target the Nuclear Factor-κB (NF-κB) pathway. The NF-κB pathway plays a critical role in regulating the immune response to infection. NF-κB is a crucial mediator of inflammation-induced tumor growth and progression, as well as an important modulator of tumor surveillance and rejection [83].

2.2. Indirect Transformational Activities

In addition to these direct mechanisms underlying virus-induced carcinogenesis, virus-induced chronic infection and inflammation can also function as indirect transforming agents [84]. In fact, Colotta et al. in a 2009 paper in Carcinogenesis refer to inflammation as the seventh hallmark of cancer [85]. Evidence of the carcinogenic potential of inflammation is provided by chronic inflammatory bowel diseases, such as chronic ulcerative colitis and Crohn’s disease, which can lead to colon cancer, even in the absence of other factors [86]. Another well-known example is the chronic infection and inflammation induced by Helicobacter pylori, which results in stomach cancer worldwide [87]. Inflammation can also be caused by infection with viruses, thereby providing another linkage between infection and carcinogenesis. For example, chronic HBV or HCV infection can lead to hepatocellular carcinoma (HCC) through a process that induces cell death, regeneration, cirrhosis and, finally, cancer [55,88,89]. In many cases, the inflammation induced by chronic infection creates a microenvironment that favors expression of viral oncogenes. For example, in HBV-induced HCC, HBV is clonally integrated into host DNA, and the integrated HBV sequences encode HBV X (HBx) and/or truncated envelope pre-S2/S proteins in a large portion of the HCC. These oncoproteins are thought to participate in directly promoting transformation of hepatocytes to HCC [55]. Another example of the way in which virus infections can indirectly contribute to carcinogenesis is provided by cutaneous papillomavirus infections, which contribute indirectly to skin carcinogenesis by blocking apoptosis in cells exposed to ultraviolet light and, thus, permitting the survival of UV-damaged cells [90].

3. DNA Damage and Viruses

p53 has been termed “the guardian of genome integrity.” Its depletion or inactivation by virus proteins, as well as the inflammation produced as a result of chronic infection, both lead to an accumulation of point mutations, genomic instability and DNA damage. DNA damage starts with a chemical modification to a base of DNA that induces a break in either one or both strands of the DNA [91]. In uninfected cells, DNA repair systems can recognize DNA damaged bases as abnormal structures and repair the damage prior to replication. The cell can call on several systems to repair DNA damage caused by both endogenous and exogenous factors, and these DNA repair pathways recognize both single-strand breaks (SSB) and double-strand breaks (DSB). SSB DNA repair mechanisms include direct repair, nucleotide excision repair, mismatch repair and base excision repair. The DSB repair mechanisms include homologous recombination and nonhomologous end joining [92]. If the DNA damage cannot be completely repaired, this damage can lead to a deregulated cell cycle, genomic instability [47] and mutations associated with the development of cancer [93,94]. However, activation of DNA damage response and cell cycle regulation by virus proteins benefit virus production by providing an S-phase-like replication environment, preventing apoptosis and promoting episome maintenance [47].

DNA Damage and Virus Infection

During viral infections, the host must maintain genome integrity through the activation of its surveillance network for detecting and repairing DNA damage. Many viruses can employ direct and/or indirect mechanisms to activate DNA damage signaling pathways [47], and this DNA damage signaling can be activated either by virus infection itself or the expression of viral proteins. One example of DNA damage signaling activated by virus infection is seen in the case of EBV infections, where infection induces the cellular DNA damage response and activates the ataxia telangiectasia-mutated (ATM) signal transduction pathway [95]. ATM is the primary signal kinase activated after sensor proteins detect DNA damage. Autophosphorylation of ATM can activate downstream substrates, such as checkpoint kinase Chk2 and the DNA damage response protein, 53BP1. Inhibition of ATM and Chk2 significantly increases the transformation efficiency of EBV-infected primary B-cells [96]. Another example is seen in the case of SV40 infection, where infection results in activation of ATM and endogenous ATM substrates [97]. SV40 viral proteins can also induce the DNA damage response, as expression of the SV40 Tag protein activates the DNA damage response via binding to the mitotic spindle checkpoint kinase, Bub 1 [98]. Activation of the DNA damage response is in the best interest of the virus, as it promotes SV40 viral DNA replication [97]. In the case of HPV, expression of viral proteins from the HPV episome leads to ATM kinase activation, which, in turn, activates Chk2. Caspase activation due to Chk2 activity is necessary for productive viral genome amplification [99,100]. In any case, if the DNA damage repair system activated by the virus is not sufficiently efficient or if additional DNA damage induced by exogenous sources accumulates in the host genome, one side effect of this increased DNA damage in infected cells is the ability of viral genomes to become integrated into the host genome. In the case of HPV and HBV, viral genome integration is a major trigger point for the development of cancer. The role and importance of viral genome integration in cancer development will be discussed below.

4. Factors that Cause DNA Damage and that Contribute to Carcinogenesis

Factors that induce DNA damage in infected cells can be divided into endogenous, i.e., virus mediated, and exogenous factors.

4.1. Inflammation and DNA Damage Induced by Virus Itself

Inflammation is a primary immune response to infection by pathogens [101]. That process involves activation and directed migration of leukocytes from the venous system to the sites of infection; tissue mast cells also play a significant role [86]. A family of chemokines attract leukocytes, whose persistence at an inflammatory site is important in the development of chronic disease [86].
Inflammation is referred to as a cancer “promoter”, because it induces cell proliferation, recruits inflammatory cells, increases cellular levels of ROS, thereby leading to oxidative DNA damage, and reduces DNA repair [86]. The deregulation of cell death and/or repair programs results in DNA replication and proliferation in chronically-inflamed tissue [86]. Inflammation also causes resistance to apoptosis, secretion of pro-angiogenic and immunosuppressive factors, invasion and metastasis [102]. All of these processes contribute to carcinogenesis. For example, the most common type of liver cancer results from chronic liver inflammation induced by either HBV or HCV [103,104]. The enhanced DNA replication and DNA damage created during chronic inflammation increase the number of free DNA ends in host genomic DNA and promote HBV integration [105]. In case of HCV, its core protein interacts with the signal transducer and activator of transcription 3 (STAT3) protein [106], a transcription factor involved in mediating cytokine signaling [107]. This interaction results in enhanced proliferation and upregulation of Bcl-XL and cyclin-D [106]. In this way, chronic liver inflammation induced by HCV may alter the local cytokine profile and the balance between apoptosis and proliferation.

4.2. Inflammation and DNA Damage induced by Co-Infections

Co-infections with certain sexually transmitted diseases (STD) cause cervical inflammation and increase the risk of cervical cancer in HPV-infected women [101,108,109]. Furthermore, high levels of inflammatory mediators, such as cyclooxygenase (COX)-2, an enzyme responsible for prostaglandin formation, are observed in cervical cancer [110,111]. However, this inflammation is not thought to be primarily due to the HPV infection itself, but rather due to other factors, such as co-infections. One reason for the lack of HPV-induced inflammation is that HPV infects basal keratinocytes that are distant from immune centers and have short lifespans. In addition, the virus does not destroy the cells it infects, thereby avoiding the triggering of inflammation [112]. Co-infections that do trigger inflammation can be of either viral or bacterial origin. For example, studies have determined that co-infection with either Chlamydia Trachomatis or HSV increase the risk of developing cervical cancer [113], as does infection with other STDs, such as Neisseria gonorrhoeae [114]. The inflammation induced by these co-infections can induce the generation of ROS, which can contribute to carcinogenesis by damaging DNA, as described below.

4.3. Oxidative Stressors, Viruses and Cancer

ROS and reactive nitrogen species (RNS) are charged free radicals that are primarily generated in mitochondria as by-products of aerobic respiration, cytochrome P450 activity and peroxisome function [115]. In normal conditions, the pro- and anti-oxidant systems maintain ROS homeostasis. A lack of proper balance between these two sets of systems results in changes to cellular levels of ROS and can lead to oxidative stress (OS). In general, the sources of OS can be divided into two broad categories: exogenous and endogenous. Endogenous OS, as discussed above, is primarily derived from natural processes, such as cellular signaling, metabolic processes and inflammation [116,117,118,119]. Exogenous and environmental sources include ionizing radiation, such as X-, γ- and cosmic rays, α-particles from radon decay, oxidizing chemicals and UVA solar radiation. For example, ionizing radiation generates radicals, including superoxide, hydrogen peroxide and hydroxyl radicals [120], most of which are generated during the radiolysis of water. Of these, the hydroxyl radical is the most damaging species and produces mostly single-strand breaks [121]. Overall, radiation induces genetic instability and chromosomal rearrangements, and many of these rearrangements are similar to those found in human cancers [122].
Chronic exposure to viral infections also induce the constant generation of free radicals [123], which can damage cellular biomolecules, including DNA. DNA damage produced by OS results in apurinic/apyrimidinic (abasic) DNA sites, oxidized purines and pyrimidines, and single- and double-stranded DNA breaks [124]. Therefore, the ROS- and RNS-induced oxidative and nitrative DNA damage that frequently occurs during inflammation can contribute to carcinogenesis [125,126].
A connection between OS, DNA damage and the incidence of hepatocellular carcinoma has been demonstrated by Hagen and colleagues [127]. The elevated level of ROS observed in chronic HBV-infected livers causes liver cell injury [127,128] and may lead to an accumulation of repeated genetic damage and an increased risk for genomic alterations in infected hepatocytes. Moreover, cell necrosis and proliferation in chronic HBV infection in response to cell injury could allow for greater exposure of DNA to ROS and incomplete repair of DNA damage. All of these factors are predicted to increase the probability of the fixation of genetic and chromosomal abnormalities, thus causing mutations and enhancing the development of HCC [127]. Among the proteins encoded by HBV, the X gene product (HBx) is a protein that increases virus gene expression and replication to maintain viral infection by transactivating cellular promoters and enhancers [129,130]. HBx induces OS and contributes to liver disease pathogenesis associated with HBV infection [131].
High levels of ROS can directly regulate NF-κB activation. This was demonstrated in EBV-positive Burkitt’s lymphoma cells, which have elevated ROS levels and altered NF-κB activation as opposed to EBV-negative Burkitt’s lymphoma cells [132]. EBV nuclear antigen-1 is the only viral protein expressed in all EBV-carrying malignancies, and it induces chromosomal aberrations and DNA DSBs through increasing ROS production. This effect can be reversed by antioxidants [133,134].
As opposed to several of these other viruses, infection with HPV does not in itself cause significant inflammation likely to lead to OS. However, expression of the smaller splice variant of E6, E6*, is able to increase ROS levels in oral keratinocytes. This ROS level increase is likely connected to the E6*-mediated decrease in the expression of the antioxidant enzymes, superoxide dismutase (SOD) 2 and glutathione peroxidase (GPx) 1/2 [135]. The increase in the level of oxidative stress due to E6* expression also resulted in an increase in DNA damage [135].
Other environmental and lifestyle-related factors also contribute to the induction of oxidative stress in infected cells. Among the factors connected to lifestyles that can promote virus-induced tumorigenesis are alcohol consumption and tobacco smoking. For example, heavy alcohol intake, defined as ingestion of more than 50–70 g/day for prolonged periods, is a well-established HCC risk factor [136]. Alcohol-induced oxidative stress plays an important role in the development of alcohol liver disease, and alcohol metabolism via the enzyme, alcohol dehydrogenase, results in increased ROS production, hepatocyte injury and apoptosis. Interestingly, all of these reactions could be blocked by the administration of antioxidants [137,138].
Smoking is strongly associated with an increased risk of developing cervical cancer in HPV-positive women [139,140]. Smoking is known to induce inflammation and oxidative stress [141], which may lead to DNA damage, integration and carcinogenesis, as discussed later.
Many studies have demonstrated that long-term use of oral contraceptives increases the risk of cervical cancer [142,143,144]. The strongest evidence for this connection comes from the large pooled analysis of the International Agency for Research on Cancer (IARC) studies for the role of oral contraceptive use in HPV-induced carcinogenesis. Although “ever use” of oral contraceptives was only moderately associated with cancer risk, a strong dose-response relationship with increasing years of use was observed [142].
High parity may also increase the risk of cervical cancer. One possible mechanism is that pregnancy maintains the transformation zone on the exocervix for many years [145], thereby facilitating direct exposure to HPV and other cofactors. Hormonal changes induced by pregnancy, such as increased levels of estrogen and progesterone, may also modulate the immune response to HPV along with viral persistence and/or progression [146,147]. High parity may also cause cervical trauma and cellular oxidative stress, thus leading to DNA damage and carcinogenesis [148]. In the IARC-pooled analysis, the odds ratio for cervical cancer in women with seven or more full-term pregnancies was four-fold higher than that in nulliparous women, and the risk increased linearly with an increasing number of full-term pregnancies [146].

5. Integration of Viral DNA into the Human Genome

Cancer incidence is associated with integration of the viral genome for several oncogenic viruses, including HBV, HPV and MCV. It is not yet clear that integration of EBV is a mechanism for carcinogenesis. The risk of virus integration depends on the level of DNA damage, because integration requires DSBs in both the host and virus DNA [149] (see Figure 1). Furthermore, a major cause of DNA damage is OS, which can be triggered by viruses and enhanced by exogenous factors.
Figure 1. Integration of HBV, HPV and MCV viral DNAs into the human genome induces cellular and viral responses and further contributes to carcinogenesis (HBV, hepatitis B virus; HPV, high-risk human papillomaviruses; MCV, Merkel cell polyomavirus; MCC, Merkel cell carcinoma; TERT, telomerase reverse transcriptase; MLL4, Mixed-lineage leukemia 4).
Figure 1. Integration of HBV, HPV and MCV viral DNAs into the human genome induces cellular and viral responses and further contributes to carcinogenesis (HBV, hepatitis B virus; HPV, high-risk human papillomaviruses; MCV, Merkel cell polyomavirus; MCC, Merkel cell carcinoma; TERT, telomerase reverse transcriptase; MLL4, Mixed-lineage leukemia 4).
Cancers 06 02155 g001

5.1. Hepatitis B Virus (HBV)

HCC is the fifth most common cancer and the third leading cause of cancer death worldwide [136], and up to 80% of HCC is attributable to either HBV or HCV infection [150]. HBV is a partially double-stranded DNA hepadnavirus with retroviral features. The risk of HCC is increased five- to 15-fold in chronic HBV carriers compared with the general population [136], and studies have shown that integration of the HBV genome into the cellular genome is present in over 85%–90% of HBV-related HCCs. However, the integrated form of HBV is also present in non-tumor tissue of patients with chronic HBV infections. Integration of the HBV genome into hepatocytes occurs during persistent HBV infection and precedes development of HCC [151,152]. HBV integration leads to the elevated expression of several cellular cancer-related genes, such as TERT, mixed-lineage leukemia 4 (MLL4) (MLL4 is a part of the ASC-2 complex implicated in the p53 tumor suppressor pathway [153]) and CCNE1 (encoding cyclin E1) [154]. HBV integration is also associated with early onset of HCC and poor outcomes [154], and integrated HBV sequences encoding the HBx and/or truncated envelope pre-S2/S proteins are found in a large percentage of HCC [55]. Integration of the HBx sequence into host DNA in HCC promotes genetic instability through mechanisms that include the inactivation of the UV-damage DNA binding protein, so as to interfere with nucleotide excision repair, repression of p53-mediated gene transcription [155] and inactivation of p53-dependent apoptosis, cell cycle regulation, DNA repair and tumor suppression [156]. In addition, HBx transactivates several signaling pathways connected to carcinogenesis, including those mediated by protein kinase C, JAK/STAT and PI3K. [130,157]. HBx also upregulates TGF-β expression in HCC tissue and is thought to contribute to carcinogenesis through this mechanism, as well [158]. TGF-β is a cytokine that inhibits hepatocyte proliferation during liver regeneration [159,160,161] and stimulates extracellular matrix protein production by hepatocytes during liver cirrhosis [162,163]. Deletion of the preS2 region of the S2/S protein during HBV integration leads to a truncated envelope pre-S2/S protein that is frequently found in HCC samples [164] (see Figure 1). This truncated pre-S2/S product increases malignant transformation by transactivating several cellular genes, including c-myc, c-fos and c-Ha-ras [165,166]. Interestingly, the pre-S mutant large surface antigens can be retained in the endoplasmic reticulum and, thus, escape detection by the immune system. Moreover, this protein can initiate ER stress and thereby induce oxidative DNA damage and genomic instability [167]. The pre-S mutant also can upregulate COX-2 and cyclin A to induce cell-cycle progression and proliferation of hepatocytes [167]. Overall, the process of HBV integration induces additional genetic alterations, including chromosomal deletions, translocations, fusion of transcripts, amplification of DNA and generalized genomic instability [105,157], leading to overexpression of oncogenes, depletion of tumor suppressor genes and an altered microRNA profile [157].
The integration rate of HBV DNA into the host genome is significantly increased in the presence of DSBs [168]. If DNA damage is induced by the addition of H2O2, which increases cellular levels of ROS, the frequency of HBV integration also increases significantly [169]. Moreover, inhibiting poly(ADP-ribosyl)ation by adding the ADP-ribosylation inhibitor, 3-aminobenzamide, also leads to a significant increase in integration [169]. Poly(ADP-ribose) polymerase-1 (PARP-1) is an enzyme that is stimulated by DNA strand breaks caused by oxidative stress and other types of DNA damage and responds by carrying out poly(ADP-ribosyl)ation. Poly(ADP-ribosyl)ation is involved in DNA damage repair processes, such as base excision repair and suppression of genetic recombination [170,171,172,173].
Integration of HBV into the human genome can occur near or within fragile sites in genes that regulate cell signaling, proliferation and viability [174]. Common targets of integration include genes for human cyclin A2 [175], the PDGF receptor, calcium signaling-related genes, mixed lineage leukemia encoding genes, 60S ribosomal protein genes [174], human telomerase reverse transcriptase (hTERT) [176] and the retinoic acid receptor β [177]. The data produced by next-generation sequencing (NGS) now allows researchers to not only determine the sites of integration, but also to identify the mutations and to specify DNA damage types that contribute to carcinogenesis [178,179]. NGS studies have shown that HBV integrations into TERT, MLL4, CCNE1 and ANGPT1 (encoding angiopoietin 1) [154,180,181] can and do occur. NGS has also identified frequent mutations in CTNNB1 (encoding β-catenin), IRF2 (encoding interferon regulatory factor 2), TP53, ARID2 (subunit of the polybromo- and BRG1-associated factor chromatin remodeling complex [182]; functioning as a tumor suppressor gene [183]) and ARID1A (encoding a component of the SWI/SNF chromatin remodeling complex) [183,184,185]. These mutations can be associated with disease etiology. For example, ARID2 mutations are significantly enriched in HCV-associated HCC [183], and ARID1A may be crucial in HCC invasion and metastasis [185]. IRF2 mutations have been associated with hyperploidy and high chromosomal instability [184], while TP53 inactivating mutations result in an altered p53 pathway [184]. Interestingly, CTNNB1 mutations may define a homogenous subtype of HCC not related to HBV infection [184].

5.2. Human Papillomaviruses (HPVs)

HPVs are a group of circular, double-stranded DNA viruses that infect epithelial cells. More than 100 different genotypes of HPV [186] have been described; of these, a subset infects the anogenital area, and within this subset, the individual types are classified as either high risk or low risk. High-risk HPVs can cause cancerous lesions, while low-risk HPVs do not [187]. An important difference between high- and low-risk HPVs is that high-risk HPVs show a greater tendency to integrate into the host genome, thereby causing high-grade lesions and cancer, while low-risk types are preferentially maintained as extrachromosomal circular episomes [188,189].
HPV is known to cause virtually all cases of cervical cancer, which is the second most common cancer in women worldwide and the fourth most common cause of cancer death in women worldwide [190,191]. More than 270,000 women die from cervical cancer each year, and according to 2013 data from the WHO, the developing world accounts for more than 85% of these cases. More than 90% of premalignant and malignant squamous lesions in the uterine cervix are HPV DNA positive [192,193]. Moreover, HPV16, HPV18, HPV31 and HPV33 account for 90% of all cases of cervical cancer. Among these high-risk HPVs, HPV type 16 is the most prevalent type and by itself accounts for more than 50% of all cases of cervical cancer [194,195]. High-risk HPV infection is also associated with several other anogenital and oropharyngeal cancers. For example, it is thought to be responsible for more than 90% of anal cancers, 70% of vaginal and vulvar cancers, 60% of penile cancers and 63% of oropharyngeal cancers [196].
The majority of HPV infections are spontaneously cleared by human cells [197,198]. Cervical intraepithelial neoplasia (CIN) 1 lesions develop occasionally, but most of these lesions will be spontaneously cleared and fail to progress [199]. However, some HPV-infected women can be co-infected by other viruses or bacteria, then develop cervical inflammation, as noted above. The cellular proliferative and anti-apoptotic effects of inflammation, combined with low-level expression of the E6 and E7 oncogenes encoded by the episomal HPV, contribute to CIN1 to CIN2 progression [101]. At the CIN2 stage, the generation of ROS and RNS may create DSBs in both the viral and host DNA. This can then allow HPV to integrate into the human genome, thereby enabling overexpression of the E6 and E7 oncogenes, which then facilitate the transition to CIN3 and, sometimes, invasive carcinoma.
The oncogenes encoded by HPV play crucial roles in carcinogenesis. Typically, the levels of E6 and E7 oncogene expression from episomal HPV16 are low. In vitro, high-level expression of HPV oncogenes from integrated HPV forms is generally preceded by the loss of episomal HPV [200,201]. HPV-16 integration often disrupts the E2 gene (the E2 ORF has been identified as a preferential site of integration), resulting in significantly increased expression of the E6 and E7 oncoproteins due to loss of negative feedback control by the viral regulatory E2 protein [149,202] (see Figure 1). Consistent with this model, transcriptional activity of integrated HPV-16 DNA can be suppressed by E2 proteins from the episomal form [149,200,203,204,205].
Elevated expression of HPV oncogenes can also be achieved by the presence of an increased number of episomal HPV-16 copies. A high episomal HPV-16 load, combined with an absence of HPV-16 integration, is frequently observed in CIN, while a high proportion of invasive cancers contain integrated HPV-16 forms, suggesting that integration is an important factor in progression from high-grade lesions to invasive cancer and may be a potential marker for CIN progression risk assessment [206,207].
Accumulating evidence suggests that among those factors that can increase the risk of HPV integration are those factors that generate excessive amounts of ROS and RNS. Nitrative and oxidative DNA damage were found in high-risk HPV infections, especially nitric oxide synthase-dependent DNA damage, which is believed to play a critical role in inflammation-mediated cervical carcinogenesis [208]. Exposure to high and sustained levels of nitric oxide (NO) will increase DNA damage and mutation frequencies, supporting the idea that NO is a cofactor with infection in cervical carcinogenesis [209]. Furthermore, ROS and RNS have the potential to create DNA strand breaks [209,210], thereby increasing HPV-DNA integration into cellular chromatin. As discussed above, these factors may be virus mediated (i.e., E6* expression) [135], mediated through infection or co-infection (inflammation) or exogenously derived [101]. Our laboratory has demonstrated that chronic oxidative stress induced by E6* expression is able to increase DNA damage [135], and it has also been shown by other laboratories that the frequency with which foreign DNA constructs containing antibiotic resistance markers integrate into genomic DNA is increased by the induction of DSBs, either by HPV16 E6 and E7 expression or by spontaneous breakage at fragile sites in specific cancer cell lines [211,212,213]. Therefore, the rate of HPV integration seems to correlate with the level of DNA damage.
A number of pathways are affected by E6 and E7 expression [214]. As mentioned earlier, E6 mediates the rapid degradation of p53 and activates hTERT, while E7 inactivates pRB. In addition, E6 binds to IRF-3 and inhibits its transcriptional activity, which may provide HPV with the ability to circumvent the normal antiviral response [215]. E7 sequences mediate the activation of cyclin E, followed by the activation of cyclin A, which is required for transformation [216]. In addition, the cooperative interactions between the E6 and E7 proteins leads to cellular immortalization [217,218], likely through a combination of specific mechanisms, such as those noted above. In summary, the elevated expression of oncoproteins from integrated forms of HPV deregulates cellular proliferation, blocks apoptosis and increases genomic instability, all of which contribute to cellular transformation (see Figure 1).
HPV integration sites are distributed randomly throughout the host genome, without a single region predominating [219]. However, 38% of 192 integrants were found in known common fragile sites (CFSs), and there was no evidence to suggest targeted disruption or functional alteration of critical cellular genes by the integrated viral sequences [219]. However, some studies have demonstrated that high-risk HPV integration has occurred within or adjacent to known oncogenes, most commonly within intronic sequences [219,220,221,222]. For example, the region of the MYC gene at chromosomal band 8q24 is a frequently observed integration site in HPV18-positive cervical cancer [221,222,223]. Recently developed NGS-based methods now provide a very efficient method to map HPV integration sites. One NGS study, for example, showed that the 3'-breakpoints of integrated HPV16 DNA distribute preferentially within the early region E1-PAE segment in HPV 16. This indicates the importance of deregulated viral oncogene expression for carcinogenesis [224]. Interestingly, about half of the mapped HPV16 integration sites directly target human cellular genes [224]. These studies suggest that the insertional mutagenesis of the host genome may play a role in at least some cervical cancers [149]. However, many publications using NGS focus on identifying genotype and determining HPV load, rather than on identifying sites of integration [225,226]. Further studies are needed to identify integration sites, genomic mutation and DNA damage.

5.3. Merkel Cell Polyomavirus (MCV)

MCV is a double-stranded DNA polyomavirus, shown to be associated with Merkel cell carcinoma (MCC) through the use of NGS in 2008 [29]. These tumors display MCV DNA in an integrated form within the tumor cell genomes in a clonal pattern, suggesting that MCV infection and integration contribute to clonal expansion of the tumor cells [29,227] (see Figure 1). The MCV genome encodes multiple splice variants of a tumor (T) antigen protein complex that targets several tumor suppressor proteins, such as pRB and p53 [228]. One of these splice variants, the large tumor antigen, is mutated in MCV-positive MCC tumors cells, and this selective mutation affects the cellular DNA damage response to prevent auto-activation of integrated virus replication, disrupt host genomic integrity and inhibit cellular proliferation [228,229,230]. Several features of this virus, such as the frequent and selective association of MCC with MCV, integration of MCV into the host genome, the recurrent pattern of conserved viral DNA sequences and the constant expression of viral oncoproteins are very similar to those seen in high-risk HPV-induced cervical cancer [55]. However, the exact role of integration in MCC carcinogenesis requires further study.

5.4. Epstein-Barr Virus (EBV)

EBV is a double-stranded DNA herpesvirus that is primarily associated with Burkitt’s lymphoma, nasopharyngeal carcinoma and several lymphoproliferative disorders [30]. Burkitt’s lymphoma appears in three main clinical variants—the endemic, sporadic and immunodeficiency-associated variants. EBV is detected in 96% of cases of endemic variant Burkitt’s lymphoma involving the jaw, which is the most common malignancy of children in certain areas of central Africa [231]. In contrast, EBV is rarely associated with the sporadic variant of Burkitt’s lymphoma, and the jaw is less commonly involved [232]. EBV-associated Burkitt’s lymphoma is common in individuals lacking efficient T-cell function, such as AIDS patients or transplant recipients [233].
All Burkitt’s lymphomas have chromosomal translocations that place the MYC oncogene under the control of the Ig heavy chain or one of the light-chain loci, which induce MYC deregulation and contribute to the pathogenesis associated with Burkitt’s lymphoma [234,235,236]. EBV also displays transformative abilities, as EBV-encoded latent genes induce B-cell transformation into permanent, latently-infected lymphoblastoid cell lines (LCLs) in vitro by altering cellular gene transcription and activating important cell-signaling pathways [237].
EBV usually persists in an episomal state with multiple copies of circular DNA. Integration of the EBV genome into that of the host is rare and is unlikely to contribute to most cases of Burkitt’s lymphoma. Such integration has been only shown in several Burkitt’s lymphoma cell lines, such as IB4, BL-36, BL-60 and BL-137 [238,239,240,241,242]. Integrated, episomal and linear copies of EBV DNA can coexist in Burkitt’s lymphoma cells [241]. EBV integration into fragile sites of the host chromosome is associated with partial deletion of the viral genome and generates a region of enhanced chromatin instability in the host cell genome [238]. This genome instability can induce the loss of host genes, such as BACH2, which is a putative tumor suppressor gene, and this may contribute to lymphomagenesis [243].
In addition to Burkitt’s lymphoma, EBV is also associated with nasopharyngeal carcinoma (NPC). The undifferentiated form of NPC, classified by WHO as type III, shows the most consistent association with EBV worldwide, especially in particular areas of China and south-east Asia [244]. An association between EBV and WHO types I and II NPC has also been demonstrated [245]. Evidence of the importance of EBV integration for the development NPC is not conclusive, and integration of EBV was demonstrated in only a subset of studies. For example, integrated EBV was found in four of 17 pathologically-diagnosed, EBV-positive NPC biopsy samples [246]. Another study showed that EBV DNA was integrated into chromosomes in the EBV-positive NPC cell lines, HSB4 and H2B17-7. However, the exact role of integration of EBV in NPC is not clear.

6. Prevention and Reduction of Risk Factors for Cancers

Knowledge of the etiology of virus-mediated carcinogenesis, the networking of pathways involved in the transition from infection to cancer and the risk factors associated with each type of cancer, all suggest prophylactic and therapeutic strategies that may reduce the risk of virus-mediated cancer. Prophylactic vaccination provides one such strategy; HPV vaccines can effectively reduce the incidence of HPV infections and are anticipated to correspondingly reduce the incidence of HPV-associated cancers and pre-cancerous lesions over the next several decades [247]. As discussed in this current review, lifestyle choices also matter, and education on the risk factors associated with these lifestyle choices can be considered a prophylactic measure targeting cancer prevention. We also discussed factors that contribute to virus-mediated carcinogenesis, including integration of the viral genomes of HBV, HPV and MCV. Typically, cancer develops years to decades following the initial infection with HBV and HPV. This delay provides us with a unique opportunity for cancer interception, especially in cases where the tumorigenesis process requires virus integration. In these situations, approaches that would decrease the probability of integration are anticipated to reduce the number of cancer cases. Theoretically, therapeutic or dietary measures directed against oxidative stress could reduce oxidative stress in already infected cells [248] and thereby diminish the risk of viral integration. Dietary antioxidants can, in some cases, supplement the activity of endogenous antioxidants found in normal cells and fortify them against challenges posed by increased levels of ROS. In the case of cervical cancer, studies have shown that several antioxidants, such as reduced glutathione, GPx, glutathione-S-transferase, SOD and antioxidant vitamins (vitamin E, vitamin C, lutein, beta-carotene, lycopene and zeaxanthin), are reduced in the circulation of cervical cancer patients [249,250]. In contrast, the concentration of the lipid peroxidation parameter, malondialdehyde, was significantly higher in women with CIN or invasive cervical cancer [250]. That may be due to the increased utilization of the antioxidants to scavenge oxidative stress-induced lipid peroxidation and their sequestration by tumor cells. A deficiency in antioxidant vitamins may, therefore, contribute to the increased prevalence of cervical cancer observed in women with a low socioeconomic status [249]. Possible benefits from dietary antioxidant consumption, a highly researched topic for the last two decades, still remain debatable [251,252]. In addition, clinical trials addressing the effect of antioxidant therapies on cancer have not yet provided a clear answer [253]. Future development of therapies for the reduction of cellular oxidative stress must take into account the complexity of problems inherent in the regulation of redox balance.

7. Conclusions

Viruses have been demonstrated to be the causative agents of approximately 10%–15% of all cancers worldwide. Viruses, when acting as carcinogenic agents, utilize a variety of mechanisms to transform human cells. Their genome encodes proteins that reprogram the normal functioning of cells so as to favor virus production, and the most common outcome for this virus-induced reprogramming is the induction of genomic instability, as seen by the accumulation of point mutations, aberrations and DNA damage. However, the mechanisms causing this genomic instability differ between viruses. Genomic instability caused by viruses serves as a major step on the pathway leading to carcinogenesis and can be induced by viral infection and inflammation, by viral gene expression and by exogenously derived factors. All of these mechanisms can lead to increased oxidative stress, which is then able to damage host DNA. For some viruses, such as HBV, HPV and MCV, the pivotal step of the onset of tumorigenesis appears to be integration of the viral genome, which elevates viral oncogene expression and promotes cancer progression. Typically, cancer develops years to decades following the initial infection. This provides us with a unique opportunity for cancer interception, especially in cases where the tumorigenesis process requires virus integration. Our increasing understanding of the etiology of virus-mediated carcinogenesis, the networking of pathways involved in the transition from infection to cancer and information on risk factors specific for each type of cancer will continue to suggest prophylactic and therapeutic strategies to reduce the risk of virus-mediated cancer. However, the detailed mechanisms of how integration causes and accelerates carcinogenesis are still not fully understood and require further study.

Acknowledgments

We thank Whitney Evans for helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fairley, C.K.; Read, T.R. Vaccination against sexually transmitted infections. Curr. Opin. Infect. Dis. 2012, 25, 66–72. [Google Scholar] [CrossRef] [PubMed]
  2. Tohme, R.A.; Holmberg, S.D. Is sexual contact a major mode of hepatitis C virus transmission? Hepatology 2010, 52, 1497–1505. [Google Scholar] [CrossRef]
  3. Burchell, A.N.; Winer, R.L.; de Sanjose, S.; Franco, E.L. Chapter 6: Epidemiology and transmission dynamics of genital HPV infection. Vaccine 2006, 24, S52–S61. [Google Scholar] [CrossRef]
  4. Buddeberg, F.; Schimmer, B.B.; Spahn, D.R. Transfusion-transmissible infections and transfusion-related immunomodulation. Best Pract. Res. Clin. Anaesthesiol. 2008, 22, 503–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Wilkins, T.; Malcolm, J.K.; Raina, D.; Schade, R.R. Hepatitis C: Diagnosis and treatment. Am. Fam. Physician 2010, 81, 1351–1357. [Google Scholar] [PubMed]
  6. Hughes, R.A. Drug injectors and the cleaning of needles and syringes. Eur. Addict. Res. 2000, 6, 20–30. [Google Scholar] [CrossRef] [PubMed]
  7. Kazim, S.N.; Wakil, S.M.; Khan, L.A.; Hasnain, S.E.; Sarin, S.K. Vertical transmission of hepatitis B virus despite maternal lamivudine therapy. Lancet 2002, 359, 1488–1489. [Google Scholar] [CrossRef] [PubMed]
  8. Lam, N.C.; Gotsch, P.B.; Langan, R.C. Caring for pregnant women and newborns with hepatitis B or C. Am. Fam. Physician 2010, 82, 1225–1229. [Google Scholar] [PubMed]
  9. Cozad, J. Infectious mononucleosis. Nurse Pract. 1996, 21, 14–16, 23, 27–28. [Google Scholar] [CrossRef] [PubMed]
  10. Shah, K.V. Sv40 and human cancer: A review of recent data. Int. J. Cancer 2007, 120, 215–223. [Google Scholar] [CrossRef] [PubMed]
  11. Datta, S.D.; Koutsky, L.A.; Ratelle, S.; Unger, E.R.; Shlay, J.; McClain, T.; Weaver, B.; Kerndt, P.; Zenilman, J.; Hagensee, M.; et al. Human papillomavirus infection and cervical cytology in women screened for cervical cancer in the United States, 2003–2005. Ann. Intern. Med. 2008, 148, 493–500. [Google Scholar] [CrossRef] [PubMed]
  12. Shulzhenko, N.; Lyng, H.; Sanson, G.F.; Morgun, A. Menage a trois: An evolutionary interplay between human papillomavirus, a tumor, and a woman. Trends Microbiol. 2014, 22, 345–353. [Google Scholar] [CrossRef] [PubMed]
  13. WHO 2014 data: Hepatitis B. Available online: http://www.who.int/mediacentre/factsheets/fs204/en/ (accessed on 14 July 2014).
  14. Negro, F.; Alberti, A. The global health burden of hepatitis C virus infection. Liver Int. 2011, 31, S1–S3. [Google Scholar] [CrossRef]
  15. WHO 2014 data: Hepatitis C. Available online: http://www.who.int/mediacentre/factsheets/fs164/en/ (accessed on 14 July 2014).
  16. Rosen, H.R. Clinical practice. Chronic hepatitis C infection. N. Engl. J. Med. 2011, 364, 2429–2438. [Google Scholar] [CrossRef] [PubMed]
  17. Centers for Disease Control and Prevention. Recommendations for counseling persons infected with human T-lymphotrophic virus, types I and II. MMWR Recomm. Rep. 1993, 42, 1–13. [Google Scholar]
  18. Kondo, T.; Nonaka, H.; Miyamoto, N.; Yoshida, R.; Matsue, Y.; Ohguchi, Y.; Inouye, H.; Komoda, H.; Hinuma, Y.; Hanaoka, M. Incidence of adult T-cell leukemia-lymphoma and its familial clustering. Int. J. Cancer 1985, 35, 749–751. [Google Scholar] [CrossRef] [PubMed]
  19. Murphy, E.L.; Hanchard, B.; Figueroa, J.P.; Gibbs, W.N.; Lofters, W.S.; Campbell, M.; Goedert, J.J.; Blattner, W.A. Modelling the risk of adult T-cell leukemia/lymphoma in persons infected with human T-lymphotropic virus type I. Int. J. Cancer 1989, 43, 250–253. [Google Scholar] [CrossRef] [PubMed]
  20. Dalstein, V.; Riethmuller, D.; Pretet, J.L.; le Bail Carval, K.; Sautiere, J.L.; Carbillet, J.P.; Kantelip, B.; Schaal, J.P.; Mougin, C. Persistence and load of high-risk hpv are predictors for development of high-grade cervical lesions: A longitudinal french cohort study. Int. J. Cancer 2003, 106, 396–403. [Google Scholar] [CrossRef] [PubMed]
  21. Chen, C.J.; Yang, H.I.; Su, J.; Jen, C.L.; You, S.L.; Lu, S.N.; Huang, G.T.; Iloeje, U.H. Risk of hepatocellular carcinoma across a biological gradient of serum hepatitis B virus DNA level. JAMA 2006, 295, 65–73. [Google Scholar] [CrossRef] [PubMed]
  22. Poiesz, B.J.; Ruscetti, F.W.; Gazdar, A.F.; Bunn, P.A.; Minna, J.D.; Gallo, R.C. Detection and isolation of type c retrovirus particles from fresh and cultured lymphocytes of a patient with cutaneous t-cell lymphoma. Proc. Natl. Acad. Sci. USA 1980, 77, 7415–7419. [Google Scholar] [CrossRef] [PubMed]
  23. Yasunaga, J.; Matsuoka, M. Leukaemogenic mechanism of human t-cell leukaemia virus type I. Rev. Med. Virol. 2007, 17, 301–311. [Google Scholar] [CrossRef] [PubMed]
  24. Choo, Q.L.; Kuo, G.; Weiner, A.J.; Overby, L.R.; Bradley, D.W.; Houghton, M. Isolation of a cDNA clone derived from a blood-borne non-a, non-b viral hepatitis genome. Science 1989, 244, 359–362. [Google Scholar] [CrossRef] [PubMed]
  25. Marcucci, F.; Mele, A. Hepatitis viruses and non-hodgkin lymphoma: Epidemiology, mechanisms of tumorigenesis, and therapeutic opportunities. Blood 2011, 117, 1792–1798. [Google Scholar] [CrossRef] [PubMed]
  26. Donato, F.; Boffetta, P.; Puoti, M. A meta-analysis of epidemiological studies on the combined effect of hepatitis B and C virus infections in causing hepatocellular carcinoma. Int. J. Cancer 1998, 75, 347–354. [Google Scholar] [CrossRef] [PubMed]
  27. Donato, F.; Tagger, A.; Gelatti, U.; Parrinello, G.; Boffetta, P.; Albertini, A.; Decarli, A.; Trevisi, P.; Ribero, M.L.; Martelli, C.; et al. Alcohol and hepatocellular carcinoma: The effect of lifetime intake and hepatitis virus infections in men and women. Am. J. Epidemiol. 2002, 155, 323–331. [Google Scholar] [CrossRef] [PubMed]
  28. Chang, Y.; Cesarman, E.; Pessin, M.S.; Lee, F.; Culpepper, J.; Knowles, D.M.; Moore, P.S. Identification of herpesvirus-like DNA sequences in aids-associated kaposi’s sarcoma. Science 1994, 266, 1865–1869. [Google Scholar] [CrossRef] [PubMed]
  29. Feng, H.; Shuda, M.; Chang, Y.; Moore, P.S. Clonal integration of a polyomavirus in human merkel cell carcinoma. Science 2008, 319, 1096–1100. [Google Scholar] [CrossRef] [PubMed]
  30. Epstein, M.A.; Achong, B.G.; Barr, Y.M. Virus particles in cultured lymphoblasts from burkitt’s lymphoma. Lancet 1964, 1, 702–703. [Google Scholar] [CrossRef] [PubMed]
  31. Blumberg, B.S.; Alter, H.J.; Visnich, S. A “new” antigen in leukemia sera. JAMA 1965, 191, 541–546. [Google Scholar] [CrossRef] [PubMed]
  32. Di Bisceglie, A.M. Hepatitis B and hepatocellular carcinoma. Hepatology 2009, 49, S56–S60. [Google Scholar]
  33. Flores, E.R.; Lambert, P.F. Evidence for a switch in the mode of human papillomavirus type 16 DNA replication during the viral life cycle. J. Virol. 1997, 71, 7167–7179. [Google Scholar] [PubMed]
  34. Durst, M.; Gissmann, L.; Ikenberg, H.; zur Hausen, H. A papillomavirus DNA from a cervical carcinoma and its prevalence in cancer biopsy samples from different geographic regions. Proc. Natl. Acad. Sci. USA 1983, 80, 3812–3815. [Google Scholar] [CrossRef] [PubMed]
  35. Boshart, M.; Gissmann, L.; Ikenberg, H.; Kleinheinz, A.; Scheurlen, W.; zur Hausen, H. A new type of papillomavirus DNA, its presence in genital cancer biopsies and in cell lines derived from cervical cancer. EMBO J. 1984, 3, 1151–1157. [Google Scholar] [PubMed]
  36. Carbone, M.; Pass, H.I.; Rizzo, P.; Marinetti, M.; di Muzio, M.; Mew, D.J.; Levine, A.S.; Procopio, A. Simian virus 40-like DNA sequences in human pleural mesothelioma. Oncogene 1994, 9, 1781–1790. [Google Scholar] [PubMed]
  37. Bergsagel, D.J.; Finegold, M.J.; Butel, J.S.; Kupsky, W.J.; Garcea, R.L. DNA sequences similar to those of simian virus 40 in ependymomas and choroid plexus tumors of childhood. N. Engl. J. Med. 1992, 326, 988–993. [Google Scholar] [CrossRef] [PubMed]
  38. Martini, F.; Iaccheri, L.; Lazzarin, L.; Carinci, P.; Corallini, A.; Gerosa, M.; Iuzzolino, P.; Barbanti-Brodano, G.; Tognon, M. Sv40 early region and large t antigen in human brain tumors, peripheral blood cells, and sperm fluids from healthy individuals. Cancer Res. 1996, 56, 4820–4825. [Google Scholar] [PubMed]
  39. Huang, H.; Reis, R.; Yonekawa, Y.; Lopes, J.M.; Kleihues, P.; Ohgaki, H. Identification in human brain tumors of DNA sequences specific for SV40 large T antigen. Brain Pathol. 1999, 9, 33–42. [Google Scholar] [CrossRef] [PubMed]
  40. Martini, F.; Lazzarin, L.; Iaccheri, L.; Vignocchi, B.; Finocchiaro, G.; Magnani, I.; Serra, M.; Scotlandi, K.; Barbanti-Brodano, G.; Tognon, M. Different simian virus 40 genomic regions and sequences homologous with sv40 large t antigen in DNA of human brain and bone tumors and of leukocytes from blood donors. Cancer 2002, 94, 1037–1048. [Google Scholar] [CrossRef] [PubMed]
  41. Vilchez, R.A.; Madden, C.R.; Kozinetz, C.A.; Halvorson, S.J.; White, Z.S.; Jorgensen, J.L.; Finch, C.J.; Butel, J.S. Association between simian virus 40 and non-hodgkin lymphoma. Lancet 2002, 359, 817–823. [Google Scholar] [CrossRef] [PubMed]
  42. Shivapurkar, N.; Harada, K.; Reddy, J.; Scheuermann, R.H.; Xu, Y.; McKenna, R.W.; Milchgrub, S.; Kroft, S.H.; Feng, Z.; Gazdar, A.F. Presence of simian virus 40 DNA sequences in human lymphomas. Lancet 2002, 359, 851–852. [Google Scholar] [CrossRef] [PubMed]
  43. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef] [PubMed]
  44. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  45. Levine, A.J. The common mechanisms of transformation by the small DNA tumor viruses: The inactivation of tumor suppressor gene products: P53. Virology 2009, 384, 285–293. [Google Scholar] [CrossRef] [PubMed]
  46. Collot-Teixeira, S.; Bass, J.; Denis, F.; Ranger-Rogez, S. Human tumor suppressor p53 and DNA viruses. Rev. Med. Virol. 2004, 14, 301–319. [Google Scholar] [CrossRef] [PubMed]
  47. Weitzman, M.D.; Lilley, C.E.; Chaurushiya, M.S. Genomes in conflict: Maintaining genome integrity during virus infection. Annu. Rev. Microbiol. 2010, 64, 61–81. [Google Scholar] [CrossRef] [PubMed]
  48. Scheffner, M.; Werness, B.A.; Huibregtse, J.M.; Levine, A.J.; Howley, P.M. The E6 oncoprotein encoded by human papillomavirus types 16 and 18 promotes the degradation of p53. Cell 1990, 63, 1129–1136. [Google Scholar] [CrossRef] [PubMed]
  49. Friborg, J., Jr.; Kong, W.; Hottiger, M.O.; Nabel, G.J. p53 inhibition by the LANA protein of KSHV protects against cell death. Nature 1999, 402, 889–894. [Google Scholar] [PubMed]
  50. Carbone, M.; Rizzo, P.; Grimley, P.M.; Procopio, A.; Mew, D.J.; Shridhar, V.; de Bartolomeis, A.; Esposito, V.; Giuliano, M.T.; Steinberg, S.M.; et al. Simian virus-40 large-T antigen binds p53 in human mesotheliomas. Nat. Med. 1997, 3, 908–912. [Google Scholar] [CrossRef] [PubMed]
  51. Levresse, V.; Renier, A.; Fleury-Feith, J.; Levy, F.; Moritz, S.; Vivo, C.; Pilatte, Y.; Jaurand, M.C. Analysis of cell cycle disruptions in cultures of rat pleural mesothelial cells exposed to asbestos fibers. Am. J. Respir. Cell Mol. Biol. 1997, 17, 660–671. [Google Scholar] [CrossRef] [PubMed]
  52. Butel, J.S.; Lednicky, J.A. Cell and molecular biology of simian virus 40: Implications for human infections and disease. J. Natl. Cancer Inst. 1999, 91, 119–134. [Google Scholar] [CrossRef] [PubMed]
  53. Majumder, M.; Ghosh, A.K.; Steele, R.; Ray, R.; Ray, R.B. Hepatitis C virus NS5A physically associates with p53 and regulates p21/waf1 gene expression in a p53-dependent manner. J. Virol. 2001, 75, 1401–1407. [Google Scholar] [CrossRef] [PubMed]
  54. Lan, K.H.; Sheu, M.L.; Hwang, S.J.; Yen, S.H.; Chen, S.Y.; Wu, J.C.; Wang, Y.J.; Kato, N.; Omata, M.; Chang, F.Y.; et al. HCV NS5A interacts with p53 and inhibits p53-mediated apoptosis. Oncogene 2002, 21, 4801–4811. [Google Scholar] [CrossRef] [PubMed]
  55. Tsai, W.L.; Chung, R.T. Viral hepatocarcinogenesis. Oncogene 2010, 29, 2309–2324. [Google Scholar] [CrossRef] [PubMed]
  56. Hickman, E.S.; Moroni, M.C.; Helin, K. The role of p53 and prb in apoptosis and cancer. Curr. Opin. Genet. Dev. 2002, 12, 60–66. [Google Scholar] [CrossRef] [PubMed]
  57. Hamid, N.A.; Brown, C.; Gaston, K. The regulation of cell proliferation by the papillomavirus early proteins. Cell Mol. Life Sci. 2009, 66, 1700–1717. [Google Scholar] [CrossRef] [PubMed]
  58. Dyson, N.; Howley, P.M.; Munger, K.; Harlow, E. The human papilloma virus-16 E7 oncoprotein is able to bind to the retinoblastoma gene product. Science 1989, 243, 934–937. [Google Scholar] [CrossRef] [PubMed]
  59. Munger, K.; Werness, B.A.; Dyson, N.; Phelps, W.C.; Harlow, E.; Howley, P.M. Complex formation of human papillomavirus e7 proteins with the retinoblastoma tumor suppressor gene product. EMBO J. 1989, 8, 4099–4105. [Google Scholar] [PubMed]
  60. Hwang, S.G.; Lee, D.; Kim, J.; Seo, T.; Choe, J. Human papillomavirus type 16 E7 binds to E2F1 and activates E2F1-driven transcription in a retinoblastoma protein-independent manner. J. Biol. Chem. 2002, 277, 2923–2930. [Google Scholar] [CrossRef] [PubMed]
  61. Duensing, S.; Lee, L.Y.; Duensing, A.; Basile, J.; Piboonniyom, S.; Gonzalez, S.; Crum, C.P.; Munger, K. The human papillomavirus type 16 E6 and E7 oncoproteins cooperate to induce mitotic defects and genomic instability by uncoupling centrosome duplication from the cell division cycle. Proc. Natl. Acad. Sci. USA 2000, 97, 10002–10007. [Google Scholar] [CrossRef] [PubMed]
  62. De Luca, A.; Baldi, A.; Esposito, V.; Howard, C.M.; Bagella, L.; Rizzo, P.; Caputi, M.; Pass, H.I.; Giordano, G.G.; Baldi, F.; et al. The retinoblastoma gene family pRB/p105, p107, pRB2/p130 and simian virus-40 large T-antigen in human mesotheliomas. Nat. Med. 1997, 3, 913–916. [Google Scholar]
  63. Ali, S.H.; DeCaprio, J.A. Cellular transformation by Sv40 large T antigen: Interaction with host proteins. Semin. Cancer Biol. 2001, 11, 15–23. [Google Scholar] [CrossRef] [PubMed]
  64. Seifried, L.A.; Talluri, S.; Cecchini, M.; Julian, L.M.; Mymryk, J.S.; Dick, F.A. pRB-E2F1 complexes are resistant to adenovirus ELA-mediated disruption. J. Virol. 2008, 82, 4511–4520. [Google Scholar] [CrossRef] [PubMed]
  65. Campisi, J.; Kim, S.H.; Lim, C.S.; Rubio, M. Cellular senescence, cancer and aging: The telomere connection. Exp. Gerontol. 2001, 36, 1619–1637. [Google Scholar] [CrossRef] [PubMed]
  66. Klingelhutz, A.J.; Foster, S.A.; McDougall, J.K. Telomerase activation by the E6 gene product of human papillomavirus type 16. Nature 1996, 380, 79–82. [Google Scholar] [CrossRef] [PubMed]
  67. Veldman, T.; Horikawa, I.; Barrett, J.C.; Schlegel, R. Transcriptional activation of the telomerase htert gene by human papillomavirus type 16 E6 oncoprotein. J. Virol. 2001, 75, 4467–4472. [Google Scholar] [CrossRef] [PubMed]
  68. Oh, S.T.; Kyo, S.; Laimins, L.A. Telomerase activation by human papillomavirus type 16 E6 protein: Induction of human telomerase reverse transcriptase expression through myc and GC-rich SP1 binding sites. J. Virol. 2001, 75, 5559–5566. [Google Scholar] [CrossRef] [PubMed]
  69. Veldman, T.; Liu, X.; Yuan, H.; Schlegel, R. Human papillomavirus E6 and myc proteins associate in vivo and bind to and cooperatively activate the telomerase reverse transcriptase promoter. Proc. Natl. Acad. Sci. USA 2003, 100, 8211–8216. [Google Scholar] [CrossRef] [PubMed]
  70. Terrin, L.; Dal Col, J.; Rampazzo, E.; Zancai, P.; Pedrotti, M.; Ammirabile, G.; Bergamin, S.; Rizzo, S.; Dolcetti, R.; de Rossi, A. Latent membrane protein 1 of epstein-barr virus activates the hTERT promoter and enhances telomerase activity in b lymphocytes. J. Virol. 2008, 82, 10175–10187. [Google Scholar] [CrossRef] [PubMed]
  71. Jeang, K.T.; Giam, C.Z.; Majone, F.; Aboud, M. Life, death, and tax: Role of HTLV-I oncoprotein in genetic instability and cellular transformation. J. Biol. Chem. 2004, 279, 31991–31994. [Google Scholar] [CrossRef] [PubMed]
  72. Gabet, A.S.; Mortreux, F.; Charneau, P.; Riou, P.; Duc-Dodon, M.; Wu, Y.; Jeang, K.T.; Wattel, E. Inactivation of htert transcription by tax. Oncogene 2003, 22, 3734–3741. [Google Scholar] [CrossRef] [PubMed]
  73. Foddis, R.; de Rienzo, A.; Broccoli, D.; Bocchetta, M.; Stekala, E.; Rizzo, P.; Tosolini, A.; Grobelny, J.V.; Jhanwar, S.C.; Pass, H.I.; et al. SV40 infection induces telomerase activity in human mesothelial cells. Oncogene 2002, 21, 1434–1442. [Google Scholar] [CrossRef] [PubMed]
  74. Toyooka, S.; Carbone, M.; Toyooka, K.O.; Bocchetta, M.; Shivapurkar, N.; Minna, J.D.; Gazdar, A.F. Progressive aberrant methylation of the RASSF1A gene in simian virus 40 infected human mesothelial cells. Oncogene 2002, 21, 4340–4344. [Google Scholar] [CrossRef] [PubMed]
  75. Toyooka, S.; Pass, H.I.; Shivapurkar, N.; Fukuyama, Y.; Maruyama, R.; Toyooka, K.O.; Gilcrease, M.; Farinas, A.; Minna, J.D.; Gazdar, A.F. Aberrant methylation and simian virus 40 tag sequences in malignant mesothelioma. Cancer Res. 2001, 61, 5727–5730. [Google Scholar] [PubMed]
  76. Cacciotti, P.; Libener, R.; Betta, P.; Martini, F.; Porta, C.; Procopio, A.; Strizzi, L.; Penengo, L.; Tognon, M.; Mutti, L.; et al. Sv40 replication in human mesothelial cells induces HGF/MET receptor activation: A model for viral-related carcinogenesis of human malignant mesothelioma. Proc. Natl. Acad. Sci. USA 2001, 98, 12032–12037. [Google Scholar] [CrossRef] [PubMed]
  77. Bocchetta, M.; Miele, L.; Pass, H.I.; Carbone, M. Notch-1 induction, a novel activity of Sv40 required for growth of Sv40-transformed human mesothelial cells. Oncogene 2003, 22, 81–89. [Google Scholar] [CrossRef] [PubMed]
  78. Pass, H.I.; Mew, D.J.; Carbone, M.; Matthews, W.A.; Donington, J.S.; Baserga, R.; Walker, C.L.; Resnicoff, M.; Steinberg, S.M. Inhibition of hamster mesothelioma tumorigenesis by an antisense expression plasmid to the insulin-like growth factor-1 receptor. Cancer Res. 1996, 56, 4044–4048. [Google Scholar] [PubMed]
  79. Carbone, M.; Pass, H.I.; Miele, L.; Bocchetta, M. New developments about the association of Sv40 with human mesothelioma. Oncogene 2003, 22, 5173–5180. [Google Scholar] [CrossRef] [PubMed]
  80. Chung, Y.L.; Sheu, M.L.; Yen, S.H. Hepatitis C virus NS5A as a potential viral BCL-2 homologue interacts with bax and inhibits apoptosis in hepatocellular carcinoma. Int. J. Cancer 2003, 107, 65–73. [Google Scholar] [CrossRef] [PubMed]
  81. Moore, P.S.; Chang, Y. Antiviral activity of tumor-suppressor pathways: Clues from molecular piracy by KSHV. Trends Genet. 1998, 14, 144–150. [Google Scholar] [CrossRef] [PubMed]
  82. Mosialos, G.; Birkenbach, M.; Yalamanchili, R.; VanArsdale, T.; Ware, C.; Kieff, E. The Epstein-Barr virus transforming protein LMP1 engages signaling proteins for the tumor necrosis factor receptor family. Cell 1995, 80, 389–399. [Google Scholar] [CrossRef]
  83. Karin, M.; Greten, F.R. Nf-kappab: Linking inflammation and immunity to cancer development and progression. Nat. Rev. Immunol. 2005, 5, 749–759. [Google Scholar] [CrossRef] [PubMed]
  84. Zur Hausen, H. Oncogenic DNA viruses. Oncogene 2001, 20, 7820–7823. [Google Scholar]
  85. Colotta, F.; Allavena, P.; Sica, A.; Garlanda, C.; Mantovani, A. Cancer-related inflammation, the seventh hallmark of cancer: Links to genetic instability. Carcinogenesis 2009, 30, 1073–1081. [Google Scholar] [CrossRef] [PubMed]
  86. Coussens, L.M.; Werb, Z. Inflammation and cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef] [PubMed]
  87. Ernst, P.B.; Gold, B.D. The disease spectrum of helicobacter pylori: The immunopathogenesis of gastroduodenal ulcer and gastric cancer. Annu. Rev. Microbiol. 2000, 54, 615–640. [Google Scholar] [CrossRef] [PubMed]
  88. Seeger, C.; Mason, W.S. Hepatitis B virus biology. Microbiol. Mol. Biol. Rev. 2000, 64, 51–68. [Google Scholar] [CrossRef] [PubMed]
  89. Mason, W.S.; Liu, C.; Aldrich, C.E.; Litwin, S.; Yeh, M.M. Clonal expansion of normal-appearing human hepatocytes during chronic hepatitis B virus infection. J. Virol. 2010, 84, 8308–8315. [Google Scholar] [CrossRef] [PubMed]
  90. Jackson, S.; Storey, A. E6 proteins from diverse cutaneous HPV types inhibit apoptosis in response to UV damage. Oncogene 2000, 19, 592–598. [Google Scholar] [CrossRef] [PubMed]
  91. Hoeijmakers, J.H. DNA damage, aging, and cancer. N. Engl. J. Med. 2009, 361, 1475–1485. [Google Scholar] [CrossRef] [PubMed]
  92. Basu, B.; Yap, T.A.; Molife, L.R.; de Bono, J.S. Targeting the DNA damage response in oncology: Past, present and future perspectives. Curr. Opin. Oncol. 2012, 24, 316–324. [Google Scholar] [CrossRef] [PubMed]
  93. Marnett, L.J. Oxyradicals and DNA damage. Carcinogenesis 2000, 21, 361–370. [Google Scholar] [CrossRef] [PubMed]
  94. Valko, M.; Rhodes, C.J.; Moncol, J.; Izakovic, M.; Mazur, M. Free radicals, metals and antioxidants in oxidative stress-induced cancer. Chem. Biol. Interact. 2006, 160, 1–40. [Google Scholar] [CrossRef] [PubMed]
  95. Kudoh, A.; Fujita, M.; Zhang, L.; Shirata, N.; Daikoku, T.; Sugaya, Y.; Isomura, H.; Nishiyama, Y.; Tsurumi, T. Epstein-Barr virus lytic replication elicits ATM checkpoint signal transduction while providing an S-phase-like cellular environment. J. Biol. Chem. 2005, 280, 8156–8163. [Google Scholar] [CrossRef] [PubMed]
  96. Nikitin, P.A.; Yan, C.M.; Forte, E.; Bocedi, A.; Tourigny, J.P.; White, R.E.; Allday, M.J.; Patel, A.; Dave, S.S.; Kim, W.; et al. An ATM/CHK2-mediated DNA damage-responsive signaling pathway suppresses Epstein-Barr virus transformation of primary human B cells. Cell Host Microbe 2010, 8, 510–522. [Google Scholar] [CrossRef] [PubMed]
  97. Shi, Y.; Dodson, G.E.; Shaikh, S.; Rundell, K.; Tibbetts, R.S. Ataxia-telangiectasia-mutated (ATM) is a T-antigen kinase that controls Sv40 viral replication in vivo. J. Biol. Chem. 2005, 280, 40195–40200. [Google Scholar] [CrossRef] [PubMed]
  98. Hein, J.; Boichuk, S.; Wu, J.; Cheng, Y.; Freire, R.; Jat, P.S.; Roberts, T.M.; Gjoerup, O.V. Simian virus 40 large t antigen disrupts genome integrity and activates a DNA damage response via BUB1 binding. J. Virol. 2009, 83, 117–127. [Google Scholar] [CrossRef] [PubMed]
  99. Moody, C.A.; Laimins, L.A. Human papillomaviruses activate the ATM DNA damage pathway for viral genome amplification upon differentiation. PLoS Pathog. 2009, 5. [Google Scholar] [CrossRef] [PubMed]
  100. Moody, C.A.; Fradet-Turcotte, A.; Archambault, J.; Laimins, L.A. Human papillomaviruses activate caspases upon epithelial differentiation to induce viral genome amplification. Proc. Natl. Acad. Sci. USA 2007, 104, 19541–19546. [Google Scholar] [CrossRef] [PubMed]
  101. Williams, V.M.; Filippova, M.; Soto, U.; Duerksen-Hughes, P.J. HPV-DNA integration and carcinogenesis: Putative roles for inflammation and oxidative stress. Future Virol. 2011, 6, 45–57. [Google Scholar] [CrossRef] [PubMed]
  102. Peebles, K.A.; Lee, J.M.; Mao, J.T.; Hazra, S.; Reckamp, K.L.; Krysan, K.; Dohadwala, M.; Heinrich, E.L.; Walser, T.C.; Cui, X.; et al. Inflammation and lung carcinogenesis: Applying findings in prevention and treatment. Expert Rev. Anticancer Ther. 2007, 7, 1405–1421. [Google Scholar] [CrossRef] [PubMed]
  103. Di Bisceglie, A.M. Hepatitis C and hepatocellular carcinoma. Hepatology 1997, 26, 34S–38S. [Google Scholar] [CrossRef] [PubMed]
  104. Brechot, C.; Kremsdorf, D.; Soussan, P.; Pineau, P.; Dejean, A.; Paterlini-Brechot, P.; Tiollais, P. Hepatitis B virus (HBV)-related hepatocellular carcinoma (HCC): Molecular mechanisms and novel paradigms. Pathol. Biol. (Paris) 2010, 58, 278–287. [Google Scholar] [CrossRef]
  105. Bonilla Guerrero, R.; Roberts, L.R. The role of hepatitis B virus integrations in the pathogenesis of human hepatocellular carcinoma. J. Hepatol. 2005, 42, 760–777. [Google Scholar] [CrossRef] [PubMed]
  106. Yoshida, T.; Hanada, T.; Tokuhisa, T.; Kosai, K.; Sata, M.; Kohara, M.; Yoshimura, A. Activation of STAT3 by the hepatitis C virus core protein leads to cellular transformation. J. Exp. Med. 2002, 196, 641–653. [Google Scholar] [CrossRef] [PubMed]
  107. Bromberg, J.; Darnell, J.E., Jr. The role of STATs in transcriptional control and their impact on cellular function. Oncogene 2000, 19, 2468–2473. [Google Scholar] [CrossRef] [PubMed]
  108. Castle, P.E.; Giuliano, A.R. Chapter 4: Genital tract infections, cervical inflammation, and antioxidant nutrients—Assessing their roles as human papillomavirus cofactors. J. Natl. Cancer Inst. Monogr. 2003, 29–34. [Google Scholar]
  109. Castle, P.E.; Hillier, S.L.; Rabe, L.K.; Hildesheim, A.; Herrero, R.; Bratti, M.C.; Sherman, M.E.; Burk, R.D.; Rodriguez, A.C.; Alfaro, M.; et al. An association of cervical inflammation with high-grade cervical neoplasia in women infected with oncogenic human papillomavirus (HPV). Cancer Epidemiol. Biomarkers Prev. 2001, 10, 1021–1027. [Google Scholar] [PubMed]
  110. Kulkarni, S.; Rader, J.S.; Zhang, F.; Liapis, H.; Koki, A.T.; Masferrer, J.L.; Subbaramaiah, K.; Dannenberg, A.J. Cyclooxygenase-2 is overexpressed in human cervical cancer. Clin. Cancer Res. 2001, 7, 429–434. [Google Scholar] [PubMed]
  111. Kim, G.E.; Kim, Y.B.; Cho, N.H.; Chung, H.C.; Pyo, H.R.; Lee, J.D.; Park, T.K.; Koom, W.S.; Chun, M.; Suh, C.O. Synchronous coexpression of epidermal growth factor receptor and cyclooxygenase-2 in carcinomas of the uterine cervix: A potential predictor of poor survival. Clin. Cancer Res. 2004, 10, 1366–1374. [Google Scholar] [CrossRef] [PubMed]
  112. Stanley, M. Immune responses to human papillomavirus. Vaccine 2006, 24, S16–S22. [Google Scholar] [CrossRef] [PubMed]
  113. Finan, R.R.; Musharrafieh, U.; Almawi, W.Y. Detection of chlamydia trachomatis and herpes simplex virus type 1 or 2 in cervical samples in human papilloma virus (HPV)-positive and HPV-negative women. Clin. Microbiol. Infect. 2006, 12, 927–930. [Google Scholar] [PubMed]
  114. Hawes, S.E.; Kiviat, N.B. Are genital infections and inflammation cofactors in the pathogenesis of invasive cervical cancer? J. Natl. Cancer Inst. 2002, 94, 1592–1593. [Google Scholar] [CrossRef]
  115. Jezek, P.; Hlavata, L. Mitochondria in homeostasis of reactive oxygen species in cell, tissues, and organism. Int. J. Biochem. Cell Biol. 2005, 37, 2478–2503. [Google Scholar] [CrossRef] [PubMed]
  116. Cadet, J.; Douki, T.; Ravanat, J.L. Oxidatively generated base damage to cellular DNA. Free Radic. Biol. Med. 2010, 49, 9–21. [Google Scholar] [CrossRef] [PubMed]
  117. Altieri, F.; Grillo, C.; Maceroni, M.; Chichiarelli, S. DNA damage and repair: From molecular mechanisms to health implications. Antioxid. Redox Signal. 2008, 10, 891–937. [Google Scholar] [CrossRef] [PubMed]
  118. De Bont, R.; van Larebeke, N. Endogenous DNA damage in humans: A review of quantitative data. Mutagenesis 2004, 19, 169–185. [Google Scholar] [CrossRef] [PubMed]
  119. Sedelnikova, O.A.; Redon, C.E.; Dickey, J.S.; Nakamura, A.J.; Georgakilas, A.G.; Bonner, W.M. Role of oxidatively induced DNA lesions in human pathogenesis. Mutat. Res. 2010, 704, 152–159. [Google Scholar] [CrossRef] [PubMed]
  120. Riley, P.A. Free radicals in biology: Oxidative stress and the effects of ionizing radiation. Int. J. Radiat. Biol. 1994, 65, 27–33. [Google Scholar] [CrossRef] [PubMed]
  121. Wallace, S.S. Enzymatic processing of radiation-induced free radical damage in DNA. Radiat. Res. 1998, 150, S60–S79. [Google Scholar] [CrossRef] [PubMed]
  122. Huang, L.; Snyder, A.R.; Morgan, W.F. Radiation-induced genomic instability and its implications for radiation carcinogenesis. Oncogene 2003, 22, 5848–5854. [Google Scholar] [CrossRef] [PubMed]
  123. Georgakilas, A.G.; Mosley, W.G.; Georgakila, S.; Ziech, D.; Panayiotidis, M.I. Viral-induced human carcinogenesis: An oxidative stress perspective. Mol. Biosyst. 2010, 6, 1162–1172. [Google Scholar] [CrossRef] [PubMed]
  124. Kryston, T.B.; Georgiev, A.B.; Pissis, P.; Georgakilas, A.G. Role of oxidative stress and DNA damage in human carcinogenesis. Mutat. Res. 2011, 711, 193–201. [Google Scholar] [CrossRef] [PubMed]
  125. Kawanishi, S.; Hiraku, Y.; Pinlaor, S.; Ma, N. Oxidative and nitrative DNA damage in animals and patients with inflammatory diseases in relation to inflammation-related carcinogenesis. Biol. Chem. 2006, 387, 365–372. [Google Scholar] [CrossRef] [PubMed]
  126. Ohshima, H.; Tatemichi, M.; Sawa, T. Chemical basis of inflammation-induced carcinogenesis. Arch. Biochem. Biophys. 2003, 417, 3–11. [Google Scholar] [CrossRef] [PubMed]
  127. Hagen, T.M.; Huang, S.; Curnutte, J.; Fowler, P.; Martinez, V.; Wehr, C.M.; Ames, B.N.; Chisari, F.V. Extensive oxidative DNA damage in hepatocytes of transgenic mice with chronic active hepatitis destined to develop hepatocellular carcinoma. Proc. Natl. Acad. Sci. USA 1994, 91, 12808–12812. [Google Scholar] [CrossRef] [PubMed]
  128. Shimoda, R.; Nagashima, M.; Sakamoto, M.; Yamaguchi, N.; Hirohashi, S.; Yokota, J.; Kasai, H. Increased formation of oxidative DNA damage, 8-hydroxydeoxyguanosine, in human livers with chronic hepatitis. Cancer Res. 1994, 54, 3171–3172. [Google Scholar] [PubMed]
  129. Keasler, V.V.; Hodgson, A.J.; Madden, C.R.; Slagle, B.L. Enhancement of hepatitis C virus replication by the regulatory x protein in vitro and in vivo. J. Virol. 2007, 81, 2656–2662. [Google Scholar] [CrossRef] [PubMed]
  130. Feitelson, M.A.; Reis, H.M.; Tufan, N.L.; Sun, B.; Pan, J.; Lian, Z. Putative roles of hepatitis C X antigen in the pathogenesis of chronic liver disease. Cancer Lett. 2009, 286, 69–79. [Google Scholar] [CrossRef] [PubMed]
  131. Waris, G.; Huh, K.W.; Siddiqui, A. Mitochondrially associated hepatitis B virus X protein constitutively activates transcription factors STAT-3 and NF-kappa B via oxidative stress. Mol. Cell Biol. 2001, 21, 7721–7730. [Google Scholar] [CrossRef] [PubMed]
  132. Cerimele, F.; Battle, T.; Lynch, R.; Frank, D.A.; Murad, E.; Cohen, C.; Macaron, N.; Sixbey, J.; Smith, K.; Watnick, R.S.; et al. Reactive oxygen signaling and mapk activation distinguish Epstein-Barr virus (EBV)-positive versus EBV-negative burkitt’s lymphoma. Proc. Natl. Acad. Sci. USA 2005, 102, 175–179. [Google Scholar] [CrossRef] [PubMed]
  133. Gruhne, B.; Sompallae, R.; Marescotti, D.; Kamranvar, S.A.; Gastaldello, S.; Masucci, M.G. The Epstein-Barr virus nuclear antigen-1 promotes genomic instability via induction of reactive oxygen species. Proc. Natl. Acad. Sci. USA 2009, 106, 2313–2318. [Google Scholar] [CrossRef] [PubMed]
  134. Gruhne, B.; Sompallae, R.; Masucci, M.G. Three Epstein-Barr virus latency proteins independently promote genomic instability by inducing DNA damage, inhibiting DNA repair and inactivating cell cycle checkpoints. Oncogene 2009, 28, 3997–4008. [Google Scholar] [CrossRef] [PubMed]
  135. Williams, V.M.; Filippova, M.; Filippov, V.; Payne, K.J.; Duerksen-Hughes, P. Human papillomavirus type 16 E6* induces oxidative stress and DNA damage. J. Virol. 2014, 88, 6751–6761. [Google Scholar] [CrossRef] [PubMed]
  136. El-Serag, H.B.; Rudolph, K.L. Hepatocellular carcinoma: Epidemiology and molecular carcinogenesis. Gastroenterology 2007, 132, 2557–2576. [Google Scholar] [CrossRef] [PubMed]
  137. Adachi, M.; Ishii, H. Role of mitochondria in alcoholic liver injury. Free Radic. Biol. Med. 2002, 32, 487–491. [Google Scholar] [CrossRef] [PubMed]
  138. Bailey, S.M.; Cunningham, C.C. Contribution of mitochondria to oxidative stress associated with alcoholic liver disease. Free Radic. Biol. Med. 2002, 32, 11–16. [Google Scholar] [CrossRef] [PubMed]
  139. Plummer, M.; Herrero, R.; Franceschi, S.; Meijer, C.J.; Snijders, P.; Bosch, F.X.; de Sanjose, S.; Munoz, N. Smoking and cervical cancer: Pooled analysis of the iarc multi-centric case—Control study. Cancer Causes Control 2003, 14, 805–814. [Google Scholar] [CrossRef] [PubMed]
  140. Deacon, J.M.; Evans, C.D.; Yule, R.; Desai, M.; Binns, W.; Taylor, C.; Peto, J. Sexual behaviour and smoking as determinants of cervical HPV infection and of CIN3 among those infected: A case-control study nested within the manchester cohort. Br. J. Cancer 2000, 83, 1565–1572. [Google Scholar] [CrossRef] [PubMed]
  141. Tollefson, A.K.; Oberley-Deegan, R.E.; Butterfield, K.T.; Nicks, M.E.; Weaver, M.R.; Remigio, L.K.; Decsesznak, J.; Chu, H.W.; Bratton, D.L.; Riches, D.W.; et al. Endogenous enzymes (NOX and ECSOD) regulate smoke-induced oxidative stress. Free Radic. Biol. Med. 2010, 49, 1937–1946. [Google Scholar] [CrossRef] [PubMed]
  142. Moreno, V.; Bosch, F.X.; Munoz, N.; Meijer, C.J.; Shah, K.V.; Walboomers, J.M.; Herrero, R.; Franceschi, S. Effect of oral contraceptives on risk of cervical cancer in women with human papillomavirus infection: The iarc multicentric case-control study. Lancet 2002, 359, 1085–1092. [Google Scholar] [CrossRef] [PubMed]
  143. Appleby, P.; Beral, V.; Berrington de Gonzalez, A.; Colin, D.; Franceschi, S.; Goodhill, A.; Green, J.; Peto, J.; Plummer, M.; Sweetland, S. Cervical cancer and hormonal contraceptives: Collaborative reanalysis of individual data for 16,573 women with cervical cancer and 35,509 women without cervical cancer from 24 epidemiological studies. Lancet 2007, 370, 1609–1621. [Google Scholar] [CrossRef] [PubMed]
  144. Smith, J.S.; Green, J.; Berrington de Gonzalez, A.; Appleby, P.; Peto, J.; Plummer, M.; Franceschi, S.; Beral, V. Cervical cancer and use of hormonal contraceptives: A systematic review. Lancet 2003, 361, 1159–1167. [Google Scholar] [CrossRef] [PubMed]
  145. Autier, P.; Coibion, M.; Huet, F.; Grivegnee, A.R. Transformation zone location and intraepithelial neoplasia of the cervix uteri. Br. J. Cancer 1996, 74, 488–490. [Google Scholar] [CrossRef] [PubMed]
  146. Munoz, N.; Franceschi, S.; Bosetti, C.; Moreno, V.; Herrero, R.; Smith, J.S.; Shah, K.V.; Meijer, C.J.; Bosch, F.X. Role of parity and human papillomavirus in cervical cancer: The IARC multicentric case-control study. Lancet 2002, 359, 1093–1101. [Google Scholar] [CrossRef] [PubMed]
  147. Sethi, S.; Muller, M.; Schneider, A.; Blettner, M.; Smith, E.; Turek, L.; Wahrendorf, J.; Gissmann, L.; Chang-Claude, J. Serologic response to the E4, E6, and E7 proteins of human papillomavirus type 16 in pregnant women. Am. J. Obstet. Gynecol. 1998, 178, 360–364. [Google Scholar] [CrossRef] [PubMed]
  148. Castellsague, X.; Munoz, N. Chapter 3: Cofactors in human papillomavirus carcinogenesis—Role of parity, oral contraceptives, and tobacco smoking. J. Natl. Cancer Inst. Monogr. 2003, 20–28. [Google Scholar]
  149. Pett, M.; Coleman, N. Integration of high-risk human papillomavirus: A key event in cervical carcinogenesis? J. Pathol. 2007, 212, 356–367. [Google Scholar] [CrossRef]
  150. Perz, J.F.; Armstrong, G.L.; Farrington, L.A.; Hutin, Y.J.; Bell, B.P. The contributions of hepatitis B virus and hepatitis C virus infections to cirrhosis and primary liver cancer worldwide. J. Hepatol. 2006, 45, 529–538. [Google Scholar] [CrossRef] [PubMed]
  151. Brechot, C.; Pourcel, C.; Louise, A.; Rain, B.; Tiollais, P. Presence of integrated hepatitis B virus DNA sequences in cellular DNA of human hepatocellular carcinoma. Nature 1980, 286, 533–535. [Google Scholar] [CrossRef] [PubMed]
  152. Shafritz, D.A.; Shouval, D.; Sherman, H.I.; Hadziyannis, S.J.; Kew, M.C. Integration of hepatitis B virus DNA into the genome of liver cells in chronic liver disease and hepatocellular carcinoma. Studies in percutaneous liver biopsies and post-mortem tissue specimens. N. Engl. J. Med. 1981, 305, 1067–1073. [Google Scholar] [CrossRef] [PubMed]
  153. Lee, J.; Kim, D.H.; Lee, S.; Yang, Q.H.; Lee, D.K.; Lee, S.K.; Roeder, R.G.; Lee, J.W. A tumor suppressive coactivator complex of p53 containing ASC-2 and histone h3-lysine-4 methyltransferase MLL3 or its paralogue MLL4. Proc. Natl. Acad. Sci. USA 2009, 106, 8513–8518. [Google Scholar] [CrossRef] [PubMed]
  154. Sung, W.K.; Zheng, H.; Li, S.; Chen, R.; Liu, X.; Li, Y.; Lee, N.P.; Lee, W.H.; Ariyaratne, P.N.; Tennakoon, C.; et al. Genome-wide survey of recurrent HBV integration in hepatocellular carcinoma. Nat. Genet. 2012, 44, 765–769. [Google Scholar] [CrossRef] [PubMed]
  155. Lee, S.G.; Rho, H.M. Transcriptional repression of the human p53 gene by hepatitis B viral x protein. Oncogene 2000, 19, 468–471. [Google Scholar] [CrossRef] [PubMed]
  156. Kremsdorf, D.; Soussan, P.; Paterlini-Brechot, P.; Brechot, C. Hepatitis B virus-related hepatocellular carcinoma: Paradigms for viral-related human carcinogenesis. Oncogene 2006, 25, 3823–3833. [Google Scholar] [CrossRef] [PubMed]
  157. Feitelson, M.A.; Lee, J. Hepatitis B virus integration, fragile sites, and hepatocarcinogenesis. Cancer Lett. 2007, 252, 157–170. [Google Scholar] [CrossRef] [PubMed]
  158. Yoo, Y.D.; Ueda, H.; Park, K.; Flanders, K.C.; Lee, Y.I.; Jay, G.; Kim, S.J. Regulation of transforming growth factor-beta 1 expression by the hepatitis B virus (HBV) X transactivator. Role in HBV pathogenesis. J. Clin. Invest. 1996, 97, 388–395. [Google Scholar] [CrossRef] [PubMed]
  159. Braun, L.; Mead, J.E.; Panzica, M.; Mikumo, R.; Bell, G.I.; Fausto, N. Transforming growth factor beta mrna increases during liver regeneration: A possible paracrine mechanism of growth regulation. Proc. Natl. Acad. Sci. USA 1988, 85, 1539–1543. [Google Scholar] [CrossRef] [PubMed]
  160. Fausto, N.; Mead, J.E. Regulation of liver growth: Protooncogenes and transforming growth factors. Lab. Invest. 1989, 60, 4–13. [Google Scholar] [PubMed]
  161. Nakamura, T.; Tomita, Y.; Hirai, R.; Yamaoka, K.; Kaji, K.; Ichihara, A. Inhibitory effect of transforming growth factor-beta on DNA synthesis of adult rat hepatocytes in primary culture. Biochem. Biophys. Res. Commun. 1985, 133, 1042–1050. [Google Scholar] [CrossRef] [PubMed]
  162. Czaja, M.J.; Weiner, F.R.; Flanders, K.C.; Giambrone, M.A.; Wind, R.; Biempica, L.; Zern, M.A. In vitro and in vivo association of transforming growth factor-beta 1 with hepatic fibrosis. J. Cell Biol. 1989, 108, 2477–2482. [Google Scholar] [CrossRef] [PubMed]
  163. Nakatsukasa, H.; Nagy, P.; Evarts, R.P.; Hsia, C.C.; Marsden, E.; Thorgeirsson, S.S. Cellular distribution of transforming growth factor-beta 1 and procollagen types i, iii, and iv transcripts in carbon tetrachloride-induced rat liver fibrosis. J. Clin. Invest. 1990, 85, 1833–1843. [Google Scholar] [CrossRef] [PubMed]
  164. Tai, P.C.; Suk, F.M.; Gerlich, W.H.; Neurath, A.R.; Shih, C. Hypermodification and immune escape of an internally deleted middle-envelope (m) protein of frequent and predominant hepatitis B virus variants. Virology 2002, 292, 44–58. [Google Scholar] [CrossRef] [PubMed]
  165. Luber, B.; Arnold, N.; Sturzl, M.; Hohne, M.; Schirmacher, P.; Lauer, U.; Wienberg, J.; Hofschneider, P.H.; Kekule, A.S. Hepatoma-derived integrated HBV DNA causes multi-stage transformation in vitro. Oncogene 1996, 12, 1597–1608. [Google Scholar] [PubMed]
  166. Schluter, V.; Meyer, M.; Hofschneider, P.H.; Koshy, R.; Caselmann, W.H. Integrated hepatitis B virus X and 3' truncated preS/S sequences derived from human hepatomas encode functionally active transactivators. Oncogene 1994, 9, 3335–3344. [Google Scholar] [PubMed]
  167. Wang, H.C.; Huang, W.; Lai, M.D.; Su, I.J. Hepatitis B virus pre-S mutants, endoplasmic reticulum stress and hepatocarcinogenesis. Cancer Sci. 2006, 97, 683–688. [Google Scholar] [CrossRef] [PubMed]
  168. Bill, C.A.; Summers, J. Genomic DNA double-strand breaks are targets for hepadnaviral DNA integration. Proc. Natl. Acad. Sci. USA 2004, 101, 11135–11140. [Google Scholar] [CrossRef] [PubMed]
  169. Dandri, M.; Burda, M.R.; Burkle, A.; Zuckerman, D.M.; Will, H.; Rogler, C.E.; Greten, H.; Petersen, J. Increase in de novo HBV DNA integrations in response to oxidative DNA damage or inhibition of poly(ADP-ribosyl)ation. Hepatology 2002, 35, 217–223. [Google Scholar] [CrossRef] [PubMed]
  170. Burkle, A. Physiology and pathophysiology of poly(ADP-ribosyl)ation. Bioessays 2001, 23, 795–806. [Google Scholar] [CrossRef] [PubMed]
  171. Trucco, C.; Oliver, F.J.; de Murcia, G.; Menissier-de Murcia, J. DNA repair defect in poly(ADP-ribose) polymerase-deficient cell lines. Nucleic. Acids. Res. 1998, 26, 2644–2649. [Google Scholar] [CrossRef] [PubMed]
  172. Yoshida, S.; Simbulan, C.M. Interaction of poly(ADP-ribose)polymerase with DNA polymerase alpha. Mol. Cell Biochem. 1994, 138, 39–44. [Google Scholar] [CrossRef] [PubMed]
  173. Meyer, R.; Muller, M.; Beneke, S.; Kupper, J.H.; Burkle, A. Negative regulation of alkylation-induced sister-chromatid exchange by poly(ADP-ribose) polymerase-1 activity. Int. J. Cancer 2000, 88, 351–355. [Google Scholar] [CrossRef] [PubMed]
  174. Murakami, Y.; Saigo, K.; Takashima, H.; Minami, M.; Okanoue, T.; Brechot, C.; Paterlini-Brechot, P. Large scaled analysis of hepatitis B virus (HBV) DNA integration in HBV related hepatocellular carcinomas. Gut 2005, 54, 1162–1168. [Google Scholar] [CrossRef] [PubMed]
  175. Wang, J.; Zindy, F.; Chenivesse, X.; Lamas, E.; Henglein, B.; Brechot, C. Modification of cyclin a expression by hepatitis B virus DNA integration in a hepatocellular carcinoma. Oncogene 1992, 7, 1653–1656. [Google Scholar] [PubMed]
  176. Horikawa, I.; Barrett, J.C. Cis-activation of the human telomerase gene (hTERT) by the hepatitis B virus genome. J. Natl. Cancer Inst. 2001, 93, 1171–1173. [Google Scholar] [CrossRef] [PubMed]
  177. Yaginuma, K.; Kobayashi, H.; Kobayashi, M.; Morishima, T.; Matsuyama, K.; Koike, K. Multiple integration site of hepatitis B virus DNA in hepatocellular carcinoma and chronic active hepatitis tissues from children. J. Virol. 1987, 61, 1808–1813. [Google Scholar] [PubMed]
  178. Barzon, L.; Lavezzo, E.; Militello, V.; Toppo, S.; Palu, G. Applications of next-generation sequencing technologies to diagnostic virology. Int. J. Mol. Sci. 2011, 12, 7861–7884. [Google Scholar] [CrossRef] [PubMed]
  179. Li, S.; Mao, M. Next generation sequencing reveals genetic landscape of hepatocellular carcinomas. Cancer Lett. 2013, 340, 247–253. [Google Scholar] [CrossRef] [PubMed]
  180. Fujimoto, A.; Totoki, Y.; Abe, T.; Boroevich, K.A.; Hosoda, F.; Nguyen, H.H.; Aoki, M.; Hosono, N.; Kubo, M.; Miya, F.; et al. Whole-genome sequencing of liver cancers identifies etiological influences on mutation patterns and recurrent mutations in chromatin regulators. Nat. Genet. 2012, 44, 760–764. [Google Scholar] [CrossRef] [PubMed]
  181. Jiang, Z.; Jhunjhunwala, S.; Liu, J.; Haverty, P.M.; Kennemer, M.I.; Guan, Y.; Lee, W.; Carnevali, P.; Stinson, J.; Johnson, S.; et al. The effects of hepatitis B virus integration into the genomes of hepatocellular carcinoma patients. Genome Res. 2012, 22, 593–601. [Google Scholar] [CrossRef] [PubMed]
  182. Yan, Z.; Cui, K.; Murray, D.M.; Ling, C.; Xue, Y.; Gerstein, A.; Parsons, R.; Zhao, K.; Wang, W. PBAF chromatin-remodeling complex requires a novel specificity subunit, BAF200, to regulate expression of selective interferon-responsive genes. Genes Dev. 2005, 19, 1662–1667. [Google Scholar] [CrossRef] [PubMed]
  183. Li, M.; Zhao, H.; Zhang, X.; Wood, L.D.; Anders, R.A.; Choti, M.A.; Pawlik, T.M.; Daniel, H.D.; Kannangai, R.; Offerhaus, G.J.; et al. Inactivating mutations of the chromatin remodeling gene ARID2 in hepatocellular carcinoma. Nat. Genet. 2011, 43, 828–829. [Google Scholar] [CrossRef] [PubMed]
  184. Guichard, C.; Amaddeo, G.; Imbeaud, S.; Ladeiro, Y.; Pelletier, L.; Maad, I.B.; Calderaro, J.; Bioulac-Sage, P.; Letexier, M.; Degos, F.; et al. Integrated analysis of somatic mutations and focal copy-number changes identifies key genes and pathways in hepatocellular carcinoma. Nat. Genet. 2012, 44, 694–698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Huang, J.; Deng, Q.; Wang, Q.; Li, K.Y.; Dai, J.H.; Li, N.; Zhu, Z.D.; Zhou, B.; Liu, X.Y.; Liu, R.F.; et al. Exome sequencing of hepatitis B virus-associated hepatocellular carcinoma. Nat. Genet. 2012, 44, 1117–1121. [Google Scholar] [CrossRef] [PubMed]
  186. Calleja-Macias, I.E.; Kalantari, M.; Allan, B.; Williamson, A.L.; Chung, L.P.; Collins, R.J.; Zuna, R.E.; Dunn, S.T.; Ortiz-Lopez, R.; Barrera-Saldana, H.A.; et al. Papillomavirus subtypes are natural and old taxa: Phylogeny of human papillomavirus types 44 and 55 and 68a and -b. J. Virol. 2005, 79, 6565–6569. [Google Scholar] [CrossRef] [PubMed]
  187. Ault, K.A. Human papillomavirus vaccines and the potential for cross-protection between related HPV types. Gynecol. Oncol. 2007, 107, S31–S33. [Google Scholar] [CrossRef] [PubMed]
  188. Scheurer, M.E.; Tortolero-Luna, G.; Adler-Storthz, K. Human papillomavirus infection: Biology, epidemiology, and prevention. Int. J. Gynecol. Cancer 2005, 15, 727–746. [Google Scholar] [CrossRef] [PubMed]
  189. Arends, M.J.; Buckley, C.H.; Wells, M. Aetiology, pathogenesis, and pathology of cervical neoplasia. J. Clin. Pathol. 1998, 51, 96–103. [Google Scholar] [CrossRef] [PubMed]
  190. Parkin, D.M.; Bray, F. Chapter 2: The burden of HPV-related cancers. Vaccine 2006, 24, S3/11–S3/25. [Google Scholar] [CrossRef] [PubMed]
  191. World Health Organization. Cancers of the female reproductive organs. In World Cancer Report 2014; Stewart, B.W., Wild, C.P., Eds.; IARC: Lyon, France, 2014; pp. 465–481. [Google Scholar]
  192. Zur Hausen, H. Human papillomaviruses in the pathogenesis of anogenital cancer. Virology 1991, 184, 9–13. [Google Scholar] [CrossRef] [PubMed]
  193. Clifford, G.M.; Rana, R.K.; Franceschi, S.; Smith, J.S.; Gough, G.; Pimenta, J.M. Human papillomavirus genotype distribution in low-grade cervical lesions: Comparison by geographic region and with cervical cancer. Cancer Epidemiol. Biomarkers Prev. 2005, 14, 1157–1164. [Google Scholar] [CrossRef] [PubMed]
  194. Liu, Y.; Tergaonkar, V.; Krishna, S.; Androphy, E.J. Human papillomavirus type 16 E6-enhanced susceptibility of l929 cells to tumor necrosis factor alpha correlates with increased accumulation of reactive oxygen species. J. Biol. Chem. 1999, 274, 24819–24827. [Google Scholar] [CrossRef] [PubMed]
  195. Lowy, D.R.; Kirnbauer, R.; Schiller, J.T. Genital human papillomavirus infection. Proc. Natl. Acad. Sci. USA 1994, 91, 2436–2440. [Google Scholar] [CrossRef] [PubMed]
  196. Wu, X.C.; Watson, M.; Wilson, R.; Saraiya, M.; Cleveland, J.L.; Markowitz, L. Human Papillomavirus-Associated Cancers—United States, 2004–2008. MMWR Morb. Mortal. Wkly. Rep. 2012, 61, 258–261. [Google Scholar] [PubMed]
  197. Vinokurova, S.; Wentzensen, N.; Kraus, I.; Klaes, R.; Driesch, C.; Melsheimer, P.; Kisseljov, F.; Durst, M.; Schneider, A.; von Knebel Doeberitz, M. Type-dependent integration frequency of human papillomavirus genomes in cervical lesions. Cancer Res. 2008, 68, 307–313. [Google Scholar] [CrossRef] [PubMed]
  198. Song, S.H.; Lee, J.K.; Oh, M.J.; Hur, J.Y.; Park, Y.K.; Saw, H.S. Risk factors for the progression or persistence of untreated mild dysplasia of the uterine cervix. Int. J. Gynecol. Cancer 2006, 16, 1608–1613. [Google Scholar] [CrossRef] [PubMed]
  199. Ostor, A.G. Natural history of cervical intraepithelial neoplasia: A critical review. Int. J. Gynecol. Pathol. 1993, 12, 186–192. [Google Scholar] [CrossRef] [PubMed]
  200. Pett, M.R.; Herdman, M.T.; Palmer, R.D.; Yeo, G.S.; Shivji, M.K.; Stanley, M.A.; Coleman, N. Selection of cervical keratinocytes containing integrated hpv16 associates with episome loss and an endogenous antiviral response. Proc. Natl. Acad. Sci. USA 2006, 103, 3822–3827. [Google Scholar] [CrossRef] [PubMed]
  201. Peitsaro, P.; Hietanen, S.; Johansson, B.; Lakkala, T.; Syrjanen, S. Single copy heterozygote integration of HPV 33 in chromosomal band 5p14 is found in an epithelial cell clone with selective growth advantage. Carcinogenesis 2002, 23, 1057–1064. [Google Scholar] [CrossRef] [PubMed]
  202. Von Knebel Doeberitz, M. New markers for cervical dysplasia to visualise the genomic chaos created by aberrant oncogenic papillomavirus infections. Eur. J. Cancer 2002, 38, 2229–2242. [Google Scholar] [CrossRef] [PubMed]
  203. Herdman, M.T.; Pett, M.R.; Roberts, I.; Alazawi, W.O.; Teschendorff, A.E.; Zhang, X.Y.; Stanley, M.A.; Coleman, N. Interferon-beta treatment of cervical keratinocytes naturally infected with human papillomavirus 16 episomes promotes rapid reduction in episome numbers and emergence of latent integrants. Carcinogenesis 2006, 27, 2341–2353. [Google Scholar] [CrossRef] [PubMed]
  204. Hafner, N.; Driesch, C.; Gajda, M.; Jansen, L.; Kirchmayr, R.; Runnebaum, I.B.; Durst, M. Integration of the HPV16 genome does not invariably result in high levels of viral oncogene transcripts. Oncogene 2008, 27, 1610–1617. [Google Scholar] [CrossRef] [PubMed]
  205. Bechtold, V.; Beard, P.; Raj, K. Human papillomavirus type 16 E2 protein has no effect on transcription from episomal viral DNA. J. Virol. 2003, 77, 2021–2028. [Google Scholar] [CrossRef] [PubMed]
  206. Hopman, A.H.; Smedts, F.; Dignef, W.; Ummelen, M.; Sonke, G.; Mravunac, M.; Vooijs, G.P.; Speel, E.J.; Ramaekers, F.C. Transition of high-grade cervical intraepithelial neoplasia to micro-invasive carcinoma is characterized by integration of HPV 16/18 and numerical chromosome abnormalities. J. Pathol. 2004, 202, 23–33. [Google Scholar] [CrossRef] [PubMed]
  207. Guo, M.; Sneige, N.; Silva, E.G.; Jan, Y.J.; Cogdell, D.E.; Lin, E.; Luthra, R.; Zhang, W. Distribution and viral load of eight oncogenic types of human papillomavirus (HPV) and HPV 16 integration status in cervical intraepithelial neoplasia and carcinoma. Mod. Pathol. 2007, 20, 256–266. [Google Scholar] [CrossRef] [PubMed]
  208. Hiraku, Y.; Tabata, T.; Ma, N.; Murata, M.; Ding, X.; Kawanishi, S. Nitrative and oxidative DNA damage in cervical intraepithelial neoplasia associated with human papilloma virus infection. Cancer Sci. 2007, 98, 964–972. [Google Scholar] [CrossRef] [PubMed]
  209. Wei, L.; Gravitt, P.E.; Song, H.; Maldonado, A.M.; Ozbun, M.A. Nitric oxide induces early viral transcription coincident with increased DNA damage and mutation rates in human papillomavirus-infected cells. Cancer Res. 2009, 69, 4878–4884. [Google Scholar] [CrossRef] [PubMed]
  210. Ziech, D.; Franco, R.; Pappa, A.; Panayiotidis, M.I. Reactive oxygen species (ROS)—induced genetic and epigenetic alterations in human carcinogenesis. Mutat. Res. 2011, 711, 167–173. [Google Scholar] [CrossRef] [PubMed]
  211. Yu, T.; Ferber, M.J.; Cheung, T.H.; Chung, T.K.; Wong, Y.F.; Smith, D.I. The role of viral integration in the development of cervical cancer. Cancer Genet. Cytogenet. 2005, 158, 27–34. [Google Scholar] [CrossRef] [PubMed]
  212. Matzner, I.; Savelyeva, L.; Schwab, M. Preferential integration of a transfected marker gene into spontaneously expressed fragile sites of a breast cancer cell line. Cancer Lett. 2003, 189, 207–219. [Google Scholar] [CrossRef] [PubMed]
  213. Kessis, T.D.; Connolly, D.C.; Hedrick, L.; Cho, K.R. Expression of HPV16 E6 or E7 increases integration of foreign DNA. Oncogene 1996, 13, 427–431. [Google Scholar] [PubMed]
  214. Zur Hausen, H. Papillomaviruses causing cancer: Evasion from host-cell control in early events in carcinogenesis. J. Natl. Cancer Inst. 2000, 92, 690–698. [Google Scholar] [CrossRef] [PubMed]
  215. Ronco, L.V.; Karpova, A.Y.; Vidal, M.; Howley, P.M. Human papillomavirus 16 E6 oncoprotein binds to interferon regulatory factor-3 and inhibits its transcriptional activity. Genes Dev. 1998, 12, 2061–2072. [Google Scholar] [CrossRef] [PubMed]
  216. Zerfass, K.; Schulze, A.; Spitkovsky, D.; Friedman, V.; Henglein, B.; Jansen-Durr, P. Sequential activation of cyclin e and cyclin a gene expression by human papillomavirus type 16 E7 through sequences necessary for transformation. J. Virol. 1995, 69, 6389–6399. [Google Scholar] [PubMed]
  217. Band, V.; Zajchowski, D.; Kulesa, V.; Sager, R. Human papilloma virus DNAs immortalize normal human mammary epithelial cells and reduce their growth factor requirements. Proc. Natl. Acad. Sci. USA 1990, 87, 463–467. [Google Scholar] [CrossRef] [PubMed]
  218. Munger, K.; Phelps, W.C. The human papillomavirus E7 protein as a transforming and transactivating factor. Biochim. Biophys. Acta 1993, 1155, 111–123. [Google Scholar] [PubMed]
  219. Wentzensen, N.; Vinokurova, S.; von Knebel Doeberitz, M. Systematic review of genomic integration sites of human papillomavirus genomes in epithelial dysplasia and invasive cancer of the female lower genital tract. Cancer Res. 2004, 64, 3878–3884. [Google Scholar] [CrossRef] [PubMed]
  220. Thorland, E.C.; Myers, S.L.; Gostout, B.S.; Smith, D.I. Common fragile sites are preferential targets for HPV16 integrations in cervical tumors. Oncogene 2003, 22, 1225–1237. [Google Scholar] [CrossRef] [PubMed]
  221. Ferber, M.J.; Thorland, E.C.; Brink, A.A.; Rapp, A.K.; Phillips, L.A.; McGovern, R.; Gostout, B.S.; Cheung, T.H.; Chung, T.K.; Fu, W.Y.; et al. Preferential integration of human papillomavirus type 18 near the c-myc locus in cervical carcinoma. Oncogene 2003, 22, 7233–7242. [Google Scholar] [CrossRef] [PubMed]
  222. Wentzensen, N.; Ridder, R.; Klaes, R.; Vinokurova, S.; Schaefer, U.; Doeberitz, M. Characterization of viral-cellular fusion transcripts in a large series of HPV16 and 18 positive anogenital lesions. Oncogene 2002, 21, 419–426. [Google Scholar] [CrossRef] [PubMed]
  223. Peter, M.; Rosty, C.; Couturier, J.; Radvanyi, F.; Teshima, H.; Sastre-Garau, X. Myc activation associated with the integration of HPV DNA at the MYC locus in genital tumors. Oncogene 2006, 25, 5985–5993. [Google Scholar] [CrossRef] [PubMed]
  224. Xu, B.; Chotewutmontri, S.; Wolf, S.; Klos, U.; Schmitz, M.; Durst, M.; Schwarz, E. Multiplex identification of human papillomavirus 16 DNA integration sites in cervical carcinomas. PLoS One 2013. [Google Scholar] [CrossRef]
  225. Meiring, T.L.; Salimo, A.T.; Coetzee, B.; Maree, H.J.; Moodley, J.; Hitzeroth, I.I.; Freeborough, M.J.; Rybicki, E.P.; Williamson, A.L. Next-generation sequencing of cervical DNA detects human papillomavirus types not detected by commercial kits. Virol. J. 2012. [Google Scholar] [CrossRef] [Green Version]
  226. Conway, C.; Chalkley, R.; High, A.; Maclennan, K.; Berri, S.; Chengot, P.; Alsop, M.; Egan, P.; Morgan, J.; Taylor, G.R.; et al. Next-generation sequencing for simultaneous determination of human papillomavirus load, subtype, and associated genomic copy number changes in tumors. J. Mol. Diagn. 2012, 14, 104–111. [Google Scholar] [CrossRef] [PubMed]
  227. Sastre-Garau, X.; Peter, M.; Avril, M.F.; Laude, H.; Couturier, J.; Rozenberg, F.; Almeida, A.; Boitier, F.; Carlotti, A.; Couturaud, B.; et al. Merkel cell carcinoma of the skin: Pathological and molecular evidence for a causative role of MCV in oncogenesis. J. Pathol. 2009, 218, 48–56. [Google Scholar] [CrossRef] [PubMed]
  228. Shuda, M.; Feng, H.; Kwun, H.J.; Rosen, S.T.; Gjoerup, O.; Moore, P.S.; Chang, Y. T antigen mutations are a human tumor-specific signature for merkel cell polyomavirus. Proc. Natl. Acad. Sci. USA 2008, 105, 16272–16277. [Google Scholar] [CrossRef] [PubMed]
  229. Li, J.; Wang, X.; Diaz, J.; Tsang, S.H.; Buck, C.B.; You, J. Merkel cell polyomavirus large t antigen disrupts host genomic integrity and inhibits cellular proliferation. J. Virol. 2013, 87, 9173–9188. [Google Scholar] [CrossRef] [PubMed]
  230. Demetriou, S.K.; Ona-Vu, K.; Sullivan, E.M.; Dong, T.K.; Hsu, S.W.; Oh, D.H. Defective DNA repair and cell cycle arrest in cells expressing merkel cell polyomavirus t antigen. Int. J. Cancer 2012, 131, 1818–1827. [Google Scholar] [CrossRef] [PubMed]
  231. Zur Hausen, H.; Schulte-Holthausen, H.; Klein, G.; Henle, W.; Henle, G.; Clifford, P.; Santesson, L. Ebv DNA in biopsies of burkitt tumours and anaplastic carcinomas of the nasopharynx. Nature 1970, 228, 1056–1058. [Google Scholar] [CrossRef] [PubMed]
  232. Xing, L.; Kieff, E. Epstein-barr virus bhrf1 micro- and stable RNAs during latency iii and after induction of replication. J. Virol. 2007, 81, 9967–9975. [Google Scholar] [CrossRef] [PubMed]
  233. Kieff, E.; Rickinson, A.B. Epstein-barr virus and its replication. In Fields Virology, 5th ed.; Knipe, D.M., Howley, P.M., Eds.; Lippincott Williams and Wilkins: Philadelphia, PA, USA, 2007; pp. 2603–2654. [Google Scholar]
  234. Polack, A.; Hortnagel, K.; Pajic, A.; Christoph, B.; Baier, B.; Falk, M.; Mautner, J.; Geltinger, C.; Bornkamm, G.W.; Kempkes, B. C-MYC activation renders proliferation of epstein-barr virus (EBV)-transformed cells independent of EBV nuclear antigen 2 and latent membrane protein 1. Proc. Natl. Acad. Sci. USA 1996, 93, 10411–10416. [Google Scholar] [CrossRef] [PubMed]
  235. Kovalchuk, A.L.; Qi, C.F.; Torrey, T.A.; Taddesse-Heath, L.; Feigenbaum, L.; Park, S.S.; Gerbitz, A.; Klobeck, G.; Hoertnagel, K.; Polack, A.; et al. Burkitt lymphoma in the mouse. J. Exp. Med. 2000, 192, 1183–1190. [Google Scholar] [CrossRef] [PubMed]
  236. Li, Z.; van Calcar, S.; Qu, C.; Cavenee, W.K.; Zhang, M.Q.; Ren, B. A global transcriptional regulatory role for c-MYC in burkitt’s lymphoma cells. Proc. Natl. Acad. Sci. USA 2003, 100, 8164–8169. [Google Scholar] [CrossRef] [PubMed]
  237. Young, L.S.; Rickinson, A.B. Epstein-barr virus: 40 years on. Nat. Rev. Cancer 2004, 4, 757–768. [Google Scholar] [CrossRef] [PubMed]
  238. Jox, A.; Rohen, C.; Belge, G.; Bartnitzke, S.; Pawlita, M.; Diehl, V.; Bullerdiek, J.; Wolf, J. Integration of epstein-barr virus in burkitt’s lymphoma cells leads to a region of enhanced chromosome instability. Ann. Oncol. 1997, 8, S131–S135. [Google Scholar] [CrossRef]
  239. Matsuo, T.; Heller, M.; Petti, L.; O’Shiro, E.; Kieff, E. Persistence of the entire epstein-barr virus genome integrated into human lymphocyte DNA. Science 1984, 226, 1322–1325. [Google Scholar] [CrossRef] [PubMed]
  240. Wolf, J.; Pawlita, M.; Jox, A.; Kohls, S.; Bartnitzke, S.; Diehl, V.; Bullerdiek, J. Integration of epstein barr virus near the breakpoint of a translocation 11;19 in a burkitt’s lymphoma cell line. Cancer Genet. Cytogenet. 1993, 67, 90–94. [Google Scholar] [CrossRef] [PubMed]
  241. Delecluse, H.J.; Bartnizke, S.; Hammerschmidt, W.; Bullerdiek, J.; Bornkamm, G.W. Episomal and integrated copies of epstein-barr virus coexist in burkitt lymphoma cell lines. J. Virol. 1993, 67, 1292–1299. [Google Scholar] [PubMed]
  242. Popescu, N.C.; Chen, M.C.; Simpson, S.; Solinas, S.; DiPaolo, J.A. A burkitt lymphoma cell line with integrated epstein-barr virus at a stable chromosome modification site. Virology 1993, 195, 248–251. [Google Scholar] [CrossRef] [PubMed]
  243. Takakuwa, T.; Luo, W.J.; Ham, M.F.; Sakane-Ishikawa, F.; Wada, N.; Aozasa, K. Integration of epstein-barr virus into chromosome 6q15 of burkitt lymphoma cell line (RAJI) induces loss of BACH2 expression. Am. J. Pathol. 2004, 164, 967–974. [Google Scholar] [PubMed]
  244. Yu, M.C.; Yuan, J.M. Epidemiology of nasopharyngeal carcinoma. Semin. Cancer Biol. 2002, 12, 421–429. [Google Scholar] [CrossRef] [PubMed]
  245. Pathmanathan, R.; Prasad, U.; Chandrika, G.; Sadler, R.; Flynn, K.; Raab-Traub, N. Undifferentiated, nonkeratinizing, and squamous cell carcinoma of the nasopharynx. Variants of epstein-barr virus-infected neoplasia. Am. J. Pathol. 1995, 146, 1355–1367. [Google Scholar] [PubMed]
  246. Kripalani-Joshi, S.; Law, H.Y. Identification of integrated epstein-barr virus in nasopharyngeal carcinoma using pulse field gel electrophoresis. Int. J. Cancer 1994, 56, 187–192. [Google Scholar] [CrossRef] [PubMed]
  247. Saslow, D.; Castle, P.E.; Cox, J.T.; Davey, D.D.; Einstein, M.H.; Ferris, D.G.; Goldie, S.J.; Harper, D.M.; Kinney, W.; Moscicki, A.B.; et al. American cancer society guideline for human papillomavirus (HPV) vaccine use to prevent cervical cancer and its precursors. CA Cancer J. Clin. 2007, 57, 7–28. [Google Scholar] [CrossRef] [PubMed]
  248. Garcia-Closas, R.; Castellsague, X.; Bosch, X.; Gonzalez, C.A. The role of diet and nutrition in cervical carcinogenesis: A review of recent evidence. Int. J. Cancer 2005, 117, 629–637. [Google Scholar] [CrossRef] [PubMed]
  249. Manju, V.; Kalaivani Sailaja, J.; Nalini, N. Circulating lipid peroxidation and antioxidant status in cervical cancer patients: A case-control study. Clin. Biochem. 2002, 35, 621–625. [Google Scholar] [CrossRef] [PubMed]
  250. Kim, Y.T.; Kim, J.W.; Choi, J.S.; Kim, S.H.; Choi, E.K.; Cho, N.H. Relation between deranged antioxidant system and cervical neoplasia. Int. J. Gynecol. Cancer 2004, 14, 889–895. [Google Scholar] [CrossRef] [PubMed]
  251. Halliwell, B. Oxidative stress and cancer: Have we moved forward? Biochem. J. 2007, 401, 1–11. [Google Scholar]
  252. Manda, G.N.; Marina, T.; Neagu, T.-M. Reactive oxygen species, cancer and anti-cancer therapies. Curr. Chem. Biol. 2009. [Google Scholar] [CrossRef]
  253. Bjelakovic, G.; Nikolova, D.; Simonetti, R.G.; Gluud, C. Antioxidant supplements for prevention of gastrointestinal cancers: A systematic review and meta-analysis. Lancet 2004, 364, 1219–1228. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Chen, Y.; Williams, V.; Filippova, M.; Filippov, V.; Duerksen-Hughes, P. Viral Carcinogenesis: Factors Inducing DNA Damage and Virus Integration. Cancers 2014, 6, 2155-2186. https://doi.org/10.3390/cancers6042155

AMA Style

Chen Y, Williams V, Filippova M, Filippov V, Duerksen-Hughes P. Viral Carcinogenesis: Factors Inducing DNA Damage and Virus Integration. Cancers. 2014; 6(4):2155-2186. https://doi.org/10.3390/cancers6042155

Chicago/Turabian Style

Chen, Yan, Vonetta Williams, Maria Filippova, Valery Filippov, and Penelope Duerksen-Hughes. 2014. "Viral Carcinogenesis: Factors Inducing DNA Damage and Virus Integration" Cancers 6, no. 4: 2155-2186. https://doi.org/10.3390/cancers6042155

Article Metrics

Back to TopTop