Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Cancer Letters 345 (2014) 174–181

Contents lists available at ScienceDirect

Cancer Letters
journal homepage: www.elsevier.com/locate/canlet

Mini-review

Virus induced inflammation and cancer development


Scott A. Read a, Mark W. Douglas a,b,*
a
Storr Liver Unit, Westmead Millennium Institute, University of Sydney at Westmead Hospital, Sydney, Australia
b
Centre for Infectious Diseases and Microbiology, Marie Bashir Institute for Infectious Diseases and Biosecurity University of Sydney at Westmead Hospital, Sydney, Australia

a r t i c l e i n f o a b s t r a c t

Keywords: Chronic inflammation as a result of viral infection significantly increases the likelihood of cancer devel-
Cancer opment. A handful of diverse viruses have confirmed roles in cancer development and progression, but
Viruses
the list of suspected oncogenic viruses is continually growing. Viruses induce cancer directly and indi-
Inflammation
rectly, by activating inflammatory signalling pathways and cytokines, stimulating growth of infected cells
Review
and inhibiting apoptosis. Although oncogenic viruses induce inflammation by various mechanisms, it is
generally mediated by the MAPK, NFjB and STAT3 signalling pathways. This review will explore the
unique mechanisms by which different oncogenic viruses induce inflammation to promote cancer initi-
ation and progression.
Ó 2013 Elsevier Ireland Ltd. All rights reserved.

1. Introduction are present in a typical cancer genome, including 10–20 in


genes actively contributing to oncogenesis [7,8]. This implies
Viruses contribute significantly to cancer development that cancer development is usually a slow process, largely unaf-
worldwide, with up to 15% of human cancers attributed to chronic fected by acute infections which resolve. For this reason, viruses
viral infection [1,2]. Human papillomavirus (HPV) and hepatitis B and bacteria with the potential to develop chronic infections are
virus (HBV) are the most significant contributors, but the list of significantly over-represented among tumourigenic pathogens
tumourigenic viruses is much longer [1]. Epstein–Barr virus [9].
(EBV), human T-lymphotropic virus-I (HTLV-I), hepatitis C virus Cancer initiation during viral infection can generally be di-
(HCV), Kaposi’s sarcoma herpesvirus (KSHV) and Merkel cell vided into two categories, based on the capacity of the virus
polyomavirus (MCV) all have confirmed roles in cancer induction, to directly contribute to oncogenesis. Oncogenic viruses encode
with a significantly longer list of viruses also suspected to play a proteins that stimulate cell proliferation and/or interfere with
role in tumourigenesis. cell apoptosis, and therefore play a direct role in carcinogene-
The acute inflammatory response to virus infection is an sis [10]. Most viral oncogenes target similar cell growth path-
essential component of the antiviral response, inducing genes ways to counteract the growth arrest that occurs in response
responsible for antiviral activity, immune cell recruitment and to viral infection. Other viral oncogenes inhibit apoptosis
cell fate [3]. Unfortunately, excessive inflammation can arise dur- and immune cell recognition, thus allowing ongoing viral
ing persistent infection and is generally damaging, having muta- replication.
genic effects on the host genome [4]. Chronic inflammation Viral carcinogenesis can also occur indirectly, if chronic infec-
plays a role in a number of disease states, including diabetes, tion leads to oxidative stress and inflammation [10]. Cytokine
arthritis, Alzheimer’s disease, and over a longer period, cancer secretion by the infected cell and surrounding tissues creates an
[5]. In the context of cancer development, chronic inflammation inflammatory milieu which can promote cancer development,
has been linked not only to tumourigenesis, but also to increased principally mediated by IL-1B, TNF-a and IL-6 [4,11]. Inflammatory
cell proliferation, survival, invasion, angiogenesis and metastasis stimuli and direct effects of the virus activate signalling pathways
[5,6]. responsible for cancer development, including NF-jB, STAT3 and
Tumourigenesis usually follows the accumulation of muta- the MAPKs [12,13].
tions in gene coding regions, resulting in uncontrolled cell Inflammation alone can increase the oncogenic potential of a
growth. It is estimated that around 100 coding region mutations cell, but the combination of direct and indirect factors present dur-
ing chronic viral infections explains the high rates of virus-associ-
ated cancers. This review will focus on the role of virus-induced
* Corresponding author. Storr Liver Unit, Westmead Millennium Institute, inflammation in cancer development, highlighting the similarities
University of Sydney at Westmead Hospital, Sydney, Australia. Tel.: +61 298457705. and differences among oncogenic viruses.
E-mail address: mark.douglas@sydney.edu.au (M.W. Douglas).

0304-3835/$ - see front matter Ó 2013 Elsevier Ireland Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.canlet.2013.07.030
S.A. Read, M.W. Douglas / Cancer Letters 345 (2014) 174–181 175

2. Oncogenic viruses of the liver: HBV and HCV HCV core induces lipid accumulation, increases ROS and inflamma-
tion due to lipid peroxidation, and promotes development of HCC
HBV and HCV are both hepatotropic viruses which can establish in transgenic mice [51,52]. Inflammation associated NF-jB not
persistent infection. HCV is unusual among oncogenic viruses, as it only promotes growth and suppresses apoptosis, but stimulates
does not integrate into the host genome or demonstrate viral la- growth factor production, resulting in oncogenesis and cancer pro-
tency, yet sustains chronic infection in approximately 50–80% of gression [53]. Pro-apoptotic JNK signalling is also activated by
untreated individuals [14]. Chronic HCV infection dramatically in- inflammation, but during chronic viral hepatitis is likely inhibited
creases the likelihood of developing liver fibrosis (scarring) and by viral oncoproteins [16]. Since HBV and HCV do not infect the en-
steatosis (fat accumulation), with approximately 20% of patients tire liver, JNK signalling in uninfected ‘‘bystander’’ hepatocytes
developing cirrhosis, and 4% developing hepatocellular carcinoma may result in cell death, as they lack ‘‘protective’’ viral oncopro-
(HCC) [15]. In countries where HBV is not endemic, including the teins [54,55]. The death of uninfected hepatocytes appears to drive
United States, Europe, Egypt and Japan, HCV contributes to 60% compensatory cell proliferation, further stimulating cancer pro-
of diagnosed HCC [16]. Conversely, approximately 54% of HCC gression [56].
cases worldwide are attributed to HBV, particularly in endemic re- Taken together, these data demonstrate that chronic HBV/HCV
gions such as East Asia and sub-Saharan Africa where up to 8% of infection induces an inflammatory phenotype, characterised by
the population have chronic HBV infection [17–19]. perpetual cell death and regeneration in the liver. Antiviral treat-
HCV (Flaviviridae family) and HBV (Hepadnaviridae family) have ment to reduce viral replication can considerably reduce inflam-
significantly different genomes, which contribute to their unique mation in the liver, but is expensive and can be associated with
mechanisms of establishing chronic infection in the liver and even- side effects [57]. There is nonetheless hope for reduced HCC rates
tually cancer. The HCV genome is a single positive RNA strand, in the future as antiviral treatments improve. The development
which is replicated by a low fidelity RNA polymerase, creating a of direct acting antivirals against HCV has increased cure rates to
diverse population of quasispecies capable of evading the immune over 70%. The likely availability of all oral, interferon free options
response [20,21]. In contrast, after entering the nucleus, the double in the next few years would provide effective, convenient treat-
stranded DNA HBV genome forms covalently closed circular DNA ment with fewer side effects, which should increase community
(cccDNA), which is very stable and persists as an episome take up rates and reduce HCC incidence. New oral antivirals against
[18,22]. HBV is capable of integrating into the host genome [23], HBV effectively inhibit viral replication for many years, with min-
which is not required for HBV oncogenesis, but is seen in approx- imal side effects, reducing the risk of HCC. However complete erad-
imately 80% of HBV related HCCs, and increases in frequency with ication of HBV with treatment is still rare, reinforcing the
chronic inflammation [24,25]. importance of HBV vaccination and education programs [58,59].
Both HBV and HCV encode proteins that affect signalling path-
ways mediating cell growth, apoptosis and antiviral immunity, 3. Human T lymphotropic virus (HTLV-1)
thus favouring viral replication and facilitating chronic infection.
Most notable are the HCV proteins core, NS5A and NS3/4A; and Following an exhaustive worldwide search for a human retrovi-
HBV proteins HBx and pre-S2. These proteins restrict immune rec- rus, HTLV-1 was identified in 1980 [60,61]. HTLV-1 was the first
ognition and response, increase cell survival and proliferation, and retrovirus to be associated with human cancer, termed adult T-cell
inhibit apoptosis [16,26]. However, HCC resulting from HCV and leukaemia (ATL) [62]. The HTLV-1 genome is positive sense RNA,
HBV infection typically takes decades to develop, and is often asso- which is reverse transcribed and integrates randomly into the gen-
ciated with cirrhosis and steatosis [18,27,28]. Therefore the main ome of predominantly CD4+ T cells [63]. Following integration, clo-
driving force in HBV and HCV induced liver cancer is most likely nal expansion of the infected T cells is induced by the viral trans-
chronic liver damage, resulting from years of inflammation, oxida- activator protein Tax, which interferes with the NF-jB, AKT, p53
tive stress, cell death and regeneration [16,29,30]. and pRb signalling pathways, to name a few [64]. Although Tax
A major source of harmful, pro-inflammatory cytokines during is required to immortalise cells in vitro, it is often silenced follow-
chronic HBV/HCV infection is liver infiltrating lymphocytes [31]. ing progression of ATL in vivo, and is seen in only 40% of cases
Viral persistence causes enrichment in the liver of both pro-inflam- [65,66]. Conversely, viral HTLV-1 basic leucine zipper factor
matory T cells and innate immune lymphocytes, including natural (HBZ) transcripts are found in all ATLs, and it is thought that HBZ
killer (NK) cells [32,33]. Many infiltrating lymphocytes display a may be responsible for maintaining the leukaemic state [67].
high level of activation and low level of specificity, resulting in HTLV-1 Tax and HBZ contribute directly to T cell oncogenesis
the inflammatory pathologies associated with chronic infection by increasing growth potential, modulating immune recognition
[31]. Following HBV and HCV infection, hepatocytes and infiltrat- and increasing genetic instability. Nonetheless, HTLV-1 induced
ing lymphocytes secrete increased amounts of TNF-a, IL-6 and ATL is rare, occurring in only 5–10% of infected individuals over
IL-1b, all of which are associated with HCC development and pro- a lifetime [60,63]. Persistent activation of NF-jB by the Tax pro-
gression [34–39]. tein occurs following HTLV-1 infection in vitro and in vivo, and
Inflammation and leukocyte infiltration trigger activation of contributes significantly to the oncogenic potential of the virus
resident hepatic stellate cells, making them responsive to trans- [68,69]. Tax activates IKK, which phosphorylates the IjB regula-
forming growth factor b (TGF-b) [40,41]. TGF-b promotes hepatic tors of NF-jB, targeting them for proteosomal degradation and
fibrogenesis and apoptosis, and is associated with HCC carcinogen- activating NF-jB [70–72]. HBZ on the other hand does not affect
esis, progression and prognosis [42,43]. Interestingly, both HBV the regulatory components of NFjB signalling, and instead inhib-
and HCV increase TGF-b mediated signalling, but shift it from a tu- its the canonical NF-jB pathway by interacting directly with p65,
mour suppressive (apoptotic) phenotype to a proliferative, pro-fi- leaving the non-canonical pathway stimulated by the Tax protein
brotic phenotype [41,44,45]. alone [73].
HBV and HCV induced inflammation is tightly linked to oxida- HTLV-1 stimulates the expression of a number of cytokines that
tive stress [18,30]. Both HBV and HCV infection induce endoplas- stimulate both the canonical and non-canonical NF-jB pathways,
mic reticulum (ER) stress, which increases intracellular levels of including IL-1, IFN-c, TNF-a/b, and Bcl-3 [74–78]. However, few
reactive oxygen species (ROS), leading to an increase in inflamma- studies have examined the association between inflammatory
tory gene expression by activating NF-jB, AP-1 and STAT3 [46–50]. cytokine production and ATL initiation and progression. Elevated
176 S.A. Read, M.W. Douglas / Cancer Letters 345 (2014) 174–181

TNF-a expression, as a result of genetic polymorphism of the TNF- affinity for the IL-10R and functions primarily to induce B cell pro-
a gene, has been shown to enhance the susceptibility of some Jap- liferation [107,108].
anese patients to ATL [79]. Inflammatory cytokines IFN-c and TNF- Approximately a third of HL cases are associated with EBV
a produced by infected CD4+ T cells have in fact been shown to infection, however HL by nature is mediated by inflammation
weaken the anti-inflammatory Th2 immune response to infection, [109]. Inhibitors of NF-jB and STAT3 significantly decrease the
which may indirectly promote the oncogenic potential of HTLV-1 viability of lymphoma cells, both in vitro and in LMP1 transgenic
[80,81]. mice, supporting the role of inflammatory stimuli in EBV medi-
In summary, it appears that HTLV-1 associated ATL results pri- ated cancer progression [110,111]. The particular reliance of
marily from proviral oncogene expression, but may be facilitated HRS cell on inflammatory mediators and immune cell recruit-
by virus-induced inflammation. Viral Tax and HBZ proteins are ex- ment for survival warrants further study, supporting the clinical
pressed at different phases during the progression of ATL, and ap- investigation of inflammatory signalling inhibitors for HL
pear to have opposing effects on a number of cell growth, immune treatment.
response and inflammatory pathways [82]. Additional research is
required to further understand the influence of these viral proteins
as well as external inflammatory stimuli on ATL development and 5. KSHV
progression.
Like EBV, KSHV is a gamma herpes virus that usually remains
latent as a nuclear episome, with occasional viral reactivation fol-
lowed by lytic replication. However unlike EBV, KSHV is usually
4. EBV oncogenic only in immunosuppressed patients, either as a result
of AIDS or following organ transplant [112–114]. KSHV is associ-
EBV infects 90% of the world’s population, primarily as an ated with the development of Kaposi’s sarcoma (KS), primary effu-
asymptomatic infection in childhood [83]. Infection at a later stage sion lymphoma (PEL) and multicentric Castleman’s disease (MCD)
(often adolescence) results in infectious mononucleosis, character- [115]. KSHV infection is more common in HIV infected homosexual
ised by fever, pharyngitis and/or tonsillitis, and tender cervical men, with prevalence rates of up to 30–65%, compared with 10–
lymphadenopathy [84]. EBV preferentially infects B lymphocytes, 15% in the general population [116].
but has also been shown to infect other cells types, primarily epi- The KHSV DNA genome consists of approximately 140 kilobases
thelial cells [85,86]. Following acute infection, EBV establishes la- and encodes over 90 open reading frames (ORFs) [117]. A signifi-
tent infection in B lymphocytes, which persists for life. Infected cant number of KSHV genes (representing 30% of the viral genome)
cells carry multiple episomal copies of the EBV genome, which ex- encode viral homologues of human genes, many of which interact
press different subsets of viral genes, depending on the degree of with inflammatory and immune signalling pathways, including
latency [87]. viral interferon regulatory factors (vIRFs) 1–4, viral interleukin 6
EBV infection is linked to a number of cancers, including Hodg- (vIL-6) and viral Fas-associated death domain-like IL-1-converting
kin’s lymphoma (HL), Burkitt’s lymphoma (BL), lymphomas in enzyme inhibitory protein (vFLIP) [118]. vFLIP in particular has
immunosuppressed individuals, and a number of carcinomas been shown to induce NF-jB activation, possibly through the TNFR
[87]. Interestingly, none of the EBV associated cancers are strongly signalling pathway, which increases survival and tumour forma-
linked to inflammation apart from HL, which is strongly associated tion in KSHV induced lymphomas [119–122]. Interestingly, vFLIP
with an inflammatory phenotype [88,89]. and NF-jB activation correlate with in vitro expression of the che-
A number of viral genes have been linked to the oncogenic ef- mokine CCL20 and its receptor CCL6, both of which are upregulated
fect of EBV, but only the viral proteins latent membrane protein in clinical samples of KS [123]. CCL20 is responsible for dendritic
1 (LMP1) and Epstein–Barr nuclear antigen 2 (EBNA2) induce cell cell and lymphocyte recruitment, suggesting that it may contribute
immortalisation [90–93]. EBNA2 regulates viral gene expression to the inflammatory infiltrate observed in KS lesions prior to trans-
and activates the notch signalling pathway to stimulate cell prolif- formation [123,124].
eration, but is highly immunogenic and is rarely expressed in lym- The KS protein vIL-6 is a homologue of the human IL-6 protein;
phomas [88,94]. LMP1 activates the TNF-a signalling pathway, by it is expressed at low levels during latent infection but increases
interacting with signalling intermediates downstream of the during lytic replication [125]. vIL-6 has been shown to promote
TNF-a receptor (TNFR), subsequently activating NF-jB and result- angiogenesis and tumour growth in a T cell deficient mouse model,
ing in production of IL-6 [95,96]. In fact, NF-jB appears to be but its limited expression during KSHV latency suggests that it
central to the development and pathogenesis of HL, inducing the contributes differently to development of KS, PEL and MCD [126].
expression of a number of pro-inflammatory and immune stimula- In KS tissue, vIL-6 is rarely detected; however cellular IL-6 is ex-
tory cytokines [97,98]. pressed, induced by viral proteins vFLIP and vPCR, via activation
Hodgkin’s lymphoma is characterised by malignant B cells of JNK signalling [127,128]. PEL tissues contain high levels of vIL-
termed Hodgkin and Reed–Sternberg (HRS) cells, surrounded by 6, stimulated by hypoxia-induced reactivation of KSHV [129,130],
a large infiltrate of diverse, non-malignant inflammatory cells. as well as cellular IL-6 and IL-10, which promote autocrine growth
Cytokine signalling between HRS and the inflammatory infiltrate of PEL cells [130,131]. The waxing and waning inflammation that
is crucial to the survival of HRS cells, as cell death results from often occurs in patients with MCD is attributed to vIL-6, IL-6 and
its disruption in vitro [99]. Inflammatory cytokines, including IL- IL-10, expression of which correlate with the degree of inflamma-
5, IL-6, IL-7, IL-13, TNF-a, CCL5 (RANTES) and CCL28, contribute di- tion [132].
rectly to disease progression. These cytokines recruit eosinophils, Thus, although KS is clearly associated with KSHV infection,
monocytes, T cells and mast cells to the site of tumour develop- immunosuppression favouring viral replication and oncogene
ment, inducing their differentiation and proliferation [100–104]. expression may be the most important contributor to cancer devel-
EBV infected HRS cells also express significantly more IL-10 than opment. Combined antiretroviral therapy (cART) for HIV infected
uninfected lymphocytes, inhibiting T cell activation and prolifera- individuals has significantly reduced KS incidence, following an
tion [105,106]. A viral homologue of IL-10 (vIL-10) is also ex- increase of the CD4+ cell count [133]. Conversely, when KS occurs
pressed by EBV infected cells, but demonstrates a much weaker in the post-transplant setting due to cyclosporine induced
S.A. Read, M.W. Douglas / Cancer Letters 345 (2014) 174–181 177

immunosuppression, cyclosporine can be changed to rapamycin, a fected carcinoma cells, but with minimal downstream activation
mammalian target of rapamycin (mTOR) inhibitor, which of STAT3 [160]. It has been suggested that E6/E7 induced inter-
demonstrates anti-tumour as well as immunosuppressive activity ruption of IL-6 signalling suppresses expression of monocyte che-
[134]. mo-attractant protein-1, limiting monocyte recruitment
[161,162]. Nonetheless, like TNF-a, elevated levels of IL-6 contrib-
6. HPV ute to the increased incidence of cervical cancer associated with
HPV infection [163–165].
HPVs belong to the Papillomaviridae family, and are a diverse Unlike other oncogenic viruses, HPV replicates as the infected
group of small non-enveloped dsDNA viruses. The 120 types of keratinocytes differentiate, without virus particle release into the
HPV identified to date can be subdivided based on their tissue tro- bloodstream, and thus minimal immune recognition. To facilitate
pism, infecting either cutaneous or internal mucosal surfaces this replication strategy, HPV down-regulates inflammatory signal-
[135]. The mucosal HPV types have further been divided based ling, but inflammatory cytokines still seem to contribute to cancer
on their oncogenic potential, with low risk types generally causing progression. In high risk individuals with HPV, co-infection with
benign genital warts, and high risk types associated with neoplasia other sexually transmitted infections induces inflammation, and
and subsequent cancers. High risk HPV types 16 and 18 account for may therefore contribute to cancer initiation and progression
approximately 50% and 20% of cervical and ano-genital cancers, [166].
respectively [136].
The potent tumourigenic potential of HPV is primarily the re- 7. Conclusions
sult of its tissue tropism, followed by its ability to establish per-
sistent infection. HPV infects basal keratinocytes, which become Chronic viral infection is increasingly being recognised as a
mitotically inactive during differentiation, and therefore under leading cause of cancer worldwide. As well as the direct oncogenic
highly regulated transcriptional control. To facilitate the replica- role of some viruses, ongoing inflammation stimulated by chronic
tion of its own DNA, HPV initiates cell growth by stimulating the viral infections also contributes significantly to tumour formation.
production and activation of cellular transcription factors In addition to the viruses discussed above, mounting evidence now
[137,138]. Growth deregulation seems to be initiated by the implicates other viruses in the development of human cancers.
HPV proteins E6 and E7, which interact with cellular proteins These include Merkel cell polyomavirus, which has been associated
p53 and pRb, are present in all cervical tumour samples, and with the majority of Merkel cell carcinomas following immunosup-
whose expression can immortalise primary human keratinocytes pression [167,168], as well as members of the Adenoviridae family,
in vitro [137,139–141]. Fortunately, the expression of E6 and E7 human mammary tumour virus (HMTV), and human endogenous
is not sufficient to immortalise cells in vivo, and additional host retroviruses (HERVs) [169].
cell mutations are required [10,142,143]. E5, E6 and E7 from the Identifying viruses that induce cancer provides an opportu-
highly oncogenic HPV type 16 have also been shown to induce nity for cancer prevention, either by preventing or treating
cyclooxygenase (COX) expression, with subsequent synthesis of the virus infection. Vaccines against HBV and HPV are becom-
prostaglandins (PG) [144–146]. Prostaglandins are lipid media- ing increasingly available worldwide, and have significantly re-
tors of tissue remodelling associated with inflammation, and duced transmission of these viruses [58,170]. Long term
have been shown in many models to contribute to cancer forma- antiviral therapy against HBV inhibits virus replication and re-
tion and progression [147]. duces the risk of HCC development and recurrence [171–173].
Following HPV infection, numerous cytokines are produced by Public Health interventions can also reduce transmission of
the infected and neighbouring keratinocytes, as well as by infil- viruses such as HCV and HTLV-1, but have proven less effective
trating immune cells, including macrophages, natural killer cells against endemic viruses such as EBV. Conversely, for viruses
and lymphocytes [148]. Type I/II IFNs, TNF-a and IL-1 are all where vaccines and/or specific antiviral treatments are not
produced in response to HPV infection, and all demonstrate anti- available, drugs targeting host inflammatory pathways, medi-
viral activity in vitro, yet HPV is able to avoid the immune sys- ated by NF-jB and STAT3, can significantly reduce cancer
tem and establish persistent infection in a small subgroup of development and progression [174,175]. The benefits of low
people [149–151]. It remains poorly understood why some pa- dose aspirin in particular are becoming increasingly apparent
tients develop chronic HPV infection, but the mechanisms of vir- for the prevention and treatment of multiple cancers
al persistence likely include avoiding immune recognition and [174,176]. Unfortunately, often the countries with the highest
inhibiting the immune response. HPV proteins E6 and E7 are rates of chronic viral infections are those least able to afford
most likely involved, as they can inhibit IFN promoter activation, expensive antiviral treatment, which is typically prolonged. In-
through interactions with a number of signalling intermediates, creased dietary intake of natural anti-inflammatory agents
including interferon regulatory factors (IRFs) and STATs (re- may provide some protection against virus induced cancers,
viewed in [152]). as has been demonstrated with curcumin in turmeric and res-
Interestingly, it has been suggested that both TNF-a and IL-6 veratrol in grapes [177,178]. Furthermore, inhibitors of inflam-
signalling are interrupted in HPV-16 infected cells, thus increas- mation often target pathways utilised by oncogenic viruses to
ing their growth potential. Resistance to TNF-a and resulting establish persistent infections, suggesting that they may not
NF-jB activation appears to be mediated by both E6 and E7, only reduce the incidence of virus mediated cancer, but also
which limit respectively its apoptotic and anti-proliferative ef- synergise with ongoing antiviral treatment.
fects [153–155]. Other reports suggest that HPV-18 but not In summary, many cancers are caused by viral infection, either
HPV-16 immortalized cell lines are resistant to TNF-a [156,157]. through direct oncogenic effects or via virus-induced inflammation
This is further complicated by evidence suggesting that TNF-a (see Figure 1). A better understanding of the interactions between
stimulates the proliferation of both HPV-16 and -18 transformed viruses and their hosts therefore provides an opportunity to pre-
keratinocytes [158,159].The role of TNF-a in HPV oncogenesis vent many cancer deaths around the world, through a combination
clearly requires further study in relation to viral genotype and cell of Public Health interventions, anti-inflammatory treatment and
line resistance to TNF-a. High levels of IL-6 are seen in HPV in- specific antiviral drugs.
178 S.A. Read, M.W. Douglas / Cancer Letters 345 (2014) 174–181

Fig. 1. Inflammation resulting from chronic viral infection contributes to cancer development. Following infection, oncogenic viruses utilise a number of mechanisms to
evade the host immune system, occasionally resulting in chronic infection. Chronic activation of inflammatory signalling can result, mediated directly by the virus or
indirectly as a result of viral propagation. Direct activation of inflammation is generally mediated by viral proteins capable of activating inflammatory signalling cascades and/
or host inflammatory cytokines. Conversely, some oncogenic viruses possess viral homologues of human inflammatory cytokines including vIL-6 (KSHV) and vIL-10 (EBV).
The hepatotrophic viruses HBV and HCV especially, induce significant indirect inflammation as a result of viral replication induced ROS, lipid accumulation and immune
recognition. As a result, a substantial amount of cell death and resulting hepatocyte regeneration occurs that contributes to cancer progression.

Conflict of interest G.R. Ordonez, L.J. Mudie, C. Latimer, S. Edkins, L. Stebbings, L. Chen, M. Jia, C.
Leroy, J. Marshall, A. Menzies, A. Butler, J.W. Teague, J. Mangion, Y.A. Sun, S.F.
McLaughlin, H.E. Peckham, E.F. Tsung, G.L. Costa, C.C. Lee, J.D. Minna, A.
None. Gazdar, E. Birney, M.D. Rhodes, K.J. McKernan, M.R. Stratton, P.A. Futreal, P.J.
Campbell, A small-cell lung cancer genome with complex signatures of
tobacco exposure, Nature 463 (2010) 184–190.
References [8] P.J. Stephens, D.J. McBride, M.L. Lin, I. Varela, E.D. Pleasance, J.T. Simpson, L.A.
Stebbings, C. Leroy, S. Edkins, L.J. Mudie, C.D. Greenman, M. Jia, C. Latimer,
[1] D.M. Parkin, The global health burden of infection-associated cancers in the J.W. Teague, K.W. Lau, J. Burton, M.A. Quail, H. Swerdlow, C. Churcher, R.
year, Int. J. Cancer 118 (2006) (2002) 3030–3044. Natrajan, A.M. Sieuwerts, J.W. Martens, D.P. Silver, A. Langerod, H.E. Russnes,
[2] H. zur Hausen, Viruses in human cancers, Curr. Sci. 81 (2001) 523–527. J.A. Foekens, J.S. Reis-Filho, L. van’t Veer, A.L. Richardson, A.L. Borresen-Dale,
[3] S.I. Grivennikov, F.R. Greten, M. Karin, Immunity, inflammation, and cancer, P.J. Campbell, P.A. Futreal, M.R. Stratton, Complex landscapes of somatic
Cell 140 (2010) 883–899. rearrangement in human breast cancer genomes, Nature 462 (2009) 1005–
[4] A. Kuraishy, M. Karin, S.I. Grivennikov, Tumor promotion via injury- and 1010.
death-induced inflammation, Immunity 35 (2011) 467–477. [9] C. de Martel, S. Franceschi, Infections and cancer: established associations and
[5] B.B. Aggarwal, S. Shishodia, S.K. Sandur, M.K. Pandey, G. Sethi, Inflammation new hypotheses, Crit. Rev. Oncol. Hematol. 70 (2009) 183–194.
and cancer: how hot is the link?, Biochem Pharmacol. 72 (2006) 1605–1621. [10] P.S. Moore, Y. Chang, Why do viruses cause cancer? Highlights of the first
[6] A. Mantovani, Cancer: inflammation by remote control, Nature 435 (2005) century of human tumour virology, Nat. Rev. Cancer 10 (2010) 878–
752–753. 889.
[7] E.D. Pleasance, P.J. Stephens, S. O’Meara, D.J. McBride, A. Meynert, D. Jones, [11] Y. Ben-Neriah, M. Karin, Inflammation meets cancer, with NF-kappaB as the
M.L. Lin, D. Beare, K.W. Lau, C. Greenman, I. Varela, S. Nik-Zainal, H.R. Davies, matchmaker, Nat. Immunol. 12 (2011) 715–723.
S.A. Read, M.W. Douglas / Cancer Letters 345 (2014) 174–181 179

[12] Y. Fan, R. Mao, J. Yang, NF-kappaB and STAT3 signaling pathways factor-beta1 to immunity and prognosis in unresectable hepatocellular
collaboratively link inflammation to cancer, Protein Cell 4 (2013) 176–185. carcinoma, Liver Int. 24 (2004) 21–28.
[13] T. Chiba, H. Marusawa, T. Ushijima, Inflammation-associated cancer [44] K. Matsuzaki, M. Murata, K. Yoshida, G. Sekimoto, Y. Uemura, N. Sakaida, M.
development in digestive organs: mechanisms and roles for genetic and Kaibori, Y. Kamiyama, M. Nishizawa, J. Fujisawa, K. Okazaki, T. Seki, Chronic
epigenetic modulation, Gastroenterology 143 (2012) 550–563. inflammation associated with hepatitis C virus infection perturbs hepatic
[14] S.M. Kamal, Acute hepatitis C: a systematic review, Am. J. Gastroenterol. 103 transforming growth factor beta signaling, promoting cirrhosis and
(2008) 1283–1297. hepatocellular carcinoma, Hepatology 46 (2007) 48–57.
[15] D.L. Thomas, L.B. Seeff, Natural history of hepatitis, C. Clin. Liver. Dis. 9 (2005) [45] S. Battaglia, N. Benzoubir, S. Nobilet, P. Charneau, D. Samuel, A.L. Zignego, A.
383–398. Atfi, C. Brechot, M.F. Bourgeade, Liver cancer-derived hepatitis C virus core
[16] A. Arzumanyan, H.M. Reis, M.A. Feitelson, Pathogenic mechanisms in HBV- proteins shift TGF-beta responses from tumor suppression to epithelial–
and HCV-associated hepatocellular carcinoma, Nat. Rev. Cancer 13 (2013) mesenchymal transition, PLoS One 4 (2009) e4355.
123–135. [46] G. Gong, G. Waris, R. Tanveer, A. Siddiqui, Human hepatitis C virus NS5A
[17] J. Ferlay, H.R. Shin, F. Bray, D. Forman, C. Mathers, D.M. Parkin, Estimates of protein alters intracellular calcium levels, induces oxidative stress, and
worldwide burden of cancer in 2008: GLOBOCAN, Int. J. Cancer 127 (2010) activates STAT-3 and NF-kappa B, Proc. Natl. Acad. Sci. USA 98 (2001) 9599–
(2008) 2893–2917. 9604.
[18] G. Fallot, C. Neuveut, M.A. Buendia, Diverse roles of hepatitis B virus in liver [47] H.C. Wang, W. Huang, M.D. Lai, I.J. Su, Hepatitis B virus pre-S mutants,
cancer, Curr. Opin. Virol. 2 (2012) 467–473. endoplasmic reticulum stress and hepatocarcinogenesis, Cancer Sci. 97
[19] L.R. Roberts, G.J. Gores, Hepatocellular carcinoma: molecular pathways and (2006) 683–688.
new therapeutic targets, Semin. Liver Dis. 25 (2005) 212–225. [48] H.K. Cho, K.J. Cheong, H.Y. Kim, J. Cheong, Endoplasmic reticulum stress
[20] T. Suzuki, H. Aizaki, K. Murakami, I. Shoji, T. Wakita, Molecular biology of induced by hepatitis B virus X protein enhances cyclo-oxygenase 2
hepatitis C virus, J. Gastroenterol. 42 (2007) 411–423. expression via activating transcription factor 4, Biochem. J. 435 (2011)
[21] P. Simmonds, Genetic diversity and evolution of hepatitis C virus – 15 years 431–439.
on, J. Gen. Virol. 85 (2004) 3173–3188. [49] I. Qadri, M. Iwahashi, J.M. Capasso, M.W. Hopken, S. Flores, J. Schaack, F.R.
[22] D. Ganem, A.M. Prince, Hepatitis B virus infection – natural history and Simon, Induced oxidative stress and activated expression of manganese
clinical consequences, N. Engl. J. Med. 350 (2004) 1118–1129. superoxide dismutase during hepatitis C virus replication: role of JNK, p38
[23] M. Nassal, Hepatitis B viruses: reverse transcription a different way, Virus MAPK and AP-1, Biochem. J. 378 (2004) 919–928.
Res. 134 (2008) 235–249. [50] K.D. Tardif, K. Mori, A. Siddiqui, Hepatitis C virus subgenomic replicons
[24] R. Bonilla Guerrero, L.R. Roberts, The role of hepatitis B virus integrations in induce endoplasmic reticulum stress activating an intracellular signaling
the pathogenesis of human hepatocellular carcinoma, J. Hepatol. 42 (2005) pathway, J. Virol. 76 (2002) 7453–7459.
760–777. [51] K. Moriya, H. Fujie, Y. Shintani, H. Yotsuyanagi, T. Tsutsumi, K. Ishibashi, Y.
[25] C. Brechot, Pathogenesis of hepatitis B virus-related hepatocellular Matsuura, S. Kimura, T. Miyamura, K. Koike, The core protein of hepatitis C
carcinoma: old and new paradigms, Gastroenterology 127 (2004) S56–61. virus induces hepatocellular carcinoma in transgenic mice, Nat. Med. 4
[26] S.F. Wieland, F.V. Chisari, Stealth and cunning: hepatitis B and hepatitis C (1998) 1065–1067.
viruses, J. Virol. 79 (2005) 9369–9380. [52] M. Okuda, K. Li, M.R. Beard, L.A. Showalter, F. Scholle, S.M. Lemon, S.A.
[27] S. Caldwell, S.H. Park, The epidemiology of hepatocellular cancer: from the Weinman, Mitochondrial injury, oxidative stress, and antioxidant gene
perspectives of public health problem to tumor biology, J. Gastroenterol. 44 expression are induced by hepatitis C virus core protein, Gastroenterology
(Suppl. 19) (2009) 96–101. 122 (2002) 366–375.
[28] J.R. Pekow, A.K. Bhan, H. Zheng, R.T. Chung, Hepatic steatosis is associated [53] S.P. Hussain, L.J. Hofseth, C.C. Harris, Radical causes of cancer, Nat. Rev. Cancer
with increased frequency of hepatocellular carcinoma in patients with 3 (2003) 276–285.
hepatitis C-related cirrhosis, Cancer 109 (2007) 2490–2496. [54] J.D. Stiffler, M. Nguyen, J.A. Sohn, C. Liu, D. Kaplan, C. Seeger, Focal
[29] B. Rehermann, M. Nascimbeni, Immunology of hepatitis B virus and hepatitis distribution of hepatitis C virus RNA in infected livers, PLoS One 4 (2009)
C virus infection, Nat. Rev. Immunol. 5 (2005) 215–229. e6661.
[30] D.R. McGivern, S.M. Lemon, Virus-specific mechanisms of carcinogenesis in [55] E. Rodriguez-Inigo, L. Mariscal, J. Bartolome, I. Castillo, C. Navacerrada, N.
hepatitis C virus associated liver cancer, Oncogene 30 (2011) 1969–1983. Ortiz-Movilla, M. Pardo, V. Carreno, Distribution of hepatitis B virus in the
[31] N.M. Valiante, A. D’Andrea, S. Crotta, F. Lechner, P. Klenerman, S. Nuti, A. liver of chronic hepatitis C patients with occult hepatitis B virus infection, J.
Wack, S. Abrignani, Life, activation and death of intrahepatic lymphocytes in Med. Virol. 70 (2003) 571–580.
chronic hepatitis C, Immunol. Rev. 174 (2000) 77–89. [56] F. Chen, JNK-induced apoptosis, compensatory growth, and cancer stem cells,
[32] L.G. Guidotti, T. Ishikawa, M.V. Hobbs, B. Matzke, R. Schreiber, F.V. Chisari, Cancer Res. 72 (2012) 379–386.
Intracellular inactivation of the hepatitis B virus by cytotoxic T lymphocytes, [57] I. Gentile, M.A. Carleo, F. Borgia, G. Castaldo, G. Borgia, The efficacy and safety
Immunity 4 (1996) 25–36. of telaprevir – a new protease inhibitor against hepatitis C virus, Expert Opin.
[33] S. Nuti, D. Rosa, N.M. Valiante, G. Saletti, M. Caratozzolo, P. Dellabona, V. Investig Drugs 19 (2010) 151–159.
Barnaba, S. Abrignani, Dynamics of intra-hepatic lymphocytes in chronic [58] J.J. Ott, G.A. Stevens, J. Groeger, S.T. Wiersma, Global epidemiology of
hepatitis C: enrichment for Valpha24+ T cells and rapid elimination of hepatitis B virus infection: new estimates of age-specific HBsAg
effector cells by apoptosis, Eur. J. Immunol. 28 (1998) 3448–3455. seroprevalence and endemicity, Vaccine 30 (2012) 2212–2219.
[34] K. Falasca, C. Ucciferri, M. Dalessandro, P. Zingariello, P. Mancino, C. Petrarca, [59] W.P. Hofmann, S. Zeuzem, A new standard of care for the treatment of chronic
E. Pizzigallo, P. Conti, J. Vecchiet, Cytokine patterns correlate with liver HCV infection, Nat. Rev. Gastroenterol. Hepatol 8 (2011) 257–264.
damage in patients with chronic hepatitis B and C, Ann. Clin. Lab. Sci. 36 [60] R.C. Gallo, Research and discovery of the first human cancer virus, HTLV-1,
(2006) 144–150. Best Pract. Res. Clin. Haematol. 24 (2011) 559–565.
[35] R. Shukla, J. Yue, M. Siouda, T. Gheit, O. Hantz, P. Merle, F. Zoulim, V. [61] B.J. Poiesz, F.W. Ruscetti, A.F. Gazdar, P.A. Bunn, J.D. Minna, R.C. Gallo,
Krutovskikh, M. Tommasino, B.S. Sylla, Proinflammatory cytokine TNF-alpha Detection and isolation of type C retrovirus particles from fresh and cultured
increases the stability of hepatitis B virus X protein through NF-kappaB lymphocytes of a patient with cutaneous T-cell lymphoma, Proc. Natl. Acad.
signaling, Carcinogenesis 32 (2011) 978–985. Sci. USA 77 (1980) 7415–7419.
[36] M.I. Radwan, H.F. Pasha, R.H. Mohamed, H.I. Hussien, M.N. El-Khshab, [62] Y. Hinuma, K. Nagata, M. Hanaoka, M. Nakai, T. Matsumoto, K.I. Kinoshita, S.
Influence of transforming growth factor-beta1 and tumor necrosis factor- Shirakawa, I. Miyoshi, Adult T-cell leukemia: antigen in an ATL cell line and
alpha genes polymorphisms on the development of cirrhosis and detection of antibodies to the antigen in human sera, Proc. Natl. Acad. Sci.
hepatocellular carcinoma in chronic hepatitis C patients, Cytokine 60 USA 78 (1981) 6476–6480.
(2012) 271–276. [63] L.B. Cook, M. Elemans, A.G. Rowan, B. Asquith, HTLV-1: persistence and
[37] K. Machida, H. Tsukamoto, J.C. Liu, Y.P. Han, S. Govindarajan, M.M. Lai, S. pathogenesis, Virology 435 (2013) 131–140.
Akira, J.H. Ou, C-Jun mediates hepatitis C virus hepatocarcinogenesis through [64] M. Boxus, L. Willems, Mechanisms of HTLV-1 persistence and transformation,
signal transducer and activator of transcription 3 and nitric oxide-dependent Br. J. Cancer 101 (2009) 1497–1501.
impairment of oxidative DNA repair, Hepatology 52 (2010) 480–492. [65] T. Akagi, H. Ono, K. Shimotohno, Characterization of T cells immortalized by
[38] Y. Wang, N. Kato, Y. Hoshida, H. Yoshida, H. Taniguchi, T. Goto, M. Moriyama, Tax1 of human T-cell leukemia virus type 1, Blood 86 (1995) 4243–4249.
M. Otsuka, S. Shiina, Y. Shiratori, Y. Ito, M. Omata, Interleukin-1beta gene [66] Y. Furukawa, R. Kubota, M. Tara, S. Izumo, M. Osame, Existence of escape
polymorphisms associated with hepatocellular carcinoma in hepatitis C virus mutant in HTLV-I tax during the development of adult T-cell leukemia, Blood
infection, Hepatology 37 (2003) 65–71. 97 (2001) 987–993.
[39] R. Saxena, Y.K. Chawla, I. Verma, J. Kaur, Interleukin-1 polymorphism and [67] Y. Satou, J. Yasunaga, M. Yoshida, M. Matsuoka, HTLV-I basic leucine zipper
expression in hepatitis B virus-mediated disease outcome in India, J. factor gene mRNA supports proliferation of adult T cell leukemia cells, Proc.
Interferon Cytokine Res. 33 (2013) 80–89. Natl. Acad. Sci. USA 103 (2006) 720–725.
[40] V. Hernandez-Gea, S.L. Friedman, Pathogenesis of liver fibrosis, Ann. Rev. [68] M. Watanabe, T. Ohsugi, M. Shoda, T. Ishida, S. Aizawa, M. Maruyama-Nagai, A.
Pathol. 6 (2011) 425–456. Utsunomiya, S. Koga, Y. Yamada, S. Kamihira, A. Okayama, H. Kikuchi, K. Uozumi,
[41] K. Matsuzaki, Modulation of TGF-beta signaling during progression of chronic K. Yamaguchi, M. Higashihara, K. Umezawa, T. Watanabe, R. Horie, Dual
liver diseases, Front Biosci. 14 (2009) 2923–2934. targeting of transformed and untransformed HTLV-1-infected T cells by DHMEQ,
[42] B.A. Teicher, Malignant cells, directors of the malignant process: role of a potent and selective inhibitor of NF-kappaB, as a strategy for chemoprevention
transforming growth factor-beta, Cancer Metastasis Rev. 20 (2001) 133–143. and therapy of adult T-cell leukemia, Blood 106 (2005) 2462–2471.
[43] K. Okumoto, E. Hattori, K. Tamura, S. Kiso, H. Watanabe, K. Saito, T. Saito, H. [69] S.C. Sun, S. Yamaoka, Activation of NF-kappaB by HTLV-I and implications for
Togashi, S. Kawata, Possible contribution of circulating transforming growth cell transformation, Oncogene 24 (2005) 5952–5964.
180 S.A. Read, M.W. Douglas / Cancer Letters 345 (2014) 174–181

[70] K. Leung, G.J. Nabel, HTLV-1 transactivator induces interleukin-2 receptor lymphoma-derived tissue serially transplanted into severe combined
expression through an NF-kappa B-like factor, Nature 333 (1988) 776–778. immunodeficient mice, Blood 82 (1993) 1247–1256.
[71] D.W. Ballard, E. Bohnlein, J.W. Lowenthal, Y. Wano, B.R. Franza, W.C. Greene, [100] N. Tedla, P. Palladinetti, D. Wakefield, A. Lloyd, Abundant expression of
HTLV-I tax induces cellular proteins that activate the kappa B element in the chemokines in malignant and infective human lymphadenopathies, Cytokine
IL-2 receptor alpha gene, Science 241 (1988) 1652–1655. 11 (1999) 531–540.
[72] Z. Qu, G. Xiao, Human T-cell lymphotropic virus: a model of NF-kappaB- [101] H. Hanamoto, T. Nakayama, H. Miyazato, S. Takegawa, K. Hieshima, Y. Tatsumi,
associated tumorigenesis, Viruses 3 (2011) 714–749. A. Kanamaru, O. Yoshie, Expression of CCL28 by Reed–Sternberg cells defines a
[73] T. Zhao, J. Yasunaga, Y. Satou, M. Nakao, M. Takahashi, M. Fujii, M. Matsuoka, major subtype of classical Hodgkin’s disease with frequent infiltration of
Human T-cell leukemia virus type 1 bZIP factor selectively suppresses the eosinophils and/or plasma cells, Am J Pathol 164 (2004) 997–1006.
classical pathway of NF-kappaB, Blood 113 (2009) 2755–2764. [102] M. Samoszuk, L. Nansen, Detection of interleukin-5 messenger RNA in Reed–
[74] Y. Yamada, Y. Ohmoto, T. Hata, M. Yamamura, K. Murata, K. Tsukasaki, T. Sternberg cells of Hodgkin’s disease with eosinophilia, Blood 75 (1990) 13–16.
Kohno, Y. Chen, S. Kamihira, M. Tomonaga, Features of the cytokines secreted [103] H.D. Foss, H. Herbst, E. Oelmann, J. Samol, M. Grebe, T. Blankenstein, J.
by adult T cell leukemia (ATL) cells, Leuk Lymphoma 21 (1996) 443–447. Matthes, Z.H. Qin, B. Falini, S. Pileri, et al., Lymphotoxin, tumour necrosis
[75] E.W. Harhaj, N.S. Harhaj, C. Grant, K. Mostoller, T. Alefantis, S.C. Sun, B. factor and interleukin-6 gene transcripts are present in Hodgkin and Reed–
Wigdahl, Human T cell leukemia virus type I Tax activates CD40 gene Sternberg cells of most Hodgkin’s disease cases, Br. J. Haematol. 84 (1993)
expression via the NF-kappa B pathway, Virology 333 (2005) 145–158. 627–635.
[76] M. Higuchi, T. Matsuda, N. Mori, Y. Yamada, R. Horie, T. Watanabe, M. [104] H.D. Foss, M. Hummel, S. Gottstein, K. Ziemann, B. Falini, H. Herbst, H. Stein,
Takahashi, M. Oie, M. Fujii, Elevated expression of CD30 in adult T-cell Frequent expression of IL-7 gene transcripts in tumor cells of classical
leukemia cell lines: possible role in constitutive NF-kappaB activation, Hodgkin’s disease, Am. J. Pathol. 146 (1995) 33–39.
Retrovirology 2 (2005) 29. [105] H. Herbst, H.D. Foss, J. Samol, I. Araujo, H. Klotzbach, H. Krause, A.
[77] E.P. Cowan, R.K. Alexander, S. Daniel, F. Kashanchi, J.N. Brady, Induction of Agathanggelou, G. Niedobitek, H. Stein, Frequent expression of interleukin-
tumor necrosis factor alpha in human neuronal cells by extracellular human 10 by Epstein–Barr virus-harboring tumor cells of Hodgkin’s disease, Blood
T-cell lymphotropic virus type 1 Tax, J. Virol. 71 (1997) 6982–6989. 87 (1996) 2918–2929.
[78] H. Albrecht, A.N. Shakhov, C.V. Jongeneel, Trans activation of the tumor [106] G.E. Brito-Melo, V. Peruhype-Magalhaes, A. Teixeira-Carvalho, E.F. Barbosa-
necrosis factor alpha promoter by the human T-cell leukemia virus type I Stancioli, A.B. Carneiro-Proietti, B. Catalan-Soares, J.G. Ribas, O.A. Martins-
Tax1 protein, J. Virol. 66 (1992) 6191–6193. Filho, IL-10 produced by CD4+ and CD8+ T cells emerge as a putative
[79] K. Tsukasaki, C.W. Miller, T. Kubota, S. Takeuchi, T. Fujimoto, S. Ikeda, M. immunoregulatory mechanism to counterbalance the monocyte-derived
Tomonaga, H.P. Koeffler, Tumor necrosis factor alpha polymorphism TNF-alpha and guarantee asymptomatic clinical status during chronic
associated with increased susceptibility to development of adult T-cell HTLV-I infection, Clin. Exp. Immunol. 147 (2007) 35–44.
leukemia/lymphoma in human T-lymphotropic virus type 1 carriers, Cancer [107] S.I. Yoon, B.C. Jones, N.J. Logsdon, B.D. Harris, S. Kuruganti, M.R. Walter,
Res. 61 (2001) 3770–3774. Epstein–Barr virus IL-10 engages IL-10R1 by a two-step mechanism leading
[80] J.B. Guerreiro, S.B. Santos, D.J. Morgan, A.F. Porto, A.L. Muniz, J.L. Ho, A.L. to altered signaling properties, J. Biol. Chem. 287 (2012) 26586–26595.
Teixeira Jr., M.M. Teixeira, E.M. Carvalho, Levels of serum chemokines [108] M.T. Bejarano, M.G. Masucci, Interleukin-10 abrogates the inhibition of
discriminate clinical myelopathy associated with human T lymphotropic Epstein–Barr virus-induced B-cell transformation by memory T-cell
virus type 1 (HTLV-1)/tropical spastic paraparesis (HAM/TSP) disease from responses, Blood 92 (1998) 4256–4262.
HTLV-1 carrier state, Clin. Exp. Immunol. 145 (2006) 296–301. [109] R.F. Jarrett, Viruses and Hodgkin’s lymphoma, Ann. Oncol. 13 (Suppl. 1)
[81] S. Tattermusch, C.R. Bangham, HTLV-1 infection: what determines the risk of (2002) 23–29.
inflammatory disease?, Trends Microbiol 20 (2012) 494–500. [110] E. Cahir-McFarland, E. Kieff, NF-kappaB inhibition in EBV-transformed
[82] T. Zhao, M. Matsuoka, HBZ and its roles in HTLV-1 oncogenesis, Front lymphoblastoid cell lines, Recent Results Cancer Res. 159 (2002) 44–48.
Microbiol. 3 (2012) 247. [111] K.H. Shair, K.M. Bendt, R.H. Edwards, E.C. Bedford, J.N. Nielsen, N. Raab-Traub,
[83] B.G. Bajaj, M. Murakami, E.S. Robertson, Molecular biology of EBV in relationship EBV latent membrane protein 1 activates Akt, NFkappaB, and Stat3 in B cell
to AIDS-associated oncogenesis, Cancer Treat Res. 133 (2007) 141–162. lymphomas, PLoS Pathog. 3 (2007) e166.
[84] T. Bravender, Epstein–Barr virus, cytomegalovirus, and infectious [112] I. Penn, Kaposi’s sarcoma in organ transplant recipients: report of 20 cases,
mononucleosis, Adolesc. Med. State Art Rev. 21 (2010) 251–264 (ix). Transplantation 27 (1979) 8–11.
[85] C.M. Borza, L.M. Hutt-Fletcher, Alternate replication in B cells and epithelial [113] A.E. Friedman-Kien, Disseminated Kaposi’s sarcoma syndrome in young
cells switches tropism of Epstein–Barr virus, Nat. Med. 8 (2002) 594– homosexual men, J. Am. Acad. Dermatol. 5 (1981) 468–471.
599. [114] Y. Chang, E. Cesarman, M.S. Pessin, F. Lee, J. Culpepper, D.M. Knowles, P.S.
[86] M.A. Epstein, B.G. Achong, Y.M. Barr, Virus particles in cultured lymphoblasts Moore, Identification of herpesvirus-like DNA sequences in AIDS-associated
from Burkitt’s lymphoma, Lancet 1 (1964) 702–703. Kaposi’s sarcoma, Science 266 (1994) 1865–1869.
[87] L.S. Young, A.B. Rickinson, Epstein–Barr virus: 40 years on, Nat. Rev. Cancer 4 [115] A. Hansen, C. Boshoff, D. Lagos, Kaposi sarcoma as a model of oncogenesis and
(2004) 757–768. cancer treatment, Expert Rev. Anticancer Ther. 7 (2007) 211–220.
[88] N. Raab-Traub, Novel mechanisms of EBV-induced oncogenesis, Curr. Opin. [116] J.N. Martin, Kaposi sarcoma-associated herpesvirus/human herpesvirus 8 and
Virol. 2 (2012) 453–458. Kaposi sarcoma, Adv. Dent. Res. 23 (2011) 76–78.
[89] G. Khan, Epstein–Barr virus, cytokines, and inflammation: a cocktail for the [117] S.C. Verma, E.S. Robertson, Molecular biology and pathogenesis of Kaposi
pathogenesis of Hodgkin’s lymphoma?, Exp Hematol. 34 (2006) 399–406. sarcoma-associated herpesvirus, FEMS Microbiol. Lett. 222 (2003) 155–163.
[90] K.M. Kaye, K.M. Izumi, E. Kieff, Epstein–Barr virus latent membrane protein 1 [118] R. Holzerlandt, C. Orengo, P. Kellam, M.M. Alba, Identification of new
is essential for B-lymphocyte growth transformation, Proc. Natl. Acad. Sci. herpesvirus gene homologs in the human genome, Genome Res. 12 (2002)
USA 90 (1993) 9150–9154. 1739–1748.
[91] D. Wang, D. Liebowitz, E. Kieff, An EBV membrane protein expressed in [119] I. Guasparri, H. Wu, E. Cesarman, The KSHV oncoprotein vFLIP contains a
immortalized lymphocytes transforms established rodent cells, Cell 43 TRAF-interacting motif and requires TRAF2 and TRAF3 for signalling, EMBO
(1985) 831–840. Rep. 7 (2006) 114–119.
[92] J.I. Cohen, F. Wang, J. Mannick, E. Kieff, Epstein–Barr virus nuclear protein 2 is [120] P. Chugh, H. Matta, S. Schamus, S. Zachariah, A. Kumar, J.A. Richardson, A.L.
a key determinant of lymphocyte transformation, Proc. Natl. Acad. Sci. USA 86 Smith, P.M. Chaudhary, Constitutive NF-kappaB activation, normal Fas-induced
(1989) 9558–9562. apoptosis, and increased incidence of lymphoma in human herpes virus 8 K13
[93] W. Hammerschmidt, B. Sugden, Genetic analysis of immortalizing functions transgenic mice, Proc. Natl. Acad. Sci. USA 102 (2005) 12885–12890.
of Epstein–Barr virus in human B lymphocytes, Nature 340 (1989) 393–397. [121] P.M. Chaudhary, A. Jasmin, M.T. Eby, L. Hood, Modulation of the NF-kappa B
[94] U. Zimber-Strobl, L.J. Strobl, EBNA2 and Notch signalling in Epstein–Barr virus pathway by virally encoded death effector domains-containing proteins,
mediated immortalization of B lymphocytes, Semin. Cancer Biol. 11 (2001) Oncogene 18 (1999) 5738–5746.
423–434. [122] I. Guasparri, S.A. Keller, E. Cesarman, KSHV vFLIP is essential for the survival
[95] A.G. Eliopoulos, M. Stack, C.W. Dawson, K.M. Kaye, L. Hodgkin, S. Sihota, M. of infected lymphoma cells, J Exp. Med. 199 (2004) 993–1003.
Rowe, L.S. Young, Epstein–Barr virus-encoded LMP1 and CD40 mediate IL-6 [123] V. Punj, H. Matta, S. Schamus, T. Yang, Y. Chang, P.M. Chaudhary, Induction of
production in epithelial cells via an NF-kappaB pathway involving TNF CCL20 production by Kaposi sarcoma-associated herpesvirus: role of viral
receptor-associated factors, Oncogene 14 (1997) 2899–2916. FLICE inhibitory protein K13-induced NF-kappaB activation, Blood 113
[96] G. Mosialos, M. Birkenbach, R. Yalamanchili, T. VanArsdale, C. Ware, E. Kieff, (2009) 5660–5668.
The Epstein–Barr virus transforming protein LMP1 engages signaling proteins [124] B. Ensoli, M. Sturzl, Kaposi’s sarcoma: a result of the interplay among
for the tumor necrosis factor receptor family, Cell 80 (1995) 389–399. inflammatory cytokines, angiogenic factors and viral agents, Cytokine
[97] M. Hinz, P. Lemke, I. Anagnostopoulos, C. Hacker, D. Krappmann, S. Mathas, B. Growth Factor Rev. 9 (1998) 63–83.
Dorken, M. Zenke, H. Stein, C. Scheidereit, Nuclear factor kappaB-dependent [125] S. Sakakibara, G. Tosato, Viral interleukin-6: role in Kaposi’s sarcoma-
gene expression profiling of Hodgkin’s disease tumor cells, pathogenetic associated herpesvirus: associated malignancies, J. Interferon Cytokine Res.
significance, and link to constitutive signal transducer and activator of 31 (2011) 791–801.
transcription 5a activity, J. Exp. Med. 196 (2002) 605–617. [126] Y. Aoki, E.S. Jaffe, Y. Chang, K. Jones, J. Teruya-Feldstein, P.S. Moore, G. Tosato,
[98] E. Maggio, A. van den Berg, A. Diepstra, J. Kluiver, L. Visser, S. Poppema, Angiogenesis and hematopoiesis induced by Kaposi’s sarcoma-associated
Chemokines, cytokines and their receptors in Hodgkin’s lymphoma cell lines herpesvirus-encoded interleukin-6, Blood 93 (1999) 4034–4043.
and tissues, Ann. Oncol. 13 (Suppl. 1) (2002) 52–56. [127] J. An, Y. Sun, R. Sun, M.B. Rettig, Kaposi’s sarcoma-associated herpesvirus
[99] U. Kapp, J. Wolf, M. Hummel, M. Pawlita, C. von Kalle, F. Dallenbach, M. encoded vFLIP induces cellular IL-6 expression: the role of the NF-kappaB and
Schwonzen, G.R. Krueger, N. Muller-Lantzsch, C. Fonatsch, et al., Hodgkin’s JNK/AP1 pathways, Oncogene 22 (2003) 3371–3385.
S.A. Read, M.W. Douglas / Cancer Letters 345 (2014) 174–181 181

[128] S. Montaner, A. Sodhi, J.M. Servitja, A.K. Ramsdell, A. Barac, E.T. Sawai, J.S. modulate proliferation and apoptotic pathways in human keratinocytes
Gutkind, The small GTPase Rac1 links the Kaposi sarcoma-associated expressing the human papillomavirus-16 E7 oncoprotein, J. Biol. Chem. 276
herpesvirus vGPCR to cytokine secretion and paracrine neoplasia, Blood (2001) 22522–22528.
104 (2004) 2903–2911. [155] P.J. Duerksen-Hughes, J. Yang, S.B. Schwartz, HPV 16 E6 blocks TNF-mediated
[129] D.A. Davis, A.S. Rinderknecht, J.P. Zoeteweij, Y. Aoki, E.L. Read-Connole, G. apoptosis in mouse fibroblast LM cells, Virology 264 (1999) 55–65.
Tosato, A. Blauvelt, R. Yarchoan, Hypoxia induces lytic replication of Kaposi [156] L.L. Villa, K.B. Vieira, X.F. Pei, R. Schlegel, Differential effect of tumor necrosis
sarcoma-associated herpesvirus, Blood 97 (2001) 3244–3250. factor on proliferation of primary human keratinocytes and cell lines containing
[130] Y. Aoki, R. Yarchoan, J. Braun, A. Iwamoto, G. Tosato, Viral and cellular human papillomavirus types 16 and 18, Mol. Carcinog. 6 (1992) 5–9.
cytokines in AIDS-related malignant lymphomatous effusions, Blood 96 [157] K.B. Vieira, D.J. Goldstein, L.L. Villa, Tumor necrosis factor alpha interferes
(2000) 1599–1601. with the cell cycle of normal and papillomavirus-immortalized human
[131] K.D. Jones, Y. Aoki, Y. Chang, P.S. Moore, R. Yarchoan, G. Tosato, Involvement keratinocytes, Cancer Res. 56 (1996) 2452–2457.
of interleukin-10 (IL-10) and viral IL-6 in the spontaneous growth of Kaposi’s [158] C.D. Woodworth, E. McMullin, M. Iglesias, G.D. Plowman, Interleukin 1 alpha
sarcoma herpesvirus-associated infected primary effusion lymphoma cells, and tumor necrosis factor alpha stimulate autocrine amphiregulin expression
Blood 94 (1999) 2871–2879. and proliferation of human papillomavirus-immortalized and carcinoma-
[132] E. Oksenhendler, M. Duarte, J. Soulier, P. Cacoub, Y. Welker, J. Cadranel, D. derived cervical epithelial cells, Proc. Natl. Acad. Sci. USA 92 (1995) 2840–
Cazals-Hatem, B. Autran, J.P. Clauvel, M. Raphael, Multicentric Castleman’s 2844.
disease in HIV infection: a clinical and pathological study of 20 patients, AIDS [159] D. Gaiotti, J. Chung, M. Iglesias, M. Nees, P.D. Baker, C.H. Evans, C.D.
10 (1996) 61–67. Woodworth, Tumor necrosis factor-alpha promotes human papillomavirus
[133] A.S. Semeere, N. Busakhala, J.N. Martin, Impact of antiretroviral therapy on (HPV) E6/E7 RNA expression and cyclin-dependent kinase activity in HPV-
the incidence of Kaposi’s sarcoma in resource-rich and resource-limited immortalized keratinocytes by a ras-dependent pathway, Mol. Carcinog. 27
settings, Curr. Opin. Oncol. 24 (2012) 522–530. (2000) 97–109.
[134] S.M. Hosseini-Moghaddam, A. Soleimanirahbar, T. Mazzulli, C. Rotstein, S. [160] S. Hess, H. Smola, U. Sandaradura De Silva, D. Hadaschik, D. Kube, S.E. Baldus,
Husain, Post renal transplantation Kaposi’s sarcoma: a review of its U. Flucke, H. Pfister, Loss of IL-6 receptor expression in cervical carcinoma
epidemiology, pathogenesis, diagnosis, clinical aspects, and therapy, cells inhibits autocrine IL-6 stimulation: abrogation of constitutive monocyte
Transpl. Infect. Dis. 14 (2012) 338–345. chemoattractant protein-1 production, J. Immunol. 165 (2000) 1939–1948.
[135] H.U. Bernard, R.D. Burk, Z. Chen, K. van Doorslaer, H. Hausen, E.M. de Villiers, [161] K. Kleine-Lowinski, J.G. Rheinwald, R.N. Fichorova, D.J. Anderson, J. Basile, K.
Classification of papillomaviruses (PVs) based on 189 PV types and proposal Munger, C.M. Daly, F. Rosl, B.J. Rollins, Selective suppression of monocyte
of taxonomic amendments, Virology 401 (2010) 70–79. chemoattractant protein-1 expression by human papillomavirus E6 and E7
[136] D.M. Parkin, F. Bray, Chapter 2: The burden of HPV-related cancers, Vaccine oncoproteins in human cervical epithelial and epidermal cells, Int. J. Cancer
24 (Suppl. 3) (2006). S3/11-25. 107 (2003) 407–415.
[137] K. Munger, B.A. Werness, N. Dyson, W.C. Phelps, E. Harlow, P.M. Howley, [162] E. Boccardo, A.P. Lepique, L.L. Villa, The role of inflammation in HPV
Complex formation of human papillomavirus E7 proteins with the carcinogenesis, Carcinogenesis 31 (2010) 1905–1912.
retinoblastoma tumor suppressor gene product, EMBO J. 8 (1989) 4099– [163] P.E. Castle, S.L. Hillier, L.K. Rabe, A. Hildesheim, R. Herrero, M.C. Bratti, M.E.
4105. Sherman, R.D. Burk, A.C. Rodriguez, M. Alfaro, M.L. Hutchinson, J. Morales, M.
[138] D.J. McCance, R. Kopan, E. Fuchs, L.A. Laimins, Human papillomavirus type 16 Schiffman, An association of cervical inflammation with high-grade cervical
alters human epithelial cell differentiation in vitro, Proc. Natl. Acad. Sci. USA neoplasia in women infected with oncogenic human papillomavirus (HPV),
85 (1988) 7169–7173. Cancer Epidemiol. Biomarkers Prev. 10 (2001) 1021–1027.
[139] M. Scheffner, B.A. Werness, J.M. Huibregtse, A.J. Levine, P.M. Howley, The E6 [164] G. Castrilli, D. Tatone, M.G. Diodoro, S. Rosini, M. Piantelli, P. Musiani, Interleukin
oncoprotein encoded by human papillomavirus types 16 and 18 promotes 1alpha and interleukin 6 promote the in vitro growth of both normal and
the degradation of p53, Cell 63 (1990) 1129–1136. neoplastic human cervical epithelial cells, Br. J. Cancer 75 (1997) 855–859.
[140] C.L. Halbert, G.W. Demers, D.A. Galloway, The E7 gene of human [165] E. Tartour, A. Gey, X. Sastre-Garau, C. Pannetier, V. Mosseri, P. Kourilsky, W.H.
papillomavirus type 16 is sufficient for immortalization of human epithelial Fridman, Analysis of interleukin 6 gene expression in cervical neoplasia using
cells, J. Virol. 65 (1991) 473–478. a quantitative polymerase chain reaction assay: evidence for enhanced
[141] K. Munger, W.C. Phelps, V. Bubb, P.M. Howley, R. Schlegel, The E6 and E7 interleukin 6 gene expression in invasive carcinoma, Cancer Res. 54 (1994)
genes of the human papillomavirus type 16 together are necessary and 6243–6248.
sufficient for transformation of primary human keratinocytes, J. Virol. 63 [166] V.M. Williams, M. Filippova, U. Soto, P.J. Duerksen-Hughes, HPV-DNA
(1989) 4417–4421. integration and carcinogenesis: putative roles for inflammation and
[142] J.M. Arbeit, P.M. Howley, D. Hanahan, Chronic estrogen-induced cervical and oxidative stress, Future Virol. 6 (2011) 45–57.
vaginal squamous carcinogenesis in human papillomavirus type 16 [167] M.E. Spurgeon, P.F. Lambert, Merkel cell polyomavirus: a newly discovered
transgenic mice, Proc. Natl. Acad. Sci. USA 93 (1996) 2930–2935. human virus with oncogenic potential, Virology 435 (2013) 118–130.
[143] A. Saha, R. Kaul, M. Murakami, E.S. Robertson, Tumor viruses and cancer [168] R. Arora, Y. Chang, P.S. Moore, MCV and Merkel cell carcinoma: a molecular
biology: modulating signaling pathways for therapeutic intervention, Cancer success story, Curr. Opin. Virol. 2 (2012) 489–498.
Biol. Ther. 10 (2010) 961–978. [169] M.E. McLaughlin-Drubin, K. Munger, Viruses associated with human cancer,
[144] J.M. Oh, S.H. Kim, E.A. Cho, Y.S. Song, W.H. Kim, Y.S. Juhnn, Human Biochim. Biophys. Acta 1782 (2008) 127–150.
papillomavirus type 16 E5 protein inhibits hydrogen-peroxide-induced [170] A.C. Tricco, C.H. Ng, V. Gilca, A. Anonychuk, B. Pham, S. Berliner, Canadian
apoptosis by stimulating ubiquitin-proteasome-mediated degradation of oncogenic human papillomavirus cervical infection prevalence: systematic
Bax in human cervical cancer cells, Carcinogenesis 31 (2010) 402–410. review and meta-analysis, BMC Infect. Dis. 11 (2011) 235.
[145] J.M. Oh, S.H. Kim, Y.I. Lee, M. Seo, S.Y. Kim, Y.S. Song, W.H. Kim, Y.S. Juhnn, [171] C.Y. Wu, Y.J. Chen, H.J. Ho, Y.C. Hsu, K.N. Kuo, M.S. Wu, J.T. Lin, Association
Human papillomavirus E5 protein induces expression of the EP4 subtype of between nucleoside analogues and risk of hepatitis B virus-related
prostaglandin E2 receptor in cyclic AMP response element-dependent hepatocellular carcinoma recurrence following liver resection, JAMA 308
pathways in cervical cancer cells, Carcinogenesis 30 (2009) 141–149. (2012) 1906–1914.
[146] K. Subbaramaiah, A.J. Dannenberg, Cyclooxygenase-2 transcription is [172] Y.F. Liaw, J.J. Sung, W.C. Chow, G. Farrell, C.Z. Lee, H. Yuen, T. Tanwandee, Q.M.
regulated by human papillomavirus 16 E6 and E7 oncoproteins: evidence Tao, K. Shue, O.N. Keene, J.S. Dixon, D.F. Gray, J. Sabbat, Lamivudine for
of a corepressor/coactivator exchange, Cancer Res. 67 (2007) 3976–3985. patients with chronic hepatitis B and advanced liver disease, N. Engl. J. Med.
[147] D. Wang, R.N. Dubois, Prostaglandins and cancer, Gut 55 (2006) 115–122. 351 (2004) 1521–1531.
[148] P. Bhat, S.R. Mattarollo, C. Gosmann, I.H. Frazer, G.R. Leggatt, Regulation of [173] C.J. Chen, H.I. Yang, J. Su, C.L. Jen, S.L. You, S.N. Lu, G.T. Huang, U.H. Iloeje, Risk
immune responses to HPV infection and during HPV-directed of hepatocellular carcinoma across a biological gradient of serum hepatitis B
immunotherapy, Immunol. Rev. 239 (2011) 85–98. virus DNA level, JAMA 295 (2006) 65–73.
[149] P. Delvenne, W. al-Saleh, C. Gilles, A. Thiry, J. Boniver, Inhibition of growth of [174] P.M. Rothwell, F.G. Fowkes, J.F. Belch, H. Ogawa, C.P. Warlow, T.W. Meade,
normal and human papillomavirus-transformed keratinocytes in monolayer Effect of daily aspirin on long-term risk of death due to cancer: analysis of
and organotypic cultures by interferon-gamma and tumor necrosis factor- individual patient data from randomised trials, Lancet 377 (2011) 31–41.
alpha, Am. J. Pathol. 146 (1995) 589–598. [175] X. Wang, P.J. Crowe, D. Goldstein, J.L. Yang, STAT3 inhibition, a novel
[150] S. Kyo, M. Inoue, N. Hayasaka, T. Inoue, M. Yutsudo, O. Tanizawa, A. Hakura, approach to enhancing targeted therapy in human cancers (review), Int. J.
Regulation of early gene expression of human papillomavirus type 16 by Oncol. (2012).
inflammatory cytokines, Virology 200 (1994) 130–139. [176] A.M. Algra, P.M. Rothwell, Effects of regular aspirin on long-term cancer
[151] K.Y. Kim, L. Blatt, M.W. Taylor, The effects of interferon on the expression of incidence and metastasis: a systematic comparison of evidence from
human papillomavirus oncogenes, J. Gen. Virol. 81 (2000) 695–700. observational studies versus randomised trials, Lancet Oncol. 13 (2012)
[152] R.W. Tindle, Immune evasion in human papillomavirus-associated cervical 518–527.
cancer, Nat. Rev. Cancer 2 (2002) 59–65. [177] M. Jang, L. Cai, G.O. Udeani, K.V. Slowing, C.F. Thomas, C.W. Beecher, H.H.
[153] E.R. Vandermark, K.A. Deluca, C.R. Gardner, D.F. Marker, C.N. Schreiner, D.A. Fong, N.R. Farnsworth, A.D. Kinghorn, R.G. Mehta, R.C. Moon, J.M. Pezzuto,
Strickland, K.M. Wilton, S. Mondal, C.D. Woodworth, Human papillomavirus Cancer chemopreventive activity of resveratrol, a natural product derived
type 16 E6 and E 7 proteins alter NF-kB in cultured cervical epithelial cells from grapes, Science 275 (1997) 218–220.
and inhibition of NF-kB promotes cell growth and immortalization, Virology [178] S. Bengmark, Curcumin, an atoxic antioxidant and natural NFkappaB,
425 (2012) 53–60. cyclooxygenase-2, lipooxygenase, and inducible nitric oxide synthase
[154] J.R. Basile, V. Zacny, K. Munger, The cytokines tumor necrosis factor-alpha inhibitor: a shield against acute and chronic diseases, JPEN J. Parenter
(TNF-alpha) and TNF-related apoptosis-inducing ligand differentially Enteral Nutr. 30 (2006) 45–51.

You might also like