You are on page 1of 11

Seminars in Immunology 32 (2017) 14–24

Contents lists available at ScienceDirect

Seminars in Immunology
journal homepage: www.elsevier.com/locate/ysmim

Review

Interplay between viruses and bacterial microbiota in cancer development MARK


a,1 a,1 b,⁎ a,⁎
Dariia Vyshenska , Khiem C. Lam , Natalia Shulzhenko , Andrey Morgun
a
College of Pharmacy, Oregon State University, 1601 SW Jefferson Way, Corvallis, OR 97331, USA
b
College of Veterinary Medicine, Oregon State University, 208 Dryden Hall, Corvallis, OR 97331, USA

A R T I C L E I N F O A B S T R A C T

Keywords: During the last few decades we have become accustomed to the idea that viruses can cause tumors. It is much
Microbiota less considered and discussed, however, that most people infected with oncoviruses will never develop cancer.
Oncovirus Therefore, the genetic and environmental factors that tip the scales from clearance of viral infection to devel-
Microbiome opment of cancer are currently an area of active investigation. Microbiota has recently emerged as a potentially
Virus-associated cancer
critical factor that would affect this balance by increasing or decreasing the ability of viral infection to promote
Transkingdom interactions
carcinogenesis. In this review, we provide a model of microbiome contribution to the development of oncogenic
viral infections and viral associated cancers, give examples of this process in human tumors, and describe the
challenges that prevent progress in the field as well as their potential solutions.

1. Introduction Studies thus far have placed emphasis on gastrointestinal micro-


biota and its role in the development and progression of gut-associated
As the human lifespan lengthens, the incidence of cancer worldwide malignancies [14]. For example, Helicobacter pylori causes gastric ade-
is also increasing. The World Health Organization predicts the fre- nocarcinoma and is a classic case of oncogenic bacteria [15]. Intestinal
quency of cancer occurrence to increase by 70% over the next two infections with other bacteria such as Salmonella typhi [16] and Strep-
decades [1,2], indicating a rise in the global cancer epidemic. One of tococcus gallolyticus (bovis) [17] were also linked to the development of
the established causes of cancer is viral infection, which is responsible hepatobiliary and colorectal cancers, respectively. These studies re-
for 20% of the global cancer burden [3]. Among these infections, the present additional support for the hypothesis that some microbiota
most common are Human Papilloma Virus (HPV) and Hepatitis C/B members as an understudied environmental factor contributing to
viruses with other less frequent contributors being Epstein - Barr virus protection from or the development of virus-associated cancers. Even
(EBV), Human Immunodeficiency Virus (HIV), and Kaposi Sarcoma though the gut microbiome represents the majority of bacteria in the
Herpesviruses (KSHV) [4]. These viruses use two different strategies to human microbiome [18], other body sites such as the vagina and oral
cause cancer: first, by directly affecting host cell machinery (e.g. HPV); cavity have been explored for their participation in cancer development
and second, indirectly, by inhibiting the human immune system (e.g. and progression.
HIV) [5,6]. It is common knowledge that the development of some Although oncogenic properties of virus and bacteria, individually,
cancers require viral infection, such as HPV for cervical cancer. It is less are popular topics, the interaction between these in the context of
known, however, why most people infected with oncogenic viruses will cancer has not been well investigated. In recent years, we have wit-
never develop cancer. nessed an increasing number of attempts to interrogate this three-way
A hint in solving this puzzle may come from studies demonstrating interaction, particularly the influence of microbiota on the progression
the crucial role of microbiota (collection of microorganisms living with and acquisition of oncogenic viral infections. However, the question
the host) in the course of viral infections [7–10]. Moreover, microbiota whether bacteria are beneficial or harmful in this context remains un-
have been recently implicated in different diseases associated with answered for many cancers. In this review, we provide a model of
aberrant immune responses ranging from diabetes and autoimmunity to microbiome contribution to the development of oncogenic viral infec-
obesity and cancer [11,12]. For example, a recent epidemiologic study tions, discuss examples of this process in human tumors, and describe
reported an association between antibiotic exposure and the develop- obstacles in the field and their potential solutions.
ment of several malignancies such as esophageal, gastric, pancreatic,
lung, prostate, and breast cancers [13].


Corresponding authors.
E-mail addresses: natalia.shulzhenko@oregonstate.edu (N. Shulzhenko), Andriy.morgun@oregonstate.edu (A. Morgun).
1
Equal contributors.

http://dx.doi.org/10.1016/j.smim.2017.05.003
Received 19 February 2017; Received in revised form 3 May 2017; Accepted 30 May 2017
Available online 09 June 2017
1044-5323/ © 2017 Elsevier Ltd. All rights reserved.
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

2. Model description in disease development and progression. Among most prominent ex-
amples are diabetes [28], irritable bowel disease [25], and cancer
The role of bacteria (and bacterial microbiota) in viral infections [29,30]. Additionally, intestinal microbiota has been implicated in
leading to cancer can be assigned to two broad categories: bacteria that modulating the effect of different anti-cancer treatments [31–33]. Heli-
influence viral particles, and bacteria that affect host interaction with cobacter species, in particular Helicobacter pylori, is the most well studied
viral infection. bacterial member of the gut community that causes cancer [15].
On the one hand, healthy microbiota was shown to contribute to
infections by interacting with viruses directly and through bacterial by- 3.1. Hepatitis viruses
products [7–10]. It was reported that commensal microbiota augment
the transmission of mouse mammary tumor virus [19,20], bacterial The second leading cause of cancer mortality is liver cancer [34]. The
lipopolysaccharides enhance virion stability of poliovirus [21], and most prevalent histologic type of primary liver cancer is hepatocellular
enteric bacteria promote norovirus infection through histo-blood group carcinoma (HCC) [35] primarily caused by chronic infection with hepa-
antigen-like substances [9,22]. On the other hand, healthy microbiota titis B (HBV) or hepatitis C (HCV) virus [36]. Although both viruses can
are critical for immune system development, especially on mucosal cause cancer, HCV currently attracts stronger interest from the scientific
surfaces [23]. Antibiotic-treated mice exposed to mucosal influenza community possibly due to the absence of a vaccine against HCV.
virus were observed to have impaired innate and adaptive antiviral The pathogenicity of both HBV and HCV involves a combination of
immune responses and delayed clearance of the virus [24]. These and direct and indirect mechanisms. The HCV encoded core, nonstructural
other studies define microbiota as a putatively important factor for the protein 5A (NS5A) and NS3, and HBV encoded X antigen (HBx) are able
development of virus-associated cancers. to promote host cell proliferation. Both viruses are similarly capable of
Herein we propose a model for the three-way interaction between blocking cell immune response, inhibiting apoptosis while promoting
bacteria, virus, and mammalian host, highlighting two distinct me- angiogenesis and metastasis. By contrast, chronic inflammation caused
chanisms for the contribution of microbiota to virus-associated cancers. by oxidative stress also contributes to the process of carcinogenesis [6].
The first involves the direct effect of interaction of bacteria and bac- While HCV is oncogenic, not all patients suffering from chronic
terial products on viruses, primarily affecting their infectivity (Fig. 1A). hepatitis C will develop cancer. One of the first indications that bacteria
The second involves bacteria-host interactions leading to changes in may be a critical parameter in liver cancer came with the observation
host gene expression and subsequent activation/repression of viral ex- that mice infected with Helicobacter were developing strong in-
pression or direct promotion of inflammation synergizing with the tu- flammatory responses leading to hepatocellular carcinoma [37]. An-
morigenic effects of a virus (Fig. 1B). There is evidence to suggest that other group later found an association between H. pylori specific anti-
the role of bacteria can be positive or negative in terms of disease body levels and HCV associated hepatocellular carcinoma [38].
progression with each of these cases. In this review, we discuss con- Helicobacter DNA was also found in liver and was associated with he-
ventional tumor viruses and explore the role of gut, vaginal, and oral patitis C induced cirrhosis [39], which indicates the ability of H. pylori
microbiota components in both of these mechanisms. to invade the liver and putatively contribute to disease development
(Fig. 2B). However, a more recent study conducted on HCV transgenic
mice colonized with H. pylori found no indication of bacteria translo-
3. Gastrointestinal microbiome cation into the liver and no promotion of tumorigenesis [40]. Whether
this experimental system failed to promote carcinogenesis, or H. pylori
The gut microbiome is the largest microbial community in the human does not contribute to HCV-associated liver cancer, remains unknown.
body. Recent discoveries show its involvement in a variety of functions, Another Helicobacter species, Helicobacter hepaticus, has been shown
including immune system training and metabolism regulation [25–27]. to cause chronic hepatitis and liver cancer in rodents [37]. In the fol-
Separate members of gut microbiota as well as dysbiosis (i.e. non-specific lowing study, Fox et al. have shown that presence of H. hepaticus in the
alterations of mammalian microbial communities) have been implicated gut lumen promotes development of hepatocellular carcinoma in HCV
infected mice, acting synergistically with viral infection [41]. This
process did not require bacterial invasion. Fox et al. also showed that in
an aflatoxin B1 HCC mouse model, H. hepaticus induced the nuclear
factor-κB (NF-κB) along with downstream innate and Th1-type adaptive
immunity. In a HCV transgenic mouse model, they also observed in-
creased gene expression of an NF-kB-dependent inflammatory chemo-
kine (CXCL9) in mice colonized with H. hepaticus [6,41,42] (Fig. 2C, D).
Concordantly in both models, tumor growth was accelerated in the
presence of H. hepaticus. These bacteria, detected in human intestine
[43] and liver [44–46], is suspected to contribute not only to cirrhosis
[47] and HCC [44], but also to a set of other conditions such as in-
flammatory bowel disease [43] and prostate cancer [48]. Consequently,
this data suggests a synergistic relationship between H. hepaticus and
HCV in human cancers.
The link between HBV related oncogenesis and gut microbiota has also
been reported. In 2011, Chen et al. found that enteric fungi diversity was
positively correlated with a worsened disease state in chronic HBV in-
fection [49]. More recently, Chou et al. showed that in adult C3H/HeN
mice transfected with HBV, elimination of gut microbiota with antibiotics
resulted in viral persistence phenotypes, including prolonged HBV surface
Fig. 1. Model of bacteria-virus interactions in cancer development and progression: A)
antigens (HBsAg) presence, impaired anti-HBs production, and limited
direct interaction between bacteria or bacterial by-products and virus resulting in in- HBV core antigen (HBcAg)-specific IFN-γ–secreting splenocytes [50]. Al-
hibition or promotion of viral infection into host cell; B) indirect interaction between though these results suggest that the persistence of HBV infection in an-
bacteria and virus mediated by host response to bacterial stimuli through activation of tibiotic-treated mice is attributed to an ineffective adaptive humoral and
various pattern recognition receptors. cellular immune response, specific bacteria were not identified in this

15
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

Fig. 2. The interplay between gut microbiome, host


and human hepatitis viruses (HCV and HBV): A)
healthy gut microbiota stimulates host immune
system resulting in HBV infection clearance; B)
Helicobacter pylori invades liver and contributes to
chronic inflammation induced by HCV; C, D)
Helicobacter hepaticus upregulates NF-κB dependent
pathways in mouse gut (C) and liver (D), synergizing
with HCV-related inflammation in the development
of hepatocellular carcinoma.

study. Other studies observed that a decrease in fecal Bifidobacterium po- hrHPV infections remain the main predictors for cervical in-
pulations was associated with liver disease progression of HBV infection in traepithelial neoplasia (CIN), precursor to tumor, and cervical cancer
humans [51,52]. In addition, Bifidobacterium species have been shown to development itself [74–76]. HPV infects basal epithelial cells. After
decrease the protein and transcript levels of HBsAg in a HBV-transfected infection HPV either exist in episomal form or can integrate into the cell
human hepatoma cell line [53]. Interestingly, Bifidobacterium can also genome causing genomic instability. Expression of HPV oncogenes E6
increase host gene expression of IFN-signaling components such as STAT1 and E7 is also dependent on integration: in episomal form expression of
[53] and lower serum cholesterol levels [54–56]. Furthermore, IFN sti- E2 protein keeps expression levels of E6/7 low, but during integration
mulation have been shown to downregulate lipid metabolic pathways open reading frame of E2 gene is disrupted and E2 no longer is able to
[27], known in host cells to be necessary for HBV production [57,58]. control HPV oncogenes. E6, E7 and E5 HPV oncoproteins retain kera-
Therefore, we propose that Bifidobacterium species are able to stimulate tinocytes in proliferative state, avoiding apoptosis and clearance by the
IFN-dependent pathways which results in a downregulation of lipid me- immune system. It was also reported that HPV is able to promote an-
tabolism and a reduction in HBV infection (Fig. 2A). Further experi- giogenesis and deregulate cellular energetics [77].
mentation is warranted to establish how the antiviral and antitumor ef- Although hrHPV type infections are necessary for development of
fects of these bacteria contribute to overall protection against malignancy. cervical cancer, in most cases the virus is cleared by the host. Only
0.3–1.2% of women eventually develop cervical cancer [78] which
4. Cervicovaginal microbiome means that additional risk factors contribute to the disease progression
and vaginal microbiota is possibly one of them. It appears that both, the
Despite the fact that microbiota in the vagina is less diverse than in protective role of normal vaginal microbiome and the contribution of
the gut it can also contribute to protection against or susceptibility to certain pathogens, may play a role in development of cervical in-
some illnesses, especially sexually transmitted diseases. Healthy stable traepithelial lesions and cancer.
vaginal microbiome is thought to be a first line of defense [59,60] against While Lactobacillus gasseri-dominated vaginal bacterial communities
diseases caused by opportunistic and true pathogens: outcompeting are associated with faster clearance of HPV infection [79], dysbiosis, and
dangerous bacteria or protecting the host with bacterial by-products, in bacterial vaginosis are associated with CIN development and progression
particular lactic acid [61–64] and hydrogen peroxide [60,65,66]. [80,81] (Fig. 3D). Furthermore, it is unclear whether bacteria in the
Lactobacilli, commonly considered as such beneficial microbes, re- disrupted vaginal microbiome affect host susceptibility, virus survival, or
present the dominant genus [67,68] in healthy vaginal microbiota. infectivity. Some evidence (discussed below), however, points to the
However, this does not seem to be universally applicable as Gardner- possible involvement of Prevotella genus that contains Prevotella bivia, a
ella-dominated microbiomes were observed as another frequent type in microbe known to be associated with bacterial vaginosis.
some populations [69]. Disruption of healthy vaginal microbiota Normally inhibited by Lactobacillus [65], Prevotella species may
(bacterial vaginosis - BV) [67] may increase risks of sexually trans- become abundant when the homeostasis of the vaginal microbial
mitted infections (STI) [70], and even contribute to the disease pro- community is disrupted by such factors as diet or hormone status [82]
gression, putatively, including cancer development. (Fig. 3E). Recent findings also suggest host genetics as an important
factor in Prevotella outgrowth [82]. Increasing in abundance, Pre-
4.1. Human papillomavirus votella species may provide nutrients (e.g. ammonia and amino acids),
to other members of microbial community such as Gardnerella vaginalis
Human papillomavirus (HPV) is the most common STI in the United and Peptostreptococcus anaerobius [83,84], and thus diversify the vaginal
States and although the majority of HPV types are non-carcinogenic, microbial landscape [69] (Fig. 3F). Furthermore, multiple studies of
there are at least 13 high-risk oncogenic types (hrHPV), with HPV16 Prevotella associations with bacterial vaginosis and cervicitis
and HPV18 leading the list [71,72]. One of the most common cancers [68,82,85] point to Prevotella as a conductor orchestrating the state of
caused by HPV infection remains cervical squamous cell carcinoma. vaginal microbiomes. Additionally, a clear link between Prevotella
However, HPV is also linked to anal, vulvar, vaginal, penile and head- genus and HPV infection [86], in particular with high risk HPV types
and-neck carcinomas [73]. [87], has been established. Adding to the potential importance of these

16
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

Fig. 3. Cervicovaginal microbiota and pathogens can influence the progression of viral infection associated with cancer development: A) Lactobacillus spp., predominant bacteria in the
vaginal microbiome, secretes lactic acid capable of inactivating HIV; B) butyric acid secreted by vaginal microbiota induces HIV replication in Human CD4+ T lymphocyte and
macrophage/monocyte cell lines harboring latent HIV; C) Neisseria gonorrhoeae induces IL-6, IL-8 and TNFα production in genital epithelial cells and upregulates HIV gene expression in
T-cells via heptose-monophosphate induced NF-kB and AP-1 pathways; D) domination of vaginal microbiota by Lactobacillus gasseri is associated with faster HPV clearance; E)
Lactobacillus spp. is able to inhibit the viability of Gardnerella vaginalis and Prevotella bivia; F) Prevotella spp. produces ammonia and different amino acids that benefit G. vaginalis and
Peptostreptococcus anaerobius, contributing to bacterial vaginosis; G) Chlamydia trichomatis may influence cervical cancer development, either through synergy with HPV or by con-
tributing to local chronic inflammation.

microbes, Prevotella was cultured from cervical cancer samples in 1993 4.2. Human immunodeficiency virus
[88], and more recently we detected it as the most abundant genus in
cervical cancer biopsies (unpublished). Finally, increased expression of Human Immunodeficiency Virus (HIV) is a carcinogenic virus in its
NF-κB, Toll-like receptor (TLR), NOD-like receptor, and TNF-α sig- ability to act as a co-factor for EBV and KSHV in carcinogenesis [93].
naling pathways in antigen presenting cells from blood, and increased Immunosuppression, chronic antigenic stimulation, and cytokine dys-
levels of pro-inflammatory cytokines from vaginal lavage have been regulation are reported to contribute to HIV-associated cancer devel-
associated with microbial communities that include Prevotella [69]. opment. However, the main cause of HIV-related cancer development
Thus, it is likely that Prevotella or Prevotella-driven vaginal micro- remains immunosuppression: the inability of the immune system to
biome may act in favor of persistent HPV infection promoting cervical recognize and clear host cells infected with potentially oncogenic
cancer development through upregulation of cell proliferation and viruses enables those viruses to express oncogenic proteins, which in its
chronic inflammation. turn leads to cancer development [6]. Due to the absence of vaccines
Aside from the common members of vaginal microbiota, pathogens and ways to eliminate HIV, understanding risk factors contributing to
have also been suspected in the promotion of HPV infection. For ex- the pathogenesis of HIV infections becomes a priority.
ample, Chlamydia trachomatis, has been studied for some time as a Sexual transmission is one of the most common ways to acquire
potential co-factor of HPV in the process of tumorigenesis [89,90]. In- HIV, suggesting that vaginal microbiota may influence this process.
vestigation of the potential mechanism of how C. trachomatis infection Indeed, several studies found an association between the risk of HIV
may influence HPV infection and cancer development are underway. infection and bacterial vaginosis [94–96]. However, it is still unclear
For example, it was demonstrated that C. trachomatis can decrease the whether the absence or overgrowth of some bacteria in the vaginal
expression of caveolin-1 (tumor suppressor) and increase C-myc mRNA microbiota would make women more susceptible to or protected from
levels (oncogene) [91]. A study conducted by Paba et al. found a cor- HIV infection.
relation between C. trichomonas infection and upregulation of cyto- In 2013, Aldunate et al. showed that physiological concentrations of
plasmic and nuclear NF-kB, VEGF-c and survivin in HPV-positive CINs lactic acid can inactivate different HIV subtypes [97]. Thus, Lactobacillus
and cervical cancer [92], which points out to the possibility that C. spp. which is known to produce lactic acid, may play an important role in
trachomatis can also act through the NF-kB pathway, promoting local protection against sexually transmitted diseases, including HIV [98]
inflammation, cell survival and proliferation (Fig. 3G). Despite many (Fig. 3A). In contrast, bacterial vaginosis predisposes females to sexually
studies and even mechanistic research, the scientific community has not transmitted diseases [99], and can reactivate latent HIV infection
yet reached a consensus about whether C. trachomatis plays a causal [100,101]. Evidence suggests that bacteria are not acting directly on HIV,
role in aiding HPV in carcinogenesis. Thus epidemiological studies that but rather via bacterial by-products (e.g. butyric acid) [105,106]. Speci-
would investigate the temporal relationship between C. trachomatis and fically, immunosuppression caused by HIV is an important step in the
HPV infections are required to settle this long standing debate. establishment of lifelong latent infection and avoidance of host immune

17
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

Fig. 4. The role of periodontal disease in


head and neck cancers can be mediated by
the effects of HIV and herpesviruses (EBV
and KSHV): A) butyric acid produced by
periodontal pathogens such as Fusobacterium
nucleatum and Porphyromonas gingivalis
competitively inhibit Histone Deacetylases
(HDACs), resulting in the reduced deacety-
lation of histone proteins and sustaining
viral gene expression B) activation of HIV
proviral gene expression through HDAC in-
hibition increases viral progeny production;
HIV infection suppresses systemic immune
responses and aides in the acquisition of
other viral infections; C) inhibition of HDAC
causes transcriptional activation of the
BZLF1 promoter in EBV infected cells and
results in production of ZEBRA, known
switch for lytic cycle activation; D) inhibi-
tion of HDACs in KSHV-associated cells re-
sult in the induction of MAPK and expres-
sion of lytic genes. Both EBV (C) and KSHV
(D) mechanisms promote the spread of viral
infection and contributes to tumor develop-
ment.

response. Latently infected cells harbor the HIV proviral DNA genome tract, the connection between HIV, oral microbiome, and periodontal
integrated into heterochromatin, allowing the persistence of tran- disease has also drawn the attention of the scientific community
scriptionally silent proviruses [102]. Hypoacetylation of histone proteins [113–116] [105,106]. Recent works have found associations between
by histone deacetylases (HDAC) is involved in the maintenance of HIV periodontitis, a chronic inflammatory disease of the periodontium oc-
latency by repressing viral transcription [103,104]. Interestingly, butyric curring in response to bacterial infection, and several types of oral
acid-producing bacteria (Fusobacterium nucleatum, Porphyromonas gingi- cancers [117–120]. Periodontal disease is marked by a disruption in the
valis, Clostridium cochlearium, Eubacterium multiforme, and Anaerococcus oral microbiome and is often associated with a shift to anaerobic bac-
tetradius), residing in different mucosal environments (gut, vaginal, and teria such as Porphyromonas gingivalis, that has also been directly im-
oral cavities), are capable of reactivating latent HIV infection through host plicated in cancers [121,122]. Although common risk factors for these
HDAC pathways [105,106] (Figs. 3B, 4B). Butyric acid is known to inhibit cancers include smoking and alcohol abuse, there are an increasing
the enzymatic activity of HDAC by directly competing with HDAC sub- number of cases where no significant smoking or drinking history has
strate at the catalytic center of the enzyme [107] (Fig. 4A). Butyrate ori- been reported. Among other risk factors are viral infections HIV, HPV,
ginating from microbiota has been shown to promote im- herpesviruses (EBV, HCMV, KSHV) and oral hygiene [123]. Herpes-
munosuppressive/immunoregulatory effects [108]. Considering this, it is viruses, particularly EBV and KSHV, are known to cause cancer in the
plausible that in addition to reactivating viruses, butyric acid can inhibit context of immunodeficiency caused by HIV/AIDS [124].
antiviral immunity in the host directly.
Additionally, the epidemiological synergy of sexually transmitted in- 5.1. Herpesviruses
fections (STI) with HIV transmission was reported in multiple studies.
Ferreira et al. found that stimulation of HIV infected T-cells with TLR3, 4, Herpesviruses have recently been implicated in the progression of
or 5 ligands leads to enhanced HIV-1 replication via direct activation of periodontitis in the oral cavity [125–129]. Among these viruses, KSHV,
HIV long terminal repeats, or by inducing secreted factors that promote EBV, and HCMV have also been associated with head and neck cancers,
HIV replication [109]. Among the pathogens suspected to contribute to primarily those found in the mouth. A hallmark of herpesviruses is their
HIV infection and reactivation, one player stands out - Neisseria gonor- ability to establish life-long latent infection in host cells. During latent
rhoeae: it was found to be highly associated with HIV infection [110] and infection, certain viral genes are repressed and viral progenies are not
activation of HIV expression [109,111,112]. Not only does N. gonorrhoeae produced. The reactivation of lytic cycle genes in virus-infected cells
drives an increase of activated CD4+ T-cells [111] and pro-inflammatory marks the production of viral progeny which ultimately leads to host
cytokines in genital epithelial cells such as IL-6, IL-8 and TNFα [109], but cell lysis. Both latent and lytic cycle genes are critical for tumorigenesis
exposure to the pathogen directly drives HIV expression in T cells in NF- and the evasion of host immune response. In addition, studies exploring
kB and AP-1 dependent manner [109,111]. Malott et al. showed that N. co-infection and crosstalk between these viruses as well as with HIV has
gonorrhoeae by-product heptose-monophosphate is necessary for invoking shown direct (virus-virus) and indirect (through host immunity) inter-
these host responses [111] (Fig. 3C). Thus, the course of HIV infection can actions, most notably affecting the transition between latent and lytic
be altered negatively (by lactate producers) or positively (by butyrate viral cycles [124,130].
producers or N. gonorrhoeae), preventing or favoring, respectively, de- Human cytomegalovirus (HCMV) is not usually regarded as an onco-
velopment of immunodeficiency, which may increase susceptibility to genic virus. However, HCMV infections have been implicated in malignant
EBV and KSHV infection, which are responsible for a substantial pro- diseases from different cancer entities and can cause fatal diseases in
portion of HIV-related cancers. immunocompromised patients [131]. Some evidence also suggests its role
in sustaining chronic inflammation in the progression of cancer [132].
5. Oral microbiome However, in the case of oral squamous cell carcinomas, Saravani et al.
reported that only 6.3% of 48 patient samples were found to have de-
Although many studies have been devoted to HIV in the genital tectable HCMV [133]. Although this patient population had low incidence

18
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

Table 1
Summary of findings for each virus discussed in this review. Information presented is limited to the scope of the article.

Virus Viral Location Cancer Type Bacteria-Virus Microbiome Bacteria Bacterial By-products Bacteria Impact on Host
Interaction Location

HCV liver hepatocellular indirect gut, liver Helicobacter pylori [37–39], activation of NF-κB [6,41]
carcinoma Helicobacter hepaticus cirrhosis [47]
[37,41] HCC [44]
HBV liver hepatocellular indirect gut Bifidobacterium [51,52] HBV clearance [50,53]
carcinoma anti-tumor immunity [160]
up-regulation of IFN-dependent
pathways [53]
lower serum cholesterol [54–56]
HPV vaginal cervical indirect vaginal Lactobacillus gasseri [79] decrease caveolin-1 and increase
C-myc [91]
BV-associated [80,81] upregulation of NF-kB, VEGF-c
Chlamydia trachomatis and survivin [92]
[74–76,89,174,175]
Prevotella [86,87] upregulation of NF-κB, Toll-like
receptor (TLR), NOD-like
receptor, and TNF-α signaling
pathways [69]
HIV lymphoid cells head & neck, cervical, direct, indirect gut, oral, BV-associated [100,101] lactic acid [97] TLR3,4,5 [109]
Kaposi's sarcoma vaginal activated CD4+ T-cells [111]
pro-inflammatory cytokines [109]
Neisseria gonorrhoeae heptose- NF-kB and AP-1 [109,111]
[109–112] monophosphate
[111]
butyric acid producing butyric acid inhibit HDAC [105,106]
bacteria [105,106] [105,106]
HCMV oral head & neck oral Porphyromonas gingivalis
[126–129]
KSHV oral Kaposi's sarcoma indirect oral Porphyromonas gingivalis butyric acid [136] inhibit HDAC and activate p38
[125–128,136] [136]
Fusobacterium nucleatum
[136]
EBV B lymphocytes nasopharyngeal indirect (host) oral Porphyromonas gingivalis butyric acid [138] inhibit HDAC and activate BZLF1
carcinoma [138] promoter [138]
HERVs indirect gut upregulation of type I interferon
pathways [145–147]
T-cells specific antitumor
immunity [148]

of HCMV detection, it does not discount the possibility of synergy between in humans and is associated with a wide range of human cancers ori-
HCMV and periodontal disease (not explored) in the development of oral ginating from epithelial cells, lymphocytes and mesenchymal cells
cancer subtypes. In fact, HCMV has been associated with active period- [137]. Infection is transmitted from host to host via salivary contact,
ontal disease and P. gingivalis [126–129]. Thus, strengthening the argu- and the virus passes through the oropharyngeal epithelium to B lym-
ment to keep HCMV among potential co-factors contributing to chronic phocytes, where it establishes a lifelong latent infection. EBV has three
inflammation that leads to oral tumorigenesis. main latency patterns, each of which have a role in avoiding host re-
Kaposi's sarcoma-associated herpesvirus (KSHV), also known as sponse while promoting B-cell survival/proliferation [6]. Although not
human herpesvirus 8 (HHV-8), is a herpesvirus that has become a well- deeply studied, there is increasing evidence that EBV early lytic genes,
known oncovirus in immunocompromised individuals infected with particularly those encoding homologues of Bcl-2 (BALF1), IL-10
HIV [134]. This virus is consistently associated with Kaposi's sarcoma, a (BCRF1) and c-fms receptor (BARF1), may be involved in oncogenesis
cancer developed from cells that line the lymph or blood vessels and as well as in promoting viral infection. Reactivation of the virus and
usually appears as a tumor on the skin or mucosal [135]. Latent tran- production of progeny contributes to several human diseases including
scripts in KSHV include genes and miRNAs that favor viral persistence nasopharyngeal carcinomas and lymphomas [138]. Similar to KSHV
and replication while promoting host-cell proliferation and survival. and HIV, the latent virus was found to be reactivated upon stimulation
Furthermore, lytic cycle genes favor viral replication by affecting the with supernatant from P. gingivalis (containing high concentrations of
DNA damage response, reprogramming metabolism, promoting sur- butyric acid) through the inhibition of HDACs [138]. This inhibition
vival, and mediating immune evasion, all of which have a role in pro- resulted in increased acetylation of adjacent histone and transcriptional
moting cancer development [6]. In 2007, Morris et al. explored the activation of the BZLF1 promoter, whose gene product ZEBRA is known
effects of supernatants from cultures of different bacteria on KSHV in- to be the master regulator in EBV transition from latency to lytic state
fected BCLBL-1 cells [136]. They found that metabolic end products [138] (Fig. 4D). Taken together with the case of KSHV, there is com-
from pathogens (gram-negative anaerobic periodontopathogens F. nu- pelling evidence for the ability of butyric acid producing bacteria such
cleatum and P. gingivalis) induced lytic replication of KSHV through the as P. gingivalis (main player in periodontitis) to reactivate latent her-
activation of a stress-activated MAPK pathway in host cells. More spe- pesvirus infection and contribute to the development of cancer in the
cifically, butyric acid from bacteria supernatants were found to inhibit oral cavity.
cellular HDACs and activate the p38 kinases pathway, resulting in hy- As described above, butyric-acid producing bacteria (especially the
peracetylation of histones on immediate early viral promoters (Fig. 4C). periodontopathogens P. gingivalis) can regulate the viral life cycle in
The targeting of HDACs by P. gingivalis is also seen in the reactivation of host cells. Periodontopathogens can induce a reactivation of HIV which
HIV as we discussed previously. in turn may lead to the opportunistic infection of herpesviruses due to
Epstein–Barr virus (EBV) was the first virus shown to cause cancer immunodeficiency. Latent infection of EBV and KSHV can be

19
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

reactivated by the same bacteria leading to the induction of oncogenes liver cancer, for which diabetes and obesity have been identified as risk
and transformation to malignancy. Independent of bacteria, a model of factors [149,157]. Interestingly, in obese people among many changes
KSHV and EBV co-infection in vitro has shown complementing lytic in the gut microbiome, a decrease in Bifidobacterium was reported
activation [139], suggesting a more complex model of reactivation in [158]. Because this bacterium has been shown to play a positive role in
which periodontopathogens may act in tandem with direct inter-viral elimination of HBV (as described in Section 3.1), this may partially
interaction to regulate infection and the development of cancers. Al- explain the association between diet-induced obesity and liver cancer.
though HCMV has not been studied extensively in the same context, its Although environmental and hereditary factors make virus-bacteria-
association with P. gingivalis and periodontal disease offers hints of host interactions even more complex to investigate, it is crucial to re-
viral-bacterial interaction potentially involving mechanisms used by member that understanding the role of these external factors may be-
other herpesviruses. come a powerful tool for cancer prevention and treatment.

6. Human endogenous retroviruses 8. Summary

Endogenous retroviruses (ERVs) are endogenous viral elements lo- It has recently become evident that progress in the field of virus-
cated in the genome of jawed vertebrates, including humans, and are associated cancers can be enhanced by elucidation of how bacterial
thought to be relics of ancestral infectious retroviruses [140]. Human microbiota contributes to virus-host interaction. Although the dissec-
endogenous retroviruses (HERVs) compose about 4–8% of human tion of these transkingdom interactions is evolving [159], we are still in
genome [141]. While there is not enough evidence to solidify the role of the early stages of this journey. Indeed, epidemiological and even some
HERVs in causing human cancers, the connection between the two is mechanistic studies are accumulating. However, the scientific com-
constantly been researched [140]. Reactivation of HERV expression in munity has not yet reached a unanimous agreement on whether any
cancers [140,142] can contribute to genome instability and is suspected specific cancer requires bacteria-virus interaction for carcinogenesis.
to be a prominent player in disease development. Furthermore, resur- Herein, we have described a model of bacteria-virus interactions in
rection of murine leukemia virus (MLV) in immunocompromised mice the development of cancer. We identify two main mechanisms; the first
was found to be dependent on intestinal microbiota [143]. In humans, is accomplished by the interaction between bacteria and virus resulting
environmental and intestinal microbes were able to modulate the in changes in the course of viral infection. The most notable example of
transcriptional activity of endogenous retroviruses [144]. It is not clear such mechanism is the inactivation of HIV via bacteria-derived lactic
whether this process contributes to cancer development. Nonetheless, acid [97]. The second mechanism is an effect of the bacteria on the host
multiple studies show that upregulated HERV expression in cancerous resulting in alterations of host susceptibility to viral infection. This can
cells can drive inflammatory responses by upregulating type I inter- happen in several ways: a) bacteria inducing pro-tumor chronic in-
feron pathways [145–147] and eliciting T-cell specific antitumor im- flammation (e.g. activation of NF-kB pathway [6,41,69,92,109,111]),
munity [148]. These findings may indicate the role of the immune b) commensal bacteria promoting antiviral and antitumor immunity
system in recognizing the reactivation of “sleeping” HERVs as a sign of (e.g. Bifidobacterium in HBV [53,160]), and c) bacterial metabolite
cell transformation and clearing the potential threat before it develops reactivating oncoviruses (butyric acid inhibition of host HDAC path-
into malignancies. ways reactivating HIV [105,106], EBV [138], KSHV [136]) (Table 1).
In addition to proposing the model, we identified key questions that
7. Risk factors of virus-associated cancers and microbiome have to be answered in order to move the field forward (Box 1) and
major technological/logistical solutions that are required to answer
Risk factors for cancer, such as age, lifestyle, diet and genetics, have these questions (Box 2).
been identified but the mechanisms underlying their contribution to Finally, treatment of cancer with chemicals and radiation are cur-
disease are not always well understood [149,150]. The discovery of the rently the most popular and efficient strategies [161,162]. More re-
role of commensal microbiota and pathobionts in cancer development cently, our armamentarium was advanced with vaccines against on-
and progression suggests that at least some risk factors may be linked to coviruses [163] and immunotherapy [164], both using the immune
the disease through microbiota. For example, while smoking is a risk system to prevent or kill cancer. Unfortunately, neither of these stra-
factor for nasopharyngeal carcinoma [151], it is also associated with tegies allowed us to eliminate completely any oncovirus, nor to cure
dysbiosis in oral microbiota [152], which in its turn can contribute to some cancers.
periodontal disease leading to cancer development [153]. In cervical Therefore, novel approaches are required to significantly change the
cancer, a high number of sexual partners is a well-known risk factor. status quo of this field. We believe that this change should come with
This association has been commonly attributed to increased chances to methods promoting a healthy microbiome, development of next gen-
be exposed to high risk HPV [154]. However, women with multiple eration antibiotics targeting individual bacteria, new probiotics, and
sexual partners also present disruption of vaginal microbiota (bacterial companion diagnostics that will define a course of personalized/pre-
vaginosis) [155,156] that may facilitate chronic HPV infection and cision medicine for each individual patient based on their respective
cervical cancer as discussed earlier (Section 4.1). Another example is transcriptome and microbiome.

Box 1
Current trending questions:

- Which shifts in microbial communities (presence/absence of particular bacterial species or alterations in microbial community structures)
can enhance oncogenic virus infection progression or, on the contrary, help to eliminate it?
- Which host molecular pathways involved in viral infections and tumorigenesis are altered by microbiota?
- Can host genetics shape microbiota toward being anti- or pro-tumorigenic?
- How do microbiota affect the success of anticancer therapies?
- How do antitumor treatments affect microbiota? Are these effects relevant for efficient treatments and treatment-related co-morbidities?

20
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

Box 2
Emerging technologies and other solutions:

- Experimental approaches for the generation of different types of omics data that would allow simultaneous assessment of functional states
of each of viruses, host and bacteria (e.g. new and improved single cell technologies for both eukaryotic and prokaryotic cells).
- Computational tools for the analysis of multi-omics datasets that would provide robust predictions of regulatory relationships between all
three kingdoms (e.g. transkingdom networks [165,166], LEfSe [167]).
- Generation of new experimental models: a) humanized animal models (e.g. gnotobiotic animal models for cancer caused by oncogenic
viruses harboring human microbiota and immune system); b) cell lines and tissues with virus and bacteria present (e.g. differentiated 3D
cell aggregates colonized with specific bacteria [168–173]); c) next generation in vitro models such as organ-on-chip and human-on-chip
microfluidic devices that allow application of robotic systems.
- Creation of multidisciplinary teams (involving oncologists with expertise in basic and clinical science, systems biologist, virologist, micro-
biologists, and engineers) through special funding mechanisms or even the creation of new institutions.

Funding [20] D.P. Page Thomas, B. King, T. Stephens, J.T. Dingle, In vivo studies of cartilage
regeneration after damage induced by catabolin/interleukin-1, Ann. Rheum. Dis.
50 (2) (1991) 75–80.
This work was supported by the following grants: NIH U01 [21] C.M. Robinson, P.R. Jesudhasan, J.K. Pfeiffer, Bacterial lipopolysaccharide
AI109695 (AM) and R01 DK103761 (NS) binding enhances virion stability and promotes environmental fitness of an enteric
virus, Cell Host Microbe 15 (1) (2014) 36–46.
[22] F.M. Wilson, 2nd, ostler HB: superior limbic keratoconjunctivitis, Int. Ophthalmol.
Acknowledgements Clin. 26 (4) (1986) 99–112.
[23] A.J. Macpherson, N.L. Harris, Interactions between commensal intestinal bacteria
and the immune system, Nat. Rev. Immunol. 4 (6) (2004) 478–485.
We would like to thank members of the Morgun and Shulzhenko [24] M.C. Abt, L.C. Osborne, L.A. Monticelli, T.A. Doering, T. Alenghat,
laboratories for their discussions and support in the review. Special G.F. Sonnenberg, M.A. Paley, M. Antenus, K.L. Williams, J. Erikson, Commensal
consideration to Karen DSouza, Kimberly White, and Nicholas Brown bacteria calibrate the activation threshold of innate antiviral immunity, Immunity
37 (1) (2012) 158–170.
for their assistance in text editing.
[25] A.B. Shreiner, J.Y. Kao, V.B. Young, The gut microbiome in health and in disease,
Curr. Opin. Gastroenterol. 31 (1) (2015) 69–75.
References [26] R.L. Greer, A. Morgun, N. Shulzhenko, Bridging immunity and lipid metabolism by
gut microbiota, J. Allergy Clin. Immunol. 132 (2) (2013) 253–262 (quiz 263).
[27] N. Shulzhenko, A. Morgun, W. Hsiao, M. Battle, M. Yao, O. Gavrilova, M. Orandle,
[1] GLOBOCAN 2012: Estimated Cancer Incidence Mortality and Prevalence L. Mayer, A.J. Macpherson, K.D. McCoy, Crosstalk between B lymphocytes, mi-
Worldwide in 2012, http://globocan.iarc.fr/Pages/fact_sheets_cancer.aspx? crobiota and the intestinal epithelium governs immunity versus metabolism in the
cancer=all. gut, Nat. Med. 17 (12) (2011) 1585–1593.
[2] Cancer, http://www.who.int/cancer/en/. [28] G. Musso, R. Gambino, M. Cassader, Obesity, diabetes, and gut microbiota: the
[3] G.G. Luo, J.H. Ou, Oncogenic viruses and cancer, Virol. Sin. 30 (2) (2015) 83–84. hygiene hypothesis expanded? Diabetes Care 33 (10) (2010) 2277–2284.
[4] D.M. Parkin, The global health burden of infection-associated cancers in the year [29] P. Louis, G.L. Hold, H.J. Flint, The gut microbiota, bacterial metabolites and
2002, Int. J. Cancer 118 (12) (2006) 3030–3044. colorectal cancer, Nat. Rev. Microbiol. 12 (10) (2014) 661–672.
[5] P.S. Moore, Y. Chang, Why do viruses cause cancer? Highlights of the first century [30] H. Tlaskalova-Hogenova, R. Stepankova, H. Kozakova, T. Hudcovic, L. Vannucci,
of human tumour virology, Nat. Rev. Cancer 10 (12) (2010) 878–889. L. Tuckova, P. Rossmann, T. Hrncir, M. Kverka, Z. Zakostelska, The role of gut
[6] E.A. Mesri, M.A. Feitelson, K. Munger, Human viral oncogenesis: a cancer hall- microbiota (commensal bacteria) and the mucosal barrier in the pathogenesis of
marks analysis, Cell Host Microbe 15 (3) (2014) 266–282. inflammatory and autoimmune diseases and cancer: contribution of germ-free and
[7] R.D. Situnayake, D.I. Thurnham, S. Kootathep, S. Chirico, J. Lunec, M. Davis, gnotobiotic animal models of human diseases, Cell Mol. Immunol. 8 (2) (2011)
B. McConkey, Chain breaking antioxidant status in rheumatoid arthritis: clinical 110–120.
and laboratory correlates, Ann. Rheum. Dis. 50 (2) (1991) 81–86. [31] N. Iida, A. Dzutsev, C.A. Stewart, L. Smith, N. Bouladoux, R.A. Weingarten,
[8] V. Buchli, W.B. Pearce, Listening behavior in coorientational states, J. Commun. D.A. Molina, R. Salcedo, T. Back, S. Cramer, Commensal bacteria control cancer
24 (3) (1974) 62–70. response to therapy by modulating the tumor microenvironment, Science 342
[9] H.R. Konrad, C.C. Rattenborg, Combined action of laryngeal muscles, Acta (6161) (2013) 967–970.
Otolaryngol. 67 (6) (1969) 646–649. [32] S. Viaud, F. Saccheri, G. Mignot, T. Yamazaki, R. Daillere, D. Hannani, D.P. Enot,
[10] M.T. Baldridge, T.J. Nice, B.T. McCune, C.C. Yokoyama, A. Kambal, M. Wheadon, C. Pfirschke, C. Engblom, M.J. Pittet, The intestinal microbiota modulates the
M.S. Diamond, Y. Ivanova, M. Artyomov, H.W. Virgin, Commensal microbes and anticancer immune effects of cyclophosphamide, Science 342 (6161) (2013)
interferon-lambda determine persistence of enteric murine norovirus infection, 971–976.
Science 347 (6219) (2015) 266–269. [33] M. Karin, C. Jobin, F. Balkwill, Chemotherapy, immunity and microbiota–a new
[11] R.L. Greer, X. Dong, A.C. Moraes, R.A. Zielke, G.R. Fernandes, E. Peremyslova, triumvirate? Nat. Med. 20 (2) (2014) 126–127.
S. Vasquez-Perez, A.A. Schoenborn, E.P. Gomes, A.C. Pereira, Akkermansia mu- [34] Cancer Fact Sheet, http://www.who.int/mediacentre/factsheets/fs297/en/.
ciniphila mediates negative effects of IFNgamma on glucose metabolism, Nat. [35] K.J. Lafaro, A.N. Demirjian, T.M. Pawlik, Epidemiology of hepatocellular carci-
Commun. 7 (2016) 13329. noma, Surg. Oncol. Clin. N. Am. 24 (1) (2015) 1–17.
[12] C. Xuan, J.M. Shamonki, A. Chung, M.L. Dinome, M. Chung, P.A. Sieling, D.J. Lee, [36] C. de Martel, D. Maucort-Boulch, M. Plummer, S. Franceschi, World-wide relative
Microbial dysbiosis is associated with human breast cancer, PLoS One 9 (1) (2014) contribution of hepatitis B and C viruses in hepatocellular carcinoma, Hepatology
e83744. 62 (4) (2015) 1190–1200.
[13] B. Boursi, R. Mamtani, K. Haynes, Y.X. Yang, Recurrent antibiotic exposure may [37] J.M. Ward, J.G. Fox, M.R. Anver, D.C. Haines, C.V. George, M.J. Collins Jr.,
promote cancer formation–Another step in understanding the role of the human P.L. Gorelick, K. Nagashima, M.A. Gonda, R.V. Gilden, Chronic active hepatitis
microbiota? Eur. J. Cancer 51 (17) (2015) 2655–2664. and associated liver tumors in mice caused by a persistent bacterial infection with
[14] H.B. Mabrok, R. Klopfleisch, K.Z. Ghanem, T. Clavel, M. Blaut, G. Loh, Lignan a novel Helicobacter species, J. Natl. Cancer Inst. 86 (16) (1994) 1222–1227.
transformation by gut bacteria lowers tumor burden in a gnotobiotic rat model of [38] N. Leone, R. Pellicano, F. Brunello, M.A. Cutufia, M. Berrutti, S. Fagoonee,
breast cancer, Carcinogenesis 33 (1) (2012) 203–208. M. Rizzetto, A. Ponzetto, Helicobacter pylori seroprevalence in patients with cir-
[15] D.B. Polk, R.M. Peek, Jr.: Helicobacter pylori: gastric cancer and beyond, Nat. Rev. rhosis of the liver and hepatocellular carcinoma, Cancer Detect. Prev. 27 (6)
Cancer 10 (6) (2010) 403–414. (2003) 494–497.
[16] J.C. Welton, J.S. Marr, S.M. Friedman, Association between hepatobiliary cancer [39] M. Rocha, P. Avenaud, A. Menard, B. Le Bail, C. Balabaud, P. Bioulac-Sage,
and typhoid carrier state, Lancet 1 (8120) (1979) 791–794. D.M. de Magalhaes Queiroz, F. Megraud, Association of Helicobacter species with
[17] A.S. Abdulamir, R.R. Hafidh, F. Abu Bakar, The association of Streptococcus hepatitis C cirrhosis with or without hepatocellular carcinoma, Gut 54 (3) (2005)
bovis/gallolyticus with colorectal tumors: the nature and the underlying me- 396–401.
chanisms of its etiological role, J. Exp. Clin. Cancer Res. 30 (2011) 11. [40] A. Garcia, Y. Feng, N.M. Parry, A. McCabe, M.W. Mobley, K. Lertpiriyapong,
[18] R. Sender, S. Fuchs, R. Milo, Revised estimates for the number of human and M.T. Whary, J.G. Fox, Helicobacter pylori infection does not promote hepatocel-
bacteria cells in the body, PLoS Biol. 14 (8) (2016) e1002533. lular cancer in a transgenic mouse model of hepatitis C virus pathogenesis, Gut
[19] J. Wilks, E. Lien, A.N. Jacobson, M.A. Fischbach, N. Qureshi, A.V. Chervonsky, Microbes 4 (6) (2013) 577–590.
T.V. Golovkina, Mammalian lipopolysaccharide receptors incorporated into the [41] J.G. Fox, Y. Feng, E.J. Theve, A.R. Raczynski, J.L. Fiala, A.L. Doernte, M. Williams,
retroviral envelope augment virus transmission, Cell Host Microbe 18 (4) (2015) J.L. McFaline, J.M. Essigmann, D.B. Schauer, Gut microbes define liver cancer risk
456–462.

21
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

in mice exposed to chemical and viral transgenic hepatocarcinogens, Gut 59 (1) bacterial vaginosis, BMC Genomics 11 (2010) 488.
(2010) 88–97. [69] M.N. Anahtar, E.H. Byrne, K.E. Doherty, B.A. Bowman, H.S. Yamamoto,
[42] M. Hiroi, Y. Ohmori, Constitutive nuclear factor kappaB activity is required to M. Soumillon, N. Padavattan, N. Ismail, A. Moodley, M.E. Sabatini, Cervicovaginal
elicit interferon-gamma-induced expression of chemokine CXC ligand 9 (CXCL9) bacteria are a major modulator of host inflammatory responses in the female
and CXCL10 in human tumour cell lines, Biochem. J. 376 (Pt 2) (2003) 393–402. genital tract, Immunity 42 (5) (2015) 965–976.
[43] R.J. Cahill, C.J. Foltz, J.G. Fox, C.A. Dangler, F. Powrie, D.B. Schauer, [70] R.M. Brotman, Vaginal microbiome and sexually transmitted infections: an epi-
Inflammatory bowel disease: an immunity-mediated condition triggered by bac- demiologic perspective, J. Clin. Invest. 121 (12) (2011) 4610–4617.
terial infection with Helicobacter hepaticus, Infect. Immun. 65 (8) (1997) [71] J.M. Walboomers, M.V. Jacobs, M.M. Manos, F.X. Bosch, J.A. Kummer, K.V. Shah,
3126–3131. P.J. Snijders, J. Peto, C.J. Meijer, N. Munoz, Human papillomavirus is a necessary
[44] Y. Huang, X.G. Fan, Z.M. Wang, J.H. Zhou, X.F. Tian, N. Li, Identification of he- cause of invasive cervical cancer worldwide, J. Pathol. 189 (1) (1999) 12–19.
licobacter species in human liver samples from patients with primary hepatocel- [72] F.X. Bosch, M.M. Manos, N. Munoz, M. Sherman, A.M. Jansen, J. Peto,
lular carcinoma, J. Clin. Pathol. 57 (12) (2004) 1273–1277. M.H. Schiffman, V. Moreno, R. Kurman, K.V. Shah, Prevalence of human pa-
[45] H.O. Nilsson, J. Taneera, M. Castedal, E. Glatz, R. Olsson, T. Wadstrom, pillomavirus in cervical cancer: a worldwide perspective: international biological
Identification of Helicobacter pylori and other Helicobacter species by PCR, hy- study on cervical cancer (IBSCC) Study Group, J. Natl. Cancer Inst. 87 (11) (1995)
bridization, and partial DNA sequencing in human liver samples from patients 796–802.
with primary sclerosing cholangitis or primary biliary cirrhosis, J. Clin. Microbiol. [73] M.E. McLaughlin-Drubin, J. Meyers, K. Munger, Cancer associated human pa-
38 (3) (2000) 1072–1076. pillomaviruses, Curr Opin Virol 2 (4) (2012) 459–466.
[46] J.G. Fox, F.E. Dewhirst, Z. Shen, Y. Feng, N.S. Taylor, B.J. Paster, R.L. Ericson, [74] A.J. Remmink, J.M. Walboomers, T.J. Helmerhorst, F.J. Voorhorst, L. Rozendaal,
C.N. Lau, P. Correa, J.C. Araya, Hepatic Helicobacter species identified in bile and E.K. Risse, C.J. Meijer, P. Kenemans, The presence of persistent high-risk HPV
gallbladder tissue from Chileans with chronic cholecystitis, Gastroenterology 114 genotypes in dysplastic cervical lesions is associated with progressive disease:
(4) (1998) 755–763. natural history up to 36 months, Int. J. Cancer 61 (3) (1995) 306–311.
[47] M. Rowen, M. Myers, R.A. Williamson, Emphysematous gastritis in a leukemic [75] H.C. Chen, M. Schiffman, C.Y. Lin, M.H. Pan, S.L. You, L.C. Chuang, C.Y. Hsieh,
child, Med. Pediatr. Oncol. 2 (4) (1976) 433–437. K.L. Liaw, A.W. Hsing, C.J. Chen, Persistence of type-specific human papilloma-
[48] T. Poutahidis, K. Cappelle, T. Levkovich, C.W. Lee, M. Doulberis, Z. Ge, J.G. Fox, virus infection and increased long-term risk of cervical cancer, J. Natl. Cancer Inst.
B.H. Horwitz, S.E. Erdman, Pathogenic intestinal bacteria enhance prostate cancer 103 (18) (2011) 1387–1396.
development via systemic activation of immune cells in mice, PLoS One 8 (8) [76] K.S. Cuschieri, H.A. Cubie, M.W. Whitley, G. Gilkison, M.J. Arends, C. Graham,
(2013) e73933. E. McGoogan, Persistent high risk HPV infection associated with development of
[49] Y. Chen, Z. Chen, R. Guo, N. Chen, H. Lu, S. Huang, J. Wang, L. Li, Correlation cervical neoplasia in a prospective population study, J. Clin. Pathol. 58 (9) (2005)
between gastrointestinal fungi and varying degrees of chronic hepatitis B virus 946–950.
infection, Diagn. Microbiol. Infect. Dis. 70 (4) (2011) 492–498. [77] C.A. Moody, L.A. Laimins, Human papillomavirus oncoproteins: pathways to
[50] H.H. Chou, W.H. Chien, L.L. Wu, C.H. Cheng, C.H. Chung, J.H. Horng, Y.H. Ni, transformation, Nat. Rev. Cancer 10 (8) (2010) 550–560.
H.T. Tseng, D. Wu, X. Lu, Age-related immune clearance of hepatitis B virus in- [78] N. Shulzhenko, H. Lyng, G.F. Sanson, A. Morgun, Menage a trois: an evolutionary
fection requires the establishment of gut microbiota, Proc. Natl. Acad. Sci. U. S. A. interplay between human papillomavirus, a tumor, and a woman, Trends
112 (7) (2015) 2175–2180. Microbiol. 22 (6) (2014) 345–353.
[51] H. Lu, Z. Wu, W. Xu, J. Yang, Y. Chen, L. Li, Intestinal microbiota was assessed in [79] R.M. Brotman, M.D. Shardell, P. Gajer, J.K. Tracy, J.M. Zenilman, J. Ravel,
cirrhotic patients with hepatitis B virus infection: intestinal microbiota of HBV P.E. Gravitt, Interplay between the temporal dynamics of the vaginal microbiota
cirrhotic patients, Microb. Ecol. 61 (3) (2011) 693–703. and human papillomavirus detection, J. Infect. Dis. 210 (11) (2014) 1723–1733.
[52] M. Xu, B. Wang, Y. Fu, Y. Chen, F. Yang, H. Lu, Y. Chen, J. Xu, L. Li, Changes of [80] J.J. Platz-Christensen, E. Sundstrom, P.G. Larsson, Bacterial vaginosis and cervical
fecal Bifidobacterium species in adult patients with hepatitis B virus-induced intraepithelial neoplasia, Acta Obstet. Gynecol. Scand. 73 (7) (1994) 586–588.
chronic liver disease, Microb. Ecol. 63 (2) (2012) 304–313. [81] H.Y. Oh, B.S. Kim, S.S. Seo, J.S. Kong, J.K. Lee, S.Y. Park, K.M. Hong, H.K. Kim,
[53] D.K. Lee, J.Y. Kang, H.S. Shin, I.H. Park, N.J. Ha, Antiviral activity of bifido- M.K. Kim, The association of uterine cervical microbiota with an increased risk for
bacterium adolescentis SPM0212 against hepatitis B virus, Arch. Pharm. Res. 36 cervical intraepithelial neoplasia in Korea, Clin. Microbiol. Infect. 21 (7) (2015)
(12) (2013) 1525–1532. (674 e671-679).
[54] D.K. Lee, S. Jang, E.H. Baek, M.J. Kim, K.S. Lee, H.S. Shin, M.J. Chung, J.E. Kim, [82] J. Si, H.J. You, J. Yu, J. Sung, G. Ko, Prevotella as a hub for vaginal microbiota
K.O. Lee, N.J. Ha, Lactic acid bacteria affect serum cholesterol levels, harmful fecal under the influence of host genetics and their association with obesity, Cell Host
enzyme activity, and fecal water content, Lipids Health Dis. 8 (2009) 21. Microbe 21 (1) (2017) 97–105.
[55] H.S. Shin, S.Y. Park, D.K. Lee, S.A. Kim, H.M. An, J.R. Kim, M.J. Kim, M.G. Cha, [83] V. Pybus, A.B. Onderdonk, Evidence for a commensal, symbiotic relationship be-
S.W. Lee, K.J. Kim, Hypocholesterolemic effect of sonication-killed tween Gardnerella vaginalis and Prevotella bivia involving ammonia: potential
Bifidobacterium longum isolated from healthy adult Koreans in high cholesterol significance for bacterial vaginosis, J. Infect. Dis. 175 (2) (1997) 406–413.
fed rats, Arch. Pharm. Res. 33 (9) (2010) 1425–1431. [84] V. Pybus, A.B. Onderdonk, A commensal symbiosis between Prevotella bivia and
[56] D.K. Lee, S.Y. Park, S. Jang, E.H. Baek, M.J. Kim, S.M. Huh, K.S. Choi, M.J. Chung, Peptostreptococcus anaerobius involves amino acids: potential significance to the
J.E. Kim, K.O. Lee, The combination of mixed lactic acid bacteria and dietary fiber pathogenesis of bacterial vaginosis, FEMS Immunol. Med. Microbiol. 22 (4) (1998)
lowers serum cholesterol levels and fecal harmful enzyme activities in rats, Arch. 317–327.
Pharm. Res. 34 (1) (2011) 23–29. [85] Z. Ling, X. Liu, X. Chen, H. Zhu, K.E. Nelson, Y. Xia, L. Li, C. Xiang, Diversity of
[57] Y.L. Lin, M.S. Shiao, C. Mettling, C.K. Chou, Cholesterol requirement of hepatitis B cervicovaginal microbiota associated with female lower genital tract infections,
surface antigen (HBsAg) secretion, Virology 314 (1) (2003) 253–260. Microb. Ecol. 61 (3) (2011) 704–714.
[58] R.L. Arenson, S.J. Dwyer 3rd, H.K. Huang, H.L. Kundel, S.B. Seshadri, Computer [86] J.E. Lee, S. Lee, H. Lee, Y.M. Song, K. Lee, M.J. Han, J. Sung, G. Ko, Association of
applications and digital imaging, Radiology 178 (3) (1991) 913–914. the vaginal microbiota with human papillomavirus infection in a Korean twin
[59] C. Nardis, L. Mosca, P. Mastromarino, Vaginal microbiota and viral sexually cohort, PLoS One 8 (5) (2013) e63514.
transmitted diseases, Ann. Ig. 25 (5) (2013) 443–456. [87] E.O. Dareng, B. Ma, A.O. Famooto, S.N. Akarolo-Anthony, R.A. Offiong,
[60] S.P. Voravuthikunchai, S. Bilasoi, O. Supamala, Antagonistic activity against pa- O. Olaniyan, P.S. Dakum, C.M. Wheeler, D. Fadrosh, H. Yang, Prevalent high-risk
thogenic bacteria by human vaginal lactobacilli, Anaerobe 12 (5–6) (2006) HPV infection and vaginal microbiota in Nigerian women, Epidemiol. Infect. 144
221–226. (1) (2016) 123–137.
[61] D.E. O'Hanlon, T.R. Moench, R.A. Cone, Vaginal pH and microbicidal lactic acid [88] H. Mikamo, K. Izumi, K. Ito, K. Watanabe, K. Ueno, T. Tamaya, Internal bacterial
when lactobacilli dominate the microbiota, PLoS One 8 (11) (2013) e80074. flora of solid uterine cervical cancer, Kansenshogaku Zasshi 67 (11) (1993)
[62] D.E. O'Hanlon, T.R. Moench, R.A. Cone, In vaginal fluid, bacteria associated with 1057–1061.
bacterial vaginosis can be suppressed with lactic acid but not hydrogen peroxide, [89] E. Samoff, E.H. Koumans, L.E. Markowitz, M. Sternberg, M.K. Sawyer, D. Swan,
BMC Infect. Dis. 11 (2011) 200. J.R. Papp, C.M. Black, E.R. Unger, Association of Chlamydia trachomatis with
[63] H. Borgdorff, E. Tsivtsivadze, R. Verhelst, M. Marzorati, S. Jurriaans, persistence of high-risk types of human papillomavirus in a cohort of female
G.F. Ndayisaba, F.H. Schuren, J.H. van de Wijgert, Lactobacillus-dominated cer- adolescents, Am. J. Epidemiol. 162 (7) (2005) 668–675.
vicovaginal microbiota associated with reduced HIV/STI prevalence and genital [90] J.S. Smith, N. Munoz, R. Herrero, J. Eluf-Neto, C. Ngelangel, S. Franceschi,
HIV viral load in African women, ISME J. 8 (9) (2014) 1781–1793. F.X. Bosch, J.M. Walboomers, R.W. Peeling, Evidence for Chlamydia trachomatis
[64] E.R. Boskey, R.A. Cone, K.J. Whaley, T.R. Moench, Origins of vaginal acidity: high as a human papillomavirus cofactor in the etiology of invasive cervical cancer in
D/L lactate ratio is consistent with bacteria being the primary source, Hum. Brazil and the Philippines, J. Infect. Dis. 185 (3) (2002) 324–331.
Reprod. 16 (9) (2001) 1809–1813. [91] T. Schlott, H. Eiffert, W. Bohne, J. Landgrebe, E. Brunner, B. Spielbauer, B. Knight,
[65] F. Atassi, D. Brassart, P. Grob, F. Graf, A.L. Servin, Lactobacillus strains isolated Chlamydia trachomatis modulates expression of tumor suppressor gene caveolin-1
from the vaginal microbiota of healthy women inhibit Prevotella bivia and and oncogene C-myc in the transformation zone of non-neoplastic cervical tissue,
Gardnerella vaginalis in coculture and cell culture, FEMS Immunol. Med. Gynecol. Oncol. 98 (3) (2005) 409–419.
Microbiol. 48 (3) (2006) 424–432. [92] P. Paba, D. Bonifacio, L. Di Bonito, D. Ombres, C. Favalli, K. Syrjanen, M. Ciotti,
[66] S.J. Klebanoff, S.L. Hillier, D.A. Eschenbach, A.M. Waltersdorph, Control of the Co-expression of HSV2 and Chlamydia trachomatis in HPV-positive cervical cancer
microbial flora of the vagina by H2O2-generating lactobacilli, J. Infect. Dis. 164 and cervical intraepithelial neoplasia lesions is associated with aberrations in key
(1) (1991) 94–100. intracellular pathways, Intervirology 51 (4) (2008) 230–234.
[67] B. Ma, L.J. Forney, J. Ravel, Vaginal microbiome: rethinking health and disease, [93] F.X. Mbopi-Keou, L. Belec, C.G. Teo, C. Scully, S.R. Porter, Synergism between HIV
Annu. Rev. Microbiol. 66 (2012) 371–389. and other viruses in the mouth, Lancet Infect. Dis. 2 (7) (2002) 416–424.
[68] Z. Ling, J. Kong, F. Liu, H. Zhu, X. Chen, Y. Wang, L. Li, K.E. Nelson, Y. Xia, [94] L. Myer, L. Denny, R. Telerant, M. Souza, T.C. Wright Jr., L. Kuhn, Bacterial va-
C. Xiang, Molecular analysis of the diversity of vaginal microbiota associated with ginosis and susceptibility to HIV infection in South African women: a nested case-

22
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

control study, J. Infect. Dis. 192 (8) (2005) 1372–1380. (2012) 1055–1058.
[95] J. Atashili, C. Poole, P.M. Ndumbe, A.A. Adimora, J.S. Smith, Bacterial vaginosis [122] N.H. Ha, B.H. Woo, D.J. Kim, E.S. Ha, J.I. Choi, S.J. Kim, B.S. Park, J.H. Lee,
and HIV acquisition: a meta-analysis of published studies, AIDS 22 (12) (2008) H.R. Park, Prolonged and repetitive exposure to Porphyromonas gingivalis in-
1493–1501. creases aggressiveness of oral cancer cells by promoting acquisition of cancer stem
[96] B.E. Sha, M.R. Zariffard, Q.J. Wang, H.Y. Chen, J. Bremer, M.H. Cohen, G.T. Spear, cell properties, Tumour Biol. 36 (12) (2015) 9947–9960.
Female genital-tract HIV load correlates inversely with Lactobacillus species but [123] S.M. Gondivkar, R.V. Parikh, A.R. Gadbail, V. Solanke, R. Chole, M. Mankar,
positively with bacterial vaginosis and Mycoplasma hominis, J. Infect. Dis. 191 (1) S. Balsaraf, Involvement of viral factors with head and neck cancers, Oral Oncol.
(2005) 25–32. 48 (3) (2012) 195–199.
[97] M. Aldunate, D. Tyssen, A. Johnson, T. Zakir, S. Sonza, T. Moench, R. Cone, [124] S.R. da Silva, de Oliveira DE: HIV, EBV. and KSHV: viral cooperation in the pa-
G. Tachedjian, Vaginal concentrations of lactic acid potently inactivate HIV, J. thogenesis of human malignancies, Cancer Lett. 305 (2) (2011) 175–185.
Antimicrob. Chemother. 68 (9) (2013) 2015–2025. [125] H. Li, V. Chen, Y. Chen, J.C. Baumgartner, C.A. Machida, Herpesviruses in en-
[98] C. Gosmann, M.N. Anahtar, S.A. Handley, M. Farcasanu, G. Abu-Ali, B.A. Bowman, dodontic pathoses: association of Epstein-Barr virus with irreversible pulpitis and
N. Padavattan, C. Desai, L. Droit, A. Moodley, Lactobacillus-Deficient cervi- apical periodontitis, J. Endod. 35 (1) (2009) 23–29.
covaginal bacterial communities are associated with increased HIV acquisition in [126] A. Kubar, I. Saygun, A. Ozdemir, M. Yapar, J. Slots, Real-time polymerase chain
young south african women, Immunity 46 (1) (2017) 29–37. reaction quantification of human cytomegalovirus and Epstein-Barr virus in per-
[99] J.H. van de Wijgert, H. Borgdorff, R. Verhelst, T. Crucitti, S. Francis, iodontal pockets and the adjacent gingiva of periodontitis lesions, J. Periodontal
H. Verstraelen, V. Jespers, The vaginal microbiota: what have we learned after a Res. 40 (2) (2005) 97–104.
decade of molecular characterization? PLoS One 9 (8) (2014) e105998. [127] M. Sabeti, Y. Valles, H. Nowzari, J.H. Simon, V. Kermani-Arab, J. Slots,
[100] F.B. Hashemi, M. Ghassemi, S. Faro, A. Aroutcheva, G.T. Spear, Induction of Cytomegalovirus and Epstein-Barr virus DNA transcription in endodontic symp-
human immunodeficiency virus type 1 expression by anaerobes associated with tomatic lesions, Oral Microbiol. Immunol. 18 (2) (2003) 104–108.
bacterial vaginosis, J. Infect. Dis. 181 (5) (2000) 1574–1580. [128] I. Saygun, A. Kubar, A. Ozdemir, M. Yapar, J. Slots, Herpesviral-bacterial inter-
[101] S. Cu-Uvin, J.W. Hogan, A.M. Caliendo, J. Harwell, K.H. Mayer, C.C. Carpenter, relationships in aggressive periodontitis, J. Periodontal Res. 39 (4) (2004)
H.I.V.E.R. Study, Association between bacterial vaginosis and expression of human 207–212.
immunodeficiency virus type 1 RNA in the female genital tract, Clin. Infect. Dis. 33 [129] J. Slots, J.J. Kamma, C. Sugar, The herpesvirus-Porphyromonas gingivalis-peri-
(6) (2001) 894–896. odontitis axis, J. Periodontal Res. 38 (3) (2003) 318–323.
[102] S.A. Williams, W.C. Greene, Host factors regulating post-integration latency of [130] S. Thakker, S.C. Verma, Co-infections and pathogenesis of KSHV-Associated ma-
HIV, Trends Microbiol. 13 (4) (2005) 137–139. lignancies, Front. Microbiol. 7 (2016) 151.
[103] S.A. Williams, L.F. Chen, H. Kwon, C.M. Ruiz-Jarabo, E. Verdin, W.C. Greene, NF- [131] M. Michaelis, H.W. Doerr, J. Cinatl, The story of human cytomegalovirus and
kappaB p50 promotes HIV latency through HDAC recruitment and repression of cancer: increasing evidence and open questions, Neoplasia 11 (1) (2009) 1–9.
transcriptional initiation, EMBO J. 25 (1) (2006) 139–149. [132] G. Herbein, A. Kumar, The oncogenic potential of human cytomegalovirus and
[104] G. Jiang, A. Espeseth, D.J. Hazuda, D.M. Margolis, c-Myc and Sp1 contribute to breast cancer, Front. Oncol. 4 (2014) 230.
proviral latency by recruiting histone deacetylase 1 to the human im- [133] S. Saravani, H. Kadeh, E. Miri-Moghaddam, A. Zekri, N. Sanadgol, A. Gholami,
munodeficiency virus type 1 promoter, J. Virol. 81 (20) (2007) 10914–10923. Human cytomegalovirus in oral squamous cell carcinoma in southeast of Iran,
[105] K. Imai, K. Yamada, M. Tamura, K. Ochiai, T. Okamoto, Reactivation of latent HIV- Jundishapur J. Microbiol. 8 (8) (2015) e21838.
1 by a wide variety of butyric acid-producing bacteria, Cell. Mol. Life Sci. 69 (15) [134] L.E. Cavallin, P. Goldschmidt-Clermont, E.A. Mesri, Molecular and cellular me-
(2012) 2583–2592. chanisms of KSHV oncogenesis of Kaposi's sarcoma associated with HIV/AIDS,
[106] K. Imai, K. Ochiai, T. Okamoto, Reactivation of latent HIV-1 infection by the PLoS Pathog. 10 (7) (2014) e1004154.
periodontopathic bacterium Porphyromonas gingivalis involves histone mod- [135] O. Radu, L. Pantanowitz, Kaposi sarcoma, Arch. Pathol. Lab. Med. 137 (2) (2013)
ification, J. Immunol. 182 (6) (2009) 3688–3695. 289–294.
[107] M.G. Riggs, R.G. Whittaker, J.R. Neumann, V.M. Ingram, n-Butyrate causes his- [136] T.L. Morris, R.R. Arnold, J. Webster-Cyriaque, Signaling cascades triggered by
tone modification in HeLa and Friend erythroleukaemia cells, Nature 268 (5619) bacterial metabolic end products during reactivation of Kaposi's sarcoma-asso-
(1977) 462–464. ciated herpesvirus, J. Virol. 81 (11) (2007) 6032–6042.
[108] J.K. Nicholson, E. Holmes, J. Kinross, R. Burcelin, G. Gibson, W. Jia, S. Pettersson, [137] Y.H. Ko, EBV and human cancer, Exp. Mol. Med. 47 (2015) e130.
Host-gut microbiota metabolic interactions, Science 336 (6086) (2012) [138] K. Imai, H. Inoue, M. Tamura, M.E. Cueno, H. Inoue, O. Takeichi, K. Kusama,
1262–1267. I. Saito, K. Ochiai, The periodontal pathogen Porphyromonas gingivalis induces
[109] V.H. Ferreira, A. Nazli, G. Khan, M.F. Mian, A.A. Ashkar, S. Gray-Owen, R. Kaul, the Epstein-Barr virus lytic switch transactivator ZEBRA by histone modification,
C. Kaushic, Endometrial epithelial cell responses to coinfecting viral and bacterial Biochimie 94 (3) (2012) 839–846.
pathogens in the genital tract can activate the HIV-1 LTR in an NF{kappa}B-and [139] S. Spadavecchia, O. Gonzalez-Lopez, K.D. Carroll, D. Palmeri, D.M. Lukac,
AP-1-dependent manner, J. Infect. Dis. 204 (2) (2011) 299–308. Convergence of Kaposi's sarcoma-associated herpesvirus reactivation with Epstein-
[110] L.V. Torian, H.A. Makki, I.B. Menzies, C.S. Murrill, D.A. Benson, F.W. Schween, Barr virus latency and cellular growth mediated by the notch signaling pathway in
I.B. Weisfuse, High HIV seroprevalence associated with gonorrhea: new York City coinfected cells, J. Virol. 84 (20) (2010) 10488–10500.
Department of Health, sexually transmitted disease clinics, 1990–1997, AIDS 14 [140] G. Kassiotis, Endogenous retroviruses and the development of cancer, J. Immunol.
(2) (2000) 189–195. 192 (4) (2014) 1343–1349.
[111] R.J. Malott, B.O. Keller, R.G. Gaudet, S.E. McCaw, C.C. Lai, W.N. Dobson-Belaire, [141] E.S. Lander, L.M. Linton, B. Birren, C. Nusbaum, M.C. Zody, J. Baldwin, K. Devon,
J.L. Hobbs, F. St Michael, A.D. Cox, Moraes TF et al: Neisseria gonorrhoeae-derived K. Dewar, M. Doyle, W. FitzHugh, Initial sequencing and analysis of the human
heptose elicits an innate immune response and drives HIV-1 expression, Proc. Natl. genome, Nature 409 (6822) (2001) 860–921.
Acad. Sci. U. S .A. 110 (25) (2013) 10234–10239. [142] E. Lee, R. Iskow, L. Yang, O. Gokcumen, P. Haseley, L.J. Luquette 3rd, J.G. Lohr,
[112] A. Chen, I.C. Boulton, J. Pongoski, A. Cochrane, S.D. Gray-Owen, Induction of C.C. Harris, L. Ding, R.K. Wilson et al, Landscape of somatic retrotransposition in
HIV-1 long terminal repeat-mediated transcription by Neisseria gonorrhoeae, AIDS human cancers, Science 337 (6097) (2012) 967–971.
17 (4) (2003) 625–628. [143] G.R. Young, U. Eksmond, R. Salcedo, L. Alexopoulou, J.P. Stoye, G. Kassiotis,
[113] J.O. Kistler, P. Arirachakaran, Y. Poovorawan, G. Dahlen, W.G. Wade, The oral Resurrection of endogenous retroviruses in antibody-deficient mice, Nature 491
microbiome in human immunodeficiency virus (HIV)-positive individuals, J. Med. (7426) (2012) 774–778.
Microbiol. 64 (9) (2015) 1094–1101. [144] G.R. Young, B. Mavrommatis, G. Kassiotis, Microarray analysis reveals global
[114] H. Tenenbaum, R. Elkaim, F. Cuisinier, M. Dahan, P. Zamanian, J.M. Lang, modulation of endogenous retroelement transcription by microbes, Retrovirology
Prevalence of six periodontal pathogens detected by DNA probe method in HIV vs 11 (2014) 59.
non-HIV periodontitis, Oral Dis. 3 (Suppl 1) (1997) S153–155. [145] T. Hung, G.A. Pratt, B. Sundararaman, M.J. Townsend, C. Chaivorapol,
[115] P.A. Murray, J.R. Winkler, W.J. Peros, C.K. French, J.A. Lippke, DNA probe de- T. Bhangale, R.R. Graham, W. Ortmann, L.A. Criswell, G.W. Yeo, The Ro60 au-
tection of periodontal pathogens in HIV-associated periodontal lesions, Oral toantigen binds endogenous retroelements and regulates inflammatory gene ex-
Microbiol. Immunol. 6 (1) (1991) 34–40. pression, Science 350 (6259) (2015) 455–459.
[116] T.E. Rams, M. Andriolo Jr., D. Feik, S.N. Abel, T.M. McGivern, J. Slots, [146] D. Roulois, H. Loo Yau, R. Singhania, Y. Wang, A. Danesh, S.Y. Shen, H. Han,
Microbiological study of HIV-related periodontitis, J. Periodontol. 62 (1) (1991) G. Liang, P.A. Jones, T.J. Pugh, DNA-Demethylating agents target colorectal
74–81. cancer cells by inducing viral mimicry by endogenous transcripts, Cell 162 (5)
[117] M. Tezal, M.A. Sullivan, M.E. Reid, J.R. Marshall, A. Hyland, T. Loree, C. Lillis, (2015) 961–973.
L. Hauck, J. Wactawski-Wende, F.A. Scannapieco, Chronic periodontitis and the [147] K.B. Chiappinelli, P.L. Strissel, A. Desrichard, H. Li, C. Henke, B. Akman, A. Hein,
risk of tongue cancer, Arch. Otolaryngol. Head Neck Surg. 133 (5) (2007) N.S. Rote, L.M. Cope, A. Snyder, et al., Inhibiting DNA methylation causes an
450–454. interferon response in cancer via dsRNA including endogenous retroviruses, Cell
[118] M. Tezal, M.A. Sullivan, A. Hyland, J.R. Marshall, D. Stoler, M.E. Reid, T.R. Loree, 164 (5) (2016) 1073.
N.R. Rigual, M. Merzianu, L. Hauck, Chronic periodontitis and the incidence of [148] E. Cherkasova, Q. Weisman, R.W. Childs, Endogenous retroviruses as targets for
head and neck squamous cell carcinoma, Cancer Epidemiol. Biomarkers Prev. 18 antitumor immunity in renal cell cancer and other tumors, Front. Oncol. 3 (2013)
(9) (2009) 2406–2412. 243.
[119] M. Tezal, S.G. Grossi, R.J. Genco, Is periodontitis associated with oral neoplasms? [149] G. Danaei, S. Vander Hoorn, A.D. Lopez, C.J. Murray, M. Ezzati, Comparative Risk
J. Periodontol. 76 (3) (2005) 406–410. Assessment collaborating g: causes of cancer in the world: comparative risk as-
[120] M. Kruger, T. Hansen, A. Kasaj, M. Moergel, The correlation between chronic sessment of nine behavioural and environmental risk factors, Lancet 366 (9499)
periodontitis and oral cancer, Case Rep. Dent. 2013 (2013) 262410. (2005) 1784–1793.
[121] J. Ahn, S. Segers, R.B. Hayes, Periodontal disease, Porphyromonas gingivalis [150] B. Vogelstein, K.W. Kinzler, Cancer genes and the pathways they control, Nat.
serum antibody levels and orodigestive cancer mortality, Carcinogenesis 33 (5) Med. 10 (8) (2004) 789–799.

23
D. Vyshenska et al. Seminars in Immunology 32 (2017) 14–24

[151] C.J. Chen, K.Y. Liang, Y.S. Chang, Y.F. Wang, T. Hsieh, M.M. Hsu, J.Y. Chen, [164] I. Mellman, G. Coukos, G. Dranoff, Cancer immunotherapy comes of age, Nature
M.Y. Liu, Multiple risk factors of nasopharyngeal carcinoma: epstein-Barr virus, 480 (7378) (2011) 480–489.
malarial infection, cigarette smoking and familial tendency, Anticancer Res. 10 [165] A. Morgun, A. Dzutsev, X. Dong, R.L. Greer, D.J. Sexton, J. Ravel, M. Schuster,
(2B) (1990) 547–553. W. Hsiao, P. Matzinger, N. Shulzhenko, Uncovering effects of antibiotics on the
[152] J. Wu, B.A. Peters, C. Dominianni, Y. Zhang, Z. Pei, L. Yang, Y. Ma, M.P. Purdue, host and microbiota using transkingdom gene networks, Gut 64 (11) (2015)
E.J. Jacobs, S.M. Gapstur, Cigarette smoking and the oral microbiome in a large 1732–1743.
study of American adults, ISME J. 10 (10) (2016) 2435–2446. [166] X. Dong, A. Yambartsev, S.A. Ramsey, L.D. Thomas, N. Shulzhenko, A. Morgun,
[153] S.G. Grossi, J.J. Zambon, A.W. Ho, G. Koch, R.G. Dunford, E.E. Machtei, Reverse enGENEering of regulatory networks from big data: a roadmap for biol-
O.M. Norderyd, R.J. Genco, Assessment of risk for periodontal disease: i. Risk ogists, Bioinform. Biol. Insights 9 (2015) 61–74.
indicators for attachment loss, J. Periodontol. 65 (3) (1994) 260–267. [167] N. Segata, J. Izard, L. Waldron, D. Gevers, L. Miropolsky, W.S. Garrett,
[154] L.A. Brinton, R.F. Hamman, G.R. Huggins, H.F. Lehman, R.S. Levine, K. Mallin, C. Huttenhower, Metagenomic biomarker discovery and explanation, Genome
J.F. Fraumeni, Jr.: Sexual and reproductive risk factors for invasive squamous cell Biol. 12 (6) (2011) R60.
cervical cancer, J. Natl. Cancer Inst. 79 (1) (1987) 23–30. [168] A.L. Radtke, M.M. Herbst-Kralovetz, Culturing and applications of rotating wall
[155] S. Smart, A. Singal, A. Mindel, Social and sexual risk factors for bacterial vaginosis, vessel bioreactor derived 3D epithelial cell models, J. Vis. Exp. 62 (2012).
Sex Transm. Infect. 80 (1) (2004) 58–62. [169] S.Y. Doerflinger, A.L. Throop, M.M. Herbst-Kralovetz, Bacteria in the vaginal
[156] E. Filipp, P. Raczynski, A. El Midaoui, A. Pawlowska, U. Tarnowska-Madra, microbiome alter the innate immune response and barrier properties of the human
A. Scholz, K.T. Niemiec, J. Chamerski, Chlamydia trachomatis infection in sexually vaginal epithelia in a species-specific manner, J. Infect. Dis. 209 (12) (2014)
active adolescents and young women, Med. Wieku Rozwoj. 9 (1) (2005) 57–64. 1989–1999.
[157] S.C. Chuang, C. La Vecchia, P. Boffetta, Liver cancer: descriptive epidemiology and [170] B.E. Hjelm, A.N. Berta, C.A. Nickerson, C.J. Arntzen, M.M. Herbst-Kralovetz,
risk factors other than HBV and HCV infection, Cancer Lett. 286 (1) (2009) 9–14. Development and characterization of a three-dimensional organotypic human
[158] M. Million, M. Maraninchi, M. Henry, F. Armougom, H. Richet, P. Carrieri, vaginal epithelial cell model, Biol. Reprod. 82 (3) (2010) 617–627.
R. Valero, D. Raccah, B. Vialettes, D. Raoult, Obesity-associated gut microbiota is [171] N. Li, D. Wang, Z. Sui, X. Qi, L. Ji, X. Wang, L. Yang, Development of an improved
enriched in Lactobacillus reuteri and depleted in Bifidobacterium animalis and three-dimensional in vitro intestinal mucosa model for drug absorption evalua-
Methanobrevibacter smithii, Int. J. Obesity (2005) 36 (6) (2012) 817–825. tion, Tissue Eng. Part C Methods 19 (9) (2013) 708–719.
[159] R. Greer, X. Dong, A. Morgun, N. Shulzhenko, Investigating a holobiont: micro- [172] Y. Chen, Y. Lin, K.M. Davis, Q. Wang, J. Rnjak-Kovacina, C. Li, R.R. Isberg,
biota perturbations and transkingdom networks, Gut Microbes 7 (2) (2016) C.A. Kumamoto, J. Mecsas, D.L. Kaplan, Robust bioengineered 3D functional
126–135. human intestinal epithelium, Sci. Rep. 5 (2015) 13708.
[160] A. Sivan, L. Corrales, N. Hubert, J.B. Williams, K. Aquino-Michaels, Z.M. Earley, [173] C. Meyers, Organotypic (raft) epithelial tissue culture system for the differentia-
F.W. Benyamin, Y.M. Lei, B. Jabri, M.L. Alegre, Commensal Bifidobacterium tion-dependent replication of papillomavirus, Methods Cell Sci. 18 (3) (1996).
promotes antitumor immunity and facilitates anti-PD-L1 efficacy, Science 350 [174] J.S. Smith, C. Bosetti, N. Munoz, R. Herrero, F.X. Bosch, J. Eluf-Neto, C.J. Meijer,
(6264) (2015) 1084–1089. A.J. Van Den Brule, S. Franceschi, R.W. Peeling, Chlamydia trachomatis and in-
[161] A. Urruticoechea, R. Alemany, J. Balart, A. Villanueva, F. Vinals, G. Capella, vasive cervical cancer: a pooled analysis of the IARC multicentric case-control
Recent advances in cancer therapy: an overview, Curr. Pharm. Des. 16 (1) (2010) study, Int. J. Cancer 111 (3) (2004) 431–439.
3–10. [175] M.M. Madeleine, T. Anttila, S.M. Schwartz, P. Saikku, M. Leinonen, J.J. Carter,
[162] L.E. Gaspar, M. Ding, A review of intensity-modulated radiation therapy, Curr. M. Wurscher, L.G. Johnson, D.A. Galloway, J.R. Daling, Risk of cervical cancer
Oncol. Rep. 10 (4) (2008) 294–299. associated with Chlamydia trachomatis antibodies by histology, HPV type and
[163] J.T. Schiller, D.R. Lowy, Vaccines to prevent infections by oncoviruses, Annu. Rev. HPV cofactors, Int. J. Cancer 120 (3) (2007) 650–655.
Microbiol. 64 (2010) 23–41.

24

You might also like